Vous êtes sur la page 1sur 951

1

2
INDEX

Basic Completion Design & Practices


Topic Summary: Introduces features used to categorize well completions, emphasizing the importance of correct
tubing design and techniques. Presents downhole components common in producing wells. Outlines the major
completion operations: cementing, logging, perforating, stimulation and sand control.

Estimated Time to Complete:


2 hours to 4 hours. (average of 20 to 40 minutes per Subtopic)

Subtopic Listing:
General Design Criteria
Introduction
Well and Reservoir Parameters
Functional and Well Servicing Requirements
Drilling Considerations
Specifications and Regulations

Basic Downhole Configurations


Openhole Completions
Liner Completions
Cemented & Perforated Completions
Single String Completions (Single Zone)
Single String Completions (Selective)
Multiple String Completions

Lift Methods
Flowing Wells
Rod Pump
Gas Lift
Electrical Submersible Pump (ESP)
Hydraulic Pump
Plunger Lift

Subsea Completions
Overview

Completion Productivity
Introduction
Inflow Performance Relationships
Tubing Performance
Matching Completion and Reservoir Performance
Artificial Lift Requirements; Targets & Allowables
Formation Damage Considerations
Drilling Fluids Selection
Perforating
Completion Fluids Selection and Treatment
Stimulation

3
Completion Planning
Preplanning
Role of Experience and Analogy
Post-Drilling Analysis
Production Reviews
Sources of Geological and Engineering Data
Regulatory Constraints
Programming and Supervision

Cementing
Topic Summary: Covers procedures and equipment used in primary and secondary cement jobs.
Reviews the basic chemistry of various bulk cement and cement additives, as well as industry-standard mixing
procedures. Describes surface and downhole equipment, precise placement techniques, quality control and safe
work practices.

Estimated Time to Complete:


3 hours 40 minutes to 7 hours 20 minutes. (average of 20 to 40 minutes per Subtopic)

Subtopic Listing:
Introduction
History and Overview
Primary Cementing
Squeeze Cementing
Plug Cementing

Manufacture & Classification of Oilwell Cements


Cement Manufacture & Chemistry
API Cement Types
Specialty Cements
Slurry Formulation

Cement Testing Procedures


Overview
Soundness & Fineness Tests
Filtration Rate Test
Compressive Strength Test
Thickening Time Test
Slurry Density Test

Cement Additives
Introduction
Accelerators
Retarders
Extenders (Density-Reducing Additives)
Weighting Materials (Density-Increasing Additives)
Filtration Control Additives
Friction Reducing Additives
Lost Circulation Materials
Salt
Compressive Strength Stabilizers

4
Other Additives

Primary Cementing Equipment


Bulk Loading and Transport
Cement Mixing & Pumping Equipment
Cementing Head
Float Equipment, Guide Shoe
Wiper Plugs
Centralizers
Wipers, Scratchers, Agitators

Primary Cementing Operations


Introduction
Temperature, Slurry Volume and Depth
Axial Loads, Hydraulic Bonding and Casing Support
Single-Stage Cementing
Multiple-Stage Cementing
Special Equipment for Multiple-Stage Cementing
Conductor Casing
Surface Casing
Intermediate Casing
Production Casing
Liner Cementing Procedures
Liner Cementing Equipment
Liner Cementing in Deep Wells
Inner String Cementing
External Cement-Filled Packer
Effect of Hole Conditions
Mud Contamination
General Job Guidelines

Squeeze Cementing
Introduction
Basic Concepts & Common Misconceptions
High-Pressure Methods
Low-Pressure Methods
Squeeze Techniques
Job Evaluation

Cement Plugs
Reasons for Setting Plugs
Job Design
Placement Techniques
Testing Cement Plugs
Plug Failure

Evaluating/Testing the Cement Job


Introduction
Temperature Logs and Tracers
Acoustic and Other Logging Measurements
Inflow Tests & Production Testing/Logging

5
Fluid Flow Properties & Mud Displacement
Introduction
Flow Property Measurement: Slurry Preparation
Calculation Formulas & Sample Problems
Mud Conditioning & Displacement
Improving Mud Displacement

References & Additional Information


References
Additional Reading

Completion Equipment
Topic Summary: Presents downhole components common in producing wells.

Estimated Time to Complete:


2 hours 20 minutes to 4 hours 40 minutes. (average of 20 to 40 minutes per Subtopic)

Subtopic Listing:
Tubulars
API Specifications
Tubing String Design
Production Casing
Couplings

Packers
Packer Functions
Packer Types and Applications
Tubing/Packer Forces and Movement

Artificial Lift Equipment


Anchors
Pumps

Downhole Completion Accessories


Seating Nipples
Sliding Sleeves
Side Pocket Mandrels
Blast Joints & Flow Couplings

Subsurface Safety Valves (SSSVs)


Applications
Flow-Controlled SSSVs
Surface-Controlled SSSVs
Bottomhole Chokes and Regulators

Wellhead Equipment
Wellheads
Tubing Heads and Hangers
Christmas Trees
Beans and Chokes

References & Additional Information


Nomenclature

6
References
Additional Reading
Additional Reading
Additional Reading
API Publications Related to Completion Design
Selected List of Published Multiphase Gradient Curves

Perforating
Topic Summary: Describes the operation of shaped charges, as well as basic categories of perforating guns and perforating meth
safety guidelines.

Estimated Time to Complete:


2 hours to 4 hours. (average of 20 to 40 minutes per Subtopic)

Subtopic Listing:
Introduction
History and Background
Basic Perforating Methods
Fundamentals of Shaped Charge Operation
Explosives Used in Perforating

Perforating Gun Types


Retrievable vs. Expendable Guns
Casing Guns
Through-Tubing Guns
Casing Deformation Considerations

Perforating & Production Performance


API Testing Procedures
Perforation Productivity Considerations
Relative Importance of Perforating Parameters
Perforation Cleanup Considerations
Perforating for Gravel Packing
Perforating for Fracturing

Electric Wireline Perforating


Equipment
Pressure Control System
Perforating Procedures and Depth Control
Positioned or Oriented Perforating
Job Planning

Tubing-Conveyed Perforating
Applications
Perforating Equipment
Perforating Procedures
Measurement While Perforating

References & Additional Information


References
Perforating Safety

Acidizing & Other Chemical Treatments

7
Topic Summary: Introduces safe, cost-effective acid treatment design. Covers three basic treatments: acid washing, matrix acid

Estimated Time to Complete:


3 hours to 6 hours. (average of 20 to 40 minutes per Subtopic)

Subtopic Listing:
Introduction to Acidizing
Background & Definitions
Types of Acid Treatments

Acid Treating Solutions


Mineral Acids
Organic Acids
Specialty Acids

Additives in Acidizing Fluids


Corrosion Inhibitors
Surfactants
Iron Sequestering Agents
Fluid Loss Control Agents
Friction Reducers
Clay Stabilizers
Bacteria Control Agents

Matrix Acidizing
Introduction
Surface Equipment
Sandstone Matrix Treatments
Carbonate Matrix Treatments

Fracture Acidizing
General Considerations
Fracture Geometry
Formation Characteristics
Treatment Materials and Procedures

Diverting Materials in Acidizing


Introduction
Solid Diverters
Gel Diverters
Ball Sealers
Selection of Diverters

Wellbore Cleanout and Scale Removal


Wellbore Cleanout: General Description
Materials Used in Wellbore Cleanout
Wellbore Cleanout Procedure
Scale Problems
Scale Removal
Scale Prevention

Paraffins and Asphaltenes


Basic Properties

8
Factors Affecting Deposition
Removal and Control Techniques

References & Additional Information


References
Diluting & Testing Hydrochloric Acid Mixtures
Additional Reading

Hydraulic Fracturing
Topic Summary: Introduces three aspects of hydraulic fracture stimulation: theory, design, and execution.
Reviews rock mechanics and fracture modeling. Discusses hydraulic fracture stimulation treatments,
including determining fracture dimensions and selecting proppants and chemical additives. Presents case
study of an actual fracturing job.

Estimated Time to Complete:


2 hours 20 minutes to 4 hours 40 minutes. (average of 20 to 40 minutes per Subtopic)

Subtopic Listing:
Hydraulic Fracturing Fundamentals
Objectives & Economic Considerations
Fracture Initiation Orientation & Growth
Role of Formation Properties in Fracturing
Quantitative Description of Fracture Growth

Hydraulic Fracturing Fluids


Fluid Functions
Fluid Loss Control
Rheology
Proppant Carrying Ability & Friction Loss
Pressure/Temperature/pH Limitations
Types of Fluids
Fracturing Additives

Proppant Agents
Introduction
Proppant Pack Permeability & Fracture Conductivity
Properties Affecting Proppant Performance
Types of Proppants

Acid Fracturing
Introduction
Acid Fluid Loss Control
Acid Reaction Rate
The Changing Role of Acid Fracturing

Hydraulic Fracture Treatment Design & Execution


Well Completion Considerations
Basic Design
Material Selection
Equipment Selection
Operational Considerations
Fracture Calibration Treatment

9
Main Treatment
Treatment Evaluation

Hydraulic Fracturing: Emerging Technologies


High Permeability Fracturing (Frac & Pack)
Fracturing in Horizontal & Deviated Wells

References & Additional Information


Nomenclature
References
Additional Reading
Units, Conversion Factors & Formulas

Sand Control
Topic Summary: Covers sand control topics from rock mechanics to equipment. Emphasizes estimation and
control methods. Presents gravel-pack design and placement procedures in detail, describing surface equipment
used, as well as downhole screens and associated equipment.

Estimated Time to Complete:


2 hours 40 minutes to 5 hours 20 minutes. (average of 20 to 40 minutes per Subtopic)

Subtopic Listing:
Sand Production in Oil & Gas Wells
Sand-Related Production Problems
Causes of Sand Production
Classification of Sands and their Failure Mechanisms
Rock Mechanics Considerations
Sand Failure Prediction

Completion & Production Practices


Completion Efficiency
Rate Control & Arching Effects
Selective Perforation

Liners & Screens


Plain Slotted Liners
Wire-Wrapped Screens
Applications and Limitations of Plain Screens
Prepacked Screens

Gravel Packing
Overview
Cased Hole (Internal) Gravel Packs
Open Hole (External) Gravel Packs
Well Productivity
Gravel Selection, Sizing and Quality
Liner/Screen Selection
Downhole Tools
Surface Equipment
Perforation Design and Cleanup Procedures
Gravel Placement Techniques (Cased Hole)
Gravel Placement Techniques (Open Hole)

10
Example: Cased Hole Gravel Pack Design
Slurry Requirements
Outline Procedures for Gravel Packs
Fluids Used in Gravel Pack Operations
Formation Damage & Fluid Filtration
Multiple Zone Completions
Deviated and Horizontal Wells
Quality Control

Frac-and-Pack Techniques
Overview

Consolidation Techniques
Chemical Consolidation
Consolidated Gravel Packs

Sand Control in Wells Producing Heavy Oil


Introduction
Gravel Packing
Other Control Methods

References & Additional Information


References
Additional Reading

Horizontal Wells: Completion & Evaluation


Topic Summary: Introduces the deliverability equation for horizontal wells. Compares horizontal and vertical
well performance. Identifies formation characteristics that favor horizontal wells, and discusses methods for their
determination. Describes basic well configurations and completion designs. Looks at practical aspects of cementing,
zonal isolation, casing design and sand control. Address guidelines for matrix acid treatments and hydraulic fracture
stimulations.

Estimated Time to Complete:


2 hours to 4 hours. (average of 20 to 40 minutes per Subtopic)

Subtopic Listing:
Horizontal Wells in a Reservoir Management Strategy
Introduction
Reservoir Candidate Recognition
Horizontal Well Deliverability
Anisotropic Formations
Horizontal Wells vs. Fractured Vertical Wells
Multilateral Wells
Formation Evaluation for Optimal Well Completions

Horizontal Well Completions


Well Curvature
Completion Types
Zonal Isolation Techniques
Casing Design
Cementing
Sand Control

Matrix Stimulation of Horizontal Wells

11
Introduction
Damage Characterization
Fluid Placement Techniques
Partial Stimulation
Stimulation of Sandstone Reservoirs
Stimulation of Carbonate Reservoirs
Optimization of Matrix Treatments
Acidizing Physics

Hydraulic Fracturing of Horizontal Wells


Introduction
General Considerations
Performance of Fractured Horizontal Wells
Fracturing Pressures
Fracture Design
Fracture Execution and Perforating Considerations

Multilateral, Multibranched and Multilevel Wells


Introduction
Reservoirs and Well Architecture
A Production Prediction Model
Reservoir Performance, Selected Configurations
Partially Completed Wells
Completions for Complex Well Architecture
References

References & Additional Information


References
Nomenclature
Additional Reading
SI Metric Conversion Factors for Common Field Units

12
BASIC COMPLETION & DESIGN PRACTICES

GENERAL DESIGN CRITERIA

Basic Completion Categories


There is a significant diversity in the type of completions being used around the world.
However, in general they are variations on a few basic designs. The most common
criteria for classifying completions include

The Interface between the Wellbore and Reservoir

openhole completions
liner completions
perforated completions

The Production Method

artificial lift
flowing

The Number of Tubing Strings

tubingless
single string
multiple strings

The Surface Location

onshore
offshore (platform)
offshore (subsea)

The Stage of Completion

initial completion
recompletion
workover

Single-zone completions include downhole commingling of production from several intervals and
may be designed to allow sequential development of successive reservoirs. Multizone completions
include not only the separation of various zones but also segregation of individual sand units within
a thick pay section for reservoir control purposes.

Beyond these major classifications, the completion complexity is largely a function of


the problems encountered and the prevailing economic constraints.

13
COMPLETION SELECTION AND DESIGN CRITERIA
Well completion designs will vary significantly with:

gross production rate;


well pressure and depth;
rock properties;
fluid properties;
well location.
Typical ranges for various classes of completions and the design implications are presented in
Table 1. This table, of course, represents a partial list of well parameters; there are many other
variables that figure into a given completion design. Given the variety of production conditions
around the world, definition of the thresholds is naturally somewhat nebulous (a low production rate
in a Middle Eastern well would be considered a very respectable rate in many North American
fields). However, this table gives a general idea of the range of design considerations.

Table 1: Completion Design Considerations


Well Parameters Design Implications
High Production Rate: Significant frictional pressure losses; Large diameter tubing
(1500-10,000 B/D liquid [160- (>2 7/8 in. or 73 mm); Large diameter casing (>5 1/2 in. or
16,000 m3/d]; 35-140 140 mm); Special artificial lift equipment; Thermal
MMSCF/d gas [1 - 4 106
contraction/expansion equipment; Erosion control equipment
m3/d]).

Low Production Rate: Artificial lift required; Paraffin buildup problems; Special
(<30 B/D liquid [5 m3/d]; < 1 attention to operating costs required.
MMSCF/d gas [30 103 m3/d]).

Very High Pressure: Special stress checks required during completion; High-
(10,000-25,000 psi [70-175 MPa] strength tubulars required; Special high-performance
packers/accessories required; Problems with H2S
aggravated by high pressure requiring special tubular steel

High Pressure: Flanged, rather than threaded, wellheads required; Well-


(3000-10,000 psi [20-70 MPa] killing capabilities required

Low Pressure : Threaded wellheads may be used; Artificial lift required;


(< 1000 psi [< 7 MPa] Greater risk of damage/fracturing during completion process

14
Deep Wells: Problems associated with high pressures; Tubular
(> 10000 ft [ >3000 m] weight/tension must be considered; Casing size/liner usage
must be considered; Hydraulic piston pumps or gas lift more
likely to be used as artificial lift; External corrosion of
tubulars may be a problem due to higher pressure and
temperature

Carbonate Reservoirs: Acid wash required upon completion; Difficulty identifying


water contact--need formation or drillstem tests

Very Low Permeability (<1 md): Fracturing required upon completion

Low Permeability (1-50 md): May need fracturing upon completion

Moderate Permeability ( >50 md): Little benefit from fracturing; Matrix acidizing may be
necessary; Moderate pressure drawdown across
perforations

High Permeability ( >1000 md ): Lost circulation a problem; Sand strength may not be great
enough to support high velocity flow; Easily damaged

Unconsolidated sandstone: Sand control (screens or gravel pack) probably required

Partially consolidated and friable Sand control possibly required; Minimize drawdown to
sandstone: prevent sand production; Maximize sand exposed to flow;
(acoustic log reads >100s/ft Selective perforation required; Difficult to fracture
[328s/m]; compressive strength successfully
<1000 psi [<7 MPa]; (poor sidewall
core recovery

Hydrogen Sulfide (H2S) present: Special HSE regulations/procedures; Corrosion inhibitors


may be required; Gas usually considered sour if H2S partial
pressure is 70.05 psia (0.3 kPa)

Carbon Dioxide (CO2) present: Consider inhibitor or special steel if CO2 partial pressure is
>10 psi (70 kPa)

Water production : Scaling and/or corrosion may be a problem; Special artificial


lift equipment may be required

Water injection : Consider oxygen corrosion prevention requirements;


Consider backflush requirements

Well location : Offshore--Special HSE regulations; Subsurface safety


valve requirments; Well servicing and access
constraints

Urban/populated areas--Special HSE regulations;


Noise and height limits

Mountainous areas--Potential for wellhead damage

15
due to landslides

Functional and Well Service Requirements

Definition of the functional and well servicing requirements at the outset can
considerably simplify selection of preliminary completion concepts and will highlight the
key trade-offs needing further evaluation. Table 1 is a checklist for identifying the
critical concerns for a completion design; it illustrates the use of such a checklist in
designing a specific subsea oilwell. The completions engineer relies on experience and
judgment to prepare the initial input at the concept stage. However, as development
plans become more clearly defined, it is often possible to quantify the requirements,
based on the results of the initial wells or of detailed design or field studies.

Completion Importance Completion


Considerations or Need Design
Implications

Rates

High None

Moderate w/chokes High favors two small


tubing strings

Low Possible

Variable Critical

Pressures

High None

Low Probable artificial lift


required

Producing Characteristics

Multiple zones Possible stack completions

Minimize costs Moderate review costs

Access difficulty High TFL/new technology

16
Uptime High minimize difficulty of future workovers

Rate control Critical chokes needed

Rate stability Critical wellhead chokes needed

Long life Unlikely carbon steel sufficient

Density of kill fluid Moderate kickoff w/gas lift

Safety during vessel Critical 2 SSSVs and kill system


reentry

Wellhead damage Possible annular SSSV

Monitoring

Test frequency High critical choke bean or dedicated

flow line

Pressure Moderate TFL access for downhole tools


measurement

Special BHP surveys Some needed TFL access for downhole tools

Log contacts Critical vertical access required

Production logs Some needed vertical access required

Tubing investigation High TFL access &/or vertical access

Artificial Lift

Intermittent High gas lift is optimal method


w/maintenance

via TFL and vertical


access

Continuous Possible

Increasing gross rate High

Pressure depletion Possible

17
Kick-off

Initial completion Moderate use gas lift system

Routine High gas compressor supply


operations required

Depleted Possible
conditions

High water cut High

Critical rate High GLV maintenance system

Frequency High

Gas supply Moderate gas compressor special


volume requirements

Gas supply Design


pressure variable

Repairs

Cement High future concurrent production and

workover operations; easy access;

robust tubing joints

Gravel pack Critical

SSSV Probable

Tubulars Low

New interval Possible multizone completion design

Recompletions

Uphole Moderate large casing preferable

Deepen None limit depth of rathole

Sidetrack Possible maximize casing size

Function change Moderate large CSG preferable

18
Well Kill

Frequency High or operations procedure


low

Difficulty Mod.- alternate methods


high

Production Problems

Sand control Critical gravel pack required

Paraffin Possible TFL access for scraping

Emulsions Possible chemical injection


capability

Water cut High artificial lift required

Scale Possible TFL access

Corrosion Moderate carbon steel & downhole


chemical

inhibitor injection

Erosion Low

Fines Probable frequent acid jobs


required

GP failure Moderate TFL w/annular kill valve

Table 1: Subsea oilwell functional requirements.

It is important for the completion design engineer to have some appreciation for the
relative impact of production revenue, capital costs, and operating costs on project
economics. In a high tax environment they are usually in the order of importance listed
above, with the revenue stream being the most critical. Installation costs are only
significant to the extent that special completion requirements have a significant impact
on the overall drilling and completion time. The actual cost of the completion
equipment is often relatively insignificant compared to the value of incremental
production from improved potential or increased uptime. However, production
engineers must not take this argument too far. It is important to remember that, in
most cases, downtime only results in deferred production. (An exception is the case of
competitive production along lease lines.) Nevertheless, for subsea developments in

19
hostile environments, it is reasonable to assume that a premium can be paid for
minimizing the frequency of reentry and for equipment reliability and durability.

To a large extent, reservoir, geological, and economic considerations will dictate the
functional requirements of a completion and the relative significance of major and
minor workovers. These requirements have to be anticipated at an early stage since
the techniques to be employed (wireline, service rig reentry, TFL, coiled tubing, etc.)
are limited by the tubing design and packer/tubing configurations of the completion.

The completion design of a well is also influenced by the well service requirements.The
general term "well servicing" covers a broad range of activities, which can be broken
down into five major functions:

1. routine monitoring (e.g., being able to run production logs, shoot fluid levels, etc.)

2. wellhead and flow line servicing (e.g., designing components for easy
isolation)

3. minor workovers (e.g., through-tubing operations, wireline work, TFL)

4. major workovers (e.g., tubing-pulling operations)

5. emergency situations (e.g., well-killing operations)

While to some extent these apply to all oil and gas developments, their relative importance,
frequency, complexity, and cost are functions of reservoir conditions, governmental regulations,
operating philosophy, and geographic and environmental considerations. For example, it should be
self-evident that the options for reentry of subsea wells in deep water are limited and are going to
be expensive. This is true to a certain extent for any offshore well. The designer must therefore look
carefully at the functions that can be built into the completion and wellhead to minimize well service
requirements.

It is probable that at least three different generic types of systems will be involved in
well servicing: those with functions built into the producing facilities; service units; and
workover rigs.

From a completion design viewpoint, it is also important to appreciate what capabilities


are already inherently available. For example, all wells have the potential for "bull-
heading" kill or treatment fluids through the tubing, although it becomes more difficult
to control the operation and ensure an efficient displacement as the tubing size and
deviation increases. Similarly, with relatively shallow dry gas wells, it should be
possible to estimate the bottomhole pressure fairly accurately from tubing head
pressure measurements, avoiding the need to run bottomhole surveys. Another built-in
function in all offshore wells is the ability to achieve a subsurface shut-off using the
government-regulation-required subsurface safety valve.

As completion designs become more sophisticated, they can provide an increased


number of integrated service functions, up to the ultimate multizone, full TFL
completion with downhole pressure monitoring capability. The economic and technical
justification for this type of completion must be based on a detailed functional analysis

20
of the reservoir, completion lifetime, and well service economics. Moreover, increased
sophistication also introduces higher risks of completion problems or subsequent
failures, requiring improved quality control and materials selection.

Drilling Considerations

Several drilling considerations can influence the type of completion installed,


particularly for exploration and delineation wells. Conversely, completion
considerations will help to determine drilling practices in development and infill wells.
Factors to be considered include

1. Probable extent of drilling damage and the resulting requirements for special perforating
or stimulation techniques, or the selection of special drilling fluids, or both.

2. The evaluation program, particularly the need for precompletion testing, to


determine if special logs or tools like the repeat formation tester (RFT) can
reduce testing requirements.

3. The size and weight of the production casing. Table 1 illustrates the
limitations this imposes on the type of completion that can be installed. The
heavyweight tubular casing used in high pressure wells has reduced drift
diameters (internal diameters, or IDs) , which imposes limitations on the
packers and accessories that can be used. For example, the use of 7-in (178-
mm) production casing precludes the use of a dual tubing string with 2 7/8 x 2
7/8-in (73 73-mm) or larger tubing diameter. Depending on the production
capacities and reserves of the various producing zones, a single-string,
multizone completion with larger diameter tubing may be better.

4. The burst and collapse strength of the production casing. The casing must be
able to withstand the maximum closed-in tubing pressures in case of a tubing
break at surface. Similarly, if the well is to be pumped off with an open annulus,
the casing must have adequate collapse strength. Casing strength often
dictates stimulation design, kill procedures, and selection of annulus pressure
operated tools.

5. Wear or corrosion of the production casing must be evaluated in liner


completions, especially for deep wells, and, if necessary, a tie-back string must
be installed. However, use of a tie-back string may limit throughput capacity by
limiting the diameter of the production tubing.

6. In sour (H2S) environments, or where conditions could become sour,


production casing materials should conform to NACE specifications. This is

21
critical in deep, high pressure wells where very small amounts of H2S can result
in a stress cracking risk.

7. The coupling used on the production casing needs to be carefully selected


where high differential pressures (>5000 psi or >34 MPa), high temperatures
(>300 F or >422 K), or high compressional or tensional loads are expected
(e.g., deep wells, high rate wells, thermal wells). Where a gas-tight seal is
essential (e.g., sour or high pressure gas wells or wells with high pressure gas-
lift systems), premium couplings are generally recommended.

8. Proper cementation of the production casing is the key to successful zonal


isolation and avoidance of many production problems.

Table 1: Tubing size and production rate limits based on casing diameter.

Casing Size Maximum Tubing Maximum Maximum Theoretical


Size Theoretical Liquid Gas Rate*
Rate*

(in) (mm) (in) (mm) (b/d) (m3/d) (MMScf/d) (


103m3/d)

4 102 2 3/8 60 2000 300 15 400

4 1/2 113 2 7/8 73 5000 800 25 700

5 1/2 140 3 1/2 89 7500 1200 40 1100

6 5/8 168 4 1/2 114 15,000 2400 80 2300

7 5/8 194 5 1/2 140 20,000 3200 120 3400

9 5/8 244 7 178 60,000 9550 100 2800


*IPR, THP, GLR, and conduit length often prevent such high rates being achieved in specific cases.

b. Casing Requirements for Dual Tubing

Table 1, continued: Casing requirements for dual tubing

Casing Maximum Dual Tubing

(in) (mm) (in) (mm)

9 5/8 244 3 1/2 x 3 1/2 89 x 89

8 5/8 219 3 1/2 x 2 7/8 89 x 73

7 5/8 194 2 7/8 x 2 7/8 73 x 73

7 178 2 7/8 x 2 3/8 73 x 60

22
2 7/8 x 5 concentric 73 x 127 concentric

5 1/2 140 2 1/16 x 1.9 52 x 48

c. Artificial Lift Requirements

Table 1, continued: Artificial lift requirements

Casing Size Nominal Tubing Size Tubing Pump Capacity


Size

(in) (mm) (in) (mm) (in) (mm) (b/d) (m3/d)

Rod Pumps

3 1/2 89 1.9 48 1.50 38 550 100

4 102 2 3/8 60 1.75 44 800 150

4 1/2 113 2 7/8 73 2.25 57 1300 200

5 1/2 140 3 1/2 89 2.75 70 1900 300

Electrical
Submersible

4 1/2 113 2 7/8 73 1750 300

5 1/2 140 3 1/2 89 4000 650

7 178 5 127 10,000 1600

9 5/8 244 7 178 35,000 5550

Based on 144-in stroke and 15 spm 100% efficiency.

Rounded off to nearest 50 m3/d.

Based on a net lift of 3000 ft.

Specifications and Regulations

23
In many well completion situations (e.g., high pressure wells, deep wells, sour gas
wells, and offshore and subsea completions) the design options are constrained by
government regulations, company operating philosophies, and company design
specifications.

In addition, designers are expected to conform to the standards of "good oilfield


practice," which are often embodied in agreements and regulations. Generally, this is
interpreted to mean keeping the well under control with two lines of defense, so that a
single failure or human error will not cause serious injury or environmental damage.
Typical provisions for a moderate to high pressure well are presented in Table 1.

1. During Production
a. Surface

Internal: Xmas-tree wing and master valves and offshore Xmas tree and
SSSV
External: packer and wellhead

b. Subsurface: tubing and casing (check valve and casing for side pocket mandrel
devices)
2. During Drilling and Workover
a. Surface

Internal: mud/workover fluid and BOPs


External: cement and wellhead

b. Subsurface: mud/workover fluid and casing/shoe strength


3. During Lifting BOPs/Xmas Tree
a. Surface

Internal: two plugs or SSSV and plug


External: packer and wellhead, including annular access shutoff via a
valve, plugs, or annular SSSV

b. Subsurface: As in 1b
4. Long-Term Suspension of Completed Well
a. Surface

Internal: deep-set plug and SSSV


External: deep-set plug and packer

b. Subsurface: as in 1b
5. Long-Term Suspension of Uncompleted Well
a. Surface

Internal: two cement and/or bridge plugs


External: as in 2a (external)

b. Subsurface: plug and casing/shoe strength


6. Temporary Suspension of Uncompleted Well
a. Internal: as in 5a (internal); or casing/cement and a kill string/tubing hanger

24
Table 1: Typical provisions of a two-barrier safety philosophy for a moderate to high pressure well.

Even if the well has such low pressures that it tends to kill itself, wellsite personnel
should always be able to rely on a second line of defense (wellhead, BOP, etc.).
Switching off the artificial lift system or lift gas supply can sometimes be considered a
line of defense in pressure control, if this action would normally cause the well to die.

The major design specifications commonly used by the oil industry worldwide are those
issued by the American Petroleum Institute (API). In general the specifications address
the manufacture and testing of components; however, a number of Bulletins and
Recommended Practices address the performance that can be assumed for design
purposes and the procedures to be adopted in implementing that design. The API
specifications of particular relevance to completion design are detailed in Appendix A.
Materials used in sour wells should conform to NACE Specification MR-01-75.

25
BASIC DOWNHOLE CONFIGURATION
Openhole Completions

In an openhole or barefoot completion, the production casing is set in the caprock


above or just into the top of the pay zone, while the bottom of the hole is left uncased
( Figure 1 , (a) openhole completion; (b) uncemented liner completion; (c) perforated
completion).

Figure 1

Often, the final drilling of the pay zone is carried out with special non-damaging drilling
fluids or an underbalanced mud column.

This form of well completion dates back to the days of cable tool drilling, but is rarely
used today. Nonetheless, openhole completions offer certain advantages in thick,
relatively competent formations:

Exposure of entire pay zone to the wellbore;

26
No perforating expense;
less critical need for precise log interpretation;
Reduced drawdown because of the large inflow area;
Slightly reduced casing cost;
Ease of deepening the well;
relative ease of converting the well to a liner completion;
no risk of formation damage resulting from cementing casing.

Unfortunately, the disadvantages and limitations of openhole completions outweigh


these benefits in most cases. Some of these disadvantages are as follows:

inability to control excessive gas-oil and/or water-oil ratios (except in the case of bottom
water);
need to set casing before drilling or logging the pay;
difficulty of controlling the well during completion operations;
unsuitability for producing layered formations consisting of separate reservoirs with
incompatible fluid properties;
inability to selectively stimulate separate zones within the completion interval ;
need for frequent clean-outs if the producing sands are not completely competent or if the
shoulder of the caprock between the shoe and top of the pay is not stable.

Liner Completions

To overcome the problems of collapsing sands plugging the production system, the
early oil producers placed slotted pipe or screens across the openhole section as a
downhole sand filter ( Figure 1 (a) openhole completion; (b) uncemented liner
completion; (c) perforated completion).

27
Figure 1

The simplest and oldest liner completion method involves running slotted pipe into the
open hole interval. The slots are cut small enough that the produced sand bridges off
on the opening rather than passing through. This method is till used in some areeas
today, but because it entails many of the same disadvantages inherent in openhole
completions (i.e., lack of control), its use is not widespread.

For very fine sands, wire-wrapped screens or sintered bronze are used in place of
machine-cut slots. This technique is a reasonably effective sand control method in
uniform coarse sands with little or no fine particles (e.g., in California). Sometimes this
is the only sand control system that can be used because of pressure loss and
placement considerations (e.g., in unconsolidated heavy oil sands).

In general, however, the uncemented liner completion is no longer recommended for


the follwoing reasons:

Sand movement into the wellbore tends to cause permeability impairment by the
intermixing of sand sizes, and of sand and shale particles.
Fine formation sands tend to plug the slots or the screen.
At high rates, the screen often erodes as formation sand moves into the wellbore.

28
Poor support of the formation can cause shale layers to collapse and plug the slots or
screen.
Formation failure can cause the liner itself to collapse.

To overcome these problems, operators have resorted to more effective sand control
methods such as gravel packing, in which the annulus between the screen and the
openhole is filled with coarse, graded sand, or the use of pre-packed screens. In some
cases, even where sand control is planned, it may be best to employ a cased and
perforated completion with an external gravel pack--this configuration has become the
norm for light iol and gas developments because of the flexibility it provides.

Cemented and Perforated Completions

By far the most common type of completion today involves cementing the production
casing (or liner) through the pay zone, and subsequently providing communication
with the formation by perforating holes through the casing and cement ( Figure 1 : (a)
openhole completion; (b) uncemented liner completion; (c) perforated completion).

Figure 1

29
Ideally, perforations should penetrate any damaged zone around the original wellbore
and create a clean conduit within the undamaged formation. If the well is cased and
unperforated during the early stages of the completion operation, well control is easier
and completion costs may be reduced.

Using various depth control techniques, we can select which sections of pay should be
perforated and opened to flow, thereby avoiding undesired fluids (gas, water), weak
zones that might produce sand, and unproductive sections or shale barriers.

This selectivity, which is completely dependent on a good cement job and adequate
perforating, also allows a single well-bore to produce several separate reservoirs
without their being in communication. This is done by setting isolating packers within
an unperforated section of the pipe. Selective perforation can also be used to control
the flow from, or stimulation of, various parts of the pay. By shutting off or partially
plugging selected perforations, injected fluids (water, stimulation fluids, or cement)
can be diverted into less permeable zones.

Cementing casing at TD rather than completing the well openhole can reduce the
likelihood of well control problems. Moreover, the decision to set production casing can
be deferred until the openhole logs of the prospective pay zone have been evaluated,
substantially reducing the dry-hole costs if the hole is dry.

In summary, the advantages of cased and perforated completions include:

Safer operations
More informed selection of the zones to be completed
Reduced sensitivity to drilling damage
Facilitation of selective stimulation
Possibility of multizone completions
Easier planning of completion operations

This type of completion is generally used unless there is a specific reason to prefer an
openhole or uncemented liner completion. Even where sand control is planned,
perforated completions with internal gravel packs have become the norm for light oil
and gas developments because of the flexibility provided.

Single-String Completions (Single Zone)

Producing a well through a tubing string protects the casing from formation fluids and
maximizes flow efficiency. The tubing also provides a means of circulating fluids in the
well.Single-tubing-string completions may or may not use a packer, depending on the
well conditions and the completion method used. Some examples of single-string
completions are shown in Figure 1 ,

30
Figure 1

Single-string flowing wells: (a) temporary; (b) tubingless gas well; (c) simple low cost;
and Figure 2 , (d) high pressure; (e) high-rate liner completion).

31
Figure 2

The complexity of tubing and packer installations is driven by functional requirements


and economic considerations. Since a number of useful features can be installed at
very low incremental cost, the designer should consider these options and possibilities:

Simplification of the completion and future workover operations


Optimum tubing size for maximum long term flowrate
Future artificial lift needs
A "bomb" well for future bottomhole pressure surveys
Use of a permanent packer and tailpipe to protect the formation during workovers or to
facilitate "killing" the well;
The need for moving seals and/or a slip joint to accommodate tubing elongation and
contraction caused by thermal stresses
Anchoring the tubing to the packer
Availability of a downhole sliding sleeve for removing or adding fluid to the tubing (kick-off
or killing operations)
The need for downhole corrosion inhibitor injection
An additional packer and nipple between sets of perforations for future recompletion
operations;

32
Use of tubing-conveyed perforating gun and/or through-tubing guns for underbalanced
perforating to improve completion efficiency.

For single-tubing-string, liner completions, a polished bore receptacle in the liner


hanger is often used in place of a packer. This is simply a polished internal section at
the top of the liner, into which the tubing string is inserted, much as it would be into a
packer. This is useful both for deep wells where tubing/casing clearances are often
small and for very high productivity wells where the use of a packer might cause a
restriction on well productivity. In this type of completion it is useful to incorporate a
landing nipple in the liner string for future isolation of the producing zone via wireline.

Tubingless completions are a particularly low-cost installation method used in marginal


flow operations, such as low rate gas development. While they are also used in high
gas-oil ratio oilfields, problems develop when artificial lift is required. Hollow sucker
rod pumps or small diameter "macaroni" gas-injection tubing strings must be installed.
Minor liquid buildup in gas wells can usually be easily blown out of the well with flexible
small diameter tubing.

Single String Completions (Selective)

Selective completions include both single-string configurations, such as those shown


in Figure 1 (Multiple-zone completions: (a) tubingless; (b) low rate, single string; (c)
high rate, single string) and multiple-string arrangements such as those shown in
Figure 2 (Multiple-zone completions: (d) dual string (parallel gas lift); (e) concentric).

33
Figure 1

34
Figure 2

Single string completions are often preferable to multiple-string completions because


the casing size in multiple-string completions limits the diameter, which, in turn, limits
the flow rate obtainable through each string. Single-string completions may also be
used where segregation is required purely for reservoir control (e.g., in a case where
zones will be commingled at some stages, but shut in during other periods because of
high gas-oil ratio, high water cut, or for some other reason). These completions may
also be used to minimize completion costs, which is also often the reason for limiting
the size of the production casing.

Multiple String Completions

Multiple tubing string completions are generally more expensive than single-string
completions, and more complicated to install and service. They do, however, offer the
ability to simultaneously produce from or inject into different zones, and to allocate
production or injection for each zone.

35
Dual-string completions may be parallel or concentric. Where artificial lift may be
required, parallel strings are usually used. Concentric strings require less clearance and
can often achieve a higher overall flow capability.

Triple-string completions have also been used in some areas, but they are usually too
restrictive of well capacity to be economically attractive as a conventional completion.
The difficulty of future remedial workovers of wells thus completed also prevents their
widespread use.

Multistring tubingless completions are sometimes used for completing stacked deltaic
reservoirs (e.g., the U.S. Gulf Coast) that have low individual reserves and normal
pressures. These completions are particularly attractive for depleting small oil
accumulations below a large gas reservoir and for low cost gas developments. The
improvement in the design and equipment quality of more conventional casing
completions has resulted in a decrease in the popularity of this last type of installation.

36
LIFT METHODS

Flowing Wells

Most flowing well completions employ at least one string of tubing and one or more
packers ( Figure 1 ,

Figure 1

Single-string flowing wells: (a) temporary; (b) tubingless gas well; (c) simple low cost;
and Figure 2 , (d) high pressure; (e) high-rate liner completion).

37
Figure 2

Although a single-string flowing well can be completed without a packer, this can result
in heading due to the large volume of highly compressible gas stored in the annulus.
Heading can result in alternating slugs of liquids and gas being produced up the tubing.
Under most conditions, it is therefore advisable to include a packer in a flowing well
installation. Use of a packer also protects the casing from corrosive fluids and high
pressures. Since artificail lift is likely to become part of a well's producing life at one
point or another, it is common to plan the completion for eventual conversion to
artificial lift.

Beam Pumping Completions

Many low to moderate rate wells (less than 2000 b/d, [320 m3/d]) are produced with
beam pumps because of their ability to provide high formation drawdowns and their
familiarity to North American operators (many low rate wells exist in North American
oil fields).

Figure 1 (Rod pump completions: (a) simple; (b) improved) includes schematics of
pumping well completions.

38
Figure 1

There are three main elements: the bottomhole pump, the rod string, and the pumping
unit.

From a completion viewpoint, the key considerations are

an open annulus
anchoring of the tubing (if required)
adequate pump diameter to displace the target, or maximum production, volume
a pump-seating nipple
a gas anchor (if required)
easy installation and pulling of tubing, especially if a tubing pump is used
properly sized rods
Sucker rods transmit the tensional loads from the plunger to the pumping unit. Their main design
criteria are dynamic effects including fatigue, stretch, and rod fall. Since the mathematical
description of these physical phenomena is complex, rod strings and pumping loads are normally
designed using semiempirical techniques which have been well documented by the API (API RP
11L, Bul. 11L3 and 4). Computer programs, which are available from various pumping unit
manufacturers, can be used to solve the partial differential equations directly.

39
Individual sucker rods are generally 25 ft long and come in diameters from 5/8 in. to 1
1/8 in. A combination of sizes (a tapered string) is normally used. Continuous rod is
also available and becoming more common, although a special service rig is needed.

There are two principal grades of steel sucker rods: "C" Grade, with a tensile strength
of 90,000 psi (0.6 GPa), used in shallow wells; and "D" Grade, with a tensile strength
of 115,000 psi (0.8 GPa), used in deeper wells. Fiber glass rods are also sometimes
used in deep wells or severely corrosive environments.

Table 1, below, provides some useful rules of thumb on design of a pumping system.

Tubing Size (in) 1.25 1.50 2.00 2.50 3.00 4.00


API Pump Type Pump Bores (in)
Rod pump Heavy wall 1 1/16 1 1/4 1 1/2 1 3/4 2 1/4
Rod pump Thin wall 7/8 1 1/4 1 1/2 2 2 1/2 3 1/4
Tubing pump Heavy wall 1 3/4 2 1/4 2 3/4 3 3/4
Approximate Fluid
50 100 300 500 700 1500
rate (BPD)
Normal depths (ft) 10,000 8000 6000 5000 4000 3000

These are typical values, not maximums.

Rules of Thumb

1. First 100 BPD requires 1-1 1/4-in bore pump. Each additional 100 BPD requires 1/4-in
addition to pump bore.

2. Speed times stroke length is usually less than 1500 in/min as a practical
limit.

3. Pump stroke

3 ft at depths less than 3000 ft

4 ft at depth of 4000 ft

5 ft at depth of 5000 ft

6 ft at depth of 6000 ft

4. Pump rate (N) is a function of stroke length, viscosity, etc., but is usually between 8 and
20 SPM.

5. Pump efficiency (Ep) = 75 to 95%, but is usually 90% (a function of rod


stretch)

40
7. Pump displacement (PD) = Ep x 0.1166 S Es N D2

Table 1. Rules of thumb for beam pump applications

Rod pump wells are generally completed with an open annulus through which the gas
is bled off at the surface ( Figure 1 ). All pumping systems (except plunger lift) become
increasingly inefficient in the presence of gas.

Failure to anchor the tubing in rod pumping installations decreases the efficiency of the
pump because of tubing stretch, and can result in rod or tubing wear because of
buckling. However, a tubing anchor may not be necessary where clearances between
casing and tubing are small (less than 0.3 in or 0.76 cm), provided the pump is
modified accordingly. Shallow wells (< 3000 ft or 900 m) often are not equipped with
tubing anchors.

Gas Lift Completions

In essence, continuous gas lift is simply a matter of supplementing the natural flow
process by the addition of compressed gas as either a constant stream or as
intermittent injections ( Figure 1 , Gaslift system schematics).

41
Figure 1

The gas lightens the fluid head in the tubing, reduces the fluid's viscosity, reduces
friction, and supplies potential energy in the form of gas expansion. It can be seen
from the gradient curves presented in Figure 2 (Vertical flowing pressure gradient
curve) that for each tubing size and production rate, there is a decrease in pressure
loss with increasing gas-liquid ratio. An upper limit exists for each given set of
conditions.

42
Figure 2

From a completion viewpoint the key issues in gas-lift design are sizing the tubing and
determining the number and setting depths of the necessary gas-lift valves. Figure 3
(Gaslift design) presents the gas-lift system design as a depth versus pressure plot.

43
Figure 3

Intersection of the optimum gas-liquid ratio curve (drawn down from the THP) and the
formation GLR gradient curve (drawn up from the FBHP) gives the position of the
operating valve. This position may have to be modified if available gas pressure is
limited and the designated depth cannot be reached.

Additional valves are generally required uphole to "kick the well off," particularly if the
tubing will be loaded with kill fluid from time to time. These unloading or "kickoff"
valves will normally be closed during routine gas-lift operations and only reopen when
the tubing pressure exceeds a certain preset limit.

The minimum depth of the first kickoff valve is determined by

the kill fluid gradient

the system pressure the well is to be kicked off against (flow line pressure,
blowdown separator pressure, service line pressure, etc.)

the gas supply pressure during kickoff. (Some companies use special high
pressure compressors for this purpose.)

As shown in Figure 3 , this depth is determined by drawing in a workover fluid pressure


gradient down to the point where it intersects the gas supply pressure gradient line. If

44
some other method is going to be used to remove workover or kill fluids (i.e., tubing,
swabbing), the depth of this valve is determined by the gradient of dead formation
fluid that will exist after the producing well is closed in. Normally, a mini-mum gradient
of 0.5 psi/ft (11 kPa/m) is used for this purpose.

The additional unloading valves required are set at depths determined by the
intersecting of successive kill fluid gradient lines drawn between the design gradient
and the gas design pressure.

Figure 4 (Gaslift design for future development) illustrates a design for the same well,

Figure 4

based on the assumptions that future conditions will include an increase in gas supply
pressure (e.g., installation of new compressor or high pressure gas source) and
increasing water cuts (typical in older fields with active water drive production

45
mechanism). The additional valves that are required for future conditions are
accounted for in many gas-lift installations by including side pocket mandrels holding
dummy valves at the appropriate depths. When conditions change, the dummy valves
are replaced with functioning valves to allow continued efficient operation.

Some gas-lift completions utilize a packer to isolate the annulus from well fluids. In
continuous gas-lift installations, the use of a packer will not prevent the bottomhole
gas injection pressure from being transmitted to the formation and lessening the
drawdown that could be achieved. The production casing must be able to contain the
lift gas pressure without leakage and some companies therefore use premium
couplings for the production casing in gas-lifted wells operated under high lift gas
supply pressures. High pressures are needed in particularly deep wells and where
single point gas lift is used to avoid servicing and leakage problems at multiple
unloading valves.

Servicing of the gas-lift valves is typically needed every 18 to 24 months. If


conventional tubing-mounted gas-lift valves are used, this means a service rig job.
However, wireline-retrievable gas-lift valves are more commonly used today.
Retrievable valves are positioned into a side pocket mandrel by means of a kick-over
tool. Such tools can be either wireline- or TFL-operated.

Electrical Submersible Pump Completions

Electrical submersible pumps have their greatest application in moving large volumes
of low gas-liquid ratio fluids from reservoirs with temperatures below 250 F (390 K).
They are a particularly attractive artificial lift method for water supply wells, high water
cut producers, and high deliverability, undersaturated oil wells. Modified design
procedures and equipment improvements are allowing effective operation at up to
1000 scf/bbl (200 m3/m3). High rate units are best suited to a total head of around
4000 to 5000 ft (1200 to 1500 m). Figure 1 (Submersible pump lift capacity for various
casing sizes) shows that the critical limitation is the casing size, which will determine
the available horsepower and capacity.

46
Figure 1

Submersible pumps are available from a number of manufacturers (TRW-Reda,


Centrilift-Hughes, Westech-Dresser, etc.) all of whom provide design services,
engineering manuals, and customer support. It is usually advantageous to select the
supplier at an early stage and then to work with that supplier in finalizing the design
and equipment selection. To do this effectively and fairly the production engineer
needs to have some appreciation of submersible pump design and electrical
engineering. If a submersible system is selected, it is strongly recommended to include
electrical engineers in the design and operations groups, since this can lead to
considerable improvements in performance and overall costs. Moreover, a total
systems approach to power generation (or purchase), distribution, and utilization can
then be implemented.

47
Most submersible pump installations are simply run on the tubing and landed off above
the perforations or openhole section ( Figure 2 , Components of the electrical
submersible pump system).

Figure 2

The motor is run on bottom so that the produced fluid will dissipate the heat
generated. If the pump has to be set below the perforations, a shroud (or flow tube) is
used to draw the produced fluid past the motor at sufficient velocity to provide cooling.
Bottom discharge pumps are also available for injection wells in powered "dump"
waterfloods.

The cable-suspended pump has been developed to reduce pulling costs in offshore
applications, although results to date have been discouraging. The pump is locked into
a shoe with dogs, while a sealing element supports the fluid head. A safety valve
operated by the pump discharge pressure is often mounted below the pump to provide
a positive shutoff when the pump is pulled and to meet SSSV regulations. The main
problems with the cable-suspended pumps have been failures and shorts in the cable,
pot-head, and wellhead splice. Gas penetration and armor corrosion are the main
causes, since the cable is exposed to direct contact with the produced fluids at high
temperatures.

48
Effective fluid removal and correct choice of the insulation materials for the actual
operating temperatures are often cited as keys to efficient submersible pumping
operations. The clearance past the pump should be small enough for the fluids to flow
by at more than 1 ft/sec (0.3 m/s) but not so small as to generate significant friction.
Operating outside of the design range and, particularly, below the design rate, results
in increased heating of the produced fluid and greater heat transfer to the cable. If this
is coupled with the well being nearly pumped off, with resultant reduced heat transfer
capacity to the formation via the annulus, serious earth leakage problems can be
expected. It is therefore imperative to understand the proper operating conditions and
design for them.

Submersible pump completions should have

open annuli for gas venting (exceptions to this at offshore locations)

special wellheads for sealing around the cable

enough clearance for the pump and cable

adequate cable protection

adequate cooling of the motor

adequately sized tubing for handling the large volumes of produced fluid, with
a minimum of backpressure on the pump, to minimize wasted energy

Hydraulic Pump Completions

There are three types of hydraulic pumping systems: the piston pump, the jet pump,
and the hydraulic turbine.

Hydraulic pumping has a wide variety of applications. The piston pump ( Figure 1 and
Figure 2 ,

49
Figure 1

Schematic of closed power fluid surface system) is particularly attractive for deep
wells.

50
Figure 2

In the moderate to high volume range, the jet pump is used where flowing gas-liquid
ratios are moderate to low.

Hydraulic pumps are well suited to offshore applications, crooked holes, heavy oils,
and variable production conditions. Their major disadvantages are the high initial
investment and the maintenance costs of the high pressure power fluid supply system.
The critical factor determining operating efficiency is the maintenance of a clean,
solids-free power fluid.

In open power fluid (OPF) systems, the spent power fluid is mixed with the produced
fluid in the return line. This reduces the number of conduits required, and therefore the
operating pressures, and permits a higher production rate. The closed power fluid
(CPF) system reduces the cleaning and chemical treatment requirements and costs,
especially when water is used as the power fluid. As a result, equipment maintenance
and metering problems tend to be reduced.

So-called free pumps are run in the power conduit and can be pumped in and out of
the well for servicing by simply reversing the flow. The ease with which the bottomhole
operating assembly can be recovered, inspected, and replaced, and the

51
interchangeability of jet and piston pumps with well conditions, make this type of
hydraulic pump particularly attractive. This option comes at the cost of reduced pump
capacity, and such trade-offs must be considered in the design.

A subsurface safety valve can be mounted under the hydraulic pump operating unit
and controlled by power fluid pressure. This can be recovered for servicing by either
wireline or TFL methods.

From a completion design viewpoint, the critical considerations for hydraulic pumping
are

the number of fluid conduits

pressure losses in the power and return lines

whether the produced fluid can return up the casing

the need for a packer

the need for lubricator access and/or flow loops to pump in the jet or piston
unit

the fact that hydraulic turbine pumps require large casings like electrical
submersible pumps

whether or not an SSSV is needed

Plunger Lift Installation

The plunger lift is a very low liquid rate lift system in which stored gas energy is used
to drive a piston, or plunger, carrying a small slug of liquid, up the tubing when the
well is opened at the surface ( Figure 1 , Typical plunger installation).

52
Figure 1

The well is then closed in and the plunger falls back to the bottom. This lift method is
particularly useful for dewatering low rate gas wells. Plunger lift efficiency decreases
with depth and productivity, but increases with tubing size. Requirements include the
following:

tubing must be carefully drifted and swaged prior to installation

the annulus must be open to store the required lift gas

a nipple or collar stop must be installed to support a catcher and shock


absorber

53
SUBSEA COMPLETIONS
A subsea completion is one which the wellhead and christmas tree are located on the
sea floor. This is in contrast to offshore wells completed on platforms or floating
facilities, where wellheads and trees are above the water. From a completions
standpoint, subsea wells present a number of challenges relating to wellhead
accessibility, fluid production, well maintenance and equipment reliability.

Subsea production began in the 1960s, became a significant part of the offshore
industry during the 1980s, and increased dramatically during the 1990s (Table 1).

Table 1: Subsea Completions by Decade


Decade Completions
1960s 68
1970s 87
1980s 426
1990s 1092
2000s (actual, 2005) 1347

(forecast, 2005) 2147


Source: Quest Offshore Resources, Inc.

Early subsea completions were installed in shallow water, and conventional water
depths remained the norm until the 1990s. The North Sea boasted the greatest
number of subsea completions through the 1990s, followed by Brazil. After 2000, West
Africa and the Gulf of Mexico began to overtake the North Sea in the number of subsea
wells actually completed, and deepwater completions claimed a greater share of the
total. The annual water depth record has passed back and forth between Brazils
Campos Basin and the Gulf of Mexico since 1995. As of 2005, this record stood at a
little more than 7,000 feet (McCabe, 2006: Offshore Production Facilities. IPIMS e-
Learning system).

Subsea wells can be located long distances from a main platform or other host facility,
beyond those attainable by extended-reach drilling. In the North Sea, maximum
distances are 30 miles (48.3 km) for a gas reservoir and 12 miles (19.3 km) for an oil
reservoir. Although most subsea wells have produced by natural flow, artificial lift
methods are becoming more common, along with the use of water injection to
maintain reservoir pressure.

These wells employ a variety of production configurations, including single satellite


wells consisting of wellhead assemblies on individual guide bases; steel template
structures with production manifolds; and clustered well systems (i.e., single satellite
wells connected to a subsea manifold).

Installing subsea completions can be cheaper than drilling highly deviated platform
wells, and more economical than constructing additional surface facilities. They do,
however, have a number of drawbacks and limitations. For example, in long flowlines,
substantial pressure losses can occur between the well and the host facility. Also, the
cooling effects of seawater can result in poor oil flow properties and gas hydrate

54
formation. Most important, well maintenance and workover cost are high because wells
are not easily accessible--rig work requires moving in a floating drilling unit or jackup.
Subsea wells thus require extensive, careful planning to ensure adequate equipment
and operational reliability in the harsh sea floor environment.

For additional information on subsea wells, refer to the IPIMS topic Offshore
Production Facilities, which is found in the Petroleum Engineering discipline under
the Series heading Offshore Operations.

55
COMPLETION PRODUCTIVITY

Sizing the Tubing


It is important to remember that the primary function of the tubing is to provide a
conduit for transportation of hydrocarbons or injection water. Undersizing the tubing is
the most common and costly mistake made by many completion designers. Undersized
tubing will limit the amount of production or injection that can be achieved, or result in
inefficient or unnecessary artificial lift. On the other hand, oversizing the tubing can
also cause liquid hold-up problems and unnecessarily increase well and equipment
costs.

The production of hydrocarbons from the reservoir, through the tubing, gathering
lines, and facilities, to the sales point makes up an integrated system. The
performance of each element in this system is a function of both its own design and
the performance of other elements. Computer simulation techniques, when properly
applied, are enormously valuable tools for analyzing production problems. However,
the application of simple analytical techniques often provides a better understanding of
the situation and may quickly identify the major problem areas. This may be all that is
required, saving both time and cost. If further analysis is needed, however, a data
base and focus have been developed which often reduce the cost and improve the
value of the simulation work.

Pressure maintenance operations are particularly interesting from a systems viewpoint,


especially offshore and in single ownership pools, since the operation is essentially a
closed loop ( Figure 1 and Figure 2 , Production system pressure distribution, including
injection).

56
Figure 1

Thus the injection pump rating is often directly related to the separator pressure,
especially prior to installation of artificial lift.

Figure 2

57
This becomes particularly important in offshore developments where facility and well
cost optimization can heavily influence the overall economics. Obviously, there are
external parameters that limit flexibility (such as the initial reservoir pressure,
formation breakdown pressure, bubble-point pressure, reservoir injectivity and
productivity, and the oil properties). However, there is a definite value in making a
cost-benefit analysis on the effects of various completion practices on operating
conditions, development costs, reservoir performance, and oil yields.

Most production systems, however, have an open-ended dependency on reservoir


performance. We must therefore consider this as the first step in any completion or
artificial lift system design.

Inflow Performance Relationships

Since all producers must flow hydrocarbons through the reservoir at least to the
bottom of the well if not to surface, we define the bottomhole pressure under
producing conditions as the flowing bottomhole pressure (pwf). For a pumping well
this is the pump intake pressure.

The difference between the flowing bottomhole pressure and the average reservoir
pressure is termed the drawdown and determines the production rate. The relationship
between rate and drawdown will vary depending on the flowing conditions and the type
of fluid.

Productivity Index (PI)

In calculating oil well productivity, it is commonly assumed that production is directly


proportional to drawdown. The constant of proportionality is termed the productivity
index, and is commonly denoted as PI or J.

From Darcy's semisteady state flow equation, the PI for a well producing 100% oil is

or, in oilfield units:

where ko = effective permeability to oil (kro k)


h = reservoir thickness

o= oil viscosity

Bo = oil formation volume factor

58
re = effective drainage radius

rw = effective wellbore radius

s = skin factor

These terms typically have the following orders of magnitude:

ko = f (Sw, Sg) (0.7 to 0.3) (k)

o = (API, GOR) 0.5 to 5.0 cp

Bo = f (API, GOR) 1.0 to 2.0

[ln re/rw - 0.75 + S] 10

Since PI relates to the total fluid produced, the magnitude of the PI can change as the water cut
changes. This can be important for sizing artificial lift and treating facilities to handle expected fluid
production after water breakthrough on a flood operation.

We should note from the above equations that the skin (S) is a parameter we can alter
by our completion practices. (We can also increase rw by drilling larger diameter holes
or increase the effective rw by fracturing the well.) Example 1 illustrates the effect of
damage and stimulation.

Effect of Skin on PI

Example of a PI calculation showing effects of

(a) wellbore damage (skin = +5)

(b) fracture stimulation (skin = -5)

(c) a good normal completion (skin = 0)

re = well drilled on 160 acre (64 ha) spacing, 1320 ft (402 m) radius

o = oil viscosity, 1.5 cp (mPa.S)

k = permeability of rock, 500 md

kro = relative permeability at Sw = Swc, 0.8

h = thickness of pay, 50 ft (15 in)

rw = wellbore radius, 0.4 ft (0.12 m)

S = variable

59
= average reservoir pressure, 2900 psi (20,000 kPa)

Bo = oil formation volume factor, 1.3 v/v

Using Equation 2

(a) S = +5 J = 5.88 b/d/psi

(b) S = -5 J = 30.89 b/d/psi

(c) S = 0 J = 9.88 b/d/psi

or in the SI system

(a) S = +5, J = 0.1334 m3/d/kPa = 13.34 m3/d/B

(b) S = -5, J = 0.697 m3/d/kPa = 69.67 m3/d/B

(c) S = 0, J = 0.2238 m3/d/kPa = 22.38 m3/d/B

Effect of damage = (c-a) /c = 40% loss in PI

Reward for stimulation = (c-a) /c = 68% increase in PI

Although in theory the negative skin demonstrates the effects of a fracture stimulation,
in practice it is not possible to achieve an adequate permeability contrast between the
fracture and formation in high permeability zones (>75 md) for this to be achieved.

Oilwell Inflow Performance Relationship (IPR)

The straight line PI relationship should not be expected to hold when two-phase gas
and liquid flow exists in the reservoir. Gilbert (1954) recognized the PI variation with
drawdown and proposed the use of a bottomhole pressure versus producing rate plot
for well analysis. He termed this curve the inflow performance relationship, or IPR, of a
well ( Figure 1 ).

60
Figure 1

Several techniques have been proposed for determining the IPR for a well below the
bubble-point of the oil, where multiphase flow exists. Vogel (1966), using a computer
model of a solution gas drive reservoir, developed a generalized IPR reference curve.
Using this general curve, a specific IPR curve can be constructed for a well knowing
only the static pressure and a flowing bottomhole pressure at one producing rate. For
those who prefer to do all their work on a calculator, at or below the bubble point,
Vogel found that the IPR curve can be approximated by the expression

where qmax = maximum producing rate at Pwf = 0

With a curved IPR it is obviously more difficult to predict the effects of damage and/or
the improvements to be expected from stimulation. However, Standing (1970)
published a modification to the Vogel curve accounting for changes in flow efficiency.

61
Fetkovich (1973) showed--both theoretically and from numerous oil well tests--that oil
well backpressure curves follow an IPR equation of a form commonly used for gas
wells:

The exponent n and intercept J' are usually determined from a multipoint or isochronal

backpressure test, where is plotted against q on log-log paper.

Gas Well Inflow Performance Relationship

The most common method of estimating gas well IPRs is the "backpressure" method of
Rawlins and Schellhardt (1936) where

The well is flowed for a fixed period at different rates. Using the bottomhole flowing pressure at
equal flow times, a plot of

log

versus log qg is prepared. The slope gives a value for 1/n ( Figure 2 , Plot for a
conventional well test example) and using this, C can be calculated. The exponent n
varies from 1.0 for laminar flow to 0.5 for fully turbulent conditions.

62
Figure 2

It is important to remember that this IPR relationship is empirical and that C is a


function of flow time; its value under semisteady state conditions must either be
calculated or determined from an extended flow period. At low rates, where n 1.0, we
may calculate C

or, in the SI system:

The absolute open flow potential (AOF) is defined as the rate corresponding to Pwf = 0. It will be a
function of flow time. Production engineers need to be aware of this and clarify the meaning of
quoted AOF values. Also remember that a value for AOF calculated using flowing tubing pressures
rather than flowing bottomhole pressures is distorted by tubing performance.

63
Another method of determining the IPR for a gas well is to plot

versus q from the generalized semisteady state flow equation

The slope will give a value for F, the non-Darcy or turbulence-dependent coefficient, and the
intercept will give a value for B, the Darcy coefficient. Dake (1978) provides formulas for estimating
B and F from core data or build-up analyses. More correctly, B and F should be calculated from
pseudopressures (m(p)) to be independent of variations in gas viscosity and deviation factor, at
which point they can be used to predict future performance accurately. Theoretically, this method is
still not absolutely correct, but in the majority of cases it is a perfectly adequate description of the
inflow performance. Stimulation of gas wells will affect not only their skin factor (S) and therefore
their Darcy coefficient (C or B) but also the non-Darcy coefficient (n or F).

From a completion engineering viewpoint, the following concepts are fundamental to


proper well design:

the inflow performance of a well is largely determined by reservoir parameters

the skin factor, and the turbulence coefficient in high rate wells, especially gas
wells, are the only parameters we can normally affect by completion efficiency
and stimulation

test results alone may not adequately describe the long-term inflow
performance of a producer unless corrected for

- semisteady state conditions

- curving of the IPR in oil wells below the bubble-point and in gas wells

- expected skin (this is a function of perforation length, perforation


efficiency, stimulation, damage, etc.)

Tubing Performance

Most wells produce under conditions of two- or three-phase flow in the tubing and flow
line. Multiphase flow analysis is complex, and the pressure loss from the bottom of the
tubing to the top is a function of the fluid head, the friction, slippage between phases,
and the flow regime.

These flow parameters, in turn, are affected by the producing conditions:

pipe diameter (d)

flow rate (q)

64
gas liquid ratio (GLR)

water cut (WC)

fluid density ()

fluid viscosity ()

pressure (p)

temperature (T)

While some effects are readily apparent (e.g., pressure losses increase with increasing water cut) ,
others are less obvious (e.g., increasing the GLR initially reduces the pressure loss, but eventually
the trend reverses because of friction at high velocities).

Because of this complexity, empirical and semiempirical analysis techniques have been
used to develop relationships among the producing conditions listed. There are a
number of correlations available as computer programs or as published gradient
curves. For preliminary work, any of these correlations are satisfactory. However, since
they give somewhat different results ( Figure 1 , Comparison of gradient curves), for
more detailed work the engineer should establish a match with field test data and
choose the most appropriate correlation.

65
Figure 1

Usage of the gradient curves is illustrated in Figure 2 (Vertical flowing pressure


gradient curve).

66
Figure 2

The important thing to remember is to enter the curve at a point defined by the rate,
GLP and flowing tubing pressure, or BHP (THP equivalent to 1000 ft in Figure 3 , Effect
of tubing size on productivity of gas-liftedoilwell), and then move along the appropriate
GLR line by an increment equivalent to the depth (i.e., from 1000 to 8000 ft for a
7000-ft deep well).

67
Figure 3

Do not just read the BHP conditions at a given depth this merely corresponds to a
value of 0 THP. The other important considerations are that you use the correct water
cut and adjust the GOR to a GLR:

GLR = (1 - WC) GOR


For deviated wells, it may be necessary to use a computer or to interpolate between true vertical
depth and measured depth by deducting the additional head effects using an average effective
density.

Presentation of the tubing performance (vertical lift performance) data depends on the
problem being addressed. For well design, the most useful presentation of tubing
performance is to plot flowing bottomhole pressure (pwf) versus rate (q) for various
tubing sizes and gas liquid ratios.

For any tubing size there is a minimum flow rate that is required for continuous
removal of the liquids from the well. This is the rollover point in the tubing
performance curves, which is not easily identified without a computer simulation,
although the rule of thumb is a velocity of about 5 ft/s (1.6 m/s). Below this rate the
well will be unstable. This phenomenon is referred to as liquid holdup, and is due to

68
slippage of the gas phase through the liquid. The larger the tubing diameter, the
higher will be the liquid holdup rate.

Matching Completion and Reservoir Performance

Having developed inflow (IPR) and tubing performance curves for a given system, the
final step in determining maximum system productivity is to combine them and
identify the intersection points. This is probably one of the most important production
engineering design functions; not only does the tubing size define the system flow
rate, but it dictates the sizing of all other downhole equipment. While this requirement
is obvious for flowing wells, gas-lift operations, and injection wells, it is often forgotten
when other artificial lift systems are used.

Obviously, tubing size is constrained by the size of the production casing. Therefore,
this type of analysis must be part of the planning for the well's development drilling
phase. It is also important to consider that the production system curve will most likely
change over time, and that we will need to optimize the tubing size over the life of the
well.

To construct a system curve, one typically assumes a well-head or separator pressure


and four or five rates that adequately span the expected productivity as estimated
from the IPR curve. The production target rate and the expected water cut and GLR
behavior will also be constraining factors that must be evaluated.

Figure 1 (Effect of tubing size on productivity of gas-lifted oilwell) illustrates a field


where the operator was investigating apparent poor performance of the gas-lift
system.

69
Figure 1

With a minor amount of effort, it became apparent that the major problem was
undersized tubing. Moreover, poor completion practices had resulted in severe skin
damage. After stimulation and installation of 3 1/2 in (73 mm) tubing this well could
triple its rate.

Figure 2 (Effect of tubing size on gaswell productivity) presents an overlay of the


individual inflow performance curves and a selection of tubing performance curves for
15 wells in a high-deliverability, relatively low-pressure reservoir.

70
Figure 2

The operator wanted to be able to maximize short-term production rate from the
reservoir, while deferring the installation of a compressor for as long as possible. Since
9 5/8 in. and 10 3/4 in. production casing had been installed, very large tubing sizes
could be considered. Note how the performance of the best wells (4, 14, 10) is quite
sensitive to tubing size (e.g., 7 5/8-in. tubing produces 33% more than 5-in. tubing in
well 10), while on the worst well (13) the 5-in. tubing would actually produce slightly
more than the 7 5/8-in. tubing because of reduced liquid holdup.

Figure 3 illustrates the effect of tubing size on a gas injector.

71
Figure 3

Here, the IPR curve is reversed, with increasing injection rates corresponding to
increasing bottomhole pressures. The tubing performance curves display the
bottomhole pressure and rate that correspond to various tubing diameters and a
constant 5000 psi injection pressure at the surface. A curve is also plotted for 5 1/2-in.
tubing and a 6000 psi injection pressure. The increase in injection pressure appears to
have a much greater effect than increasing tubing size.

In each of these displays, it is important to remember that the system definition


determines the intersection of the reservoir and tubing performance curves and thus
the rate and flowing bottomhole pressure. In order to change those values we must
change the system.

Artificial Lift Requirements

Most artificial lift textbooks and manuals emphasize the importance of knowing and
designing for what the well can produce. This is just as important for rod pumping a
stripper well as for the design of a high-volume gas-lift system. In fact, in many rod
pumping installations more work is done in moving the sucker rods than in moving the
oil. This is not a problem, providing the well is achieving its target efficiency. However,

72
in many cases, the amount of fluid produced is curtailed by inadequately sized pump,
tubing, rods, or surface unit.

It is also important to remember that an artificial lift system is only a method of adding
energy to the system; thereafter the produced fluids still have to "flow" out of the well
to the separator, and are therefore subject to the same pressure losses as a flowing
well. Evaluation of these losses is particularly important for high-rate lift systems. In
fact, Kermit Brown (1982) has suggested that to compare different lift methods, quasi-
tubing performance curves should be generated for each lift system ( Figure 1 , Tubing
intake curves for artificial lift systems).

Figure 1

(Note that ESP and beam pump installations can be designed for the same Dp and
therefore the same production rate). Figure 1 also illustrates another important
consideration in designing a well that is, that artificial lift can often be usefully
applied to wells capable of flow to enhance their offtake rate and accelerate income.
The type of artificial lift system selected will affect the tubing and casing size required
in a completion, and vice versa.

When an artificial lift technique is employed, the added operating costs may preclude
attaining the maximum production rate. Some maximum economic rate must be
achieved.

73
Targets and Allowables

Another set of limitations that the completion designer must be aware of are those
imposed on the production rate by regulatory authorities, reservoir control
requirements, facility limitations (e.g., water or gas disposal capacity) , and market
constraints (e.g., gas contracts).

There is obviously no point in spending a lot of money to provide a capacity that


cannot be processed or sold. This argument must, however, be used with care, since it
has been used in the past as an excuse for adopting practices that resulted in wells
being permanently damaged and unable to recover their reserves economically. It is
also important to consider the changes in production conditions likely to occur before
the next major workover (i.e., over five to ten years). For example, will there be
pressure depletion or water breakthrough? How will this affect the well's capability to
meet its target?

Formation Damage Considerations


Formation damage is the term used when we physically cause an additional pressure
drop (i.e., a true skin) in the wellbore area ( Figure 1 , Effect of damage on pressure
transients around a producing well).

Figure 1

The following are some causes of formation damage:

74
invasion of drilling mud solids into the formation (especially into fractures);

drilling mud filtrate invasion into the formation;

cement losses into fractures;

cement filtrate invasion into the formation;

plugged perforations (often due to overbalanced perforating);

inadequate perforations (size, number or penetration);

partial penetration of the producing zone (i.e., not opening the total pay);

crushing and compaction of formation matrix surrounding a perforation;

invasion of solids in completion or workover fluids into the formation or


perforations;

invasion of completion or workover fluids into the formation;

plugging of the formation from the swelling of water-sensitive native clays;

asphaltene or paraffin precipitation in the formation or perforations;

scale precipitation in the formation or perforations;

creation of an emulsion in the formation;

injection of acids or solvents that contain solids or precipitate solids;

sand fill in the wellbore;

injection of an oil-wetting surfactant into the formation;

excessive drawdown that causes movement of formation fines, compaction of


a weak formation, or instigates water production.

From the work of Abrams (1975), Darley (1965) , Tuttle and Bark-man (1974), and Nowak and
Kruegar (1951), et al., it can be concluded that solids entrained in the drilling or completion fluids

are the main cause of impairment. If = mean pore size (microns) of the formation, then the
effects of invading materials and the treatment for their removal can be characterized as follows:

Diameter of Invading
Effect Treatment
Solid

75
Bridging (e.g., drilling mud filter cake) Backflow
>33%

<33% and Shallow invasion (e.g., skin caused by solids in Acidize or


completion fluid) Reperforate
>10%

Probably not harmful


<10%
Two rules of thumb for estimating pore size of sandstones:

25% of mean grain size

Carbonate rocks do not generally develop consistent pore-size relationships, but fortunately pores
are generally quite small (10-20 microns).

Roughly speaking, we may say that unfiltered water (with 2 micron particles) will
damage moderately permeable sandstone formations, (k = 30 to 400 md). Mud filtrate
(with 3 micron particles) will damage good permeability formations, (k = 80 to 900
md). Mud solids (with 9 micron particles) will damage very good formations (k > 725
md). Remember, it is easier to damage good reservoirs; therefore, filtering of well
servicing fluids is essential during completions and workovers.

Injection water must be well filtered. Drilling fluids must be designed for bridging on
the formation face to prevent solids invasion.

Drilling Damage

With modern drilling muds and perforating techniques, drilling damage is not as
serious a concern as it once was in most perforated well completions, because the
perforation tunnel will extend beyond the damaged zone. However, deep mud damage
can be a serious problem and is particularly common:

in high permeability reservoirs, especially in vugular or fractured carbonates, where the


fluid invades the flow channels. Since plugging is very difficult to remove, it is often
necessary to do a clean-up acidization or mini-fracturing treatment. In sandstone, seepage
losses can be prevented by properly sizing the mud solids to form a filter cake;

in low porosity and low permeability reservoirs, where an equal volume of


filtrate has to occupy a greater radius. This damage can be minimized with a
low permeability mud cake and nondamaging filtrate. Although it is usually
necessary to fracture these reservoirs in any event, the probability of damage
should be recognized in the prefracture condition and taken into account in
estimating potential fracture results;

in regions of slow drilling where mud filtrate losses are of necessity much
greater;

76
in lost circulation zones. In sandstone, severe lost circulation is usually the
result of inadvertently fracturing the formation. Since fracture pressures are a
function of pore pressure, this is a particularly severe problem in partially
pressure depleted reservoirs.

Obviously, in openhole, uncemented liner, or gravel-packed completions, drilling damage is a major


concern. Four approaches are common for minimizing drilling damage, and are often used in
combination with one another:
1. Drilling with clean, filtered, clear fluids.

2. Drilling with nondamaging fluids weighted with acid-soluble, lost circulation


materials and weighting materials (e.g., graded limestone).

3. Drilling underbalanced under pressure.

4. Post-drilling stimulation.

In addition to mud and mud filtrate invasion, the other major cause of drilling damage is the effect of
incompatible filtrates on water-sensitive clays, particularly when drilling with fresh water-base muds.
This is a major concern when the formation contains significant amounts of kaolinite, illite, chlorite,
smectite, or montmorillonite clays. Permeability reduction occurs as a result of dispersion or
swelling of the clays. Extensive research has been done on the subject of clay chemistry, but suffice
it to say that once such damage occurs, it is very difficult to remove. Therefore, it is essential to take
preventative action when selecting fluids that will contact a pay zone containing water-sensitive
clays. Low concentrations of potassium or calcium chloride (2%) in completion fluids are very
effective and better than higher concentrations of sodium chloride.

Perforation Damage

Although the perforation process itself always causes some degree of damage due to
the crushing of the surrounding rock by the high pressure jet, the major causes of
perforation damage are:

overbalanced perforating;
solids plugging;
inadequate cleanup;
inadequate shot density (<4 shots per ft);
partial completion.
Figure 2

77
Figure 2

and Figure 3 (Perforation mechanics showing cleanup of plugged perforations after overbalanced
perforation or after a well kill with mud)

78
Figure 3

and Figure 4

79
Figure 4

and Figure 5 (Perforation cleanup mechanics showing cleanup after underbalanced perforating)

80
Figure 5

review the mechanics of perforation cleanup and demonstrate both the inherent superiority of
underbalanced perforating and the problems involved with trying to clean up plugged perforations.
Once one or two perforations begin to flow, the drawdown on the others is reduced to ps.

Figure 6 ,

81
Figure 6

Figure 7 (Effect of perforation density and penetration on productivity) illustrate the


importance of having the perforations penetrate the drilling damage and having
adequate shot density (4-12 shots per ft).

82
Figure 7

In general, within practical limits, an increase in shot density alone cannot overcome
the combined effects of drilling damage and perforating damage. Deeply penetrating
perforations are necessary if the productivity is to approach that of an undamaged,
openhole completion.

Partial penetration of the pay zone (i.e., not perforating all the net pay) results in an
additional pressure drop near the wellbore, often referred to as a geometric skin. This
effect is often overlooked by production engineers, who attempt to acidize away the
"apparent damage." Analytical techniques are available for estimating this effect and
are discussed in reservoir engineering texts such as that of Dake (1978).

Fluids Damage

In addition to the damage caused by entrained solids and the effects of fluids on clay
chemistry, injected fluids can cause damage by incompatibility with the formation
fluids. These incompatibilities manifest themselves in:

precipitates (scale, salt crystallization, etc.);


sludges or emulsions (especially from acid);

83
insoluble residues (from gelling agents);
relative permeability effects (water blockage, wettability reversal, etc.).
Because of these considerations it is important to test properly all fluids that are to be displaced into
the formation. API RP 42 presents a procedure for testing the surface active agents used in well
stimulation. It is important to remember that these tests should be run on the entire treatment that is
to be used because some additives will have an affect on others. If acid is to be used, both the live
and spent acid should be tested.

Sand Fill

Sometimes poor productivity may result from fill covering the perforations or pay
interval. It is therefore advisable to make at least an annual check for sand fill, when
some other well operation is being undertaken (e.g. , wireline runs, pump changes,
workovers) . This is critical after a fracturing job or gravel pack, or in marginal sand
producers. Normally a wireline drift run to check the holdup depth is the cheapest and
easiest technique. In anticipation of this need, the completion designer should provide
easy wire-line access to the bottom of the tubing, consider the hydraulics and
mechanics of possible clean out operations (coiled tubing circulation), and leave an
adequate sump below the bottom perforation.

Selection of Drilling Fluids

There are five key considerations for designing a drilling fluid that will not inhibit well
productivity at the time of initial completion:

1. Minimizing the overbalance.


2. Monitoring the system hydraulics to avoid washouts.
3. Ensuring an effective filter cake.
4. Treating the filtrate to avoid clay problems.
5. Effectively removing fine particles from the mud.
Since damage is often a function of the overbalance between the wellbore and the reservoir, the
density of the drilling fluid should usually be adjusted to maintain only 200 to 500 psi (1 to 3.5 MPa)
overbalance. Underbalance drilling (e.g., with air, gas, clear fluids) may be used in low permeability
reservoirs to avoid damage, but usually has to be specially licensed. This approach is claimed to be
more effective than postdrilling stimulation in some reservoirs. Normally, however, well kicks pose a
much greater risk to safety, overall costs, and project economics than minor drilling damage;
therefore, drilling with a small overbalance is normal procedure. Overbalance is also useful for
controlling shale problems in areas of high tectonic stresses. Table 1, below, lists some of the
typical drilling fluids and the pore pressures they generally offset.

Pore Pressure
Drilling Fluid Densities
Gradient
AIR
At up to 1000 psi
0-0.025 psi/ft
Annular velocity 3000 ft/min
No oil or water influx can be tolerated
FOAM
0.025-0.35 psi/ft Annular velocities 800 to 1300 ft/min
Rule of thumb 200 ft3/min/1000 ft of hole

84
Good fluid loss control
OIL/AERATED WATER/WATER/BRINE
0.35-0.5 psi/ft Annular velocity 100 to 125 ft/min
Fluid loss can be a problem in permeable zones
UNWEIGHTED MUDS
Annular velocities are function of viscosity (100 ft/min)
0.45-0.5 psi/ft
Filter cake necessary to control fluid loss
Control drilled solids buildup to maintain correct density and drilling rate
BARITE-WEIGHTED MUD
Annular velocities are function of viscosity (100 ft/min)
0.5-1 psi/ft Fine drilled solids difficult to remove from mud without also removing barite
Overweighted mud causes slow penetration rate and lost circulation because
of fracturing
HIGH-DENSITY MUDS
0.9-1.6 psi/ft
Use galena or iron compounds in place of barite
Table 1. Pressure gradients of typical drilling fluids

The openhole drilling time through the pay section should be as short as possible to
minimize invasion and shale deterioration. On the other hand, it is essential that a high
penetration rate does not result in poor hole conditions due to hole erosion by the mud
stream in the restricted annulus around the drill collars. Similarly, particular attention
needs to be paid to controlling pipe pulling rates during trips, because of the potential
for damage to the pay section from swab and surge pressures.

Abrams (1977) recommends that at least 5% of the mud solids volume should be
more than one-third the formation median pore size to ensure that mud solids do not
penetrate more than 1 in. (25 mm) into the formation. Silica, calcium carbonate, wax,
and asphalt solids are available in suitable graded ranges to maintain this level.

It is better practice to build a filter cake than to drill with losses, because of the
damaging effect of the drilling fluid and the well control risks. However, if it is planned
to drill without returns, the fluids lost to the formation should be filtered and kept as
clean as possible.

The most cost-effective inhibition of shale sloughing is generally achieved with


additions to the drilling fluid of 5 to 10% NaCl, 1 to 3% CaCl2, or 1 to 3% KCl. Lower
concentrations can be used where the mud has low filter loss properties, although this
may be considered an unwarranted risk.

To minimize solids damage, we must look not only at maintaining a good filter cake
but also at keeping down the solids that can pass through the filter cake with the
filtrate. This is achieved by having the rig equipped with adequate mud cleaning
equipment.

The use of non-damaging fluids should likewise figure into plans for the initial
completion and perforating operations. In some cases, it may be advantageous to plan
stimulation work as part of the initial completion.

85
Perforating

In most wells, the initial completion method involves running and cementing casing,
perforating the casing, and installing the tubing. During perforation and tubing
installation, the formation pressure must be controlled by the completion fluid (i.e.,
overbalance conditions).

The problem with perforating while overbalanced is the potential for formation damage
and plugged perforations. There is also the risk of excessive lost circulation, which
could result in a "kick," with no pipe in the hole. The use of mud as a perforating fluid
reduces these latter risks but increases the chances of poor cleanup and productivity
damage.

Specially formulated perforating fluids (generally brines) should be selected based on


the results of compatibility tests with the water, reservoir fluids, and the formation.
After the brine has been prepared, it should be filtered through two-micron filters to
remove entrained solids, undissolved chemicals, and contaminants. This should be
carried out even though it is planned to add materials subsequently to control fluid
loss. The completion fluid may need to be treated with a bactericide and corrosion
inhibitor.

Even with specially formulated, filtered perforating fluids, there is still a perforation
skin effect inherent in the perforation process itself. Typically, the crushed zone
surrounding a perforation may be only 10-60% of the formation permeability. Three
techniques that have been used to remove the damage resulting from this skin effect
are (1) perforating in acid or using an acid wash in carbonate formations, (2)
backflushing the perforations in sandstones, and (3) perforation washing.

Backsurging involves running DST-type equipment and exposing the sandface to a


sudden high drawdown for a limited period of time. A backsurge volume of at least one
gallon per perforation is required.

Perforation washing is primarily used prior to gravel packing and is a very effective
method of removing damage. The idea is to pump clean fluid into one set of
perforations and out of another, creating a void behind the pipe which will
subsequently be filled with gravel.

Perforating underbalanced is the logical extension of the back-surge concept.


Underbalanced perforating is generally superior to conventional overbalanced
perforating for the following reasons:

all perforations will see a high differential pressure and have an opportunity to
backflow

sensitive formations need not be exposed to completion fluid

cleanup is faster and more effective

the productive potential of marginal zones can be established

86
test results are clear-cut

the safety hazards involved with setting production equipment in a potentially


live well are reduced

Underbalanced perforating can be achieved by

through-tubing perforating

tubing-conveyed perforating with

- mechanical firing using a drop bar system

- pressure firing by either tubing pressure or annulus pressure

- electric firing methods

Completion Fluids Selection and Treatment

Overall, brine is usually the best choice as a completion fluid, if a clear fluid is to be
used. It should be filtered through a two-micron filter to avoid deposition of fines in the
pore throats and flow channels of the formation. In high permeability zones, it is often
necessary to add graded material to the brine to form a filter cake. This material
should be easily removable by backflow, and soluble in acid (CaCO3 chips) or oil
(benzoic acid flakes). Formation brine is often a highly attractive completion fluid
provided solids, precipitates, and fine oil droplets are filtered out. Crude oil is also a
good completion fluid; however it is dangerous, messy, and environmentally unsafe. In
general, minimizing overbalance is a good practice.

Stimulation

The classic solution to maximizing a well's productivity is to stimulate it. However, as


discussed earlier, the basis for selecting stimulation candidates should be a review of
the well's actual and theoretical IPR. Low permeability wells often need fracturing on
initial completion. In low permeability zones, additional poststimulation production can
be significant to the economics; however, the production engineer needs to make
management aware of the true long term potential or else overly optimistic projections
can easily be made.

A few useful rules of thumb in selecting candidates include the following:

within the production targets, the best wells make the best stimulation
candidates, but only if they show evidence of skin damage that can actually be
removed

fractured carbonates nearly always require an acid job to remove drilling


damage

87
there is no use in stimulating a well if the real problem is a lack of reservoir
pressure (i.e., depletion)

inadequate or plugged perforations are often the main cause of damage, and
may need a hydraulic shock to ensure breakdown (alternatively, the zone
should be reperforated)

accurate placement of fluids and fluid diverters is often the key to success

the best candidates should have a low chance of stimulating beyond the
producing zone or into a water zone

In the final analysis the decision must be an economic one. Historically, about 90% of
treatments result in a productivity increase, with acid jobs giving a 10% to 50%
increase. Much greater increases can be realized on specific zones, but this is a
function of local circumstances, particularly the presence of a significant, removable
skin.

The specific stimulation treatment depends on the nature of the problem and the
formation type. There are four main types of treatment:

matrix acidizing, for damage removal

acid fracturing, for low permeability carbonates

propped fracturing, for low permeability sandstones

propped fracturing, ("frac and pack") treatments in moderate to high


permeability reservoirs

These main types of treatments encompass a wide variety of alternative treatment


designs; therefore it is important for the engineer to select the treatment carefully.
Sophisticated formulations and techniques are usually needed only for difficult
problems, and the simplest, cheapest treatment that does not introduce problems is
usually the best.

Special additives other than a corrosion inhibitor need real justification (e.g., an iron-
sequestering agent is needed only if the formation contains iron or if the tubing is
badly corroded). Any additives should be tested for compatibility with the reservoir
fluids.

It is also important to properly design the pumping hydraulics so that matrix acidizing
treatments avoid fracturing the formation and fracturing treatments exceed the
pressure required to part rock.

88
COMPLETION PLANNING

Preplanning

The items on the following list should be considered during the design of any
completion. Some will have greater importance than others in a given situation, but all
should be given thought and, in some cases, significant effort:

production objectives and target rate


reservoir properties (pressure, fluids, drive mechanism, permeability)
inflow performance
tubing size and artificial lift method (if any)
casing and drilling considerations
functional specifications (servicing, anticipated production problems, corrosion tendencies)
regulations and general corporate operations philosophy (SSSV, pressure ratings, master
valves)
special fluid requirements and packer fluid selection
stimulation or sand control requirements
well-killing requirements

The completion design procedure will typically take the following path:

tubing and wellhead design (burst, collapse, tension)


preliminary packer selection
tubing movement force calculations
final packer selection and definition of landing conditions
gravel pack or stimulation design
artificial lift design
preparation of the authorization for expenditures (AFE) or budget
equipment ordering
specification of additional data or logging required
drilling program check
writing of completion program
test program design

Rule of Experience and Analogy

Completion engineers tend to fall into two groups: those who copy the previous or
offset well program, and those who redesign each well from first principles. Both
routes miss out on some very valuable information.

Wells differ significantly, even within a given field, and conditions change. Moreover,
our expertise and our understanding of the field constantly grows as new information
becomes available. It is therefore very valuable to reexamine our assumptions,
philosophies, and standard designs from time to time.

On the other hand, nothing speaks like success and it is patently a waste of time and
effort to try to reinvent the wheel. Standard designs and programs should obviously be
developed and used for extensive development programs. However, it is advisable to

89
reexamine these decisions in the face of new reservoir data and the changing
performance of existing wells.

Engineers should also keep themselves current with developments elsewhere in the
industry by attending local operator and professional meetings, talking to service and
field hands, reading literature, and attending courses when possible. Keep a close
check on competitors' activities, especially where new techniques appear to have been
used. It is important to remember that the production engineer's responsibilities
include

looking for ways to improve revenue generation by increasing production and minimizing
costs
evaluating well and field performance and improving it where possible with workovers,
stimulations, facility improvements, or changes in operating procedure
planning for future production needs and policies (e.g., artificial lift, water disposal, offtake
and injection control)
preparing "post mortems" on drilling and workover activities so as to improve future
performance. All major errors should be highlighted and recommendations made.
reviewing and, where applicable, promoting the application of new equipment and
techniques (and reporting on the results achieved, both good and bad)

Post-Drilling Analysis

Before beginning to design a completion program, the engineer should collect and
review

the location data, especially the ground and derrick floor elevations

the casing and wellhead specifications

the drilling fluid data corresponding to any pay zones (weight, type, losses)

the drilling highlights (TD, plugback depth, problems, drilling time in


production casing, cement tops, well status)

the core data and geological description of the pay zones

the pressure data collected in this well and offsets

the logs and their interpretation, especially the resistivity, density, sonic,
gamma ray, and caliper, remembering to indicate which is the reference log

the data from any DSTs, RFTs, or kicks

the casing leakoff tests

the cementation data on the production casing (pressure, volumes, cement


type)

90
any special completion requirements from the geologist or reservoir engineers
(required separation of zones, expected water or gas coning)

any reservoir fluid analysis

the performance of any offset wells

The engineer should then prepare:


1. A depth versus pressure plot showing mud weights, reservoir pressures, and fracture
pressure ( Figure 1 , Depth-pressure plot). On this he or she should plot the operating and
kill conditions, the completion fluid gradient, and the perforating conditions.

Figure 1

2. A copy of the reference log over the pay zone showing any tests, problems,
fluid contacts, casing points, cement tops, plugback depths, etc.

3. A status diagram of the present condition of the entire well showing the
casing points and cement tops.

91
PRODUCTION REVIEWS
If a new well is not the only well in the reservoir, the completion engineer should
ascertain:

the performance data in the offsets (rate, water cut, pressure


( Figure 1 , Reservoir performance map));

Figure 1

the completions used on offsets;


the production decline curves for the offsets, including pressure, water cut, production, and
GOR;
well test analyses;
the reservoir development plan and drive mechanism (i.e., can depletion or water
breakthrough be expected; are recompletions planned).

Many completion operations are the result of workovers (recompletions), in which case
there is production data on the subject well. The checklists below list issues to consider
for field and well performance studies in workover reviews. Many of the same
questions could be asked in offset reviews when a new completion is being planned.

92
CHECKLIST FOR FIELD STUDIES FOR SELECTING WORKOVER
CANDIDATES

1. Review shut-in wells.

Tabulate location, last production (date, rate, cumulative production, water cut,
GOR/GLR, history of producing problems, workovers, reason well was shut in,
possible causes).
Review options (e.g., repair, abandon, convert for injection).

2. Review wells off for reasons of reservoir control.

Tabulate location, production data, reason well was shut in.


Review decision to leave well shut in.

3. Classify producing wells.

Prepare tabulations according to production and potential: wells producing


<60% target; 60-85% target; 85-120% target; potential >120% target.

4. Review limitations of facility.

Tabulate throughput and design capacity for oil, gas, and water; injection and
water disposal equipment.

5. Review decline curves.

Project group plot.


Project individual well plots
Identify anomalies, steep declines, slow declines (<10% per year), low
indicated reserves, well interference, flush production.

6. Review operating costs.

Breakdown costs according to well operations, production processing, water


disposal, personnel needs.

7. Review reservoir pressure and effects on well performance.

Prepare isobar maps, decline trends.

8. Prepare rate, GOR, and water-cut maps ( Figure 1 ).

Plot cumulative production as a circle that is scaled in proportion to total


reservoir production around each well.

93
REVIEW OF INDIVIDUAL WELLS FOR SELECTION OF WORKOVER
CANDIDATES

1. Is well safe? (If not, what is the minimum that must be done?)

2. Is well producing?

What was expected?


Target?

3. Is wireline work required?

Should a BHP survey or production log be run?


Should a fluid level survey or tubing drift be run?
Should an impression block be run to determine depth and type of obstruction?

4. If performance is poor, is this related to IPR or lift system?

5. If IPR is poor, is this due to low pressure, low permeability, or skin effect?

6. Is skin effect due to well geometry, turbulence (gas wells), or damage?

7. What is required to achieve target or maximum potential?

Reperforation (underbalance, increase shots per foot, per penetration, or per perforated
interval)
Acid job (simple cleanout, matrix, deep)
Fracture job (minifrac <25 ft, conventional frac <500 ft, large frac 500-1000 ft, MHF >1000
ft)
Ability to handle or need to reduce gas or water production
Recompletion to a different zone (uphole, downhole)

8. Is workover the best option?

Is workover economic?
Will it be economic in the future?
Is redrilling a new well another option for recovering reserves?
What is cost of abandonment?

9. If the well is located on an offshore platform or at a remote location, can several


wells be scheduled for workovers simultaneously to distribute rig costs of moving and
to minimize deferral of production?

Sources of Geological and Engineering Data


Table 1 lists the sources of technical data pertaining to field production history and
well performance.

94
Field Data

company reserves report

field development plan

annual performance reviews

field studies

company field books

company field files

operations department
production department
exploration department
field office
Well Data Production
government publications

commercial data banks

monthly reports

decline curves

test reports

well files

operations, production, exploration, drilling departments

log data

well files
log data bank
log analysts

core data

description and geological analysis


engineering analysis
special core tests

95
test data

DSTs
production tests
monthly tests

status diagram

Table 1: Sources of geological and engineering data

Well Logs

The critical logs over the pay from a production engineering viewpoint are, in order of
priority

the reference log


the gamma ray-casing collar log
the porosity log (if it is not the reference log)
the cement bond log
the resistivity log
other cased-hole logs
other openhole logs
The engineer should maintain a marked-up section of the reference log in a personal work file.

Status Diagram

A good production engineer should have a book of status diagrams for each well under
his or her supervision. This should be a schematic of the well completion, showing

casing sizes and depths


total depth, plugback depth, holdup depth
tubing and accessories with ID and ODs
perforations (previous and existing)
known casing damage
junk
packers and their specifications
kelly bushing elevation

Each schematic should be dated and initialed. These should be included as a part of every well
report or proposal and automatically updated with any new information from wireline work, well
servicing, major workovers, or production log results.

Regulatory Constraints

Offshore Constraints

96
In most countries, offshore operations are subject to a much stricter set of regulations
than those governing land operations. These often have specific requirements that
govern how wells can be completed, and the production engineer must become familiar
with them. Common requirements include the following:

the design assumptions and minimum design factors for the tubular goods used in a well

the use of a single-zone or segregated multizone completion where different


pressure and inflow characteristics of two zones could detrimentally affect
recovery as a result of interflow

the use of a packer with tubing flow unless specially licensed

an uphole backpressure valve or tubing plug installed and tested prior to


removal of the master valve or valves

before pulling tubing from a hole, both the tubing and tubing-casing annulus
must be circulated full of a fluid that will overbalance the formation pressure

the packer must be set within a specified distance of the top of the uppermost
interval to be perforated

any offshore well capable of flowing must be equipped with a subsurface


safety valve installed at least 100 ft (30 m) below the seafloor

the specifications, design, installation, operation, and testing of subsurface


safety valves shall be in accordance with the relevant section of the American
Petroleum Institute Specifications for Subsurface Safety Valves API Spec 14A
(9th edition, Dec. 1, 1994) and American Petroleum Institute Recommended
Practice for Design, Installation and Operation of Subsurface Safety Valve
Systems, (API RP 14B 4th edition, July 1, 1994)

no operator shall produce a well on which the subsurface valve is not in good
working order

all production casing strings shall limit combined stress to less than 80% of
casing yield

the production casing shall be cemented in a manner that provides for good
cement bonding across the pay zone to a minimum of 200 ft (60 m) above and
to 100 ft (30 m) below the zone or the casing shoe, whichever is the lesser

every operator shall ensure that the wellhead and related equipment in a
completed well has a working pressure that is greater than the initial reservoir
pressure in any productive interval

approval must be obtained for certain unusual operations and programs prior
to their implementation

97
Onshore Constraints

Onshore regulations are usually less stringent, but often specify

approval of programs and licensing of certain operations (especially in sour wells)

minimum casing and cementing requirements

the conditions under which a packer must be installed (e.g., sour wells, salt
water disposal wells, high pressure wells)

the conditions under which tubing may be omitted or casing flow used (sweet
gas, freshwater injection, low pressure oil production)

special requirements for sour wells (e.g., must conform to NACE


specifications; must include a packer, two master valves, and sometimes a
SSSV)

the BOP requirements

Programming and Supervision

Writing a Procedure

The quality of a completion, test, or workover depends not only upon the equipment
and techniques selected but also upon the experience and attitudes of the crew and on
the guidance and supervision provided. A clearly written, easy-to-use completion
procedure is invaluable as a guide to the completion program. A good procedure will
detail exactly how a job is to be done, anticipating problems and providing
contingencies. The engineer must know enough about well operations to organize his
or her thinking properly within the physical constraints of the well and must be
cognizant of what is feasible and safe. Rig time is usually the most costly item in a
completion and careful planning can avoid inefficiency.

In writing a procedure, operations can conveniently be divided into six parts:

1. preparation procedure (rigging up, testing equipment and BOPs, etc.)

2. operating policy and procedures (safety, well control, achieving objectives,


gathering data, maximizing productivity)

3. flanging procedures (removing BOPs, installing and testing wellhead and


surface equipment)

4. stimulation procedure (a complete program on its own)

5. kickoff and cleanup operations

98
6. initial well testing (to assess success of job or if required for reservoir or
production purposes)

It is often useful when developing the procedure to put the program aside for a short while and
come back to it to see if it still hangs together, or to have another engineer read it over.

Each procedure should include these particulars:

Location
The field, well name and identifier, access constraints, and timing constraints (lease
expiration, etc.).

Objectives
The data to be collected, action to be taken, results to be expected, and
economic justification.

Well-Status Diagram
The current and final status, and the depth pressure plot.

Logs of Pay Zone


The resistivity, porosity, GR/CCL, CBL/ VDL, and any other logs. (Mark-up
present and future perforations, look for contacts, etc., be sure to identify
reference log.)

Outline of Procedure
The steps that will maximize productivity, the steps that will maximize data.

Requirements
The surveys or logs, tubular goods, completion equipment, wellhead equipment,
artificial lift equipment, kickoff method, special designs for stimulation, cement
job and artificial lift, and special tools.

Potential Problems
Any formation damage, well control (killing, circulation losses, etc.).

AFE (Authorization for Expenditure)


The cost estimate, approval level, and budget.

Cost Estimate

Most operating agreements call for partner approval of expenditures in excess of a


specified level, except in emergencies. This means that AFEs need to be prepared in
plenty of time. All engineers should familiarize themselves with the AFE procedure and
forms for their respective companies. It is important to note the level of internal
approval that will be needed, since this will dictate how long the approval process can
be expected to take.

It is extremely useful to develop an AFE estimate form as a memory aid to help break
down the costs into their component parts (i.e., rig, services, tubing, pumps,
chemicals, gravel, etc.). This is not only useful for the engineer to make sure that

99
nothing is forgotten, but also to accounting and administration for tax and incentive
purposes.

The contingency cost should be varied according to the realism of the cost estimate. If
it is impossible to tell if a job will take six or fourteen days, build the estimate on the
expectation of ten days and increase the contingency provision.

Explanatory notes will help management in their review process and speed up the
overall approval.

In preparing cost estimates from suppliers' bids and quotes, it is important to build in
provision for

vendor bias
company supplied items and supervision
transportation and installation
commissioning
variations in exchange rate
differences between the actual job and that bid (i.e., last minute change of orders).
These should not be part of the contingency cost since you can reasonably expect to spend this
money even if the job goes as planned.

Contingency cost provisions need to include such items as weather downtime, standby
time (e.g., waiting on orders or supplies), unidentified problems, and uncertainties.
Contingency percentages therefore should not necessarily be standard, but may
change from job to job. A minimum of 10% and more commonly 20% is
recommended, but your company most certainly has its own policy.

Scheduling Equipment

A major cost saving can be made by making sure that major rental equipment is not
called out until it is needed and that downtime is avoided. A chart of planned
operations is invaluable in this respect. It also is important to recognize that there is a
specific lead time involved in getting some equipment. In fact, certain items, such as
custom-designed subsea trees, for example, have long lead times for manufacturing.
Even when items are immediately available, purchasing departments can often get a
better deal if the material is prepurchased in a bulk order. Budget estimates should
therefore include equipment identification.

Wellsite Supervision

While the direct responsibility for the overall operation and safety generally lies with
the operations group, the production engineer remains responsible for the technical
and economic success of the operation. He or she should therefore be on site at least
for key operations such as perforating, production logging, stimulation, and major well
tests.

Needless to say, it is essential that all operations and events be reported in detail and
that tally sheets and diagrams of all downhole equipment are maintained not only in
the field and district office files but also in the central production files.

100
Unless crews are experienced with running a standard completion, tool specialists from
the equipment supply companies should be on site to assist in makeup, testing, and
installation of their equipment and/or pumping of their treatment fluids.

Office Supervision

Even when on-site engineering presence is not considered necessary, the field
engineer must keep abreast of the well operations in his or her area of responsibility.
For complex jobs, the engineer should maintain a diagram of the current well status
and a depth-pressure plot showing pressure and fluid information from the well (e.g.,
reservoir pressure, loss conditions, fracture or pressure test data).

In particular, the engineer should review the costs reported and compare these to the
AFE. If a problem occurs or an overrun is apparent, a review should be made of the
revised economics and a recommendation made to management of action to be taken.
The best way to monitor costs is by graphing the predicted and actual cumulative time
and cost both for the entire job and to specific milestones.

101
CEMENTING

INTRODUCTION

History and Overview

The basic principle of oil well cementing involves displacing cement slurry down the
casing to a predetermined point in the well. The slurry is formed by mixing water with
Portland cement, or with cement blended with additives. This procedure controls
gas/oil and water/oil ratios, and is used in various types of liner jobs and remedial
work. The casing must be cemented to exclude water and other unwanted fluids.
Cement slurry is forced into the annular space between the casing and the wall of the
hole, where the cement can set and form a permanent barrier against water and other
fluids.

Protection and Prevention

Protection of the well and prevention of possible value loss are two vital considerations
in cementing.

Protection Against Pressure, Corrosion, and Shock Loading; Maintenance of Casing


Support

Cement that is pumped down the casing and into the annulus between the casing and
wellbore is used as a sealant to help protect:

casing and wellbore from external pressure that could collapse the pipe or cause a blowout
oil- and gas-producing strata from extraneous fluids
casing from possible corrosion and electrolysis caused by formation waters and physical
contact with various strata
drillstring against loss in a key-seat or sticky hole
downhole production and drilling equipment
pipe from the stresses of formation movement

Cement also bonds the pipe to the formation for support and minimizes the danger of
blowouts from high-pressure zones.

Prevention of Fluid Migration, Lost Circulation, Pollution, and Blowouts

Lost-circulation cement, placed at critical points behind the casing and in open hole, is
also used to help prevent:

unwanted migration of fluids from one formation to another


pollution of valuable oil or gas zones
lost circulation, by sealing off lost-circulating zones and other potentially troublesome
formations as a prelude to deeper drilling

102
contamination of freshwater zones that may be used for domestic supply; protect other
formation strata such as coal, potash, etc.
blowouts from high-pressure gas zones behind the casing
cave-in of the hole during drilling
unscrewing of the bottom joints in surface and intermediate strings

103
MANUFACTURE & CLASSIFICATION OF OILWELL CEMENTS

Cement Manufacture and Chemistry


Cement is manufactured with materials and methods that have changed little over the
last 160 years. Joseph Aspdin, a builder from Leeds, UK, was granted a patent in 1924
for "a cement of superior quality resembling Portland stone." Asp-din's cement was
prepared by sintering fixed proportions of calcareous materials (limestone, chalk,
seashells, etc.) with aluminosilicates (clays) in a kiln at temperatures of 1425 to 15350
C. The resulting material, which is called clinker, is then cooled and ground with
gypsum to form Portland cement. Gypsum is added to prevent flash set. In addition to
the basic raw materials cited above, other materials such as sand, bauxite, iron oxide,
etc., may be used in the kiln feed to adjust the elemental ratios in the resulting clinker.

Portland Cement Components

The principal components of common Portland cement are

50% tricalcium silicate/C3S* (3CaO SiO2)

25% dicalcium silicate/C2S* (2CaO SiO2)

10% tricalcium aluminate/C3A* (3CaO Al2O3)

10% tetracalcium aluminoferrite/C4AF* (4CaO Al2O3 Fe2O3) 5% other


oxides

* C3S, C2S, C3A, and C4AF are commonly used abbreviated notations for the cement
components.
The relative amounts of the above compounds may be varied depending on the intended
application of the cement. The component having the greatest effect on the overall strength of
Portland cement is C3S; it is also responsible for the early strength (1 to 28 days) of the set mass.
C3S is the slow-reacting component that accounts for the gradual gain in strength of the cement
which occurs over an extended period of time. Of the above components, C3A shows the fastest
rate of hydration. The initial set and thickening time of cement, as well as the sensitivity of the set
cement to sulfate-containing waters, are influenced by the concentration of this component. C4AF is
the low heat of hydration compound in cement; the addition of Fe2O3 to the kiln feet favors the
production of C4AF at the expense of C3A in the clinker. This procedure is commonly followed in the
production of medium sulfate resistant (MSR 3 to 8% C3A) and high sulfate resistant (HSR up
to 3% C3A) cements. A description of the manufacture and composition of cement may be found on
pages 3 to 5 of the Halliburton Cementing Tables.

Hydration of Cement

Cement is composed principally of a blend of anhydrous metallic oxides. The addition


of water to this material converts these compounds to their hydrated form. After a
period of time, the hydrates form an interlocking crystalline structure which is
responsible for the set cement's strength and impermeability. The formation of this
structure, as the cement hydrates, is shown schematically and actually (by scanning

104
electron micrograph) in Figure 1 (The setting of portland cement ).

Figure 1

Water in excess of that required for cement to set is necessary for formation of slurry
(a suspension of solids in a continuous liquid medium) and gives the slurry its required
mobility. Removal of the excess water by filtration results in loss of slurry mobility and
in deposition of the wetted cement solids as filter cake.

The amount of water required for a given slurry depends on the type of cement and
the amount and type of additives used. For example, the water requirements for API
class "G" cement slurry, containing varying amounts of bentonite are as follows:

% bentonite water requirement in U.S. gals/94 lb. Sack

0 5.0

2 6.3

4 7.6

If the amount of water is correct for a given system, the slurry will have acceptable
fluidity and will not show excessive solids-settling or free-water breakout. Water
requirements for cement and most of the commonly used additives can be found in
service company publications (e.g. Halliburton cementing tables (1981)).

Temperature Effects

105
As is true of most chemical reactions, the hydration of cement is accelerated by
increased temperature. Figure 2 (The effect of temperature and pressure on Portland
cement thickening time ( after Bearden, 1959)) shows the effect of temperature on the
thickening time of various cement slurries. The reaction of cement with water begins
when they are first mixed.

Figure 2

As the reaction proceeds, but before the system achieves sufficient set to have
measurable strength, the slurry shows an increase in viscosity. When the slurry
reaches a viscosity equivalent to 100 poise, it is said to be no longer pumpable. The
length of time required to achieve this viscosity under a given set of conditions is
referred to as the thickening time.

API Cement Types


Service company literature, such as Halliburton Cementing Tables , describes the
various API cement types. Section 2 of API Spec 10 describes not only the various API
cement types but the chemical ( API Table 2 .1) and physical requirements ( API Table
2 .2) for each. Note that not all countries use the API classification system. RHC (Rapid

106
Hardening Portland Cement) and HSC (High Strength Portland Cement), for example,
ar two non-U.S. cement classifications that correspond roughly to API class C cement.

The API classifications and their general applications are summarized as follows: for
operations outside the United States, the predominant API Classes are A and G (Smith,
1987).

API Class A cement also referred to in the United States as "construction cement" or
"common cement" may normally be used, mixed neat, from surface to a depth of
6000 ft (1830 m) if no special properties are required. This cement is widely available
and its performance can be modified by the use of additives (see the Halliburton
Cementing Tables, Section 230, Blue Subsection II). Additive response and product
uniformity of Class A cement are generally inferior to those of

API Class G cement.

API class B cement is intended for use from surface to 6,000 ft depth, where conditions
require moderate to high sulfate resistance, while API class C is used this same depth
range where early high strength is needed.

API Class D cement is intended for use from 10,000 to 14, 000 ft. depth, under
conditions for moderate to high pressure and temperature. It is available in both
moderately and highly sulfate - resistant types.

API Class E is intended for use from 10,000 to 14,000 ft. depth, for conditions of high
temperature and pressure, and is available in both moderately and highly sulfate
resistant types.

API Class F is for use in the 10,000 to 16,000 ft. depth range, under conditions of
extremely high temperatures and pressures, and is available in both moderately and
highly sulfate -resistant types.

API Class G is intended for use in neat slurries from the surface to a depth of 8000 ft
(2438 m). This cement is available in both moderately sulfate-resistant and highly
sulfate-resistant forms and may be modified by additives for use in a wide range of
well depths pressures and temperatures. Specific formulations and resulting properties
of Class G slurries are found in section 230 of the Halliburton Cementing Tables,
(Subsection IV ).

API Class H cement is similar in form and application to class G, but is usually more
coarsely ground.

Most oil well-cementing operations worldwide use slurries based on API Classes A, G,
H, and C (high-early-strength cement). API Classes D, E, and F are intended for high-
temperature applications, and each contains an organic set retarder added by the
manufacturer. Classes D, E, and F are not widely distributed, and in some instances
the set retarder is not compatible with materials added on-site.

API Class J cement is specifically intended for high-temperature applications; it does


not set at BHSTs less than 230 F (110 C), and does not require the addition of silica

107
to protect the set mass from high-temperature degradation. Class J is not widely
available.

Specialty Cements
These include such materials as these cements: pozzolan , resin or plastic , gypsum ,
diesel oil, expanding, calcium aluminate, latex, and permafrost. Although they do not
comprise the bulk of cements used in the petroleum industry, they have certain
properties that are useful in various applications. These are summarized as follows:

Specialty Cement Type Application

Pozzolan High temp.; low weight requirements

Resin or Plastic Selective plugging/squeezing operations

Gypsum Temporary remedial work requiring rapid setting, high strength

Diesel Oil Control of water zones, lost circulation

Expanding Special downhole conditions

Calcium Aluminate Extreme high temperature (i.e., in-situ combination wells)

Latex Improved bonding/filtration control requirements

Permafrost Frozen environments

Formation of Cement Slurries


Cement slurries are usually prepared in the field with a jet mixer. The components and
operation of this continuous-mixing system are shown in Figure 1 (Jet mixing of oilwell
cement slurries ) Jet mixers are simple, dependable, and are not volume-limited.

108
Figure 1

Their principal disadvantage is an inability to closely control slurry density. Various


types of batch mixers utilizing either mechanical agitation or circulation are available
for mixing up to 100 bbl (16 m3) of cement slurry. This type of mixer can control
density within a range of 0.2 lb/gal. For some operations it is useful to prepare the
slurry with a jet mixer that discharges into a batch mixer, where final density
adjustments may be made before the slurry is pumped into the well. Recirculating-type
jet mixers can also prepare slurries with a more uniform density.

Neat Slurries

A suspension of cement in water, containing no other components, is called a neat


cement slurry. From the standpoints of ease of mixing and of economy, it would be
desirable to use such slurries extensively in oil well cementing. This is usually not
possible, however, since control of the properties of a neat slurry can be achieved only
by slight changes in liquid/solids ratio and by choice of API cement type. Table 1 ,
below, gives the preferred water/cement ratio for the various API cement types, as
well as the resulting slurry densities.

API Cement Type Recommended Water/Cement Slurry Density (lb/gal)


Ratio (gal/sack)

109
A&B 5.2 15.6

C 6.3 14.8

D&F 4.3 16.4

G 5.0 15.8

H 4.3 16.4

J 4.9 15.4

Table 1. Neat cement slurries

The range of densities of these slurries results from the different degrees of fineness of
the various types of API cement, which affect water requirement. Of the commonly
used API types, Class C is the most finely ground, Classes A and G are intermediate,
and Class H is the most coarsely ground. Slurries prepared with the finer grinds tend to
show slightly lower cement-slurry-filtration rates. The thickening time of neat cement
slurries may be estimated from the thickening-time requirements for the various API
cement classes given in Section 8, Table 8 .3, Schedules 1, 4, 5, 6, 8, and 9 of API
Spec 10. API Classes A and C require a minimum thickening time of 90 minutes when
tested according to Schedule 1 (final temperature 80 F [27 C]) and Schedule 4 (final
temperature 113 F [45 C]). API Classes G and H must indicate a thickening time
between 90 and 120 minutes when tested according to Schedule 5 (final temperature
125 F [52 C]). This information is useful for comparing the relative reactivity of the
various API classes; however, a thickening-time test under simulated well conditions is
an important part of most cementing jobs.

For most applications, sufficient control over slurry density, filtration rate, and
thickening time cannot be obtained simply by choosing the appropriate API cement
class; additives must be used to achieve slurry-property modification. Choice of the
proper API cement type (when a choice is possible) can make modification by the use
of additives easier.

Additive-Containing Slurries

Additives are necessary for the use of most oil well cements. This is not surprising If
one considers that Portland-type cements are designed primarily for construction
applications at surface conditions of temperature and pressure. For simplicity and
economy, it is good practice to minimize the number of additives, thereby facilitating
mixing and reducing the chances of undesirable interactions. Also, it is important to
utilize optimum concentrations of additives for each situation; for this reason, slurry
should be tested using on-site components and realistic procedures. The effects of a
variety of additives on the properties of oil well cement, according to Dwight K. Smith,
are given in Table 2 (Effects of additives on cement properties )

110
Table 2

Alteration of Rate of Set

Additives are used extensively to alter the rate of reaction between cement and mix
water. Acceleration of the reaction rate decreases the thickening time of the slurry and
thus the WOC time (time required for the cement to attain a compressive strength of
500 psi [3450 kPa]). Retardation of the rate increases both thickening time and WOC
time. At shallow well depths temperatures are relatively low, pumping times are short,
and it is often necessary to accelerate the set of the cement to reduce WOC time. This
must be done within the constraint of retaining adequate thickening times. For deep,
hot wells slurries are usually retarded to attain sufficient thickening time (plus a safety
factor), but over-retardation must be avoided or WOC time will be unduly prolonged.

111
CEMENT TESTING PROCEDURES

Cement-Testing Procedures
API Spec 10 covers in detail the laboratory methods used to perform API cement tests.
Properties that are covered under the API test procedures include soundness and
fineness, free water content (i.e., filtration rate), compressive strength and thickening
time. Tests for these physical properties involve both chemical and physical analysis.

Some general observations about cement-slurry testing are worthy of mention here:

Testing serves both as a tool for pre-job quality control and as a means of post-job
evaluation and/or troubleshooting, as well as future job planning.
API specifications do not cover all the properties of cements over the wide range of
conditions experienced in the field.
The do, however, list properties for various classes of cements that will fit most well
conditions (Smith,1987).
Cement characteristics vary from one manufacturer to another, and can vary over time from
the same manufacturer. Also, mix-water chemistry can have a significant influence on
cement properties. The information in service company tables or API bulletins is helpful in a
general way, but most jobs require testing samples of on-site components.
If cement-slurry components are sent to a service company laboratory for testing, the
quantity and quality of the samples should be adequate. At least 5 kg of dry cement is
needed to perform the usual set of tests. The cement should be placed in a clean, dry,
airtight metal or plastic container. If a mix-water sample is sent, at least 4 liters should be
placed in a plastic or glass container; glass should be carefully packed to prevent breakage.
Metal is not used because of the reaction between the water and the container.

Fluid-Loss Test

The rate at which a cement slurry loses the water required for its fluidity through a
permeable barrier is called filtration rate or fluid-loss rate. Appendix F of API Spec 10
describes procedures for performing both the low-pressure (100 psi [689 kPa]/room
temperature) and the well-simulation (1000 psi [6890 kPa]/simulated BHST) fluid-loss
tests. The API high-pressure test (performed according to Paragraph F.2 of API Spec
10) has been found to produce more useful information than the API low-
pressure/room-temperature test. Data from the low-pressure/room-temperature test
should be used with caution. The only advantage of the low-pressure/ room-
temperature test is that it can be performed in the field; because the high-pressure
test requires a consistometer, it must be done in a cement laboratory.

Operating Compressive-Strength Tests

Cement-strength data, described in Appendix D of API Spec 10, is useful for


establishing WOC time and monitoring the stability of the set material. Several authors
have addressed the question of what degree of cement strength is required to allow
the safe resumption of operations. Although certain general guidelines have been

112
established with repsct to cement compressive strength, it is best to refer to specific
government regulations, as they may vary among different locations.

The final curing temperatures found in Table D.1 of API Spec 10 ("Well-Simulation Test
Schedules for Curing Strength Specimens") are intended to simulate BHSTs, whereas
comparable temperatures for thickening-time tests (Section 8 and Appendix E of API
Spec 10) simulate bottomhole circulation temperatures (BHCT).

Operating Thickening-Time Tests


The thickening time of a slurry when tested according to simulated well conditions is
one of the more important properties of an oil well cementing slurry. Insufficient
thickening time can result in cement deposits in the casing, tubing, or drillpipe, and
has economically disastrous consequences. Excessively long thickening time
necessitates unduly long waiting on cement (WOC) time, and in extreme cases of over-
retardation can result in cement that never sets. By accepted practice, cement-slurry
thickening time (when tested realistically) is the estimated job time plus a safety factor
usually one hour. Estimated job times may vary considerably, however, and one
hour may not necessarily be a good safety-factor general guideline.

Application Schedule Pages in API Spec 10


Numbers

Specification testing 1-9 29-31

Casing cementing 1g-11g 52-57

Liner cementing 22-32 58-61

Squeeze cementing and 12.21 62-70


plug setting

Table 1: API Thickening-time schedules.

Depth - Ft Factor Depth - Ft Factor

1,000 1.020 11,000 1.160

2,000 1.035 12,000 1.165

3,000 1.050 13,000 1.160

4,000 1.070 14,000 1.150

5,000 1.085 15,000 1.140

6,000 1.100 16,000 1.125

7,000 1.115 17,000 1.105

113
8,000 1.130 18,000 1.1085

9,000 1.140 19,000 1.060

10,000 1.155 20,000 1.030

Table 2: Factors for converting log temperature to bottomhole static


temperature. (From Halliburton).

Section 8 and Appendix E of API Spec 10 describe accepted equipment and laboratory
procedures used for oil well cement-thickening-time testing, including several
categories of time/ pressure/temperature schedules. Section 8 describes specification
tests, and Appendix E lists various operating thickening-time tests. The proper
application of these testing schedules is summarized in Table 1.

Selection of an appropriate liner cementing (No. 22-32) or casing cementing (No. 1g-
11g) thickening-time schedule depends on a knowledge of bottomhole static
temperature (BHST), so the temperature gradient in F/100 ft of vertical depth should
be calculated. If the BHST is unknown, it may be estimated by use of the empirical
relationship between logging and static temperatures in Table 2. The correct
application of the API squeeze, liner, or casing thickening-time schedules (or
interpolations thereof) produces reliable data for establishing retarder requirements. In
some cases, where unusual geothermal gradients are encountered (outside the 0.9 to
1.9 F/100 ft range) or for cementing through a long riser in very deep water (>1000
ft or 305 m), the use of a special time/pressure/temperature schedule may be
required.

Determination of Slurry Density

Cement-slurry density measurements can and should be used in the field to monitor
cement slurry-quality. Appendix C of API Spec 10 describes (with reference to API RP
13B) the accepted procedure for making this measurement. The pressurized-fluid
density balance described in Paragraph C.5 of API Spec 10 should be employed during
density measurement of a slurry containing entrapped air (e.g., a salt-containing
slurry).

114
CEMENT ADDITIVES

Cement Additives

Almost all cement used in oil and gas wells is Portland cement. However, neat cement
is seldom used throughout a job since various additives are usually necessary to
modify the properties of either slurry or set cement. With basic cements (API Class G
or H) and the use of additives, cement slurries can be tailored for any specific
requirement.

Most additives in current use are free-flowing powders that are dry-blended with the
cement prior to its transportation to the well. When necessary, some powdered
additives can be dispersed in mixing water at the site. Liquid additives are more
commonly used offshore and in remote land locations where dry cement blending and
storage are a problem.

Properties that are modified by additives are shown below:

For the slurry:

thickening time (acceleration, retardation)


density (extenders, weight increase/reduction)
friction during pumping
fluid loss (by filtrate)
lost-circulation resistance (whole slurry loss)

For set cement:

compressive strength
strength retrogression (loss with time)
expansion/contraction

Accelerators
Cement-setting time is accelerated to reduce WOC time and to increase early strength.
This is desirable for surface pipe, in shallow (cooler) wells, and for setting plugs.

General pressure recommendations are as follows:

pipe support and zonal isolation 100 psi (690 kPa)

drilling out 500 psi (3450 kPa)

perforating

115
bullets 500 psi (3450 kPa)

hollow carrier or expendable jets 2000 psi (13,800 kPa)

whipstock plug 2500 psi (17,200 kPa) or greater (or harder than formation)
The most common accelerators are calcium chloride, sodium silicate, sodium chloride (low
concentrations), seawater, hemihydrate forms of gypsum, and ammonium chloride. Table 1 shows
typical amounts used per sack.

Type Amount used per sack (% by


weight)

Accelerators

Calcium Chloride (CaCl2) (flake, powder, an hydrous) 2-4

Sodium Chloride (NaCl) 3 - 10 (water)

1.5 - 5 (cement)

Hemihydrate forms of 20 - 100

Gypsum (plaster of Paris)

Sodium Silicate (Na2SiO3)w 1 - 7.5

Cements with dispersants and reduced water 0.5 - 1.0

Sea Water

Retarders

Calcium-Sodium Lignosulfonate 0.1 - 1.0

Calcium Lignosulfonate 0.1 to 1.0

Calcium Lignosulfonate 0.1 to 2.5

plus organic acid

CMHEC 0.1 to 1.5

Saturated Salt 15 to 17 lb/sk


Table 1. Commonly used accelerators and retarders. (Source: Halliburton Services).
Retarders
Cement-thickening time is slowed primarily to allow the slurry to be pumped and
displaced into position before setting. Retarder additives include calcium
lignosulfonate, organic blends, carboxymethylhydroxyethyl cellulose (CMHEC), borax,
sodium chloride (in high concentrations), and most fluid-loss agents
(see Table 1).Thickening time is a function of both temperature and pressure, and

116
these effects must be predicted before additives are selected

Type Amount used per sack (% by weight)

Accelerators

Calcium Chloride (CaCl2) (flake, powder, an 2-4


hydrous)

Sodium Chloride (NaCl) 3 - 10 (water)

1.5 - 5 (cement)

Hemihydrate forms of Gypsum (plaster of 20 - 100


Paris)

Sodium Silicate (Na2SiO3) 1 - 7.5

Cements with dispersants and reduced water 0.5 - 1.0

Sea Water

Retarders

Calcium-Sodium Lignosulfonate 0.1 - 1.0

Calcium Lignosulfonate 0.1 to 1.0

Calcium Lignosulfonate plus organic acid 0.1 to 2.5

CMHEC 0.1 to 1.5

Saturated Salt 15 to 17 lb/sk

Table 1: Commonly used accelerators and retarders. (Source: Halliburton


Services).

Thickening time can also be shortened by interruption of pumping (loss of agitation).


API tests may be done in this manner to simulate actual interruptions during
squeezing.

An increase in water volume increases the thickening time of unretarded cement


(Classes A, C, G, and H). With retarded cements (Classes D, E, and F), however,
increased water or solids may decrease thickening time by reducing the concentration
of retarder.

The thickening time of a slurry under realistic conditions must be established to ensure
adequate pumping time for slurry placement.

Excessive thickening time must be avoided to prevent:

117
delays in resuming drilling operations
settling and separation of slurry components
formation of free-water pockets
loss of hydrostatic head and gas cutting

Density-Reducing Additives
Slurry density may be reduced with extenders such as bentonite, pozzolan,
diatomaceous earth, and anhydrous sodium metasilicate. Table 1 shows typical
additive concentrations.

Low-density slurry is frequently preferred, to decrease the likelihood of breaking down


the formation and causing lost circulation. In addition, low-density slurries cost less
per cubic foot because yield per sack is increased.

Density decrease results in large part from increased water content. Extenders, with
their high surface area to "tie up" water, permit water addition without separation.
Cement strength is reduced approximately in proportion to water-content increase.
However, we shall see later that high strength is not always required.

Type Amount used per sack


(% by weight)

Density reducers/extenders

Bentonite 2 to 16

Attapulgite 1/2 to 4

Diatomaceous Earth (Diacel D) 10, 20, 30 or 40

Pozzolan, Artificial (fly ash) 74 lb/sk

Natural hydrocarbons:

Gilsonite 1 to 50 lb/sk

Coal 5 to 50 lb/sk

Pozzolan-Bentonite Cement Variable

Sodium Silicate 1 to 7.5 lb/sk

Expanded Perlite 5 to 20 lb/sk

Hollow Spheres Variable

Density increasers

Sand 5 to 25

118
Barites 10 to 108

Ilmenite (iron-titanium oxide) 5 to 100

Hematite 4 to 104

Salt 5 to 16

Friction Reducers 0.05 to 1.75

Table 1: Materials used to vary slurry density.( Source: Halliburton Services).

For years, bentonite has been the most commonly used additive for filler-type cement.
In addition to its effect on density, yield, and cost, bentonite increases viscosity and
gel strength, which reduces settling of high-density particles (e.g., weight material,
cement), or floating of low-density particles (e.g., perlites, pozzolan, gilsonite, crushed
coal, hollow spheres). Bentonite also reduces API fluid loss. However, cements
containing bentonite are more permeable and have lowered sulfite resistance.

Pozzolans increase slurry viscosity and provide low permeability. Sodium meta-silicate
provides a very low density slurry with early compressive strength; this material and
calcined shale-cement (a special cement, not an extender) are becoming popular,
particularly in offshore applications.

Very light slurries (less than 8 lb./gal.) have been made using hollow spheres. These
new cements are useful in underpressured, hot geothermal wells and other special
applications.

Density-Increasing Additives
High density cement sluries are often necessary to offset the high pressures that are
frequietly encontered in deep or abnormally pressured fromations. Density may be
increased with weight material such as sand, barite, hematite or ilmenite, and/or salt
dissolved in the mix water, as shown in Table 1.

Type Amount used per sack


(% by weight)

Density reducers/extenders

Bentonite 2 to 16

Attapulgite 1/2 to 4

Diatomaceous Earth (Diacel D) 10, 20, 30 or 40

Pozzolan, Artificial (fly ash) 74 lb/sk

Natural hydrocarbons:

119
Gilsonite 1 to 50 lb/sk

Coal 5 to 50 lb/sk

Pozzolan-Bentonite Cement Variable

Sodium Silicate 1 to 7.5 lb/sk

Expanded Perlite 5 to 20 lb/sk

Hollow Spheres Variable

Density increasers

Sand 5 to 25

Barites 10 to 108

Ilmenite (iron-titanium oxide) 5 to 100

Hematite 4 to 104

Salt 5 to 16

Friction Reducers 0.05 to 1.75

Table 1: Materials used to vary slurry density.( Source: Halliburton Services).

Available densities and effects on compressive strength are shown in Table 2.

Material Spec. Gravity Max. Effect on


(lb./gal) Density compressive strength

Ottawa Sand 2.63 18 None

Barite 4.25 19 Reduce

Coarse Barite 4.00 20 None

Hematite 5.02 20 None

Ilmenite 4.45 20 None

Dispersant 17.5 Increase

Salt 18 Reduce

120
Table 2: Densities of weight materials and their effect on compressive strength.

A density of 22 lb/gal can be obtained with hematite or ilmenite plus friction-reducing


additives. Fine barite requires a large amount of water, which reduces compressive
strength and retards thickening time.

Slurry weighted with solids must have adequate viscosity and gel strength to carry and
suspend high-specific-gravity solids. In addition, some additives (e.g., fluid-loss
agents, retarders, water) tend to significantly thin or thicken a slurry.

High slurry densities (up to 17.5 lb/gal) may be obtained by (1) using heavy additives
and/or (2) adding dispersants to achieve pumpability at lower-than-normal
water/cement ratios. The latter is more expensive, but it yields the highest
compressive strength.

Pretesting of such high-density slurries should include measurement of density,


thickening time, compressive strength, settling, free water, and viscosity.

Filtration-Control Additives
Fluid loss, or the premature escape of mix water from the slurry before chemical
reaction occurs, can cause many downhole problems, including

differential sticking of casing and decentralization

formation damage by filtrate (if not controlled by mud cake)

loss of pumpability

cement bridging above gas zones and gas cutting from hydrostatic pressure
loss

improper or premature dehydration during squeezing

Filtration-control additives in present use and their recommended concentrations are listed in Table
1.

Type Amount used per sack


(% by weight)

Fluid-loss additives 0.5-1.5%

Organic polymers (cellulose), form


micelles

Organic plymers (dispersants), 0.5 - 1.25%


size distribution and form micelles

Carboxymethyl hydroxyethyl 0.3 - 1.0%


cellulose, from micelles

121
Latex additives, form films 1.0 gal/sk

Bentonite cement with dispersant 12-16% gel


Table 1: Materials to reduce filtrate loss, friction.

These materials function by forming micelles or films, and/or by improving particle-size


distribution, which holds liquids.

A neat Class G or H slurry has an API 30-minute filter loss of over 1000 ml. Figure 1
shows the effectiveness of high-molecular-weight synthetic polymer in reducing filter
loss.

Figure 1

Friction Reducers
Friction reducers or dispersants are commonly used to lower viscosity, yield point and
gel strength of the slurry to reduce friction in pipe, and thus allow turbulent flow to

122
occur at reduced pump rates. For example, to achieve turbulent flow with 7 5/8 in.
casing in a 8 5/8 in. hole requires a rate of over 600 gal/min. With 0.5, 0.75, and
1.0% friction reducing additives (FRA), the required rate is only 530, 300, and 210
gal/min, respectively. These additives also permit slurries to be mixed at lower
water/cement ratios so that higher densities may be achieved.

Some common dispersants are alkylaryl sulfonate, polyphosphate, lignosulfonate, salt,


and organic acid. Table 1 shows typical concentrations. Turbulent-flow additives tend
to cause settling and excessive free water. These effects should be tested in the lab
prior to field use.

Type Amount used per sack


(% by weight)

Friction reducers/ dispersants

Polymer: blend 0.5 to 0.3 lb/sk

Polymer: long chain 0.5 to 1.5 lb/sk

Calcium lignosulfonate 0.5 to 1.5 lb/sk

(organic acid)

Sodium Chloride 1 to 16 lb/sk

Organic acid 0.1 to 0.3 lb/sk

Table 1: Materials to reduce filtrate loss, friction.

Lost-Circulation Materials
"Lost circulation" or "lost returns" refers to the loss to formation voids of either whole
drilling fluid or cement slurry used during the course of drilling or completing a well. It
should not be confused with volume decrease caused by filtration.

Drilling fluids or slurries are usually lost to either natural or induced formation
fractures. These fluids may also be lost through highly permeable formations those
starting at about 5 darcies for drilling fluid with a maximum particle size of 0.002 in.
(300 mesh). Cement, with its larger particle size (neat cement has 2.6 to 18%
particles larger than 200 mesh) is less susceptible to loss in permeable formations.

The best time to treat the formation to reduce such fracture or formation permeability
is during drilling, when high concentrations of bridging materials and various types of
plugs (pills) may be utilized.

During primary cementing, concentrations of such materials must be carefully


controlled to avoid bridging the casing or liner-borehole annulus, or plugging downhole

123
equipment such as bottom wiper plugs, small-diameter stage tools, and float
equipment.

Types of lost-circulation additives available for cement are blocky-granular materials


(walnut shells, gilsonite, crushed coal, perlite-expanded and perlite-semiexpanded)
which form bridges, and laminated materials (cellophane flakes) which form flake-type
mats. In laboratory studies, granular material was found to be best suited for bridging
fractures ( Figure 1 ,

Figure 1

the performance of LCM materials in sealing simulated

Type Generic Type Volumes used, typical range


material name particle

Granular Gilsonite Graded 5-50 lb/sk

124
Perlite Expanded 1/2-1 cu ft/sk

Walnut shells Graded 1-5 lb/sk

Coal Graded 1-10 lb/sk

Lamellated Cellophane Flakes 1/8-2 lb/sk

Fibrous Nylon Short fibers 1/8-1/4 lb/sk

Table 1: Ranges of lost-circulation material (LCM) volumes used per sack.

Fibrous materials (such as nylon fibers) are used in drilling fluid for sealing large
openings but are not normally used in cement because they tend to plug surface and
downhole cementing equipment. Also, most other fibrous materials contain organic
chemicals that can seriously retard cement-thickening time.

Ranges of lost-circulation material (LCM) values used per sack are listed in Table 1.

Salt
Salt has many different properties. In addition to its uses as a dispersant and in slurry
densification, it may be used as a cementing additive in the following ways:

Bonding to Salt Formations Saturated-salt slurries are the best overall choice for
cementing across salt zones because they do not dissolve the salt zone and thus give a
better formation-to-cement bond.

Protecting Clay and Shale Formations Small amounts of sodium chloride


(NaCl) and potassium chloride (KCl) help protect clay and shale formations that
are otherwise susceptible to crumbling and sloughing.

Acceleration Use of salt may be used as an accelerator. KCl in small amounts


also promotes early-strength development.

Retardation Saturated salt is an effective retarder in circulating temperatures


up to about 23 to 260 F (100 to 127 C)

Expansion Salt can be used to cause linear expansion of cement to occur long
after the cement has set. This effect is minor but beneficial in obtaining a better
formation-to-cement bond.

Compressive-Strength Stabilizers
Four variables composition, temperature, pressure, and time affect compressive
strength. However, at high temperatures, cement compositions may retrogress (lose
strength) after reaching a high value and never attain the strength reached at lower
curing temperatures ( The effect of curing: Figure 1 , pressure and Figure 2 ,

125
temperature on cement strength.

Figure 1

The high temperatures cause strength retrogression).

126
Figure 2

This strength retrogression is accompanied by increased cement permeability, e.g., a


neat retarded cement with 0.02+ md permeability at 290 F (143 C) after three days
may have 8+ md at 320 F (160 C) after seven days.

Retarded cement (used in high-temperature applications) and high-water-content


cement seem particularly subject to strength retrogression. For cement types used in
deep and/or hot wells, the phenomenon begins at around 260 F (127 C), and
becomes severe at 290 F (143 C).

Silica flour in high percentages inhibits strength retrogression and produces


compressive strength far in excess of that of neat cement. Silica flour also reduces
permeability of set cement; for instance, its addition to cement cured at 350 F (177
C) reduces permeability to less than 0.001 md.

Usually 30 to 40% silica flour is used. Silica sand ground to 200 mesh reacts with
cement in the same way as fine-ground 325-mesh silica flour. Sand is used when high
density is desired, and flour when low density is adequate. The different densities are
achieved because of the different water requirements of the sand and the flour.

127
Compositions containing silica sand or flour can be retarded effectively for high-
temperature wells.

Natural pozzolans and fly ash produce a strong material with silica-stabilized cements
up to 450 F (232 C). At a temperature of 600 F (315 C), fly ash and, to a lesser
degree, natural pozzolans can cause cement to weaken and become more permeable.

Other Additives
Other additives are materials that do not fit readily into classifications. They are
usually compatible with classified additives, and include the following:

Radioactive tracers are added to serve as markers that can be detected by logging
devices. Radioactive tracers include iodine 131 (8-day half-life), scandium 46 (85-day half-
life), liquid iridium 192 (74-day half-life), and sand for squeeze work.

Anti foam agents are available in dry or liquid form to combat air entrainment.
They are used routinely in many cements to facilitate mixing. Some additives
can cause excessive air entrainment, making it difficult to achieve the desired
slurry density without the addition of anti foam agents.

Dyes are occasionally employed as an aid to determine the effectiveness of


mud displacement during cementing operations.

Mud decontaminants neutralize certain mud-treating chemicals that could


have a detrimental effect on the cement. They are used primarily in openhole
plug-back and liner jobs, squeeze jobs, and tail-in primary casing jobs.

Gypsum additives, which create thixotropic properties in cement slurry, help


combat lost circulation through rapid gel-strength development. Viscosity
increases and slurry gelling is induced when shear rate is reduced. Gypsum
additives decrease cement mobility and thus its setting time. In addition, these
additives have expansion properties to improve bonding in set cement.

128
PRIMARY CEMENTING EQUIPMENT

Surface Mixing and Pumping Facilities

Major components of surface-cementing equipment are:

mixers or blenders

pumping/displacing unit

cementing or plug-release head

Mixing Equipment

Dry cement must be mixed with the proper amount of water to ensure that slurry and
set-cement properties are as designed.

For most slurries, the jet mixer produces a uniform mixture. Special mixing equipment
is sometimes required for high-density cement, high-viscosity cement, and jobs for
which precise composition and blending of all additives is particularly critical (such as
liner and squeeze-cementing operations).

The jet mixer induces a partial vacuum at the venturi throat, drawing in the dry
cement. High stream turbulence then provides thorough mixing
( Figure 1 , Schematic of jet mixer and typical job-site setup). The jet mixer is simple,
reliable, and rugged, and can handle 50 sacks of cement per minute.

Figure 1

129
Batch mixing and/or blending is achieved through use of propeller- or impeller-type
mixers, paddle mixers, ribbon blenders, pneumatic mixing, and rotation of the cement
tank (similar in appearance to those used in construction). Figure 2 (pneumatic
blender ) and Figure 3 (ribbon blender ) illustrate pneumatic and ribbon types.

Figure 2

130
Figure 3

Only a limited amount of cement can be mixed in a batch unit. However, several units
can be combined to provide continuous operation on large jobs. Batch mixers provide
the most accurate and thorough mixing of all slurry components.

Density measurements are made to gauge consistency and control the mixing
operation. Variations during a job can result from nonuniform blending of dry
components, changes in water/cement ratio, or air entrainment in the sample.

Density is measured either on samples with balances (two types), or continuously with
radioactive devices or a force-balanced U-tube.

Pump-Skid Truck

The typical slurry-pumping unit ( Figure 4 : Courtesy of Halliburton Energy Services) is


truck-mounted, and contains diesel engines and displacement tanks that are accurately
graduated so that water or mud volumes can be controlled to place the slurry
downhole properly.

131
Figure 4

In operation, the pumper draws slurry from the mixer in a predetermined volume
calculated from sacks and yield. Meanwhile, displacing fluid is drawn from storage to
two open tanks on the pumper. The slurry is followed immediately by the fluid, which
is gauged by the alternate draining and filling of the two tanks.

Cementer or pump-skid styles vary greatly, from small portable units to large truck-
mounted systems. Most work is done at less than 5000 psi, but pressures up to 20,000
psi (137,900 kPa) can be handled with proper equipment. Pumping rates are
dependent on mixing capacity. High rates are 10 to 15 bpm. When nearing the end of
displacement, slow rates are preferred, to decrease the possibility of plug or casing
damage.

Cementing Heads

Cementing heads provide a connection for pump-truck and rig-pump lines into the
casing, and a receptacle for the plugs. Modern heads provide a quick-change cap that
can be removed to insert the cement plug
( Figure 1 , cement head with manual plug release).

132
Figure 1

Most cementing heads are designed to hold one or more plugs and are loaded before
the actual cementing operation. Plugs are selectively released into the casing from the
head. When casing rotation is desired, an adapter swivel is used between the topmost
collar and the head, and the casing is suspended by the rotary table slips. The
advantage of using a cementing head that provides space for two plugs is that it allows
for continuous pumping; that is pumping of the cement can proceed while the top plug
is being dropped.

Conventional Casing Jobs and Equipment Applications

The conventional two-plug casing job is illustrated in Figure 1 (Principal equipment


used in typical two-plug primary cement job ).

133
Figure 1

A predetermined volume of slurry is pumped into the casing between two wiper plugs.
The bottom plug ruptures when it seats; the top plug is solid. The top plug is displaced
with mud or completion fluid. Flow stops and pressure builds when the top plug lands.
Check valves in the float shoe to prevent backflow of the heavier column of slurry in
the annulus.

Guide/Float Shoes and Collars

In most cases, except in certain shallow wells, a round-nosed shoe is run on the
bottom joint to guide the casing past borehole irregularities encountered while the
string is run. Three types of shoes are commonly used:

guide shoes without valves of any kind

float shoes with a check valve that prevents slurry backflow

differential or automatic fill-up types

134
Collars have basically the same features as shoes Figure 2 (Regular type guide shoe )

Figure 2

Figure 3 (Down jet type guide shoe )

135
Figure 3

Figure 4 (Insert self fill-up float valve )

136
Figure 4

Figure 5 (Standard float collar )

137
Figure 5

Figure 6 (Standard float collar with double demale threads )

138
Figure 6

Figure 7 (Insert valve float shoe with self fill-up unit )

139
Figure 7

Figure 8 (Cement baffle collar with latchdown plug and sealing sleeve ).

140
Figure 8

They are commonly known as baffle collars (without valves), float collars, and
differential or automatic fill-up collars. Located one or more joints above the shoe, the
collar, in addition to float and fill-up functions, acts as a seat for pump-down wiper
plugs.

Since cement immediately below the wiper plug may be contaminated, the collar
should be positioned to minimize the amount of contaminated cement pumped out
around the shoe.

When float equipment is used, it prevents fluid from backflowing up the casing as the
casing is run in the hole. This causes the casing to 'float' downhole because it is
partially empty and somewhat buoyant. To control buoyancy effects with fluid when
using float shoes or float collars from a surface fill-up line.

Wiper Plugs and Displacement Concepts

Wiper plugs are used to separate cement from preceding or following fluids ( Figure 1 ,
top and bottom wiper plugs). The bottom plug also removes mud from the wall of the

141
casing, and prevents this mud from accumulating beneath the top plug and being
deposited around the lower casing joints.

Figure 1

When the two-plug system is used, the operator should verify that the bottom plug is,
in fact, placed in the bottom position in the cementing head. A mechanical device
should be used to give visual proof when the top plug leaves the head.

A bottom plug is not recommended with large amounts of lost-circulation material in


the slurry or with badly rusted or scaled casing, since such material may collect on the
ruptured diaphragm.

Displacement of the top plug should be carefully monitored. Fluid behind the plug
should be determined from calibrations on cementing-unit tanks or by measuring out
of a storage tank. Another method is to count pump strokes and convert to volume. If
available, a flowmeter can be used to verify volumes pumped.

If the top plug does not bump at a calculated volume (allowing for displacement-fluid
compressibility), displacement should be stopped. Troubleshooting the situation, as
well as keeping track of the top plug, requires that we have accurate measurements of
the volume of surface lines and equipment, the capacity of the casing, the depth to the
top of the float collar, and the amount of fluid pumped. Surface equipment capacity
should be estimated for individual locations ( although this capacity will be small
compared to the hole volume), while the volume inside the casing is base on its lined
capacity (i.e.ft. per linear ft.), which is available from service company tables or
cementing handbooks.

142
Casing Centralizer Design/Selection Factors

Casing centralizers are used to

improve displacement efficiency

prevent differential pressure sticking

keep casing out of key seats

Two general types of centralizers are spring-bow and rigid. The spring-bow type has greater ability
to provide standoff where the borehole is enlarged. The rigid type provides more positive standoff
where the borehole is "to gauge." Special close-tolerance centralizers may be used on liners.

Important design considerations are positioning, method of installation, and spacing.


Centralizers should be positioned on casing

through intervals requiring effective cementing

adjacent to (sometimes passing through) intervals subject to differential


sticking and may be used on casing that passes through the doglegs, and is
thereby subject to key-seat sticking.

Correct positioning requires a caliper log of the wellbore so that locations correspond
with to-gauge sections of the borehole. Installation method depends on the type of
centralizer: solid body, split body, or hinged. The hinged type is most commonly
installed.

Centralizers are held in their relative positions on the casing by either the casing
collars or mechanical stop collars. The restraining device (collar or stop collar) should
always be located within the bow-spring type centralizer so the centralizer will be
pulled, not pushed, into the hole. Therefore, the bow-spring type centralizer should not
be allowed to ride free on a casing joint.

Load-deflection curves may be used for determining the spacing required to achieve
desired standoff ( Figure 1 , centralizer with proper stop collar location and example
load-deflection curve). The standoff required to prevent differential-pressure sticking is
normally less than that required to centralize casing for good displacement efficiency.

143
Figure 1

The lateral load imposed on a casing centralizer is the combined effect of centralizer
spacing, casing weight, hole angle, and weight of casing below the centralizer and
dogleg (a minor effect).

Equations to calculate lateral load are available in World Oil's Cementing Handbook
(Suman et.al., 1977), p. 44. These can be programmed for calculators, and are
effective design aids. Data is also available from manufacturers and from API Spec
10D.

Field practices for centralizer spacing are summarized in Table 1 below.

Surface casing One centralizer should be placed immediately above the shoe and one at
the top of each of the bottom six joints, to ensure centralization and uniform placement of
cement in this critical section. Centralizers may also be installed to improve cement
placement around any critical water sands.

Intermediate casing One centralizer should be placed immediately above the


shoe and one at the top of each of the bottom six joints. Centralizers may also

144
be placed within the cement interval to ensure uniform cement distribution
opposite critical zones.

Production casing Place one centralizer immediately above the shoe and one at
the top of each of the bottom six joints. They should be placed on every joint
through the producing zones and extending 100 feet above (and below, if
applicable). Other potential problem zones, key seats, sticking areas, etc.
should also be protected with centralizers.

Liners Use centralizers if clearance and hole conditions permit.

Stage cementing Centralizers should be spaced over the cemented interval


above the stage collar and one joint below, since there is no casing movement
in such jobs. When used, the external packer acts as the lower centralizer.

Table 1: Rules of thumb for centralizer spacing in vertical holes. Source: World Oil Cementing
Handbook.

Wipers and Scratchers

Wipers and scratchers are used primarily to remove borehole mud cake. They also aid
in breaking up gelled mud. Both rotating and reciprocating styles are available. See (
Figure 1 ,

145
Figure 1

part a,b and c: rotating types ; part d: rotating type and Figure 2 , part e and f:
reciprocating types ).

146
Figure 2

These devices are rarely used on liners because of close clearances.

Rotating-type wipers or scratchers are run across the zone of interest plus an
additional 20 ft (6 m) above and below the zone. Reciprocating-type scratchers are
generally spaced at 5- to l5-ft (1.5- to 4.5-m) intervals throughout the zone and an
additional 20 ft above and below.

When reciprocating, the vertical casing movement should always exceed the distance
between wipers or scratchers. If reciprocal movement equals the spacing, removed
mud cake and cuttings can accumulate at the end of each stroke.

When wipers or scratchers are used, mud circulation should always be started before
the pipe is moved; and initially, the pipe should be moved slowly. If no pipe movement
is planned, these devices should not be run.

147
PRIMARY CEMENTING OPERATIONS

Primary Cementing
Primary cementing is the cementing operation performed immediately after casing has
been run in the hole. This basic principle varies with the many materials used to
perform the many cementing operations. More and deeper wells are being drilled that
have extreme temperatures, both hot and cold, in new and more hostile environments.
This presents a constant challenge to successful primary cementing. Meeting this
challenge has led to an increasing number of ingenious and complex materials, tools,
equipment, and techniques.

Primary cementing uses several basic techniques. The most widely used procedure is
the single-stage primary cement job using a two-plug method. Cement is pumped
down the casing between two rubber plugs. The plugs are equipped with wiping fins to
help prevent contamination of the cement by drilling mud, and to help clean the
interior wall of the casing.

The use of other common techniques depends on well depth and completion
requirements. Two-, three-, and four-stage cementing procedures decrease the
hydrostatic pressure of the fluid column in the annulus, help protect weak zones
against excessive high pressures, and help prevent circulation loss. In addition to its
economic advantages (i.e., it is not necessary to cement the entire string back to the
surface), multiple-stage primary cementing is also important in wells where two or
more zones are separated by long intervals.

Terminology
Setting the casing at or near the bottom and perforating it for expected production is
common practice in the industry. Sometimes the casing is suspended and cemented
above the producing formation and the well is produced from open hole; this is called
an openhole completion. The final casing string is called the flow string or oil string.
The field terms long string and production string are used in some areas. The term
casing string denotes the total footage of casing run in the well at one time.

The phrase "waiting on cement," or WOC, has long been a misnomer in most
instances: the nonproductive and expensive time spent waiting has usually not been
necessary. In most cases, the cement has firmly set some time before operations are
resumed. In the early days of cementing, standards for curing concrete in the
construction industry were adapted as appropriate WOC time for oil wells. The first
wells were shut down for 28 days to allow the cement to set. Ten years later, oil field
operators were reducing the time to three weeks, and some were resuming operations
after only two weeks. Ten years after that, 10 days was considered sufficient. Some 30
years after cement was first used, most operators accepted three days as sufficient
WOC time. Accelerators are currently added to cement, and make it possible to resume
operations within a few hours.

Important considerations in the determination of WOC time are:

148
how much strength the cementing composition must develop before drilling
can continue

the strength development characteristics of the commonly used cementing


compositions, as well as those of available materials that are not being used to
their best advantage

cement-curing temperatures that exist under wellbore conditions

By using curing pressures that closely simulate those found in oil or gas wells, it is possible to have
a better understanding of additive performances and more realistic WOC time for cementing
compositions.

Downhole Temperatures

Hole conditions and curing environments for cement slurries vary in temperature from
below freezing in permafrost zones to 700 F (371 C) in geothermal steam wells. The
capabilities and versatility of most API cements can be extended by using additives. A
blend of additives usually produces an optimum range of cement qualities.

Temperature studies conducted along the Gulf Coast of Texas and Louisiana in the
early 1950s formed the basis of API testing schedules and cement specifications. The
schedules are based on bottomhole temperatures: F = 80 F + [0.015 x depth in
feet]. The cooling effect of mud displacement lowers considerably the circulating
temperature of the hole during casing cementing. During squeeze cementing, there is
less cooling because there is less well fluid preceding the slurry. Thus, a cementing
composition can be pumpable longer during casing cementing than during squeeze
cementing at the same depth.

The bottomhole cementing temperature may be determined from logs and API
temperature data. Log temperatures taken approximately 24 hours after the last
circulation ended may be considered static for use in API Spec 10.

Slurry Volume

The amount of cement used in creating a slurry depends on the estimated total slurry
volume. If a caliper log is available, an allowance of 10% above the volume calculated
from the caliper information is generally acceptable. If volume from the bit size of the
drilled hole must be used, allow from 50% to more than 100% above the volume
calculated from the bit size. The excess-cement percentages of other wells in the area
of interest can be used as a guideline.

In some areas, regulatory requirements dictate how far the top of the cement must be
above the uppermost pay zone. Three hundred to five hundred feet of cement above
the top of the pay is typical for many locations. This volume is contingent upon contact
time the time it takes for cement to flow past a given point in the annulus. The
greater the contact time, the greater the chance of removing the drilling mud from the
annulus. Only a specific part of the contact time should be considered in calculating
slurry volume: when the cement is pumped at high velocity past the highest point in
the annulus where good zonal isolation is needed. Any time spent displacing at a low

149
rate just before bumping the top plug need not be included. Once we have an estimate
of total slurry volume,we can easily determine our bulk dry cement requirements and
water requirements along with the amounts of additives needed, by going to service
company cementing tables.

Suppose for example, that we have calculated a slurry volume of 1,000 ft.3 of class 'G '
neat cement for a surface string of 10 3/4" casing in a 14 3/4"

hole. From the cementing tables (Halliburton, 1981) we know that 1 sack of class 'G '
cement yields 1.15 ft3 of slurry, and that this slurry requires 5.0n gal per sack of
cement. Our cement and water requirements, therefore, are 870 sacks and 4,350
gals., respectively.

Depth

The first practical instruments for investigating the effect of pressure on cement
thickening time were developed by R. F. Farris (1946). He indicated that the reduction
in thickening time caused by the increase in pressure from atmospheric to 5000 psi
averages about 35% for all cements. Pressure imposed on a cement slurry by the
hydrostatic load of well fluids also reduces the pumpability of cement. In deep wells,
hydrostatic pressure plus surface pressure during placement can exceed 20,000 psi
(137,900 kPa). Increased pressure seems to alter the forces of the contact surfaces of
the water and cement, which accelerates hydration.

Strength to Support Axial Loads

Laboratory Tests

High axial loads may be imposed on the casing string and/or surrounding cement by
landing and suspension methods and later operations. The cement strength required to
support such axial casing loads has been determined through shear-bond tests (
Figure 1 , Lab test setups to measure casing-cement bonding characteristics).

150
Figure 1

The axial load which breaks the cement bond was measured and the ability of the
cement to support axial casing loads was found to be proportional to the area of
contact between cement and the casing. Therefore, "support coefficient," "shear bond,"
or "sliding resistance," as described by various investigators, is the load required to
break the bond divided by the surface area between cement and pipe.

Based on worst-case results, Bearden and Lane provided a relationship for determining
the support capability of a cement sheath, conservatively utilizing results for mud-
wetted and nondisplaced conditions. Modifying their relationship to utilize compressive
strength (assumed to be 10 times tensile strength), gives the formula (in conventional
oilfield units),

F = 0.969 ScdH (1)

where:

F = force or load to break cement bond, lb

Sc = compressive strength, psi

151
d = outside diameter of casing, in.

H = height of cement column, ft

For example, for one bonded foot of 7-in. casing, using 500-psi compressive strength
cement,

F = 0.969 x 500 x 7 x 1 = 3390 lb

Required Strength

The load to break the cement bond during hanging and drilling-out operations normally
does not exceed the weight of the casing string (such as surface pipe) plus
miscellaneous loads (such as weight on bit when drilling out the shoe joint).

Therefore, the load capacity noted above (3390 lb/ft [5045 kg/m] of cement column),
provided by the relatively low compressive strength of 500 psi (3450 kPa), should be
more than adequate to handle anticipated axial loads.

Cement composition normally can be formulated to rapidly develop adequate strength


for casing landing loads. This allows drilling operations to proceed with little or no WOC
time.

Also, low-strength filler cements, which are relatively inexpensive and of low density,
and are less likely to induce lost circulation when high cement columns are required,
may have adequate compressive strength to meet axial-load support requirements.

Needs for Zonal Isolation/Hydraulic Bonding

High Strength Needed

Although low-compressive-strength cement may be adequate to handle axial and


rotational casing loads, ultimate high strength may be required for zone isolation and
to support the borehole. Therefore, cement compositions should be selected that
provide both adequate immediate compressive strength for drilling operations and
adequate ultimate strength for production operations.

A comprehensive study of factors governing zone isolation under downhole conditions


would be very complex. Therefore, only qualitative judgments have been attempted in
studies to date, and these usually relate to the hydraulic bond that indicates adhesion
between casing and cement or between cement and formation.

The actual relationship between hydraulic bond measured in the laboratory and
downhole zone isolation has not been reported.

Bonding Tests

Various investigators have measured hydraulic bond. Test arrangements are shown in
Figure 1 (Lab test setups to measure casing-cement bonding characteristics).

152
Pressure is applied to the exterior surface of the casing, causing the casing to become
smaller in diameter and pull away from the cement, forming a microannulus that
permits leakage. Results of the setup in Figure 1 are shown in Table 1, below.

Type mud Hydraulic


bond
(psi)

Surface finish Wetting Water Gas

New mill-varnished None 200-250

Varnish removed (chemical) None 300-400

Varnish removed (sandblast) None 500-700 150

Varnish removed (sandblast) Fresh water 100 50

Varnish removed (sandblast) Invert oil 100 50


emulsion

Varnish removed (sandblast) Oil base 100 50

Resin-sand coat (new, None 1,000- 450


sandblast) 2,000

Resin-sand coat (new, Fresh water 100 55


sandblast)

Resin-sand coat (new, Invert oil 100 45


sandblast) emulsion

Resin-sand coat (new, Oil base 100 45


sandblast)

Cement: API Class A

Water content: 5.2 gal/sk

Curing temperature: 80 F

Curing time: 24 hours

Casing size: 2 in. inside 4 in.

Table 1 : Hydraulic bond vs. casing surface and type of fluid wetting.

Surface condition Surface coating Hydraulic bond


(psi)

Dry Mill varnish < 20

153
Mud film Mill varnish < 20

Dry Rusty 350-450

Mud film Rusty 20-50

Dry Acid-etched 250-400

Mud film Acid-etched 40-50

Dry Sandblasted 500-600

Mud film Sandblasted 50-60

Dry Epoxy-coated, 6-12 700-950


mesh sand

Mud film Epoxy-coated, 6-12 500-600


mesh sand

Curing time: 24 hours

Curing temperature:
120 F

Table 2: Effect of mud film on bond strength.

Annular Devices Help

The pressure at which failure of the hydraulic bond occurred in the test can be
increased by:

preventing formation of the microannulus by controlling pressure differential


across the casing as the cement sets, and/or;

attaching seal rings of deformable rubber (similar to those available for field
installation) to the exterior of the casing.

However, zone isolation is routinely obtained in the field at greater differential


pressures than those causing failure in these hydraulic-bond tests. Therefore, such
tests are probably not completely representative of downhole conditions.

Mud wetting and Ruff Cote

Further tests were conducted to more directly measure adhesion between cement and
pipe. These tests showed an advantage to the resin-sand (Ruff Cote) exterior in the
mud-wetted condition, which was not apparent in the previously discussed test (see
Table 2, above). When resin-sand coatings are used downhole, however, their
effectiveness should be increased by removing mud from the casing surface with
preflushes ahead of the cement, and by cement scouring.

154
Casing with Ruff Cote should be well centralized to avoid the embedment of mudcake
or shale into its roughened surface. This may not be possible in irregular, doglegged,
or high-angle holes, or where mud is poorly conditioned.

One advantage of the resin-sand is that it inhibits formation of a microannulus under


certain pressure and temperature conditions. This appears to be verified by cement
bond logs.

Casing Support in the Borehole

The cement sheath can protect the casing against several types of downhole damage,
including:

deformation by perforating guns;

formation movement, salt flows, etc.;

bottom-joint loss on surface/intermediate strings during drilling.

However, added resistance to casing collapse for design purposes is questionable. In


fault-slippage zones, doglegs, and certain sand-control failures, the cement sheath
may contribute to problems.

Perforating - Expendable versus Carrier Guns

The cement sheath tends to minimize casing damage caused by expendable


perforating charges. Expendable guns of nominal charge for example, through-
tubing guns may be used in cemented pipe with little or no danger of serious casing
damage. However, expendable charges may split casing collars that are unsupported
by cement, and expendable gun charges of over 20 g frequently damage partially
supported or unsupported casing.

Figure 2 (Cement sheath affects casing deformation by perforating with expendable


guns ) (top) shows lab tests on casing deformation with 20-g charges and three cases
representing no cement (top curve), 3/4-in.

155
Figure 2

(1.9-cm) sheath supported by thin steel (middle curve), and a strongly encased sheath
(lower curve). Compressive strength of the cement in the sheath had little influence on
results, as shown in the bottom figure.

Conventional hollow carrier, shaped-charge guns cause only slight casing deformation
and essentially no damage regardless of support, because most forces from the
exploding charges are contained by the carrier body.

Salt Flow

Casing damage can be caused by lateral loads resulting from flow of salt formations.
Salt may flow in various ways, and it may not be economically practical to design
casing for the most severe situations of nonuniform loading possible, such as the
flattening effect illustrated in Figure 3 (No cement or partial sheath results in eccentric
pipe loading )

156
Figure 3

and Figure 4 (With fault slippage, unconstrained pipe may minimize damage ).

Figure 4

157
However, when the annulus is completely filled with cement, casing is subject to a
nearly uniform loading approximately equal to the overburden pressure. Although the
modes of failure may differ, casing designs for withstanding salt pressure can be
computed on the same basis as those for withstanding fluid pressure.

Casing failure caused by formation movement along natural or induced fault planes
as opposed to salt flow is best handled by elimination of cement through the
affected interval and perhaps by opening the hole to enable fault slippage to occur
without loading the casing in shear ( Figure 3 and Figure 4 ).

Other downhole conditions, such as borehole doglegs and sand-control failures, also
may cause casing damage similar to the types described above. Knowledge of the
failure mechanism is essential to the selection of the failure-prevention method-i. e..,
cement sheath or no cement sheath.

Adequate cement strength and good cementing and operational practices may be
required to prevent parting or other failure in bottom joints of surface and intermediate
casing strings.

Visual inspection of a joint failure reveals whether the casing is unscrewed or broken.
Unscrewing occurs because of high-level torque impulses transmitted to the casing by
the bit as it hangs up while drilling cement and cementing equipment out of the
bottom joints. It can also be caused later by tool joint torque action on the lowermost
joints as drilling proceeds ahead.

The problem is usually prevented by welding or by using thread-locking compounds on


the connections of the lowermost two or three joints and controlling rotary speed.
Other prudent cementing practices include the following:

Apply standard good practices when cementing: e.g., maximize casing


movement; use high-rate displacement, centralization, and proper washes and
flushes.

Around the shoe, use quality cement with early high-compressive strength.

Use two plugs to prevent mud fill around the shoe joint, and do not
overdisplace even if a top plug is used.

Release pressure to avoid microannulus formation (if this is compatible with


landing methods).

Keep drill pipe out of the hole until the cement has adequate initial set.
Minimum strength for drillout is 500 psi (possibly 1000 psi [6895 kPa]).

A lowered casing design safety factor in collapse (perhaps 0.85 versus 1.125) is
sometimes considered for casing to be used below the cement top, on the assumption
that cement will provide additional support. This is not a valid practice.

158
According to Cheatham and McEver, cement in the annulus between salt and casing is
compressed by salt pressure, reducing stress transmitted to the casing. However, this
reduction is less than 5% for 8 5/8-in. casing in a 12-in. hole, or about 200 psi (1380
kPa) for 6000 psi (41,370 kPa) acting on the cement. Furthermore, this load reduction
depends on uniform placement of cement, a condition not normally achieved.

Other tests suggest that a cement sheath may provide greater collapse-resistance
support for lower-grade casing (H-40, J-55). However, minor radial or longitudinal
discontinuities in the cement sheath eliminate this support.

Stage-Cementing Equipment and Methods

Multiple-stage cementing consists of conventional placement of cement slurry around


the lower portion of a casing string followed by placement of successive upper stages
through ports in a stage or port collar. Although most stage cementing is done in two
stages, additional stages are possible.

Stage cementing may be appropriate:

when a thick interval requires cementing and a weak formation will not
support it

when two or more widely separated intervals must be cemented

in special cases (such as in arctic wells in permafrost) where cement is placed


near surface to help suspend casing

in deep, hot wells, to place faster-setting slurry above retarded compositions


in lower, hotter zones

Various tools allow flexibility and variety of application.

Stage collars are most commonly used. The stage collar contains ports that are initially
isolated by a sliding sleeve. The sleeve can be moved downward to open the ports
and moved up to close them with a special bomb or tripping plug. Figure 1
(application of stage tools - displaced type method )

159
Figure 1

and Figure 2 (application of stage tools: free fall plug method )shows two types of
collar-opening methods: displaced plugs and free-fall plugs.

160
Figure 2

Port collars can be opened or closed by mechanical action of an inner string, such as
drillpipe. Two types of port collars are shown schematically in Figure 3 (two types of
port collars opened and closed by pipe
movement ); these are actuated by inner-string rotation or vertical manipulation.

161
Figure 3

The tools shown in Figure 3 are applicable to specialized jobs such as the placement of
fluids behind pipe for corrosion protection or sand control, or cementing with treating
packers.

Petal-Basket Packer

Mechanical devices are frequently used below stage tools to prevent upper-stage slurry
from dropping through the mud. They may also be used below conventional shoes
where casing is landed above the open hole.

The most common support device is the metal petal basket attached around the casing
exterior. The basket allows vertical fluid movement but opens against the borehole to
prevent downward movement. Strength and sealing ability limit its use to shallow
depths.

A more rugged support device is the solid rubber or inflatable external casing packer.
Typically, this is placed below the cement-outlet port and inflated with mud or cement
prior to the opening of the cementing port

162
( Figure 4 , External packer supports cement column over a weak zone at shoe or for
stage cementing).

Figure 4

Types of Casing

Conductor

The conductor pipe ( Figure 1 , conductor pipe) is the first string set in the well. It may
be set by the rotary rig drilling the hole, or by a smaller rig (or rathole machine) before
the larger rotary rig is moved in.

163
Figure 1

This casing string serves as a conduit to raise circulating fluid high enough to return it
to the pit. Conductor casing also supports part of the well load where ground support is
inadequate or where the well is going to be drilled to a great depth.

The purposes of conductor casing are:

to prevent washing out under a rig

to provide elevation for the flowline

to provide support for part of the wellhead

A blowout preventer (BOP) is not usually attached to the conductor casing.

Some characteristics of conductor casing and its placement are as follows:

Casing is usually large: 20 to 30 in. in diameter.

164
The hole for the casing may be severely eroded.

Casing can be easily pumped out, and is usually tied down.

Setting depth can vary from as little as 20 ft to as much as a few hundred


feet.

The most common pipe and hole sizes are 16-in. pipe in a 20-in. hole, and 20-
in. pipe in a 26-in. hole.

Recommended cements for use with conductor casing are:


accelerated neat

ready-mix concrete

thixotropic cement

LCM additives

Typical slurries for conductor-casing application include API Class A, C, G, or H with 2% calcium
chloride as the accelerator. Lost-circulation additives such as sand, gilsonite, and cellophane may
be added without significant effect on the slurry-thickening time or compressive strength. Where lost
circulation is severe, a thixotropic cement can be used.

The following is a brief summary of conductor-casing cementing practices:

The amount of cement used should be sufficient to provide returns to surface

The casing is often cemented through drillpipe with a sealing sleeve.

When cementing down casing, plugs may not be used; cement is simply
placed.

Large-diameter (30-, 26-, and 20-in.) casing plugs are wooden body plugs. If
bumped on baffle or float, care must be taken in pressuring up to prevent
bypassing the plug with displacement fluid.

The amount of excess cement is usually determined by experience in the area


of interest.

The following factors must be considered in the selection of the correct slurry composition for a
conductor casing:
The set cement must have a compressive strength high enough to support an appreciable
wellhead load; therefore, high-compressive-strength completion cements are best.

Since the temperature at shallow depth is usually only 80 or 90 F (27 or 32


C), the cement should be accelerated for a shorter thickening time and earlier
compressive-strength development than is possible with neat cement. This

165
early strength saves rig time and ensures sufficient strength to support
continued drilling operations.

When the cement cannot be circulated back to surface, a top-out slurry is


usually pumped into the annulus through small tubing on top of the primary
cement.

Surface Casing

Surface casing ( Figure 1 , surface casing) is usually the second string of pipe set in
the well. However, when a conductor casing is not set because the well is on firm
ground or will not be drilled to great depth, the surface pipe is the first string set.

Figure 1

The purpose of surface casing is to

protect freshwater sands, particularly underground sources of drinking water

case unconsolidated formations

166
provide primary pressure control (BOP is usually nippled up on surface casing)

support future casings

case-off potential lost-circulation zones

The casing head and other fittings used for completing the well are attached later.

Some characteristics of surface casing and its application are as follows:

Casing sizes normally range from 7 5/8 in. on shallow wells to 20 in. on deep,
multistring wells.

The hole in which it is set may be severely eroded.

Shallow strings can be easily pumped out.

Drilling muds are often viscous, with little water-loss control.

Casing may stick easily in unconsolidated formations.

Loss of circulation may be a problem.

Most areas require that cement be circulated.

A guide shoe (or float shoe), float collar, scratchers, and centralizers are
commonly used.

Casing may be set from a few hundred feet to several thousand feet; the
depth depends on the proposed total well depth, the competency of shallow
formations encountered, and state regulations regarding protection of
freshwater zones.

Recommended Cements

Shallow surface casing is cemented in the same manner as conductor casing.


Completion cements with accelerated thickening times and compressive strengths are
used. Top-out slurries are used on surface jobs when cement is not circulated back to
surface.

For deeper strings of surface casing, a lightweight lead cement is used, followed by
heavier-weight completion cement. Sometimes, when zones are penetrated by long
surface casings, a lightweight lead cement may help keep these formations from
breaking down under the hydrostatic pressure exerted by a long column of cement.
The bottom of the surface casing around the shoe is cemented with the high-strength
completion cement.

This creates a strong seal with the pipe and formation for solid support of the casing.

167
There is a cost advantage to using a high-yield completion slurry to cement the entire
string. When the well is shallow and a significant load is to be placed on the wellhead,
a densified filler slurry can be used to cement the entire casing. Omission of the tail-in
slurry is not economically advantageous.

Recommended cement types include:

The amount of cement used should be sufficient to provide returns to surface.


This is a regulatory requirement in many locations.

filler cements (with a high water content) followed by neat or high-strength


tail-in

accelerated cements

LCM additives

high-strength cements, which are often used on deep-well surface casing to


support future strings

The following is a brief summary of surface-casing cementing practices:

Large-diameter strings are often cemented through drillpipe with a sealing


sleeve.

Both bottom and top plugs should be used to prevent mud contamination.

The bottom joints and thread lock should be centralized to prevent backing off
during drilling.

Regulator rules usually require WOC of 8 hours, or 500 psi minimum


compressive strength.

During displacement with mud, a float collar placed two joints above the guide shoe
helps prevent mud from contaminating the cement around the shoe joint. Scratchers
and centralizers are the final consideration of casing equipment. Scratchers are
sometimes used to help clean the mud from the formation face. Centralizers center the
casing in the hole to help place the cement completely around the pipe.

A top and a bottom plug should be used to wipe the pipe clean ahead of the cement
when mud is the drilling fluid. If only a top plug is used, the mud wiped off the casing
builds up behind the cement and contaminates the cement around the shoe.

Intermediate Casing

The intermediate casing ( Figure 1 .,intermediate casing) is the first string of pipe set
after the surface casing.

168
Figure 1

It is sometimes called the protection casing. Intermediate casing strings extend from
the surface to a formation able to hold the mud weights expected at greater depth.
This depth can vary by several thousand feet in a single-stage job. When a second
intermediate string is set, the casing is run to just below the weak zone to a competent
formation and is cemented at that point.

The purpose of intermediate casing is to

separate the hole into workable increments for drilling

case-off lost-circulation zones, water flows, etc.

isolate salt sections

protect the open hole from increases in mud weight

prevent flow from high-pressure zones if mud weight must be reduced

control pressure; the BOP is always installed

169
support subsequent casings

Some characteristics of intermediate casing and its application are the


following:

Pipe and hole sizes are determined largely by the number of casing strings to
be run below the intermediate string.

Casing sizes range from 6 5/8 in. to 20 in. Most common are 9 5/8 in. 10 3/4
in., and 13 3/8 in.

Some sections particularly salt sections may erode severely.

Strings may be very heavy and set on bottom.

Both extremely weak zones and high-pressure zones are covered by


intermediate strings.

Cement volume is dictated by wellbore condition.

A guide shoe (or float shoe) and float collar are commonly used.

Cement volumes are usually largest in the well.

Intermediate casing is often cemented in stages.

Prolonged drilling may be done through this casing, and damage is common.

Completion may be made in intermediate casing.

Recommended Cements

Because of the large volume of cement required, and the types of formation to be
covered, both filler and composition cements are used to cement most intermediate
casing. Sometimes, as many as three different slurries are needed. Formation-fracture
gradients, lost-circulation zones, formation temperatures, possible future producing
zones, and well depth determine the number and types of slurries to use.

The slurry requirements for a single-stage cementing job are similar to those for a long
surface job. The filler slurry needs to be light enough not to break down the weaker
formations. The completion slurry needs to have enough strength to hold the pipe and
provide a good seal between the pipe and the formation.

The bottom of the pipe is cemented (usually at 100 to 3000 ft) in a single-stage
intermediate job because of cost considerations, or because the uncemented sections
of casing may be reclaimed from the well later and reused. In the latter case, only a
high-strength completion slurry with a retarder is needed. Retarders ensure sufficient
pumping time to get the slurries in place, and also impart some friction-reducing
properties to the slurry.

170
Unlike the conductor and surface casings, additives such as friction reducers, fluid-loss
additives, and retarders are required for intermediate slurries. Where the annulus is
small, friction reducers lower pump pressures and reduce the chance of losing fluids in
a lost-circulation zone. Fluid-loss additives prevent slurry loss into lost-circulation
zones and dehydration in the annulus caused by permeable zones, and also give better
bonding results.

The following is a brief summary of intermediate-casing cementing practices:

Both bottom and top plugs should be used to minimize contamination of the
cement.

Stage tools are used occasionally in cementing long strings of pipe where
there is risk of breaking down a weak formation.

The number of slurries required may be determined by possible production,


weak zones, and wellbore temperatures.

Scratchers, centralizers, and flushes can be important in the successful


completion of an intermediate-casing cementing job.

Production Casing

The production casing ( Figure 1 , production casing) is the last full string of pipe set in
the well, and extends to the surface.

171
Figure 1

Production tubing, downhole pumps, and other equipment needed for the production of
oil and gas are housed in this casing. The production-casing cement must give a
pressure-tight seal between the formations and the production casing. It is essential to
isolate the reservoir from fluids both within the producing zone itself and from other
zones penetrated by the wellbore. These fluids (e.g., oil, water, gas) can create
emulsions, scale deposits, paraffin deposits, severe corrosion, and a decline in
production. Besides primary producing operations, remedial workover jobs such as
squeezes or chemical treatments are also run through the production string.

The purpose of production casing is to

complete the well for production

effect zonal isolation

protect pay zones from unwanted fluids

provide pressure control

cover worn or damaged intermediate casing

172
Since the production casing may extend from total depth to surface, the setting depth
can vary from a few thousand feet to as much as 14,000 ft (4270 m). Below 14,000 ft,
liners may be set to reduce cost and because less pipe weight is needed. The size of
the casing depends on the number of strings of production tubing to be run into the
well and the size of production equipment used.

The following are some characteristics of production casing and its application:

Common casing sizes are 4 1/2 in., 5 1/2 in., and 7 in.

Drilling mud is usually in good condition.

The cement job is usually not circulated, but cemented back to intermediate
casing depth.

A good cement job is vital to a successful completion.

To achieve a pressure-tight seal and protect the reservoir, special consideration must
be given to the production-casing cement properties. As with intermediate casing and
long surface pipe, both filler and completion cement are usually employed. The filler
cement needs good fluid-loss control. It must have enough compressive strength to
protect upper, potentially productive zones which might be completed in the future.
The completion slurry needs to have good fluid control and sufficient compressive
strength to hold the weight of the pipe and to bond the formation to the pipe. The
setting times of both slurries should be minimized to help prevent cement
contamination from formation fluids and formation contamination by cement filtrate.

Recommended types of cement are

filler cements with high-strength tail-in

low-water-ratio cements (for all potential pay zones)

densified cements (for high competency and pressure control)

fluid-loss control additives

Summary of Production Casing

Batch mixing or continuous batch mixing is recommended for all large or critical jobs.
Since the production string affects the success of the well more than any other casing,
a good job may mean the difference between success and failure of the well.

Efficient removal of the mud is essential. Spacers or flushes may be used to remove
mud and to water-wet the pipe and formation face for good cement bonding. Usually
casing reciprocation or rotation is used. A float shoe and float collar, centralizers,
scratchers, and pipe movement should be used.

Liner Cementing

173
A liner is a string of casing that does not extend up to the wellhead. It is used to case-
off the open hole below an existing casing string.

Several types of liners may be categorized by their function:

Drilling liners permit deeper drilling operations by isolating lost circulation or highly
pressured intervals and controlling sloughing or plastic formation. In lieu of full-length
casing, the drilling liner improves drilling hydraulics, i.e., the greater cross section above the
liner top enables the use of larger drillpipe and/or reduces annular pressure drop.

Production liners provide isolation and support functions when casing has been
landed above the producing interval.

A tie-back stub liner extends from the top of a liner to a point uphole, inside
another string of casing or liner. The stub liner is used to cover damaged or
worn casing above an existing liner, and to provide added protection against
corrosion and/or pressure.

Tie-back casing extends a liner to the wellhead. It is used primarily for the
same purposes as the tie-back liner. Running such a string at the end of a
drilling operation ensures that the completion will be run in unworn casing.

Figure 1 (Example of deep-well tubulars, liner/tieback application ) shows the tubular program of a
modern deep well using two liners and tie-back casing.

174
Figure 1

Cementing Procedures

Cementing procedure is illustrated in Figure 2 ,

175
Figure 2

Figure 3 ,

176
Figure 3

Figure 4 , and Figure 5 (Typical liner cement procedure).

177
Figure 4

With the liner hung in the casing but still attached to the drillpipe, slurry is pumped
into the drillpipe without a bottom wiper plug.

178
Figure 5

The cement is followed by a drillpipe wiper plug that latches into a liner wiper plug,
positioned below the liner hanger. The combination plug then wipes the liner clean and
finally latches into a landing collar to complete slurry placement.

The following are important details in cementing procedure:

With the liner in position, mud is circulated to ensure that the liner and the float equipment
are free of any foreign material, and to condition the mud. A clean mud system is important
so that materials will not fall out on top of the liner-running assembly during the cement job.

The cement can be batch-mixed, circulated through a holding tank or ribbon


blender, and/or double-pumped to obtain the desired cement-slurry properties.

Cement slurry should be pumped in turbulent flow, or as fast as possible, refer


to the heading titled "Fluid Flow Properties and Mud Displacement". Such flow
minimizes excess cement-volume requirements. Most operators prefer to limit
excess cement volume, which, of course, is pumped into the drillpipecasing
annulus. It is usually desirable to pump some type of spacer fluid (buffer)
ahead of the cement.

If no bottom plug is used, the drillpipe and liner plugs wipe mud film off the ID
of the drillpipe and liner. This mud collects below the plugs and can
contaminate cement in the bottom of the liner. Spacing between landing collar

179
and float shoe should be adequate to keep contaminated cement out of the
liner-openhole annulus.

With cement in place, it is standard procedure to pull the liner-setting


assembly out of the liner hanger. With the tailpipe of the liner-setting assembly
above the liner top, excess cement can be reversed out. However, reverse
circulation places an extra pressure on the annulus that must be controlled to
prevent formation breakdown. A liner packer keeps reverse-circulation
pressures off the formation.

One method is to pull the drillpipe all the way out of the hole and leave
cement inside the casing to be drilled out. WOC time depends on cement
composition and hole conditions.

Liner-Cementing Equipment

A liner is usually run on drillpipe that extends from the liner-setting tool to surface.
Special tools perform various running, setting, and cementing operations ( Figure 1 ,
Typical equipment used to install and cement liners).

Figure 1

180
Shoes/Collars

A float shoe placed at the bottom of the liner contains a check valve designed to
prevent backflow of the cement. A float collar can be run above the shoe to provide a
back-up check valve. Automatic fill-up-type float equipment may be selected, but it is
rarely run on liners.

A landing collar is usually run one joint above the float collar or two or more joints
above the float shoe to provide space for mud-contaminated cement inside the liner.
The collar's function is to latch and seal the liner wiper plug. It prevents the liner wiper
plug from moving uphole if a check valve fails, and prevents it from rotating, which
aids drillout.

Liner wiper plugs can be attached to the end of the tailpipe or slick joint with a shear-
pin arrangement. The selection of the proper shear rating is very important in the
prevention of premature shearing and release of the liner wiper plug.

The liner wiper plug can also be latched to the tailpipe to prevent premature shearing.
Release of this type can only be effected by engagement of the drillpipe water plug.

Hangers and Setting Tools

The liner hanger is installed at the top of the liner. Hangers are usually classified by
the method used to wedge slips against the casing wall; two such classifications are
mechanical and hydraulic.

The presence of slips between liner and casing reduces the bypass area for circulating.
This reduced area can create high-pressure loss during circulation and cementing.
Hangers are available with multiple split slips that increase bypass area and provide
increased slip-contact area.

The liner-setting tool, a rental item furnished by the liner service company, provides
the connection between drillpipe and liner. Swab cups attached to tailpipe, or a packoff
bushing and slick joint, are inserted into the liner to provide a seal between setting
tool and liner.

Once the liner is hung, the setting tool can be released and picked up a short distance
to confirm, by indicator weight loss, that the setting tool has separated. A new
retrievable packoff bushing eliminates bushing drillout.

Liner packers can be installed at the top of liners to seal between liner and casing,
after cement placement. Seal elements may be rubber or lead, or a combination of the
two. They may be run as an integral part of the liner hanger and set by manipulation
of the liner-running tool. However, this type of packer should be considered only if
clearance is such that the hole can be circulated at desired rates without increasing
back-pressure excessively.

181
Special packers can be set in conjunction with a tie-back sleeve after cementing and
cleanout operations have been completed. These packers seal both in the tie-back
sleeve and against the suspending casing.

External casing packers have been used on liners to isolate between zones in open
hole. They are inflated following cement displacement before the cements set up to
provide more effective zone isolation.

Setting on Bottom

Except in unusual cases where buckling is not expected or can be prevented through
centralization, liners to be cemented should be suspended from slips set in existing
casing, or the drilling liner. However, equipment is available for the special application
in which liners are cemented and set on bottom.

A special float shoe can be run on the bottom of the liner with an extra internal left-
hand thread. First the liner is run into the well and hung from surface slips. Then the
cementing string is run and engaged into the thread at the shoe. The liner is run to
bottom on the cementing string, and the cement job is completed. The inner string is
disconnected from the shoe by rotating to the right.

Typical Problems in Deep, Hot Wells

Liner cementing is a major problem in deep wells for the following reasons:

Tight Holes. As seen in Table 1 (below), liner-to-borehole clearances can be very small.
This is a highly undesirable situation that frequently results from poor planning, misguided
economics in selecting well tubular/bit programs, or unforeseen downhole conditions.
Small clearances require flush joint liners that approach drill-collar size. Key seats and
collar-worn grooves cause differential sticking and prevent effective mud removal by
cement

Liners, in. Hole, in. Casing, in. Cement


sheath
thickness,
in.
9 5/8 10 5/8 11 3/4 1/2

7 3/4 9 1/2 10 3/4 7/8

7 5/8 9 1/2 10 3/4 15/16

7 3/4 8 1/2 9 5/8 3/8

7 5/8 8 1/2 9 5/8 13/16

5 1/2 6 1/2 7 5/8 1/2

5 6 1/2 7 5/8 3/4

5 6 1/8 7 9/16

182
4 1/2 6 1/8 7 13/16

3 1/2 4 3/4 5 5/8

Table 1: Typical liner/casing and hole size combination.


Long Intervals, Mud Cake. In West Texas, liner lengths may be 2500 to 11,000 ft (762 to
3350 m); they average 8000 ft (2440 m). High temperature and prolonged exposure causes
mud to gel excessively.
Shale instability in long geopressured sections causes hole irregularities. Mud is difficult to
remove from enlarged sections.

Temperature Differential Top to Bottom. As shown in Figure 1 (Long West


Texas drilling liner has high temperature differential over its length), static
geothermal temperature may vary as much as 1200 F over the length of a long
liner.

Figure 1

Cement composition and setup time must be compatible with this gradient. Yet,
circulating temperatures may be radically different, actually causing maximum

183
temperatures to occur perhaps 2000 ft (610 m) uphole for nearly an hour after
circulating.

Gas Cutting, Liner Top Leakage. Premature cement setup uphole by high
temperature or filtrate leakoff to long permeable zones can reduce hydrostatic
pressure and allow gas to permeate the partially cured column. The resulting
channels are too small to effectively squeeze cement, but they continue to
channel gas to liner tops. Two compositions that minimize problems are (1)
fluid-loss additive spotted across upper zones and (2) retarder that keeps slurry
fluid, then sets up rapidly, rather than thickening gradually. Special
compressible cement that maintains its volume during setting to prevent gas
leakage and thixotropic cementing compositions are also available.

Special Primary Cementing Methods


Inner-String Cementing

Cementing large-diameter casing requires some special considerations. Such casing is


subject to being pumped out of the hole. This occurs when pump or hydrostatic
pressure acting on the cementing-head area and on the bottom of the hole through the
shoe opening provides an upward force exceeding the buoyed weight of the casing.
Pressure increase on bumping a plug is, of course, offset, and does not contribute to
the problem.

Large casing can also be floated out of the hole if the weight of the casing and the mud
in the pipe do not exceed the buoyancy provided by the annular column of cement.
The possibility of casing collapse must also be considered. Heavy mud may be required
to prevent these occurrences.

Inner-string or stab-in cementing is now a fairly common practice for large-diameter


casing. The string is cemented through drillpipe stuck into a special sealing sleeve in
the shoe ( Figure 1 , stab-in cementing technique for large-diameter casing).

184
Figure 1

Small-diameter plugs can be used. The drillpipe can be raised from the seal and the
excess cement reversed back.

In geothermal wells where no voids can be tolerated outside the casing because of
later heating and boiling problems, slurry can be continuously pumped until good
circulation to surface is established. Conventional methods, conversely, pump a
calculated volume that is difficult to determine in surface holes without the use of
caliper logs.

Various adaptations are possible using cup packers, etc. with stab-in methods. Port
collars can be opened or closed, and external packers can be inflated to permit stage
cementing of long, large-diameter pipe.

External Cement-Filled Packer

Extra-long (20- or 40-ft) elastomer-sheath covered inflatable packers can be run as


part of the casing string. One packer (or more) is landed across the productive zone to
be perforated and the primary job is completed conventionally.

185
When initial slurry displacement is complete and the top plug is bumped, increased
pressure opens shear-pin controlled valves in the packers and additional cement is
pumped into the packer elements to expand them tightly against the borehole wall.
Inflation cement is pumped down the casing between the first top wiper plug and a
second top wiper plug.

After curing, the cement-filled packer and the casing joint mandrel on which it is run
are perforated by conventional methods. Inflation cement transport and the completed
production system are illustrated in Figure 1 (Inflatable external cement-filled packers
during running and after perforating ).

Figure 1

Two cement-filled packers can be used effectively to straddle a productive interval for
zone isolation.

External cement-filled packers offer several advantages in primary cementing:

Complete mud channel removal is ensured by the application of high internal pressure to
the end-reinforced element, to squeeze mud from the rubber-formation interface.

186
A quality pipe/cement bond is ensured by the use of uncontaminated cement,
mud-free pipe, and pressure setting. Interzonal flow and the need for squeezing
may be eliminated.

Producing zones are supported by the pressure-set cement and will not dilate
or flow when fluid inflow causes a pressure differential.

Without conventional slurry circulation first, cement may be placed solely in


the packers, and thus not contact water-sensitive zones.

System limitations include the following:


Long elastomer-covered packers are durable in properly conditioned holes but are
inherently sensitive to restricted-clearance holes and problem holes with casing burrs,
broken centralizer pieces, nontapered liner shoulders, etc.

Reliable supply and pumping methods are required to prevent long delays or
job interruptions that may complicate inflation-procedure control after
protective knock-off plugs are removed by the bottom wiper plug.

ECD completions may prove less flexible for long-term production adjustments
than conventionally cemented wells.

Factors That Affect Primary Cementing

Hole Conditions

Sloughing

In many cases, this is the reason for setting an intermediate casing. Sloughing can
create several cementing problems: bridging the annulus, sticking the casing, and
increasing the annular hydrostatic pressure.

Drill pipe Drag

The cause and location of the drag could be very significant. Drag may indicate the
need for centralized casing or fluid-loss control cements.

Low-Pressure Zone

One of the most persistent problems is an incompetent formation that will not support
effective columns of cements. This most commonly occurs in intervals covered by
surface and intermediate casing.

Mud Condition

A well-conditioned mud greatly increases the mud-removal capability of flushes and


cement slurries.

Fluid Movement

187
Zone isolation fails any time fluid movement is allowed to occur in a cement slurry
before it is completely set. If the cement moves during the hardening process, it will
not set properly.

Formation Movement

The most common formation movement occurs with salt intrusions.

Mud-Contamination Effects

The possibility of mixing cement and mud always exists during pumping and
displacement. Such contamination can result in

accelerated or retarded thickening times

reduced compressive strength

reduced bond strength

increased filtrate loss (higher than in either mud or cement)

severe thickening (with oil-base mud)

Table 1 (below) shows typical mud additives and their effects on cement.

Inorganic chemicals have an erratic effect on oilwell cements, but generally tend to
accelerate thickening times. Organic chemicals generally retard, and may completely
inhibit thickening in some instances.

Severe thickening occurs with oil muds in cement mixing because these muds are
thickened by water-wet solids that are readily available in the high-solids-content
cement. The problem is most serious when mud and cement slurry densities are high.
Also, oil-emulsion muds often contain calcium chloride in the water phase, which can
accelerate setting.

Additive Purpose Effect on cement

Barium sulfate (BsSo4) Weighing agent Increases density


Reduces strength

Caustic (NaOH, Na2Co3, etc. pH adjustment Acceleration

Calcium compounds Conditioning Acceleration

CaO, Ca(OH)2, CaCl2, CaSO4 pH control


and 2H2O

Hydrocarbons (diesel oil, Fluid-loss control, Decreases density

188
lease crude oil) lubrication

Sealants (scrap, cellulose, Seal against leakage to Retardation


rubber, etc.) formation

Thinners, (tannins, Disperse mud solids Retardation


lignosulfonates,quebracho,
lignins, etc

Emulsifiers lingnosulfonates, Form oil-in-water or water-in- Retardation


alkyl ethylene oxide adducts oil muds
hydrocarbon sulfonates)

Bactericides (substituted Protect organic additives Retardation


phenols, formaldehyde, etc.) against bacterial
decomposition

Fluid -loss control additives, Reduce fluid loss from mud to Retardation
CMC,starch, guar, formation
Ployacrylamides Lignosulfonate

Table 1: Effects of mud additives on cement.

Mud-Contamination Prevention

To prevent mud/slurry problems, it is best to minimize contact. The bottom wiper plug
prevents contamination in the casing, and a spacer fluid reduces cement/mud contact
in the annulus.

Two bottom plugs may be required one preceding, and one behind the spacer fluid
to prevent mud/cement contact if contamination is likely to create serious problems,
and the spacer fluid does not by itself strip the mud film from the casing bore.

A single bottom plug, ahead of the cement, removes the film and accumulated mud
ahead of the plug and behind the spacer fluid. This accumulated mud can then
contaminate the cement ( Figure 1.

189
Figure 1.

, The lack of bottom wiper cases mud accumulation below top plug).

A variety of spacer or preflush fluids are available, including water, brine, solutions of
acid phosphates, diesel oil (weighted or unweighted), oil-base fluids, and emulsions
(oil in water, water in oil).

Compatibility of both spacer and mud, and spacer and cement should be verified on
every cement job. Selection of the amount and type of spacer depends on the type of
mud and on potential reaction problems between the cement and the mud.

A water flush, normally in turbulent flow, may aid mud-displacement efficiency. Salt
water has less tendency to cause shales to swell or slough. However, fresh water, salt
water, or fluids containing dispersing surfactant should not immediately precede a
high-density cement slurry; this can cause thinning and weight material settling

Casing

The following are general rules for casing preparation and application:

Tally all pipe; count, number, and rabbit (gauge) all casing joints on the pipe rack.

Check all casing threads for cleanliness and damage. Additionally, check the
threads on all crossover equipment for proper thread type and cleanliness.

190
Identify all joints by weight and thread type, and place them in proper order
for running into the hole.

Landing joints should be spaced out so the cementing head can be installed
from the stabbing board or rig floor after the casing is landed.

Floating Equipment

The floating equipment (float collars and float or guide shoes) must be on location and
in good working condition. Check operational features if differential floating equipment
is used. Measure and prepare stage and floating cementing equipment separately. Use
thread-locking compound or tack weld (if necessary) on all field makeup connections
between the floating equipment, plus one or two joints above.

Running Casing

The following are general rules for running casing:

Use of a movable stabbing board can minimize downtime.

A safety valve is advised for long production casing strings or suitable


pressure rating.

Control running speed of casing to prevent fracturing and lost circulation.

Circulating Time

The importance of circulation before cementing is recognized by all operators, but


there are considerable differences of opinion regarding optimum circulating time. Many
believe that because the number of variables affecting the success or failure of a
cement job is so great, it does not seem possible to correlate the degree of success
with the amount of time spent in precementing circulation.

The following are circulation guidelines:

Condition the drilling mud with good rates up to anticipated cementing rates.

High circulation rates remove gel led mud that develops during static periods
because of temperature and fluid loss.

Begin pipe movement and mud conditioning immediately after the casing is on
bottom.

Apply scratching technique when wall cake and cuttings in the mud returns
have either virtually stopped or declined rapidly in volume.

The casing moves at the start of circulation and continues throughout the
circulation period. With reciprocating scratchers, the pipe is commonly moved
through a 20-ft stroke, with a 2-min interval for the cycle. If rotating scratchers

191
are used, the pipe is rotated as slowly as possible, usually between 10 and 20
rpm.

Cementing Composition, Volume, and Slurry Weight

In primary cementing, the cement slurry should have a viscosity that will give the most
efficient mud displacement and still permit a good bond between the formation and the
pipe. The following are some cementing guidelines.

Determine the maximum allowable downhole density to prevent fracturing. The density of
cement should be at least 1 lb/gal (preferably 2 or 3 lb/gal) heavier than the drilling mud.

Design fluid loss using differential pressure of 1000 psi. To prevent gas
channeling, design on 20 cc/30 min or less.

Design cement slurry to be displaced in turbulent flow for a minimum of 10 to


20 min contact time at the top of the pay zone, if possible.

For slurries to be placed across salt formations, use saturated sodium chloride.

Use 35% silica at static temperatures above 230 F (110 C).

Control free water to 1% or less for normal slurries. To prevent gas


channeling, control free water to zero.

Determine cement-slurry thickening time at bottomhole cementing


temperature and pressure. Minimum thickening time should be job time plus
one hour of thickening time to a consistency of 50 Bc. (Bearden Units of slurry
consistency are dimensionless units formerly called "poises.") Minimum
thickening time is the time required to mix the slurry, and pump it down the
hole and up the annulus behind the pipe.

Important Tips

Use top and bottom wiper plugs, and inspect the plugs before loading. The bottom
(hollow) plug is loaded first, then the top (solid) plug. Do not slit the diaphragm of the
bottom plug with a knife before loading.

Use a two-plug cementing head:

Displace the top plug out of the cementing head without shutting down operations. Do not
open the cementing head to drop the top plug or a vacuum will be created and the well will
take in air.

Pump preflush or spacer ahead of the bottom plug. If you use two bottom
plugs, put the first bottom plug in first, then the preflush or spacer, and then
place the second bottom plug just before the cement.

192
Before mixing, check calibration of all density devices with fresh water for
proper calibration.

Hook up bulk tanks to the cement mixer. The rate of delivery of cement to
mixer should be sufficient to maintain pump rate in the annulus at the design
rate.

Batch mix all cement slurries by using a ribbon or batch blender. This
operation is extremely important for good control of slurry properties.

A bottom plug is not recommended for use with slurry containing large amounts of lost-circulation
material or with badly rusted or scaled casing. Such material may collect on the ruptured
diaphragm, bridge the casing, and thus prevent total displacement.

Personnel
The supervisor and the person on the pump throttle should understand the importance
of the pumping rate to the success of a job. They both need to know:

why variations in cement density should be held to close limits

control parameters, so they will not be easily satisfied with less control

how to change over from pumping to displacement in 15 seconds, rather than


2 minutes

Maintain a log of operations that includes time, density measurements, mixing rate and
displacement rate, wellhead pressure, operation in progress, volume of fluid pumped, etc. Record
pump speed (strokes per minute) and total strokes. Insist on a properly operating pressure-
recording chart from the operator.

Make sure that all service personnel involved are given ample notice of
commencement of casing operations, so they can be available and rigged up before
this time.

For the following items, prepare and record data on the appropriate company casing
cementing report as required:

Determine the elapsed time and volume or strokes required for the cement slurry to leave
the casing shoe after the start of displacement; to reach the pressure equalization point
after the start of displacement; and to displace the top plug to bump float. Note that when a
stage-cementing collar is used, the above calculations should be made for both phases of
the cementing operation.

Determine the theoretical weight of the casing in 1000-ft intervals. When


differential fillup equipment is used, use all available literature regarding
percent fillup and record the weight at 1000-ft (305-m) intervals during casing
descent.

193
Determine the number of barrels or pump strokes needed to displace the pipe
after the casing is landed and to circulate one full hole volume, and the number
of barrels of mud required to displace the cement.

Estimate the rate of cement mixing and displacement, plus annular velocities.

194
SQUEEZE CEMENTING

Squeeze Cementing

Squeeze cementing is the process of forcing a cement slurry through holes in the
casing. Its primary objective is to create a seal in the casing-wellbore annulus. The
basic components and concept of squeezing are illustrated in Figure 1 .

Figure 1

(concept of cement squeezing ).

The most common purposes for squeeze cementing are:

repair of a primary cement job that failed because of cement bypassing mud
(channeling) or insufficient cement height (fillup) in the annulus

195
elimination of water intrusion from above, below, or within the hydrocarbon-
producing zone commonly called "block squeezing"

reduction of the producing gas/oil ratio by isolating gas zones from adjacent
oil intervals

repair of casing leaks caused by corrosion or split pipe

plugging of all or part of one or more zones in a multizone injection well to


direct injection into desired intervals

plugging and abandonment of a depleted or watered-out producing zone

Basic Concepts and Misconceptions


There have been persistent misconceptions about squeeze cementing, including the
following:

Cement squeezed through holes (perforations) in casing under high pressure generally
forms a horizontal cement pancake opposite the holes, thereby developing a barrier to
vertical fluid movement.

Injecting drilling mud into perforations at high pressure opens all perforations.

High final squeeze is a positive indication of a successful job.

In zones with good permeability, cement penetrates the formation without


fracturing.

These have been disproved by field experience.

Filter Cake

A cement slurry consists of finely divided solid particles dispersed in liquid. Such
particles cannot be displaced into normal formation permeability, since a permeability
greater than 100 darcies would be required to allow a normal slurry to penetrate a
sand formation without fracturing.

Therefore, when slurry is forced against a permeable formation, solid particles filter out
on the formation face as filtrate is forced into the formation permeability. The filter
cake has much lower permeability than most sand formations and, as cake forms on
part of the formation, slurry may be diverted to other exposed zones.

A properly designed squeeze job causes dehydrated cement to fill the opening(s)
between formation and casing and, if allowed to cure, the dehydrated filter cake will
form a nearly impermeable solid.

In cases where slurry is to be placed in a fractured interval (either natural or induced),


cement solids have to develop a cake on the fracture faces and/or bridge the fracture.

196
Most successful squeezes in fractured formations have used a staging technique in
which a highly accelerated slurry, or a slurry with bridging agents such as gilsonite or
sand, is followed by a second stage of moderate fluid-loss slurry. This system
encourages bridging and filter-cake development and helps divert movable slurry to
unsealed fractures.

Fluid Loss

If fluid loss is uncontrolled, cement may dehydrate and bridge off the upper portion of
a perforated interval before slurry is displaced to the lower perforations. Conversely,
very low fluid loss can result in very slow filter-cake development and unacceptably
long placement operations.

Slurry-fluid (filtrate) loss can be varied and controlled with cement additives as
required over the wide range of temperature and pressure conditions normally
encountered in oil and gas wells.

In formations with unimpaired natural permeability, slurry with a water-to-solids ratio


of 0.4 (by weight) and a low fluid loss of 50 to 150 cc in 30 minutes under 1000 psi
differential should provide satisfactory caking for most low-pressure squeeze jobs.

When squeezing against shales, dense limestones, dolomites or permeable


formations where natural permeability is plugged with mud a low-fluid-loss cement
may not be desirable. In these situations, a high-pressure squeeze job is usually
performed, and low-fluid-loss slurry could be undesirable because its restricted filtrate
loss could inhibit filter-cake development.

High-Pressure Methods
High-pressure squeeze cementing is defined as a job in which fluid pressure in the
wellbore exceeds formation-fracture pressure prior to, or during, the time that cement
slurry is in contact with the formation.

High-pressure methods are recommended only for squeezing relatively impermeable


zones, or where squeezing is conducted with drilling mud in the hole.

Fracturing of the formation permits displacement of mud or workover fluid through


holes in the casing. The slurry then displaces this fluid into the fractures, permitting
development of cement filter cake on the fracture surfaces.

Potential Problems

High-pressure squeezes offer no control of either the location or orientation of the


generated fracture. The fracture will be oriented perpendicular to the least principal
stress

Horizontal fractures will not be created if fracture pressure is less than overburden
pressure. Thus, horizontal fractures containing cement pancakes cannot be generated
by high-pressure squeeze cementing in deep wells.

197
Fracturing during high-pressure squeezing may be counterproductive, since fractures
induced in formations deeper than 3000 ft are nearly always vertical. Even if the
casing-wellbore annulus is sealed, vertical communication between zones may be
established in the fracture
( Figure 1 , probable result of fracture-type squeeze-cement job).

Figure 1

Other problems are

large slurry volumes required to fill fractures: 100 to 150 sacks may be lost in a job

resistance of mud-filled perforations to fracturing: many may not readily


receive cement

Recommendations

198
Generally, it is recommended that solids-free workover fluids be used whenever fluid
has to be displaced into the formation ahead of cement. Acid or chemical washes can
also be used ahead of the slurry.

Low-Pressure Methods
Low-pressure squeeze cementing jobs are those in which fluid pressure in the wellbore
is maintained below fracture pressure of exposed formations prior to, and during, the
time slurry is in contact with the formation.

Low-pressure squeeze cementing methods we generally preferred to high--pressure


squeezes, because they are more effective and less potentially damaging to the
formation.

In practice, safe squeeze pressure is usually specified as some value below established
fracture pressure 300 psi has been used in some areas.

Low-pressure squeeze cementing utilizes a small volume of low-fluid-loss slurry placed


against exposed permeable formations with a moderate squeeze pressure. Filtrate
from the slurry is forced into formation permeability, allowing buildup of cement filter
cake. Low fluid loss reduces dehydration rate and discourages bridging as the slurry is
forced along openings or channels.

A properly designed slurry will leave only a small cement filter-cake bump

(node) inside the casing after excess slurry has been circulated out. Improperly
designed slurries can result in excessive caking with enlarged nodes, or inadequate
caking and inability to hold pressure ( Figure 1 , cement filter-cake node buildup after
45-minute squeeze).

199
Figure 1

The casing can be left with cement nodes small enough that drilling-out is not
required. Also, the ability to reverse-out excess cement in many applications makes
low-pressure squeezing compatible with through-tubing techniques.

Field Practices

In low-pressure squeezes, perforations and channels must be clear of mud and other
solids. If the well has been on production, such openings may have been purged. If the
job is to be performed through new perforations, results may be enhanced by
perforating in a solids-free, nondamaging fluid with pressure underbalanced to permit
purging of perforation cavities. In existing perforations, pressure/ suction washing with
or without acid may be considered.

Summary

In practice, the following steps may be used as a guideline for conducting a low-pressure
squeeze job:

200
Initiate injection. Determine downhole injection pressure.

Circulate slurry to desired location in the casing.

Apply moderate squeeze (downhole) pressure.

Restore squeeze pressure by engaging the pump as bleed-off occurs.

Gradually increase downhole pressure to 500 to 1000 psi above the pressure
required to initiate flow. When bleed-off ceases for about 30 minutes, stop
displacing cement slurry and hold the pressure. Do not exceed safe squeeze
pressure.

Reverse-circulate excess cement from casing, or pull work string leaving


cement to be drilled out later, if necessary.

Squeeze Techniques

Hesitation Techniques

The most important principle of hesitation techniques is the alternation of pumping and
hesitation. The hesitation is to encourage cement filter-cake buildup. Hesitation
methods can be used in either high- or low-pressure applications.

Hesitation procedures are much more of an art than a science, since the operator
observes hesitation time and pressure changes during pumping and waiting, and varies
these on subsequent jobs, according to experience. The alternation of pumping and
hesitation is continued until the desired final squeeze pressure is obtained ( Figure 1 ,
example pressure response to hesitation-type cement squeezing).

201
Figure 1

Final squeeze pressure may be misleading. Years ago, high final squeeze pressure was
one primary indicator used to measure success. However, high final pressure may
occur because dehydrated cement has bridged off the casing or perforations; and mud-
cake-filled perforations are also capable or withstanding high differential pressure,
particularly in the direction of the formation. Thus, high final squeeze pressures can be
achieved where the squeeze was unsuccessful.

Bradenhead versus Packer Methods

The Bradenhead squeeze technique is normally used on low-pressure formations.


Usually, the interval to be squeezed is at or near the bottom of the well. The
operational steps of the general procedure ( Figure 2 , Bradenhead squeeze method
applicable to competent casing strings) are as follows:

Circulate cement across the zone to be squeezed.

202
Figure 2

Pull drillpipe (or tubing) above cement.

Close BOPs or annulus valve and apply pressure to cement through drillpipe.

Reverse out excess or WOC and drill out.

Squeeze pressure is limited by casing-string and wellhead-burst strength, so the technique is


usually used with a low-pressure squeeze. It is not a precise cement-placement technique, and is
not generally recommended with several open intervals and only one to be squeezed, or where
casing is not pressure-tight.

Packer-squeeze techniques permit precise slurry placement and isolate high pressure
from casing and wellhead while high squeeze pressures are applied downhole. A packer
squeeze can be conducted with either drillable or retrievable squeeze packers.

Wellbore fluid below the packer is usually displaced through perforations ahead of the
cement when this method is used. Dirty fluid may block flow of cement to a portion of
any exposed permeability. Figure 3 (cement squeeze using retrievable packer and
bridge plug ), is one example of the many tool configurations possible with packer

203
squeezing.

Figure 3

Packer location should be carefully considered and may vary depending on the type of
job. If set too far above perforations or holes to be squeezed, excessive volumes of
either workover fluids or mud must be displaced into the formation ahead of the
cement, or the slurry may channel through the mud ( Figure 4 , possible problems
caused by setting squeeze packer too high).

204
Figure 4

Conversely, a packer set too close to perforations or holes could become stuck if
pressure on the outside of the casing is transmitted above the packer and causes the
casing to collapse.

Usually, the packer should be set 30 to 60 ft (9.14 to 18.28 m) from perforations. If


corrosion holes or split pipe are being squeezed, more space is recommended.

It is desirable to test and then maintain some pressure on the casing annulus above
the packer. Observation of this pressure can be a check for leaks in squeeze string,
packer, or casing. Annulus pressure can also prevent casing-collapse pressure during
high-pressure jobs.

Squeeze cementing in permanent and tubingless completions requires some special


precautions, but basic techniques are similar to those used in conventional wells, and
normally only low-pressure jobs are attempted.

A permanent completion is one in which tubing and welihead remain in place during
well life. Squeeze cementing can be performed with concentric small-diameter tubing.

205
Through-tubing tools such as inflatable bridge plugs and packers can be run on
wireline or small-diameter tubing strings to permit conventional but small-scale
operations.

Squeezing through small tubing uses very small slurry volumes which are susceptible
to contamination. Maintaining accurate volume control is particularly important in
ensuring proper slurry placement.

Job Evaluation

Proof of a successful squeeze comes when we apply pressure to the set cement. It is
best to test the squeeze job before removing the rig, just in case the test fails and re-
squeezing becomes necessary.

Squeeze jobs are most commonly tested by applying pressure from the rig or
cementing unit pumps. A better way to test the squeeze however, is to create a
pressure differential in the wellbore by swabbing, by artificially lifting fluid from the
well, or y circulating a lighter fluid down the tubing and closing the circulation ports
above the packer. The pressure differential should be less than or equal to the
expected drawdown pressure under producing conditions.

In some production wells, it may be impractical to unload the wellbore without


returning the well to production. In these cases, a positive pressure test that does not
exceed formation-fracture pressure should be conducted after cement has set and, if
required, after drillout.

In squeeze jobs where cement is to be drilled out, the way the cement drills is an
indication of success. If it drills hard all the way, results may be good. Soft spots or
voids usually indicate an unsuccessful job.

206
CEMENT PLUGS

Reasons for Setting


By far the most common application of cement plugs - particularly in mature areas - is
well abandonment.

Some other reasons for setting plugs are:

to cure lost circulation during drilling

for directional drilling and sidetracking (or whipstocking)

to provide zone isolation

to provide a seat for openhole test tools

Abandonment

To seal off a dry hole or depleted well, cement plugs are placed at required depths (
Figure 1 , Cement plugs for abandonment).

207
Figure 1

These plugs prevent zone communication and any fluid migration that might infiltrate
underground freshwater sources or cause undesirable surface conditions. Well plugging
regulations are driven to a large extent, by requirements relating to the protection of
underground sources of drinking water. Most states, fro example, require a bottom
plug, a plug across casing stubs, spacer plugs, and perhaps remedial squeezes at
abandonment.

Lost-Circulation Control

If drilling-fluid circulation is lost during drilling, it can sometimes be restored by


spotting a cement plug across the lost-circulation zone
( Figure 2 , Cement plug for lost circulation control) and later drilling back through the
plug.

208
Figure 2

Directional Drilling and Sidetracking

To sidetrack a hole around unrecoverable junk or for an undesirable direction or poor


structural position, a cement plug is placed at a specific depth. This plug helps support
the whipstock for directing the bit into the desired area ( Figure 3 , Cement plug for
directional drilling and sidetracking).

209
Figure 3

Zone Isolation

In a well with two or more producing zones, it is sometimes beneficial to abandon a


depleted or unprofitable producing zone by placing a cement plug above it ( Figure 4 ,
Cement plug for zone isolation). This plug prevents possible production loss into, or
fluid migration from, the lower interval.

Figure 4

Formation Testing

A cement plug is sometimes placed below a zone to be tested that is a considerable


distance from total depth ( Figure 5 , Cement plug for formation testing). This plug is

210
necessary when a straddle packer with sidewall anchor or bridge plug is not possible or
practical.

Figure 5

Job Design
Cement plugging consists of placing a cement column in open or cased hole. Although
this sounds relatively simple, many problems may be encountered during the plugging
operation. Major difficulties include fluid migration, cement contamination, and poor
cement-slurry design.

Nevertheless, cement-plug failures may be minimized by:

using a caliper log to determine hole gauge

211
carefully determining cement, water, and displacement volumes, and always
planning to use more than enough cement

using spacer or preflush ahead of and behind the cement column to minimize
contamination

rotating drillpipe/tubing and using centralizers and scratchers while placing


cement

placing plugs with care, and moving pipe slowly out of cement to minimize
contamination

using diverter tools and viscous pill spacers

In addition to the precautions listed above, there are several other important factors to consider in
designing for a cement plug: slurry composition and volume, government regulations, placement
technique, and well conditioning.

Cement volume needed for a specific plugging operation depends on plug length and
hole diameter. Government regulations may also dictate plug length for well
abandonment. For sidetracking, 200 to 300 ft is generally required to minimize drilling-
fluid contamination that occurs at the plug top. Moreover, allowances are usually made
for dressing off drilling-fluid-contaminated cement before attempting to sidetrack.

Fluid spacers should be used both ahead of and behind cement slurry to minimize
mixing of cement and drilling fluid. Also, spotting a viscous pill spacer at the intended
plug bottom can improve cement-plug stability. Use of a diverter tool, which forces
fluid to flow directly at the wellbore face, provides for a more uniform placement of
spacer, viscous pill spacer and cement slurry. A diverter-tool application to the
balanced method placement technique is illustrated by these cement plugs: Figure 1
(Idealized case ),

212
Figure 1

Figure 2 (Experimental results ) and Figure 3 (Recommended technique ).

213
Figure 2

214
Figure 3

Cement-Slurry Design

Cement-slurry design is closely related to objectives of the plug and also depends on
well depth, BHCT, and drilling-fluid properties. Some key design considerations are

rheology

density and compressive strength

thickening time and WOC time

For lost-circulation plugs, slurry density should be controlled to reduce bulk slurry loss. Bentonite
and silicate extenders are useful slurry-density-reduction additives, and they lower ultimate cement
compressive strength. Gilsonite, a granular hydrocarbon, is also a lightweight cement additive used
in lost-circulation plugs. Low compressive strength is desirable in a lost-circulation cement plug
because sidetracking is more likely to occur off a high-compressive-strength plug. Thixotropic
cement slurries are also used as lost-circulation cement plugs. Self-supporting properties of these
slurries help prevent total cement loss into the lost-circulation zone.

215
A sidetracking or whipstock plug is set to enable a new hole to be drilled away from
the original hole. Since the success of a whipstock plug depends to a large extent on
high compressive strength, use of a densified cement (API Class A, G, or H) is
recommended. Densified cements:

provide high strength

tolerate drilling-fluid contamination

expand upon setting

decrease fluid loss

have negligible permeability when set

provide strong bonding

Drilling-fluid contamination is always possible when placing an openhole cement plug. When drilling
fluid commingles with cement slurry, slurry retardation as well as dilution of the cement plug can
occur. The effect of drilling-fluid contamination on two different density slurries is shown in Table 1
(below).

Because of the properties listed above, densified or reduced-water-ratio cements


generally produce more successful results. Table 2 (below) shows typical compressive
strengths of densified Class G and H cements used for setting sidetracking cement
plugs.

Additives such as calcium chloride, dispersant, or retarder can be used in densified


slurries. However, because slurries with dispersants have lower consistency and
viscosity, the amount of dispersant should be kept to a minimum for better plug
stability. Laboratory testing for maximum viscosity, ample thickening time, and
compressive strength should be performed before each job.

Some operators using whipstock plugs claim that higher compressive strengths and
better success ratios are achieved when 10 to 20% sand and/or silica flour are
included in the cement composition. A possible explanation of this phenomenon is that
sand may improve drilling-fluid removal by its scouring action, and thus may reduce
drilling-fluid contamination. Therefore, sand and/or silica flour may affect compressive
strength in a way unrelated to silica/cement reaction.

To further reduce mud contamination, it is advisable to pump a spacer fluid ahead of


and behind the cement. The density of the spacer fluid should be equal to or greater
than that of the drilling fluid. It is important to note that because of the density
difference between mud and cement. The cement will tend to migrate downward in the
well. This is another reason for including a preflush in the plugging treatment.

Drilling-Fluid contamination Compressive strength (psi) at 230o F


(%) for 12 hours

216
15.6 lb./gal slurry 17.4 lb./gal *

0 2,910 7,010

10 2,530 5,005

30 1,400 2,910

60 340 2,315

* contains dispersant
Table 1: Effect of drilling-fluid contamination on cement compressive strength.

Slurry weight

(lb./gal) 100 F 140 F 170 F 200 F

1,600 psi 3,000 psi 3,000 psi 3,000 psi

After 12 hours

16.5 2,075 4,000 7,800 9,035

17.0 2,850 6,535 8,375 10,025

17.5 3,975 6,585 8,550 10,675

After 24 hours

16.5 5,475 8,985 9,750 10,460

17.0 6,035 9,060 11,075 12,660

17.5 7,025 10,125 11,860 12,875

Table 2: Typical compressive strengths of API Class G and H cements at API curing
conditions.

Mixing

A poorly mixed cement slurry may lead to cement-plug failure, so a high-quality slurry
should be prepared. Because of the small cement volumes associated with plug
cementing, the slurry should be batch mixed or mixed through a ribbon blender, if
possible. These methods lead to a very uniform cement slurry. Should either of these
mixing techniques be impractical, the cement should be mixed at a rate that will
ensure uniform slurry density.

Placement Techniques

217
There are three basic techniques for placing cement plugs: the balanced method, the
dump-bailer method, and the two-plug method.

The Balanced Method

The balanced method involves pumping cement slurry down the drillpipe or tubing until
the level of cement outside the workstring is equal to that inside. The workstring is
then pulled slowly from the slurry, leaving the plug in place. No special equipment is
required for plug placement other than a cementing service unit.

For this method to succeed, the drilling fluid or other well-bore fluid must circulate
freely. It is also important that the well be static, neither gaining nor losing returns,
while the cement plug is being placed or is setting. Density differences between drilling
fluid and cement slurry should be minimized if possible. Otherwise, a viscous pill
spacer and a cement plug should be spotted through a diverter tool. The viscous pill
spacer may be prepared from either bentonite (25 to 30 lb./bbl [71.325 to 85.6
kg/m3] total to a portion of the existing drilling fluid) or downhole intermixing of 20%
calcium chloride and sodium silicate.

The height of the balanced cement column may be calculated by

H = Vc/(Va + Vt)
where (in oilfield units):
H = height of balanced cement column (ft)

Vc = volume of cement slurry (bbl)

Va = volume per foot of annulus (bbl/ft)

Vt = volume per foot of workstring (bbl/ft)

For example, determine the volume of cement slurry required to set a 500-ft cement plug at 8000 ft
in 7-in., 26-lb./ft casing. The workstring is 2 7/8-in. 8.6-lb./ft tubing with an ID of 2.259 in.

The required volume of cement slurry for such a job is 19.1 bbl:

Vc = (500 ft)(0.0382 bbl/ft) = 19.1 bbl

Va = 0.0302 bbl/ft

Vt = 0.00496 bbl/ft

Therefore,
H = (19.1 bbl)/(0.0302 bbl/ft + 0.00496 bbl/ft) = 543 ft
Displacement fluid volume is 37 bbl:
(8000 ft - 543 ft) (0.00496 bbl/ft) = 37 bbl
The Dump-Bailer Method

The dump-bailer method ( Figure 1 , cement plug- dump-bailer method placement) is


usually used in low-pressure, cased holes at shallow depths. Under these conditions,

218
drilling fluid is not normally required, since the well can be controlled with produced
brines.

Figure 1

Unless the well is to be plugged from total depth, a bridge plug is normally placed at
the base of the intended cement plug. Cement slurry is then lowered in a dump bailer
on a wireline. The bailer is opened by touching the bridge plug, and the slurry is then
dumped by raising the bailer.

Two advantages of the dump-bailer method are lower cost and easy control of cement-
plug depth. However, it is not readily adaptable to deep wells, contamination with
drilling fluid is possible, and the slurry quantity is limited to the dump-bailer volume.

The Two-Plug Method

The two-plug method ( Figure 2 , cement plug - two-plug method placement), which
uses a plug-catcher-type tool, offers a means of cement plug placement with less
likelihood of overdisplacement and contamination. Cement plug top is also easily

219
established.

Figure 2

The tool consists of a bottomhole sub installed at the bottom of the workstring,
aluminum tailpipe, bottom wiper plug (which carries a dart), and top wiper plug. The
bottom plug is pumped ahead of the cement slurry to its seat. It cleans the drillpipe
and isolates cement from drilling fluid. A shear pin connecting the dart to the plug is
broken by increased pump pressure and pumped down through the aluminum tailpipe.
The top plug is pumped behind the cement to isolate cement slurry from displacement
fluid. Another increase in surface pressure indicates when the plug has arrived at its
seat. Drillpipe is pulled up until the lower end of the tailpipe is at the calculated depth
for the cement plug top. (Should the aluminum tailpipe get stuck in cement, an
increase in pull will break the tailpipe and free the drillpipe.) The shear pin between
the catcher sub body and sleeve is then broken, allowing the sleeve to slide down and
open reverse-circulation ports.

Some advantages of the two-plug method are:

it provides isolation ahead of and behind the slurry

220
the pipe is cleaned to the bottom of the tailpipe

breakable tailpipe can be readily abandoned if stuck

reverse-circulation is accomplished through the bottom of the tailpipe

it allows accurate placement of the cement plug

it shows positive surface indications

Additional Planning

Regardless of the method, if a cement plug is to be set near the bottom of the
wellbore, sand or gravel can be used to fill the hole to the desired depth. This approach
may be very useful for plugging back to recomplete a higher zone, or for cement plugs
that will be drilled out at a later date.

When drillpipe is used to place a cement plug, centralizers and rotating scratchers may
be placed on the lower end of the workstring to help minimize drilling-fluid
contamination. Scratcher rotation cleans drilling fluid from the wellbore, promoting
better bonding, and allows bypassed drilling fluid to mix uniformly with the cement
slurry. This procedure helps eliminate mud channels in the unset cement.

Testing Cement Plugs


After placing the plug and waiting on cement, the first step in testing the plug is to
'tag' the top of the cement with the drill pipe or tubing. This step is very important,
because the actual top of cement is different from that at which the plug was placed.
The usual test for a cement plug is a performance test. For example, if a cement plug
has been drilled for lost circulation and full returns are observed during further drilling,
the operation is considered to be successful.

It is important to allow adequate setting time before tagging a cement plug. Normal
WOC time ranges from 12 to 36 hours. However, by using a densified cement or an
accelerator, this time may be reduced to between 8 and 18 hours. When the BHST is
above 230 F, 35% silica flour in the cement composition produces a high-strength,
low-permeability cement plug after a minimum placement time.

Operations may resume following primary cementing after the cement has a
compressive strength of approximately 500 psi (3447 kPa). But, when sidetracking,
cement-plug compressive strength must be substantially higher before the plug is
kicked off. In general, for a successful sidetrack, the cement plug must be harder than
the formation. Past experience in a given location should give an indication of the
compressive requirement for a whipstock plug.

Failure of Cement Plugs

Depending on the purpose of a cement plug, failure can occur for the following
reasons:

221
lack of adequate compressive strength (sidetracking)

poor isolation (plug back, abandonment)

wrong depth (all plugs)

the plug is not in place because


it has sunk to the bottom (all plugs)
it has been lost to a thief zone (lost circulation)

Reasons for failure can be traced to the following:

inadequate compressive-strength slurry design

insufficient WOC time

inaccurate BHST

drilling-fluid contamination during cement displacement and POH

cement slurry not designed for the specific problem

inadequate cement volume

too great a difference between cement slurry and well-fluid densities

improper placement technique

222
EVALUATING/TESTING THE CEMENT JOB

Evaluating/Testing Primary Cementing


Evaluation of a primary cement job can determine the existence of one or more of
these potential problems:

failure of the cement to fill the casing-borehole annulus to minimum acceptable height
failure to provide a seal at the casing shoe, or at the top of a liner
failure to provide effective isolation of the zone(s) of interest

If any of these failures is detected, squeeze cementing remedial operations are usually required.

Common evaluation techniques include:

temperature surveys
radioactive tracer logs
cement bond logging
pressure/inflow tests
production testing, production logging

Temperature Logs and Acoustic Bond Logs

Temperature Logs

Temperature surveys are used to detect maximum height of cement in the casing-
wellbore annulus. Although reasonably accurate in this application, such surveys
cannot measure cement quality or its effectiveness in preventing vertical fluid
migration.

The method consists of running a recording thermometer in the casing after the
cementing operation. Setting cement generates "heat of hydration," which increases
the temperature of adjacent fluid in the casing by several degrees ( Figure 1 , idealized
temperature logs in homogeneous lithologies).

223
Figure 1

Maximum temperature anomalies may range from 10 to 40 F (-12 to 5 C). The


magnitude depends on the thickness of the cement behind the casing, as well as the
thermal diffusivity of the surrounding formation. Where lithology is fairly uniform, the
temperature log indicates relative thickness of the cement behind the casing. Caliper
surveys can be particularly helpful in analyzing the temperature survey.

To locate the cement top, the survey should begin either at surface or at least 1000 ft
(305 m) above where the top is expected to be. The survey should be run at 5 F/in.
(6 C/cm) sensitivity under normal conditions. Well conditions must remain static from
the time the plug is bumped until the survey is completed.

The rate at which temperature changes depends on the temperature to which the
cement is exposed. This is usually a function of the depth of the cement job ( Figure 2
, cement-temperature rise as a function of depth and temperature of environment ).

224
Figure 2

Temperatures often peak 4 to 12 hours after the start of mixing, and remain elevated
for more than 24 hours. Therefore, temperature surveys should be run between 8 and
24 hours after the cement is mixed.

Addition of radioactive tracer material to the lead portion of the cement slurry provides
a positive indicator of the cement top. Either long or short half-life material can be
used. Several radioactive materials used as tracers have half-lives of 8 to 80 days.

Principal drawbacks of radioactive tracers include possible health hazards, interference


with natural radioactive surveys, and high cost.

Acoustic Bond Logs

Sonic signals from acoustic cement bond logs are transmitted to a receiver that is
acoustically isolated within a combination tool. In traversing through casing, signal
amplitude is attenuated to varying degrees, depending on the material outside the
casing. Attenuation effect is greater if that material is solid and bonded to the casing.

Signal amplitude is converted to electronic signals and varies inversely with degree of
attenuation. Thus, a high-amplitude casing signal indicates that there is no bond

225
between cement and casing ( Figure 1 , cement-bond log signature and variable-
density comparison).

Figure 1

When cement is firmly bonded to casing and formation, there is a low casing signal,
and the signal received is characteristic of formation behind pipe. When cement is
bonded to pipe but not formation, both casing and formation signals have low
amplitude ( Figure 1 ).

An example bond log is shown in Figure 2 (example response of amplitude and


acoustic signature tracks).

226
Figure 2

This example illustrates a half-wave signature in the right column, which is an


oscillograph picture of amplitudes shown schematically in the previous figure. Other
logs may display the acoustic signature log in the right column as a variable density
log ( Figure 1 ).

CBL interpretation has been a subject of controversy ever since the tools first appeared
in the early 1960's Lack of industry standards for tools and procedures, inadequate
information on log headings, miscalibration of tools, lack of effective tool centering and
poor running procedures have all fueled this controversy. Proper recognition of the
tool's characteristics, however, and an understanding of how hole characteristics affect
them, can help greatly in obtaining a valid log. Properly run and interpreted, the CBL
has been found to provide correctly more than 90% of the time. (Pilkington, 1992).

Downhole conditions that can cause errors in acoustic CBL interpretation (and
recommended preventive measures) follow:

The extent to which the cement is set (hardness) affects sonic signal velocity
and amplitude. As cement hardens, acoustic transmittability increases and
casing signal is dampened out. Therefore, it is best to run CBLs at least 24 to

227
36 hours after the job, or when compressive strength reaches 1000 psi (6894
kPa).

Cement composition affects acoustic transmission. If a high degree of


sensitivity is applied where low-density cementing materials have been used,
poor bonding may be indicated although good bonding actually exists.
Conversely, tests indicate that small sections void of cement can only be
located by high sensitivity. Therefore, voids or channels may not always be
indicated on CBLs unless the proper sensitivity is selected with respect to
cement composition.

Cement-sheath thickness may vary, causing changes in attenuation rate. Lab


tests indicate that a thickness of 3/4 in. (2 cm) or more is required to achieve
full attenuation.

Cement compressive strength and the percent of casing circumference that is


bonded affect CBL amplitude. It is not possible to determine the difference
between a job in which cement strength is lower than anticipated, and one in
which cement strength is as estimated but small mud channels exist. Vertical
zone isolation does not exist in the latter case.

A microannulus is a very small gap between casing and cement. This gap
would affect the CBL presentation although the presence of a microannulus
does not usually prevent isolation between zones, and it usually heals with
time. Running the CBL under pressure can help eliminate the microannulus.

Attention Rate Tools: Attention rate tools, unlike acoustic CBL's are not affected
by fluid travel time. These tools measure amplitude at two receivers located 1
ft. Apart, and then calculate a compensated attention rate. The quantity
measured is the acoustic signal as it travels down the 1 ft. Of casing in the well.
High attenuation rates generally reflect a good pip-to-cement bond. Attenuation
logs share some of the same limitations as CBL, including microannulus
problems and the need to carefully center the tool in the well.

Circumferential Bond Logs: These tools are designed to survey the complete
circumference of the casing, and to detect smaller channels that can be found
using an acoustic CBL. Circumferential tools include multi-pad attenuation-type
tools, as well as tools that employ rotating pulse-echo transducers.

Inflow Tests and Production Testing/Logging

Water Shutoff

Pressure tests are conducted to verify integrity of the casing following primary cement
jobs. Government regulations specify procedures in most locations. The casing-
pressure test is conducted after cement has set but prior to drilling out the cement
shoe. These tests are not an indication of cement effectiveness.

In some locations, regulations require that the casing be perforated and tested by
either bailing or inflow evaluation tests, called water shut-off tests. In California, the

228
tests provide assurance that a cement seal of the annulus exists, to protect shallower
freshwater reservoirs.

Perforating and checking the rate and content of inflow have been used to verify a
cement seal above or below hydrocarbon-producing zones in many areas. The
advantage of this technique is that if a failure is indicated, a cement squeeze of the
holes will be opposite a nonproductive formation.

Casing Seat Test After Drillout

When casing is cemented in an impermeable formation where additional drilling is to


be conducted, the casing shoe can be drilled out and the casing seat inflow tested. The
amount of open hole to be drilled below casing is specified by law in some locations,
but is usually 5 to 10 ft (1.5 to 3.0 m). Tests can be made with a conventional tester
set near bottom.

A pressure test can also be conducted after drilling out the casing shoe and 5 to 10 ft
of open hole. Two objectives of this procedure are to (1) test the effectiveness of the
cement seal at the casing shoe and (2) determine the formation strength (fracture
gradient) of the shoe.

Liner-Top Testing

Pressure testing the overlap to check the seal at the top of the liner prior is usually
preferable to cleaning out the float collar or float shoe. If the liner is not sealed, a
cement squeeze is most easily applied at this time. The overlap pressure tests may use
applied internal pressure to create a differential toward the formation.

Where high formation pressures exist, low-density fluid inside the liner may provide
sufficient differential toward the borehole to indicate leakage. High differential could be
ensured by using a drillstem test (DST) tool. Such differential pressure should be equal
to or greater than the maximum differential expected during future drilling or
production operations.

Production Testing/Production Logging

The most positive evaluation of cement effectiveness has been obtained by production
testing and production logging methods used after completion. These methods include
one or more of the following techniques:

production tests for flow rate and content (water, oil, gas, and solids, if any)

inflow evaluation by determination of flow rates and content versus surface


flowing pressures

evaluation of historical production data, and comparison between wells with


common completions (production surveillance)

pressure buildup and fall-off measurements with downhole pressure recorders

229
inflow versus depth measurements with downhole flowmeters

flowing and/or static temperature versus depth measurements with high-


resolution surface-recording thermometers

flowing fluid density versus depth measurements

downhole fluid samples

radioactive fluid injection and surveys to identify injection points and possible
presence of migration channels behind casing

noise logging to detect behind-casing fluid movement

Generally, a combination of production tests and production logging procedures are required to
identify and locate channels or other problems associated with lack of effective zone isolation.

Documentation/Trouble-Shooting

Primary cementing has been called the critical period of drilling and completion
operations. There is universal agreement that effective primary cementing is a critical
requirement for effective completions.

Perhaps the most important primary-cementing procedure is accurate and detailed


documentation of casing running and cementing operations. This information is
invaluable for evaluating cement jobs, and is essential for the development of
improved equipment and procedures.

230
FLUID FLOW PROPERTIES & MUD DISPLACEMENT

Introduction: Fluid Flow Types

The character of flowing fluid is described by the relationship between flow rate (shear
rate) and pressure (shear stress) that caused the movement.

There are two basic fluid types: Newtonian and non-Newtonian. Newtonian fluids, such
as water, exhibit a straight-line relationship between flow rate (shear rate) and
pressure (shear stress) while the fluid is in laminar flow. They begin to flow when
pressure is applied. As pressure increases, flow velocity increases from laminar,
through a transition zone (part laminar, part turbulent), to fully developed turbulent (
Figure 1 , flow regimes and velocity profile for water-type fluids)

Figure 1

Drilling muds and oilwell-cement slurries are non-Newtonian. These fluids are more
complex; they may exhibit resistance to flow (gel strength) when pressure is applied.
Fluids with gel strength can flow at very low rates in a solid or pluglike manner. Such

231
fluids thus have three flow regimes plug, laminar, and turbulent with transition
zones between each ( Figure 2 , flow regimes and velocity profile for cement).

Figure 2

Extensive study has resulted in the development of mathematical models that can be
used to predict flow properties and pressure-velocity relationships of such muds and
cements. The Bingham plastic model and the Power law model are most commonly
used. The former has been used for drilling-fluid analysis since the mid-1940s. Power
law model equations presented in the late 1950s are generally considered to be more
accurate than those of the Bingham model.

These models attempt to describe the relationship of shear and shear stress for muds
and slurries. Although very useful in analyzing the displacement process, they are not
precise techniques. They should be used to determine flow regime and pressure
requirements for displacement, but results should be considered more qualitative than
quantitative i.e., if analysis indicates a potential displacement problem, believe it; if
it shows acceptable displacement conditions, attempt to enhance displacement
anyway.

232
Flow Properties of Wellbore Fluids and Mud Displacement

Slurry Preparation
The procedure outlined here helps determine two slurry properties: flow behavior index
(n') and consistency index (K'). These factors then allow estimation of frictional
pressure loss and prediction of flow velocity for turbulence or plug flow for non-
Newtonian fluids. It is also advantageous to have n' and K' values for drilling mud in
the hole. When a sample of the mud is available, these data can be determined by the
same viscometer procedure used for cement slurries. When plastic viscosity and yield
point are available, the values can be converted to their equivalent n' and K' values.

Because of the thixotropic, or gelling, tendency of many brands of cement, it is


important that slurry-mixing methods be standardized so that reproducible data can be
obtained. It should also be recognized that many additives used in cement exhibit
time-temperature dependency characteristics somewhat different from those of
conventional slurries such as neat cement and gel cement.

Since the advent of friction-reducing cement additives, and with the increased
application of turbulent flow, a modification of slow-speed mixing may be desirable.
Because emphasis in turbulent flow cementing should be placed on annular conditions
and properties of the slurry in the annulus, it is generally preferable to use Steps 2b or
2c, rather than 2a (shown below):

1. Mix the slurry initially for 35 seconds in a Waring blender by the method described in API
Spec 10.

2.

a. To produce a slurry similar to that existing on the discharge of the displacing


pump, immediately transfer the slurry to a Halliburton consistometer for a 5-minute
period of slow-speed mixing at 80 F (27 C), then proceed to Step 3.

b. To simulate average displacement conditions that influence time-


temperature dependency, transfer the slurry to a Halliburton
consistometer for a 20-minute period of slow-speed mixing at 100 F
(38 C), then proceed to Step 3. This procedure is most commonly used
for slurries containing fluid-loss additives, cement-friction reducers, and
other additives when specific well conditions are not known.

c. To simulate specific well conditions and determine the flow properties


n' and K' at the time of slurry entrance into the annulus, transfer the
slurry either to a Halliburton consistometer or the pressure-temperature
consistometer for a period of slow-speed mixing under a time-
temperature schedule duplicating that expected during placement of the
slurry. In this case, the time is that required for the first sack of cement

233
to reach the bottom of the casing, and the temperature should be the
bottomhole circulating temperature.

3. Pour the slurry from the consistometer slurry container into a Fann V-G Meter cup with a
minimum time lag prior to starting the Fann instrument. Although the slow-speed mixing
period has taken the slurry past its rapid initial gel, gelling tendencies of a lesser order still
exist and may result in variations in data if the slurry were allowed to remain in the static
state between operations. Under conditions in 2c (above) it may be desirable to have the
Fann cup, rotor, and bob at an elevated temperature.
Filling the Fann Apparatus

The slurry to be tested should be poured into the sample cup to the line etched inside
the cup. With the rotor turning at 600 RPM, the cup containing the slurry should be
raised until the liquid level is at the designated level on the rotor (indicated by an
etched line). The instrument should be run at 600 RPM during this operation to ensure
filling of the annular space between the rotor and bob with the slurry.

Shear Tests

Dial readings should be observed at 600, 300, 200, and 100 RPM, and in that order
only. The initial reading at 600 RPM should be taken 60 seconds after the sample cup
has been raised into proper position. The rotor speed should be shifted to the
succeeding lower speeds at 20-second intervals, and each dial reading taken just
before shifting to the next-lower speed.

Interpretation of Data

The following graphs or calculations should be made to obtain n' and K' in conventional
oilfield units.

The flow curve is prepared by plotting on logarithmic coordinate paper, as ordinate (Y


axis)

shear stress (lb. force/sq. ft) = (8-1)

where N is the range extension factor of the torque spring.

As abscissa (X axis), rate of shear (Sec-1) may be determined from the following
table:

Speed (RPM) Rate of Shear (Sec-1)

600 1022

300 511

234
200 341

100 170

When all four data points do not form a straight-line flow curve, the best straight line
through these values should be drawn and extrapolated to the shear stress axis. A
sample plot is shown in Figure 1 (flow curve for non-Newtonian fluid ).

Figure 1

The flow behavior index (n') may be determined as follows:

n' (dimensionless) = slope of flow curve


The consistency index (K') may be determined as follows:
K' (lb.-sec'/ft.2) = intercept of flow curve at unity rate of shear

235
The slope of the flow curve, n' can be calculated from two points on the plotted line, but it is
necessary to use logarithms of the points since the data were plotted on logarithmic coordinates.
The equation below can be used by substituting for the 600 and 300 RPM readings the actual shear
stress from the plot for any two points where the ratio of shear rates is 2.

If only the field model of the Fann viscometer is available (rather than Model 35), the
600 RPM and 300 RPM readings are the only ones that can be made and the following
formulas can be used to calculate n' and K':

(8-2)

(8-3)

For calculations where plastic viscosity and yield point are known:

(8-4)

(8-5)
where:

PV = plastic viscosity, cp

YP = yield point, lbf /l00 ft2

* The multiplier, 1.066, to be used if PV and UP are obtained from dial readings of Fann V-G Meter.

For Newtonian fluids, n' = 1.0, and

(8-6)

Figure 2 (classical Reynolds number friction factor correlation for Newtonian fluids
flowing in turbulence in circular conduits ) is the classical Reynolds number-friction
factor correlation for Newtonian fluids flowing in the turbulence in circular conduits.

236
Figure 2

This correlation is presented for comparison of Newtonian data with the non-Newtonian
behavior of the fluids described in Figure 3 (correlation shown in Fig.

237
Figure 3

2, presented for comparison of Newtonian data with the non-Newtonian behavior of


fluids)and Figure 4 (correlation shown in Fig 2,

238
Figure 4

presented for comparison of Newtonian data with the non-Newtonian behavior of fluids
) These figures were prepared from experimental pressure-drop measurements taken
while cement slurries were pumped through the indicated pipe sizes. The data may be
expected to differ from similar data obtained using other slurries or geometries. The
solid curve in these figures describes the equations commonly used for pressure-drop
calculations and allows a visual comparison of actual data with predicted values.

The Reynolds number transition limits shown in Table 1 (below) were obtained by
procedures designed to minimize the effects of fluid time and energy dependency, as
well as the effect of flow-system geometry, since the laminar flow parameters n' and K'
do not adequately predict these limits. However, the experimentally derived turbulent
flow parameters used for this analysis require such complex equipment that it has
been impractical to obtain such data on all the cement-slurry variations that might be
encountered in practice. The data points shown in Figure 3 and Figure 4 indicate the
deviations encountered, since these points were calculated from experimental data.

Figure no. Slurry Type Reynolds number


transition limits

8.3 Neat API Class A 1800-2500


cement

8.4 12% gel cement 2400-3500

239
Table 1: Reynolds number transition limits.

Formulas for Making Calculations (in Conventional Oilfield Units)

Displacement Velocity

(8-7)

where:

V = velocity, ft/s
Qb = pumping rate, bbl/min
Qc f = pumping rate, ft3/min
D = inside diameter of pipe, in.

For the annulus:


D2 = DO2 - DI2

where:

DO = outer pipe inside diameter (or hole size), in.


DI = inner pipe outside diameter, in.

Reynolds Number

(8-8)

where:

NRe = Reynolds number, dimensionless


V = velocity, ft/s
= slurry density, lb/gal
n' = flow behavior index, dimensionless
K' = consistency index, lb-sn'/ft2
D = inside diameter of pipe, in.

For the annulus:


D = DO - DI

Frictional Pressure Drop

240
(8-9)

where:

Pf = frictional pressure drop, psi


L = length of pipe, ft
= slurry density, lb/gal
V = velocity, ft/s
f = friction factor, dimensionless
D = inside diameter of pipe, in.

For the annulus:


D = DO - DI

Turbulent Friction Factor


- for slurries containing no bentonite

f = 0.0303/NRe 0.1612 (8-10)

- for slurries containing bentonite

f = 0.00454 + 0.645/NRe0.7 (8-11)

Plug and Laminar Friction Factor

f = 16/NRe (8-12)

Velocity at Some Specific Reynolds Number

For generalized calculations:


NRe for plug flow = 100 (maximum)
NRe for turbulence = 3000

(8-13)

where:

V = velocity, ft/s
K' = consistency index, lb-sn'/ft2
n' = flow behavior index, dimensionless
= slurry density, lb/gal
D = inside diameter of pipe, in.
NRe = specified Reynolds number, dimensionless

241
For the annulus:
D = DO - D I

Hydrostatic Pressure

Ph = 0.052 H (8-14)

where:

Ph = hydrostatic pressure, psi


= fluid density, lb/gal
H = height of column, ft

Newtonian Fluids

Turbulent flow for Newtonian fluids such as mud flush, phosphate washes, and water
may be obtained with very little effort. Values have been calculated for the velocity
necessary to obtain turbulence in the annular space for about five different pipe and
hole sizes. In all cases, these values were below a flow rate of one-tenth of a barrel
per minute. Therefore, if any of these fluids are pumped at a rate greater than 0.1 to
0.2 bbl/min in any combination of pipe and hole sizes commonly used, turbulent flow
should result.

Typical Flow-Calculation Problems

The following data consist of four example problems which have been worked out in
detail. These problems are normally encountered in flow calculations using different
cementing compositions, and are (1) primary cementing, (2) multiple tubingless
completions, (3) primary cementing Newtonian fluid, and (4) liner job.

Problem 1 - Primary Cementing

Calculate

1. Pumping rate for turbulence of slurry in the annulus.

2. Frictional pressure drop of slurry in annulus and pipe.

3. Hydraulic horsepower to overcome friction losses.

Well conditions:

hole size 10 in.


casing size 7 in. - 35 lb
depth 5000 ft
slurry neat API Class A cement

(From Halliburton Cementing Tables, Section 210:

242
7-in. x 35 lb casing ID = 6.004)
n' = 0.30
K' = 0.195
= 15.6 lb/gal

1. Pumping rate for turbulence

Equation 8-13 for velocity:

For the annulus:

D = DO-DI = 10 in. - 7 in. = 3 in.

Vc = 10.8 ft/s

Equation 8-7 (rearranged):

For the annulus:

D2 = DO2 - DI2 = 100 - 49 = 51

Qb = = 32.1 bbl/min

2. Frictional pressure drop

In the annulus:

NRe = 3000

Equation 8-9:

243
where D = DO - DI = 10 in. - 7 in. = 3 in.

NRe = 3000

In-casing pumping rate = 32.1 bbl/min

Equation 8-7 for velocity - D = 6.004 in.

Equation 8-8 for Reynolds number:

Equation 8-10:

NRe =6668

Equation 8-9:

244
To make a complete hydraulic analysis of the system, it is necessary to have n', K',
and the density of both the mud in the well and the fluid used to displace the top plug.
After calculations similar to those above have been made on these fluids, it is possible
to make wellhead-pressure calculations dependent upon the location of the cement
slurry in the well, and to estimate what the maximum wellhead pressure will be during
the cementing operation. From this figure, the hydraulic horsepower can be calculated
by the formula:

(8-15)

where:

pressure is in lb/ft2
pumping rate is in ft3/min
HP = 0.0245 x PSI x BPM
HP = (0.0245) (862) (32.1) = 678 HP required

Problem 2 - Multiple-Tubingless Completions

Calculate

1. pumping rate for turbulence in annulus

2. frictional pressure drop in pipe

3. theoretical horsepower required to overcome friction loss in pipe

Well conditions:

hole size 12 1/4 in.


casing size 4 1/2 in., 11.6 lb, 2 strings
depth 12,000 ft
slurry POZMIX A cement + 2% gel + 0.5% retarder

(From Halliburton Cementing Tables, Section 210:

4 1/2-in. casing, 11.6 lb, ID = 4 in.)


n' = 0.23
K' = 0.155
= 14.1 lb/gal

1. Pumping rate for turbulence

245
Equation 8-13 for velocity:

(8-13)

D = equivalent diameter = 4 x hydraulic radius

hydraulic radius =

area of flow = area of hole - 2 (area of pipe)

wetted perimeter = perimeter of hole + 2 (perimeter of pipe)

Qcf = flow rate in ft3/min = VA

where:

V = velocity, ft/min

A = cross-sectional area, ft2

246
V = (7.42) (60) = 445 ft/min

Qc f = (445) (0.597) = 265.7 ft3/min

Qb = (265.7) (0.178) = 47.3 bbl/min for turbulence in annulus

47.3 bbl/min = 23.65 bbl/min down each pipe

Equation 8-7:

Equation 8-8:

Equation 8-9:

From Equation 8-11:

f = 0.00454 + 0.645/(24,888)0.7 = 0.00508

= 5385 psi pressure drop

247
For one pipe:

where:

pressure is in lb/ft2

rate is in ft3/min

or HP = 0.0245 x psi x BPM = (0.0245) (5385) (23.65)

= 3120 HP required per pipe, or 6240 total HP required for both pipes

Problem 3 - Primary Cementing - Newtonian Fluid

Calculate the pump rate required to get an aqueous-base flush into turbulent flow.

Well conditions:

hole size 6 in.


casing size 4 1/2 in. (11.6 lb/ft)
depth 5000 ft

It is desirable to run mud flush ahead of cement.

where:

D = diameter, in.
V = velocity, ft/s
= viscosity, cp
= density, lb/gal

Turbulence begins at NRe = 2100

248
D = DO - DI

= 6.0 - 4.5 = 1.5 in.

For mud flush:

= 1 cp

= 8.33 lb/gal

D2 = DO2 - DI2 = 36 - 20.25 = 15.75

= 0.166 bbl/min required for turbulence in annulus

Problem 4 - Liner Job

Calculate

1. Pumping rate for turbulence of slurry in annulus

2. Frictional pressure drop of slurry in pipe

3. Hydraulic horsepower to overcome friction losses

Well conditions:

hole size 6 3/4 in.


last casing 3628 ft of 7 5/8 in.
total depth 6010 ft
drillpipe 3 1/2 in (11.2 lb/ft) 3577 ft
liner size 5 1/2 in. (15.5 lb/ft) 2433 ft
slurry neat API Class A cement

(From Halliburton Cementing Tables, Section 210:

5 1/2 in. - 15.5 lb casing - ID = 4.95 in.,


3 1/2 in. - 11.2 lb drillpipe - ID = 2.90 in.)

249
n' = 0.30
K' = 0.195
= 15.6 lb/gal

1. Pumping rate for turbulence

Equation 8-13 for velocity:

For the annulus:

D = DO - DI = 6.75 in. - 5.50 in. = 1.25 in.

From Equation 8-7:

where:

D2 = DO2 - DI2

= (6.75)2 - (5.5)2 = 15.3

= 11.24 bbl/min necessary to obtain turbulence in annulus

2. Frictional pressure drop in drill pipe

Equation 8-7 for velocity:

250
Equation 8-8 for Reynolds number:

Equation 8-10:

f = 0.0303/(10,690)0.1612 = 0.00679

Equation 8-9:

= 2677 psi

Frictional pressure drop in 5 1/2-in. liner

Equation 8-8 for velocity:

Equation 8-8 for Reynolds number:

From Equation 8-12:

f = 16/2039 = 0.0079

Equation 8-9:

251
= 146 psi

Total frictional pressure drop = 2677 + 146 = 2823 psi

Hydraulic horsepower = 0.0245 x PSI x BPM

= 0.0245 (2823) (11.24) = 780

Mud-Displacement Principles
Mud conditioning is the single most significant factor in obtaining effective mud
displacement. Findings of studies in wellbores indicate that fluid loss control is
important in preventing excess filter cake; and that immobile mud filter cake cannot be
completely displaced by the cement slurry even under turbulent flow conditions.

In short, "there is no substitute for maintaining drilling fluid properties that enhance
the mobility of the mud, enabling displacement by the cement slurry (Halliburton)."

Two basic forces associated with drilling-mud displacement during primary cementing
are differential pressure and cement-on-mud (fluid-on-fluid) drag forces. To displace
muds effectively, oilwell cements must exert a combination of differential pressure and
drag forces of sufficient magnitude to overcome forces resisting displacement. Here we
discuss properties that influence these two forces.

Figure 1 (forces acting on bypassed mud column to resist or cause displacement by


slurry ) illustrates a bypassed vertical mud channel and displacing/resisting forces of
flurry and mud, respectively.

252
Figure 1

Introduction to Fluid Flow Types

The character of flowing fluid is described by the relationship between flow rate (shear
rate) and pressure (shear stress) that caused the movement.

As previously mentioned, there are two basic fluid types: Newtonian and non-
Newtonian. Newtonian fluids, such as water, exhibit a straight-line relationship
between flow rate (shear rate) and pressure (shear stress) while the fluid is in laminar
flow. They begin to flow when pressure is applied. As pressure increases, flow velocity
increases from laminar, through a transition zone (part laminar, part turbulent), to
fully developed turbulent ( Figure 2 , flow regimes and velocity profiles for water-type
fluids).

253
Figure 2

Drilling muds and oilwell-cement slurries are non-Newtonian. These fluids are more
complex; they may exhibit resistance to flow (gel strength) when pressure is applied.
Fluids with gel strength can flow at very low rates in a solid or pluglike manner. Such
fluids thus have three flow regimes plug, laminar, and turbulent with transition
zones between each ( Figure 3 , flow regimes and velocity profiles for cements)

254
Figure 3

Extensive study has resulted in the development of mathematical models that can be
used to predict flow properties and pressure-velocity relationships of such muds and
cements. The Bingham plastic model and the Power law model are most commonly
used. The former has been used for drilling-fluid analysis since the mid-1940s. Power
law model equations presented in the late 1950s are generally considered to be more
accurate than those of the Bingham model.

These models attempt to describe the relationship of shear and shear stress for muds
and slurries. Although very useful in analyzing the displacement process, they are not
precise techniques. They should be used to determine flow regime and pressure
requirements for displacement, but results should be considered more qualitative than
quantitative i.e., if analysis indicates a potential displacement problem, believe it; if
it shows acceptable displacement conditions, attempt to enhance displacement
anyway.

Factors Affecting Cement-Mud Drag Forces

255
Resisting drag forces exist at contact planes between mud and borehole wall, and
between mud and casing. When casing is not centered, resisting drag-force effects are
not uniform across the annular flow area. This difference increases with
decentralization, and increases the likelihood of bypassing mud on the narrow side of
the annulus.

An indicator of the degree of decentralization is percent standoff, and investigations


have shown that standoff increases the velocity required to initiate mud flow from the
narrow side of the annulus ( Figure 4 , effect of decentralization on velocity in narrow
side of annulus). A stand off of at least 60 percent is recommended, and 70% is
preferred. These may be hard to achieve in horizontal wells.

Figure 4

In Figure 4 , the effect of decentralization is indicated as a ratio of fluid velocity in the


narrow portion (Wn) to average fluid velocity. For example, with 50% standoff, fluid
movement at 110 bpm average rate is zero on the narrow side, and never exceeds

256
60% of average at any higher rate. (Note: percent standoff = 100 x Wn/borehole ID
radius - casing OD radius.)

Contact time is the period during which a position in the annulus (generally above the
zone of interest) remains in contact with a cement slurry that is in turbulent flow. If
cement-on-mud drag forces are high enough to cause mud erosion, and contact time is
long enough, complete mud removal should be achieved. However, those conditions
are most likely to exist when cement in turbulent flow has adequate contact time with
mud having a significantly lower velocity.

The resisting drag force between mud and casing can be altered to a positive mud-
displacing force by rotating the casing while displacing cement. This positive effect is
illustrated in Figure 5 , (Pipe rotation aids mud removal on narrow side of annulus ).

Figure 5

Reciprocation moving casing up and down exerts a somewhat less positive


displacing-drag force. However, reciprocation also affects velocity of cement and mud.

Improving Mud Displacement


Assuming an effort has been made to drill the hole properly, the following practices
should contribute to optimum primary cementing:

Center the pipe in the borehole.

257
Move the pipe during mud conditioning and cementing.

Know formation-pressure limits in the borehole.

Condition the mud.

Prevent cement/mud reactions.

Control displacement rates and slurry rheology.

Additional discussion of these guidelines follows.

Centering Pipe

Centering pipe in the borehole creates a uniform annular flow area perpendicular to
flow direction, and minimizes variation of resistive drag forces across this flow area.
Centralizers do not provide perfect casing/borehole concentricity. They do substantially
improve standoff conditions, however, since a casing without a centralizer lies against
the borehole wall.

The use of centralizers is strongly discouraged under certain conditions by some


drilling personnel. Generally, their concern is that the device will hang up, and prevent
the casing from being run to the desired depth. Unfortunately, the conditions that
generate the greatest concern such as highly deviated wells with numerous
washouts are often those that require centralizers. In some cases, centralizers can
actually increase the chances of properly running the casing.

Rotating versus Reciprocating

Either type of pipe movement alters drag effects between mud and casing from a
resistive displacement force to a positive one. However, model studies have shown
that rotation appears to be more effective than reciprocation for removing bypassed
mud where casing is severely off-center ( Figure 1 , pipe rotation aids mud removal on
narrow side of annulus).

258
Figure 1

Rotation provides more effective cement/casing drag forces than reciprocation, as it


seems to pull the cement into the bypassed mud column. Rotating at 15 to 25 rpm
provides more pipe movement relative to annular fluids than does reciprocating 20 ft
(6.1 m) on a 1-minute cycle.

Reciprocating can cause lateral casing movement, or changes in standoff, as


centralizers move across wellbore irregularities. This lateral movement alters the flow
area and encourages bypassed mud displacement.

Pressure Velocity Surges

Reciprocal movement also affects flow rate and velocity of fluid in the annulus. During
the upstroke, velocity in the annulus decreases, because part of the fluid pumped out
of the shoe occupies the volume previously occupied by the casing.

On the downstroke, the casing acts like a piston, displacing fluid in the wellbore below
the shoe up the annulus, along with the volume of fluid being pumped through the
shoe. This motion creates substantial pressure and velocity surges in the well-bore (
Figure 2 , Measured pressure surges while lifting and running one casing joint smoothly
into the hole.), which improve the erosional effect of cement on bypassed mud by
substantially increasing displacing-drag forces.

259
Figure 2

However, it is important to know the magnitude of pressure changes to avoid breaking


down the formation and causing lost circulation.

Fracture-gradient information is important in the determination of safe slurry density


and/or pump rate, and of whether stage equipment is needed. A profile showing
fracture gradient is helpful; however, accurate profiles are not always available.
Indications of fracture gradient for a given area are obtained through evidence of lost
circulation during drilling and records of breakdown pressure encountered during
stimulation and squeeze operations.

Conditioning Mud Before Cementing

Reducing gel strength and plastic viscosity greatly improves displacement efficiency
and reduces the pressures required at the cement/mud interface to displace mud. It
also reduces displacement drag forces required to erode and remove bypassed mud by
reducing resistive drag-force effects.

Under certain well-defined pressure-window limits, it may be desirable to lower mud


density, along with gel strength and plastic viscosity, nearly to the minimum wellbore-

260
pressure limit. This would permit a larger pressure increase for displacement purposes.
If this is done, pipe should only be rotated (and not reciprocated), to prevent a
swabbing action that may reduce pressure to below the lower limit.

In most cases, mud circulation should be adequate for cleaning up the hole and
removing cuttings from the mud if favorable mud properties are maintained while
drilling the final portion of the hole.

Displacement Rate and Rheology

High displacement rates improve displacement efficiency if cement can be in turbulent


flow up the annulus. Conditions that may prevent such flow include:

limited displacement-rate capability (pumping equipment)

breakdown (pressure-window) restrictions that limit pressure

improper flow (rheological) properties of mud and/or slurry

If wellbore conditions can tolerate high displacement pressures, the provision of extra pumping
equipment is basically an economic decision. Formation conditions that determine the pressure
window are fixed, and attempts to exceed those limits may create serious problems.

Dispersants versus Turbulence

Fluid properties of the slurry can also be altered i.e., dispersants can be added to
lower gel strength to attain turbulent flow at lower rates. This can be desirable where
high pump rates would otherwise be required. By adding dispersant and lowering
pump rate, an increase in effective contact time can be effected, along with the desired
velocity profile. However, if turbulence can be achieved at reasonable pump rates
without dispersants, the resulting displacement should be better; i.e., turbulent flow is
better than laminar flow, but additional turbulence may not be even more effective.

Once turbulent flow is established, displacement efficiency increases with increased


slurry-flow resistance, as displacing drag forces increase with increasing contact
pressure at the cement/mud interface. Thus, thinning the slurry to increase turbulence
is not recommended.

Buoyancy Effect

The buoyancy effect of higher-density cement slurry on lower-density mud is


controversial. Such effects should provide a positive displacing force on bypassed mud
as long as there is vertical continuity of the mud column to the top of the rising
cement/mud interface.

Contact pressure at the base of the bypassed cement/mud interface increases with
increasing height of cement. This increased contact pressure near the bottom of the
bypassed mud column should increase both displacing pressure and erosional effects.
However, if the cement bypasses a portion of the mud and then reestablishes complete

261
displacement of the movable mud in the annulus above the bypassed mud, displacing
drag forces may be the only effective force working to remove the mud.

Plug Flow

When wellbore conditions are such that turbulence cannot be achieved, displacing with
cement in a plug-flow regime can maintain a flatter velocity profile in the annulus.
Although the resulting drag forces are not as effective as those of turbulence, they can
be maximized by increasing cement gel strength as much as possible, particularly in
the lead part of the slurry. Also, cement density can improve plug flow displacement
when it is maintained at least two pounds per gallon heavier than the mud.

Improved Laminar-Flow Displacement

Wellbore and/or surface conditions that prohibit turbulent flow may also prohibit plug
flow. When these uncommon circumstances exist, an alternative is to alter cement
rheological properties to increase apparent slurry viscosity.

Even in laminar flow, displacement can be effective if the slurry is thicker (i.e., has
higher yield strength and plastic viscosity) than the mud, and if sufficient volumes are
used to obtain the desired cement height on the narrow side of an eccentric annulus.

One guide for cement rheological design is to have cement yield strength exceed mud
yield strength by a factor equal to maximum annulus clearance divided by minimum
annulus clearance.

262
COMPLETION EQUIPMENT

TUBULARS

API Specifications for Oilfield Tubulars

The American Petroleum Institute (API) has defined certain standards for oilfield
tubular goods, such as tubing and casing. The API has defined ten grades of steel:
H40, J55, K55, C75, L80, N80, C90, C95, P105, and P110. The number indicates the
API minimum yield strength in thousands of psi. The letters H, J, and N are primarily to
minimize verbal confusion, while the others have an additional meaning:

K has higher ultimate strength than J

C, L "restricted yield strength" with tighter specifications

P high strength

The behavior of tubular goods under stress conditions is a basic problem in strength of
materials. The API has developed a set of standard formulas that are used throughout
the oil industry to predict the minimum load-carrying capacity to be expected from a
particular grade and weight of pipe (API Bulletin 5C3). Tables of casing and tubing
strengths based on the formulas are also published by the API (Bulletin 5C2) and in
various manufacturers' and service companies' handbooks. Remember that the API
formulas are modified from time to time and it is important to make sure that the
performance data used is taken from the most recent version.

The major failure modes that we are concerned with are

burst

collapse

tension failure of the coupling or pipe

There is always some debate as to whether the API formulas are the best theoretical
basis for computing a particular strength parameter (e.g., for burst, a modified
Barlows equation is used instead of Lame). However, each company's assessment of
the conservative nature or inadequacies of the API formulas is generally reflected in
the design factor and design assumptions that they apply in using the API Strength
Criteria.

Tubing Design Concept

The uncertainties regarding actual loading conditions and the state of the tubing (e.g.,
corrosion, anomalies due to poor handling) considerably exceed our analytical
capabilities to determine the resultant stresses. The tendency in the oil industry

263
therefore has been not to be overly sophisticated in analyzing an extremely complex
system, but rather to make designs on the basis of a set of idealized loading conditions
that have proven adequate in the past, such as those presented in Table 1. It is
important to remember that each company has its own philosophy, criteria, and design
factors to consider. The balance of design assumptions versus actual conditions is
depicted in Figure 1 (The balance of design assumptions versus actual conditions).

Figure 1

While this may lead to a tendency to overdesign, the relative cost of the convenience is
generally fairly small. Extreme caution should therefore be used in making
modifications to the idealized loading assumptions. For special, severe loading
conditions (e.g., ultra deep >20,000 ft (6000 m), very high pressure >10,000 psi (70
MPa), very hot >300 F (150 C)), it is necessary to make a detailed computer-
assisted stress analysis.

Condition Loading Design Criteria Typical Design Factor

Burst Internal Kill pressure on 1.125

264
hydrocarbon-
filled tubing

External Packer fluid and


zero annulus
pressure

Considerations Check effects of


compression

Collapse External Casing head 1.125


pressure= shut-
in tubing
pressure

Internal Tubing empty


and
depressured

Considerations Check effects of


tension

Tension Running Buoyant weight Body: 1.333


in completion
fluid

Joint: 1.8*

Tension and Operating Cold stimulation Body: 1.125


Compression and hot
production
conditions

Joint: 1.333

Considerations Check effects of


temperature
and pressure
changes

Total Stress Triaxial Max. stress 80% yield

*Assuming separate checks are not planned on shock and bending effects; otherwise
use 1.5.

Table 1: Typical criteria for tubing design on a flowing well

In many field situations and preliminary estimates, to establish the weight and
strength of the tubing it is sufficient simply to look at the tubing rating and to apply
the corporate design factor. However, it must be recognized that loading conditions
vary over the length of the tubing string, and to properly visualize this it is generally
advantageous to carry out a graphical tubing string design.

265
Graphical Tubing String Design

This is a convenient way of understanding loading conditions and presenting design


results. The technique is presented in Example 2 (part 1) and illustrated in Figure 2
(Graphic tubing design estimated operating pressures),

Figure 2

Figure 3 (Graphic tubing design burst loads), and Figure 4 (Graphic tubing design
tubing selection).

266
Figure 3

Abbreviations are presented in the Nomenclature.

267
Figure 4

Example 2 (part 1)

Graphical Tubing Design

Planning Data

KBE: 3000 ft (915 m)

TD: 11,500 ft (3500 m)

Tbg: 2 7/8 in. OD (73 mm)

Closed-in bottomhole 5500 psi (38 MPa)


pressure:

estimated from mud weight

268
Formation breakdown 12,500 psi (86 MPa)
pressure:

estimated from offset well

Fracture propagation pressure: 9200 psi (63 MPa)

estimated from offset well

Packer fluid: inhibited oil (0.38 psi/ft)

Production: expect sour gas

(gas gravity = 0.80 reservoir)

(gas gravity = 0.70 separator)

J55 or L80 tubular to be used

Stimulation: fracture expected (assume 20 barrels per minute); maximum


allowable annulus pressure is 2000 psi (13,790 kPa)

THP Estimate

Depth of Hole Gas Gravity

(ft) (m) 0.60 0.65 0.70 0.80

1000 305 .979 .978 .976 .973

2000 610 .959 .956 .953 .946

3000 915 .939 .935 .930 .920

4000 1219 .920 .914 .907 .895

5000 1524 .901 .893 .885 .870

6000 1830 .883 .873 .854 .847

7000 2133 .864 .854 .844 .823

8000 2438 .847 .835 .823 .801

9000 2743 .829 .816 .804 .779

10,000 3048 .812 .798 .764 .758

11,000 3353 .795 .780 .766 .737

12,000 3660 .779 .763 .747 .717

269
13,000 3962 .763 .746 .729 .697

14,000 4267 .747 .729 .712 .678

15,000 4572 .732 .713 .695 .659

16,000 4876 .717 .697 .670 .641

17,000 5181 .702 .682 .652 .624

18,000 5486 .687 .656 .645 .607

19,000 5791 .673 .652 .631 .590

20,000 6097 .659 .637 .615 .574

Table 2: Ratio between surface pressure and bottomhole pressure in gas wells for a
range of gas gravities

At a gas gravity = 0.8, CITHP = 0.727 CIBHP = 3999 psi

At a gas gravity = 0.7, CITHP = 0.757 CIBHP = 4164 psi

For a kill situation:


bottomhole injection pressure = CIBHP + 2000 psi = 5500 psi + 2000 psi = 7500 psi

If gas gravity = 0.8, THIP = 0.727 BHIP = (0.727) (7500)

=5453 psi

Assumed Fracture Conditions


1. Formation breakdown achieved with water

2. Fracture job carried out with water-base fluid

Friction loss in 2 7/8 in. tubing at 20 BPM using water with friction reducer is 350 psi/1000 ft for
11,500 ft (Dowell Handbook)
FPP = 9200 psi

Friction = +4025 psi (350 11.5)

Head = -5175psi (0.45 11,500)

Frac THP= 8050 psi

Prepare a depth pressure plot ( Figure 2 ) in the following manner:


1. Plot the closed-in bottomhole pressure (CIBHP).

2. Plot the formation breakdown pressure (FBP) and the fracture


propagation pressure (FPP).

270
3. Plot the packer fluid gradient, fracture fluid gradient, and water
gradient.

4. Estimate wet and dry gas gradients and plot these up from the
closed-in bottomhole pressure.

5. Establish the closed-in tubing head pressure for normal production


conditions (i.e., oil or, as in this case, wet gas) and for worst case design
assumption (usually dry gas).

6. Establish maximum THP for which completion is to be designed, which


normally will be kill or stimulation conditions (fluid gradient through FBP,
FPP, or specified differential above CITHP). For Example 2, the graphical
design should now look like Figure 2 .

7. Establish through inspection the greatest differential pressure at


surface and downhole (usually stimulation conditions). Determine what
steps can be taken to reduce loading (e.g., maintaining maximum
allowable annulus pressure during stimulation). Plot adjusted annulus
pressure line ( Figure 3 ).

8. Plot burst load line (BLL) as difference between most critical tubing
and annulus pressures. The BLL is a function of the relative densities in
the tubing and annulus. BLL will generally, but not always, decrease with
depth ( Figure 3 ).

9. Plot critical collapse load conditions (CLL). Normally we assume that a


slow leak has changed the CHP to CITHP and that tubing is empty and
depressured. This can occur in gas wells if the tubing becomes plugged
or a downhole safety valve is closed. Conditions can approach this
situation in oil wells after a fracture treatment if operators commence
kickoff before bleeding off annulus pressure. (In some cases this may be
a more critical load ( Figure 4 ).)

10. Plot pressure test conditions (PT). This is often the most critical load
to which a completion is subjected. Consider timing of the pressure test
and density of fluids in the tubing and annulus at time of test.

11. Look up tubing performance data in API Bulletin 5C2.

12. Adjust API internal yield (burst) and collapse resistance


specifications with design factor (see Figure 1 and API Bulletin 5C2).

13. List resulting tubing capabilities ( Figure 4 ).

14. Compare design loads with tubing capabilities and select tubing. In
most cases the optimum tubing grade and weight will vary with depth.
To minimize costs and/or tensional loads, such variations may be
incorporated, although there will then be a constraint on pressure-
testing capabilities. However, most operators prefer to use a common

271
weight and grade throughout the completion, if possible. This reduces
the risk of installation and operating errors. When regulations permit,
the designer may be able to compromise slightly on accommodating
loading conditions deep in the hole, if the associated design assumption
is extremely unrealistic (e.g., a completely empty tubing in a high
productivity oil well). However, the designer must first check on how
critical the actual biaxial (or triaxial) loading conditions are likely to be
and make appropriate notes in the well file.

With reference to Example 2, in Figure 4 the options include the following:


1. full string of 6.4 lb/ft L80 tubing

2. 0 to 6500 ft = 6.4 lb/ft J55


6500 to TD = 6.4 lb/ft L80

3. full string of 6.4 lb/ft J55 with modified collapse design criteria of
2000 psi as maximum CHP with an empty tubing

Since 2000 psi is the maximum allowable annulus pressure during stimulation, option 3 may be an
acceptable design. Since the differential cost of J55 and L80 is around $3 per ft, the potential
saving of $34,500 between options 1 and 3 may justify further detailed engineering work. On the
other hand, if the wellstream is expected to be extremely corrosive, the higher grade tubing may be
selected in any case to provide a corrosion allowance.

The key things to note from Figure 5 (Effect of buoyancy on axial load) are

the most severe burst loadings occur at surface

272
Figure 5

the most severe burst and collapse loadings occur during pressure testing,
well kill, and stimulation

the most severe collapse loading occurs downhole

additional annulus pressure can be used to reduce burst loading, provided the
casing is strong enough

the tubing-head pressure during kill operations (THIP) often approximates or


exceeds the reservoir pressure (CIBHP)

With relatively small tubing strings (<3.5 in. or 90 mm), the inherent burst and collapse strength is
so high that some engineers do not bother with tubing design in wells with depths of less than 8000
ft (2500 m), unless overpressures are expected.

Simplified Tensional Strength Design

273
Although burst and collapse resistance may not be significant considerations in
pumping wells, tensional strength is a critical design parameter for all wells. Coupling
leakage and failure, which accounts for 80% of the problems in well tubulars, often
may be the result of inadequate tensional design rather than a burst or sealing
problem. In this respect, it is particularly important to remember that test pressures
impose substantial piston forces on the tubing (e.g., a 2000 psi (13.8 MPa) pressure
test on a plug set inside 2 7/8-in. (73-mm) tubing will increase the tension on the
hanger by (2000)( /4)(2.44l)2 = 9360 lb (42 kN).

It is also important to recognize that, unlike other strength parameters, the API joint
strength is based on a failure condition rather than the onset of plastic deformation.
The failure condition is either an unzipping of the pin and box in the case of API
threads, because of yielding (also called "jump-out"); or breakage of reduced cross
section at the threads in the case of square threads.

Finally, there are all sorts of additional tensional loads that we do not normally analyze
in detail (e.g., shock loading and drag forces during running, bending stresses,
buckling, cross-sectional piston forces, changes in buoyancy).

Since it is common practice to make a preliminary tensional design using tubing weight
loading only, a higher design factor is used for tension and especially for joint strength
(Table 1, above). Some companies and more conservative engineers will even ignore
the potential benefits of buoyancy. Buoyancy results in a piston force on the lower end
of the tubing and as a first approximation it may normally be assumed that

(12)
where:
WB = buoyant weight

WN = weight in air

= density of steel ( 8 gm/cc)

= density of fluid

Figure 5 graphically depicts the tensional and compressional forces at work on a tapered string of
tubular goods. The load resulting from the weight of the pipe is shown for a string weighed in air,
and with the buoyant forces accounted for as piston forces or approximated using Equation 12. We
can see that at a point approximately midway in the length of the heavier pipe at the bottom of the
string there is a change from compression to tension. This is also the concept which guides the
design of drillstrings with the purpose of keeping the drillpipe in tension while using the heavier drill
collars to maintain a compressional load on the bit.

Part 2 of Example 2 gives the preliminary tension design considerations for the
completion already covered in part 1.

Example 2 (part 2)

274
Preliminary Tension Design

Tubing weight: 6.4 lb/ft

Tubing length: 11,500 ft

Packer fluid: inhibited oil


0.38 psi/ft = 0.88 gm/cc

WN = 6.4 11,500
= 73,600 lb

= 0.89 73,600

= 65,504 lb

Joint Specifications

J55 L80

EUE HYD CS EUE HYD A95

API joint strength 99.7 100 135.9 128


(Klb)

Design factor 1.8 1.8 1.8 1.8


(Table 1)

Design capacity 55.4 55.6 75.5 71.1


(Klb)
Tubing Tension Design Considerations
1. Requires L80 tubing at surface

2. Requires joint strength capability of EUE or equivalent

3. In view of pressures, depth, and H2S would probably select premium


grade coupling

Many companies have these design techniques programmed for the computer and use the same
general technique for both tubing and casing designs.

Tubing Design Parameters

It is important to remember that while the primary function of the tubing is as a


conduit for hydrocarbon production or for injection of water or gas, the most severe
loadings often occur during well service or killing operations, or during pressure tests.
It is therefore prudent to make provision for these operations when designing a
completion, and to check out the tubing limitations when planning a well servicing

275
operation (e.g., a stimulation or a workover) . Care must be taken not to increase
completion costs excessively by trying to make provisions for all sorts of unlikely, but
possible, occurrences. It must also be remembered that there are steps that can be
taken to mitigate the induced stresses during many operations (e.g., applying annular
pressure or heating fracturing fluids). On the other hand, the consequential costs of a
failed tubing string, or of having to run a special working string, in terms of deferred
production and rig time, can be quite substantial. Assessment of the most cost
effective solution is generally a judgment call based on the engineer's experience and
on corporate attitudes and policy. A typical set of parameters has already been
illustrated in Example 2.

Burst

The tubing and wellhead should be designed for squeeze and kill conditions. Since fines
in the perforations or oil can sometimes cause a "check valve" effect when attempting
to squeeze back liquids, many completion designers like to have the flexibility of being
able to raise the bottomhole pressure to the FBP or at least to the FPP. However, with
high permeability reservoirs or gas wells in which fracture stimulation is unlikely,
completion engineers are often satisfied with a certain minimum differential for
injection. The value selected varies from area to area and from company to company,
but is commonly either around 1000 psi (7 MPa), or 33% of the reservoir pressure.
The author suggests

1. FBP where k1 < 100 md


kg < 50 md

2. FPP for squeezing liquids, where k1 > 100 md

3. CIBHP + 1000 psi (7 MPa) for squeezing gas, where


kg > 50 md; or for squeezing liquids, where k1 > 1000 md

From the rock mechanics theory presented by Geertsma (1978) and others it may be deduced that
in a tectonically relaxed area, a provisional estimate of the fracture propagation gradient (FPG) can
be obtained from the equation

(13)

FPG < FBG < 1.1 psi/ft (25 kPa/m) (14)

where:
sv = overburden stress ( ~1 psi/ft depth)

p = pore pressure, psi

D = depth, ft

FPG = formation propagation gradient, psi/ft

FBG = formation breakdown gradient, psi/ft

276
The specification of the pressure test conditions is often critical to burst design. Government
regulations sometimes specify pressure test conditions (e.g., to at least 90% of the reservoir
pressure or to 1000 psi (5 MPa) over the maximum differential pressure expected at the packer). If
no regulations exist, most operators test to their tubing design conditions.

Collapse

Severe collapse loads on the tubing can occur

in gas wells and high GOR oil wells with low-flowing bottom-hole pressures and deep-set
safety valves, after blowdown to test a plug, etc.

during annulus pressure tests, or operation of shear circulation devices

where there are pressured annuli

during underbalance perforating or testing at high drawdown

during tubing blowouts

It is important to remember that tension reduces collapse strength. This biaxial effect should be
examined for large diameter tubings, especially if reduced collapse design assumptions and/or a
deep-set safety valve is used.

Tension

Tubing strings are not only subjected to running tensions with all the associated shock
and acceleration loadings, but also to varying operating stresses due to piston forces
on the steel and/or any plugs, pumps, standing valves, and the like in the tubing.
Moreover, if the tubing is anchored or held by a packer, its operating tension will vary
as a result of

thermal effects (hot production or cold kill fluid)

piston effects (changes in buoyancy and forces at joint upsets)

ballooning effects (changes in internal or external pressure)

buckling effects (longitudinal instability)

These potential problems are listed in terms of their most common relative magnitude (although the
relative importance of piston and ballooning effects is variable).

Combined Loading

While the designer of tubular goods normally talks in terms of burst, collapse, and
tension compression as if they were independent, it is obvious that in most actual
loading situations they occur simultaneously. Precise stress analysis should really
consider a triaxial loading situation.

277
The simultaneous solution of all the associated equations is rather complicated. A
number of computer programs are available, but for most field engineers they will be a
"black box" solution. This can be dangerous. It is important to check that the formulas
are properly handled, particularly with respect to collapse, which is a "stability effect."
Therefore it is usual for critical stress analyses (e.g., for ultra deep, high pressure, or
sour wells) to be undertaken by a specialist consultant, research group, or
intracompany task force. Moreover, since this is not the routine design technique,
design factors are less well proven (although a value of 1.25 is often used).

A more convenient approach for the intermediate range, moderately complex design
problem is to use the ellipse of biaxial yield stress proposed by Holmquist and Nadai
(1939). The critical relationships are (a) tension reduces collapse resistance; (b)
compression reduces burst resistance.

The other important concept in the consideration of triaxial loads is that pressure
changes affect axial stresses or cause tubing movement. This has been extensively
discussed in SPE papers by Lubinski (1962), Hammerlindl (1977), and Stillebroer
(1967).

Bending

Bending stresses can be significant in large tubulars. They are compressive in the inner
wall and tensional in the outer wall, the most detrimental being

(15)
where:
R = the radius of curvature (ft)

sb = bending stress

E = Young's modulus (for steel, E = 30 106 psi)

do = outside diameter of the tubular

Bending stresses result from both hole curvature and buckling. The effects of doglegs need only be
considered if they are very severe (>10/100 ft; 10/30 m) or if very large tubing (5 1/2 to 7 in.; 140
to 178 mm) is being used.

Production Casing

The production casing must be adequately sized for the planned completion. It will
obviously affect the size of the other required casing strings, the bit selection, the
capacity of the rig, and the overall well costs. The production casing must be designed
for the loads that may be imposed during the producing life of the field. It is similar to
tubing design in several ways.

278
Burst

Production casing must be designed to withstand the maximum closed-in tubing


pressure that can be expected. If a packer has been used, this pressure is assumed to
be applied at the top of a full column of packer fluid (i.e., for the case of a tubing
failure at the surface). We usually assume that the external pressure resisting burst is
a water gradient. If the packer fluid is heavier than water the burst load will increase
with depth.

In the event that a snubbing operation could not be conveniently attempted if a tubing
break occurs at the surface, the casing must be strong enough to withstand a bullhead
squeeze on the live tubing string, in which case this would be the design criteria for the
casing and wellhead.

In many cases it may be necessary to design the casing for loads imposed during
stimulation and pressure testing. Conversely, the casing capacity must be checked
when designing a fracturing treatment. This is particularly important in wells where no
packer is used.

Collapse

Production casing may be subject to complete evacuation during production operations


if the well is operated on gas lift or pumped-off, or if the packer or workover fluid is
lost into a depleted zone. As the pipe may have deteriorated before this occurs, a
higher design factor (1.125+) is often used for production casing.

Severe collapse loads may occur in situations in which thermal expansion of the
annular fluid between the production and intermediate strings cannot be bled off (e.g.,
in some subsea wells).

Increased loading should be assumed if live annuli are a feature of the area. Reduced
loadings may be assumed if the wells will not be pumped off, gas lifted, or severely
depleted.

Severe collapse loads may exist in the pay section during high drawdown,
underbalanced perforating and testing, and squeeze either or both cementation and
fracturing ( Figure 1 , Collapse loads in the pay). It is highly advisable to maintain
some set casing/tubing annulus pressure during such service operations.

279
Figure 1

Tension/Compression

In high rate production areas and thermal wells, expansion of the production tubing
may impose additional tension on the casing strings, via the packer.

Couplings

In high pressure (>5000 psi; 34 MPa), high temperature (>300 F; 422 K) and/or
severely sour conditions, premium casing couplings are recommended.

Material Selection

In sour environments, material specification must consider the chances of H2S


contamination of the casing/tubing annulus and the added possibility of temperature
changes during stimulation affecting the stress corrosion tolerance of the pipe.

Couplings

280
There are many forms of coupling available, some of which have been dedicated to the
public through the auspices of the API, while others are produced by, or under license
from, a specific manufacturer. It was, in fact, the need to obtain a standardization of
thread forms and diameters that led to the formation of the API Committee on
Standardization of Tubular Goods in 1924.

The API couplings ( Figure 1 ,

Figure 1

Figure 2 and Figure 3 ,

281
Figure 2

Cutaways of basic types of couplings) are of three basic types: external upset (EUE),
nonupset (NU), and integral joint.

282
Figure 3

Threads are of two main forms. The round API threads are weaker than the pipe body
(i.e., <100% efficient). The buttress threads were developed by the National Tube
Division of the United States Steel Corporation to provide a high strength coupling for
deep, high pressure well service. This thread is used in a number of proprietary
couplings, e.g., Hydril, VAM, Atlas-Bradford.

All tapered threads achieve a seal by driving the pin and box surfaces together under
sufficient stress to generate a bearing pressure exceeding any differential that is to be
subsequently applied. However, a small spiral void is always left between the mating
surfaces, and must be filled with solids in the form of thread compound. (This is the
reason for careful specification of the compound in Bul 5 A2.) On API round threads,
this void occurs between the crest and root of the mating threads, while on buttress
threads it extends over the whole flank of the thread on the beveled side.

To improve leak resistance, especially at elevated temperatures and under high


pressure differential, the so-called "premium" seals were developed. These consist of
either metal-to-metal seals on tapered portions of the pin and box surfaces or an
elastomer seal ring, or both. This type of seal requires a high quality finish and precise

283
gauging and inspection. The coupling is therefore more costly. Since API and buttress
threads have proven to be very reliable in the field, the decision to use the more
expensive "premium" seals requires careful economic justification. In general, their
application has proven valuable in highly corrosive conditions in high pressure gas
wells or high pressure/high GOR oil wells, and in thermal wells subject to high
compressive loads. They may be used where workover costs are high (e.g., offshore).
In general, API tubulars are adequate for differential pressures of less than 5000 psi
(34.4 MPa) and temperatures of less than 300 F (150 C), using high temperature
thread compound. For corrosive conditions and continuous gas service, the pressure
limit is often reduced to 2500 psi (17 MPa).

284
PACKERS

Packer Functions

A packer is a subsurface tool that provides a seal between the tubing and
casing,thereby preventing the vertical movement of fluids across this sealing point.

Packers are used for the following reasons:

to improve safety by providing a barrier to flow through the annulus

to keep well fluids and pressures isolated from the casing

to improve flow conditions and prevent heading

to separate zones in the same wellbore

to place kill fluids or treating fluids in the casing annulus

to pack off perforations rather than use squeeze cementing

to keep gas lift or hydraulic power fluid injection pressure isolated from the
formation

to anchor the tubing

to install a casing pump

to minimize heat losses by allowing the use of an empty annulus or thermal


insulator

to isolate a casing leak or leaking liner lap

to facilitate temporary well service operations (e.g., stimulations, squeezes)

Packer Types

There are many packer manufacturers, some of whom offer an extensive variety of
packers, with each differing to some degree from those of the other manufacturers.
This rather bewildering array can, however, be grouped into principal classes or types,
and may be further categorized by method of setting, by direction of pressure across
the packer, and by the number of bores through the packer.

Packers can be primarily classified as either retrievable, or permanent, or permanent-


retrievable, or inflatable.

Retrievable Packers

285
This type of packer is run on the tubing ( Figure 1 , Retrievable packer).

Figure 1

After setting, it can be released and recovered from the well on the tubing. Since it is
an integral part of the tubing string, the tubing cannot be removed from the well
without pulling the packer, unless a detachable packer head is used.

Retrievable packers may be designed to be set mechanically or hydraulically.


Mechanical setting methods include rotation of the tubing string, reciprocation of the
tubing string, or the application of tension or set-down weight. With mechanical
packers, the tubing is usually set in compression.

Hydraulic packers are set by applying hydraulic pressure through the tubing string, but
once set they hold the set position mechanically. The tubing is usually in tension.

Retrievable packers are usually used for complex multizone and multistring
completions. Their main limitation was in their limited ability to accommodate tubing
stress changes without unsetting; the availability of effective slip joints and detachable
heads has eased this situation. An historical problem was failure of the internal

286
elastomer seals, but this technology also has improved markedly in the last decade. All
metal-to-metal seal packers are available but are expensive.

One disadvantage of retrievable packers is that if they fail to retrieve, they must be
removed by milling them out of the casing with an abrasive milling head and a
drillstring. They are very difficult to mill. Generally, this type of packer is used under
nonsevere conditions (differential pressures less than 5000 psi (34.4 MPa),
temperatures less than 300 F (422 K).

Because of the setting mechanism, retrievable packers tend to have a restricted bore,
compared with other packers designed for the same casing size. This factor may
restrict flow or limit wireline operations below the packer depth.

Permanent Packers

Permanent packers are independent of the tubing and may be run on tubing or on
wireline ( Figure 2 and Figure 3 ).

Figure 2

287
The tubing can be released from the packer and can be pulled, leaving the packer set
in the casing.

Figure 3

Tubing can subsequently be run back and resealed in the packer. The packer may thus
be considered an integral part of the casing. It is sometimes called either a production
packer or a retainer-production packer.

The permanent packer cannot be recovered as such, but it can be destructively


removed (e.g., by milling) ( Figure 4 and Figure 5 ).

288
Figure 4

If the packer includes a tailpipe and must be recovered, a millout extension is needed
on the packer for the "packer picker," or catch sleeve, on the mill to engage. In other
cases, it may be adequate to simply push the packer to the bottom of the casing after
milling.

289
Figure 5

Permanent packers can be set using an electric wireline setting tool, a hydraulic setting
tool run on drillpipe or tubing, or by a combination of rotation and pull.

Permanent packers are typically used when

formation, treating, or swabbing differential pressures will be high

it is desirable to pull the tubing without unseating the packer

it might be desirable to convert the packer to a temporary or permanent


bridge plug

high bottomhole temperatures exist

tubing operating stress variations would not be accommodated with a


retrievable packer without making it impossible to pull

a retrievable packer would have an inadequate bore

290
Permanent-Retrievable Packers

A recent arrival, this type of packer has the same characteristics as the permanent
packer, but it can, when desired, be released with a special pulling tool and recovered.

Inflatable Packers

These are packers with a flexible sealing element that can be expanded hydraulically
using either completion fluid or cement. They are used as openhole packers, or when
the casing is buckled or collapsed, preventing the usage of conventional packers.

Inflatable packers cannot stand high pressure differentials and are generally limited to
special applications, such as drillstem testing.

Packer Failures

The major causes of so-called "packer failures" relate to

use of the packer outside its operating range

unsetting of the packer or seal assembly as a result of pressure or


temperature changes

using or setting the packer incorrectly

the packer being in poor condition when run

Difficult operating conditions require more expensive equipment and more rigorous
design work. It is important to remember that a packer is effectively a piston within
the casing and will therefore be heavily affected by changes in differential pressure.
Pressure from below or increasing string tension due to tubing contraction, or both,
tend to unset the following:

weight-set packers

hydraulically set packers not equipped with holddown buttons

locator seal assemblies or overshots

Pressure from above tends to unset tension-set packers or cause collapse-type failures
or leakage at seal assemblies under high pressure differentials.

Tubing/Packer Forces and Movement

Changes in temperature and pressure inside and outside the tubing affect the length of
the tubing string (if the string is designed to permit motion) and the forces exerted at
the packer (if no motion is permitted).

291
These changes in tubing length or force between producing and pump-in conditions
can be large, and should be considered in choosing a packer. This is especially
important in high temperature (usually deep) wells, and may limit the use of
retrievable packers. Design of slip joints and seal assemblies must also consider these
forces.

(The following section draws heavily on the work of D.J. Hammerlindl (1977) and Arco
as published in JPT February 1977, and on that of Arthur Lubinski, W. S. Althouse, and
T. L. Logan (1962).)

Factors Causing Packer Forces or Tubing Movement

If the tubing string is free to move, its length will change as a result of temperature
and pressure influences, which may be subdivided into thermal, piston force,
ballooning, and buckling effects. Consideration of these effects will determine the seal
length required and/or slip joint design.

If the tubing is anchored to the packer, these effects will result in a change In the axial
tension in the tubing. This can affect not only the design of the uppermost tubing joint,
but also packer shear pin rating and the degree of buckling above the packer and
therefore the through-bore access. A mechanical force is also involved in this situation.

To consider these effects, it is necessary to define the critical conditions to be


examined. These normally include:

landing conditions;

operating conditions;

shut-in conditions;

killing/stimulating conditions;

pressure test conditions.

Initial calculations are made for a set of assumed landing conditions, e.g.,
-50,000 to +50,000 lb (-225 to +225 kN) tension (-) or compression (+). For convenience a designer
may sometimes choose to examine only the differences between what are considered to be the
most severe conditions (i.e., hot producing to cold stimulating); however, these are not always
apparent (e.g., pressure test loads can be the most severe).

Mechanical Forces

Mechanical forces can be subdivided into tension and compression. Tension results in
the stretching of tubing string. The elongation due to tension forces can be determined
by Hooke's law, which states that the change in length is directly proportional to the
applied force. The equation for Hooke's law in this application is as follows:

292
(16)
where:
= length change (inches)
L = tubing length (inches)
Ft = tensional force (-) (lbf)
E = Young's modulus (30 106 psi)
As = cross-sectional area of the tubing wall (in2)
This relationship is the basis of the tubing stretch tables and graphs published by various
equipment manufacturers.

Temperature or Thermal Effects

The length change due to change in temperature is equal to the length of the tubing,
times the coefficient of thermal expansion for steel, times the change in average
temperature:

Since

(17)
then
(18)
where:
= change in tensional force in tubing at the surface due to temperature change
As the coefficient of thermal expansion for steel () is 6.9 l0-6/F (12.42 l0-6/C) and Young's
modulus (E) is 30 x 106 psi (207 GPa):
= -207 As T lb/F

Ft = -2.57 As T N/C

In most cases, it is adequate to deal with the change in the average string temperature. It is often
assumed that
the completion is at geothermal gradient when landed (1 to 2 F/100 ft, or 1.8 to 3.6
C/100 m)

during stimulation/killing it will stabilize within 20 F (10 C) of ambient


temperature

The temperature during producing conditions is a function of flow rate, gas expansion, geothermal
gradient, thermal insulation of the tubing, and the like. Various computer programs are available to
calculate this in critical cases, but it is usually adequate to use data from offsets producing under
similar conditions. Production test data should be used judiciously since test rates and flow periods
are often too low and too short for thermal stabilization to occur. As a first approximation designers
will often assume a flowing gradient for high rate wells of 0.4 F/100 ft, or 0.7 C/100 m.

Piston Force Effects

293
The most familiar form of piston force is that of the stretch and/or stress induced when
making an internal pressure test against a plug set inside a string of tubing.

The force against the plug is equal to the pressure applied times the cross-sectional
area of the tubing ID:

Fp = pt Ai (19)
For example, a 10,000 psi (69,000 kPa) pressure test on 3 1/2-in. (89 mm) tubing (I.D. 2.992 in.)
should result in a force of

Fp = 10,000 (2.992)2
= 70,000 lb (311,000 N)
Where the tubing is inserted into a packer we must similarly consider the piston effects on the cross
section of the steel as it is affected by changes in the internal and external pressure.

The piston force at the packer related to a change in the inside and outside pressures
is

Fp = (Ap - Ai) pi - (Ap - Ao) po (20)


where:
Ap = area of packer bore

Ai = area of tubing ID

Ao = area of tubing OD

The change in tubing length related to this piston force is

(21)
where the pi and po changes are considered positive if they correspond to an increase and
negative if they correspond to a decrease in pressure ( Figure 1 , Schematic showing piston force at
the packer related to a change in the inside and outside pressures).

294
Figure 1

The piston forces are, in essence, the change in the buoyant force on the tubing.

Ballooning Effects

As the pressure inside the tubing increases, the pipe expands radially. This will cause
an axial shortening. Application of pressure to the annulus will cause the tubing
diameter to contract (reverse ballooning). This will result in a tubing elongation.

The length change accounts for the change in radial pressure forces due to surface
pressure changes (pis and pos) and fluid density changes (i and o) as well as flow
inside the tubing ( ). The formula for calculating the change in length due to
ballooning is

295
(22)
where:
= Poisson's ratio of the material (for steel = 0.3)

R = ratio OD/ID of the tubing


= drop in pressure in the tubing per unit length due to flow. The
pressure drop is positive when the flow is downward and zero when
there is no flow.
i = density of fluid inside tubing

o = density of fluid outside tubing

pis = surface tubing pressure

pos = surface annulus pressure

Buckling Effects

Buckling is caused by two effects: applied longitudinal compression loads and internal
pressure.

The first is easy to understand as a logical consequence of the loading of a column.


However, the buckling of a pipe under tension as a result of internal pressure is a
difficult concept to appreciate. One way to visualize this type of buckling is to consider
a slightly banana-shaped tubing joint subject to internal pressure ( Figure 2 and Figure
3 , Causes of buckling).

Figure 2

296
There will be a small unopposed area upon which the pressure will act to cause
distortion of the pipe.

Figure 3

The resulting buckling will reflect the balance of strain and bending energy. This is
similar to having a "fictitious force" (Ff) acting on the end of the pipe.

To better understand buckling resulting from compression loading, consider a string of


tubing freely suspended inside the casing. Now consider an upward force, F, applied to
the lower end of the tubing. This force compresses the string and buckles the lower
portion of the string into a helix. The neutral point is where the buckling stops. This
force decreases with increasing distance from the bottom of the string and becomes
zero at the neutral point. The distance, n, from the bottom of the tubing to the neutral
point is calculated from the following formula:

(23)
where:
W = Ws + iAi - oAo, representing the buoyed weight of the tubing per unit length
F = the force applied to the lower end
It should be understood that in the presence of fluids the neutral point is not the point at which there
is neither tension nor compression, but it is the point below which the string is buckled and above
which the string is straight.

If the neutral point is within the string, then the shortening of the string due to
buckling (Lb) is as follows:

297
(24)
where I is the moment of inertia of tubing cross section with respect to its diameter:

I= (D4 - d4)

When the calculated value of the neutral point is above the upper end of the string, the
entire string buckles into a helix.

When the pressure inside the tubing (Pi) is greater than the pressure outside (po) at
the packer, a shortening will occur due to helical buckling. It was determined by
Lubinski, Althouse, and Logan (1962) that, as far as the buckling is concerned, the
tubing behaves as if subject to the following "fictitious" force:

Ff = Ap(pi - po) (25)


Because part of this force, Ff, appears to be nonexistent, the entire force, F, was given the name
fictitious. Further, it was determined that the string would buckle if Ff is positive and would remain
straight if Ff is zero or negative. Substitution into Equation 24 gives

(26)
where Lb is the change in length with respect to the length of the tubing when landed with pi = po

If the pressure outside the tubing is greater than the pressure inside the tubing at the
packer (po > pi), there is no helical buckling due to pressure.

The total change in tubing length as a result of these various factors (mechanical and
piston forces, and thermal, ballooning, and buckling effects) may be expressed as

L = Lm + Lt + Lp + LB + Lb (27)
The effect of this net change in tubing length on the tubing stress will depend upon the amount of
motion permitted by the packer design.

Permanent Corkscrewing

If the buckling results in the yield strength of the tubing being exceeded, permanent
corkscrewing can occur. In addition to the stresses of helical buckling, the tubing is
subject to elongational stresses, as well as tangential and radial stresses due to
pressure inside and outside the tubing. A triaxial stress analysis must be made where
conditions suggest that significant tubing stress will result. All the major equipment
suppliers have computer and hand calculator programs available for making these
computations. This service is usually available free to purchasers. The production
engineer should spend some time determining the pressure and temperature
conditions to be expected. Typically, a series of calculations is made for various landing
assumptions and the results plotted for analysis. It is also useful to make at least one
hand calculation as a cross-check on the data received.

298
ARTIFICIAL LIFT EQUIPMENT

Anchors

Unanchored tubing in a rod-pumped well will be subject to constant movement. The


tubing will buckle on the upstroke and stretch on the downstroke. This is sometimes
called "breathing." This movement leads to wear and fatigue problems and can result
in inefficient use of the available pumping unit stroke. Tubing anchors are used to
minimize this movement.

Tubing anchors are classified as follows:

tension anchors, which permit the tubing to elongate but not to shorten

compression anchors, which permit shortening but not elongation

fixed anchors

Compression anchors reduce the breathing problems but do not prevent buckling and
are therefore rarely used. Tension anchors, which gradually "walk down" the inside of
the casing as the pump starts to operate and then set the tubing at its maximum
elongation, may damage the casing by repeated slight movement. Therefore, most
anchors used today are of the fixed type (often miscalled tension anchors) ( Figure 1 ,

299
Figure 1

Mechanically set anchor, Figure 2 ,

300
Figure 2

Dual hydraulically set anchor, and Figure 3 , Single hydraulically set anchor).

301
Figure 3

There are two main setting mechanisms:

rotation set

hydraulic set

It is best to select an anchor with a back-up retrieving mechanism so that if the


primary one (i.e., rotation) fails, a secondary release (i.e., shear pins) can be used.

Some hydraulic anchors depend only upon the differential pressure between the tubing
and casing. These have an additional application in preventing seal assemblies or
packers from becoming unlatched during stimulation operations.

Bottomhole Pumps

Details of the bottomhole pump for rod-pumped wells are set out in API Spec 11AX,
which includes a 12-character code to specify each pump type. The most critical is the

302
second group which is the pump bore (125 to 275 corresponding to 1 1/4 in. to 2 3/4
in. or 32 to 70 mm) and the pump type (R-rod and T-tubing).

The pump displacement in oil field units can be obtained from

PD = Ep 0.1166 S Es N D2 (28)
where:
S = stroke (in.); say, 74 in.
Ep = pump efficiency; assume 90%
N = pump rate; say, 20 SPM
Es = stroke efficiency (or rod stretch); assume 80%
D = pump bore; say, 2 in.
SN < 1500 in./minute (maximum desirable for rod fall)
Using the numbers given above, the pump rate would be
PD = 497 B/d (79 m3/d)

303
DOWNHOLE COMPLETION ACCESSORIES

Seating Nipples

There are three main types of seating nipple used as integral parts of the tubing
string:

pump-seating nipples

selective landing nipples

nonselective or no-go landing nipples

Seating nipples, which are used to accommodate a pump, plug, hanger, or flow control
device, consist of a polished bore with an internal diameter just less than the tubing
drift diameter. Usually a lock profile is also required, especially for landing nipples.
Heavy duty tubing sections, called flow couplings, are often run on either end of a
seating nipple to minimize the effects of turbulence ( Figure 1 , Landing nipple and flow
coupling installation).

304
Figure 1

Seating nipples and the devices that are set inside them are used for the following
purposes:

to facilitate pressure testing of the bottomhole assembly and tubing couplings,


and the setting of hydraulic packers

to land and seal off a bottomhole pump (pump seating nipple)

to isolate the tubing if it is to be run dry for high draw-down perforating

to land wireline retrievable flow controls, such as plugs, tubing safety valves,
bottomhole chokes, and regulators

to plug the well if the tree must be removed

to land bottomhole pressure bombs

to pack-off across blast joints

to install a standing valve for intermittent gas lift

to plug the tailpipe below packer in order to pull the tubing without killing the
well

to temporarily plug the well while the rig is moved on or off the well

Selective Landing Nipples

Selective landing nipples are nipples with a common internal diameter. In some, the
lock profile is varied for easy identification ( Figure 2 ,

305
Figure 2

Figure 3 ,

306
Figure 3

Figure 4 and Figure 5 , Landing nipples and locking mandrels).

307
Figure 4

Others are accessed by tripping the lock mechanism at the selected depth.

308
Figure 5

Selective nipples are used when more than one nipple is required within a single string
of tubing, and the designer wishes to maintain maximum throughbore. They should be
no closer than 30 ft (10 m) from a similar profile, and at least 10 ft (3 m) from any
change in diameter.

No-Go Landing Nipples

No-go landing nipples are designed with an ID that is slightly restricted to provide a
positive shoulder to locate a locking mandrel. The ID of these nipples should be
checked against the dimensions of any through-tubing equipment that may be used.
This type of nipple is usually located at the bottom of the tubing string or tailpipe and
at least 5 ft below any profile change.

In tailpipe installations, it is best to include a sliding sleeve above the nipple in case
debris prevents the pulling of any plug set in the nipple by regular wireline methods.
Alternatively, a mechanical perforator may be used to punch a hole above the plug.

Sliding Sleeves

309
Also referred to as sliding side doors or circulating sleeves, these tubing components
are used to obtain access from the tubing to the tubing/casing annulus either for fluid
circulation or to permit a previously isolated zone to be produced ( Figure 1 , Sliding
sleeves). They are opened and closed with a wireline tool that has a locating key that
engages the profile in the sleeve. A TFL version is also available for subsea
completions.

Figure 1

These devices are typically placed above each packer in the well. Obviously they are an
essential requirement of multizone completions scheduled for selective production.
Many producers run sliding sleeves in each string in a multistring completion to
increase production flexibility.

A sleeve above the upper packer is particularly useful for the following operations:

kick-off by displacing the tubing contents with a low density fluid, thereby
avoiding the use of coiled tubing within the tubing

310
well killing prior to a tubing pulling job or workover

circulating out completion fluid with a packer fluid (e.g., from mud to brine or
from water to inhibited brine)

testing of subsurface safety valve (SSSV)

temporarily producing a selective zone into the tubing so it can be tested or so


a bottomhole pressure survey can be obtained.

The quality of the elastomer seals in sliding sleeves has improved greatly over the last
decade. They are now much easier to open and less prone to failure. Special
elastomers are needed for some well fluids and suitable design procedures are now
available for elastomers.

A ported nipple is sometimes used in place of a sliding sleeve, although this makes it
necessary to pull the tubing string in order to stop annular access. Alternatively, some
completion engineers prefer to use a side pocket mandrel and valve as a circulation
point above the packer. Note, however, that side pocket mandrels offer a reduced area
to flow and restrict circulation rates.

Side Pocket Mandrels

Side pocket mandrels are a special eccentric nipple that can accommodate a valve in
parallel to the tubing to control access to the annulus ( Figure 1 , Side pocket
mandrels). They are used to install wireline retrievable gas-lift valves, circulation
devices, flow control valves, and injection valves.

311
Figure 1

The location of side pocket mandrels for gas-lift valves will be determined by the lift
gas pressure available and kickoff requirements.

It is highly desirable to have one or two mandrels located just above the top packer in
high pressure gas well completions. These are used to facilitate a controlled circulation
kill in the event the sliding sleeve is inaccessible or if corrosion-inhibitor injection is
required. Inhibitor may be supplied either through the annulus or via a special control
line, continuously or in batch treatments. Some operators also use side pocket
mandrels to install a pressure and temperature sensor that can transmit data to the
surface via a cable attached to the outside of the tubing.

Some engineers prefer to use a side pocket mandrel instead of a sliding sleeve above
the top packer, since the elastomer seals on a side pocket circulation valve are easily
retrieved and redressed using wireline, while repair of those in a sliding sleeve requires
a workover. However, most circulation valves have a limited throughput capacity (0.5
b/m or 5 m3/hr) and some operators therefore have a tendency to pull the valve to
increase circulation capability. This can result in a cutting out of the valve seat in the
mandrel, which inevitably requires a workover to replace the mandrel.

312
Blast Joints and Flow Couplings

Blast joints and flow couplings are special joints having the same nominal ID as the
tubing, but a greater OD. They are usually manufactured from special heat-treated
steel ( Figure 1 ,

Figure 1

Blast joints, Figure 2 ,

313
Figure 2

Polished nipples, and Figure 3 , Schematic of polished nipple run to provide sealing
surface in case of blast joint erosion).

314
Figure 3

While they do not prevent erosion from occurring, their greater thickness can delay the
time to erosion-caused failures.

Blast Joints

Blast joints are used to increase the abrasion resistance of the tubing string against
the jetting action of a producing formation. Blast joints should be located in the tubing
string opposite all upper perforations spanned by the tubing. Blast joints should also be
used in the wellhead area where abrasive fracturing fluids may be pumped into the
casing access. Polished nipples are sometimes included in the tubing string on either
end of a blast joint in order to provide sealing surfaces for a spacer pipe should the
blast joint fail.

Flow couplings

Flow couplings should be run immediately above each selective or no-go landing nipple
in the tubing string that may be used to locate a flow control device. In high rate or
corrosive gas wells, flow couplings should be used above and below all upsets or profile
changes to reduce erosion, especially if the turbulent fluid contains abrasive particles.
Since most flow controls restrict the tubing ID, the tubing above and below the
controls should be protected by use of a flow coupling.

315
SUBSURFACE SAFETY VALVES

Application of Downhole Safety Valves

An SSSV must be installed in all offshore wells capable of flow and at onshore locations
in high pressure or sour gas wells in close proximity to housing, public roads, or rock
slide areas. These requirements are often dictated by government regulations and/or
corporate policy.

The objective is to provide a downhole shut-off that will limit the magnitude and
consequences of the hydrocarbon emission if the primary well control device at the
surface (e.g., Christmas tree) is damaged or cannot be operated. This could occur if a
platform was damaged by a storm or major vessel impact, an explosion, a blowout, or
by foundation instability. Similarly, on land a landslide or vehicle impact might knock
off the well-head. There are several different types of subsurface safety valves.

Flow-Controlled Safety Valves

These are usually deep-set valves whose operation is directly controlled by the well
stream. They are normally wireline retrievable since they must be reset from time to
time, especially as well conditions change. They are designed to be open normally, but
to snap shut if the tubing pressure dips below a threshold or the production rate
exceeds a preset limit ( Figure 1 and Figure 2 , Flow-controlled safety valves).

316
Figure 1

317
Figure 2

They normally use spring tension to hold the valve open. The flow passes through a
flow tube containing a bean. If the pressure drop across the bean exceeds the spring
tension the valve will snap closed. The valve can be a ball, flapper, or stem type. The
safety valve is reopened by raising the pressure on the downstream side in excess of
the closed-in bottomhole pressure. Obviously, setting this type of valve requires an
accurate knowledge of well behavior, temperature, and flow conditions.

The major advantage of these valves is that they are cheap and can be set deep in the
well below the packer, protecting both the tubing and annulus. The main
disadvantages are the servicing and design requirements, the restrictions to flow
capacity and flexibility, and the risk of inadvertent reopening as a result of fluids lost
into the wellbore (e.g., seawater or mud in the event of an offshore collision of a boat
with a wellhead).

Surface-Controlled Subsurface Safety Valves (SCSSSVs)

The SCSSSV is a fail-close valve that is held open by a high pressure control line (
Figure 1 , Tubing-retrievable, surface controlled, subsurface safety valve).

318
Figure 1

If the control line is severed in an event that damages the tubing or wellhead, pressure
will leak off and the safety valve will close. The control line is generally connected to
an emergency shutdown system to give automatic closure during unsafe or alarm
conditions (such as fire or detection of gas). There is usually a control panel with
pressure gauges and control valves for all of the wells on an offshore platform.
Surface-controlled valves are the type of downhole safety valve most commonly
favored today. In some countries, the regulations require the use of this type of valve
in all offshore wells and onshore sour wells capable of flow. There are two basic types
of SCSSSV: wireline retrievable (TFL retrievable) and tubing retrievable.

With the wireline retrievable valve, an SSSV landing nipple is installed in the tubing
string. This is basically a landing nipple with a port through which the control line
enters between a set of packings on the SSSV. The SSSV is installed across this nipple.
This type of valve can have a service life of 18 to 24 months or longer, although many
valves fail during periodic testing. The relative ease of servicing and replacing wireline
retrievable valves therefore offers distinct advantages. The main disadvantages of
wireline retrievable SSSVs are that

the service life is short

319
the restricted throughbore means the valve has to be pulled for deep wireline
or through-tubing work

the turbulence in the flow stream increases pressure loss and erosion
problems

we are forced to rely on a lock to make sure the valve is not blown out of the
well on closure

On the other hand, the initial costs are relatively low and servicing can be undertaken
with minimal disruption to production.

The tubing-retrievable SSSV valve is an integral part of the tubing string. As a result, it
generally has a larger through-bore than the retrievable-type valves and may even be
designed with an internal diameter that is the same as the tubing (a fullbore valve).
Also, it is not so dependent upon elastomer seals and therefore has a much longer
service life (5 to 20 years, depending on design and materials selected) . Tubing-
retrievable valves with all metal-to-metal seals are available for severe environments.

Since most valve failures are caused by elastomer problems and since tubing-
retrievable valves require a rig entry for repair work, these valves are often backed up
with a nipple section positioned to accept a wireline-retrievable valve. Use of this valve
insert minimizes the impact of a valve failure on production. This feature is
incorporated in tubing retrievable valves together with a lock-out sleeve for the original
valve, to lock it permanently open. Service procedures have been developed for
installing the insert valve using both TFL and wireline techniques.

To reduce the number of critical seals, many companies prefer single-control line
valves. Although spring design does limit the depth to which this type of valve can be
set, technology improvements have pushed the limit from around 650 ft (200 m) to in
excess of 3250 ft (1000 m). Balance-line valves (valves with two control lines, one to
close and one to open), which were developed to overcome the earlier depth
limitations, are therefore becoming less popular. However, they have an advantage in
that they can be pumped closed to facilitate the cutting of an obstruction in the valve,
such as wireline.

A balance-line valve can also be set at any depth, the critical issue being the closing
time and control line efficiency. Protection systems for the control lines have been
considerably improved over the last decade so that installation damage is much less
common. New hydraulic fluids have also reduced control line pressure losses, so that it
is possible to set a balance-line valve at the packer level. Field trials are also in
progress on the use of electrical control systems for deep-set surface-controlled
subsurface safety valves.

The SCSSSV may incorporate a ball valve or a flapper valve. Ball valves are often
considered more robust and can sometimes cut wireline when they are closed with it
across the valve. However, they are prone to damage by sand and improper operation.
The simpler flapper valve has the advantage of always being reopenable (mechanically
if necessary) should it become stuck in the closed position. Extensive studies have
shown that flapper valves are more reliable than ball valves; as a result, most

320
operators run flapper style SCSSSVs. Operators also have the choice of running
tubing-retrievable SCSSSVs with lockout option. These valves offer redundancy (i.e.,
the ability to lock open one valve and run a second valve inside). Sleeve-type valves
are also available.

Setting Depth

The recommended setting depth for a safety valve is often a matter of company
philosophy and operating procedure. Where concurrent drilling and production is
undertaken, many companies like to set the safety valve beyond the kickoff point so
that it can be used to shut in the well during the tophole drilling and kickoff of an
adjacent well. Other designers prefer not to subject the valve to significant bending
and therefore favor its installation nearer to the surface. Subsurface safety valves are
normally set at least 150 ft (50 m) below the surface or sea floor.

Bottomhole Chokes and Regulators

These are not safety devices in the strict sense of the word; however they are often
used with the objective of limiting the rate at which a well could produce and therefore
blowout, limiting the surface operating pressures in high pressure wells, and limiting
the drawdown rates on wells that have a tendency to produce sand.

They are also used to take maximum advantage of the formation temperature to avoid
hydrate formation during choking of a high pressure gas stream at surface
temperatures. Chokes are designed to give a constant rate, while regulators give
constant choking or pressure differential.

The other safety device that is often used in injection wells is a simple check valve that
will permit injection but not production. This device may be used to prevent backflow
of water and formation sand, should injection cease.

All of these devices are normally wireline retrievable.

321
WELLHEAD EQUIPMENT

Wellheads

Wellheads typically are the joint responsibility of the production department (tubing
head and Christmas tree) and drilling department (casing head and intermediate
casing head). Figure 1 shows a typical flanged wellhead.

Figure 1

The size and pressure ratings of wellheads are dictated by the design considerations
for the tubulars (e.g., tubing size, casing size, kill and stimulation pressure
requirements, flowing pressure requirements). However, government regulations
sometimes require that the rating of the upper part of the wellhead be at least equal to
the reservoir pressure.

Wellhead specifications are presented in API Spec. 6A. The standard wellhead ratings
are 1000, 2000, 3000, 5000, 10,000, 15,000, and 20,000 psi (7, 14, 21, 34, 69, 103,
130 MPa).

322
Wellhead components are generally flanged, although threaded components are
permitted on low pressure wellheads with pressures less than 2000 psi (14 MPa).
Threaded valve and choke connections can be used with pressures of up to 5000 psi
(34 MPa) in wells less than 12,000 ft (3700 m) deep, but are not recommended.
Clamped connections are sometimes used in the intermediate pressure range 2000 to
10,000 psi (14 to 69 MPa) ( Figure 2 , Wellhead and Christmas tree for a dual-tubing
completion utilizing clamp-type connections). For wells producing H2S gas, the
wellhead materials must conform to NACE specifications.

Figure 2

Tubing Heads and Hangers

The tubing head packs off around the production casing ( Figure 1 , Tubing head and
tubing hanger installation).

323
Figure 1

It should have an outlet for access to the tubing/casing annulus, with an internal
thread to receive a plug when redressing of the side outlet valve is necessary. The
rating of the upper flange must be compatible with the Christmas tree. The tubing
head should have lock-down screws for the hanger, and the lower flange size and
rating must be compatible with the casing head flange.

The bore and size of the top flange are generally determined by completion and well
servicing requirements (BOP size, packer, and tool ODs) rather than the Christmas tree
flange size.

Like the casing, the pressure rating of the tubing head spool is often dictated by
stimulation pressure requirements and may therefore be of a higher rating than the
Christmas tree, which can be removed or protected during stimulation.

Offshore, a compact wellhead, or unihead, is often used to combine both the casing
and tubing spool's function and reduce the overall height of the wellhead.

Three types of tubing hangers are commonly used:

324
the boll weevil (also called a threaded mandrel) hanger, which is an integral
part of the tubing string and therefore a fixed point that shoulders into the
tubing head spool ( Figure 2 , Boll weevil tubing hanger)

Figure 2

the wrap around hanger, which is hinged to permit installation onto any part
of the tubing other than a coupling

the dual hanger, either multibore mandrel or split hanger

The mandrel types are the most common.

It is highly desirable to have an internal thread in the tubing hanger to allow the
installation of a back pressure valve while removing, repairing, or pressure testing the
tree. This can be installed and removed under pressure with a special tool.

Christmas Trees

325
(see Figure 1 , Typical flanged wellhead)

Figure 1

There are three main types of trees: the assembled tree, the solid block tree, and the
control head tree (often found on thermal wells). The major components (from bottom
up) are

the flange

the master valve(s)

the tee or flow cross

the swab valve

the crown plug

326
the wing valve

the bean box or choke

the flow line valve

For high rate wells the flow tee is often Y-shaped to reduce turbulence and erosion. Similarly, a flow
control valve may be installed in a straight run rather than in the conventional right-angled bean
arrangements.

A second side outlet is often used on high pressure wells as a connection for a tubing
kill line. Similarly, two master valves are often used in severe operating conditions.
This is often a regulatory requirement in sour or high pressure wells.

Tubing Rating Tree Tree


Size Bore Drift

(in) (mm) (psi) (MPa) (in) (mm) (in) (mm)

2 3/8 (60) 15,000 (103) 2 1/16 (52) 1.901 (48)

2 7/8 (73) 15,000 (103) 2 9/16 (65) 2.347 (60)

3 1/2 (89) 5,000 (34.4) 3 1/8 (79) 2.867 (73)

3 1/2 (89) 15,000 (103) 3 1/16 (78) 2.867 (73)

Table 1: Christmas tree specifications

Full opening gate valves are used for the master and swab valves
( Figure 2 ,

327
Figure 2

Manually-operated gate valve, Figure 3 ,

328
Figure 3

Pressure-actuated gate valve in open position, and Figure 4 , Pressure-actuated gate


valve in closed position).

329
Figure 4

These should not be opened when a significant differential pressure exists across the
closed valve.

The throughbore of the tree is specified by the API and is generally 1/16 in. larger than
the tubing ID to facilitate installation of a back pressure valve in the tubing hanger.
Tree sizes are shown in Table 1.

Although the body of a Christmas tree is normally pressure tested to twice the working
pressure for trees rated at 5000 psi (34.4 MPa) and less, and 1.5 times the working
pressure for 7500 to 20,000 psi (52 to 140 MPa) ratings, the flange bolts and valves
may not necessarily have the same rating. Therefore, it is extremely imprudent to
overload Christmas trees when stimulating a well. Similarly, many valves are
unidirectional and this should be taken into account when planning pressure test
sequences. Valve gates can be damaged by applying significant pressure from the
wrong side.

Beans and Chokes

330
In flowing wells, rate is controlled by a bean, choke, or flow control valve.
Traditionally, the most common was the fixed bean operating under critical flow
conditions (i.e., at sonic velocity). Under these conditions the upstream pressure, or
tubing head pressure (THP) , is independent of the downstream pressure, or flow line
pressure (FLP). To achieve this, THP must be greater than or equal to 2.0 times the
flow line pressure. The advantages of operating under these conditions include the
following:

over the short term (generally one to three months) the well rate is fixed, and
a single monthly test is representative of the entire producing period

test separator conditions need not be the same as the bulk separator to
ensure a representative test, since fluctuations in downstream pressure do not
affect THP at sonic velocity

well flow rate is limited in event of a line break

lower pressure ratings can be used for flow lines and separators

the sand face is not subjected to production surges in event of a production


facility fluctuation (this point is particularly important in weak formations)

choke performance can be used as an indication of production rate

The disadvantages relate primarily to lower pressure wells and gas wells:

the choke introduces a major pressure loss into the system

flow lines may need to be larger to accommodate the higher flow velocities
without excessive erosion or pressure loss

associated cooling can cause hydrate formation at the choke

choke beans are inconvenient for changing production rates in accordance


with changes in gas sales requirements

To meet the last objection, motorized or manual variable chokes or flow control valves
are often used on key wells so that the operator can quickly change the field flow rate.

331
APPENDIX A
API Publications Related to Completion Design

Key References for Completion Engineers

RP 5A4 Care and Use of Reinforced Thermosetting Resin Casing and Tubing

RP 5A5 Field Inspection of New Casing, Tubing, and Drill Pipe

RP 5C1 Care and Use of Casing and Tubing

Bull 5C2 Performance Properties of Casing, Tubing and Drill Pipe

Bull 5C4 Round Thread Casing Joint Strength with Internal Pressure and
Bending

Spec 6A Wellhead Equipment

RP 11L Recommended Design Calculations for Sucker Rod Pumping


Systems

Other Related Documents

Std 4A Specification for Steel Derricks (Including Standard Rigs)

Std 4D Specification for Portable Masts

Spec 4E Drilling and Well Servicing Structures

Spec 5A Casing, Tubing, and Drill Pipe

Spec 5AC Restricted Yield Strength Casing and Tubing

Spec 5AR Reinforced Thermosetting Resin Casing and Tubing

Spec 5AX High Strength Casing, Tubing and Drill Pipe

Bull 5AZ Thread Compounds

Std 5B Specifications for Threading, Gauging and Thread Inspection of


Casing, Tubing and Line Pipe

RP 5B1 Thread Inspection on Casing, Tubing and Line Pipe

Bull 5C3 Formulae and Calculations for Casing, Tubing, Drill Pipe and Line
Pipe Properties

332
Spec 6AB 30,000 PSI Wellhead Equipment

RP 6F Fire Test for Valves

RP 6G Through Flowline (TFL) Pump Down Systems

Spec 10 Materials and testing for Well Cements

Bull 10C Oil Well Cement Nomenclature

Spec 10D Casing Centralizers

RP 10E Application of Cement Lining to Steel Tubular Goods: Handling,


Installation and Joining (Tentative)

Spec 11AX Subsurface Sucker Rod Pumps and Fittings

RP 11AR Care and Use of Subsurface Pumps

Spec 11B Sucker Rods

RP 11BR Care and Handling of Sucker Rods

Std 11E Specification for Pumping Units

RP 11G Installation and Lubrication of Pumping Units

Bull 11L3 Sucker Rod Pumping System Design Book

Bull 11L4 Curves for Selecting Beam Pumping Units

Spec 11P Packaged High Speed Separable Engine Driven Reciprocating Gas
Compressors

RP 11R Electric Submersible Pump Installation

RP 11S Operation, Maintenance and Trouble Shooting of Electric Submersible


Pump Installations

Spec 14A Subsurface Safety Valve Equipment

RP 14B Design, Installation and Operation of Subsurface Safety Valve


Systems

Manual

14BM Subsurface Controlled Subsurface Safety Valve Sizing Computer


Program

RP 14C Analysis, Design, Installation and Testing of Basic Safety Systems


Offshore Production Platforms

Spec 14D Wellhead Surface Safety Valves and Underwater Safety Valves for
Offshore Services

333
RP 14E Design and Installation of Offshore Production Platform Piping
Services

RP 37 Proof-Test Procedure for Evaluation of H.P. Casing and Tubing


Connection Designs

RP 39 Standard Procedures for Evaluation of Hydraulic Fracturing Fluids

RP 41 Standard Procedures for Presenting Performance Data on Hydraulic


Fracturing Equipment

RP 42 Laboratory Testing of Surface Active Agents for Well Stimulation

RP 43 Standard Procedure for Evaluation of Well Perforators

RP 49 Safe Drilling of Well Containing Hydrogen Sulphide

RP 53 Blowout Prevention Equipment Systems

RP 56 Testing Sand Used in Hydraulic Fracturing Operations

APPENDIX B
A Selection of Published Multiphase Gradient Curves

Author Title Maximum Comment


TubingSize (in.)

K. E. Brown The Technology of Artificial 13.625 Most extensive oil, gas,


Lift Methods Vol. 3a, 3b, water, hor. Convenient
PennWell Books scale

Otis Gas Lift Flowing Gradient 4.5 Convenient scale for G/L
Curves design

Camco

H. W. Winkler S. Camco Gas Lift Manual 3.5 Convenient scale for G/L
S. Smith design

K. Aziz Gradient Curves for Well 4.5 Metric Convenient for


J. R. Eickmeier Analysis and Design CIM deep wells or HP wells
M. Fogarasi Vol. 20 Includes gas wells
G. A. Gregory

W. E. Gilbert Flowing and Gas Lift Well 3.5 Empirical Convenient for
Performance API 801-30H deep or HP wells

334
PERFORATING

INTRODUCTION

PERFORATING FUNDAMENTALS
Once a well has been cased and cemented we must provide effective flow
communication between the formation and wellbore, and this involves creating a series
of holes through casing and cement--perforating.

During the 1920s, rotary drilling was replacing cable-tool drilling and it was becoming
common to cement casing throughout the well. Accordingly, in the early 1930s, several
service companies began to offer bullet perforating. Perforating "guns," which fired
pointed projectiles at the inside of the casing, were the mainstay of the perforating
industry through the l950s. Some companies still offer bullet guns, particularly for use
in low compressive strength formations, but the vast majority of all perforating jobs
performed in the world today involve shaped charge perforating.

The shaped charge, or jet, perforator came into use in the oil field in the 1950s
following the development of explosives technology during World War II. Early
observers found that when they detonated an explosive charge that had been shaped
to form a cavity and lined with metal, the lining was transformed into a powerful
particle jet capable of penetrating armored targets. This capability was developed into
a variety of antitank weapons, including the bazooka, and ultimately into the
perforating gun. The shaped charge perforating gun largely replaced the bullet gun
because it was more reliable and because more perforating power could be packaged
into a smaller space, making the shaped charge more adaptable to the increasing
variety of completion designs.

Basic Perforating Methods


Cased holes may be perforated using any of several devices ( Figure 1 , Basic
perforating methods):

Conventional casing guns which are run into the well on electric wireline with or without
wireline pressure control equipment

335
Figure 1

Through-tubing guns which are run into the well after the tubing has been
installed, again via wireline pressure control equipment

Tubing-conveyed guns which are run on the bottom of the tubing string and
detonated using mechanical, electrical, or pressure-activated firing mechanisms

The completion designer also has the option of perforating overbalanced, with a higher pressure in
the wellbore than in the formation, or underbalanced, with a wellbore pressure lower than formation
pore pressure.

The bottomhole pressure is determined by the density and height of fluid in the
wellbore or tubing, and by any additional gas pressure added to the surface.
Depending on the option chosen and the degree of pressure differential, completion
fluid will flow into the formation, or formation fluids will flow into the wellbore when
perforations are created in a productive reservoir.

The delivery system for placing the shaped charges at the proper location in the well
via wireline may be categorized as

retrievable, consisting of a cylindrical, hollow steel charge carrier;

336
semiexpendable, where the charges are conveyed into the well on a
retrievable metal strip or wire carrier (used in through-tubing operations where
gun size is an important factor); and

fully expendable, where the charge cases and carrier linkage disintegrate and
only the wireline is retrieved (again, used in through-tubing operations).

Generally speaking, the larger-diameter casing guns run via wireline will give larger hole diameters
and greater penetration depths than the smaller through-tubing guns. Through-tubing perforating,
however, provides a safe, practical method for perforating underbalanced. Underbalanced
perforating has been generally accepted as a technique that can help to maximize well productivity.
Tubing-conveyed perforating provides a means to combine both high-performance guns and
flexible pressure conditions. However, higher cost and operational considerations are sometimes
cited as reasons for choosing the more conventional techniques.

Completion designers must consider many factors when designing a perforating


program, including

the relative importance of hole size, penetration, and shot density (shots per unit length)

the anticipated radial extent of formation damage due to drilling fluid losses

the reservoir's pressure, permeability, and susceptibility to damage from


completion fluids

the length of the perforated interval, the size of the completion string
components, and the need to perform specialized operations (e.g., sand
control, stimulation)

the safety issues related to the particular wellsite conditions

the expected pressure, temperature, and timing extremes for the perforating
operation

the relative economic importance of minimizing short-term costs versus


maximizing long-term productivity

Fundamentals of Shaped Charge Operation


Figure 1 (Shaped charge components) illustrates the relatively simple construction of
the shaped charge.

337
Figure 1

The four important components are the conical metallic liner, the main explosive
charge, the primer explosive, and the case that encloses the charge.

The firing is initiated electrically or mechanically by the detonator (blasting cap) via the
detonating cord (Primacord), which in turn sets off the primer explosive. The primer
initiates detonation of the main charge. Explosive pressure on the metal liner causes it
to collapse inwardly along its axis forming a high-velocity jet of fluidized metal ( Figure
2 , Simulated snapshots of an exploding shaped charge during the first 16.6
microseconds after detonation).

338
Figure 2

In early shaped charges a portion of this metallic liner collapsed to form a slower moving
slug ("carrot"), which would subsequently lodge in the perforation tunnel, thereby reducing
flow capabilities. Modern premium charges have eliminated this slug by incorporating
liners fabricated from powdered metal. The metal particles are spread over the perforation
tunnel rather than forming a solid slug. Earlier attempts at slug elimination utilized a
bimetallic liner with a low-melting-point outer layer that volatized at high temperatures,
reducing the size of the slug (Delacour 1958).

Liners are formed from mixtures of copper, lead, zinc, tin or tungsten. The density
distribution and dimensions of the liner must be carefully controlled to maintain
consistent, high performance.

The tip of the particle jet moves at velocities of 25,000 to 30,000 ft/s (8000 to 10,000
m/s) and develops pressures estimated to be in the order of 5 to 15106 psi (35 to 100
GPa). (The tail of the jet moves more slowly, at about 3,000 to 7,000 ft/s (914-2134
m/s).) Under these conditions, the materials in its path flow plastically under impact.
The jet crushes the reservoir rock and displaces both rock and pore fluids radially from
its axis. The entire perforating process, from initiation to completion of penetration,

339
lasts for only several 100 microseconds (Bell l972a). Although the formation rock is
crushed and compacted, it is not fused together by the jet.

The result of this split-second, high-energy operation is illustrated in Figure 4 (Typical


perforation).

Figure 4

Depending on the design of the shaped charge and the type of materials perforated,
perforation length generally varies from about 2 to 20 in. (5 to 50 cm) and entrance
hole diameter from about 0.2 to 1.0 in. (0.5 to 2.5 cm) for most oil field perforators.
The precise shape of the perforation will vary somewhat depending on charge
geometry, target characteristics, and gun positioning.

Commercial shaped charges are manufactured to deliver perforations of a certain size.


The gun's performance is affected by various charge design parameters, including

the geometry, density, thickness, dimensions, and makeup of the liner

the distribution, density, and amount of explosive

It would be easy to assume that larger loads of explosive would generate larger perforations.
However, increasing the gun size (and the associated increase in the size of the charge liner) can

340
be an important factor. Some charges with less explosive loads outperform smaller guns with higher
loads (Bell l972a).

The position of the gun in the casing is another important parameter in determining
the size of the perforation. Gun clearance, or standoff, is the distance from the casing
inner surface to the gun, along the axis of the charge. Figure 5 (Gun clearance effects)
illustrates that penetration and hole size generally decrease with increasing standoff.

Figure 5

The issue of gun clearance becomes particularly important when through-tubing


perforating is carried out using smaller-diameter guns in larger-size casing. Figure 6
shows the range of perforation sizes that might result when a 1 11/16-in.

341
Figure 6

(4.3-cm) gun is fired in 7-in. (17.8-cm) casing. Centralizing the gun would make the
perforation sizes more uniform in such a case, but a better approach is to position the
gun against the casing with all charges pointing in the same direction (zero degrees
phasing) and a low clearance. However, while this may maximize perforation length
and hole size, the positioning of all perforations along one side of the casing can
introduce a geometrical skin effect to the completion and affect productivity. The
completion designer may need to weigh the advantages of underbalanced perforating
available with through-tubing techniques against the disadvantages of reduced charge
performance or possible phasing restrictions that come with smaller gun sizes.

Another important factor in the performance of shaped charge perforators is the


strength of the target material. Early researchers determined that increases in the
compressive strength of the formation resulted in decreases in the penetration depth
of the perforation. For example, Thompson (1962) reported a correlation between rock
penetration and compressive strength as Lp/Lpr=exp[0.086(cr-c)] where Lp=total
target penetration in the formation of interest (inches); Lpr=total penetration on a
"reference" formation (inches); c=compressive strength in formation of interest
(1000psi) and cr=compressive strength of reference formation (1000psi).

Subsequent work has shown that in deep, high-strength formations, the reduction in
penetration from that published as test data can be very significant (Weeks 1974).

342
Figure 7

Figure 7 (Penetration vs. formation compressive strength) shows that a charge that
penetrates 5 in. (12.7 cm) into a Berea sandstone test target (6000 to 8000 psi
compressive strength) might be expected to penetrate less than 2 in. (5 cm) into a
Smackover sandstone target with a compressive strength of 22,000 psi. When the
effect of such reduced penetration depth is compounded by formation damage and
perforation plugging, the productivity of the completion can be drastically reduced.
More recent work (Saucier and Lands 1978) has confirmed the correlation between
reduced penetration and increased compressive strength among samples from
different formations. These same researchers determined that penetration was also
reduced with increasing effective stress (overburden pressure minus pore pressure). A
penetration plateau was achieved in some test samples, after which no further
reduction occurred.

Casing grade also affects perforation entrance holes, with increasing casing strength
causing a reduction in hole diameter. Typically, a 10% diameter decrease might be
expected in moving from J-55 to N-80 casing, with a <10% reduction in going from N-
80 to P-1l0 (Bell l972a). Variation depends on the thickness of the casing and the
quality of the cement sheath behind the casing. Multiple strings of casing also reduce

343
perforation depth, as here in Table 1, (King et al. 1984).

Casing size (in.) Charge wt (gm) Outer CSG hole (in.) Cement penetration (in.)

5.5, 7.0, 8.6 19 0.257- 0.266 16

7.0 19 0.397- 0.412 21.9

5.5, 7.0, 8.6 22 0.332 - 0.309 9

5.5 22 0.454 -0.439 20.6

Notes:
(1) same age and manufacturer for each charge weight;
(2) single casing test data are average of five shots;
(3) smallest and largest hole diameter measurements (90o)

Hydrostatic pressure, wellbore fluid density and temperature (within the charge's
operating range) do not appear to significantly affect charge performance, although
perforating in gas-filled wells can produce marked differences in penetration compared
with shooting in liquid-filled holes (Bell et al., 1995). Generally speaking, penetration
values in a gas-filled wellbore are equal to or greater than those in a liquid-filled
wellbore. High bottomhole temperatures may indirectly reduce penetration, since the
explosives required for such situations are not always as effective as those used in
lower-temperature operations (Bell 1984).

As operators have begun to realize the importance of charge performance, more


attention is being paid to the quality control standards utilized in the manufacture and
storage of the actual charges and in the assembly of the gun itself. Investigations have
shown that the penetration performance of charges stored for more than three years
prior to use varies widely from published standards. In addition, misalignment of the
charge in the carrier and charge malfunctions can affect penetration by 20 to 50%
(King et al., 1984). Generally speaking, the productivity of completions shot with
larger guns does not suffer as much from poor charge performance. But compounded
by the effects of drilling damage, debris plugging, and high in-situ formation strength,
these performance reductions can prove significant. The added cost to the operator of
an adequate quality control program for the perforating company is a wise completion
investment.

Explosives Used in Perforating

Explosives are categorized according to the ease with which they are detonated, the
speed with which they react, and the combustion pressure that they generate.

Low explosives generate relatively low combustion pressures and react at velocities in
the 500- to 1500-m/s range. The first low explosives were probably invented by the
Chinese in the 10th century, using a combination of powdered carbon and nitrates of
sodium and potassium. Gunpowder is the most common example of low explosives. On
reacting, most gunpowders liberate about 150 times their original volume in the form

344
of gas. The reaction rate depends on the grain size, temperature, degree of
confinement, and packing density. High temperatures alone will start a reaction and
some form of heat is used to initiate these explosives. Low explosives do not generally
play a role in perforating systems, but they are used in other areas of the oil fieldfor
example, in sidewall sample gun cartridges.

High explosives detonate at 3000 to 9000 m/s and generate high pressures. The first
high explosive, glyceryltrinitrate (nitroglycerin), was discovered in 1846 by an Italian
chemist, Ascanio Sobrero. In 1867 Alfred Nobel mixed it with clay to make dynamite.
High explosives are subdivided into primary high explosives (or initiator explosives)
and secondary high explosives, according to their sensitivity to detonation.

Primary high explosives are generally high-density compounds of metals and nitrogen
that instantaneously detonate when subjected to a heat source (static electricity,
impact, friction, spark, flame). Lead azide and lead styfnate are common examples.
The purpose of primary explosives is to initiate the more powerful and less sensitive
secondary high explosives. Primary explosives are found in blasting caps and boosters,
the handling of which requires special safety procedures and equipment. Lead azide is
the most commonly used initiator in oil industry tools because of its relatively high
resistance to temperature (475 F to 600 F).

Secondary high explosives have the highest reaction rate (7000 to 9000 m/s,
depending on packing density) and generate the highest pressures (millions of psi).
The volume of gases produced is 750 to 1000 times the original volume of explosive.
However, most secondary explosives are so insensitive to detonation that they can be
melted, cast, and machined. These explosives are generally organic compounds of
nitrogen and oxygen, and when a detonator initiates the breaking of the molecules'
atomic bonds, the atoms of nitrogen lock together with much stronger bonds, releasing
tremendous amounts of energy. Typical secondary explosives include

RDX
HMX
PS (picryl sulfone)
HNS (hexanitrostilbene)
Composition B (60% RDX, 40% trinitrotoluene)
Amoniumnitrate (fertilizer, also used in seismic operations)

RDX is the most commonly used secondary explosive for shaped charges. As with all
explosives, it becomes more sensitive to detonation as it becomes more finely divided
(powder versus pellets). When it is unconfined, RDX will burn slowly if lit. However, if
sufficiently confined an explosion can occur. Despite its "insensitivity," RDX should be
handled with extreme care. Other types of secondary explosives are sometimes used
when extreme temperature resistance is required, as may be the case in deep wells or
TCP operations, where the guns are exposed to well temperatures for longer periods of
time.

When a perforating gun is fired, an electrical current sets off the primary explosive in
the blasting cap, which in turn sets off an RDX base or booster charge within the cap.
This in turn detonates the Primacord (5 g of RDX per foot) used to carry the detonation
to the individual charges. As the pressure wave reaches the shaped charge, it
detonates fine-grained RDX primer in the tunnel at the base of the charge casing.

345
Detonation of this primer in turn detonates the main body of RDX (generally with 1%
wax binder) and the wave achieves full speed and pressure as it bears on the metallic
liner. This pressure collapses the liner, forming the "jet" of high-speed material that
strikes the target at high pressure and does the penetrating. A typical charge
containing 2/3 oz. (22 g) of RDX might penetrate 6 in. of solid steel at a normal
clearance, or 1/2 in. of steel at a distance of 30 ft.

346
PERFORATING GUN TYPES

Perforating Gun Types


Perforating guns are categorized ( Figure 1 ) as:

retrievable steel hollow carrier guns;

Figure 1

semiexpendable wire or strip carrier guns;

fully expendable guns.

With the retrievable hollow carrier gun, the charges are positioned within a steel cylinder. Within the
carrier, each charge is surrounded by air at surface pressure, and is aligned with a threaded port
plug or a thinner portion of the carrier wall (scallop gun) ( Figure 2 , Steel hollow carrier guns: (a)
threaded port plugs; (b) scallop gun).

347
Figure 2

Upon detonation, the jet pierces the plug, providing a positive indication of firing when the gun is
retrieved. The carrier cylinder may expand slightly due to the explosive force, but most of the debris
is recovered within the gun.

Expendable guns are designed to partially or completely disintegrate upon firing. They
typically consist of a series of individually sealed charge cases constructed of a
frangible material (aluminum, ceramic, glass, or cast iron). After firing, only the
wireline remains to be retrieved from the well. Semiexpendable guns are designed with
a strip or wire carrier that is retrieved after firing. Some advantages and disadvantages
of each gun type are listed below (from Bell l972a):

Retrievable hollow carrier guns

Advantages

high reliability because all components are protected within carrier (generally gastight and
chemical resistant)

heavier and stronger than expendable types, permitting rougher treatment


and faster wireline speeds during running procedure

348
generally higher temperature and pressure ratings

very little debris left in well

explosive force is retained within carrier to eliminate casing deformation

port plug provides positive indication of firing (and, in some cases, evidence of
charge performance)

Disadvantages
may have trouble running through-tubing retrievable guns in buckled tubing due to rigidity
of carrier

may have smaller charge components than comparable expendable guns, with
some resulting reduction in penetration

weight may limit length of assembly and, therefore, of perforation interval


that can be shot with single gun run

Fully expendable guns

Advantages

generally cheaper and easier to assemble at wellsite

lighter and more flexible than steel hollow carrier guns, facilitating longer
perforation intervals

size for size, these offer more penetration than comparable retrievable guns
(up to 25% increase for very small diameter guns)

Disadvantages
can deform casing when detonated

leave substantial amounts of debris in well

components are exposed to wellbore fluids, temperature, and pressure,


resulting in reduced reliability

charge cases may not be leakproof, allowing gas or liquid to enter casing and
reduce performance

guns are not as sturdy (by design) as retrievable guns and premature
breakage may result in fishing operations; running speed usually is limited to
168 ft/min (50 m/min)

pressure and temperature limits usually are lower than those of retrievable
guns

349
no positive indication of individual charge firing

Semiexpendable guns

Advantages

same as fully expendable

some reduction in amount of debris left in wellbore due to recovery of linkage


system and carrier frame

with a metal strip or wire carrier, ceramic or glass charge cases can be used,
reducing the size of debris particles and increasing resistance to chemicals and
gas leakage

Disadvantages
generally the same as fully expendable, although ceramic casings are somewhat sturdier
and wear resistant

debris or a loose charge casing can cause gun to lodge in tubing while running
in, causing further breakage and jamming; fishing can be difficult and may
require that the tubing be pulled or casing drilled out

The strip carrier has proved popular and reliable for larger casing guns, but the metal strip's
tendency to expand after firing has made its recovery difficult in through-tubing applications. Wire
carriers are more popular in these cases.

Expendable and semiexpendable guns are generally confined to shallower completions,


although ceramic charge casings can be useful in deeper gas wells requiring through-
tubing perforating.

Casing Guns

Both retrievable and expendable guns are used for conventional casing perforation
operations. Retrievable hollow carrier guns normally range from 3 1/8 in. (7.9 cm) up
to 5 in. (12.7 cm). Since larger guns generally carry larger charges, and larger charges
generally achieve larger, longer perforations, the largest casing gun that can be run is
ordinarily used. A running clearance of 0.5 in. (1.3 cm) is a typical rule-of-thumb
allowance for trouble-free entry (Bell l972a). Because of the small clearance, the guns
do not require positioning against the casing, and phasing can be employed to
minimize geometrical skin effects during flow (Karakas and Tariq, 1991). Any shot
density can be employed to minimize geometrical effects, with 4 shots per foot being a
very common design. Perforating gun length is limited by the total perforating
assembly length that must pass through the pressure control equipment at the
surface. This generally limits the active gun length to about 30-40 ft for large-diameter
hollow carrier guns, and about 60-80 ft for small-diameter hollow carrier guns, or
expendable or semiexpendable through-tubing guns.

Expendable and semiexpendable casing guns are generally available in sizes from 3
1/8 in. (7.9 cm) to 4 in. (10 cm). Phasing and shot density options are much the same

350
as for retrievable casing guns, but much longer lengths (several hundred feet) may be
run on a single trip.

Through-Tubing Guns

Through-tubing guns have smaller diameters than casing guns to allow passage
through tubing or small-diameter casing strings. A hollow carrier gun run in 2 3/8 in.
tubing, for example, typically has a diameter of 1 9/16 in. The retrievable guns are
slightly smaller than their expendable counterparts to permit them to be pulled back
through any tubing nipples when the carrier has swelled after firing (unless they are
designed to be remotely detached). Because of the clearance problems associated with
perforating inside casing after the smaller-diameter guns have exited the tubing,
retrievable guns are often equipped with positioning devices that employ magnetic or
mechanical means to set the gun against the casing wall for firing. In addition, the
charges are often positioned for "in-line firing" (zero degrees phasing). Carrier swelling
(due to the force of explosion) may be reduced by maintaining a minimum hydrostatic
pressure on the gun during firing. This pressure is generally between 0 and 500 psi (0
to 3.5 MPa) in liquid and 500 to 4500 psi (3.5 to 31.0 MPa) in gas-filled boreholes (Bell
l982b).

The length of retrievable through-tubing guns is generally limited by the height of the
pressure control riser practically speaking, about 20 ft (6.1 m). The effective
perforating gun length is reduced further if weights are required to allow the gun to
descend.

Pressure and temperature ratings for through-tubing guns are given below in Table 1
(Ratings are primarily charge-related). The hollow steel carrier retrievable versions are
preferred for deeper, higher-pressure wells. Because the perforating debris is
minimized and retained within the carrier, retrievable guns are also preferred in
situations where the well is to be flowed immediately after perforating, prior to gun
retrieval. In such situations, semiexpendable carriers may become stuck as gun debris
settles back to the bottom after a short flow period. In addition, if the well is
perforated underbalanced and through tubing, the pressure differential must not be so
great that it causes the carrier and/ or cable to be blown up the tubing after firing.

Table 1.

Gun Type Hollow Carrier Semiexpendable Expendable

2 3/4 in.-1 3/8 in. 325oF/20,000 psi

2 1/8 in. (ceramic) 300o F/15,000 psi

2 1/8 in. (glass) 300o F/7500 psi

1 11/16 in. (ceramic) 300o F/20,000 psi

1 11/16 in. 300o F/13,000 psi

351
*2 in. 500+oF/20,000 psi

**4 in. 500+oF/25,000 psi

*high operating temperature explosives

**high operating temperature explosives, hollow carrier casing gun

Casing Deformation Considerations

Casing deformation or damage is caused by the extremely high pressures generated


by the exploding charges. Fortunately, these high pressures are short-lived. The
degree of damage depends on (1) the gun characteristics and (2) the well conditions.

Retrievable hollow carrier guns (of any size) do not produce casing damage (Bell
l972a, Bell and Shore 1964). This is because the steel carrier absorbs the energy of
the explosion, as evidenced by carrier swelling (usually 0.005 to 0.015 in., or 0.127 to
0.381 mm). Expendable or semiexpendable guns can cause some degree of casing
damage depending on the

amount of explosive
casing thickness and grade
wellbore hydrostatic pressure
degree of casing support (cement sheath thickness and strength)
The effect on the casing may range from deformation (a slight bulging adjacent to the perforation) to
damage (vertical hairline cracks extending from the perforation or extensive bulging and splitting of
the casing).

Bell and Shore (1964) observed the degree of deformation for various well conditions
in laboratory tests done with a 20.5-gram, 2.125-in. (5.4-cm) aluminum case capsule
charge fired at 1000 psi (6900 kPa) and 180 F (82 C). They investigated three
casing support conditions using with new casing in each case:

nonsupported (NS) gun fired inside casing placed within a well with water inside and
outside casing

semisupported (SS) casing surrounded by 0.75 in. (1.9 cm) of 3500 psi
(24,130 kPa) compressive strength cement retained by metal sheath

well supported (WS) same as SS with thick-walled metal retaining sheath

Figure 1 (Casing deformation vs.

352
Figure 1

explosive load (1000 psi at 180o)) shows the effect of explosive load on new J-55, 5.5-in. (14-cm),
l7-lb/ft (25-kg/m) casing. Casing rupture is possible when cement coverage is poor and charge
sizes are 25 grams or larger. Most casing and through-tubing guns employ charges with explosive
loads in the 2.5- to 22-g range.

Casing strength and thickness are important in determining the degree of casing
deformation. Figure 2 illustrates the effect of increasing casing "strength" (defined as
wall thickness times minimum yield strength).

353
Figure 2

As before, we may note the importance of a good cement job in reducing casing
deformation. In worn or corroded casing, the yield strength may be reduced from that
published, and its "strength" reduced accordingly.

Increasing hydrostatic pressure appears to help minimize casing deformation. The


effect is not significant at pressures greater than a few thousand psi.

Cement sheath thickness is a significant factor in determining the degree of


deformation. Figure 3 illustrates the decrease in deformation with increasing cement
thickness. Even a relatively thin cement sheath supports the pipe well when backed up
by the formation.

354
Figure 3

Deformation with a 0.75-in. (1.9-cm) cement sheath and 3 in. (7.62 cm) of packed
sand approaches the minimum "well supported" value.

Cement compressive strength does not appear to be an important factor in minimizing


deformation. Extensive shattering of the cement sheath does not appear to be a side
effect of perforating. In addition, the hydraulic seal of the cement bond to pipe and
formation appears to remain effective if cement compressive strength is at least 2500
psi (17,240 kPa) (Godfrey 1968).

Another effect of perforating is to reduce the collapse resistance of the casing. As


higher shot densities become a more common completion practice, this effect must be
considered. However, even at 8, 0.5-in. shots per foot (24 shots per meter [SPM]) the
reduction in collapse resistance for 7-in. (17.8-cm) casing shot with zero degree
phasing is less than 20%. hen 90 phasing is used, the reduction is negligible. If the
cement job is good and a reasonable drawdown is experienced, casing collapse should
not be a problem.

We may summarize by suggesting that retrievable hollow carrier guns should be


employed if casing deformation must be avoided. If expendable guns must be used for

355
other reasons, the explosive load should be kept as low as required performance
allows.

356
PERFORATING & PRODUCTION PERFORMANCE

API Testing Procedures


In 1962, The API Division of Production set out recommendations and procedures for
the evaluation of well perforators. The procedures were subsequently revised, with
Edition 5 being published in 1991. These procedures (API Recommended Practice 43)
provide a standard means for evaluating perforator performance under surface
conditions, as well as the physical and flow characteristics of the perforations obtained
at elevated temperature and pressure. In its fifth edition, API RP43 is divided into four
sections:

The surface field test (Section 1 Test) is carried out in a section of casing cemented inside
a concrete drum.

The laboratory flow test (Section 2) is carried out using a Berea sandstone
target that is cemented inside a core holder designed to permit flow through the
perforation.

The elevated temperature and pressure test (Section 3) involves (a) firing a
gun into a steel target after exposing the gun to rated temperature conditions
for a specified time period; (b) testing the resistance of gun components to
automatic self-detonation; (c) confirming satisfactory detonation transfer
between guns; (d) testing the pressure resistance of gun components.

The flow test (Section 4) is an optional test that measures perforation flow
performance on a Berea core or other rock sample, under simulated, site-
specific downhole conditions.

The engineer must be cautioned that the actual performance of a given perforator will be
determined by the in-situ conditions and the quality control exercised in the manufacture and
assembly of the gun components. A number of researchers (e.g., Ott et al., 1993) have put forth
methods for relating test data to actual downhole conditions. But while the API tests are not
necessarily diagnostic of how a particular charge will perform in any given formation they can be
used to compare commercially available charges. What follows is a brief summary of the API test
procedures.

Section 1 Testing

The target for this test is prepared by cementing a section of casing inside a steel
drum using a mixture of Class A cement and sand ( Figure 1 ).

357
Figure 1

A ratio of 1 part cement to 2 parts dry sand is used (i.e., one 94-lb sack of cement
with 188 1b of sand and 43.3 1b or 5.1 gal of fresh water). The concrete is required to
have cured a minimum of 28 days and have a minimum tensile strength of 400 psi
(2760 kPa). Six to eight shots are normally fired at once. Bullet guns are fired in air
and jet perforators are fired in water at surface conditions. This procedure may be
repeated for one or more gun clearances. The following values are measured and
reported on a standard form:

casing hole diameter (short axis, long axis, average)

total penetration

burr height

depth to debris or bullet

This test provides no assessment of perforation cleanliness, but it does have value in evaluating
changes in hole size due to changes in clearance. The target approximates actual well conditions
only in that it is a curved casing surface backed up by cement. Some indication of the consistency

358
of charge performance may also be obtained since several tests are reported. However, penetration
at the given cement compressive strength cannot be extrapolated to estimate penetration in other
formation types. In addition, any casing damage observed at surface conditions cannot be related
to downhole conditions.

Section 2 Testing

Two tests are defined under Section 2 of API RP43: one simulating positive
(overbalanced) perforating followed by backflow, the other simulating reverse
(underbalanced) perforating. The positive pressure test is more commonly reported.

The target in each case is a restored-state Berea sandstone core. The effective
permeability to kerosene is measured in a Hassler sleeve unit prior to mounting the
target for the test. The test mounting is shown in Figure 2 .

Figure 2

In the positive pressure test, a 1500-psi (10,350-kPa) wellbore pressure is established


in 9.6-lb/gal (1150-kg/m3) salt water ( Figure 3 , (a)Positive-pressure flow test; (b)

359
Reverse pressure flow test). A core pressure of 1000 psi (6895 kPa) is established,
giving a pressure differential of 500 psi (3450 kPa) into the core sample.

Figure 3

The charge is detonated at the appropriate clearance. Following this, the core sample
is backflowed with kerosene, using a pressure differential of 200 psi (1380 kPa). The
core pressure is 1200 psi (8274 kPa) and the well pressure is 1000 psi (6895 kPa).
Typically a volume of 10 to 50 liters of kerosene is flowed.

In the case of a reverse pressure test, the perforation differential is 200 psi (1380 kPa)
from the core into the well, with 1200 psi (8274 kPa) in the core and 1000 psi (6895
kPa) in the well. Backflow is carried out under these same conditions.

In either instance, after the perforation has been flowed for cleanup, the flowrate is
measured and the "perforation permeability" is calculated using Darcy's law. The core
flow efficiency (CFE) is determined as

where:

ko = original target permeability measured prior to perforating (saturated with salt water and
then flowed with kerosene, generally 100 to 300 md for Berea)

360
kp = perforation permeability measured after perforating and backflow using
kerosene

ki = effective permeability of a Berea sample with an "ideal" (drilled) perforation


of the same length as that in the perforated target

While determination of the core flow efficiency is part of the Section 2 procedure for Edition 4 of API
RP43, Edition 5 recommends the use of flow parameters from the Section 4 tests. In addition to
these flow data, the nonflow parameters shown in Figure 4 are also measured and entered in the
standard API data form.

Figure 4

These nonflow parameters include entrance hole diameter, total core penetration (TCP), and total
target penetration (TTP). Typically, three such tests are carried out, and the results averaged. A
single clearance setting is used for all the tests (either zero or 0.5 in. [1.27 cm]).

The Section 2 test provides some useful data regarding the penetrating ability of the
perforator in a medium-hard sandstone. Some indication of the perforation cleanup
response to backflow is also obtained for a "Berea-like" formation. Although it would be
nice if the CFE provided an absolute indication of the well productivity expected from
such perforations, the test has serious limitations (Bell l972a):

Only a single clearance setting is tested, so no idea of performance with multiphased


guns is obtainable.

Only a single test material (Berea sandstone) is used under idealized


laboratory conditions.

The accuracy of the data is only about +10%.

361
The limited invasion time for the positive pressure test results in optimistic
CFE values.

The entry hole size measured in the Section 2 test is typically larger than that
measured in the Section 1 test because a flat, mild steel plate is used; (the
Section 1 test results are more representative of perforator capability).

Section 3 Testing

The two Section 3 tests are designed to evaluate explosives stability and perforating
system performance under conditions of elevated temperature and pressure.

The first test evaluates the ability of explosives systems to withstand rated
temperature/time conditions without performance deterioration or automatic, self-
detonation of explosive components. This involves obtaining charge-penetration and
entrance-hole data on guns fired into steel targets after exposure to rated
temperature/time conditions, and comparing them with performance at ambient
conditions. It also involves determining the resistance to self-detonation, and
evaluating the success with which detonations can be transferred from gun to gun.

The second test measures the system's ability to withstand rated


pressure/temperature/time conditions without fluid leakage, significant deformation or
physical collapse. This involves testing for seal failure, hardware deformation and
collapse at rated temperature/time conditions, at 1.05 times the rated pressure.

The Section 3 tests are optional; data are neither certified or filed with the API. Section
3 information is available from perforating service companies or equipment
manufacturers.

Section 4 Testing

The Section 4 tests measure a perforation's flow performance under either of two
separate sets of conditions: (1) simulated, site-specific conditions of overburden stress
and pore, wellbore and differential flowing pressures, using Berea sandstone, other
quarry rock or actual well cores, and (2) simulated radial or axial flow at an
overburden stress of 4500 psi, pore pressure of 1500 psi, wellbore pressure of 1000
psi and differential flowing pressure of 500 psi, using Berea sandstone targets.

Data obtained from these tests include Core Flow Efficiency (CFE), which is the
observed flow under a 3000 psi effective stress divided by theoretical flow for the
given test condition, and the Permeability Reduction Factor (PRF), which is the ratio of
perforation-damaged zone permeability to original target bulk permeability. These data
may be used to compare different commercial charges (Bell et al., 1995).

Perforation Productivity Considerations


In a perforated completion there are basically four geometrical parameters that affect
productivity. Figure 1 illustrates them; shot density (SPF or number of flowing
perforations per unit length); perforation depth (penetration into formation); gun

362
phasing (angular displacement between adjacent perforations); and perforation
diameter (within the formation).

Figure 1

Together these parameters combine to create a distinctive flow geometry that creates
a "skin" caused by the convergence of flow into the perforations. This skin is
independent of permeability and can either enhance or impair performance relative to
that of an openhole completion. In addition, Figure 2

363
Figure 2

(Drilling and perforating damage--generally k1>k2>k3>k4) shows that there are


permeability reductions due to drilling damage and the crushing of the formation due
to the perforating process. The total "skin" represents the combined effect of all these
conditions (geometry, drilling damage, perforating damage).

When fluids are flowing into the wellbore at high rates, the linear laminar flow
assumptions for flow according to Darcy's law do not hold. The effect of this "non-
Darcy" flow ("turbulence") is to cause an additional "skin" factor.

The overriding question for completion design engineers is: What are the combined
effects of the various parameters on well productivity or, specifically, on the
productivity ratio (theoretical productivity of the perforated system divided by the
productivity of an ideal, undamaged, uncased wellbore)?

Bell et al. (1972b), Klotz et al. (1974), Locke (1981), McLeod (1983), Tariq et al.
(1984, 1985), and Karakas and Tariq (1991) are among those who have made
significant attempts to understand perforation behavior. Each of these studies
represents a slightly different perspective and are presented in chronological order.

364
Earlier work by Muskat (1943), Howard and Watson (1950), McDowell and Muskat
(1950), Harris (1966) , and Hong (1975) is referenced.

The focus of the work discussed here is on determining the productivity index (PI) of a
perforated completion relative to that of an open hole completion in the same interval.
This comparison of productivities is important to the engineer in determining well
potential and evaluating damage.

Bell, Brieger, and Harrigan (1972b) mathematically represented the perforation as an


ellipsoid using a finite difference technique. Their model also accounted for the radial
flow of fluids into the perforation from the surrounding reservoir rock, rather than the
linear fluid flow pattern established in the API laboratory test system ( Figure 3 ,
Pressure distribution around perforation). They found that the more representative
radial model flowed at a significantly higher rate than the linear model, for the same
penetration.

Figure 3

The researchers realized, however, that the actual perforation in the well is surrounded
by a "crushed zone" of reduced permeability. This crushed zone was assumed to be the

365
reason that measured API CFE values for various perforators generally ranged from 65
to 85%. Using their radial flow perforation model, they compared the performance of a
perforation having a typical crushed zone (calculated to have only 10 to 20% of
original permeability) with the performance of an "ideal" radial flow perforation. They
found that the "perforation efficiency" (PE as opposed to CFE in the linear test system)
was only 25 to 35%. This result was because in the radial system, most of the
pressure drop takes place across the crushed zone, making its influence greater. These
calculated data were substantiated with laboratory tests using a specially designed
core holder that permitted an actual perforated target core to be flowed radially.

These researchers also investigated the influence of varying differential pressures


(positive and reverse) on perforation flow performance. Using their radial flow target
system and typical 2-in. (5.08-cm) and 3 3/8-in. (8.57-cm) perforators, they
determined that perforation efficiency increased with increasing differential pressure
for both positive and negative pressure differentials. They felt that the relatively short
invasion time period was responsible for the efficiency increase with positive pressure
differential.

They also discovered that as the volume of fluid backflow from the perforation
increased, the perforation efficiency increased. Since the debris from the perforator is
discharged with the first 1 to 2 liters of flow, subsequent flow (10 to 50 liters)
apparently improved perforation efficiency by improving crushed-zone permeability.

They also demonstrated that injection into a perforation without backflow after
shooting results in a very low injection rate less than 1/10 the flow rate possible
after normal backflow.

Although the methods used in this study did not account for the effects of shot density
and phasing, there are implications for field completions:

The flow efficiency of perforations is reduced by the lower-permeability compacted zone


surrounding the perforation.

Perforation efficiency (PE) measured using a radial flow system is substantially


lower than that indicated by the linear API test. This calls into question the
applicability of API data.

Perforation efficiency is influenced by the level of differential pressure toward


the wellbore. Some minimum negative pressure differential is required to
achieve maximum efficiency, and this seems to be at least 200 psi (1380 kPa)
for the test systems used.

Perforations are not initially very effective but are cleaned by flow. Some
maximum efficiency is eventually reached after flow has removed debris and
increased the crushed zone permeability. Injection into a perforation prior to
backflow will result in significantly lower injection rates.

Klotz, Krueger, and Pye (1974) wished to relate the CFE measured on an API test core to the
actual flow efficiency of a perforation affected by drilling damage, perforating damage, radial flow
behavior, and nearby perforations (perforation density and phasing). They used finite element

366
computer calculations to model both the linear and radial flow systems with the ability to vary depth
and degree of perforating damage and also, in the radial case, drilling damage.

They defined well flow efficiency (WFE) as the ratio of flow rate into a cased well
through a real perforation in a zone damaged by perforating, drilling, or workover, to
the flow rate into the same well through a clean, ideal perforation of the same depth in
an undamaged zone. Using the computer models of the linear API core system and the
more realistic radial flow system, they investigated the relationship between CFE and
WFE for a variety of perforation damage situations and perforation depths. Figure 4
graphically displays the relationship that results from the following completion
scenario:

Figure 4

well diameter: 6 in. (15.24 cm)

drainage radius: 660 ft (201 m)

shot density: 4 SPF

367
phasing: 90?

damaged zone thickness: 0.5 in. (1.27 cm)

These researchers' results indicate that, for the normal range of CFE values published
for commercial perforators (0.65 to 0.85), the crushed zone permeability ranges from
about 7 to 35% of the undamaged formation permeability. The values for WFE over
the same range are 0.50 to 0.90. It is suggested that the reason these values are
much higher than the 0.25 to 0.35 (PE) calculated by Bell is that these values reflect
the effects of shot density.

The researchers went on to investigate the effects of perforation depth, drilling


damage, and crushed zone thickness on calculated WFE values. They discovered that

when formation damage is present, well flow efficiency is substantially reduced until
perforations penetrate beyond the damaged zone;

when a severely damaged crushed zone exists, well flow efficiency remains
low despite any drilling damage.

Klotz et al. also wished to compare their model's calculations with those of previous investigators
who considered the relative effects of shot density and penetration. In order to do this, they used
the results from their finite element model to calculate the well productivity ratio (productivity of a
perforated well divided by that of an open hole). The results are shown in Figure 5.

368
Figure 5.

Klotz et al. then carried these studies one step further by including the effects of perforation and
drilling damage on the shot density-versus-penetration relationship. Figure 6 and Figure 7 show the
results of these calculations.

369
Figure 6

These curves show that an increase in shot density is more beneficial when the perforation is
damaged than when it is undamaged.

370
Figure 7

Also, increasing shot density alone cannot overcome the combined effects of permeability damage
from perforating and drilling or workover fluids. Deeply penetrating perforations are required if the
productivity of a damaged well is to approach that of an undamaged openhole completion.

Locke (1981) corroborated the work of earlier researchers like McDowell and Muskat
(1950) and Howard and Watson (1952), whose research was based on physical models
using tanks of electrolyte, and the analogy between fluid flow and electricity; and of
Harris (1966), Bell (1972b), and Klotz et al. (1974), who used mathematical
techniques to model perforations. Hong (1975) had expanded the work of Harris to
consider wellbore and perforation damage and phasing, and presented a nomograph
for determining skin and productivity ratio. Locke built on the work of these earlier
researchers and extended the relationships using a more refined finite element model.
In addition, Locke presented a nomograph for determining the skin factor and
productivity ratio for a given perforation if the following are known (or estimated):

perforation length
perforation diameter
damaged zone thickness and permeability
crushed zone permeability
formation permeability

371
phasing
shot density

In general, Locke's work further substantiated the fact that

productivity is improved with increasing shot density up to at least 8 SPF (26


SPM)
penetration is more critical than perforation diameter above diameters of 0.25
in. (6.35 mm)
90? phasing is most effective (rather than 0, 120?, 180?)
damaged zone penetration is critical

McLeod (1983) took a step back from the increasing complexity of the finite element perforation
models. Specifically concerned about the effects of perforating efficiency on high-rate gas wells,
McLeod applied the idea of radial turbulent flow in the compacted zone around each perforation,
and came up with a simplified model for determining the productivity of such completions. Using
McLeod's equations, an engineer can evaluate the turbulent flow coefficient in the radial gas flow
equation, based on his or her knowledge of completion geometry. McLeod also presents a
procedure for separating the geometrical, drilling damage, and perforating damage components of
the skin factor. These equations are easy to program for computer analysis of gaswell productivity,
but are not generally as accurate as the nomographs of Hong or Locke.

Tariq (1984)employed a more sophisticated finite-element model to evaluate the


effects of non-Darcy flow ("turbulence") in high-rate wells. Tariq qualitatively
confirmed Locke's earliest findings, but determined that his nomograph was 5 to 10%
optimistic in its prediction of productivity ratio. Another conclusion was that angular
phasing between adjacent perforations and perforation length was more important
than shot density in reducing the pressure drop in high-rate wells. Tariq's conclusion
was that turbulence is responsible for a portion of the pressure drop historically blamed
on permeability damage in the crushed zone. The importance of angular phasing in
completion design was highlighted by Tariq's analysis, particularly for high-rate
completions.

Figure 8 (Productivity vs.

372
Figure 8

perforation length) compares Tariq's analysis with earlier researchers. Even without
the result of non-Darcy flow effects, Tariq's estimates of productivity ratio are lower
than those of Hong (Harris) and Locke. With high rates the difference is even more
significant.

Figure 9 shows the effect of angular phasing on productivity based on Tariq's analysis.
The graph indicates that angular phasing is an important parameter influencing
productivity.

373
Figure 9

Roughly 20% improvement in productivity can be expected by perforating with 90?


phasing rather than 0? phasing, given the same penetration. We also see that with
angular phasing, improvements to productivity by increasing shot density are
magnified.

Figure 10 shows that, for perforation diameters above 0.4 in.

374
Figure 10

(10 mm), there is very little productivity improvement with increasing diameters.

All the studies mentioned up to this point have assumed that the formation being
perforated was homogeneous and isotropic, at least near the wellbore. In most real
situations this is not the case. Tariq, Ichara and Ayestaran (1985) used finite element
analysis to evaluate the relative importance of various perforating parameters when
perforating laminated formations (sand-shale sequences) and various horizontal and
vertical fracture systems. The principal conclusions were that

high shot densities are important in anisotropic and laminated formations

perforation performance varies with the type, orientation, and interval of


natural fractures

different perforation parameters (penetration, shot density, phasing) assume


different significances for different fracture systems.

Figure 11 (Productivity vs.

375
Figure 11

perforation length for anisotropic reservoir (kv<kh)) illustrates the importance of increasing shot
density when vertical permeability declines. Figure 12 (Productivity ratio vs.perforation length for
various phasings in anisotropic reservoir) illustrates the relative insignificance of phase angle in
such a situation.

376
Figure 12

The types of fracture-matrix systems studied by Tariq et al. are shown in Figure 13.

377
Figure 13.

The most commonly used representation of a naturally fractured reservoir is the case
of three sets of mutually perpendicular fractures ( Figure 13 d).

Figure 14 (Three sets of mutually orthogonal fractures) shows the importance of


fracture block size (fracture interval) in determining productivity.

378
Figure 14

Figure 15 shows the relative insignificance of shot density, especially when large
matrix blocks are present.

379
Figure 15

Figure 16 illustrates the fact that, for small fracture blocks, perforated completions
perform better than openhole completions, particularly when penetration is increased.

380
Figure 16

Karakas and Tariq (1991) developed a step-by-step, semianalytical procedure for


approximating skin in a perforated completion, incorporating simple algebraic
expressions for determining the discrete components of total skin, and combining
these components into an expression for total skin. Using finite-element models, they
presented results for two separate cases:

two-dimensional, steady-state flow, through which they established the dependency of


perforation skin on gun phasing, penetration and wellbore radius

three-dimensional flow, in which they quantified vertical flow effects in terms


of a vertical pseudoskin, and then extended the model results to include the
effects of reduced permeability in the crushed zone and formation anisotropy.

They then examined the combined effects of formation damage and perforations on well
productivity.

The semianalytical expressions developed in this study indicate that at conditions of


negligible vertical flow, a perforated completion's productivity is governed mainly by

381
the perforation penetration and phasing, with multidirectional perforations offering
significant advantages over unidirectional (0o) perforations. Where vertical flow is
significant, converging flow effects can be quantified in terms of a vertical pseudoskin
based on the dimensionless perforation spacing and radius. Effects resulting from the
crushed zone around perforations can also be quantified in terms of an additional
vertical skin. For wells where the perforations terminate within the damaged zone, this
study also provides relationships for combining openhole damage skin with perforation
skin. Because the productivity relationships presented here are verified extensively
with a large set of simulated productivity data, they are considered reliable estimates
of skin in perforated completions.

Relative Importance of Perforation Parameters


The relative importance of the four basic perforating parameters (shot density,
penetration, hole size, phasing) depends on the type of formation, the expected
productivity, and the degree of damage. We can, however, state some general
guidelines:

High shot density is most important, particularly in anisotropic or laminated reservoirs.


Penetration becomes more important than shot density the farther that drilling damage
extends into the formation. Hydraulically fractured wells are an exception, in that perforation
length is minor compared to the fracture length.
Some type of angular phasing is important when high rates are expected and, in general,
phasing will improve productivity by 10 to 20%, although the angle (90, 120, 180) is not
critical. Generally speaking, phasing should not be gained by sacrificing penetration, unless
some type of computer simulation supports such a decision.
Hole size is relatively unimportant beyond a minimum diameter of about 0.25 in. (0.64 cm)
unless the well must be gravel packed or fractured, in which case hole size becomes
critical.

It is important to emphasize that shot density is defined here as the number of perforations open to
flow per unit length. Ensuring that the maximum number of clean perforations is open to flow
becomes very important when shot density is shown to be critical to overall productivity.

Perforation Cleanup Considerations


According to Bell (1982b), the factors influencing the number of clean, open
perforations are

type of formation
quality of the charge
type of completion fluid
level and direction of differential pressure
flow time for cleanup
The type of formation will influence the depth and competency of the perforation and the amount
and severity of rock crushing, with the general trend being reduced penetration with increasing
strength and overburden stress. Charge quality will influence the uniformity of the perforation tunnel
and the amount and type of debris that must be removed.

382
Positive-pressure (overbalanced) perforating can result in damaged and/or plugged
perforations. Figure 1 illustrates the effects of perforating overbalanced in drilling mud.

Figure 1

When such a well is swabbed in, the pressure reversal opens up a few perforations in
the better permeability zones. This in turn reduces the differential across the remaining
plugged perforations, leaving them plugged. The effective shot density is reduced and
the productivity seriously curtailed, even though penetration depths may be large.
Postperforating acidizing is one remedial method for improving effective shot density,
but is not 100% effective.

If clean, solids-free, formation-compatible fluids are used when perforating


overbalanced, the degree of damage can be decreased, especially when special
backflow techniques are used to clean up the perforations (backsurging). Regardless of
the type of fluid in the hole, the use of tubing conveyed perforating guns or through-
tubing perforating (with the wellhead installed) can greatly enhance perforation
cleanup by making it possible to perforate underbalanced (Bell 1982a). Using these
methods to attain a minimum "threshold" value of underbalanced ensures perforation

383
cleanup and avoids the problems that might occur with excess underbalance (sanding,
fines migration, equipment failure, casing collapse, safety risks). This threshold level
varies with formation type, porosity, and permeability. Bell has suggested the figures
given in Table 1, below, based on empirical data. Other researchers have developed a
correlation between sonic travel time and maximum underbalance for unconsolidated
formations where the risk of sanding is greatest ( Table 2 , Selection of underbalanced
pressure differential for tubing-conveyed perforating.) King et al.

Table 2

(1985) have suggested that the correlation between formation permeability and
underbalance is as shown in Figure 2 (Underbalance vs.

384
Figure 2

formation permeability showing distribution of high effective shot density completions


in oil wells)and Figure 3 (Underbalance vs.

385
Figure 3

formation permeability showing distribution of high effective shot density completions


in gas wells). Those wells in which acid did not show improvement are assumed to
have been efficiently perforated due to proper underbalance. This correlation, based on
information from 90 wells, generally corroborates Bell's ranges. An important point
highlighted by Figure 3 is that below 2 md, much higher underbalance is required to
obtain clean perforations. This could be because at low permeabilities there is
insufficient flow to clean the perforations, regardless of the magnitude of the
underbalance.

Liquid Gas

Higher Permeability 200-500 psi 1000-2000 psi


(>100 md)

(1.38 to 3.45 MPa) (6.9 to 13.8 MPa)

Lower Permeability 1000-2000 psi 2000-5000 psi


(<100 md)

386
(6.9 to 13.8 MPa) (13.8 to 34.5 Mpa)

Table 1. Typical range of reverse pressure differential for perforation cleanup.

Bell (1972b) has pointed out that a minimum volume must be backflowed from
perforations in order to remove perforating damage. This is particularly important if a
well is to be injected into or to be subjected to overbalanced pressures prior to being
put on production. Thus, it is necessary to flow the well for a sufficient time to remove
debris.

Methods for achieving clean perforations include

through-tubing wireline perforating (underbalanced)


tubing-conveyed perforating (underbalanced)
positive-pressure/reverse surge technique
positive-pressure/perforation washing
We shall review each procedure. Note that through-tubing wireline perforating and tubing-conveyed
perforating are the only safe methods for perforating underbalanced.

Through-Tubing Wireline Perforating Techniques

In this technique, the tubing and packer are run and set in the cased, unperforated
well. After pressure control equipment is installed and tested, the appropriate
differential pressure is obtained by circulating fluids into the tubing. Wireline plugs can
also be used to allow tubing to be run and set with a minimum fluid cushion inside. The
gun assembly (carrier, weights, positioning tool, collar locator) are run into the tubing
on small-diameter cable (0.18 to 0.22 in. [4.6 to 5.6 mm]). After firing, the well is
typically flowed for 15 to 30 minutes before the gun is recovered. If multiple gun runs
are required, it is necessary to flow the well after each successive gun is positioned to
provide sufficient drawdown of the bottomhole flowing pressure before firing. When a
range of permeabilities exists over a long productive zone, it is preferable to perforate
low-permeability intervals first, followed by the other intervals of higher permeability.
In this way, high bottomhole backpressure will not prevent the low-permeability zones
from flowing with sufficient drawdown.

In very high permeability oil wells, the well may be flowed just long enough for the
tubing to fill up. For example, a 10,000-ft (3048-m) well flowing 5000 BOPD (795
m3/D) through 2 7/8-in. (73-mm) tubing will displace the tubing volume in about 15 to
20 minutes. The gun is then removed prior to flowing the well.

When perforating underbalanced with through-tubing wireline equipment, it is


important to prevent the carrier and cable from being blown up the hole when flow
suddenly begins. This is caused by the frictional forces imposed by the flowing cushion
fluid and well fluid. The tendency to move the gun carrier and cable will depend on the
flow rate, fluid properties, dimensions of the gun and well tubulars, and the weight of
the tool string and cable. Prevention of gun movement can be accomplished by
maintaining a minimum amount of tension in the cable. Once the cable is no longer
stretched taut, the frictional forces acting on the buckled cable rapidly create an
unstable situation. This minimum cable tension can be maintained by controlling the

387
drawdown through a combination of liquid cushion and surface tubing pressure.
Because liquid movement creates the frictional forces that move the gun, smaller liquid
cushions coupled with higher tubing pressures will allow larger drawdowns while
maintaining the minimum cable tension. Garcia and Paslay (1981) have developed a
method for determining the various combinations of cushion height and tubing
pressure that will maintain a minimum of 50 lb (23 kg) of cable tension in gas wells
perforated with a brine cushion. Below, Table 3 gives a range of combinations for a
specific perforating scenario. In medium-rate oil wells the velocity of fluid movement is
generally less than in many gas wells, allowing somewhat smaller liquid cushions (e.g.,
for the case given in Table 3, an undersaturated oil rate of 1000 STB/day [159
m3/day] would result in a bottomhole fluid velocity of about 0.8 ft/sec [0.24 m/sec],
while a gas rate of 5 MMSCFD [1.64 m3/sec] would result in a velocity of about 1.5
ft/sec [0.46 m/sec]).

Pressure/Cushion
Combination

Drawdown Wellhead Pressure (psi) Cushion Height (ft)

1000 0 7500

2000 850 4000

3000 1550 1000

The above range of values assumes the following:

productive zone 30 ft

depth 10,000 ft

permeability 10 md

formation pressure 5000 psia

casing 5 1/2 in., 17 lb/ft

tubing 2 3/8 in., 6000 ft

cushion saturated brine (0.52 psi/ft)

cable 1-22 monocable

gun 1 11/16-in., 250-lb buoyed


weight

Table 3: Wellhead pressure and cushion height required to maintain 50 lb of cable


tension when perforating through-tubing.

Tubing-Conveyed Perforating Techniques

388
In this technique, a large-diameter casing gun is run into the well on a tubing string
that includes a packer ( Figure 4 ).

Figure 4

Since a large-diameter gun is used, gun phasing can be designed to reduce "skin" (90
to 120) without sacrificing perforation penetration. The tubing may be run dry or
partially filled with a fluid cushion to establish the proper level of underbalance. The
packer is set.

After positioning, a vent is opened to equalize the pressure below the packer with the
tubing. One of a variety of firing devices may be used to detonate the charges, after
which flow is established through the vent. The gun may be dropped into the rathole.
The use of large, hollow carrier guns allows penetration depths that are typically 2.5 to
1.8 times the depth achieved using comparable through-tubing guns. Entry hole
diameter may be 1.7 to 1.3 times the diameter achievable using comparable through-
tubing guns (Bell l982a). This increase is achieved along with the added advantage of
underbalanced conditions to maintain clean perforations.

389
Productivity resulting from tubing-conveyed underbalanced perforating is generally
equal to or better than that resulting from the through-tubing wireline method (Bell
1982a). The more extensive the wellbore damage, the greater the margin of
improvement.

Compared to through-tubing perforating, we can say that tubing-conveyed perforating


(Bell l982a)

allows greater penetration and multiphasing

permits a greater degree of underbalance without risk of blowing the gun up


the hole

demands operational rig time equal to or less than wireline perforating,


particularly in deviated holes with long pay intervals

generally costs about 25% more than wireline perforating (without accounting
for possible time savings)

does not allow for a positive indication of individual charge detonation

The tubing-conveyed technique also permits the perforating of long intervals in highly deviated
wells. In certain offshore California locations, 1000-ft (300-m) sections are perforated at 12 SPF,
overbalanced. The gun is then pulled and the perforations are washed prior to gravel packing.

Positive-Pressure/Reverse Surge Technique

This method (sometimes called PACT or Positive Action Completion Technique) involves
first perforating the well under conventional overbalance conditions using large-
diameter casing guns. The tubing and packer are then run into the hole with some type
of shear-disc assembly included in the tubing string. With this disc, the tubing may be
run dry or with a water cushion. Using either a drop bar or casing pressure-operated
mechanism, the disc is sheared, exposing the perforation to a sudden drawdown, or
"surge."

This technique has generally the same advantages of the tubing-conveyed technique,
with some gains possible in cost savings and simplicity. Proper selection of the shear
disc (depending on cushion pressure) is important. A tubing plug may be run
immediately above the disc and recovered using a slickline once the packer has been
set. This procedure can protect against premature rupture of the disc.

Recent work by Bonomo and Young (1985) has compared the effects of the "perf-
surge" technique with tubing-conveyed perforating and another alternative, the "perf-
wash" technique, where the conventionally shot perforations are washed using a cup-
sealing assembly that forces a clean, nondamaging fluid into one set of perforations
and out of another ( Figure 5 ).

390
Figure 5

The study compared the performance of 45 comparable gravel-packed wells with


different completion practices (20 "perf surge," 20 "perf-wash," 12 tubing-conveyed).
Each well was assigned a "completion index" depending upon its ability to match a
theoretical maximum flow rate with test data. The tubing-conveyed technique was
shown to be the superior technique over surging or washing. TCP "completion indices"
were, on the average, twice as good as "perf-wash" indices, and 3.3 times as good as
"perf-surge" indices.

Perforating for Gravel Packing


In an ideal gravel-packed completion, the perforation tunnel is filled with a clean,
evenly sorted, high-permeability sand or "gravel." The linear flow of formation fluids
through the packed tunnel results in a substantial pressure drop. If the perforation
diameter can be increased, the pressure drop is reduced, permitting greater flow.
Gravel-pack completions are typically shot with large hole diameters (0.6 to 0.8 in. [15
to 20 mm]) and high densities (8 to 12 SPF [26 to 39 SPM]) (Bell 1982a).

391
In a typical 7- to 9 5/8-in. (178- to 244-mm) casing gravel-packed completion, either
a 5-in. (127-mm) or 6-in (l52-mm) gun is used. The 5-in. hollow carrier gun is fired at
4 SPF (13 SPM) and additional runs are made to the same interval to achieve the
desired shot density (8, 12 SPF). Maximum gun length is about 40 ft (12 m) per trip.
Over short intervals, selective firing equipment can be used to obtain high shot
densities without repeated trips (Bell 1982a).

The 6-in. (152-mm) "slick-bodied" gun fires 12 SPF (39 SPM) at one time, phased at
60? or 90?, with entry hole diameters of 0.6 to 0.8 in. (15 to 20 mm). These guns can
be run on tubing if long intervals must be perforated.

When high underbalance is used in a very unconsolidated formation, some operators


prefer to limit perforator size to allow for wash pipe to be run to recover a stuck gun.
In such cases 4- to 4.5-in. (102- to 114-mm) guns are run in 7- and 7 5/8-in. (178- to
194-mm) casing (Toney 1986). Perforator size may also be limited by packer bore
diameter in cases where guns are run on tubing.

The importance of maximizing perforation diameter when perforating for gravel


packing rearranges the priority listing established for perforating a "natural
completion." In this case, the order becomes (1) shot density, (2) perf diameter, (3)
gun phasing, and (4) penetration. As mentioned earlier, minimizing clearance is
essential to maximizing penetration. In deviated holes, the effect of decentralized
perforating (where the gun lies on the low side of the casing) on phased guns is to
decrease the average perforation diameter ( Figure 1 , Variations in high and low side
perforation diameters in deviated holes).

392
Figure 1

The alternative in these cases is to perforate using an oriented gun, which directs the
perforations against the low side (either 0? or 120? phasing) ( Figure 2, Oriented,
phased perforating for gravel packs in deviated wells).

393
Figure 2

Perforating for Fracturing


Perforations are a key consideration in the design and execution of hydraulic fracture
treatments. Bell et al. (1995) identify the following perforation parameters as
particularly significant:

casing entrance hole size

effective shot density

gun phasing

character of perforations in the formation

Casing entrance hole size: The friction pressure loss across perforations may be expressed as

pf=0.2369 qt2 o/n2d4Cd2

where pf=friction pressure in psi

Cd=discharge coefficient (dimensionless)

394
d=casing entrance-hole diameter, inches

n=number of perforations

o=liquid density, lbm/ft3

qt=total flow rate through perforations, bbl/min.

At a given flow rate, smaller-diameter perforations give a significantly higher friction


pressure loss than do larger diameter perforations. A larger frictional pressure loss
requires a larger bottomhole pressure to initiate a fracture. In multiple-zone fracturing,
this high pressure in the casing will initiate a fracture in the next zone. Therefore,
smaller perforations maximize the pressure loss at lower pump rates, minimizing the
hydraulic horsepower requirements (and thus the cost).

In selecting a minimum perforation diameter, however, we must consider the size of


the proppant so as to avoid bridging at the perforations. Also, the perforations must be
large enough to avoid shear effects that could lead to degradation of low-viscosity
fluids and subsequent proppant fallout. Typically, perforation diameters range from
0.375 to 0.5 inches (9.5 to 12.7 mm), although smaller diameters have been used in
some cases.

During a fracture treatment, the diameter of the casing entrance-hole tends to


increase because of erosion. In addition, erosion smooths out the burrs and
irregularities that are initially present, causing an increase in the value of Cd, the
discharge coefficient. These increases lead to a progressive decrease in the friction
pressure. In multiple-zone treatments, the pumping rate needs to be adjusted to
compensate for this growth.

Effective shot density: Shot density requirements are based on maintaining reasonably
low breakdown, treating and initial shut-in pressures. Typical treatments employ 4-8
shots per foot.

Gun phasing: Regardless of perforation orientation, the far-field orientation of fractures


depends entirely on formation stress charateristics. From the standpoint of maintaining
reasonably low treating pressures, it is important to minimize the angle between the
perforations and the plane normal to the minimum far-field stress (Behrmann and
Elbel, 1991). For typical fracture treatments, a phasing of 20o to 60o is recommended,
while limited-entry treatments commonly use 0o shaped charge guns for placement of
individually located perforations.

Character of perforations in the formation: Because fractures are generally initiated


from the base of perforations, depth of penetration is not as important a parameter as
it is in an unfractured completion. An intermediate penetration of about 4 to 6 inches is
generally adequate (Bell et al., 1995).

Highly Overbalanced Perforating

395
Although underbalanced methods have become generally accepted as the petroleum
industry standard for perforating oil and gas wells, some studies have indicated that
they are inconsistent in providing clean, open perforations (Dees, 1995). Highly
overbalanced perforating, in which the pressures applied to the perforations are equal
to or greater than formation fracture pressure, represents an alternative to
conventional techniques.

Highly overbalanced perforating offers two advantages: stabilization of the tunnel walls
resulting from prolonged application of high pressure, and the initiation of fractures at
the tips of the perforations. The use of a reactive fluid (e.g., acid in a carbonate
reservoir) or scouring material (e.g., fracturing proppant) may provide an added
benefit by etching the face of the created fracture. A scouring material may also help
to erode and enlarge the perforation entrance-hole diameter.

Before perforating, pressure is applied to the wellbore using an all-gas, all-liquid, or


combination gas-liquid fluid column. Perforating can be done with any type of gun,
although a hollow steel carrier is recommended because it will retain most of the
charge debris. The highest bottomhole pressure will be attainable through tubing-
conveyed perforating assemblies. If the overbalance is high enough, the rate of fluid
displacement will exceed the capacity of the perforations to accept fluid. Additional
fluid can be pumped into the formation to enhance breakdown. The highly
overbalanced stimulation technique can also be applied through existing perforations
by surging the perforations through tubing (Dees, 1995).

396
ELECTRIC WIRELINE PERFORATING

Electric Wireline Equipment


Most companies that offer perforating services also offer a variety of openhole and
cased hole logging services, and much of the basic equipment for performing any of
these wireline operations is the same, with the important addition of a pressure control
component in the case of wireline perforating. Figure 1 shows a typical setup for a
perforating job.

Figure 1

The primary components of the system are:

the perforating tools (gun cable head, collar locator, weights);

the cable;

the sheave and tension indication apparatus;

397
the logging truck (including cable winch and control panels);

the pressure control system.

The perforating guns described earlier in this manual are run into the hole along with a device for
detecting the position of each casing collar. The casing collar locator is a simple magnetic device
that responds to the increase in metallic mass associated with a casing collar coupling, and
transmits this indication to the surface where it is recorded along with depth. The casing collar
locator (CCL) is run above the perforating guns and weights ( Figure 2 ).

Figure 2

Weights must be added in cases where the well pressure acting against the cross-sectional area of
the cable is great enough to prevent the guns from traveling down the hole. Such weights may have
an electrical conductor built into them to transfer firing power to the guns, or they may simply allow
a conductor to be threaded through them. The cable head anchors the cable and attaches to the
perforating assembly by means of a threaded ring. The head provides the electrical connection
between the conductor cable and the perforating tools. Most perforating assemblies are run on a
monoconductor cable, with a single insulated conductor, as opposed to the multiconductor cables
used for well logging, which contain six or seven individually insulated conductors. The standard
cable sizes are shown in Figure 3 .

398
Figure 3

The sheaves used to suspend the cable in the well are installed above the pressure
control assembly and at the base of the wellhead or rotary table. Cable is passed
around the sheaves and attached to the perforating tool assembly before it is placed
inside the pressure control assembly. The top sheave is suspended from the rig
elevators if the well is being perforated with a rig on location. If the well is being
perforated through tubing after the rig has moved off location, some type of crane or
mast is required to install the pressure control assembly and suspend the cable.
Between the top sheave and this suspension equipment is a tension measuring device
that measures the strain on the cable and displays it in the logging truck.

Modern logging trucks generally include:

a drum of conductor cable and winch to raise and lower the cable in the well (this alone
may weigh more than 4 tons);

a depth measurement device on the cable;

a self-contained 120-volt AC generator;

399
electrical control panels;

a recording mechanism (magnetic tape and/or film).

Offshore units are generally diesel operated or designed to run from rig power, and are skid
mounted for easy transport. They may be permanently welded or bolted to the deck of a drilling
vessel or platform.

Pressure Control System


Figure 1 shows a schematic of a typical pressure control assembly.

Figure 1

It consists of a blowout preventer (BOP), a riser pipe to contain the perforating tools, a
flow tube assembly, and an upper seal or stuffing box (also called a "packoff").

The BOP encases two rubber-faced rams with cable-sized cutouts. When closed, the
rams seal tightly around the cable. Either mechanically or hydraulically operated BOPs

400
are common ( Figure 2 ).

Figure 2

The riser pipe consists of sections of pipe with quick-connect couplings. The total
height of the riser is determined by the length of the tool assembly. When cable is
moving into or out of the well under pressure, the flow-tube assembly provides the
offsetting force that prevents well fluid from flowing out of the riser. Grease is
continuously pumped into the small annular space between the cable and flow tubes,
and because the annular area is so small, the grease pressure required to balance the
well pressure is fairly small ( Figure 3 ).

401
Figure 3

At the top of the assembly is a hydraulically actuated packoff. The hydraulic pressure
forces a piston upward against a spring, compressing a rubber ring around the cable (
Figure 4 , Wireline packoff). This device is used in an emergency or when the cable is
held stationary for a long period of time. The cable is not free to move when the
packoff is sealed.

402
Figure 4

This equipment generally is designed for either low-pressure, high-pressure, or very


high pressure operations. Table 1, below, gives the specifications for these categories.
In practice, the pressure control assembly is tested to 1.2 times the expected well
pressure, after being mounted on the well.

Max. Working Test Pressure, Cable Diameters Possible (in.)


Pressure psi (MPa)
psi (MPa)

Low Pressure 5000 (34.5) 10,000 (69) 3/16, 5/16, 7/32

High Pressure 10,00 (69) 15,000 (103.5) 3/16, 7/32

Very High Pressure 15,000 (103.5) 22,500 (155) 3/16

Table 1: Pressure control assembly specifications.

403
In high-pressure situations (>5000 psi [34.5 MPa]), a special dual-ram grease-sealed
BOP can be used. This type has two sets of hydraulically operated rams in separate
chambers. Pressurized grease can be injected to prevent leakage around the ram jaws,
particularly in gas wells.

In low-pressure situations, especially when perforating overbalanced, the BOPs and the
flow tube assembly may not be required. However, one should be absolutely certain of
the expected pressures before taking any risks.

Optional items that may be found within the pressure control assembly include:

a line wiper, mounted above the packoff and used to squeeze the grease and oil from the
cable as it is pulled from the well;

a ball-type safety valve, which automatically shuts in the well should the cable
detach from the tools and be blown from the well;

a tool catcher, a hinged door mounted within the assembly below the riser to
prevent tools from falling back into the well should the cable detach when the
tool string enters the pressure control assembly.

Depending on the perforating situation at hand, the pressure control assembly will be connected to
a
drill pipe
drilling system rotating head
annular BOP stack
wellhead flange
tubing head flange
Christmas tree
In certain situations the riser might also be stabbed into and held by the annular BOP on a drilling
rig, although this is not recommended. Most often, the flange on the bottom of the pressure control
assembly BOPs will be bolted to a matching flange on the rig annular BOP, the tubing head, or a
special adapter added above the swab valve on the Christmas tree. This last arrangement allows
the well to be flowed after perforating underbalanced through tubing.

Depth Control Procedures


Perforations do absolutely no good if they are not made where they are required
within the productive zones of the well. Inaccurate depth control is responsible for a
sizable number of misplaced perforations, and the resulting poor productivity.
Reperforating zones and/or squeezing off unwanted perforations are expensive
remedial measures that can be avoided by careful depth control procedures.

The basic problem in depth control is the accurate positioning of the perforating gun
relative to the productive zones as shown on the openhole logs run prior to setting
casing. Depth measurement on these logs is related to a specific reference point,
usually the kelly bushing or rotary table. This reference is noted on the log heading,
along with its position relative to sea level and a "permanent" datum (casing head
flange, and so on). The recorded log depth is corrected for cable stretch and is checked

404
using magnetic markers on the cable. Openhole logs depths are generally accurate to
within about 3 ft in 10,000 ft.

It would not be sufficient to rely on direct line measurement to perforate after casing
has been run. Cables stretch to varying degrees when run inside casing or in the open
hole.

The standard procedure for depth control is to log through the casing using a log that
can be correlated with the openhole logs run previously. This log (normally a gamma
ray or neutron or both) is typically run along with a casing collar log. Thus we can
relate the openhole log, which indicates hydrocarbon, to the openhole gamma ray; the
cased hole gamma ray to the openhole gamma ray; and the casing collar log to the
cased hole gamma ray. Figure 1 (Perforating log showing collars recorded 3 ft deep
due to tool sensor placement) shows a simplified schematic of a gamma ray and casing
collar log, with the gamma ray tied in to the openhole measurement.

Figure 1

Due to the displacement of the gamma ray and casing collar measuring points, the

405
casing collars will be recorded off-depth (in this case 3 ft deep), and must be corrected
on the log.

These corrected collar depths are used to properly position the gun by running a collar
locator within the gun assembly. Collars are matched on the gun run and the
perforator positioned for firing. The displacement between the casing collar locator and
the perforating charges must be accounted for by subtracting the distance from the
depth odometer reading.

As indicated in Figure 1 , a distinctive pattern of irregular length casing joints within or


near the pay zone will help to eliminate difficulties in identifying casing collars.

An alternative service is offered by some companies, where the perforating gun is run
with a gamma ray tool, eliminating the casing collar correlation altogether. Generally,
however, the procedure for gun positioning and perforating is as follows:

1. Run cased hole log with gamma ray, neutron or pulsed neutron, and casing collar
locator. This will be referred to as the perforating depth control log (PDC).

2. Correct collar depths to match radioactive curve depths on PDC log.

Know previously measured distance between radioactive recording point and


casing collar locator on tool.

If CCL is above radioactive recording point, collars will be recorded too


deep on the log.

If CCL is below radioactive recording point, collars will be recorded too


shallow.

If curves are "memorized," be certain the signal from the highest tool
is "remembered" the correct distance to the lower sensor. This feature
will give a corrected collar trace on the log.

3. Correlate reference openhole log to PDC log.


Mark desired perforation intervals on reference log.

Overlay PDC log and shift until good correlation is seen between
radioactive curves.

Trace perforating depths onto PDC logdepths may not be the same
numerical value as shown on the reference log.

4. Make sketch of perforating tool assembly ( Figure 2 , Example of gun dimensions).


Measure distance from CCL to top shot of each carrier.

406
Figure 2

Measure distance from CCL to bottom of assembly.

Mark top shot depths from PDC log on sketch.

Mark "stop" depths on sketch for each firing sequence (relative to


CCL).

5. Zero CCL with respect to permanent depth datum.

6. Run assembly into well to point below bottom shooting depth.

7. Log uphole, recording sufficient casing collars for correlation.

8. Correlate casing collar log with corrected collars on PDC log and check for
depth discrepancy.

9. Make depth correction to odometer if necessary and rerun CCL log.

407
10. Correlate casing collar log with corrected collars on PDC log (check match
by sliding casing collar log up and down one joint to see if any other nearby
match is possible; here is where odd-sized casing joints help prevent errors).

11. Check for gun "creep" (movement after cable winch is stopped) and correct
if necessary.

12. Position gun for shooting.

Drop below zone and log collars upward.

Stop when depth odometer indicates correct depth for CCL when first
gun top shot is in correct position.

Fire gun.

13. After firing, continue to log upward for several collars to ensure that the
gun was fired at correct depth and that collars were properly tied in.

14. Maintain all documentation, logs, sketches, and calculations for file.

Positioned or Oriented Perforating


In addition to perforating at the correct vertical depth, we may also be concerned that
the perforating gun be positioned against the casing to minimize clearance in 0
phased through-tubing perforating, or we may wish to orient the gun to control the
direction of the perforations.

Positioning tools are normally electromagnets that are run above the charge carrier
and activated from the surface. The diameter of the positioning device must not be
less than that of the carrier if it is to function properly.

Oriented perforating may be undertaken for two reasons:

perforating the upper zone through tubing in a multiple completion (between packer
perforating;

perforating a single string in a multiple tubingless completion.

In the first case, a mechanical orienting tool is often used ( Figure 1 ).

408
Figure 1

The tool has a spring-loaded arm that extends automatically as the gun assembly drops out of the
tubing. As the gun passes alongside the adjacent tubing string, the arm ensures that the gun is
positioned at an angle away from the tubing. A switch within the tool gives an indication at the
surface that the gun is facing correctly and prevents the gun from being fired unless the orientation
is correct. The arm also acts to position the gun against the casing. A retrievable, hollow carrier gun
is used to prevent damage to the adjacent tubing and to prevent debris from settling inside the
casing on top of the lower packer.

For perforating multiple tubingless completions, a motorized orienting tool is used. This
tool ( Figure 2 ) uses a focused source of gamma rays and a focused detector to
measure the density of the material between them.

409
Figure 2

A motor rotates the source, detector, and guns. The higher the density, the more
gamma rays are absorbed and the lower the detector count rate. When displayed on
coordinate plot, the correct compass direction for firing is determined.

If the radiation pattern is difficult to interpret, a radioactive pill can be placed in the
adjacent casing string(s) to override the source signal and provide positive location of
the strings.

Job Planning
The first step in planning a perforating job is to accumulate the necessary data. Of
primary importance is the configuration of the well completion tubulars, including

casing and/or liner size and weight

tubing size and weight

all tubing restrictions, their ID and location

410
the type of packer and the pressure it can withstand without unsealing

the specifications of the wellhead or blowout preventer connection, flange and


type

pressure test specifications for tubing, casing, and wellhead equipment

the condition of the casing or tubing, excessive wear, crooked tubing,


corrosion

Other well conditions that must be specified or estimated include


the expected bottomhole temperature

the formation pressure

the bottomhole pressure (the amount and direction of pressure differential)

the type of fluid in the wellbore and in the formation

the fluid level in the tubing or casing

any surface pressure to be added to the fluid in the tubing or annulus

pressure and rate specifications of surface facilities if well is to be flowed

anticipated surface flowing pressure and rates

anticipated H2S concentration, if any

formation type

Of course, the openhole log used to pick the perforating intervals will be needed, with the intervals
clearly marked and the depths recorded. If several intervals are to be opened, the order in which
they are to be perforated is of importance.

Given the physical configuration of the well, the type of completion planned (e.g.,
"natural," gravel packed), and the well conditions, the gun type and charge
specifications should be chosen in consultation with the perforating service company
well in advance of the job. Generally speaking, one should choose the best-quality
charges and the largest-diameter gun possible. Be sure to be realistic in estimating
unknown well conditions, particularly pressure and temperature limits. The anticipated
pressures and the possibility of the presence of H2S will determine the need for and
specifications of the pressure control equipment.

When choosing a perforating company, consider the following points:

record of service in the area;

experience and training of operators;

411
equipment quality;

testing, certification, and maintenance procedures;

thorough and accurate record keeping;

accurate depth control procedures;

good charge quality control.

This last point is difficult to check since most guns arrive at the wellsite loaded into their carriers.

If visual inspection at the company's loading shop is possible, the following points,
suggested by King (1987), should be considered:

1. Are the charges fresh and stored in sealed boxes? Charges stored in unsealed boxes in
humid areas are suspect if over one year old.

2. Does a sample of the charges show any liner irregularities such as scratches,
cracks, or corrosion? Does the loading procedure prevent damage to the liner?

3. Are the correct charges being loaded? Premium powdered metal liner charges
do not exhibit a liner seam or open apex.

4. Are the primer cord and detonating cap of the proper type?

5. Are the charges loaded in a way that prevents misalignment? Scallop guns
have their charges loaded in strips, and misalignment of the charges usually
means that all the charges are off center. Beware of kinks in the detonating
cord or careless clamping of the charge rails in the carrier body.

6. If a port plug gun is used, check to see that the proper alignment sleeve is
used to hold the charge in place. Make sure that the carrier is not too badly
worn by expansion (such guns may be used 10 times or more). Bowed, split,
corroded, or badly swelled carriers should not be used.

412
TUBING CONVEYED PERFORATING

TUBING-CONVEYED PERFORATING
The technique of delivering the perforating charges on the bottom of a tubing string
and producing the well through the tubing immediately upon perforating was
considered when shaped charges first became known in the oil field. Because of safety
concerns and little real interest in improving perforating performance, the technique
was not applied. The Vann Tool Company introduced an updated version of the
technique in 1970, and its use grew over the next 10 years. The U.S. drilling boom of
the late 1970s and 1980s saw an increase in TCP completions, particularly for high
shot density, large-diameter perforating in Gulf of Mexico gravel-packed completions
and for long completion intervals in West Coast wells.

Basis for Tubing-Conveyed Perforating


The primary advantages of the tubing-conveyed technique include the following:

1. It provides an alternative to through-tubing perforating that safely combines large guns


and a high negative differential pressure. The tubing and packer are set with the wellhead
and tree installed before perforating, so there is no need to rely on a wireline blowout
preventer, as is sometimes the case with wireline through-tubing perforating.

2. It allows the easy perforation of long pay intervals. Gun length is not
constrained by weight (as is the case with conventional casing guns) or by riser
height (as in through-tubing wireline perforating). Intervals of over 1000 ft
(300 m) have been successfully perforated.

3. It allows the easy perforation of highly deviated wells since the guns are
pushed down the casing.

4. It eliminates the possibility of blowing the guns up the well, allowing larger
underbalance pressure differentials.

5. Mechanical or hydraulic detonators do not require electronic wiring and thus


eliminate the risk of premature detonation caused by stray current or radio
frequencies.

6. It permits the perforation of H2S-bearing formations with a greater degree of


safety.

7. It eliminates casing or tubing wear caused by repeated wireline trips in


multiple zone or long interval perforating.

However, there are some weaknesses in the TCP technique, specifically


1. the impossibility of confirming individual charge detonation without removing the entire
string (current research efforts are addressing this problem)

413
2. the need for a significant rathole to receive the expended guns after firing
and release or else the increased restriction of flowing the well permanently
through the vent assembly

3. the added expense when perforating a few small zones over a much larger
interval (e.g., limited entry fracturing); the ease of perforating with a
selectively fired wireline gun in such situations cannot be matched with TCP

4. the degradation of explosives with the additional time the charges must
spend at wellbore temperatures while the tubing string is run. On the whole,
tubing-conveyed perforation may take considerably less rig time than
conventional perforating, when time for multiple wireline runs, rigging up and
rigging down is considered. However, the charges are exposed to elevated well
temperatures for up to 100 times longer than wireline guns and there is a
chance of losing explosive power. For this reason, higher-temperature
explosives have been developed for TCP applications.

Tubing-Conveyed Perforating Equipment

The basic TCP assembly includes, from bottom to top ( Figure 1 ),

steel hollow carrier guns (of the expendable scallop type rather than the reusable port
plug variety)

414
Figure 1

an optional orienting sub for perforating the low side of the casing in deviated
gravel-packed applications

a firing head, operated by mechanical, hydraulic, or electronic means

an optional eccentering swivel sub, which allows smaller guns to be positioned


against the casing

production vents or valves, to permit flow into the tubing below the packer,
after firing

a gun drop sub, which permits the release of the gun assembly into the
rathole and subsequent wireline activity through the tubing opposite the
perforations

a retrievable production packer, if it is run as an integral part of the tubing


string

415
a radioactive tag sub, which permits a through-tubing gamma ray log to
precisely position the tool assembly

Spacer subs may be included between these components to provide for the correct positioning of
the guns opposite the interval to be perforated.

TCP Guns These carriers are essentially unchanged from their wireline-conveyed
counterparts. Table 1, below, shows the specifications for typical TCP gun systems
and their applications. The guns must be designed to permit vertical assembly in the
derrick. Connections that ensure 100% reliability of detonation transfer from adjacent
carriers are critical. Most TCP guns fire from the top down rather than from the bottom
up, as is the case with conventional guns. Also, the loading tubes used to align the
charges must be able to remain in place despite the repeated shock loading of tubing
string makeup. Most jobs are done with the 3 3/8-, 4-, and 5-in. guns, because most
production casing is in the 5- to 7-in. range.

Gun OD, Max Phasing, Shot Pattern Typical Application, all shot types
inches shots/ft. degrees

2 7/8 6 60 Spiral Stabbing through 7-in. packer or running


below 5-in. packer

3 3/8 6 60 Spiral Stabbing through 7-in. packer or running


below 5-in. packer

4 4 90 Spiral 5 1/2-in. or larger casing

4 6 60 Spiral 5 1/2-in. or larger casing

4 12 60 3 shots/plane 5 1/2-in. or larger casing

5 5 60 Spiral 7-in. or larger casing

5 12 120 Spiral 7-in. or larger casing

5 12 60 3 shots/plane 7-in. or larger casing

5 1/2 12 120 Spiral 7 5/8-in. or larger casing

6 12 60 3 shots/plane 7 5/8-in.or larger casing

6 12 120 Spiral 7 5/8-in. or larger casing

7 1/4 12 120 Spiral 9 5/8-in. or larger casing

Table 1: TCP Gun Systems.

Firing Systems The earliest firing system, still used today, is a drop bar which is simply
dropped down the tubing after the packer is set. The bar strikes a firing pin ( Figure 2 )
and detonates the charge.

416
Figure 2

Problems can develop when debris or crooked tubing slows down or stops the drop bar.
Accidental firing is prevented by the use of a pressure seal, a no-go restriction, and a
pressure-activated detonation transfer pin.

An alternative to the drop bar is the hydraulic firing head. The most common form
utilizes a pressure transfer sub to communicate annulus pressure above the packer to
a firing pin in the firing head ( Figure 3 ). When annulus pressure is applied at the
surface, pins are sheared and the firing piston detonates the gun.

417
Figure 3

Another version is the wet connect system. With this device a wireline electrical
connector is run inside the tubing via surface pressure control equipment. The
connector slips over a plug atop the detonator and electrical firing power comes from
the surface. This system allows "measurement while perforating" equipment to be
incorporated into the perforating program.

A second electrical detonating technique is the battery drop. A battery pack is dropped
in much the same manner as the mechanical bar drop, providing electrical power to
fire the guns.

Production Vents Situated below the packer, production vents or valves provide
communication between casing and tubing. Several types exist. A circulating sub with
disc ( Figure 4 ) allows the tubing volume to be circulated out along with any debris
that has settled out on the disc.

418
Figure 4

Below the disc the sub allows communication between tubing and casing. Once the
tubing is clear, the packer is set and the correct cushion level achieved by swabbing or
nitrogen displacement. The drop bar is then dropped through the circulating sub disc to
detonate the guns.

If the well has already been perforated it may be impossible to prevent the tubing from
filling with formation fluid, preventing underbalanced reperforating. In such a case, a
surge-disc sub may be placed above the circulating sub in place of the circulating disc.
The surge-disc isolates the tubing fluid cushion from the well fluid until it is shattered
by the drop bar.

Some companies run a simple, perforated nipple with a profile that allows a wireline
plug to be set above it. After the packer is set, the plug is pulled and the tubing is
ready for firing. This approach does not work when there is a risk that the plug will be
blown up the hole by formation pressure.

The mechanical-production valve is a valve that shifts the circulating ports open as the
drop bar passes by on its way to the firing head. A pressure-production valve is
opened by applying pressure to the tubing.

419
Gun Drop Sub This component allows the guns to be released and dropped to the
bottom of the rathole after firing, if necessary. Both mechanical and hydraulic versions
are available. Another version jettisons the guns as soon as firing occurs, before sand
influx might stick them in place.

Eccentering Swivel Sub This tool allows the guns to be laid flat against the casing when
the gun OD is small compared to the casing (such as when a gun must be stabbed
through a previously set packer). This component requires the use of pressure-
activated firing systems.

Orienting Sub To ensure that the low side of the hole is perforated in a high-angle
gravel pack completion, an orienting sub is used to signal when the rotated tubing
string is positioned with the guns pointing in the correct direction.

Shot Detectors As mentioned earlier, lack of confirmation of individual charge


detonation is a drawback of tubing-conveyed perforating. In fact, confirming that any
of the charges have fired may be impossible because tubing is not a good conductor of
mechanical shock. Some companies offer shot detectors which sense strain in the
tubing and small changes in tubing or annulus fluid columns, indications of detonation.
A new service offered by one company utilizes a wet connect electric wireline
conductor to monitor pressure, temperature, and flow rate just above the perforations.
This conductor gives a more reliable indication of firing.

Tubing-Conveyed Perforating Procedures


The basic sequence of events for a TCP job is as follows: First, the guns are assembled
vertically in the derrick with a firing head on top. Depending on the type of firing head
and the manner in which the cushion fluid is to be placed in the tubing, the vent or
production valve is made up above the firing head. A release sub may be added if the
guns are to be dropped into the rathole, and then the production packer is made up. If
a permanent packer or polished bore receptacle is already in place in the well, the
correct length of sealing element will be spaced out in the string. Spacer pipes will be
used to precisely place the guns relative to the productive zones in such cases.
Permanent packers are used perhaps 20% of the time, and while they allow precise
depth control, the packer ID normally restricts the gun size to about 3 3/8 in.

If a retrievable packer is run, the entire assembly is carefully lowered into the hole as
the remaining tubing string is made up. The guns are positioned approximately, using
the pipe tally, and then precisely located by running a gamma ray survey through the
tubing. An optional radioactive collar may be run above the packer to give a distinct
response on the gamma ray curve. Knowing the exact distance between the collar and
the top shot allows precise placement of the guns.

Before the packer is set, fluid may be circulated down the tubing, through the vent sub
and up the annulus. The packer is then set with the guns in position for firing. The
proper degree of underbalance may be established by swabbing or nitrogen
displacement.

The guns are fired, using any of the methods mentioned, and then either dropped into
the rathole or left hanging.

420
Multiple zone completions are also possible using various TCP tool configurations, as
shown in Figure 1 .

Figure 1

The TCP completion installation method depends on actual gun and tubing sizes, the
type of packers involved, whether the packers are independently set or run in one trip,
the casing dimensions, and the firing technique. In Figure 1 shown here, the
permanent packer is run and set on wireline. Both perforating guns, tubing strings,
and the retrievable packer are then run into the well on a single trip. The long string
(including the gun) is stabbed through the permanent packer and the dual packer set
and tested prior to firing the gun. Normally, the lower gun would be fired first, followed
by the upper gun. The Vann System described here includes a time-delay firing
mechanism (TDF) on the bottom of the gun assemblies, in case the primary firing
mechanism does not function correctly. Incorporation of multiple firing mechanisms
(primary and secondary) is a redundancy feature common to most commercially
available TCP systems.

Measurement While Perforating

421
Despite everyone's best efforts, completions seldom perform as expected.
Traditionally, after a well is perforated, it may be tested some time later and the test
results analyzed to determine the efficiency of the completion. Temperature,
flowmeter, or other logs may be run to determine how the perforations are performing.
Such evaluation, however, can be time-consuming and expensive.

Schlumberger's Measurement While Perforating (MWP)TM tool provides the means for
measuring bottomhole flowrate, temperature, and pressure before and/or immediately
upon perforating, without making any extra wireline trips. An analysis of the
completion efficiency can be done immediately and in certain cases problems may
even be remedied by reperforating prior to removing the original tool assembly.

The MWP tool is shown in Figure 1 ((a) Tubing-conveyed assembly without connect,
(b) Through-tubing assembly, (c) Tool Schematic).

Figure 1

It is run on a single conductor wireline and may be used to fire through-tubing wireline
guns or tubing-conveyed guns with a wet connect firing head. Up to 30 guns can be

422
included in the wireline mode and any gun can be selectively fired, from the bottom
up.

The tool assembly includes a perforating cartridge that houses the downhole firing and
switching electronics. Because a single conductor cable is used, the conductor
alternates between sending signals upward or downward. All downhole functions are
ultimately controlled by the computer system inside the logging van. The cartridge also
contains a casing collar locator for depth control. Above the cartridge are temperature
and pressure sensors, a tool centralizer, a gamma ray sonde for depth control, a flow
meter, and any weights necessary to push the through-tubing version into a high-
pressure well. Below the cartridge is a flex joint designed to isolate the delicate
downhole electronics from the shock of detonation. A magnetic positioning device may
be necessary for through-tubing perforating in larger-diameter casing. In the tubing-
conveyed system, a wet connect electronic coupling attaches to the TCP firing head.

The MWP tool string is 1 11/16 in. in diameter and houses strain gauge pressure
sensors having a rating of 10,000 psi or 20,000 psi with 0.15 or 0.30 resolution,
respectively. The most likely applications of this technology include:

1. immediate postperforation well testing

2. pre- and postperforation well testing for reperforated wells

3. improving the technique of limited entry fracturing by in-situ determination


of pressures and flow rates

4. individual perforating and testing of zones in a multi-layered reservoir

5. easier determination of vertical permeability by allowing the perforating and


testing of a portion of the formation, followed by full interval perforation and
testing

The tool provides a positive indication of gun detonation, and the temperature profile shows fluid
entry into the wellbore. Better depth control is possible since the gamma ray and CCL are run with
the guns. A precise measurement of cushion pressure is also possible.

Perforating Safety

423
Much of the responsibility for safe perforating operations falls to the service company
engineer, who follows a strict safety protocol in transporting, arming, firing, and
retrieving the potentially dangerous guns. However, the operating company
representative on location is generally responsible for safe activities on the company's
location, and together these individuals should ensure that a safety-conscious attitude
is exhibited by everyone involved. Handling of shaped charges, blasting caps, and
detonating cord can be safely accomplished if three basic rules are observed:

1. Follow the policy for safe perforating to the letter.

2. Communicate clearly with other individuals on location.

3. Concentrate on the job at hand.

The important safety considerations for perforating operations generally fall into four
categories:

1. Safe transportation of explosives

2. Detection and removal of stray electrical current, which could accidentally


detonate the blasting cap initiator

3. Safe arming of the gun prior to running in the hole

4. Safe running and retrieval procedures to eliminate the danger of premature


firing or undetonated charges

Transportation Transportation and storage of explosives must be carried out according


to all of the federal, state, and local regulations. The Department of Transportation has
strict requirements for routes, labeling, vehicle specification, and packaging and
storage of explosives. Waste explosives (unused detonating cord, blasting caps, and
unfired charges) must be stored in segregated containers and safely secured. The area
must be posted to prevent nearby use of radios, and a special smoking area should be
designated, away from the immediate location.

Stray Currents Stray currents set up by rig power malfunctions, cathodic corrosion
protectors, power malfunctions within the perforating unit, electrochemical activity
(salt mud systems), and radio frequencies (two-way radios, radar, microwave) have
the potential for detonating blasting caps. A voltmeter is used to check for voltage
between rig, casing, wellhead, catwalk, and perforating unit. If voltage exceeds 0.25 V,
the source must be eliminated. Grounding straps are used to connect the rig, wellhead
or casing, and perforating unit. Static electrical charges (electrical storms) can also
present a safety hazard. Guns should never be armed when an electrical storm is in
the vicinity.

Arming Procedures Arming the gun is a task reserved for the service company
engineer, with all unnecessary personnel cleared away. All power sources in the
perforating unit must be turned off and disconnected from the cable via a safety
switch. A special thick-walled steel loading tube is used to hold the blasting cap while
its lead wires are connected to the gun.

424
Each wire must be cut and connected individually, to avoid creating a loop antenna
that could pick up stray current. The cap is crimped to the detonating cord using a
special tool, and any unused scraps of detonating cord are stored in the appropriate
container. The bottom head connection is designed to avoid crushing of the blasting
cap and detonating cord.

Running Procedures The safety switch, which provides power to the gun via the cable,
is not turned on until the gun is at least 100 ft (30 m) into the well. Similarly, the
switch is turned off when the gun is retrieved, at the same depth. If the gun has any
unfired charges, the same safety precautions must be taken when disarming the gun
as were followed in arming the gun. The engineer will also check the gun for any
pressure buildup within the carrier, which must be bled off.

Also, precautions should be taken to control all sources of explosive initiation at all
times. Flame- and heat-producing devices (matches and cigarette lighters) and
sparking equipment (electrical power tools, steel hand tools, grinder, welding
equipment) should not be utilized near any explosive operations. Care should be taken
to avoid impact, friction, or static discharge when handling explosive devices. Always
follow standard operating procedures explicitly.

425
ACIDIZING & CHEMICAL TREATMENTS

INTRODUCTION

Introduction
Acidizing is one of the oldest techniques in well stimulation processes still being used
today, though its use only began toward the end of the last century. Table 1 (below)
lists the major events in the early history of acidizing.

Acidizing is a chemical process which dissolves an unwanted substance in the


formation, pipe, surface equipment, or flow-lines. In the case of treating formation
damage, acid removes flow-restricting particles, scale deposits, and minerals in the
reservoir or in the immediate wellbore vicinity. The acid mixture holds the dissolved
substance in solution until it is either removed from the well system or overflushed a
safe distance into the formation; then, the well can produce at its natural potential.
Little, if any, increase in productivity will result unless formation damage actually
exists.

When a fracture is created during an acidizing operation in carbonate reservoirs,


increased flow capacity can result from acid etching of the fracture faces. Here, large
increases in productivity are possible by creating a highly conductive flow path through
the formation.

Year Event

1894 Acid was used to stimulate oil production in limestone formations in Lima, Ohio.

1896 First patent related to acid treatments of wells was issued to Herman Frasch.

1928 Acid was used in Glen Pool, Oklahoma by Gypsy Oil Company, a subsidiary of
Gulf Oil Company.

1929 Subsidiary of Gulf Oil Corporation injected uninhibited hydrochloric acid under
pressure into a well in Lee County, Kentucky.

1930 Acidizing experiments were discontinued.

1932 Dow Chemical Company and Pure Oil Company used an acid treatment of
15% hydrochloric acid plus arsenic as an inhibitor on a well in the Greendale
Pool of Midland County, Michigan.

1932 Dow Chemical Company and Pure Oil Company used an acid treatment of
15% hydrochloric acid plus arsenic as an inhibitor on a well in the Greendale
Pool of Midland County, Michigan.

Patents were issued to Grebe and Sanford of Dow Chemical Company for
treatments of oil wells with chemicals containing a corrosion inhibitor.

426
R.H. Carr of the Pure Oil Company was issued a patent on the technique of
chemically treating an oil well by the balanced fluid column method.

Dowel I Chemical Company formed Dowel I Inc. (now Dowell-Schlumberger


from its Well Service Group whose original main responsibility was
stimulation of Dow's brine wells.

Carey K. West formed the Chemical Process Company, a predecessor to


Byron Jackson, Inc. (now BJ-Hughes).

1933 J.R. Wilson with Standard Oil Company filed a patent on a technique for
treating sandstone formations with hydrofluoric acid.

1935 Halliburton Oil Well Cementing Company (now Halliburton Services) began
acidizing oil wells.

1940 Dowell (now Dowell-Schlumberger) began the first commercial use of mixtures
of hydrochloric and hydrofluoric acid.

Table 1: Early History of Acidizing

427
ACID TREATING SOLUTION

Acid Treating Solutions


Whatever their specific formulation, all acids have hydrogen ions that react with
carbonate rock (principally limestone and dolomite) to form water and carbon dioxide,
and with siliceous minerals (sandstone and clays) to form fluosilicic acid and water.

Mineral Acids
Hydrochloric Acid

Hydrochloric acid (HCl) is the workhorse of the stimulation business, finding extensive
use in both carbonate and sandstone acidizing. Chemically, hydrochloric acid is
hydrogen chloride in aqueous solution. Composed of hydrogen and chlorine gas, it is
gaseous in the chemically combined state. Hydrogen chloride gas is readily soluble in
water. Concentration is described as percent by weight of HCl gas dissolved in water.
HCl acid is commercially available; its concentration has been standardized at 20 Be,
i.e., 31.45% acid, and it has a specific gravity of 1.160.

Many acidizing treatments employ HCl to some extent. Usually, it is used as a 15%
solution when treating carbonate formations. This concentration, commonly referred to
as regular acid, was originally chosen because of inadequacies in early inhibitors and
difficulty of preventing corrosion of well tubulars by more concentrated solutions.

Fifteen-percent HCl is used most frequently because it:

costs less per unit volume than stronger acids and is less costly to inhibit;
is less hazardous to handle than stronger acids;
retains larger quantities of dissolved salts in solution after spending.

Although acid concentrations higher than 15% may also be used in carbonate formations, lower
strengths are more suitable where acid dissolving power is not the sole consideration. In sandstone
acidizing, for example, 5 to 7.5% HCl is often used ahead of hydrochloric-hydrofluoric acid mixtures
to prevent precipitation of formation-plugging reaction products.

The principal disadvantage of HCl is its corrosive action on wellbore tubular goods.

HCl reacts with limestone (CaCO3) to form calcium chloride (CaCl2), water (H2O) and
carbon dioxide (CO2):

As shown, 2 moles of hydrochloric acid reacts with 1 mole of CaCO3 to produce 1 mole
each of calcium chloride, water and carbon dioxide. The molecular weights of these
compounds are as follows:

Compound Molecular Weight

428
Hcl 36.47

CaCO3 100.09

CaCl2 110.99

H2O 18.02

CO2 44.01

Thus, to dissolve one pound of limestone, we would need

[(1 lb CaCO3) (mole/100.09 lb)] [(2 mole HCl/ mole CaCO3) (36.47 lb/mole)]

= 0.729 lb HCl

On a volumetric basis, for the case of a 15 percent HCl solution (specific gravity =
1.076 at 60F; density = 8.96 lb/gal), one gallon of acid will dissolve (8.96
0.15)/(0.729) = 1.84 lb CaCO3.

The reaction of HCl on dolomite is very similar to its reaction on limestone, except for
the presence of magnesium. The chemical reaction of HCl and dolomite CaMg (CO3)2 is:

4HCl + CaMg(CO3)sub>2 => CaCl2 + MgCl2+ 2H2O + 2CO2

(2)

Hydrofluoric Acid

Hydrofluoric acid is the primary dissolving chemical used in sandstone acidizing. In


these applications, HF is usually mixed as a dilute solution with HCl, or an organic acid.
Its principal use is to dissolve siliceous minerals.

Within the chemical industry hydrofluoric acid (HF) is available commercially as a


relatively pure material in anhydrous form (where it is fuming and corrosive) or as a
concentrated (40 to 70%) aqueous solution. Anhydrous HF is not used in the field
because its low boiling point of 66.9 F is often exceeded by ambient temperatures. HF
reacts with silica and silicates, such as glass and concrete. It also attacks natural
rubber, leather, cast iron, and many organic materials.

HF is the only acid that reacts either siliceous minerals such as sand and clays. The
reaction of HF on quartz (SiO2), a primary component of sand, is:

4HF + SiO2 SiF4 + 2H2O

SiF4 + 2HF H2SiF6 (3)

429
Here, the end reaction product, fluosilicic acid (H2SiF6), is soluble in water, but its
potassium and sodium salts are insoluble. HF contact with formation water containing
CaCl2, MgCl2, NaCl or KCl thus must be avoided near the wellbore. The only salt
solution compatible with HF is ammonium chloride.

The reaction of HF on bentonite clay (Al2(Si4O10)(OH)2) is:

36HF + Al2(Si4O10)(OH)2 4H2SiF6 + 2H3AlF6 + 12 H2O (4)

Hydrofluoric acid also reacts with carbonates; its reaction with limestone is:

2HF + CaCO3 CaF2 + H2O +CO2 (5)

HF will react with calcium ions from any source to form the insoluble precipitate
calcium fluoride (CaF2), which can cause sever plugging problems in the formation. The
portion of the reaction that creates calcium fluoride is:

Ca++ + 2F- CaF2 (6)

CaF2 formation can be avoided through the proper use and optimal placement of HCl in
the acid treatment.

Hydrochloric-Hydrofluoric Acid Mixtures

Hydrochloric-hydrofluoric acid mixtures are the chief solutions used in matrix acidizing
of sandstone. These acids are very often used as mixtures because HCl ties up and
displaces anions, which in turn prevents reaction of these anions (e.g., Ca++, Mg ++,
etc.) with HF. Although HF is the only acid effective on siliceous minerals, it is
ineffective as a stimulation agent on carbonates because its reaction forms insoluble
CaF2.

The usual HF solution used in the field contains 3% HF and 12% HCl. This solution is
commonly called regular mud acid. However, concentrations of HF in HCl solutions
range from 0.5 to 9%, and some operators prefer 1-1/2% HF and 6% HCl.

These mixtures may be formed from the dilution of concentrated solutions of HF or,
more frequently, from the reaction of ammonium bifluoride (NH4HF2) or ammonium
fluoride (NH4F) with HCl. Usually, 15% HCl is used, and enough NH4HF2 or NH4F is
added to create a solution containing 3% HF. Consumption of hydrogen chloride by
these reactions leaves 12% HCl remaining in solution. Similarly, 1.5% HF can be
prepared from 7-1/2% HCl solutions where the final HCl concentration is 6%.

Corrosion characteristics of HCl-HF mixtures are comparable to those of HCl alone, and
similar corrosion inhibitors are required. Both HCl and HCl-HF acid mixtures are
stronger and more reactive than organic acids.

Organic Acids

430
Organic acids offer the benefit of relatively low corrosivity and easy inhibition at
elevated temperature. They are particularly applicable, therefore, to high-temperature
well applications.

Acetic (HC2H3O2) and formic (HCHO2) acids are the two principal organic acids used in
acidizing. These acids are weaker than mineral acids; they react more slowly and are
also less damaging. Acetic, for example, is the only acid available that will not damage
chrome plating or aluminum and magnesium surfaces. In some cases, organic acids
are mixed with HCl or HF to provide deeper acid penetration into the formation.

Acetic Acid

Acetic acid was the first organic acid to be used in appreciable volumes in acidizing. It
is easy to inhibit against corrosion and commonly used as a 10% by weight solution of
acetic acid in water. At this concentration, the reaction products are generally soluble
in spent acid. Acetic acid functions well as a perforating fluid or low-corrosivity acid
when in contact with metals that corrode easily.

Commercially available "pure" acetic acid is approximately 99% acetic acid. It is called
glacial acetic acid because icelike crystals form at temperatures around 60 F and it
solidifies near 48 F. When glacial acetic acid is mixed with water, a contraction occurs.
For this reason, the amounts of acetic acid and water normally total more than the
required volume when making dilute solutions.

Acetic acid reacts with limestone to form calcium acetate (Ca(C2H3O2)2), water and
carbon dioxide:

2HC2H3O2 + CaCO3 > Ca(C2H3O2)2 + H2O + CO2 (8)

As shown, 2 moles of acetic acid reacts either 1 mole of CaCO3 to produce 1 mole each
of calcium acetate, water and carbon dioxide. The molecular weights of these
compounds are as follows:

Compound Molecular
Weight

HC2H3O2 60

CaCO3 100.09

CaCl2 110.99

H2O 18.02

CO2 44.01

Thus, to dissolve one pound of limestone would require:

[(1 lb. CaCO3) x (mole/100.09 lb.)] x

431
[(2 mole HC2H3O2/mole CaCO3) x (60 lb./mole)]

= 1.2 lb.HC2H3O2

On a volumetric basis, for the case of a 15 percent HC2H3O2 solution (specific gravity =
1.037 at 60oF; density 8.64 lb./gal.), one gallon of acid dissolves (8.64 x 0.15)/(1.2) =
1.08 lb. of limestone. An identical concentration of HCl dissolves 1.84 lb. of limestone.
(see Table 1).

Formic Acid

Formic acid is the simplest of the organic acids. It is stronger than acetic, yet weaker
than HCl, and is most frequently used in combination with HCl as a retarded acid
system for high-temperature wells. Formic acid can be easily inhibited, but not as
effectively as acetic acid at high temperatures and long contact times. The percentage
of formic acid generally used is 8% to 10% by weight.

A 10% solution of formic acid will dissolve as much limestone as an 8% solution of HCl.
The chemical reaction of formic acid and limestone is:

2HCHO2 + CaCO3 > Ca(CHO2)2 + H2O + CO2 (9)

Acid Concentration CaCO3 CO2 CaCl2


type % dissolved formed formed
per gal. per gal. per gal.
acid- lb. acid cu. acid- lb.
Ft.

HCl* 15* 1.84 6.99 2.04

20 2.50 9.47 2.75

25 3.22 12.20 3.57

Acetic* 15* 1.08 4.09 1.71

20 1.43 5.41 2.25

25 1.80 6.82 2.84

Formic* 15* 1,42 5.38 1.84

20 1.90 7.20 2.47

25 2.40 9.09 3.12

* The most commonly used strength for HCl is 15% and 10% for acetic and formic
acids. In this table, however, the HCl reaction rate on limestone is compared to the

432
equivalent strength rates for acetic and formic acid rather than to the latter's
commonly used strengths.

Table 1: Reaction of three acids on limestone at various concentrations.

Formic acid is frequently used in reservoirs with extreme temperatures or low injection
rates. At high temperatures, organic inhibitors work more effectively on formic acid
with HCl than on HCl alone. This property minimizes the danger of hydrogen
embrittlement of steel associated with HCl treatment in high-temperature formations.

Organic acids have less dissolving capability than HCl and must be used in greater
volumes for comparable results, making such treatments more expensive.

The dissolving power of various concentrations of HCl, acetic, and formic acids is
shown in Table 1. The cost of acetic acid per unit based on dissolving power is more
expensive than either HCl or formic acid. With the exception of fracturing acidizing
treatments, dissolving power is not often the primary consideration in acid selection.

The most commonly used strength for HCl is 15% and 10% for acetic and formic acids.
In Table 1, however, the HCl reaction rate on limestone is compared to the equivalent
strength rates for acetic and formic acid rather than to the latter's commonly used
strengths.

Hydrofluoric-Organic Acid Mixtures

Acid mixtures of HF and either acetic or formic acid are used to slow the reaction of the
acid on sand and clay, and to reduce corrosiveness. These acid mixtures can be
effectively inhibited for up to 16 hours at 400 F. Mixing HF with organic acids can
provide deeper penetration and, therefore, effective removal of deep formation
damage. Organic HF mixtures are recommended at temperatures of 200 F and above;
below 200 F they can cause the formation of undesirable reaction products.

Specialty Acids
Specialty acids are designed to deal with specific formation conditions, such as deep
clay damage, paraffin blockage in the reservoir, and situations requiring a retarded
acid.

Powdered Acid

Powdered acids (sulfamic and chloroacetic) are used in applications where logistics are
of primary concern - for example, in remote locations where bulk transport of liquids is
impractical or very expensive. These crystaline powders are readily soluble in water.

Sulfamic acid (HSO3NH2) is a nonvolatile, nonhydroscopic, white crystalline or granular


solid. Its reaction products with carbonates are water soluble, and it is highly ionized in
water. Sulfamic acid is less corrosive than HCl, yet its strength is about the same. At
77 F a saturated solution of sulfamic contains about 18% hydrochloric acid equivalent.

433
Sulfamic acid reacts with limestone to form calcium sulfamate (Ca (SO3NH2)2), water
and carbon dioxide

2HSO3NH2 + CaCO3 > Ca(SO3NH2)2 + H2O + CO2 (10)

One advantage of sulfamic acid is its solid physical state. Solid sulfamic acid can be
transported and stored without special equipment or tanks, and mixed with water at or
near the wellsite as needed. Sometimes it is cast into "acid sticks" for easy
introduction into the wellbore.

The disadvantages of sulfamic acid include its decomposition at around 180 F, making
it inappropriate for applications where formation temperatures are above 160 F.
Moreover, HCl offers more dissolving power and comparable reactions at lower cost.

Chloroacetic acid is generally preferred to sulfamic acid when use of a powdered acid is
appropriate, because it is stronger and more stable.

Retarded Acids

Retarded HF acid can penetrate deeper into a formation than conventional HF to


remove siliceous solids. Retardation of HF achieves deeper penetration of unspent acid,
can aid more complete formation damage removal, and further increase production.

A number of retarded HF acid systems are commercially available several of which are:

SGMA (self-generating mud acid): The first retarded sandstone-acidizing system to be


used extensively was developed by Shell Oil Company. It involves pumping ammonium
fluoride and an organic ester, methyl formate, into the formation. (Methyl formate has a very
low flash point and should be pumped with caution.)In time, ester hydrolysis produces
formic acid. This acid reacts with ammonium fluoride to form HF, which then dissolves clays
or any siliceous minerals it contacts.
Claysol retarded HF (Halliburton): A retarded clay-dissolving system which utilizes ion-
exchange properties of clay minerals to generate HF on clay in situ. Since HF is formed on
clay surfaces, little sand is dissolved by this process. HF is created by sequentially injecting
a volume of 3% ammonium fluoride followed by an equal volume of 5% HCl. This process
dissolves clay in the formation as deep as a set of sequences can be pumped without
completely mixing together.
Clay acid (Dowell-Schlumberger): A retarded acidizing fluid using fluoboric acid (HBF4) for
matrix acidizing of sandstone formations. HBF4 is usually applied as an afterflush to an HF
treatment. Upon entering the formation, HBF4 slowly hydrolizes to generate HF. Deep acid
penetration to remove fines is possible because of the slow generation of HF. Unlike clay-
control agents, HBF4 also produces a chemical fusion of both fines and clay platelets,
which provides fines/clay stabilization.
RHF acid solution (Nowsco): A retarded HF system for treating sandstone formations
suffering deep damage caused by migration and/or swelling of siliceous minerals. This
single-stage treatment does not require sequencing or "shut-in" time for hydrolysis
reactions. Addition of aluminum chloride (AlCl3) to an HF acid solution forms aluminum
fluoride complexes, similar to those formed in spent regular mud acid, which retards the
reaction rate of HF with siliceous materials.
Retarded Hydrochloric Acid

434
The HCl reaction rate on carbonate formations can be retarded by gelled acid, oil-
wetting formation solids, or emulsifying acid with hydrocarbon.

Gelled Acid Systems: Gelled acids are used to retard acid reaction rate during fracture
acidizing. Retardation occurs because increased fluid viscosity reduces the rate of acid
contact with the fracture face.

Other related advantages include:

reduced leak-off rate

deeper penetration

better cleanup of fines or solids transport

Viscosifying agents normally associated with gel led acid systems consist of natural polymers,
synthetic polymers, cross-linking agents, and surfactant gelling chemicals.

Chemically Retarded Acids: When it is desirable to extend the spending time of an acid
system, a chemical retarding agent may be used. Most chemical retarders are anionic
surfactants (such as sulfonates or sulfates). These oil-wetting surfactants adsorb onto
a carbonate to create a physical barrier to acid contact with the rock surface. The
adsorption mechanism on a calcium crystal may look like the illustration in Figure 1
(absorption mechanism ).

435
Figure 1

Emulsified Acids: Emulsified acids, obtained by emulsifying acid and a hydrocarbon,


are effective over a wide range of bottomhole temperatures. The system may be either
oil or acid external, but the most common form is an oil-external emulsion.

Oil-external emulsion is composed of hydrocarbon-base fluid (refined or crude) as the


continuous phase and HCl as the internal phase. This acid system retards the reaction of
acid with the formation by decreasing the amount of acid in contact with the rock. As the
acid spends, the emulsifier reacts with the resulting CaCl2 solution, releasing its emulsifying
properties and therefore causing the emulsion to break.
Acid-external emulsion is composed of acid as the external phase, where selection of acid
depends on the well conditions involved. The acid phase may account for 80 to 90% of its
total volume. Ordinarily, aromatics, such as toluene or xylene, are used as the hydrocarbon
or internal phase. This acid system is used primarily to remove hydrocarbon materials like
paraffin, congealed oil, and other deposits so acid can contact acid-soluble materials.
Foamed Acid

Foamed acid has widespread application for effective fracture acidizing in carbonate
reservoirs. Both oil and gas wells have responded successfully to foamed acid
stimulation treatments. Foamed acid is particularly beneficial in low-pressure, low-
permeability, liquid-sensitive formations because of the following advantages:

High Viscosity: While often difficult to achieve with conventional gelling agents in acid,
high apparent viscosities are easily developed by foaming the acid. Higher viscosities may
result in wider fractures, better fluid-loss control, and better solids-carrying capabilities.
Reduced Fluid Loss: The high apparent viscosity of foamed acid results in reduced fluid
loss, which allows deeper acid penetration. In low-permeability reservoirs, foam bubbles
may be sufficient to decrease leakoff into the formation matrix.
Improved Solids Transport: In conventional acid treatments, large amounts of insoluble
fines can be left in the formation because of low-viscosity spent acid. These solids can
reduce fracture conductivity. Foamed acid and high viscosity suspend and remove more
fines from the well during cleanup.
Better Cleanup: Foamed acid offers built-in gas assist. This means faster cleanup. It may
even eliminate the need to swab the well after acidizing.
Less Damage: Foamed acid has a low liquid content. Normally, the volumetric gas
content (referred to as foam quality) is 60 to 80 quality (i.e., 60% to 80% gas and 40% to
20% acid).
Experience with foamed acid has shown that increasing acid viscosity with conventional gelling
agents before foaming can help increase foam stability. The longer foamed acid, live or spent, is
allowed to remain static, the easier it is for acid to drain from the foam bubbles. This foam
destabilization (liquid/gas separation) allows suspended fines to settle out.

Nitrogen is the most common gas used to make foamed acid. Foam quality is generally
between 60% and 80%, although qualities as high as 95% have been used. The acid
phase of the foam may contain 0.5% to 1.0% surfactant and 0.2% to 2.0% corrosion
inhibitor.

436
Turflo/Hydrofluoric Acid

As stated earlier, HCl-HF mixtures react with both siliceous and carbonate materials. In
most sandstone formations carbonate content is small, and an HCl preflush removes all
carbonate material. HCl may not, however, do the job effectively in formations with
high limestone or dolomite content. These carbonates can reduce the effectiveness of a
HCl-HF treatment by increasing the chance of secondary reactions and by limiting the
amount of clays removed.

Turflo/hydrofluoric acid, a phosphoric acid-HF combination, is specifically designed for


formations with high carbonate content. Tests with this system have shown its reaction
rate is much slower than that of normal 12% HCl-3% HF with limestone and nearly as
fast with silicas.

Since Turflo/HF is slightly slower reacting than HF acid, deeper penetration of HF is


possible. Turflo/HF is more expensive than conventional HF, but its limited reaction
rate with carbonates allows for improved formation damage removal in sandstone
formations with high carbonate content.. However, caution should be exercised
because precipitation problems may subsequently cause severe permeability damage.

437
ADDITIVES IN ACIDIZING FLUIDS

Corrosion Inhibitors
The principal additives to acidizing fluids are corrosion inhibitors. A corrosion inhibitor
is a chemical that slows acid corrosion on drill pipe, tubing, or any other metal that the
acid contacts during treatment. Brief explanations follow of the acid corrosion
mechanism, types of acid corrosion, and corrosion by different acid types.

Acid Corrosion on Steel: All uninhibited acid solutions corrode steel. The attack of acid
on steel occurs through the dissociated hydrogen ion of the acid solution. This results
in the oxidation and dissolution of iron at the anodic sites on the metal surface along
with the reduction of hydrogen ions and formation of hydrogen at the cathodic sites:

Anodic Reaction (Oxidation)

Metallic iron Ionic iron Electrons

Fe => Fe++ + 2e-


(14)

Cathodic Reaction (Reduction)

Ionic Electrons Hydrogen


hydrogen

2H+ + 2e- > H2


(15)

These two reactions can be summarized by the following overall reaction:

Metallic iron Ionic hydrogen Ionic Hydrogen


iron

Fe + 2H+ Fe++ H2
(16)

This electrochemical corrosion process is shown in Figure 1 (electrochemical corrosion


).

438
Figure 1

The units of measurement used to determine the amount of corrosion are lb./sq. ft. To
illustrate, one sq. ft of steel 0.001 in. thick weighs 0.041 lb.; 1.92 ft of 2-3/8 in. tubing
has an inside surface area of one sq. ft. If corrosion weight loss of this 2-3/8 in. tubing
is 0.041 lb./sq. ft, then every foot of the string has lost 0.021 lb. and is thinner by
0.001 in.

Acceptable corrosion limits are demonstrated in Table 1 (below). The range is from
0.02 to 0.09 lb./sq. ft of metal surface, depending on temperature. Keep in mind,
however, that acids corrode more evenly, with less pitting, at higher temperatures. At
elevated temperatures, then, more corrosion may be permitted with less chance of
forming a hole in the drill pipe or tubing string. A corrosion limit of 0.10 lb./sq. ft may
be acceptable with a low pitting index representing 1% metal loss. Furthermore, there
is less risk in an initial completion acidizing treatment than on tubing subjected to
repeated acidizing.

Temp. -OF Corrosion


Limit lb./sq. ft.

439
200 0.02

200-275 0.05

275-350 0.09

Table 1: Acceptable corrosion limits

Since acid corrosion is strongly influenced by the chemical composition of the steel,
metal chemistry is an important factor in corrosion studies on oilfield tubular goods.
API specifications, however, are based on physical properties of the metal (i.e., tensile
strength) rather than chemical composition. Marked differences in chemical
composition may exist in the same grade of tubing or casing from different
manufacturers, or from lot to lot from the same manufacturer.

A general rule to follow when comparing the degree of corrosivity of two types of
tubular goods is that the higher the hardness of the metal, the more prone it is to acid
attack. This rule is illustrated in Figure 2 (Effect of HCl on different types of steel ).

Figure 2

Remember, however, that there are exceptions to every rule.

440
Types of Acid Corrosion: In uninhibited acid solutions, the corrosion of steel is usually
uniform. The constant shifting of anodic and cathodic areas spreads corrosion more or
less evenly over the entire surface of the metal. Pitting of the steel surfaces occurs in
some situations due to inhibitor breakdown, insufficient inhibitor, and metal impurities.

Inhibitor Breakdown: On steel exposed to acid solutions containing an inferior


corrosion inhibitor, a pitting type of corrosion is not uncommon. All corrosion inhibitors
eventually break down after some period of time, depending on, among other factors,
temperature, acid strength, and metal type. When this point is reached, an inferior
inhibitor may actually tend to promote pitting by desorbing from the metal surface in
localized areas.

Insufficient Inhibitor: Regardless of the quality of corrosion inhibitor, pitting may also
occur if there is an insufficient amount of inhibitor to effectively coat steel surfaces
under certain conditions. Unprotected steel surface areas are rapidly attacked by the
acid solution, and pitting occurs.

Metal Impurities: Another condition that promotes pitting is the presence of impurities
or inclusions within the steel itself. For example, small pieces of slag may become
trapped during the forming of the steel, or improper heat treating or quenching of the
steel may produce discontinuities in its grain structure. These imperfections may, in
turn, become anodic to the rest of the surrounding steel structure, and thus promote
acid attack.

Hydrogen Embrittlement: This is one of the most insidious forms of corrosive attack.
Because oxidation and dissolution of iron occur at the anodic sites on the metal surface
when steel is corroded by acid, the reduction of hydrogen ions and formation of
hydrogen takes place at the cathodic sites. The cathodic reaction can be expressed as

Hydrogen Electron Atomic


ion hydrogen

H+ + e- => H (17)

Atomic Atomic Molecular


hydrogen hydrogen hydrogen

H + H => H2 (18)

Because atomic hydrogen is very small, it can penetrate the metal and become lodged
in void spaces along metal grain boundaries. Molecular hydrogen is too large to
penetrate and lodge in these spaces, and is not a problem at well temperatures.

When atomic hydrogen that has penetrated the metal structure comes in contact with
another atomic hydrogen atom, the two combine to form molecular hydrogen within
the steel structure. When the larger molecule very slowly migrates out, extremely high

441
pressures develop in the void spaces that literally split the metal open. This process is
called hydrogen embrittlement.

If hydrogen sulfide (H2S) is involved, the splitting phenomenon is known as hydrogen


sulfide cracking. H2S tends to slow down the transformation of atomic hydrogen into
molecular hydrogen on the metal surface at the cathode. It thus allows greater
penetration into the metal which, in turn, increases the probability of high-strength
steel cracking. Figure 3 , is a photomicrograph of a crack in stressed Pl05 pipe exposed
to 15% HCl saturated with H2S.

Figure 3

This cracking can occur when very little physical corrosion is taking place. Most
commercially available corrosion inhibitors are ineffective in controlling cracking of
high-strength steels in acid saturated with H2S. As a general rule, steels with a
Rockwell "C" hardness of 22 or less are not as likely to experience H2S cracking as
steels with greater hardness.

442
Corrosion by Different Acid Types The most common types of acid solutions used in the
oil field are HCl, formic, and acetic.

The degree of ionization of hydrogen from the acid molecule determines the acid
strength, which is directly proportional to its corrosivity on steel. The degree of
ionization for these acids is HCl > formic > acetic. HCl, then, is more corrosive on steel
than formic, which is more corrosive than acetic. Quite logically, the more aggressive
an acid is in its attack on steel, the more difficult it is to inhibit. The mechanism of
attack, however, is the same for all three types, as shown below.

HCl Iron Iron Hydrogen


(inorganic) chloride
acid

2HCl + Fe => FeCl2 + H2


(19)

Formic (organic) Iron Iron Hydrogen


acid formate

2HCHO2 + Fe => Fe(CHO2)2 + H2


(20)

Acetic (organic) Iron Iron formate Hydrogen


acid

2HC2H3O2 + Fe => Fe(C2H3O2)2 + H2


(21)

Inhibitor Types

There are two basic types of corrosion inhibitors: inorganic and organic.

Inorganic Corrosion Inhibitors: This class of inhibitor includes salts of zinc, nickel,
copper, arsenic, antimony, and various other metals. Of these, the most widely used
are arsenic compounds.

When these arsenic compounds are added to an acid solution, they "plate out," or form
iron sulfide at cathodic sites of exposed steel surfaces. These plates decrease the rate

443
of hydrogen ion exchange, because iron sulfide acts as a barrier between metal and
acid. It is a continuing process in which acid reacts with iron sulfide, rather than metal.

Some advantages of inorganic inhibitors are that they:

work very effectively at high temperatures for long contact times;

cost less than organic inhibitors.

Disadvantages of inorganic inhibitors are that they:


tend to lose their effectiveness in 17 to 20% HCl and stronger acid solutions;

react with H2S (via FeS ions) to form an insoluble precipitate called arsenic
sulfide. (In the reservoir this precipitate exists in a gaseous state, but when it
enters the wellbore and cools, a plugging solid forms.);

poison refinery catalysts, such as platinum;

liberate occasionally toxic arsine gas as a by-product of corrosion;

resist mixing and safe handling.

Organic Corrosion Inhibitors: These inhibitors are composed of polar organic compounds capable
of adsorbing onto the metal surface, thereby establishing a protective film that acts as a barrier
between the metal and the acid solution. They usually serve as a cathodic polarizer by limiting the
hydrogen ion mobility at cathodic sites. Organic inhibitors are composed of rather complex
hydrocarbon units with one or more polar groups made of sulfur, oxygen, or nitrogen.

Some advantages of organic inhibitors are that they:

can be used in the presence of H2S without precipitation of arsenic sulfide, which can plug
the wellbore.

do not poison refinery catalysts.

work effectively in all acid concentrations.

Some disadvantages of organic inhibitors are that they:


tend toward in-situ degradation that is, they chemically degrade with time in the
presence of an acid solution, and thus do not readily provide long-term protection at
temperatures above 220 F.

cost more than the inorganic corrosion inhibitors.

Table 2 (below) demonstrates effectiveness of organic and inorganic corrosion inhibitors at high
temperatures in 15% HCl. The protective time reflects time required for the HCl to remove 0.05
lb./sq. ft of exposed metal area.

Do not make the mistake of assuming that twice as much inhibitor provides twice as
much protection: the concentration limit varies from one inhibitor to another. The

444
general rule is that a 2% concentration is maximum. Beyond this point, no practical
benefit is gained.

Remember that the purpose of a corrosion inhibitor is to slow the acid reaction rate on
steel. Inhibitors in and of themselves do not prevent corrosion.

Listed in Table 3 are some examples of corrosion protection times available at the
more common temperatures found in acidizing treatments.

Table 3

Intensifiers

Although these additives cannot be considered inhibitors when used alone, their
effectiveness is increased when they are used in conjunction with an organic inhibitor.
The commonest intensifiers are potassium iodide, cuprous iodide, and formamide. The
addition of these synergistic chemicals to existing organic inhibitor formulations greatly
extends the range of their effectiveness, particularly in higher-temperature
applications.

Type Inhibitor Temperature Protection Time- hr.*


Inhibitor Concentration- F
%

Organic 0.6 200 24

1.0 250 10

445
2.0 300 2

Inorganic 0.4 200 24

1.2 250 24

2.0 300 12

*Time required for 15% HCl to remove 0.05 lb./sq. ft exposed metal area.

Table 2: Effectiveness of corrosion inhibitors at high temperatures in 15% HCl.

Surfactants
Surfactants, or surface-active agents, lower surface tension by adsorbing at the
interface between liquid and gas. Surfactants can also lower interfacial tension by
adsorbing at interfaces between two immiscible liquids and changing contact angles by
adsorbing at an interface between a liquid and solid. They are commonly used in
acidizing (and other completion and workover operations) to demulsify spent acid and
oil, reduce interfacial tension, speed cleanup and prevent sludge formation.

Surfactants should not be added to treatment fluids without first having a full
understanding of their effects on the reservoir oil at downhole conditions. Without this
understanding, it is better not to use surfactants at all in the formulation, particularly
when the objective is to break or prevent emulsion (Schecter, 1992).

Surface tension is a property of liquids that distinguishes them from gases. Liquid
molecules exert a mutual attractive force on each other. When this force a
combination of van der Waals' forces and electrostatic forces is not balanced, a
tension perpendicular to the surface develops. The greater the attractive force of these
molecules for one another, the greater the surface tension of the liquid, and the
greater the amount of work per unit area required to rupture the liquid surface. This
amount of work (surface tension) is specific for each liquid at a definite temperature
and is expressed in dynes/cm.

Interfacial tension is the force in the interface, or boundary layer, between two unlike
liquids. This tension results from unbalanced intermolecular forces in the boundary
region, is a factor in the formation or breaking of emulsions in the recovery of treating
fluids from matrix permeability.

In general, surfactants can be divided into four main groups: cationic, anionic,
nonionic, and amphoteric.

Cationic surfactants are organic chemicals with molecules that ionize upon contacting
water. The water-soluble group is positively charged. The positively charged particle,
consisting of oil-soluble and water-soluble ends, contains the surface-active portion of
the molecule. The active part orients itself at the liquid/air, acid/oil, or liquid/liquid
interface. Most cationic surfactants are amine compounds, such as quaternary
ammonium chloride:

446
where R represents an oil-soluble group.

Anionic surfactants are organic chemicals with molecules that also ionize upon
contacting water. But here the negatively charged particle of each ion, consisting of a
water-soluble end and an oil-soluble end, is surface active. The anionic surfactant is
negatively charged; it is this part that is active in orienting itself at the surface of the
liquid, or at the interface between the water and oil. The positively charged particle
goes off into solution and has little to do with surface tension.

Some examples of anionic surfactants are:

sulfates R-OSO3-

sulfonates R-SO3-

phosphates R-OPO3-2

phosphonates R-PO3-2

where R represents an oil-soluble group.

Of these, the most common anionic surfactants are sulfates and sulfonates. Anionic
surfactants are primarily used as nonemulsifying agents, retarding agents, and
cleaning agents.

Surfactants are organic molecules that do not ionize and therefore remain uncharged.
Most nonionic surfactants contain water groups that are polymers of either ethylene
oxide or propylene oxide.

Some examples of nonionic surfactants are:

polyethylene oxide RO(CH2CH2O)xH

polypropylene oxide

Nonionic surfactants are primarily used as foaming agents and nonemulsifiers.

447
Amphoteric surfactants are organic molecules with water-soluble groups which can be
positively charged, negatively charged, or uncharged. The specific charge depends on
system pH level. If the pH is acidic, the amphoteric surfactant takes on cationic
properties, or a positive charge. If the pH is in the neutral range, then it takes on
nonionic properties, or remains uncharged. If the pH level is basic, the amphoteric
takes on anionic negative charge properties.

Two examples of amphoteric surfactants are:

amine sulfonate RNH(CH2)ySO3H

amine phosphate RNH(CH2)zOPO3H

Some amphoteric surfactants are used as corrosion inhibitors, foamers, and silt-
suspending agents.

Properties Affected by Surfactants The main fluid properties affected by surfactants are
surface and interfacial tensions, emulsification tendency, wettability, and dispersibility.

Surface Tension : Figure 1 illustrates how surfactants lower surface tension.

448
Figure 1

Since the water-soluble group exceeds the water solubility of the oil-soluble group,
surfactant molecules orient themselves at the air/water interface (surface). The water-
soluble group is in water, and the oil-soluble group in the air. This changes the nature of the
air/water interface, making it a combination of air, water, and oil. Because oil has a much
lower surface tension than water, surface tension of the water surfactant mixture is lower
than the surface tension of pure water.

Surface tensions of some common liquids are listed in Table 1 (below).

Liquid Surface Tension - dynes/cm

Water 72.0

15% HCL 72.0

Spent 15% HCl 76.9

Octane 21.8
Table 1: Surface tension of common liquids.
Some surfactants can reduce surface tension of water to about 27 dynes/cm when used in
relatively low concentrations. One class of surfactants, called fluoro-surfactants, can lower
surface tension below 15 dynes/cm.

Interfacial Tension: Interfacial tension describes the surface energy between


two immiscible fluids, such as water and oil. This interface acts like an invisible
membrane that prevents the liquids from mixing. The force required to break
this molecular film is called interfacial tension.

The interfacial tension between kerosene and water is about 49 dynes/cm. By


the addition of a surfactant, it can be lowered to near zero. Figure 2
(surfactants lower interfacial tension) shows the surfactant being adsorbed at
the oil-water interface.

449
Figure 2

Again, the water-soluble group adheres to the water, and the oil-soluble group
is in the oil phase. The interfacial surface between the two fluids is now a
mixture of oil and water, with the interfacial tension being greatly reduced.
Even though the surfactant is water soluble, some of it migrates through the
interface and enters the oil phase.

When interfacial tension is lowered in an acidizing treatment, the water may


break up into smaller droplets that pass through the oil and flow out of the well
without forming unwanted emulsions. In the case of oil, interfacial tension can
be reduced to enable the oil to pass through rock pore spaces with greater
ease.

Emulsification Tendency: An emulsion is a mixture of two fluids in which fine


droplets of one fluid remain suspended in the other. Emulsions may be oil-
external or water-external. In the first instance, oil is the external phase with
the water droplets dispersed throughout. This is the most common emulsion
found in wellbores connected to oil reservoirs. Water-external emulsion has an
aqueous external phase with oil droplets distributed throughout.

450
Many crudes contain naturally occurring surfactants that reduce surface tension
between oil and formation water and thus promote the development of
emulsions. Treatment of the formation with certain surfactants can also lead to
the development of emulsions. However, certain surfactants can be used to
treat wells with emulsions problems.

Wettability: Oil and water are immiscible liquids that compete for a place on
the formation surface. The questions of which liquid preferentially wets the rock
surface is an important factor in acidizing. When the formation surface is
completely covered by a film of oil, the formation is termed oil-wet. Conversely,
when covered by water, the formation is water-wet.

The electrochemical approach usually explains the ability of a surfactant to


adsorb at interfaces between liquids and solids and alter the wettability of
solids. Most formations are mixtures of sand, clay, limestone, and dolomite. We
next consider each of these components.

Sand usually has a negative surface charge. When a cationic surfactant is used,
the positive water-soluble group is adsorbed by the negative silica particle. This
leaves the oil-soluble group to influence wettability. As a result, cationic
surfactants generally oil-wet sand. When anionic surfactants are used, the sand
and silicate minerals are generally left in their natural water-wet state.
Wettability characteristics exhibited by anionic and cat ionic surfactants on a
silicate particle are shown in Figure 3 and Figure 4 .

451
Figure 4

452
Figure 3

Limestone and dolomite usually have a positive surface charge. Since anionic
surfactants have a negative charge, the oil-soluble group is usually left to
influence wettability. Hence, anionic surfactants tend to oil-wet lime-stone and
dolomite formations.

Because carbonates do not usually adsorb cationic surfactants, most cationic


surfactants leave limestone and dolomite in a natural water-wet condition.
Wettability characteristics exhibited by anionic and cationic surfactants on a
carbonate particle are illustrated in Figure 5 and Figure 6 .

453
Figure 6

454
Figure 5

In the case of nonionic surfactants, the wettability of silicates and carbonates


depends primarily on the solubility of the material. If the nonionic is oil-soluble,
it will tend to oil-wet the silicates. This is because silicates are attracted by
hydrogen bonding.

Contact angles are used to study wetting of surfactants. Figure 7 and Figure 8
shows how a wetting and nonwetting drop looks on a solid surface.

455
Figure 7

By using the proper surfactant, formation wettability can be changed from oil-
wet to water-wet, and vice versa. Since oil flows more easily through water-wet
formations, selection of the proper surfactant can be an important factor in
increasing oil production.

456
Figure 8

Figure 9 shows the results of oil flow through oil-wet and water-wet sands.

457
Figure 9

As discussed earlier, certain surfactants are used to alter wetting properties of a


producing formation. These curves show that oil exhibits better flow rates
through water-wet formations. Care should be taken to have the formation in a
water-wet condition following an acidizing treatment, which is often not a
simple task. In many instances a surfactant may have been added as an
emulsion breaker, and it inadvertently acts as an oil wetter.

Dispersibility: Surfactants have been developed to inhibit the swelling,


migration, and flocculation of clays, even in high ionic strength solutions, such
as acids. This is done by using an anionic surfactant which effectively disperses
clay particles in acid mud removal solutions and minimizes the formation of
aggregates. Since the smaller particles are easier to suspend, the clay is
removed more efficiently.

Another system for removing drilling mud is an aqueous solution of surfactants


and other chemicals that reduces mud viscosity and disperses solids.

Retarding Agents

458
These surfactants are designed to slow the reaction time of acid on carbonate
formations. Reaction time is the amount of time it takes for acid to spend from its
original concentration to a concentration of about 3.2% at conditions of bottomhole
temperature and pressure. During matrix or fracture acidizing, or the pumping of large
volumes of acid into a carbonate formation, it is sometimes desirable to control the
reaction time of the acid.

The principal surfactants used as retarding agents are anionic oil-wetting surfactants,
usually sulfonates or sulfates. These function by adsorbing onto the carbonate
formation and making part of it oil-wet. Where the surfactant adsorbs most, the acid
reacts least. Consequently, an uneven surface is etched into the fracture faces. These
uneven surfaces prop the fracture upon completion, resulting in an improved fracture
permeability and increased oil and gas production. Cores etched under simulated
reservoir conditions: Figure 10 (regular 15% HCl )

Figure 10

and Figure 11 (acid retarded with anionic surfactant ) shows this uneven etching of the
acid retarded with an anionic surfactant.

459
Figure 11

Unfortunately, these additives have proven only marginally beneficial in retarding acid
reaction rates in the field, and there use is not generally recommended.

Emulsifying Surfactants

These types of surfactants form with acid either an oil-external emulsion or an acid-
external emulsion. The acid-external emulsion is commonly used with a mixture of an
aromatic and acid to form a dispersion. The solvent phase in this blend dissolves
organic residues, while the acid reacts with scale to remove deposits on the formation
to increase permeability.

The high-viscosity oil-external emulsions are more commonly used in fracture


acidizing. As the acid spends itself on the formation, the emulsion breaks. The
emulsion retards the acid formation reaction by physically preventing acid contact.

Suspending Agents

460
All calcareous formations less than 100% carbonate. Thus many insoluble fines are
released during acidizing. When the spent acid is removed from the formation, these
fines can be left behind to reduce permeability.

Two basic types of suspending agents are used in this application. One is a surfactant
that adsorbs onto these fines and keeps them in suspension by electrostatic repulsion.
The other is a polymer that creates a molecular bridging effect and, by trapping the
fines between polymers, keeps fines suspended. Suspending agents are effective when
acidizing a carbonate formation that contains HCl-insoluble materials, such as feldspar,
quartz, and clays.

Antisludging Agents

When acid contacts some crude oils, a sludge can form at the acid/oil interface. This
usually happens when the crude contains a high percentage of asphaltenes, and is
most severe with high-strength acid systems (20% or higher). Once formed, the
sludge is very difficult to dissolve into oil again. As a result, it accumulates in the
formation and decreases permeability.

To combat formation sludge, surfactants, both cationic and anionic, are used to adsorb
and provide a continuous layer of protection at the acid/oil interface. Sludge
development can also be prevented by lowering acid strength and controlling return
rates after an acid treatment. Figure 12 (acid droplet in sludge-forming oil- a skin like
precipitate forms around the acid droplet )

Figure 12

and Figure 13 (acid droplet containing anti-sludge forming oil- no precipitate forms

461
around the acid droplet ) illustrate the effectiveness of an antisludging agent.

Figure 13

Selection of a particular agent is based on tests of the acid mixture with a reservoir oil
sample.

Penetrating Agents

These surfactants are commonly used in acidizing tight gas reservoirs and in scale-
removal treatments. Their main function is to lower surface and interfacial tensions of
the acid. This allows the acid to more easily wet the formation rock and oil-wet scales,
resulting in a more uniform reaction.

Another benefit of lowering the surface tension of the treating fluid is to decrease the
fluid's likelihood to form water blocks in the formation, resulting in a faster and more
efficient cleanup after acidizing.

Most penetrating agents are nonionic surfactants which water-wet both silicates and
carbonates. These surfactants effectively lower the surface tension of acid from 72
dynes/cm to about 27 dynes/cm when used in a concentration of 0.1%.

Nonemulsifiers

One major problem encountered in acidizing is emulsification of crude oil and acid, or
of crude oil and spent acid. This phenomenon may or may not occur in the formation
but does occur in the wellbore. There are two fundamental types of emulsion: oil-
external and water-external.

462
Oil-external emulsion is commonly encountered in oil wells. It is usually a thick,
viscous fluid mixture which can reduce the flow rate of the well by plugging the
formation. Addition of proper surfactants to the acid before pumping can prevent
formation of an emulsion during the acid treatment. Obviously, the selected surfactant
must be compatible with acid, formation fluids, and all other additives in the treatment
design.

Alcohol

Alcohols are ordinarily used in acidizing fluids to remove water blocks, enhance fluid
recovery, retard acid reactivity, and decrease water content. Their effectiveness is
limited. The most common alcohols used in acidizing are isopropanol and methanol.
Their physical and chemical properties are listed in Table 2.

Properties Isopropanol Methanol

Specific Gravity at 68 F 0.785 0.792

Flash Point (closed cup) F 53.6 53.6

Solubility in Water Completely Completely

Table 2: Physical and chemical properties of isopropanol and methanol.

Isopropanol is normally used at a maximum of 20% by volume. Methanol is used at


various concentrations, commonly 30% by volume. In some cases, methanol alone is
used to dilute concentrated acid to various acid-treating solutions.

Alcohols are used in acidizing fluids for the following reasons:

Removal of Water Blocks: One problem which can severely decrease production in
producing wells is blockage of the pore spaces by water commonly known as a water
block. Water blocks may form where high capillary forces are present in porous rocks. The
most severe water-block problems occur in formations with gas permeabilities less than 120
md. The alcohol in the treating fluid reduces the capillary forces within the reservoir, thus
permitting easier removal of the liquid phases.

Enhancement of Fluid Recovery: Another problem encountered in treating oil


or gas wells is the recovery of treating fluids, especially in gas reservoirs. The
high surface tension of water or acid solution hinders their penetration and
recovery. Conventional surfactants help somewhat, although they surfactants
lose much of their surface activity by adsorption. The addition of alcohol to acid
solutions reduces their surface tension. The concentration of alcohol normally
used for this purpose is of sufficient quantity that adsorption is not a problem.
Alcohol used in acidizing fluids also increases the vaporization rate of the fluid.
This, in turn, lowers vapor pressure, allowing for a more rapid desaturation of
water from a gas reservoir.

463
Retardation of Acid Reactivity: Alcohol has a retarding effect on acid reactivity.
The retardation rate is proportional to the type and percentage of alcohol
added.

Decrease of Water in Acid: Some formations contain a large amount of water-


sensitive clays. To minimize the amount of water contained in acidizing
solutions, alcohols are used in place of part or all of the dilution water.

The major disadvantages of using alcohol in acidizing fluids are:


Effective Concentration: It takes a large amount of alcohol, 20% or more, to provide
beneficial effects.

Cost: Replacing water with alcohol in the acidizing solution makes the
treatment more expensive.

Low Flash Point: Both isopropanol and methanol, and even acid solutions
containing 20% or more by volume of either, have low flash points.

Increase in Corrosiveness: Corrosion tests have shown that alcohol-acid


mixtures require a higher concentration of inhibitor than equivalent nonalcohol-
acid mixtures.

Adverse Reactions: Formation brines with a high concentration of dissolved


salt solids can "salt out" in the presence of alcohols. To help prevent the
occurrence of this salt precipitation, do not exceed 20% by volume of
isopropanol or 50% by volume of methanol in treating solutions.

Incompatibility: Some crude oils are incompatible with both methanol and
isopropanol. Conduct compatibility tests before acidizing a well with a fluid
containing alcohol. Some formation types may be extremely sensitive to
aqueous solutions that contain high concentrations of alcohols.

In some cases, formations should not be acidized with alcohol-acid solutions. The formation of
undesirable methyl chloride is possible with HCl solutions that contain 30% or more methanol at
bottomhole temperatures above 200 F.

Mutual Solvents

A mutual solvent is a material that is soluble in both hydrocarbon and aqueous


solutions. It can help to solubilize an aqueous solution into a hydrocarbon solution or
vice versa. Common mutual solvents include ethylene glycol monobutyl ether
(EGMBE), diethylene glycol monobutyl ether (DEGMBE), modified glycol ethers (MGE),
and also alcohol.

The use of mutual solvents, particularly EGMBE and MGE, in acid stimulation of a
sandstone reservoir is a common practice. Mutual solvents have been used in
postflushes because they help to:

reduce water saturation from around the wellbore by lowering the surface tension of the
water, thus preventing water blocks

464
solubilize a portion of the water into the hydrocarbon phase, thus reducing the
amount of irreducible water saturation

provide a water-wet formation, thus maintaining the best relative permeability


for oil production

prevent insoluble fines from becoming oil-wet

stabilize emulsions

maintain the needed concentration of surfactants and inhibitors in solution by


helping to prevent adsorption of these materials onto the formation

Mutual solvents have also been found to reduce the formation damage that can result from using
corrosion inhibitors in acid treatments. The normal treating concentration of mutual solvents varies
from 2 to 50% by volume, depending on the application. Most sandstone acidizing applications use
10% or less.

Iron Sequestering Agents


Iron occurs naturally in formation water, and in formation minerals such as siderite,
hematite and pyrite. A primary source of iron, of course, is donwhole tubular goods. In
solution, iron can exist in two forms: ferrous iron (Fe2+) and ferric iron (Fe3+). Fe3+ is or
primary concern in acidizing, because it precipitates as gelatinous iron hydroxide
(Fe(OH)3) at a pH of 2.5 to 3.5. This precipitation can lead top formation damage.

When Fe3+ is present in the near-wellbore region, an iron sequestrant, or chelating


agent, may be added to the treating solution. Table 1 (below) provides information
about various sequestering agents, of which EDTA is the most commonly used.

Because these agents can themselves damage the formation, they should be used only
when pr-job evaluation and testing indicate a tendency for iron hydroxide precipitation
during acidizing.

Other methods for keeping iron in solution include pH control and the use of oxygen
scavengers. pH control is achieved by using a weak acid, such as acetic, which reacts
much more slowly than HCl on limestone and other acid-soluble materials. Thus, some
of the acid remains in solution, keeping the pH low and inhibiting iron Hydroxide
formation.

Compound Maximum Amount of Fe +2 Chelated - ppn


Solubility in
1000 gal 15%

EDTA (Ethylene Diamine 67 1,450


Tetraacetic Acid)

Sodium EDTA ~250 ~5,000

465
NTA (Nitrilo Triacetic 420 13,790
Acid)

Citric Acid 1,796* 57,292

*The solubility of citric acid is considerably higher than 1,796 lb. Theoretically, ferric
iron in 15% HCl spent on millscale can be chelated with 1,796 lb of citric acid.
However, if the acid was only partially spent on magnetite and the remainder spent on
a carbonate rock, calcium citrate could precipitate.

Table 1: Solubility and iron chelating capability of various chelating agents.

Oxygen Scavengers: An oxygen scavenger can be used to economically control


concentrations of iron up to approximately 5,000 mpl. It functions by two different
mechanisms. First, it removes free oxygen from the fluid, and then helps prevent the
oxidization of ferrous iron (F3+2) to ferric iron (Fe+3). Then, it acts as a reducing agent
and reduces any dissolved ferric iron present to ferrous iron. This helps maintain iron
in solution and helps prevent the precipitation of ferric iron. This helps maintain iron in
solution and helps prevent the precipitation of ferric iron in solution. The amount of
iron that can be reduced with an oxygen scavenger depends upon the quantity of
chemical added, but it can be used up if the fluid is aerated before it is pumped.

A comparison of the amount of ferric iron that can be retained by the various methods
and additives is illustrated in Figure 1 (iron cotnrol acid systems).

466
Figure 1

The safest way to help prevent damage to the reservoir from precipitated iron
hydroxide is to clean or "pickle" the pipe with acid before acidizing the formation. This
acid should contain large quantities of iron-control additives and should be circulated
out of the well, not pumped into the formation. In conduction with this treatment,
xylene or hydrocarbon phase should be incorporated or used as a preflush to remove
pipe dope and varnish that could plug perforations. ( In most instances, pipe dope is
not soluble in xylene or normal hydrocarbons.)

Fluid Loss Control Additives


In fracture acidizing, fluid loss control helps to extend fractures by minimizing leakoff
of the treating fluid through the fracture face. Fluid loss control is difficult to obtain
with acid solutions on carbonate formations because the acid is reacting on the surface
where the fluid loss additive is being deposited.

Fluid loss additives often are composed of two agents: an inert, solid particle that
bridges at the fracture surface and a gelatinous material that plugs the pores in the

467
solid granular material. Commonly used fluid loss additives are listed in Table 1 .

Fluid Type Solid Additives Gelatinous Additives

Aqueous pad Silica flour Guar

Calcium carbonate Cellulose

Organic polymer Polyacrylamide

Inert solid coated with


guar type material

Hydrocarbon pad Inert solid coated with


organic sulfonate

Acid Acid swellable solid Guar

Organic resin Karaya

Silica flour Cellulose

Organic polymers Polyacrylamide

Polyvinyl alcohol

Table 1 : Commonly used fluid loss additives.

In aqueous pad fluids, a combination of solid additive and polymer is usually selected.
When water without a polymer additive is used, an inert solid coated with guar-type
material may be chosen. Acid additives include acid-soluble organic polymers that
swell and become soft when contacted by acid, thereby giving both solid and
gelatinous characteristics. Other additives are similar to those chosen for use in water-
base pad fluids.

We now focus on matrix fluid loss, fluid loss in natural fractures, fluid loss in
wormholes, and viscosity effects.

Matrix Fluid Loss: Many carbonates contain little permeability within the matrix of the
rock. Because acid is a reactive substance, however, it creates more surface area as it
travels during leakoff. It is very important, therefore, to confine the acid to the
fracture. Fluid loss additives are usually sufficient to control matrix leakoff. Ideally,
they should be either degradable or slowly soluble in the acidizing solution or in
produced fluids.

Fluid Loss in Natural Fractures: Some limestone formations have unusually high
permeability in the form of natural fractures or vugular flow channels. Fluid loss control
in this type of formation cannot be achieved with ordinary fluid loss additives. A larger
diverting material is usually required to control the fluid lost to natural fractures.

468
Fluid Loss in Wormholes: Fluid lost to the matrix will react and, if lost in a highly
reactive part of the formation, may cause wormholes. These are large, highly
conductive channels which usually start when a large pore or vug grows at a rate
substantially higher than that of smaller pores and receives an increasingly larger
amount of the acid as new surface area is created. From a fluid-loss standpoint,
wormholes are similar to natural fractures since they require larger particles for
control. If little fluid leakoff occurs to the matrix, few wormholes are likely to occur.
For a more detailed discussion of wormhole formation and growth, see Economides et
al. (1995).

Viscosity Effects: The rate of leakoff is directly related to viscosity. A thicker fluid
results in less fluid lost to the formation and, therefore, improved fluid efficiency. If
efficiency increases, more live acid can be pumped deeper into the formation, thus
increasing fracture flow capacity. Viscosifying the acid can be one of the most
important steps in providing fluid-loss control in fracture acidizing treatments.

Friction Reducers

Friction reducers suppress turbulence and reduce friction pressure in fluids flowing
through tubular goods. This can either reduce horsepower requirements or increase
treating rates. Most friction reducers are long-chain natural or synthetic polymers, but
not all of them work effectively in acidizing solutions. Acid breaks down some of these
polymers at a very fast rate, thus leaving little, if any, friction-reducing properties.

Many different additives are used as friction reducers. Generic classifications of


additives for water- and oil-base pad fluids and for acids are listed in Table 1.

Fluid Type Generic Classification of Additive

Water based pad Guar and HP Guar

Polyacrylamide

Cellulose

Oil based pad Polyisobutylene

Fatty acids

Crosslinked organic polymers

Acid Guar

Gum karaya

Polyacrylamide

Cellulose

Table 1: Friction reducers.

469
Clay Stabilizers
Many clay-stabilizing materials have been developed in recent years to combat
formation damage caused by clay swelling and migration. These generally are used in
conjunction with minimum-salinity fluids such as 2% KCl, 4% NaCl, and 3% NH4Cl.

Types of Clay

Some clays, like smectite-mixed layer, swell when they contact fresh water or water
with low salt concentrations. There are several ways in which water can enter and
damage a productive zone. Some of these are:

water filtrate from drilling fluid;

loss of water from completion or workover fluids;

encroachment into the reservoir by water coning;

any type of water-base treating fluid.

Productivity also decreases when the produced fluid carries the clay fines (colloidal-size clay
particles) in suspension to the wellbore. When any type of pore restriction is encountered, the fines
bridge off in the pore, thereby plugging the pore and restricting the flow. Radial flow volume of fines
increases while the number of pore channels decreases. Some clays with migrating tendencies are
illite, kaolinite, and chlorite.

Smectite-Mixed Layer: Smectite-mixed layer, a three-layer clay, is composed of two


silica tetrahedral sheets with a central octahedral sheet ( Figure 1 , smectite -mixed
layer).

Figure 1

470
In the stacking of their mineral sheets, the proximity of two oxygen layers makes for a
weak chemical bond that allows water and other polar organic minerals to enter
between unit layers and cause the clay to swell. Figure 2 (three layer clay ) illustrates
the bonding of smectite-mixed layers. The cation exchange sites attract such positive
ions as sodium, potassium, calcium, or ammonium.

Figure 2

Illite: Illite is a clay composed of two silica tetrahedral sheets with a central alumina
octahedral sheet ( Figure 3 ).

471
Figure 3

Its structure is similar to smectite, except that potassium and oxygen bind the sheets
tightly. Hence, this clay does not expand in the presence of water. However, the
potassium ions are very susceptible to the leaching action of slightly acidic water. In
some cases, the potassium may be removed and replaced by another cation. These
altered minerals, called degraded illites, then allow entry of water and will swell as a
result.

Kaolinite: Kaolinite is composed of one silica tetrahedral sheet and one octahedral
sheet ( Figure 4 ). Their tight hydrogen bonding is not susceptible to water entry.
Although kaolinite swells very little on contact with fresh water, it is an important
component of migrating fines.

472
Figure 4

Figure 5 (two layer clay ) schematically illustrates a two-layer clay.

Figure 5

473
Kaolinite is the principal member of this family of clays. There are few cation exchange
sites: 10% of the number for smectite-mixed layer too few to cause significant
swelling.

The degree to which a clay will swell depends on the ions the clay has adsorbed, the
degree of degradation of the clay, and the amount of substitution in the clay lattice.
The amount of swelling of several salts of smectite is shown for various concentrations
in Table 1 (below). These data show that certain ions, especially sodium, allow
swelling of clays, resulting in formation damage. In Figure 6 (salinity effects on clays),
solutions of sodium chloride (NaCl) were flowed through columns of quartz sand
containing 4% by weight of various clays.

Figure 6

These curves show that the higher salinity solutions of NaCl did less damage to the
permeability of the columns than the lower salinity and fresh water systems.

474
Salt Concentration Sodium Potassium Calcium
Smectite Smectite* Smectite

Distilled Water 33 27 4

0.4% Chloride 14 2 4
Solution

2.0% Chloride 3 2 4
Solution

8.5% Chloride 2 2 4
Solution

10.0% Chloride 2 2 3 to 4
Solution

*Ammonium chloride will perform essentially the same as potassium chloride.

Table 1: Effect of salt concentration on the swelling of smectite.

The data in Table 1 and Figure 6 show good agreement. Both indicate that low
concentrations of salt swell the smectite more severely than do higher salt
concentrations. Figure 6 also indicates that the sensitivity of the clays to water swelling
is

kaolinite < illite < smectite


with kaolinite showing the least damage and smectite the most.

Clay damage may be caused by both swelling and migration. When these conditions
occur in a formation, the damage must be removed to improve production. This is
commonly accomplished by acidizing the formation with HCl-HF mixtures.

Clay Control

Formation damage due to clays must be prevented rather than cured. When clay
particles in sandstone reservoirs are rearranged or disturbed in any manner, it is
usually impossible to restore the original permeability.

X-ray diffraction analysis can easily determine the type and amount of clay in a
particular sandstone. These tests indicate which formations warrant a particular
minimum-salinity fluid, clay stabilizer, or special completion technique to avoid
formation damage.

The position of clays in the rock can be quickly determined by the use of dyes that
exhibit characteristic colors when absorbed by different types of clays. Where detailed
studies are warranted, the scanning electron microscope (SEM) permits direct
observation of certain actions on clays.

475
In a virgin formation, the swelling of a clay particle is in equilibrium with the type and
concentration of salts in the connate water. Thus, filtered formation water used in
workover operations should not disturb this balance. If clean formation water is not
available, it is necessary to synthesize an economical substitute.

Cat ionic polymers have been developed that can plate out of fresh water, brine, or
acid to stabilize clays and minimize damage. These cationic materials do not oil-wet
because the exposed polymer prefers water over oil. The cationic ends of the clay
stabilizers bind to the clay minerals in the same manner as the sodium Na+, potassium
K+, or calcium Ca++ ions bind. This leaves the water-soluble end of the molecule to form
a film over the clay minerals to prevent their swelling and migrating. The principal
cationic polymers currently available for use in sandstone acidizing include
polyquaternary amines (PQA) and polyamines (PA).

Polyquaternary Amines (PQA): PQA is a cationic organic polymer used as a clay-


stabilizing additive in sandstone acidizing applications and other treatments. It is a
clay-stabilizing chemical that can be universally placed out of aqueous solutions and
strong acids.

Since PQA does not require a shut-in period, it allows faster well cleanup and prevents
damage resulting from swelling and/or migration of clay particles and other silicate
fines. The chemical action between PQA and clays results in longer stabilized
production before migrating and/or swelling silicates begin to redamage the formation.

Figure 7 illustrates relative permeability in a Bandera sandstone core without any clay
stabilization treatment.

476
Figure 7

Notice that after the core was calibrated with 2% KCl and then exposed to fresh water,
permeability decreased about 85%.

Figure 8 demonstrates relative permeability in a Bandera sandstone core after a clay-


stabilization treatment with PQA (polyquaternary amine).

477
Figure 8

Following 2% KCl calibration and clay-stabilization treatment, fresh water had only
minimal effect on the permeability, decreasing it slightly.

Although the PWA concentration needed depends on permeability, the usual maximum
is 3% by volume.

All fluids containing PQA should be overflushed with a sufficient volume of fluid
containing PQA to ensure that they are displaced to a distance of at least two feet from
the wellbore. This results in more fines stabilizing for a given PQA concentration.

Polyamines: These are cat ionic polyelectrolytes which stabilize water-sensitive clays
against the swelling and migration problems that might occur on contact with aqueous
fluids. PA can be used in aqueous preflush and overflush fluids and acids.

Bacteria Control Agents


Types of Bacteria

Bacteria can cause problems in oil-field operations. Even though these bacteria are
extremely small (1 to 3 microns), their sheer numbers can cause plugging problems in
injection wells.

478
Generally, the worst problems result from the by-products or chemical processes of
certain organisms. Some of these problematic bacteria are sulfate-reducing bacteria,
slime-forming bacteria, and iron bacteria.

Sulfate-reducing Bacteria The bacterium causing the greatest concern to oil producers
is the sulfate-reducing organism. This bacterium, an anaerobic organism, needs an
oxygen-free atmosphere in order to spread. Under anaerobic conditions, it can produce
H2S. This gas can then corrode downhole tubulars, producing iron sulfide (FeS), which
is insoluble in water and acts as a plugging agent.

Slime-forming Bacteria Other troublesome bacteria are the slime-forming types. Under
aerobic conditions, they may produce large slime masses that cause severe plugging
problems. These bacteria are usually found where fresh or surface water comes in
contact with produced water.

Iron Bacteria A very common bacterium that can cause problems is the iron bacterium.
This aerobic bacterium has the ability to oxidize water-soluble ferrous ion to a water-
insoluble ferric ion in its metabolic processes. Iron bacteria can cause plugging of
source well, filters, and surface lines.

Bactericides

Many chemicals can be used to control the growth of microorganisms. The most
common term for chemical agents designed to kill and inhibit the growth of bacteria is
bactericide or biocide. Use of a bactericide can reduce or eliminate the severe problem
associated with bacteria.

Many bactericides, both cationic and nonionic materials, can be used in acidizing
solutions. These chemicals are effective at killing bacteria in concentrations from 1 to 2
gal/1000 gal.

479
MATRIX ACIDIZING

Matrix Acidizing
This type of acidizing treatment is designed to remove skin damage that extends
beyond the immediate surface of the perforations or the face of the producing zone.
Acid enters the pore space of the formation below fracturing pressure and dissolves
both the surface of the rock and the flow restricting contaminants within the matrix.

A matrix acidizing treatment may be effective penetrating into the formation only a few
inches or up to several feet from the wellbore, depending on formation permeability. In
wells that have suffered from skin damage, production can increase many-fold.
However, if a well had little or no skin damage, an acidizing treatment stimulates
natural production little more than one to two times, depending on its design. Figure 1
(stimulation of natural production ) illustrates this phenomenon.

Figure 1

Matrix acidizing regains permeability by removing formation damage from the


perforation tunnels and existing flow channels around the wellbore to a depth of only a
few inches. The amount of lost production is directly related to the extend and depth of

480
formation damage, as shown in Figure 2 (effect of damage on well productivity ).

Figure 2

The greatest productivity loss occurs in wells with severely damaged permeability
within the first few inches from the wellbore. According to Schecter (1992), three of
the primary mechanisms that contribute to an acidized well's productivity include
"erosion of the pore structure as the acid flows through these pores, consumption of
the acid and, selective dissolution of certain minerals. Erosion of the pore structure
leads to both increase porosity and permeability...

Matrix Acidizing of Sandstone


Matrix acidizing in sandstone achieves the natural true permeability of the formation
by removing clay damage. The stimulation fluids are pumped into the porosity (pore
spaces) of the rock below fracturing pressure.

Since the acid is exposed to and reacts on such a large area of the formation,
unreacted acid can be effective only for a short distance into the reservoir. Thus, this
type of treatment is primarily designed to treat shallow damage in the immediate

481
vicinity of the wellbore. Retarded hydrofluoric acid systems can penetrate farther. Of
course, the effective penetration is also a function of the dissolving power and reaction
rate, which depends on the properties of the acid and the mineral in the formation.

Reasons for Treatment

Sandstone formations usually have clays that swell or migrate, plugging pore spaces
and decreasing permeability. Sandstone acidizing restores natural permeability and
increases productivity by removing clay damage and controlling fines migration.

Figure 1 (effects of damage zone-permeability decrease ) demonstrates damage ratio


at various depths of damage for varying degrees of permeability decrease. Obviously,
the greatest damage occurs near the wellbore.

Figure 1

Usually, acidizing an undamaged sandstone formation is not desirable. Theoretical


production increases obtained from these treatments are rarely above 1.3 not
enough to justify the acidizing cost. In some instances, however, a highly productive

482
undamaged well can be stimulated enough by acidizing to show an economic
advantage: a production increase results from pore enlargement in the zone adjacent
to the wellbore. Figure 2 shows the effect on production of an improved permeability
zone.

Figure 2

Treating Acids

Acids used in stimulating sandstone reservoirs generally contain some form of the
highly reactive fluoride ion ( F- ). This ion is the only one that reacts significantly with
sand and clay. Treating acids containing the fluoride ion include HCl-HF and HF-organic
acid mixtures.

HCl-HF Mixtures This acid blend can be prepared by one of three methods: dilution of
concentrated HF or, as is more frequently done, by mixing Hcl with ammonium
bifluoride (NH4HF2).

483
HF-Organic Acid Mixtures: Blends of HF-organic acids, are used occasionally to retard
the reaction of acid on sand and clay, and to reduce corrosiveness. These can provide
deeper penetration of unspent HF and, consequently, effective removal of deep
damage. Either HF-formic or HF-acetic can be prepared from liquid HF, solid
ammonium bifluoride (NH4F2), or ammonium fluoride (NH4F), and the corresponding
organic acid.

All acid solutions should be prepared with fresh water. Sea water or brines should
never be used in preparing acidizing fluid for HF-type treatments. Since these waters
may contain sodium or potassium ions, they can cause formation-damaging
precipitants.

HF reacts with SiO2 to give fluosilicic acid (H2SiF6) and water. This is one of the more
important reactions of HF.

Rate of the reaction is primarily dependent on the following variables:

type of mineral present


size of sand grains
temperature
volume and/or strength of HF
concentration of HCl contained in the blend
Decreasing sand-grain size increases surface area for a given weight of sand. The larger the
surface area, the faster the acid reacts. The effect of temperature on the reaction rate is shown in
Table 1(below).

Temperature Reaction Rate Constant (cm/min)

75 F 3.89 x 10-5
-
100 F 5.59 x 10 5
-
150 F 11.27 x 10 5
-
200 F 20.05 x 10 5
-
300 F 51.46 x 10 5
Table 1: Effect of temperature on reaction of HF acid on sand.

This data show that increasing temperature significantly increases HF reaction rate on
sand. Reaction rate is about 13 times faster at 300 F than at 75 F. The primary
control on reaction rate, however, is the rate of acid movement and whether the acid
contacts a fast or slow-reacting material. HF acid, for example, reacts relatively slowly
with quartz, but much more quickly with feldspar or kaolinite. Increasing the acid
volume and/or strength increases the distance unspent HF can be pumped into the
formation.

Varying the concentration of HCl in the acid blends also affects the HF reaction rate on
sand ( Figure 3 ., influence of HCl on HF-sand reaction ). Increasing HCl concentration
in HCl-HF mixtures increases the amount of sand that can be dissolved by equivalent

484
concentrations of HF. Blends with greater HCl strengths react longer.

Figure 3

Fluosilicic and fluoaluminic acids, reaction products between HF on clay or sand, can
react with sodium ions (Na+) or potassium ions (K+) to form precipitates.

The reaction of the sodium ion is:

H2SiF6 + 2Na+ > Na2SiF6 + 2H+ (22)

H2AlF6 + 3Na+ > Na3AlF6 + 3H+ (23)

The reaction products, sodium fluosilicate and fluoaluminate, are white, gelatinous solids which can
partially plug permeability. These potentially damaging precipitants can be prevented by proper acid
placement techniques. On the other hand, ammonium fluosilicate and ammonium fluoaluminate,
reaction products with the ammonium ion, are completely soluble.

485
Reaction of HF on limestone (CaCO3) produces a very fine white precipitant: calcium
fluoride (CaF2),along with water and carbon dioxide. Calcium fluoride can cause
formation damage, but this problem is avoided by proper placement of HCl.

Treatment Technique

The three key stages of fluid placement during an HF treatment are preflush, HF acid,
and overflush or afterflush. Although various fluids may be pumped ahead of the
preflush or behind the overflush, no change should be made in these three basic
stages. This treatment technique is designed to be compatible with natural formation
fluids. Figure 4 illustrates the basic HF acid treatment design.

Figure 4

Preflush: A preflush fluid should always be pumped ahead of a fluoride-containing


solution (e.g., HF acid). It forms a vital barrier between the acid and the formation
fluids which prevents the formation of insoluble precipitants, such as sodium and
potassium fluosilicates and fluoaluminates, and calcium fluoride (a reaction product of
HF with limestone). If HCl is used as a preflush, it dissolves the limestone or dolomite,
thereby reducing the possibility of the formation of insoluble precipitants. The preflush
fluid must be compatible with both the formation fluid and the HF treatment.

486
A solution of 5 to 15% HCl or acetic acid is popular for displacing formation brines to
prevent them from mixing with reacted HF. The solution is also useful in the removal of
small amounts of calcareous cementing material. Both conditions can result in the
development of insoluble precipitants.

Ammonium chloride (NH4Cl) is useful as a preflush if calcium chloride (CaCl2) has been
used as a workover fluid. The calcium chloride should be either circulated out with a
3% NH4Cl solution or isolated behind a packer. Any calcium chloride left in the wellbore
should be preflushed away into the formation with HCl or NH4Cl solution.

Diesel oil, kerosene, and certain crude oils may also be used as preflushes. However,
their use is dependent on total compatibility with the formation fluids, the HF acid
treatment, and all surfactants included in the treatment sequence.

HF Acid: After the preflush, HF acid is pumped into the formation. The treatment fluid
should be of an adequate volume and have proper concentration. A 12% HCl-3% HF
solution is the usual concentration for damage removal in a high-quartz, low-clay
formation. A 13.5% HCl-l.5% HF solution may be used in a high-feldspar formation,
and a 6.5% HCl-l.0% solution where high clay content is a factor. In extremely "tight"
formations (i.e., those with less than 1 md permeability), a 3% to 7% HCl-0.5% HF
solution may be used. Table 2 (below) provides more detailed acid-use guidelines for
sandstone acidizing.

Situation Acid Type

HCl Solubility > 20% HCl only

High Permeability (100 + md)

High Quartz 80%, Low Clay <5% 12% HCl - 3% HF1

High Feldspar >20% 13.5% HCl - 1.5%


HF1

High Clay >10% 6.5% HCl - 1.0% HF2

High Iron Chlorite Clay 3% HCl - 0.5% HF2

Low Permeability (20- md)

Low Clay <5% 6% HCl - 1.5% HF3

High Chlorite 3% HCl - 0.5% HF4

1 Preflush with 15% HCl

2 Preflush with Sequestered 5% HCL

487
3 Preflush with 7.5% HCl or 10% Acetic

4 Preflush with 5% Acetic

Table 2: Acid use guidelines for sandstone acidizing (From McLeod, H. O.: "Matrix
Acidizing", Journal of Petroleum Technology, Dec. 1984, pp. 20552069).

As mentioned earlier, when HF is pumped into a formation, a decrease in permeability


occurs, with a possible corresponding increase in pump pressure. This condition is due
to the initial rapid reaction of the clay, precipitation of hydrous silica, and the
dislodging and migration of fines. The 3% and 6% HF concentrations demonstrate this
initial reduction in permeability, followed by overall permeability increase.
Occasionally, as demonstrated by the 9% HF concentration, the damage is complete,
with the wellbore totally plugged ( Figure 5 , change in permeability observed during
acidizing with various strength HF acid systems).

Figure 5

No more acid can be injected at this point without fracturing the rock. To maintain
permeability, 3% HF is used for most treatments; in high clay content formations,
even lower-strength HF should be used. Lower-strength HF acid concentrations
generally present fewer problems than higher concentrations.

488
Overflush: The overflush, or afterflush, is designed to minimize the precipitation of
Si(OH)4 , which can limit the success of acid treatments. The overflush is also used to:

displace unreacted HF into the formation, thus reducing corrosion, leaving more acid to
react in the formation, and allowing for a shut-in period, should one be required (e.g., ahead
of a gravel pack)

displace HF reaction products away from the wellbore

remove the oil-wet relative permeability problems caused by cationic


surfactants, such as corrosion inhibitors

stabilize clays and fines

reestablish oil or gas saturation near the wellbore

Typical overflushes for HF treatments are 3% NH4Cl, 3% to 7% HCl, organic acid, diesel oil,
kerosene, crude oil, and nitrogen (commingled with any of the above fluids, or by itself). Aqueous
ofterflushes are the most effective in displaying spent acid from the near wellbore region.

The most common is 3% NH4Cl, because it is one of the few salts that will not
precipitate insolubles with active or spent HF. Diesel is another frequently used
overflush in oil-producing wells. Often, a mutual solvent is added to the overflush to
improve saturation and relative permeability characteristics.

HF reacts very rapidly on clays. Consequently, a long shut-in time is not required. A
slow return of the treatment load should begin as soon as practically possible. This
minimizes fines movement; those that do move are more likely to come back to the
wellbore without bridging. Allowing the reacted acid system to remain in the formation
for extended periods of time increases the chance of fluid intermixing. Also, energy
used to place all treating fluids may be dissipated.

Use of Additives

During acid treatment of sandstone formations, fluids pumped into the rock pores
contact a large surface area, where capillary pressures begin to govern flow rates and
fluid recovery.

Thus, selecting the proper additives ensures that

treated fluids are used throughout the acidizing sequence

the additives are not adsorbed onto rock surfaces to leave virtually untreated
acid solutions to penetrate the deeper portions of the formation

Figure 4 (HF acid treatment design), shows the preflush fluid in direct contact with formation fluids.
To prevent emulsion problems and sludge deposits, the preflush fluids must be adequately treated
to maintain compatibility with formation crude oil.

489
If the additives have a tendency to adsorb on the rock, steps must be taken to
minimize their adsorption enough to allow adequate penetration. Surfactants can
reduce the acid's surface tension, and thus reduce capillary pressure and aid in
cleaning up the reacted acid solutions.

Treatment Design

Before designing a sandstone treatment, the engineer should gather as much


information as possible about the composition and properties of the formation.
Although well data, cores, and formation materials are not usually available, the
engineer needs to obtain or estimate information about the following:

formation permeability

fracturing pressure

condition of the tubing

wellbore geometry

Formation Permeability: Formation permeability must be estimated to calculate the injection rate
possible during treatment stages. Generally, a minimum injection rate of 1/4 BPM should be
achieved during final HF placement and overflush. Sometimes early injection rates may be very low,
making it necessary to use the hesitation-squeeze technique of acid placement. However, if a
proper analysis has been made of the wellbore condition by previous well testing or other
estimates, the injection rate can be estimate before HF reaches the formation. Then, if the
calculated treatment time is too long, a mini HF treatment can be run to open up more injectivity
before the main acidizing volume.

Fracturing Pressure: Fracturing pressure must also be estimated to determine the


allowable injection rate into a formation. Fracture gradients of new wells are usually
available, but these values decline as the formation is depleted. A useful technique is
available to estimate the effect of reservoir pressure decline on the fractured gradient.

Condition of the Tubing: Any rathole requires isolation or cleanout prior to acid
placement. Tubing should be run to bottom, and the well circulated with a formation-
compatible fluid to clean the wellbore of all debris and sludge. Otherwise, during the
treatment the more dense acid will displace the rat-hole fluid and allow this debris and
sludge to enter the perforations, potentially causing severe plugging.

Wells that periodically scale up can be treated for calcium scale by circulating acetic
acid down the annulus and up the tubing. This action removes cumulative calcium
carbonate scale from the downhole pump and around perforations. Should iron scale
exist in the tubing, acid can be pumped down the tubing through coil tubing to bypass
the inner pipe wall. Coil tubing is especially useful in injection wells, where it allows
acid to bypass (and thus protect) the injection tubing string.

Injection water often contains sufficient oxygen to cause oxidation corrosion. Acid
removes the iron oxide from tubing walls coated with iron hydroxide. This reaction
removes some protection from water with dissolved oxygen. However, ferric chloride

490
created by the acid and iron oxide reaction causes more corrosion; this type of
corrosion cannot be prevented by acid corrosion inhibitors. Bare metal can be
protected from HCl by inhibitors, but inhibitors do not prevent the oxidation-reduction
process between ferric iron and pure tubing iron.

Tubing in the well if it is known to be clean, or it can be cleaned in place can be


used to inject acid. If the tubing has considerable scale buildup or if paraffin and/or
asphaltenes have accumulated inside, the tubing should be bypassed with a concentric
string or replaced with a special treating string. If a treatment string is used, however,
pipe dope application to the threads should be closely supervised. Excessive pipe dope
can cause perforation plugging and less than desirable acidizing results. The use of
plastic-coated tubing eliminates the need to clean or "pickle" the tubing with acid prior
to acid placement.

Wellbore Geometry: The mechanical aspects of acid placement must be selected


according to well conditions and wellbore geometry.

In wells with long multiple-producing intervals, zone coverage control is needed. The
most effective treatment is with an opposed cup packer (wash tool). It can distribute
the HCl preflush over every foot of zone, followed by a similar run with the HF acid
stage and then the overflush. This technique ensures that every part of the perforated
interval is treated.

Another way to acidize the productive zone effectively is with a retrievable bridge plug
and squeeze packer positioned, respectively, below and above selected perforations.

Moreover, successful use of these techniques requires an excellent cement job. A


cement bond log (CBL) run after primary cementing can help to show whether zone
isolation has truly been achieved. If there is some question about cement job integrity
in the vicinity of the intended perforations, diverting materials should be considered.

Tubular Cleanup Fluid: If new tubing is in the well, the safest way to help prevent
formation damage is to pickle the pipe with acid containing large quantities of iron-
retention additives, and to circulate the acid out of the well. Xylene or hydrocarbon
phase should be used as a preflush for the removal of pipe dope and varnish, which
can plug perforations. A suggested tubular cleanup procedure includes 15% HCl to
remove the mill scale which is in most new tubular goods. Sequestering additives
maintain the dissolved iron in solution and the penetrating, nonionic surfactant helps
the acid contact the pipe and scale after the xylene. This cleaning and pickling process
helps prevent the deposition of debris into the formation during the stimulation
treatment.

Selecting Stage Volumes: The engineer can better select sandstone acidizing stages
and stage volumes by analyzing the well.

Formation Injection: A hydrocarbon preflush should be placed in the formation ahead of


the HCl preflush if crude oil/acid incompatibility exists. Should a problem crude and calcium
coexist in the reservoir, an aromatic acid dispersion should be squeezed into the formation
to remove both calcium carbonate and paraffin/asphaltenes.

491
Preflush: The volume of preflush fluid should be large enough to accomplish
two goals: (1) dissolve the calcium carbonate contained in the zone contacted
by unreacted HF and (2) provide sufficient fluid at least 2 ft from the wellbore
to serve as an adequate barrier between the reacted HF and formation brine.

Therefore, the volume and strength of the preflush is somewhat contingent on


the calcium carbonate content of the formation. HF reaction products (H2SiF6,
etc.) are somewhat compatible and soluble in the presence of calcium
carbonate; hence, it is only necessary to remove the calcium carbonate from
the zone contacted by HF to prevent precipitation of calcium fluoride.

HF Acid: The HF stage should be designed for at least two hours (preferably
four hours) of contact time around the wellbore.

Over flush: The overflush should be designed to displace the HF to a radius of


3 to 5 ft around the wellbore, or the overflush volume should at least equal the
HF stage. This pushes hydrous silica and other insolubles out to a safe distance,
minimizing the damaging effects from these solids during subsequent
production.

Stage volumes for the acid treatment are determined by the allowable injection rate in 8 to 10 hours
of daylight acidizing time. First priority is the overflush, second is the 2 to 4 hours of HF contact
time, and third is total acid pumping time. Nitrogen or CO2 can be added to the various stages,
depending on reservoir pressure and the anticipated difficulty in recovering the treatment.

Matrix Acidizing of Carbonate


In matrix acidizing, HCl is injected into the formation below fracturing pressure. The
objective of such a treatment is to achieve radial acid penetration and increase
permeability near the wellbore. The main application for this technique is to remove
near wellbore damage and restore productivity in highly permeable zones.

A major problem in carbonate acidizing is that HCl usually reacts too fast. Generally,
the main factors that affect HCl reaction rate on limestone and dolomite are
area/volume ratio and diffusion.

Area/Volume Ratio: HCl reaction time is indirectly proportional to the carbonate


surface area in contact with the acid. Extremely high area/volume ratios are the
general rule in matrix acidizing. Therefore, obtaining significant acid penetration before
spending during a matrix treatment is almost impossible. Figure 1 (effect of area-
volume ratio on HCl-CaCO3 reactions) shows the comparative effect of area/volume
ratio on HCl spending time in a 6-in.

492
Figure 1

wellbore 0.1-in. fracture, and matrix of a limestone formation.

Diffusion: Diffusion is the time required for spent HCl on the formation surface to be
replaced by live acid in solution.

Well Candidates

Carbonate formations damaged during drilling or completion are generally the best
candidates for matrix acidizing treatments, because natural production can usually be
restored. In some cases, however, other types of treatment may be equally effective.

If the formation is homogeneous, such that an induced fracture will heal, hydraulic
fracturing with proppant may be a better stimulation choice. Nearby water zones would
definitely be a factor in considering a matrix treatment, but, again, a low-injection
fracture acidizing treatment may yield better results.

For example, consider a fracture acidizing treatment designed for significantly


increasing production and minimizing bottom water production. This design involved a
well with a 60-ft perforated interval in a carbonate formation 200 ft above an oil-water
contact. There were two choices in attempting to stay away from the water: (1) a true

493
matrix acidizing treatment and (2) a controlled height fracture acidizing treatment
designed at a low injection rate. Matrix acidizing provides the best chance that only the
perforated interval is treated, but would require a whopping 180,000 gal of HCl to
dissolve a 5-ft wellbore radius. In contrast, fracture acidizing at 12 BPM with a viscous
preflush and 30,000 gal of HCl would create an etched fracture length of approximately
200 ft. The created flow capacity is more or less equivalent to a 100-ft effective
wellbore radius.

Acidizing Considerations

Only a few special considerations need discussion in most cases of carbonate formation
acidizing. To maximize production, acid types and additives need to be carefully
selected. From the standpoint of reaction with formation minerals, however, there is
little chance of doing more harm than good to production. Table 1 provides some
useful guidelines for acid selection.

When HCl is injected into a carbonate reservoir under matrix conditions, acid
preferentially flows into the areas of highest permeability. Acid reaction in the high-
permeability regions causes the development of large, highly conductive flow channels
called wormholes ( Figure 2 , carbonate core with large wormholes). High reaction
rates, as observed between all concentrations of HCl and carbonates, tend to favor
wormhole creation.

Figure 2

494
Wormhole length normally is controlled by the fluid-loss rate from the wormhole to the
formation matrix. It can be substantially increased by reducing the fluid loss rate from
the worm-hole to the formation with fluid-loss additives. The careful selection of type
and concentration of additive is essential. For example, too high a concentration of
fluid-loss additive can plug the formation, and too little is not effective.

The most effective fluid-loss additives are solids or acidswellable polymers used in
fracture acidizing. Emulsified acids, because of their high viscosity, often give better
results than plain HCl. On the other hand, retarded acids usually work no better than
HCl in a matrix treatment because they do not control the rate of fluid loss from a
wormhole.

Treatment Design

The design of a matrix acidizing treatment for a carbonate formation consists of


specifying acid type and volume, a maximum injection rate, and fluid pressure that will
not fracture the formation. The same calculation method used to determine maximum
injection rate and pressure in matrix acidizing of sandstone applies here. Selection of
acid type and volume, however, are more arbitrary.

Situation Acid Type

Perforating Fluid 5% Acetic

Damaged Perforations 10% Formic

10% Acetic

15% HCl

Deep Wellbore 15% HCl


Damage

28% HCl

Emulsified HCl

Table 1: Acid use guidelines for carbonate acidizing.

A good general recommendation is to inject 50 to 200 gal/ft of 15% or 28% HCl. Exact
acid volume and strength cannot be predicted without knowing specific near-wellbore
conditions. A good general rule would be to use larger acid volume in high-
temperature wells (or where deep damage is suspected), and high-strength HCl with
an effective fluid-loss additive in zones where acid can be injected at practical rates of
0.25 to 0.5 BPM.

495
FRACTURE ACIDIZING

Fracture Acidizing
Fracture acidizing is a process in which acid is injected at a high enough rate and
pressure to fracture the formation. It usually involves injecting a viscous pad fluid
ahead of the acid to initiate the fracture. Differential rock removal occurs as the acid
reacts chemically with the formation faces. This results in an uneven etching pattern
which increases flow capacity to the wellbore and provides a higher productivity ratio.

Fracture acidizing has application in carbonate reservoirs, both limestone and dolomite.
There are two main reasons for performing this type of stimulation treatment. First and
foremost is low permeability, where natural production is not economical; a conductive
fracture provides much more flow area for produced fluid or gas. The second reason is
to overcome damage that extends deeper than matrix acidizing can effectively reach.

Successful fracture acidizing depends on developing a fracture geometry with adequate


surface area. Ideally, the created fracture should cover the productive interval
vertically, be as long as the well's drainage radius, and be wide enough to aid acid
penetration.

The ideal fracture geometry is seldom, if ever, obtained, because of geological


considerations or mechanical limitations. The vertical fracture height may exceed the
productive interval and extend into a barren interval or into unwanted production. The
conductive fracture length is usually less than the created fracture length because the
fast surface-reaction rate generally spends the acid quickly. The spent acid and the
residual preflush then create additional unetched fracture area.

Fracture Geometry

In deciding whether to place a propped fracture or an acid fracture in a a carbonate


formulation, we must be guided by the treatment costs and their expected post-
treatment performances. For both types of treatments, there are optimal fracture
lengths based on the Net Present Value (NPV) of the fracture treatment (Economides
et. al., 1994). In general, acid treatments results in shorter fractures than propped
treatments, and thus would more likely be appropriate for higher-permeability
formations.

Theoretically, a vertical fracture extends an equal distance from the wellbore in


opposite directions ( Figure 1 , Fracture geometry).

496
Figure 1

Fracture geometry is defined as fracture length in one wing of the fracture times the
height and width of that wing. Factors that affect fracture geometry are:

volume of injected fluid


fluid loss
viscosity of injected fluids
injection rate
rock properties
formation fluid properties
The first four of these factors can be controlled. Volume is restricted only by economic factors;
injection rate can be in-creased by using friction reducers, more hydraulic horsepower, larger
tubular goods, or a combination of these. More information about these controllable variables
follows.

Fluid Loss: It is much more difficult to control this critical factor with acid solutions
than nonacid solutions, because the acid reacts with the formation to continually
enlarge pore sizes. Fluid loss not only affects fracture geometry, it also influences acid
penetration distance through fluid leakoff velocity. Table 1 (below) presents data
illustrating the importance of fluid loss control. In this example, the formation had an
average permeability of 1 md and 15% porosity. In each case, the only variable was

497
the fluid loss properties of the acidizing fluid.

Treating* Ceff H W** L ft. Inj % Fluid. eff.


Fluid ft/min BPM Rate
1/2

No FLA .005 75 .036 136 15 4.6

50 lb FLA .001 75 .082 501 15 38.4


/1000 gal

100 lb FLA .0001 75 .120 811 15 91.0


/1000 gal

* 10,000 gal 28% HCl

** Average Width
Table 1: Effect of fluid loss on fracture geometry.

Viscosity: Viscosity affects the reaction rate of acid on carbonate rock because it
influences the treatment fluid's capacity for turbulence, and thus the rate of contact
between the hydrogen ions and the rock surface. The higher the viscosity of the acid,
the lower its capacity for turbulent flow, and the lower the reaction rate.

Viscosity may also be the most important factor affecting the penetrating distance of
live acid in carbonate rock. High viscosity means low turbulence, which in turn means
low acid-spending rate and deeper penetration. This phenomenon is shown graphically
in Figure 2 (Acid penetration distance vs.

498
Figure 2

acid viscosity ). (The term used to describe viscosity in this figure is the "stoke," which
is equivalent to (centipoise/100) x density.) Doubling the viscosity provides a
significant increase in the penetration distance of live acid.

Another phenomenon not shown in this curve is the relation of increased viscosity to
fracture width. As viscosity increases, the fracture width at a given pump rate also
increases ( Table 2 , below). Although larger volumes of acid are regard for this
purpose, even greater penetration distance can be achieved.

Viscosity Fracture cp
Width in.

1 0.038

10 0.068

100 0.120

1000 0.214

499
Injection Rate = 10 BPM
Fracture Length = 100 ft
Young's Modulus = 10 x 106psi

Table 2: Effect of viscosity on fracture width.

Injection Rate: Generally speaking, the higher the injection rate, the greater the fluid
turbulence, and the greater the rate of contact between the hydrogen ions and the
rock surface in other words, the greater the acid spending rate. Spending rate is
governed by these factors in combination with hydrogen ion concentration. Increased
hydrogen ion concentration and turbulence work together to increase acid spending
rate and decrease penetration distance.

Increased injection rate can increase penetration distance under certain circumstances.
All other things remaining constant, the faster acid flows down a fracture, the farther it
will travel before it spends (unless it goes into turbulent flow). This is because contact
time between the acid and the formation is insufficient to allow complete reaction
between the acid and the formation.

Figure 3 (Acid penetration distance vs.

Figure 3

500
pump rate) shows both phenomena occurring. The curve for limestone shows a
decrease in penetration distance as the acid goes into turbulence. Then an increase
can be seen as the acid travels farther in a given period of time. On the dolomite
curve, the distance traveled is more important than the mixing rate because the
surface reaction rate is slower for dolomite than for limestone.

Higher injection rates also increase the fracture width, provide greater cooling effects,
and improve fluid efficiency. These effects are not shown on this curve, but all tend to
increase penetration distance.

Formation Characteristics

Other important considerations in fracture acidizing are the following formation


characteristics: rock composition, permeability, and temperature.

Rock Composition: The physical and chemical composition of the rock influences the
reaction of acid; for example, dolomite is generally slower to react with HCl than
limestone.

Carbonate formations are often mixtures of dolomite, limestone, and acid-insoluble


materials. The more insoluble material a formation contains, the slower its reaction
with acid. If the acid-soluble portion is lining fractures in an otherwise insoluble
formation, excellent stimulation may be obtained even though total solubility is low.

Although carbonates seldom contain large quantities of clays, occasionally they can
contain enough to deter the fracture acidizing treatment. If clays are present,
potassium chloride can be added to the preflush and afterflush solutions. If iron
compounds are present, iron sequestering agents added to the acid will inhibit the
secondary precipitation of iron compounds when the acid spends.

Permeability: The matrix permeability dictates both the fluid properties and the
amount of fracture conductivity needed to obtain the desired production rate. High
permeability decreases fluid efficiency.

Temperature : Under normal conditions, acid reacts more quickly with carbonate
materials as temperature increases. At extremely high temperatures, however, fluid
viscosity decreases, which decreases fracture penetration significantly. Figure 1
contains data on the relationship between acid penetration distance and temperature.

501
Figure 1

Treatment Materials and Procedures

Laboratory test data may also be important in the design of a fracture acidizing
treatment. These data may be gathered from fluid-loss tests, acid reaction rate
measurement, and acid etching tests.

Fluid-Loss Tests: Fluid-loss tests should be run on cores to determine how much fluid-
loss additive is needed. The fluid-loss coefficients are necessary to calculate created
fracture geometry in computer designs.

Acid Reaction Rate Measurement: An acid's surface reaction rate can be determined on
a formation core with a rotating core disc in a reaction cell to determine the
penetration distance of live acid in a fracture. If no cores are available, generalized
data obtained from typical limestone and dolomite formations can be substituted.

Acid Etching Tests: These very important laboratory tests help determine if fracture
conductivity can be created by acid solutions, and which acids etch best. They also
measure conductivity and the injection time required to obtain adequate conductivity.
To calculate the productivity increase resulting from a treatment, first determine the
fracture conductivity.

502
Types of Fracture Acidizing Treatments

There are several basic concepts to remember when considering types of acidizing
treatments:

Any pumped fluid follows the easiest route.

A hole cannot be dissolved without leakoff. For example, 1000 gal of 15% HCl
dissolves 10.8 cu ft (80 gal) of limestone but occupies 135 cu ft.

The higher the permeability and the solubility, the greater the leakoff.

Low-viscosity acids leak off more than high-viscosity acids.

If the leakoff is less than the injected amount, a fracture will be created.

Acid flows down a fracture easier than into the adjoining matrix.

Acid spends faster in a small hole than in a big hole, or faster in a narrow
fracture than in a wide fracture.

Low-viscosity acid spends faster than high-viscosity acid.

A 90% soluble rock is 10% insoluble, which will produce approximately 200 lb
of fines/1000 gal 15% HCl.

A number of fracture acidizing treatments types are available, each with advantages and
disadvantages. Some of the common treatments are described below: conventional treatments,
foamed acids, combination treatments, and preflushes.

Conventional Treatments: Conventional treatments run with low-viscosity acid are


easy to perform. They tend to obtain high fracture flow capacity because of the narrow
fracture and turbulent flow that develop. These treatments require less acid than most
others and the acid may be chemically retarded.

These advantages are balanced by comparable disadvantages. The acid reacts rapidly
in narrow fractures, but spends so quickly that deep penetration of the live acid is
difficult. Low-viscosity acid does little to aid in fluid-loss control, and its fines-carrying
capacity is limited.

Conventional fracture acidizing with high-viscosity acid overcomes most of the


problems of low-viscosity acid, but has other disadvantages.

Advantages include easy mixing and pumping, good fluid-loss control, and better fines-
carrying capacity. High-viscosity fracture acidizing treatments may use either gelled or
emulsified acids. Gelled acids are more effective in friction reduction. A high-viscosity
treatment is a retarded fracturing fluid system which develops a wider fracture, helps
establish laminar flow within the fracture, and has a physical surface retarding effect.

503
Disadvantages of high-viscosity acid treatments are their high cost and possible
problems of emulsion and lower fracture flow.

Foamed Acid: In addition to emulsified or gelled acids, foamed acid may be used in
fracture acidizing treatments. Foamed acid has an erosive or scouring effect on the
face of the fracture which helps develop good fracture flow capacity close to the
wellbore. The fluids are easy to recover, even in low-pressure wells. The foam is good
for removing fines and controlling fluid loss when used in low-permeability formations.

In spite of these advantages, foamed acid spends too fast to achieve deep penetration.
More pressure is required at the surface to pump it. Foamed acid has considerable
fluid-loss control capabilities, but it is more expensive than fluid-loss additives.

Combination Treatments: It is possible to run variations and combinations of the


previously described types of treatments. An alternating-phase method called the
"alpha" technique uses phases of acid (to improve fluid flow capacity) alternated with
nonacid phases that contain fluid-loss additives. These non-acid phases also act as acid
extenders, thus reducing the volume of acid required. The phases are normally
alternated in 10- to 60-minute time frames, depending on specific situations.

The density-controlled acidizing technique uses the overriding and underriding


characteristics of fluids with different densities, which are pumped in sequence into
vertical fractures. This system uses less acid than others, and keeps acid out of
unwanted zones because of viscosity differences.

When standard fracture acidizing systems do not work, or cannot be run, systems
called extended matrix acidizing and closed-fracture acidizing may be employed.
Although calculations cannot be made regarding their acid penetration into wormholes
or natural fractures, extended matrix acidizing does allow the acid, at less than
fracturing pressure, to enter natural fractures and the area of highest permeability and
solubility.

In closed-fracture acidizing, the "easy way out" routes for the acid are created
fractures that have been allowed to close, and previously etched fractures that are
closed.

When matrix acidizing is used, the acid generally goes out the area of highest
permeability because there is less resistance to flow. It also goes out the areas of
highest solubility because there are less fines released to act as fluid-loss control
additives. It goes out natural fractures which have high solubility because they channel
more readily.

Preflushes: Closed fracture acidizing can use a preflush to create a fracture rather than
relying on a natural fracture system. The procedure is to create the fracture system by
pumping a high-rate preflush, and then allow the fracture to close. Then, acid is
pumped at below fracturing pressure through perforations into the closed fracture.

A previously etched closed fracture treatment is identical to the above, except the
initial fracture is made with acid rather than with a preflush.

504
When preflushes are pumped before the treating phase, the treatment becomes more
complex, more acid is required, and the cost goes up. However, the returns from this
type of treatment are better fluid-loss control, and deeper penetration of live acid. An
excellent acid-retarding system is introduced. Also, uneven etching is achieved
mechanically, and unetched areas are able to support over-etched or acid-softened
formation faces.

A preflush may be any fluid compatible with the treating phase that serves these
intended purposes:

creating the main fracture system prior to the acid phase

controlling fluid loss ahead of the acid phase

cooling the fracture face for longer acid reaction

controlling density for optimum acid placement

A high-viscosity preflush can also


create a wider fracture, for an acid-retarding effect

allow the acid to channel, for a fingering effect

control natural fracture leakoff more effectively

Overflush fluids are those generally pumped after the treating phase. These can be any compatible
fluid that
forces the acid out farther into the formation

carries an energized fluid, such as nitrogen or CO2, for easier fluid recovery

aids in the recovery of fines, if the fluid has sufficient viscosity

Treatment Design

The usual fracture acidizing design uses a computer to calculate the volumes of fluid
necessary to obtain precise fracture geometry and productivity increases. The following
discussion, which relates to fracture acidizing computer models, describes outputs,
fracture conductivity, and inputs.

Outputs A fracture acidizing treatment creates conductivity by using acid to etch the
walls of the created fracture. Thus a treatment design program must keep track of acid
spending-how the acid reacts with the fracture walls by providing several pieces of
data:

penetration distances

amount of viscous fingering

505
viscosity of stages

acid velocities

temperatures

acid concentrations

etched fracture length

Fracture Conductivity A fracture acidizing treatment must create a conductive fracture that
increases well production. Fracture conductivity results directly from the etching process. The
conductivity at a given point in the fracture is a function of the reactivity of the acid passing that
point and the contact time between the acid and the fracture at that point.

Intuitively, it would seem that the longer the acid has contacted a portion of the
fracture, the more material it will have removed and the greater the conductivity will
be. If so, the greatest conductivity should be next to the wellbore where acid has
contacted the formation the longest. This may be far from true, since fluid temperature
is usually lowest as it enters the fracture and, if cool enough, the acid reacts very
slowly and does very little etching there. As it penetrates farther into the fracture the
acid becomes warm, increasing in effectiveness while it still has sufficient strength. The
net result may be low fracture conductivity next to the wellbore and near the etched
length, with possibly much higher conductivity somewhere in-between.

Any production increase calculations made in conjunction with a fracture acidizing


program must consider this great conductivity variation along the fracture length. A
single steady-state productivity index ratio cannot be effectively calculated.

Calculating production using a type curve simulator is difficult. Numerical simulators,


however, readily lend themselves to this type of calculation.

Inputs: Fracture acidizing design programs require the input of formation properties

fluid properties

an estimated fracture height

the pay zone location or height

They also require information about whether the treatment:


is conventional

uses the viscous fingering technique

uses foamed acid

uses any other special technique

506
Along with the treatment type, the acid type must also be selected and the acid concentration
entered.

Acid type and concentration alone do not set the reaction kinetics. The other reaction
component must be considered: the rock. If laboratory-generated reaction-rate data
are available for the specific formation, these are usually entered into the computer
program. If not, the program can empirically calculate reaction rate constants from
rock-composition (dolomite or limestone) data. The program also requires data about

well configuration

injection rate

proposed treatment designs

Treating Techniques

After selecting the type of acid and additives and using a computer design to help
choose the quantity of preflush, viscous preflush, acid, and overflush, the number of
stages must be determined. An estimate must be made of the expected vertical
fracture height, given the injection rate, viscosity, rock properties, and changes in rock
composition above and below the perforated interval.

Consider the apparent viscosity of the fluids in the treatment design. If the perforated
interval appreciably exceeds the expected vertical fracture height, then a multiple-
stage treatment is necessary. Of course, injection surveys, such as temperatures
surveys and radioactive surveys, may be useful in obtaining accurate fracture height
information after a treatment. This data can then be used for planning similar fracture
acidizing treatments in the future. Mechanical diversion or use of diverting materials
may be required to obtain the most effective zonal coverage.

507
Diverting Materials In Acidizing
To obtain maximum stimulation from a matrix or fracturing acidizing treatment, it is
necessary to treat the total productive interval. When several zones are open to the
wellbore, or when the section to be treated is massive, it is often necessary to divide
the treatment into stages. A technique that forces each stage to go into a different
zone ensures treatment of the total productive interval.

In the case of multiple intervals, the formations may differ from one another in
reservoir pressure as well as permeability.Flow often can be effectively diverted using
downhole equipment such as packers and bridge plugs, and sliding sleeves between
packers. However, since use of a workover rig to move downhole equipment can
significantly increase total cost, techniques for separting the treatment into stages
without using packers are often preferred These techniques, which use a diverting
material, temporarily plug the interval just treated, thereby diverting the next acid
stage to an untreated zone.

Diverting can be achieved completions like openhole zones, gravel-packed zones, and
perforations by varying the diverting material size distribution.

Solid Type
The most widely used method of placing treating fluids employs diverting agents
composed of solid particles. The particles create a restriction in and across the
formation sections that accept fluid most readily, forcing the breakdown of other
exposed intervals. Ideal solid-type diverting agent properties

have a wide particle-size distribution


coarse: 0.25 to 0.002 in.
fine: 0.02 to 0.0001 in.

are soluble in both water and oil

have a slow rate of solubility, so the agent remains solid during the entire
stimulation treatment

dissolve in a reasonable time, so the well can be returned to production or


injection without unnecessary delay

cause no damage to formation permeability

are suitable for use with both perforations and in openhole

block effectively across fractures and perforations at high differential


pressures

remain stable at high bottomhole temperatures

508
are compatible with all well and treating fluids, including acid

More information about the physical and chemical properties of solid-type diverting materials
follows.

Physical Properties: The melting point of the solids used in diverting should be
sufficiently higher than the formation temperature so that the solid does not melt or
soften on the formation face during the treatment. However, some solids depend on
melting with reservoir temperature to be removed. The most important property of the
diverting solids is their particle-size distribution. Diverting in all types of completions,
except gravel-packed zones, requires a uniform particle-size distribution from 40-mesh
to smaller than 325-mesh. This distribution provides sufficient variation in particle size
to treat a wide range of varying permeabilities.

To divert through a gravel pack, the solid diverting material must be fine enough to
penetrate the gravel pack and filter out on the formation face. Effectiveness of acid
distribution with and without diverting agent is demonstrated by comparing Figure 1
(distribution of acid without diverting agent )and Figure 2 (distribution of acid with
diverting agent).

Figure 1

509
Figure 2

If the diverter is not fine enough to penetrate the gravel pack, the material will bridge
on the gravel. This diverts the fluids to a different portion of the gravel pack. However,
fluid can then move vertically in the gravel pack so that all the acid is still flowing into
the same highest-permeability zones. The low-permeability zones may remain
untreated.

Chemical Properties: The diverting solids should have limited solubility in acid
solutions. To aid in their removal after the acidizing treatment, however, the solids
should be soluble in natural formation fluids, such as oil, gas, and connate water.

A material with limited solubility in water is useful for diverting acid in water wells or
injection wells. In these situations, a sufficiently high volume of water flows through
the diverted zone to eventually remove all the solid diverting material. These solid
diverting agents should be compatible with surfactants, corrosion inhibitors, and
mutual solvent additives in the acidizing fluids.

Gel Type

510
Gel-type diverting agents are available in both temporary and permanent forms.

Temporary Gels: Temporary gels are used during acidizing or fracturing treatments
that call for diverting or blocking action. They are commonly used in previously
acidized or fractured wells, with side permeability differences, that benefit more from
multiple than single extended fractures or flow channels.

Temporary gels are either oil- or water-base mixtures. They are usually pumped into
the formation as relatively thin fluids which set up quickly to form a semi-solid or thick
gel that prevents additional fluid from entering the zone. After a predetermined time
the gel breaks, reverting to a water-thin liquid that is easily produced from the well.

Diversion of a pregravel-pack acid treatment can be accomplished with 5 bbl batches


of gel led 3% NH4Cl plus 100 lb. of correctly sized gravel-pack sand. The combination
of viscosity and particles diverts treating fluid to other perforations.

Permanent Gels: Permanent blocking gels are especially useful for blocking within the
formation. On the surface they are thin, true solutions of low viscosity that can
penetrate any water-permeable formation. Once in place within the formation, they
solidify into a rigid, permanent gel that does not break.

Permanent gels are designed to

alter the injection profile of water-injection wells

shut off water zone encountered in air drilled wells

plug lost circulation zones

Ball Sealers
Ball sealers are used to temporarily seal perforations during acidizing treatments.
When seating against perforations, ball sealers direct treating fluid selectively so that
each portion of the perforated interval is treated. Two types of currently available ball
sealers are the standard weighted ball sealer and the buoyant ball sealer.

Standard Weighted Ball Sealer This type of ball sealer is put into the treating fluid
stream at regular or predetermined intervals, or is dropped in batches. Generally, at
least one ball preferably more is injected for every perforation.

Ball diameters vary from 5/8 to 1-1/4 in., and specific gravities range from 0.9 to 1.8.
Its size is proportional to that of the perforation to be sealed. Obviously, during
fracture acidizing, balls are not added until the formation has been broken down and is
accepting fluid.

When fluid injection is stopped, or upon completion of the treatment, the balls are no
longer held against the perforations and they fall to bottom. Fluid velocity and
differential pressure hold the ball sealer in position during the treatment. If the
perforation density is too great or the balls are oddly shaped, or if communication

511
exists behind the pipe, this type of ball sealer falls to bottom. Figure 1 demonstrates
the seating efficiency of ball sealers in a test case.

Figure 1

Characteristics of ball sealers are shown in Table 1(below). The type to use in a
specific application is determined by the carrying fluid, bottomhole temperature, and
maximum perforation diameter. To ensure a total "ball-off," use 200% excess over the
number of perforations.

OD Size- Specific Core Core Maximum Perforation Diameter-in.


in. Gravity Diameter- Material
in.

5/8 1.2 None Solid 0.38


Rubber

5/8 1.3 1/2 Nylon 0.38

512
3/4 1.3 5/8 Nylon 0.52

7/8 0.9 3/4 Phenolic 0.52

7/8 1.1 3/4 Nylon 0.52

7/8 1.3 3/4 Nylon 0.52

7/8 1.4 3/4 Phenolic 0.52

7/8 1.8-2.0 3/4 Aluminum 0.52

15/16 1.2 None Solid 0.52


Rubber

1 1.1 7/8 Nylon 0.63

1 1.3 7/8 Nylon 0.63

1-1/4 1.2 1.06 Nylon 0.87

1-1/4 1.3 1.06 Nylon 0.87

Table 1: Standard weighted ball sealers with rubber exterior

Buoyant Ball Sealer: This type of ball sealer, developed by Exxon Production Research,
improves the reliability of diverting techniques by properly controlling the density
between the ball sealer and the treating fluid.

Balls having a density less than the treating fluid can achieve 100% seating efficiency
on perforations that are accepting fluid, provided that they are transported to the
perforated interval. This efficiency can be maintained at very low injection rates
through individual perforations.

Buoyant ball sealers dramatically improve the ability to divert well-treating fluids, not
only in fracturing-type treatments but in matrix rate treatments as well. Using buoyant
ball sealer concepts, wells having over 1000 perforations through 250 ft of interval
have been stimulated effectively throughout their entire area. Matrix treatments have
been successfully diverted where conventional ball sealer practices would not have
been attempted.

Buoyant ball sealers are characterized by having a syntactic foam core (containing
hollow spheres bound together with resin) and a polyurethane cover. They are
generally available in the same sizes as standard weighted ball sealers. Buoyant ball
sealers have more uniform density and higher compressibility. A 50% excess over the
number of perforations is recommended.

Selection of Diverting Materials


Diverting materials are available for use in varied well conditions and over a wide
temperature range. Selection of the best agent and concentration is based on the

513
formation fluids contacted, interval length, formation temperature, physical properties
of the formation, type of well completion, and placement. Guidelines for the use of the
more common temporary diverting agents are listed as follows:

1. Graded Rock Salt h

-Melting Point: 1,472 oF

- Soluble in: water or dilute acid

-Specific Gravity: 2.164

-Type Well: Oil, Gas, Injection, Disposal

-Type Diversion: On perforations

On slots

In perforations

On formation

Concentrations to Divert:

-On Perforations (3/8" perforations) 16


lb at 1 lb/gal, 9 lb at 1 1/2 lb/gal

-In Perforations (per perf.): 1/2 to 2 lb

-Open Hole (lb/ft2 of Formation): 5 lb

2. Oil Soluble Resin

-Melting Point - 328 oF

- Soluble in: oil or distillate

-Specific Gravity: 1.062

-Type Well: Oil, Gasb

-Type Diversion: On slots

In perforations

On formation

514
Concentrations to
Divert:

-On Perforations (3/8"


perforations):2 lb at 2 lb/galf

-In Perforations (per perf.): 1/8 to


1lb

-Open Hole (lb/ft2 of Formation): 2.5


lb

3. Oil Soluble Resin in Salt Solution

-Melting Point - 328 oF

- Soluble in: oil or distillate

-Specific Gravity: 1.062

-Type Well: Oil, Gasb

-Type Diversion: Through gravel


pack

On formation

In perforations

Concentrations to
Divert:

-In Perforations (per perf.): 1/2 to


2.5 gal fluid containing 0.5%

-Open Hole (lb/ft2 of Formation): 5 to


20 lb./1000 gal fluid

4. Karaya gum

-Melting Point - None

- Soluble in: water

-Specific Gravity: 1.062

-Type Well: Injection, Disposal

515
-Type Diversion: On formation

In perforations

Concentrations to
Divert:

-In Perforations (per perf.): 1/4 to 1


lb

-Open Hole (lb/ft2 of Formation): 50


to 200 lb/1000 gal fluid

5. Graded Benzoic Acid

-Melting Point - 252oF

- Soluble in: Oil, water, acid ,gas

-Specific Gravity: 1.316

-Type Well: Oil, Gasi, Injection,


Disposal

-Type Diversion: On perforations

On Slots

In perforations

On formation

Concentrations to
Divert:

-On Perforations (3/8" perforations)

9 lb at 1/2 lb/gal, and 2 lb at 1 lb/galf

-In Perforations (per perf.): 1/4 to 1


lb

-Open Hole (lb/ft2 of Formation): 2.5


lb

6. Benzoic Acid in Alcohol Solution

-Melting Point - 252oF

516
- Soluble in: Oil, water, gas , acid

-Specific Gravity: 1.316

-Type Well: Oil, Gas, Injection,


Disposal

-Type Diversion: Through gravel


pack

On formationd

In perforations

Concentrations to
Divert:

-In Perforations (per perf.): 4%7

-Open Hole (lb/ft2 of Formation): 4%7

7. Unibead* OS-160 Buttons-


OS-160 Wide Range

-Melting Point - 192 oF

- Soluble in:Oil

-Specific Gravity: 0.92

-Type Well: Oil, Gasc

-Type Diversion: On
perforations

On slotse

Concentrations to Divert:

-In Perforations (per perf.): 1/4


to 1 lb

-Open Hole (lb/ft2 of Formation):


2.5 lb

8. Graded Napthalene

517
-Melting Point - 176oF

- Soluble in: Oil or gas

-Specific Gravity: 1.145

-Type Well: Oil, Gasi

-Type Diversion: On
perforations

On slots

In perforations

On formation

Concentrations to Divert:

-On Perforations (3/8"


perforations)

8 lb at 1/2 lb/gal or 2 lb at 2
lb/galf

-In Perforations (per perf.): 1/4


to 1 lb

-Open Hole (lb/ft2 of


Formation): 2.5 lb

9. Unibead* OS-130 Buttons-


OS-130 Wide Range

-Melting Point - 155 oF

- Soluble in: water or dilute acid

-Specific Gravity: 0.92

-Type Well: Oil, Gasc

-Type Diversion: On
perforations

On slotse

In perforations

On formation

518
Concentrations to Divert:

-In Perforations (per perf.): 1/4


to 1lb

-Open Hole (lb/ft2 of Formation):


2.5 lb

10. Unibead* OS-90 Buttons-


OS-90 Wide Range

-Melting Point - 138 oF

- Soluble in: oil

-Specific Gravity: 0.92

-Type Well: Oil, Gasc

-Type Diversion: On
perforationse

On slots

In perforations

On formation

Concentrations to Divert:

-In Perforations (per perf.): 1/4


to 1 lb

-Open Hole (lb/ft2 of Formation):


2.5 lb

* wax polymer

a. If oil or gas is not making any water, a cleanup operation with dilute acid or brine may be
required.

b. If gas well is not making distillate, oil soluble resins should not be used.

c. Unibead materials may be used in gas wells if temperature of zone exceeds


the melting point of Unibead material, but zone must be cooled by preflush or
the first stage of treating fluid to a temperature lower than the melting point.
When the zone regains its original temperature, Unibead material melts and is
removed.

519
d. Oil soluble resin in salt solution and benzoic acid in alcohol solution will pass
through a sand pack and bridges on the formation. They will also bridge in
perforation.

e. Unibead material will bridge on a 3/8 in. perforation, but the bridge will be
weak.

f. The lb represents the amount of diverting agent required to bridge on a


perforation (open behind perforation) when used at the concentration listed (lb
agent/gal of carrier fluid).

g. Each perforation or sq. ft. of formation will take approximately 25 gallons of


treating fluid containing 4% graded benzoid acid in alcohol solution before flow
will be decreased to a minimum. The treating fluid must contain enough
additional material to saturate it at bottom hole treating temperature.

h. Material should not be considered for HF acidizing operations. Other bridging


agents are compatible.

i. Subliming materials. Should receive special consideration for gas well


applications.

520
WELLBORE CLEANOUT & SCALE REMOVAL

Wellbore Cleanout
Wellbore cleanout is generally a soaking and agitation treatment designed to open up
ferforations and formation damage occurring in perforations or in the immediate
wellbore vicinity. This technique allows the treating solution to dissolve sludge and
acid-soluble precipitants (scale) and recover mud filtrate, formation fines, and other
materials that might restrict production. Removing these contaminants permits the
well to produce at its natural potential.

Special attention is required for wellbore cleanouton the increasingly rare occasions
when oil-external emulsion muds and true oil-base muds are used. Emulsions used for
drilling are usually fairly low-solids, oil-external, oil-wet, clay-stabilized emulsions. Oil-
base muds are usually prepared by gelling a viscous, highly asphaltic oil and adding to
it blown asphalt, fine calcium carbonate, and other materials that impart additional
fluid-loss properties. An approach used to successfully remove formation damage
caused by oil-base muds is discussed later.

A well is cleaned out prior to production testing (or before it is placed on initial
production) or whenever a producing interval begins to "clog up." This type treatment
is designed only to remove near-wellbore contaminants. The treating solution is not
intended to penetrate the formation beyond the immediate surface of the perforations
or face of the productive interval.

Treating Solutions

Prior to the selection of a treating solution, formation fluid samples should be taken to
determine the presence of chlorides, sulfides, sulfates, and pH. If a sample of scale
and/or sludge can also be obtained, a more thorough laboratory analysis can be
performed.

Many types of treating solutions may be used for wellbore cleanout. The constituents
are based on type of damage, formation type, general formation conditions, and the
treating technique.

Pure Solvents: Xylene or toluene alone is most effective in removing paraffinic and
asphaltic sludges and oil-base mud. However, care must be taken in the selection of a
hydrocarbon solvent when the reservoir is dry gas. Solvents such as kerosene and
diesel may cause relative permeability effects which can seriously hinder production.
For these applications, aromatic solvents, such as xylene and toluene, are preferred.
They are more easily removed from the formation than are straight-chain solvents.

Mutual Solvent Systems In acidizing, mutual solvents are used in the postflush to:

reduce surface and interfacial tensions to promote cleanup and fluid load
recovery

strip oil and sludge layers from acid-soluble materials

521
prevent emulsions from forming

break existing emulsions

assist in preventing or removing water blocks

Commonly available mutual solvents are

Amoco A-Sol: a blend of water-soluble and oil-soluble alcohols

EGMBE: ethylene glycol monobutyl ether

Amoco A-Sol A-28: a blend of alcohols

Amoco Super A-Sol: a blend of alcohols and xylene

A variety of proprietary service company mutual solvents (e.g., MS-10, MS-


12, MS15 (BJ Services), MS-200 (Nowsco), Musol A (Halliburton), MAS (Serfco),
and WSA-2, EZ-Sol and MAS (Western)).

All these mutual solvent/acid systems can break through a thin sludge coating and
allow acid to contact the surface of the solid covered by sludge. However, mutual
solvent/acid systems do not disperse sludge.

To remove thick sludge, an aromatic solvent soak must usually precede a mutual
solvent/acid treatment. A two-hour xylene or toluene soak can soften sludge layers
and allow mutual solvent/ acid penetration. Plain acid following an aromatic solvent
soak generally has no positive effect.

Oil-base muds present a special problem in acid treatment. Many of these muds
contain emulsifiers that can develop very stable emulsions with acid or spent acid. In
addition, mud solids and drill cuttings provide the fines to stabilize the emulsions. An
aromatic solvent soak followed by mutual solvent and acid usually disperses it
completely. The ideal mutual solvents for this operation are Super A-Sol, A-Sol, or
EGMBE, and the acid should be weak intensified acid (1/2% HF), if possible. The
treatment area is usually confined to the wellbore and open natural fractures, since the
fines, which are necessary for the emulsion, do not penetrate into the formation.

To avoid problems in completing a well drilled with an oil-base mud, circulate the well
with solids-free mud and then flush with xylene. The xylene dilutes the mud and
causes any remaining solids to drop out of suspension. These solids can be circulated
out or removed by an appropriate mutual solvent and acid.

Acid-External Emulsified Acid This acid system is usually composed of 10% xylene,
with a surfactant which disperses the aromatic solvent in the acid, and 90% acid phase
that is usually 15% HCl. Acid nonemulsifiers are not used because they may make the
acid dispersion unstable. Moreover, performance experience in the field shows that
nonemulsifiers have little effect even when injected into oil-producing intervals.

522
Stability of the acid-external emulsified acid is influenced by three factors: the degree
of agitation during preparation, concentration of the xylene phase, and strength of the
acid phase. Increasing the xy1ene or acid strength decreases stability. Should
separation occur, the aromatic solvent can be readily redispersed by circulating.

5 to 15% HCl This solution, with a high concentration of anionic surfactant (1 to 3% by


volume), is widely used for well-bore cleanout. The high surfactant concentration
reduces surface tension and makes the acid highly adaptable to both carbonate and
sandstone formation as a first-stage wellbore treatment for water-base mud removal.
Even when diluted with 20 volumes of brine, surface tension properties remain very
low.

Another characteristic of this solution is that clay solids, which normally have a
tendency to cluster when contacted by acid, are effectively dispersed for easier
removal. In addition to shrinking hydrated mud by acid reaction, this mixture softens
and allows the removal of harder and more tenacious filter cakes.

Weak Intensified Acid (1/2% HF) Weak intensified acid, consisting of 1/2% HF, is
useful in unplugging perforations and establishing flow for HCl or HCl-HF acid mixtures.
HF is especially useful for dissolving clays and silica, making it a good choice for
sandstone acidizing. In carbonates, however, HF must be used in extremely low
concentrations to avoid formation of insoluble precipitants such as calcium fluoride.

Treating Procedure

The number of soaking and agitation applications depends on the amount of damage
that has occurred in perforations or in the immediate wellbore area. Treating solutions
designed for suspension, aromatic-solvent acid dispersions, or cleanup types are
normally used in a soaking action. The treating solution is placed, or spotted, across
the perforations or producing interval, and is allowed to set for a fixed period of time.
The treatment load is then recovered.

Agitation can be accomplished by one of three methods. One technique is to spot the
treating solution across the perforation, allow a short soaking period, and then wash it
back through the annulus by reciprocating or rotating the work-string. Another method
is to pressure-up, staying below fracturing pressure, and then release the pressure
very quickly. This action is sometimes referred to as "backsurging the perforations."
The third method is to spot the treating solutions across the perforations, allow it to
soak for awhile, and then swab back through either the casing or the tubing, or jet the
treatment back with nitrogen.

With any of these techniques, the treating solution may have to be applied several
times before the formation is opened for fluid entry. The use of several wellbore
cleanout applications allows for a matrix or fracturing acidizing treatment to be
performed without fear of pushing unwanted plugging material into natural formation
permeability.

Scale Removal

523
Prior to production, well fluids are in a natural undisturbed state. As production occurs,
it creates a chemical disturbance and physical imbalance within the formation. The
environment changes as a result of pressure drop, which allows dissolved gases to
come out of solution.

These changes destroy the state of equilibrium, thereby creating the initial stages of
scale development as the salts reestablish equilibrium under new conditions. Scale can
restrict or block tubing, flowlines, surface equipment, downhole equipment, the
wellbore, and perforations, and thus completely choke off production.

Scale formation is usually a long and dynamic process with more than a single cause.
One direct cause of scale is the mixing of incompatible waters. Other causes include
corrosion leading to iron precipitation, and abrupt drops in pressure or temperature.

Types
Scale can be divided into three broad classes, based on their solubility in water, their
reaction with HCl, or their removability with HCl. The classifications are (1)
watersoluble, (2) acid-soluble, and (3) acid-insoluble. These classifications are
simplified, because pure calcium sulfate or pure calcium carbonate is rarely deposited.
Although scale deposit is usually a mixture of one or more major inorganic components
plus corrosion products, congealed oil, formation fines, paraffin, and other impurities,
these classes are adequate for field analysis. Common scale deposits are listed in
Table 1.

Scale Type Chemical Formula Mineral Name

Water Soluble NaCl Sodium Chloride (salt)

Acid Soluble Deposits

Calcium Carbonate CaCO3 Calcite

Iron Carbonate FeCO3 Siderite

Iron Sulfide FeS Trolite

Iron Oxides Fe3O4 Magnetite

Fe2O3 Hematite

Magnesium Hydroxide Mg(OH)2 Brucite

Acid Insoluble Deposits

Calcium Sulfate CaSO 2H2O Gypsum

Calcium Sulfate CaSO4 Anhydrite

Barium Sulfate BaSO4 Barite

524
Strontium Sulfate SrSO4 Celestite

Barium Strontium BaSr(SO4)2


Sulfate

Table 1: Common scale deposits.

Identification of Scale: One of the most important factors in dealing with scale is
accurate identification of the material being deposited. There are two fundamental
means of identifying scale: laboratory testing and fluid identification.

Laboratory Testing: The primary means of identifying scale in the laboratory are

X-ray diffraction: This method is the fastest laboratory method available and requires the
smallest sample.

chemical analysis: The scale sample is decomposed and dissolved in chemical


solution, and its various compounds are then analyzed by standard techniques
of titration or precipitation.

Field Identification: The following three-step procedure can help make a preliminary field
identification.
1. Add scale to water. If it dissolves, it is probably sodium chloride.

2. Add scale to 15% HCl.

a. If it bubbles vigorously and dissolves, it is probably calcium


carbonate.

b. If it bubbles slowly and turns the acid yellow, it is


probably iron carbonate.

c. If it bubbles and smells like rotten eggs, it is iron


sulfide.

d. If it dissolves slowly without bubbling and turns the acid


yellow, it is probably iron oxide.

e. If it dissolves slowly in acid with no bubbles and does


not color the acid, it may be magnesium hydroxide.

3. If it does not dissolve, it is one of the acid-insoluble scales. If it is gypsum, you


can scrape it or dent it; it also has the monoclinic characteristic of crystal structure.
The crystals are usually fairly large.
Anhydrite is harder than gypsum and the crystals are harder to see. They are both slowly soluble in
hot acid. Barium sulfate and strontium sulfate are very heavy, usually dark, with very fine crystals of
almost any color.

525
Sand or silicates (formation) may often be identified by its grain structure but may
look much like the above. It is lighter in weight than barium or strontium sulfate.

Scale Removal

Acid-soluble scales include the following:

calcium carbonate

iron carbonate

iron sulfide

iron oxides

calcium hydroxide

These scales can be removed by one or more of the following solutions:


15% HCl with no surfactants HCl, containing only corrosion inhibitor to lower the
corrosion rate, is a basic acid (without surfactants). It is not normally used for scale
removal.

15% HCl with 0.15% penetrating agent This acid solution contains 1.5
gal of a penetrating agent per 1000 gal acid. Penetrating agents lower
the surface tension of the acid solution, increasing the ability of acid to
wet and spread. This solution is very effective on scales containing small
amounts of iron. It can be used for all acid-soluble scale when treating
surface, gathering lines, or any other system where the solution does
not enter the formation.

15% HCl with sequestering agents This solution is used to prevent the
precipitation of iron hydroxide. It is effective in handling scale that
contains iron (Fe+2 and Fe+3) compounds. The HCl dissolves the iron
scales. Sequestering agents (see Section 4.3) prevent iron precipitation
by forming a complex with the iron, thus keeping it in solution and
preventing secondary precipitation. With most iron scales, a stronger
acid strength may also be considered.

10% acetic acid The greatest asset of this acid in scale removal is that
it does not damage "chrome," or alloy steel, found in downhole pumps.
Calcium carbonate scales readily react with this acid, as do most other
scales.

15% HCl -aroma tic dispersion This acid dispersion contains an


aromatic solvent, acid, and an emulsifying surfactant.

Acid-insoluble scales include


calcium sulfates

526
barium sulfate

strontium sulfate

barium strontium sulfate

Acid-insoluble calcium sulfate can be removed by chemical methods; barium sulfate and strontium
sulfate cannot.

There are two removal techniques for calcium sulfate (gypsum); one employs
carbonate and hydroxide converter solutions, the other organic acid salts.

Carbonate and hydroxide converter solutions convert gypsum (CaSO4 2H2O) to


calcium carbonate, which is acid soluble. However, these converting properties are
severely retarded when dense laminated gypsum scale is encountered. Because of a
lack of penetration, repeated application may be necessary. In some cases, hydroxides
form a sludge which can be pumped from the well; in others, acid must be used to
remove the reaction precipitant. Although hydroxides have proved fairly effective in
some wells, in others they tend to decrease in effectiveness with re-treatment. This is
a result of a buildup of calcium hydroxides sludge within the interstices of the gypsum
and on the surface, limiting solvent penetration.

With carbonate and hydroxide converter solutions, an acid stage is required to remove
the reaction precipitate. These jobs are usually conducted in four stages, requiring a
considerable amount of time.

Organic acid salts react with gypsum to form a reaction precipitate which tends not to
adhere to the gypsum but to form a water-dispersible sludge. The reaction precipitant
has a tendency to slough away from the gypsum surface, thereby increasing solvent
penetration. Also, since the reaction product does not adhere tightly and is readily
dispersible in water, an acid stage usually is not needed to remove the sludge.
However, should it become necessary to remove the sludge, it is soluble to the extent
of 1.2 lb/gal of 15% HCl.

When barium sulfate or strontium sulfate deposition has been diagnosed, the most
practical solution is to eliminate the cause. For instance, barium sulfate is most often
formed by the mixing of native formation water with injection or produced waters from
another interval one of which is high in barium while the other is high in sulfates
and isolating these waters can prevent the problem. The presence of barium sulfate
usually indicates either a lack of zone isolation or faulty casing.

The removal of barium sulfate or strontium sulfate is an extremely difficult problem.


Available chemicals that remove small amounts of these sulfates are usually inefficient
and expensive. These deposits are best removed mechanically, by jetting or running a
bit and scraper.

Scale Prevention
Several methods of preventing scale formation are available. The two most successful
and widely used methods are (1) the gradual placement of water-soluble

527
polyphosphates (solid scale inhibitors) into the formation and (2) the use of liquid scale
inhibitors.

Solid Scale Inhibitors: Solid polyphosphates are placed into the formation during
water-base fracturing treatments. After placement, the polyphosphates dissolve very
slowly in the water produced by the well. The amount of phosphate dissolved per unit
volume of water produced is small but usually suffices to control scale formation. When
placed properly, it can give long-term protection.

One argument against the use of polyphosphates is the placement method itself, which
is not only relatively expensive but also not always possible. The material's reversion
to inactive calcium orthophosphate (calcium phosphate scale) in the presence of HCl
also prohibits the use of acid treating solutions before and after its placement.

Liquid Scale Inhibitors: These inhibitors, including organic phosphates and synthetic
organic polyelectrolytes, have also been widely used for scale control. Successful scale
control depends on the adsorption of the inhibitor on the rock matrix and slow
desorption into the produced fluids. The adsorption properties of inhibitors vary. Many
scale inhibitors are effective in preventing scale but have poor adsorption properties,
which makes frequent re-treatment necessary.

Figure 1 shows a liquid organic scale inhibitor that effectively prevents scale deposition
in formation waters.

528
Figure 1

Its good adsorption properties provide long-lasting protection.

Placement of liquid scale inhibitors may be accomplished in three ways: (1) by the
matrix squeeze technique, (2) as an additive in stimulation treatments, and (3) by
metering the liquid into water injection systems.

529
Paraffin And Asphaltene
Paraffins are organic deposits caused by changing wellbore conditions that upset the
chemical equilibrium so that various materials in the crude oil precipitate out of
solution. These materials separate and deposit a waxy material called paraffin (in some
cases asphaltene). Figure 1 illustrates the development of paraffin in a flowing well.

Figure 1

Paraffin, usually deposited in the wellbore, extends to the top of the well and into
flowlines from a depth where the formation temperature is approximately the same as
that at which the oil became saturated with the separated wax.

Asphaltene deposits, usually hard and brittle, occur at the bottom of the wellbore
adjacent to the producing formation wall.

The buildup of paraffin or asphaltene can completely plug up the oil flow in a producing
well if the problem is ignored or left untreated.

530
Properties
Paraffins are either normal, branched, or cyclic alkanes with the chemical formula
CnH2n+2. Paraffin deposition is difficult to inhibit because these hydrocarbons are
generally very inert and therefore resistant to attack by acids, bases, and oxidizing
agents.

Crude paraffin deposits are made up of individual alkanes ranging from a carbon chain
length of 20 with a melting point of 98 F to a chain length of 60 with a melting point
of about 215 F. Crude wax or paraffin deposits consist of very small crystals that tend
to agglomerate and form granular wax particles about the size of ordinary table salt
grains. Deposited paraffin may also contain gums, resins, asphaltic material, crude oil,
sand and silt, corrosion products, scale, and often water.

The primary cause of wax separating from crude oil is loss of solubility, often resulting
from changing environmental conditions that disturb solution equilibrium. Factors
affecting this equilibrium include temperature and pressure changes, evaporation, and
loss of dissolved gases. Those paraffins having the highest melting point and molecular
weight are the first to separate from solution. The quantity of lower melting point
waxes with smaller molecular weights to separate from solution depends on how much
the oil's equilibrium changes prior to deposition that is, the solubility of different
paraffin waxes in specific crude oil at a given temperature decreases with an increase
in melting point and molecular weight.

Asphaltenes are black carbonaceous petroleum components characterized by relatively


high molecular weights and insolubility in light paraffinic hydrocarbon solvents, such as
pentane and petroleum ether. The asphaltene fraction of petroleum is made up of
polycyclic, condensed aromatic rings with few side chains. These compounds are
considered polar materials due to the presence of sulfur, nitrogen, oxygen, and
complex metals.

Asphaltenes occur in many crude oils in the form of colloidally suspended solid
particles. Asphaltene precipitation takes place when crude oil loses its ability to
disperse the particles. Asphaltene stability is dependent on crude oil composition and
the nature of reservoir rock surface. The balance that holds them in a stable
suspension is dependent on the same conditions that cause paraffin to separate from
crude.

Pour point and cloud point measurements refer to crude oil's ability to hold paraffin in
solution. Cloud point is the temperature at which paraffin begins to come out of
solution. This point is visible in clear solutions as a slight cloudiness. Since the cloud
point of a dark crude oil is not visible, instrumental methods may be used to obtain
crude oil cloud points.

If crude oil is slowly cooled below the cloud point without agitation, small wax crystals
gradually form an interlocking network which inhibits flow. The pour point is reached
when oil ceases to flow.

531
Factors Affecting Deposition
Temperature: When crude oil cools, it loses its solubility, and can no longer keep the
particles in suspension. Paraffin solidifies, its particles settle and accumulate into waxy
deposits.

Pressure: Pressure holds dissolved gases and volatile constituents in solution and helps
maintain the fluid at formation temperature. Obviously, it is impossible to produce oil
without having a considerable pressure drop; two factors, then, work simultaneously.
As pressure drops, the temperature drops, because gases evolve and expand, and
heavier components settle out, forming paraffin.

Loss of Volatile Constituents: Loss of the lighter constituents of the crude definitely
reduces the quantity of paraffin the oil can hold in solution. The reduction in the
volume of oil results in less solvent available to dissolve the same amount of wax.

Suspended Particulate Material: Evidence shows that as paraffin begins to separate


when temperature drops, formation fines such as sand and silt speed this process.

Co-produced Water: Although water has no direct effect on wax solubility in oil, it may
reduce acute paraffin problems in wells that began producing it in appreciable
amounts. One possible explanation is that steel tubing will water-wet, thus reducing
the tendency for paraffin to deposit.

Conditions Favoring Paraffin Deposition: Even though wax may separate from crude
oil, the produced paraffin does not necessarily deposit on the tubular goods and other
equipment. It seems logical, in fact, for the wax to remain suspended in the crude oil
itself. This ideal situation often exists, and in wells producing a very waxy crude, few
paraffin problems occur. Some factors that cause paraffin accumulation in a well are

alternate coating of the pipe and draining of the oil. When a surface is intermittently
coated with oil, the thin film left on the pipe surface is too thin and its movement too slow to
carry away wax particles.

presence of only an oil film contact with the pipe while the well is flowing.

contact of the oil with an unusually cold surface, such as in oil production
through aquifers; this can cause paraffin crystals to grow directly on the pipe
wall. (It is estimated that heat loss by conduction from a pipe in contact with
water is approximately eight times greater than that which would occur in
contact with air or dry earth.)

rough pipe surfaces, which often provide excellent sites for paraffin deposition.

electrical charges on various materials in the crude oil, which tend to promote
migration of the separated waxes to the pipe walls.

Removal and Control Techniques

532
There are basically three methods for removing and/or controlling paraffin deposition:
mechanical, thermal, and chemical.

Mechanical Methods: Several mechanical techniques remove paraffin deposition in


tubing, flowlines, and pipelines. These include

rod scrapers

wireline scrapers

flowline scrapers

balls, pigs, and/or soluble plugs

free-floating pistons

hollow rods for circulating hot fluids

The major advantage of using mechanical methods is that positive cleaning is assured. There are
some disadvantages, however. Application is rather limited because of the time, expense,
personnel, and special equipment involved. If tools are lost in the hole, a fishing job may be
necessary.

One preventive mechanical method is the use of smooth-surface pipe. The internal
surface of pipe can be made smooth by plastic coating or removal of mill scale.
Smooth pipe surfaces do not prevent paraffin deposition, but in most cases they slow
the process.

There are three widely used plastic coatings for pipe: phenol-formaldehyde, epoxy
phenolic, and polyurethane.

Phenol-formaldehyde has been in use longest, and is the most widely used coating in
oilfield tubulars. It is a highly crosslinked polymer, with excellent resistance to temperature,
chemicals, and infusion by small molecules (H2O, CH4, H2S). Its surface is extremely
smooth and has a high gloss. It is a brittle plastic which becomes extremely rough when
abraded by sand.

Epoxy phenolic is the second most widely used material in oilfield tubulars.
Less resistant to temperature, chemicals, and infusion than phenol-
formaldehyde, its surface is almost as smooth, but it is less brittle and does not
deform as badly when abraded by sand.

Polyurethane is a relative newcomer, and the least crosslinked of the three. It


has lower temperature, chemical, and infusion resistance than the other two,
and its surface is not quite as smooth as phenol-formaldehyde. Polyurethanes
are quite flexible, and although (like the other two materials) its film thickness
is reduced by sand abrasion, the plastic is only slightly deformed and maintains
most of its smoothness.

533
Plastic-coated pipe either works extremely well or is a dismal failure. Numerous examples of both
cases can be found in oil-field operations. Given the descriptions above, it is apparent why
particular plastics may perform well or not, depending on their application and the experience of the
particular oil operator. In areas where the oil stream contains no sand or other abrasive solids,
plastics effectively prevent paraffin buildup. When brittle plastics (such as phenol-formaldehyde) are
used in wells containing abrasive materials, the plastic deforms badly, leaving a rough, abraded
surface which only invites severe paraffin deposits.

There are three important factors regarding the use of plastic coatings in the
prevention/reduction of paraffin deposition:

Paraffinic plastics such as polytetrafluoroethylene, polyethylene, and polypropylene, no


matter how smooth, do not reduce paraffin buildup. In fact, because they are paraffins
themselves, they cause paraffin buildup by either hydrogen bonding or a phenomenon
closely resembling cocrystallization.

Smooth, nonparaffinic plastics reduce paraffin deposits only as long as they


remain smooth. Wirelines, fishing tools, piano wire equipment, and other
workover or measuring devices destroy the effectiveness of the plastic coating
by damaging the smooth surface.

All metal surfaces (nipples, valves, etc.) must be coated with plastic. One
uncoated nipple can effectively choke the well.

Thermal Methods: All thermal techniques use heat to remove the paraffin. Methods in current use
include
bottomhole heaters

hot oil, a common method which melts the waxy accumulation; lease crude oil
is run through heat exchangers and pumped into the well at temperatures
greater than 300 F

steam, which melts the paraffin so that it can be flowed with the oil from the
well

heat-liberating chemicals (such as magnesium bars) followed by HCl; these


produce chemical reactions which release heat, thereby melting the paraffin

These approaches ensure paraffin removal. In order to prevent further deposition, the working
temperature must be maintained above the melting point of the wax.

Thermal techniques have some disadvantages. Bottomhole heaters require special


equipment and consume additional power. Hot fluids pose a danger to personnel,
usually require the use of pump trucks, and sometimes dissolve the paraffin only to
redeposit it deeper than before.

Chemical Methods: Chemical methods for paraffin control are separated into two
distinct categories: solvents and inhibitors.

534
Solvents: The best solvents for paraffin removal are too toxic, hazardous to handle, or
damaging to refinery catalysts.

Two such chemicals are carbon disulfide (CS2) and carbon tetrachloride (CCl4).

Aromatic hydrocarbons, such as xylene and toluene, are acceptable solvents and
dissolve both wax and asphaltenes; they are suitable for crude paraffin deposits.
Solubility of paraffin in xylene is shown in Figure 1 .

Figure 1

Inhibitors: A large number of paraffin inhibitors are also available. Explanations of their
effectiveness vary widely. Some are said to be dispersants; others, wax crystal
modifiers; and still others claim to alter the surface conditions of equipment. Most
paraffin inhibitors also contain paraffin solvent as the carrier.

Surfactants: Of the vast number of paraffin chemicals available, wetting agents and
dispersants constitute the highest percentage. Wetting agents are designed to work much
like plastic pipe: the chemical reacts with produced water or by itself and forms an aqueous

535
film on the pipe surface. This surface film retards paraffin deposition by covering the sites
where paraffin can adhere. However, it is very difficult to get the pipe surfaces water-wet
and keep them that way. The theory behind the dispersant approach is that certain
chemicals tend to cause the crude paraffin molecules to repel each other as well as the
metal surface.

Both wetting agents and dispersants have seen some success. The most
serious drawback to their use is that no supplier can be absolutely certain of
their effectiveness. The operator has to decide whether to continue a particular
paraffin removal program or to run the trial and error method with all available
chemicals and hope to find one that is satisfactory.

Crystal Modifiers: Generally, crystal modifiers have many of the same


drawbacks as surfactants. Their effectiveness appears to be somewhat
unpredictable. With the proper placement technique, however, these chemicals
seem to work fairly well and to show the most promise among chemical
inhibitors. When viewed under a microscope, they appear to modify the wax
crystals precipitating from a wax solution ( Figure 2 , paraffin crystal clustering
and Figure 3 , effect of inhibitor on paraffin crystals).

Figure 2

536
Figure 3

Crystal modification occurs for one of two reasons: (1) the chemical comes out
of solution at a temperature slightly higher than that required for paraffin
precipitation, causing nucleation, or (2) the modifier comes out of solution and
co-crystallizes with the paraffin crystals. In either case, the normal crystal habit
of paraffin is deformed sufficiently to inhibit further growth. This causes the
crude oil to become saturated with much smaller paraffin crystals than normal
with less tendency to adhere to pipe surfaces. Although they do not completely
cure the paraffin problem, crystal modifiers offer the best chemical approach for
reducing its severity.

537
Method Of Diluting Hydrochloric Acid
The degrees Baume (Be) or specific gravity (sp gr) of the concentrated HCl must be
known in order to accurately mix HCl at 15% or any other strength. The first step in
mixing HCl to a given strength is to take the degrees Baume or specific gravity of the
concentrated HCl. These steps in order are

1. Put a sample of concentrated HCl to be tested in a 250-ml graduate cylinder and take the
degrees Baume (or specific gravity) reading with a hydrometer. Also, take the temperature
of the sample.

2. Correct the degrees Baume or specific gravity to the standard temperature of


60 F by using the temperature corrections in Table 1 (below).

3. Convert the corrected degrees Baume or specific gravity to HCl percentage


by using Figure 1 (conversion of degrees Baume or specific gravity to
percentage of HCl.

Figure 1

4. Use the dilution charts in Figure 2 and Figure 3

538
Figure 3

(Acid dilution charts) to determine how much concentrated HCl is required to


mix 1000 gal of HCl to any desired strength. Most strengths are shown on the
chart; if a desired strength is not shown, use Formula 11 given in this text
segment (below).

539
Figure 2

Different acid strengths may remain stratified in the tank and cannot be expected to mix
themselves.

The degrees Baume (or specific gravity) of concentrated HCl should be taken as often
as necessary to ensure that the weight being used to calculate the dilution is as
accurate as possible and that it never is more than 0.50 Be (0.005 sp gr) in error.

[Use this table to correct the degrees Bauma or specific gravity readings of acid to the
standard temperature, 60 F. Choose the B or specific gravity reading nearest the
observed reading obtained from the hydrometer reading.]

B Correction sp gr Correction
(b) factorc (b) factorc
(Be') for each F (sp gr) for each F

2 0.01 1.014 0.0001

540
4 .02 1.028 .0001

6 .02 1.043 .0002

8 .02 1.058 .0002

10 .03 1.074 .0002

12 .03 1.090 .0002

14 .03 1.107 .0003

16 .03 1.124 .0003

18 .04 1.142 .0003

20 .04 1.160 .0003

22 .04 1.179 .0004

24 .04 1.198 .0004

aBaum gravity is a measurement of fluid weight based on specific gravity. The Baum
gravity (B) applies to liquids heavier than water. It is calculated from specific gravity
through the relationship

Table 1 : Temperature corrections for HCl hydrometer readings:

B = 145 -

The Baume' gravity of 15% HCl acid is 10.1.

b Degrees Baume' or specific gravity nearest the observed reading.

c Correction factor for temperature(F): These temperature corrections are based on


information appearing in the International Critical Tables.

Above 60 F: If the observed temperature is above 60 F, add the amount shown


above for every 1 F above 60 F.

Below 60 F: If the observed temperature is below 60 F, subtract the amount shown


above for every 1 F below 60 F.

HCl, % oBaume Weight of Solution,b HCl/gal., lb.


specific lb/gal.
gravity

3 2.11 1.015 8.47 0.25

541
5 3.52 1.025 8.55 0.43

7.5 5.23 1.037 8.65 0.65

10 6.89 1.050 8.76 0.88

15 10.10 1.075 8.97 1.35

15.4 10.36 1.077 8.98 1.38

19.8 13.10 1.099 9.17 1.82

28 18.05 1.142 9.52 2.67

31.45 20.0 1.160 9.67 3.04

35.21 22.0 1.179 9.83 3.46

aAll figures in this column are rounded to the third decimal point.
bAll figures here are rounded to the second decimal point.

Table 2: Physical constants of acid solution strengths.

Calculating Acid Solution Ingredients


Because HCl is the primary base fluid for many acidizing systems, the definition of acid
strength needs to be well understood. Strength, or concentration, of HCl is
conveniently expressed as percentage by weight of the active materials in the acid
solution. For example, a 15% HCl solution means that for each 100 lb of solution, 15 lb
of hydrogen chloride gas is dissolved in 85 lb of water.

Concentration is expressed by using the specific gravity of the liquid or solution. The
concentration is just a measurement of the relative weight of the material in relation to
the weight of an equal volume of pure water. Pure water weighs 8.34 lb/gal at 60 F.

Table 2 lists the physical constants for some of the most frequently used HCl acid
strengths.

The percentages and specific gravities from Table 1 and Table 2 satisfy what is
needed in Equation 11 for determining concentrated HCl volume to transpose acid
strengths. Since equal weights of HCl are involved in both weak and strong acid
solutions, Equation 11 is used most often to determine the amount of ingredients for a
HCl solution.

(vol. of strong acid) x (% strength of strong acid)


x (sp gr of strong acid) = (vol. of weak acid)
x (% strength of weak acid) x.(sp gr of weak acid) (11)
This formula is symbolized as:
Vs x %s x sp grs = Vw x %w x sp gr (12)
Therefore,
Vs = (Vw x %w x sp grw)/(%s x sp grs) (13)

542
For example, if a job specifies 1000 gal of 15% HCl, volume of 200o Be HCl (31.45%) can be
calculated:
Vs = (1000 x 0.15 x l.075)/(0.3l45 x 1.16)
= 161.25/0.36482
= 442 gal

Volume of strong acid = 442 gal of 200 Be HCl

Volume of water = 1000 - 442 = 558 gal water *

Water - inhibitor - additives = required water

* When obtaining the volume of water required for proper dilution of concentrated HCl, inhibitor and
additives are considered part of the dilution water. Therefore, whatever determined volume of
additives is used in the acid treatment should be subtracted from the volume of dilution water
required.

Once you have solved the dilute acid calculations, most service company data books
can provide a listing of calculated volumes of stock acid and water volumes necessary
for a 1000 gal final volume of frequently used HCl strengths. Although the volume of
acid and water totals more than 1000 gal, only 1000 gal of dilute acid will result after
mixing. Chemical contraction of the two liquids results in this phenomenon. Service
company listings generally makes no allowance for subtraction of inhibitor and additive
volumes in the listed amounts of water. This subtraction can be solved to determine
the correct amount of dilution for the HCl solution.

Loading and Mixing the Acid Solution


The acid desired for the treatment should always be within 1% of the desired strength.
To ensure that concentration is uniform throughout the tank, the HCl, dilution water,
and all additives must be thoroughly mixed together. A good method for loading and
mixing follows.

1. Load the volume of fresh water less required additives into the acid tank.

2. Add the required inhibitor and additives separately. Never mix additives in
the same container before adding to the fresh water. If chemical reactions
should occur, emulsion tests for proper additive selection would become
useless.

3. Add the concentrated HCl required to the volume of water and additives
already in the acid tank to arrive at the total volume. Keep the acid-loading
hose end above the fluid level to aid in mixing the water, additives, and acid.

4. After all the required volume is placed in the acid tank, mix it further by
circulating it with a pump.

Keep the following general rules in mind:


Use a gauge stick to measure the volumes needed. If the total volume falls short of the
predetermined mark, add more water.

543
Measure quantities of additives using a loading system or 5-gal container.

Prepare HCl-HF mixtures by diluting concentrated HCl and liquid HF or by


adding ammonium bifluoride with water and then concentrated HCl. The
ammonium bifluoride dissolved in water releases HF when mixed with HCl.

Fresh water should always be used for mixing HCl-HF. Water containing
sulfate, calcium, sodium, or potassium ions should not be used because the
precipitate formation can reduce permeability in the reservoir.

Data is available from service companies for determining the amounts of concentrated HCl, water,
and ammonium bifluoride (NH4HF2) needed to prepare most HCl-HF mixtures, Because some of the
HCl is used to change ammonium bifluoride to HF, the HCl concentration required is greater than
the total final concentration of the mixture.

The following are needed to prepare 1000 gal of 12% HCl-3% HF:

HCl (959 gal of 15.4%)

NH4HF2 (384 lb)

With the above information and physical constants of HCl given in Table 2 , required volumes of
concentrated HCl and fresh water can be determined.

Mixing HCl-HF solutions requires rapid agitation or circulation of the water to facilitate
the dissolving of the NH4HF2 and proper mixing of all acid ingredients. Use the following
procedure to prepare HCl-HF:

1. Place the required volume of dilution water in the acid tank.

2. While agitating, add the remaining ingredients in this order to allow


complete mixing or dissolving of the additives:

(a)inhibitor, (b) additives, and (c) NH4HF2.

3. Add the proper amount of concentrated HCl and agitate the final
mixture until uniform.

Using the gauge stick, follow the same measuring procedures for loading HCl.

Testing Acids
The use of specific gravity or degrees Baume measurements for determining acid
strength is subject to error. Only the total amount of dissolved solids is measured. If
an HCl solution contains salt, the density of this solution will indicate a higher acid
strength because the total amount of dissolved solids has been increased by other ions
dissolved in the solution.

544
Here, then, is the best, most common method of determining acid strength: a simple
acid-base titration. This method can be used in the laboratory or in the field. For
mixtures of HCl-HF, titration can also be used to determine total acid concentration.

To determine acid strength by titration, the following equipment and reagents are
needed:

a syringe (1 ml or 5 ml)

a beaker (150 ml)

stirring rods (if available, use instead of swirling)

distilled water

2N sodium hydroxide (NaOH) solution

phenolphthalein indicator solution (dropper bottle)

The procedure for determining HCl strength is


1. Place 1 ml of acid to be tested into a 150-mi beaker. Dilute with distilled water to about
50 ml, and add 2 drops of phenolphthalein indicator to the acid sample.

2. Using a l-ml or 5-ml syringe, add the 2N NaOH by drops, while swirling the
acid sample, until the acid turns pink. Record the number of ml of 2N NaOH
used.

3. Determine the percentage of HCl in the acid from Figure 4 (Determination of


HCl strength with NaOH ).

545
Figure 4

4. Try to remove all air bubbles that may form in the syringes.

This procedure gives the approximate percentage of HCl (plus or minus 1%). It is ideally suited for
field use.

546
HIDRAULIC FRACTURE

HIDRAULIC FRACTURE FUNDAMENTALS

Objectives and Economic Considerations


Hydraulic fracturing is a type of well stimulation treatment designed to bypass near-
wellbore damage and improve the fluid flow path from the formation to the well (
Figure 1 , The hydraulic fracturing treatment).

Figure 1

It is an expensive and complicated engineering activity that requires, at minimum, a


knowledge of reservoir engineering, fluid flow in porous media, elasticity, fracture
mechanics, fluid rheology and fluid mechanics, solids transport, and issues related to
gel chemistry.

547
Since its first use in 1946 as a means of improving production from marginal wells in
Kansas, hydraulic fracturing has become a dominant completion technique. By 1993,
fracture treatments were being carried out on 40% of new, completed oil wells and
70% of gas wells in the USA. Of course, fracture treatments are not restricted to new
wells; they are also widely used to stimulate older wells and to re-stimulate already
treated wells.

When a hydrocarbon producing well is drilled and completed, even if the best practices
are followed, a certain "damage" is created around the wellbore ( Figure 2, Skin effect
interpreted as altered permeability zone).

Figure 2

This damage takes the form of an additional, uninvited resistance to hydrocarbon flow.
Since in most cases the flow of fluids is converging radially toward the wellbore, this
extra resistance causes a very large loss of pressure, decreasing the overall well
productivity. In a hydraulic fracturing treatment, a high-viscosity fluid is injected into
the well at treating pressures that are higher than the so-called formation breakdown
pressure (practically speaking, the minimum horizontal stress.) These high pressures
typically result in the propagation of a two-wing, vertically oriented. Fluid injection
continues for some time beyond this initial propagation, and when the created fracture
is wide enough to accept them, solid particles (sand or some other type of proppant
material) are injected simultaneously with the carrying fluid. The proppant material
gradually fills up the fracture so that when the pumps are stopped, the fracture faces
gradually close on the proppant.

The propped-open, vertically oriented fracture that results from a successful treatment
might be several dozen or several hundred feet high, and possibly several thousand
feet long. Although it will typically be only a fraction of an inch wide, it will drastically

548
change the streamline structure in the formation. Not only will it bypass near-wellbore
damage, but it will also impart a bilinear rather than a radial flow structure that can
enhance well productivity by as much as four to six times that of the initial stabilized
rate.

The objective of any fracture stimulation treatment is to improve the wells productivity
(or injectivity) index. The production engineer can then use this productivity index
improvement to achieve certain economic goals such as increasing production rate at a
certain pressure drawdown, or decreasing pressure drawdown while maintaining an
economic production rate. The latter is the basis for such non-traditional applications of
fracturing, as sand control and condensate dropout control, which are seeing ever-
increasing use.

Strictly speaking, fracture stimulation affects only the rate at which hydrocarbons are
withdrawn from the reservoir at a certain pressure drawdown. It does not increase the
total amount of petroleum that can be produced from the reservoir, provided time and
economics are not relevant factors. But in the real world, time and economics are
relevant. Once economics enters the picture, it is readily apparent that a large number
of currently producing oil and gas wells (even entire fields) could not have been
produced at all without being fractured, because of their uneconomical rates of natural
productivity. In this sense, we can consider fracturing also as a means of increasing
industrial reserves. It is estimated that over 25% of the total hydrocarbon reserves in
the United States would not have been recovered without the advent of hydraulic
fracturing. Similarly beneficial results from fracturing treatments have been realized in
many low permeability and/or already mature oil and gas fields outside the United
States.

For a long time, however, formations having an effective permeability of more than 1
md were rarely selected as candidates for extensive fracture stimulation programs. The
situation has changed dramatically with the rise of high-permeability fracturing (frac &
pack, tip screen-out technique.) Recently propped hydraulic fracturing has displaced
other stimulation and/or sand-control methods at a tremendous pace in high-
permeability reservoirs.

Productivity Index and Skin Factor

The Productivity Index of a specific well can be considered as the proportionality


constant, J in the deliverability equation between production rate and driving force
(pressure drawdown):

(1)

During its life span, a well is subject to several changes with respect to flow conditions. For the
production engineer, the most important flow regime is pseudosteady-state in a bounded circular
reservoir, where the pressure drawdown is defined as the difference of average reservoir pressure
and well flowing pressure:

549
(2)

and the Productivity Index is approximately given by

(3)

where k is the formation permeability, h is the formation thickness, B is the formation volume factor,
is the viscosity of the oil, re is the drainage radius and rw is the well radius and s is the radial flow
skin factor representing the effect of near-wellbore damage.

For gas wells, we can define a similar relation in terms of the squared pressures and
gas properties as:

(4)

and

(5)

where is average viscosity; Z is average compressibility factor; T is temperature and the index sc
refers to standard conditions.

Even using optimum drilling and completion practices, some kind of near-well damage
is present. In Equations 3 and 5, the skin-factor is dimensionless and can be
considered as the proportionality constant between the excess pressure drawdown
(due to damage) and production rate, provided both are cast into dimensionless form.
For an undamaged well, the skin factor is zero; for a damaged well, it has a positive
value. One way to visualize the skin ( Figure 2) is to consider a damaged cylindrical
region around the well, with outer radius, rs, and impaired permeability, ks. Then,
according to Hawkins (1956):

(6)

550
The Hawkins formula shows that if the permeability impairment is several-fold, a relatively small
damage radius is enough to cause significant skin effect. In other words, radial flow is very sensitive
to near-well damage.

Well Stimulation to Improve Productivity Index

Well stimulation was introduced into the petroleum industry to eliminate the effect of
near-well damage or, in terms of the skin factor, to "restore" zero skin. A typical well
stimulation technique of this type is sandstone acidizing, where the solid particles
causing the near-well damage are dissolved by the acid and the original permeability is
restored in the "skin region". Obviously, such a treatment does not change the
structure of the streamlines for the flowing hydrocarbon.

In carbonate reservoirs, matrix acidizing may not only restore the hypothetical original
state, but it may even establish larger-than-original permeability in a finite region near
the well. The streamline structure remains intact following the stimulation treatment,
but the skin factor may attain a value of less than zero. Negative skin indicates that
the treatment has created a flow situation that is better than that of the hypothetical
undamaged formation. In fact, it is not even necessary to eliminate all the damage. It
is enough that the treatment creates enough flow capacity near the well to bypass the
damaged zone.

In a hydraulic fracturing treatment, we create a conductive fracture by driving a "fluid


wedge" through the rock. We then place a solid propping agent in the created void to
prevent the fracture from healing and to provide the desired conductivity.
Alternatively, we sometimes achieve the post-treatment fracture conductivity by using
a low-pH fluid to dissolve a portion of the rock on the fracture face. Then the two
etched surfaces are unable to close and seal properly, and therefore a high-
conductivity conduit remains in the formation. This technique is termed acid fracturing.
Because of the relative significance of propped fractures, much of our remaining
discussion will concentrate on them.

With propped fracturing, we not only bypass the damaged zone, but by superposing a
highly conductive planar conduit on the formation we change the geometric structure
of the streamlines. The Productivity Index thus increases significantly. One way to
numerically characterize the effect of a propped fracture is to introduce the pseudo-
skin factor, sf (Cinco-Ley and Samaniego, 1979). We can then express the post-
treatment Productivity Index as

(7)

Having placed a propped fracture that really bypasses damage, the pseudo-skin factor will be
negative. The question is, of course, how to predict sf, knowing the relevant formation and fracture
properties. Once we can answer this question, we can optimize the treatment.

Dimensionless Fracture Conductivity

551
Following Cinco-Ley and Samaniego (1979) we assume a fully penetrating rectangular
fracture, in which the fracture height and formation thickness are equal ( Figure 3,
Fully penetrating rectangular fracture). The fracture half-length xf is the length of one
wing of a fracture.

Figure 3

Most authors assume that two symmetrical fracture wings are created simultaneously
during a fracture operation, with the total overall length equal to twice the half-length.

To obtain better insight, we rewrite Equation 7 into an expanded form:

(8)

where the fracture half-length, xf, is introduced directly into the Productivity Index. Instead of the
pseudo-skin factor, we use the dimensionless factor f, which does not contain reference to the
wellbore radius. While the two forms of the Productivity Index are algebraically equivalent, the
expanded form is physically more meaningful. The expanded form of the denominator involves
three terms. The first term, -0.75, is present because the pressure drawdown is defined in terms of
average pressure. The second term, , represents the effect of the fracture half-length. The third
term, f, represents the effect of a combination of fracture variables called dimensionless fracture
conductivity, CfD:

552
(9)

where k is the reservoir permeability, xf is the half-length of the propped fracture, kf is the
permeability of the proppant pack and w is the average fracture width. The dimensionless fracture
conductivity, CfD should not be confused with the dimensioned fracture conductivity, kf w. The
dimensionless fracture conductivity expresses how the two functions of the fracture are related. The
two functions are to
conduct the hydrocarbon inside the fracture to the well,

collect the hydrocarbon from the surrounding matrix rock.

If CfD << 1, flow within the fracture is restricted, but flow into the fracture from the surrounding matrix
is unrestricted.

If CfD >> 1, flow within the fracture is unrestricted, but flow from the surrounding
matrix into the fracture is restricted.

The expanded form of the post-treatment Productivity Index (Equation. 8) indicates


that post-treatment performance is not related to the wellbore radius and the original
skin factor. That is how it should be, because the radial streamline structure has been
changed and the fracture bypasses the pre-treatment damage.

The key to understanding hydraulic fracturing is that the dimensionless factor, f,


depends on CfD only. The most well-known graphical representation of the function f
(CfD) was given by Cinco-Ley and Samaniego, (1981) as shown in Figure 4 (Factor f,
after Cinco-Ley and Samaniego, 1981).

553
Figure 4

In modern engineering, "a programmable equation might be worth a thousand


pictures," so we also provide a simple but accurate curve fit:

(10)

When the dimensionless fracture conductivity is high (e.g., CfD > 100, which is possible for low
permeability formations having undergone a massive hydraulic fracturing treatment) the behavior is
similar to that of an infinite conductivity fracture (Gringarten and Ramey, 1974). For an infinite
conductivity fracture, the dimensionless factor f is equal to ln 2, indicating that the fractured well
produces similarly to a hypothetical well of enlarged radius equal to xf/2. Such an infinite
conductivity behavior is, however, impossible to achieve in most formations except for those with
very low permeability. In medium and high permeability formations, the propped fracture is always
of finite conductivity.

In a finite conductivity fracture, we have a case of two players (fracture length and
fracture width) competing for the same resource: an incremental amount of propping
agent. In other words, we can use the propping agent to increase fracture length or
width.

Before the advent of tip screenout (TSO) techniques, fracture extent and width were
difficult to influence separately. The TSO technique has brought a significant change to
this design philosophy. Now, fracture width can be increased without increasing the

554
fracture extent. In this context, we can formulate a strictly technical optimization
problem:

How should we select the optimum fracture length and width under the constraint
that the proppant volume is given?
The solution to this problem is of primary importance in understanding hydraulic fracturing.
Surprisingly, it was already found as early as 1961 by Prats, but, unfortunately, has since been
somewhat forgotten. Prats assumed that the volume of one wing, Vf, the fracture height, hf, and the
two permeabilities, k, and kf, are given and wanted to find the optimum width and half-length.

We can use the same propped volume to create a narrow, elongated fracture or a
wide, short fracture ( Figure 5, Two ways to place the same amount of proppant into
the formation).

Figure 5

It is convenient to select CfD as the decision variable, and then to express the fracture
half-length using the propped volume of one wing, Vf as:

(11)

Substituting Equation 11 into Equation 8, we obtain

555
(12)

in which the only unknown is CfD . Since the drainage radius, formation thickness, two permeabilities
and the propped volume are fixed, the maximum productivity index occurs when the quantity

(13)

reaches its minimum. The quantity y is also shown on Figure 4. Since it depends only on CfD, the
optimum CfD,opt = 1.6 is a given constant for any reservoir and any fixed amount of proppant.

The optimum dimensionless fracture conductivity corresponds to the best compromise


between the fractures capacity to conduct hydrocarbons and the reservoirs capacity
to deliver hydrocarbons, under the constraint that the fracture volume is fixed to any
prescribed value.

Once we know the volume of proppant that can be placed into one wing of the
fracture, Vf, we can calculate the optimum fracture dimensions as

and

(14)

Moreover, since

(15)

and yopt - 0.75 = 0.869, we obtain

(16)

556
There are several implications of the above results.
There is no theoretical difference between low and high permeability fracturing. In both
cases, there exists a technically optimum fracture, and in both cases it should have a
dimensionless fracture conductivity of 1.6. In a low permeability formation, this requirement
results in a long and narrow fracture. In high permeability formations, a short and wide
fracture will provide the same dimensionless conductivity.

Increasing the volume of proppant or the permeability of the proppant pack by


a given factor (for example, 2) has exactly the same effect on the productivity if
otherwise the proppant is placed optimally.

The skin improvement depends on the amount of proppant (or on the


proppant pack permeability) according to a log-square-root relation.

To achieve the same post-treatment skin factor in a low and a high


permeability formation the volume of proppant should be increased by the
same factor as the ratio of the formation permeabilities, provided all the other
formation and proppant parameters are the same.

The above relations shed light on the role of the individual variables and provide for the theoretical
optimum placement of a given amount of proppant. In practice, however, there may be several
factors forcing us to depart from this theoretical optimum.
Since not all proppant will be placed into the permeable layer, the optimum length and
width should be calculated with the effective volume, subtracting the proppant placed in the
non-productive layers.

In low permeability formations, the indicated fracture width might be too small
(when the permeability of the proppant pack cannot be considered constant).
Therefore, a minimum width limit should be applied.

In high permeability formations, the indicated fracture length might not be


enough to bypass the damaged zone, therefore a minimum length should be
applied.

Considerable fracture width can be lost because of proppant embedment into


soft formations.

For gas wells, non-Darcy effects may create a dependence of the apparent
permeability of the proppant pack on the production rate itself

Strictly speaking the Cinco-Ley and Samaniego pseudo skin factor is valid only
for xf<<1 and at late times. Fortunately, for moderate and high permeability
formations (k >>1md), the above conditions are likely to be fulfilled.

Transient flow regime, high penetration ratio of the fracture with respect to the drainage area, and
other phenomena may also modify the optimum compromise between width and length, but these
issues are of secondary importance.

Having settled the optimization of fracture length vs. width for a fixed proppant
volume, the remaining task is to optimize treatment size, which is best characterized

557
by proppant volume placed into the formation. The currently preferred method of
minimizing net present value by varying propped length but using an arbitrary CfD of 10
or 30 is theoretically wrong. This has to be kept in mind when studying older
publications, along with the fact that several other definitions of the dimensionless
fracture conductivity that are in use contain an additional factor of or /2.

The pseudo-skin factor is not the only possible indicator for visualizing the effect of a
fracture treatment. We could also derive all of the above results using the concept of
equivalent wellbore radius. We have to be very cautious not to use both the pseudo-
skin factor and the equivalent wellbore radius at the same time, however, because that
might lead to accounting for the same effect twice. Therefore, in this treatise we use
only the pseudo skin concept.

Sizing of Fracturing Treatments

Optimizing the treatment size (i.e., finding the optimum proppant volume) is an
economic rather than a technical optimization issue. The more proppant that is placed
in the formation (otherwise optimally), the better the performance of the well will be.
At this point, economic considerations must take over. The additional revenue from a
larger propped volume at some point becomes marginal compared to the more-than-
linearly increasing costs. This situation is properly treated by net present value (NPV)
analysis.

The NPV is the difference between the present value of all receipts and costs (both
current and future) resulting from the stimulation treatment. Future receipts and costs
are converted into present value using a discount rate and taking into account the year
in which they will appear. The NPV (as other equivalent indicators are also) is sensitive
to the selected discount rate and to the predicted future hydrocarbon prices.

When using NPV or any related economic indicator, we understand that for any given
treatment size, we find the technically optimal way to place the proppant. If we neglect
that issue, the result of the NPV optimization will be wrong. We notice that there is no
reason to fill technically oriented publications with details of the NPV technique,
because that subject is better covered in relevant economics and accounting
textbooks.

Although hydraulic fracturing was originally developed to improve oil well productivity,
it has since been found to have significant application to gas wells. The magnitude of
the hydraulic fracturing operations required in tight gas reservoirs has led to the
development of a special stimulation service termed massive hydraulic fracturing. Such
a treatment typically involves pumping very large volumes of frac fluid and proppant in
a single treatment to create an exceptionally deep-penetrating propped fracture.
Following this type of treatment, we may produce the reservoir at much higher rates
from a limited number of wells, thus avoiding the expense of extensive infill
development drilling. In view of the above derivations, however, it should be obvious
that the additional performance improvement is physically limited. In fact the
incremental benefit from an additional incremental amount of proppant is even less if
the fracture already penetrates a significant portion of the drainage area. Therefore,
oversized treatments are likely to be attractive only in periods of high gas prices.

558
Formation and Fracture Properties Affecting the Performance

At this point it is useful to make a list of properties of the formation and fracture that
directly affect well performance ( Table 1 ).

Formation Fracture

Pay thickness Extent

Permeability Proppant
Permeability

Fluid viscosity Effective Propped


Volume

Drainage radius

Table 1: Primary formation and fracture variables affecting performance

Once the optimum dimensionless fracture conductivity is understood, fracture length


and width are no longer freely selectable design parameters. From the production
engineers point of view, the amount of proppant (propped volume) should be the
primary variable characterizing the treatment size. Fracturing engineers, however,
traditionally prefer to think about fracture half-length as the main variable.

Fracture Initiation Orientation and Growth

Fracture Initiation

In most cases, a fracture may be initiated by applying hydraulic pressure to an


exposed formation. Prior to fracture initiation, a positive differential pressure will cause
the fluid to enter the formation in a radial flow pattern, with the rate of fluid flow
through the rock limited to a rate that is in compliance with Darcy's law.

Maintaining the injection rate of a fluid above the maximum matrix flow capacity of the
exposed formation area will continually increase the flowing pressure at the wellbore.
Finally, the pore pressure will be increased to the point at which the rock ruptures in
tension in a direction perpendicular to the least principal stress present in the
formation.

Fracture Orientation

Fracture orientation is directly related to the formations far field stress state. The
fracture will be oriented perpendicular to the direction of the least principal stress (
Figure 1, Fracture orientation).

559
Figure 1

This is most easily understood by realizing that for a fracture to form, a portion of the
reservoir must undergo physical deformation. The direction in which it is easiest to
push or deform the rock is the one exerting the least resistance (least stress). Thus,
the fracture will be oriented at a 90-degree angle to this stress.

The vertical stress stems from the weight of the overburden, and it is partially
translated to horizontal stresses. At sufficient depths (usually below 1000 m or 3000
ft) the minimum principal stress is horizontal; therefore, the fracture faces will be
vertical. For shallow formations, where the minimum principal stress is vertical,
horizontal (pancake) fractures will be created.

The above picture of fracture orientation is somewhat simplistic. Perforations and pre-
existing flaws and microfractures may "guide" fracture orientation, at least at the
initiation stage. Moreover, in the near-wellbore region the original stress state is
disturbed and the minimum principal stress direction might be different from the far
field direction. It is important to remember that such disturbances are localized in the
vicinity of the wellbore (say within a distance of two-to-three times the wellbore
diameter). Once the fracture extent is large enough, the far field stress state
dominates its orientation.

Fracture Growth

After breakdown, the fluid entering the fracture partly leaks off through the exposed
faces of the fracture. The other part of the fluid continues to enlarge the fracture as

560
long as sufficient hydraulic pressure is maintained and the injection rate is kept above
the rate at which the injected fluid continues to leak off into the formation. Growth is
generally confined to a single plane (perpendicular to the least principal stress), and
continues equally in all directions of the fracture plane until it encounters some barrier
limiting the growth rate in that direction ( Figure 2 , Fracture initiation and growth).

Figure 2

Simultaneously with fracture propagation, the fractures average width is also


expanding. In fact, lateral propagation, height growth and width inflation are
competitive processes. The created fracture geometry depends on how these
processes share the fluid volume left in the fracture after fluid leakoff.

We may define a fracture growth barrier as anything that limits the extension of a
fracture in any direction. Barriers may be overlying or underlying zones having
significantly different properties of elasticity than the zone being fractured (Young's
modulus of elasticity and Poisson's ratio). They may be rocks having a higher tensile
stress, high-stress loadings, or stress loadings in which the least principal stress is in a
different direction than at the wellbore. They may be rocks having higher frac
gradients, or zones having lower pore pressures. They may be slippage planesunique

561
bedding planes having no vertical bonding, in which the adjacent surfaces act almost
as if they are lubricated, and which dissipate the dynamic growth energy of a fracture.
Barriers may also be physically intruded solids, such as propping agents. In short,
fracture barriers may be combinations of any or all of these factors. Many of these
factors are difficult to measure or even estimate. The variation of minimum principal
stress from layer to layer is, however, more accessible and is considered to be the
main factor controlling height containment and growth.

Role of Formation Properties in Fracturing


The formation properties that are known to influence a fractures growth pattern,
including its height, are

Young's modulus

Poisson's ratio

Tensile strength

Fracture toughness

Permeability

Porosity

Poroelasticity constant

Along with these material properties, the actual state of the formation also affects the evolution of
the fracture. A detailed discussion regarding the effect of each of these parameters is given below.

Rock Properties

A rocks elastic properties are most often described by two different terms: Youngs
modulus and Poissons ratio.

Young's Modulus: Young's modulus (E) is essentially an index of the rocks resistance
to external force. It is defined as the ratio of the applied stress to the resulting strain:

(1)

In other words, it is a coefficient of proportionality between uniaxial stress and strain. It has the
same dimensions as pressure, and is typically measured in units of Pa or psi. Its value can be
determined from a uniaxial stress test as shown in Figure 1 (Uniaxial stress test for determining
Youngs Modulus and Poissons ratio).

562
Figure 1

Higher values of Young's modulus indicate greater stiffness. Therefore, a given amount
of fracturing fluid will create a relatively long, narrow fracture in a rock having a high
Young's modulus value. The same amount of fluid will create a shorter but wider
fracture in a rock with a low Young's modulus value, provided all other properties are
the same. This is one of the main differences in fracturing hard versus soft formations.
If the Young's modulus varies from layer to layer, it might cause a complex width
profile, with reduced widths in the layers of higher modulus values. Soft formations are
characterized by E values as low as 105 psi (Diatomite), while hard formations can
have E values as large as107 psi (hard Limestone).

Poisson's Ratio: Poisson's ratio, , is defined as the ratio of the lateral strain
demonstrated by a rock when subjected to a longitudinal load, divided by the amount
of longitudinal strain caused by the same loading. It is a dimensionless quantity,
usually ranging from 0.15 to 0.35. It can also be measured in the laboratory as shown
in Figure 1.

From a hydraulic fracturing standpoint, Poisson's ratio is primarily responsible for


translating vertical stress into horizontal stresses. It also has some (limited) influence
on fracture width.

563
Plane Strain Modulus: Most of the equations used in fracturing contain only a certain
combination of Young's modulus and Poisson's ratio, denoted by E':

(2)

The plane strain modulus is numerically very near to the Young's modulus, because the square of
the Poisson's ratio can usually be neglected with respect to one.

Shear modulus: Some authors prefer to use the shear modulus, G, which can be easily
calculated from the Young's modulus and the Poisson's ratio according to

(3)

It is important to understand that the above properties are related to the elastic behavior of the rock.
They can be measured on a core sample using static or dynamic measurement methods, or in-situ,
using dynamic (mostly sonic) methods. The "static" and "dynamic" properties may be somewhat
different.

Tensile Strength: The maximum stress that a material can tolerate without rupture in a
uniaxial tensile experiment is the tensile stress. Though the effect of the tensile
strength is minimal during fracture extension, it affects the fracture initiation
(formation breakdown) pressure.

Fracture Toughness: The critical value of the stress intensity factor, or fracture
toughness, characterizes a rocks resistance to the propagation of an existing fracture.
It is measured in Pam0.5, psift0.5, or psi inch0.5, because the stress intensity factor at the
tip of a fracture is the product of the pressure loading and the square root of a
characteristic length (such as fracture half-length). Fracture propagation occurs when
the stress intensity factor reaches its critical value. It is easier to propagate a larger
fracture than a smaller one, provided the pressure loading on the faces is the same.
When the stress intensity factor at the tip equals the fracture toughness, a special
equilibrium state is reached. Vertical fracture extension (i.e., height growth) is often
considered as a process passing through such equilibrium states. Sometimes the
lateral extension of the fracture is also considered as a sequence of such equilibrium
states. The concept itself is important, but the actual value has limited effect on our
calculations. Laboratory measurements indicate fracture toughness values ranging
from 500 to 2000 MPam0.5 (almost the same limits are obtained if expressing the
values in psiinch0.5).

Permeability: The larger the fluid leakoff, the less driving force is available for fracture
growth. Formation permeability is one of the main factors controlling fluid leakoff, and
hence, indirectly affects fracture propagation. Porosity and total compressibility have a
limited influence on fluid leakoff as well.

564
The Poroelastic Constant, , is defined by the relation:

(4)

where K is the bulk modulus (ratio of hydrostatic pressure to volumetric strain) of the dry rock
material and Ks is the same measured in a saturated sample. The saturated sample is more
resistive to compression because the fluid carries part of the load; therefore the poroelasticity
constant is less than one. Most fracturing calculations assume a value between 0.7 and 1.

All the above properties are related to the material behavior of the rock matrix.

State Variables

The following properties are related to the actual state of the rock matrix, as
determined by the in-situ conditions.

Bedding planes, layer structure: Existing large discontinuities in the rock are usually
two- dimensional structures, hence the name plane. Slippage that occurs along
bedding planes tends to dissipate the energy required for fracture propagation, and
thus reduce fracture growth in that direction. The presence of intersecting fractures or
planes of weakness hinders further fracture growth in that direction, even if growth is
not stopped completely.

The least principal stress in the reservoir (also referred to as closure stress or closure
pressure) equals the fluid pressure required to hold open an induced fracture. It is
often convenient to report this quantity as a fracture gradient, defined as the least
principal stress divided by the depth. The fracture gradient is normally calculated from
pressure measurements taken during injection tests. If it is considered a constant for
all wells in the same reservoir, the stress state around one well can be extrapolated to
another well. As the reservoir pressure varies, the least principal stress and the
fracture gradient vary simultaneously.

Pore pressure: Pore pressure influences the effective stresses in a formation via the
poroelastic constant. The change of pore pressure in the zone of interest (due to
depletion) brings about a change of the same direction (but lesser magnitude) in the
least principal stress.

The presence of higher pore pressure in an adjacent formation increases the tensile
forces present in that zone, thereby requiring a lower internal hydraulic pressure to
initiate failure caused by rupturing, which can actually cause a fracture to grow into
the adjacent formation. Conversely, an adjacent low-pressure zone (or an area of
lower pressure within a reservoir, such as that surrounding an old producing well) will
put that formation in compression and cause it to serve as a fracture barrier and stop
continued growth, or possibly divert fracture growth in another direction. Accurate
analyses of the pore pressures in the zone of interest and the surrounding formations
is one way to predict height containment of vertical fractures.

565
The stress-state of the formation is of primary importance to the fracturing engineer. It
is not necessary, however, to know every little detail. The most important issue is that
of the least principal stress and its variation with depth, often referred to as vertical
profile of the minimum stress.

Vertical Profile of Minimum Stress

At any point in the formation, the total vertical stress due to overburden, , is simply
the weight of the material above that point:

(5)

where g is the acceleration due to gravity, is the density of the rock (possibly varying with depth z)
and D is the true vertical depth.

The total stress is carried by both the "grains" and the pore fluid in a porous medium (
Figure 2, Poroelasticity).

Figure 2

566
The effective stress, , is the absolute stress minus the pore pressure (p) weighted by
the poroelastic constant ( ):

(6)

We can estimate the minimum effective horizontal stress due to the overburden by "translating" the
effective vertical stress, i.e., multiplying it by /(1-):

(7)

where is the Poisson ratio. Finally, we calculate the total horizontal stress by adding back the
poroelastic term:

(8)

In addition to the overburden, tectonic forces created by geological events have also induced
stresses in the formation, as evidenced by the many structural features present in the formation
(e.g., faults, folds, natural fractures and inhomogeneity.) Since it is not possible to fully describe the
origins and causes of these geologic events, it is also not possible to accurately predict the
magnitudes or directions of the resulting stresses. Nevertheless, we can use Equation 6 to
quantitatively explain several important phenomena:
It is often observed that the least principal stress decreases with depletion. Indeed,
Equation. 6 shows that a unit decrease of pore pressure causes
decrease in the fracturing pressure. On the contrary, injection of fluid and temporary
increase of pore pressure may increase least principal stress.

Equation 6 explains the existence of large stress contrasts between adjacent


layers. While the overburden is almost the same, the difference between the
Poisson's ratios can cause higher stresses in the layer characterized by larger
Poisson's ratio.

The fact that the least principal stress is vertical at shallow depths can be
explained if we accept that horizontal stresses are "frozen traces" of some prior
geological state, but erosion of the surface have decreased the overburden.
Consequently, for shallow formations, the vertical stress can be less than the
"frozen" horizontal stress. For deeper formations, however, the ratio of
horizontal stress to vertical stress approaches , and hence the former
is less than the latter.

The above theory is more qualitative than quantitative. Tectonic stresses due
to geologic movements can cause a large scatter around the theoretically
calculated stresses. It is generally accepted that a fracture gradient of less than
0.7 psi/ft in a tectonically relaxed reservoir indicates that a vertically oriented
fracture will be created, because it is easier to part the earth than to lift it. A

567
fracture gradient of greater than 1.1 psi/ft (which is in agreement with the
generally accepted value for the normal overburden gradient) indicates a
serious anomaly specific to the formation.

Knowledge of the vertical profile of the minimum horizontal stress is essential for two reasons. First,
the value at the center of the perforations will be the base to which net pressure is added to obtain
the fracture propagation pressure. Second, a positive stress contrast in the neighboring layers is
believed to be the most important controlling factor for height containment ( Figure 3 , Vertical
profile of minimum horizontal stress). A detailed depth-to-depth determination of the minimum
stress might be expensive and time-consuming; therefore, several methods have been suggested
for using available well logs to predict vertical stress variations.

Figure 3

Most of these methods are reliable only if used after calibration. The calibration process involves
correlating the log with known values of the minimum stress for at least one well in the formation.
The known values of the minimum stress are obtained from calibration fracturing treatments,
several variations of which are called micro fracturing, step-rate test, pump-in flow-back test, and
minifrac closure pressure determination.

In addition to the magnitude of the least horizontal stress, its orientation is also of
interest, because it governs the orientation of the induced fracture. Several methods
are available for estimating the principal stress orientation. One group of
measurements uses oriented core samples, such as an elastic strain relaxation.

568
Another group includes tiltmeter or similar measurements in open hole. Acoustic
measurements on the oriented core sample and acoustic logs of the open hole can also
provide information on the orientation of the least principal stress.

Fracturing Pressure

Concepts such as fracturing pressure are not well defined, because they are not just
properties of the formation, but are also a function of how they are determined. Except
when the injection rate is very low (at least on the order of several gallons per minute)
we have to distinguish fracture initiation pressure, fracture propagation pressure and
fracture closure pressure. These pressures, which are expressed as bottomhole values
referenced to the center of the perforations, may or may not be equal.

Fracture Initiation Pressure or breakdown pressure is the peak value of the pressure
appearing when the formation breaks down and a fracture starts to evolve. Usually it is
approximated by

(9)

where is the minimum horizontal stress, is the maximum horizontal stress, T is the
tensile stress of the rock material, is the poroelasticity constant and po is the pore pressure. In the
above equation, only the tensile stress is a material constant that can be measured in a laboratory.
The principal stresses and the pore pressure are state properties. The above theoretical breakdown
pressure might be masked by other factors such as the microstructure and material behavior of the
borehole wall, the geometry of the perforations and the properties of the fluid. The most serious
obstacle in applying Equation. 7 is that we rarely know the maximum horizontal stress. In fact,
Equation 7 is more often used in a reverse manner to approximate the maximum stress from known
minimum stress and observed breakdown pressure.

Fracture Propagation Pressure is the stabilized value of the injection pressure for a
longer period of time during which the fracture is evolving. Obviously, it is not a
material or even state property of the formation itself, because the fluid type, the
injection rate, and most important, the leakoff process may dramatically affect its
value. To understand pressure behavior, one has to be familiar with at least the
simplest mathematical models of fracture propagation.

In a narrower sense, fracture propagation pressure is associated with the so-called


step-rate test ( Figure 4 , Detection of formation breakdown from a step-rate test).

569
Figure 4

During a step-rate test, the fluid is injected into the formation while increasing the
injection rate in discrete steps. At each step, the stabilized pressure is recorded. The
plot of stabilized injection pressures versus injection rates typically shows a break
point. At low injection rates, the subsequent steps increase the pressure according to
Darcy's law. After a critical pressure is reached, a fracture is created (and propagated),
and the subsequent change in pressure with each rate step is usually much less than in
the Darcy region. The break point is usually determined by fitting two straight lines:
one through the low, and another through the high injection rate points. The procedure
calls upon the engineers subjective judgment, and therefore requires a careful
examination of all circumstances. Since the fracture propagation pressure is a state
property, its value might change during the life of a well, mostly because of pore
pressure depletion.

Fracture Closure Pressure. After a fracture calibration treatment, which is carried out
without injecting proppant material, the fracture volume gradually decreases because
of leakoff (and also because of possible back flow, if the injected fluid is flowed back
through the well). At the same time, the pressure decreases. Eventually the fracture
will close (that is, the fracture faces will contact). The bottomhole value of the pressure
at the moment of closure is the fracture closure pressure. Usually it is determined by
careful examination and processing of a pressure fall-off curve. The basis for selecting
the closure point is that before closure, the leakoff process and its combination with
the elastic behavior of the formation dominate the pressure falloff, but after closure

570
the pressure is governed by the general laws of fluid flow in porous media. Even if the
exact quantitative description of these processes is not readily available, a marked
change in the character of the pressure fall-off may be interpreted as the closure point.
This is the reason why several transformation plots are in use.

Often, a simple pressure versus time plot or pressure versus square root of time plot is
suitable to pick the closure pressure. Unfortunately, the closure pressure may be
masked by non-ideal (stochastic) events during the closure process and especially by
the particular way in which the fracture faces approach each other. The terms fracture
closure pressure and minimum principal stress are used interchangeably in the
technical literature. Figure 5 (Fracture-related pressure points: (1)

Figure 5

breakdown pressure; (2) fracture propagation pressure; (3) instantaneous shut-in


pressure; (4) closure pressure; (5) fracture reopening pressure; (6) closure pressure
from flow-back; (7) asymptotic reservoir pressure; (8) rebound pressure) shows some
of the "strategic locations" on the pressure response curves of typical fracture
calibration tests.

The pressure in a propagating fracture is higher than the closure pressure. The
difference between the actual and closure pressures is called the net pressure. In
everyday usage, the net pressure is meant at the wellbore, but in fracture propagation
models the net pressure varies along the length.

571
Quantitative Description of Fracture Growth
The increase in fracture volume over the course of a treatment is determined by the
amount of injected fluid that does not leak off.

Leakoff

The key to the material balance is fluid leakoff. Fluid leakoff is controlled by a
continuous build-up of a thin layer, or filter cake, which manifests an ever-increasing
resistance to flow through the fracture face. In reality, the actual leakoff is determined
by a coupled system, of which the filter cake is only one element. A fruitful
approximation dating back to Carter (Appendix to Howard and Fast, 1957), is to
consider the combined effect of the different phenomena as a material property.
According to this concept, the leakoff velocity, v
, is given by the Carter equation:
L

(1)

where is the leakoff coefficient (length/time0.5) and t is the time elapsed since the start of the
leakoff process. The ideas behind Carter's leakoff coefficient are that
if a filter-cake wall is building up, it will allow less fluid to pass through a unit area in unit
time; and,

the reservoir itself can take less and less fluid if it has been exposed to inflow.

Both of these phenomena can be roughly approximated as "square-root time behavior". The
integrated form of the Carter equation is:

(2)

where VLost is the fluid volume that passes through the surface AL during the time period from time

zero to time t. The first term, can be considered as width of the fluid passing through the
surface during the main part of the leakoff process. (The factor 2 appears because the integral

of is ).

The integration constant, , is called the spurt loss coefficient. It can be considered
as the width of the fluid body passing through the surface instantaneously at the very

beginning of the leakoff process. The two coefficients, and , can be determined
from laboratory tests.

A vertical hydraulic fracture has two wings. For modeling purposes, we usually assume
that these are identical, thus making it possible to model just one wing. Suppose qi is

572
the injection rate, Vi the injected volume, V the volume of fluid contained and A is the
surface area of one face, all corresponding to one wing of the fracture and a given time
t during the treatment ( Figure 1, Definition of injection rate, fracture area and
permeable height).

Figure 1

We will use the subscript e to denote the end of pumping in order to distinguish the
quantities corresponding to the time instant te. If the injection rate is constant, Vi =
qit .

The fluid efficiency defines the fraction of the fluid remaining in the
fracture: . The fracture surface, A, is the area of one face of one wing and
the average width, , is defined by the relation: .

It is often assumed that the created fracture remains in a well-defined lithological layer
(mostly the producing formation), and the fracture is therefore characterized by a
constant height, hf.

A hydraulic fracturing operation may last from tens of minutes up to several hours.
Points of the fracture face nearest to the well are opened at the beginning of pumping,
while the points at the fracture tip are "younger". To apply the following equation:

573
we must track the opening-time of the individual fracture face elements. If we divide the injected
volume by the surface area of one face of one wing, A, we obtain the so-called "would-be" width.
The would-be width can be broken down into average fracture width, leakoff width and spurt loss
width:

(3)

where the factor 2 is introduced because the fluid leaks off through both faces of one wing. The
dimensionless factor, , is the opening-time distribution factor. It reflects the effect of the distribution
of the opening-time. If all the surface is opened at the beginning of the injection, then reaches its
absolute maximum, = 2.

To obtain an analytical solution for constant injection rate, Carter considered a


hypothetical case where the fracture width remains constant during the fracture
propagation (the width "jumps" to its final value in the first instant of pumping.) Nolte
(1986) postulated a basically similar, but mathematically simpler assumption. He
assumed that the fracture surface evolves according to a power law,

(4)

where , and the exponent remains constant during the whole injection
period.

If we accept this assumption, we can easily obtain the opening time distribution factor
from the exponent . Selected values are given in Table 1.

PKN KGD Radial

: 4/5 2/3 8/9

: 1.415 1.478 1.377

Table 1: Opening time distribution factor

Once we know the opening time distribution factor, it is easy to make material balance
calculations.

Width Equations

574
Three dimensions compete for the fluid volume remaining in the fracture: lateral
extent, height and width.

Simple 2-D design models assume either that the fracture height is a given value (e.g.,
PKN and KGD models), or that the fracture is of penny shape (e.g., Radial model).
Once the problem is reduced to two dimensions, additional assumptions are applied in
order to obtain a relation between fracture extent and width. The final equations are
obtained from the conceptual model that the viscous resistance to flow gives rise to a
net pressure that is exerted on the fracture faces and keeps the fracture open.
Therefore it is not surprising that fracture width is related (among other things) to
elastic modulus (E'), injection rate to one wing (qi), viscosity of the fracturing fluid ()
and half-length (xf).

Perkins-Kern-Nordgren Width Equation

The visual representation underlying the PKN model is a two-wing rectangular fracture
of constant height. The vertical cross section of the fracture is an ellipse ( Figure 2 ,
PKN geometry.

Figure 2

575
After Perkins and Kern, 1961. Courtesy SPE). The maximum width of the ellipse at a
certain distance from the well is related to the height, plane strain modulus and net
pressure at that location. Since the net pressure is also related to injection rate and
fluid viscosity, a width equation not containing the pressure can be derived. Of
particular interest is the maximum width of the ellipse located at the wellbore, ww,0:

(5)

(The constant was originally 3.57 in the Perkins-Kern form of the equation, but the above form given
by Nordgren has become more accepted.) The average width of the fracture is related to the
maximum width according to:

(6)

Kristianovich-Zheltov-Geertsma-DeKlerk

The visual representation behind the KGD model is also a two-wing rectangular
fracture of constant height. The vertical cross section is considered, however, to be
rectangular ( Figure 3 , KGD geometry.

576
Figure 3

After Geertsma and deKlerk, 1969. Courtesy SPE). Physically, this means that the
fracture faces slip freely at the upper and lower boundary of the pay layer. The
fracture width at the wellbore is given by

(7)

(Notice that we use only one width index, because the width does not change vertically.) The
average fracture width is calculated from

(8)

Radial (Penny-shaped) Width Equation

By analogy, a radially expanding fracture ( Figure 4 ,

577
Figure 4

Radial (penny shape) geometry) has a maximum width at the wellbore:

(9)

and the average fracture width is calculated from:


(10)

Figure 5 (Comparison of PKN and KGD width equations) shows that given all the same parameters,
the curves of width versus fracture extent for the PKN and KGD models cross each other.

578
Figure 5

At smaller extent, the PKN width equation predicts larger width. The crossover occurs
approximately at the point at which a "square fracture" has been created, i.e., when .
While this fact has been used to argue for one or the other equation, the truth is that the physical
assumptions behind the KGD equation are more realistic for the small fracture extent situation. For
larger fracture extents, however, the PKN width equation is physically more sound.

No-leakoff Behavior of Width Equations

The so-called width equations relate fracture width and extent. Thus, if we know the
fracture volume, we can use the width equation to obtain the fracture dimensions. In
the particular case of negligible leakoff, the fracture volume is simply equal to the
injection rate multiplied by the injection time. Using this fact, we can derive the time
behavior of a propagating fracture for the no-leakoff case, as summarized below for
the Perkins-Kern-Nordgren, Geertsma and deKlerk, and Radial models.

1. Perkins-Kern-Nordgren model
Fracture Extent:

(the constant is 0.524 for the original PK equation)

Width:

579
(the constant is 1.91 for the original PK equation)

Net Pressure:

(the constant is 1.52 for the original PK equation)

2. Geertsma and deKlerk model


Fracture Extent:

Width:

Net Pressure:

3. Radial model

Fracture Extent:

Width:

580
Net Pressure:

Looking at the net pressure equations above, we can see that while the PKN (or PK) model predicts
an increasing treating pressure curve, the other two models predict decreasing pressure profiles. In
addition, the PKN model implies that the net pressure is higher if the injection rate is larger. The
other two models predict a net pressure varying with time independently of the injection rate, and
therefore they are of limited practical use for pressure related analysis. In general, we cannot
assume that leakoff is negligible and the above solutions in terms of time are not valid.
Nevertheless, we can combine a particular width equation with material balance relations and
obtain a closed design model.

Other Processes Controlling Fracture Extension

Our present understanding of fracture propagation is that in most cases, the simple
two-dimensional models described above predict faster fracture propagation than
actually occurs in the formation. In other words, the tip propagation is usually
retarded. This means higher-than-zero net pressure at the tip, because there is
intensive energy dissipation in the near-tip area. Several attempts have been made to
incorporate this tip phenomenon into fracture propagation models. One reasonable
approach is to introduce an apparent fracture toughness that increases with the size of
the fracture. Other approaches include a controlling relationship for the propagation
velocity, uf, incorporating some additional mechanical property of the formation
(dilatancy factor, continuum damage mechanics parameter, etc).

In principle, the lateral and vertical propagation of the fracture is subjected to the
same mechanical laws. The substantial difference is that the fracture tip meets the
same minimum stress during lateral propagation, while the vertical tip crosses several
layers with different material properties and stress state.

The equilibrium height concept of Simonson, et al. (1978) provides a simple and
reasonable method of calculating the height of the fracture if there is a sharp stress
contrast between the target layer and the over- and under-burden strata. If the
minimum horizontal stress is considerably larger in the over- and under-burden layers
(i.e., by several hundred psi), we may assume that the fracture height is determined
by the requirement of reaching the critical stress intensity factor at both the top and
bottom tips. This requirement of equilibrium poses two constraints, and so the two
penetrations can be obtained solving a system of two nonlinear equations. The solution
can be plotted as a height-map, indicating what fracture height will be reached at a
given treating pressure ( Figure 6, Height map.

581
Figure 6

The dashed line is a second (unstable) solution to the system of equations). Height-
maps are advantageous for selecting the fracture heights to be used in simple two-
dimensional design models. They also help us determine a treatment pressure limit (if,
for instance, we must avoid fracturing into a water zone).

582
HIDRAULIC FRACTURING FLUIDS

Fluid Functions
The functions of fracturing fluids are to

initiate the fracture

propagate or extend the fracture

transport the proppant where required

return to the wellbore after the treatment

These functions call for unique, extreme and sometimes contradictory fluid specifications.
Depending on what is required from a fluid at various treatment stages, it may be advantageous for
a given material property to exhibit varying values.

Fracture Initiation

The primary function of the first fluid pumped is to initiate the fracture, by increasing
the pore pressure in the rock until the tensional stress limit is exceeded and rupture
occurs. Its primary performance criterion, therefore, is a high leakoff rate. This allows
it to enter the pores and increase the pore pressure, and also to overcome the high
stress concentrations present around the wellbore due to damage caused during
drilling. To satisfy this requirement, the fluid must be what is termed a "penetrating
fluid" (acid with no fluid-loss control is an example of an excellent penetrating fluid).
Using a non-penetrating fluid to initiate a fracture would cause a higher breakdown
pressure. Weak acids and methanol-brine mixtures are typically used as breakdown
(pre-pad) fluids.

Fracture Propagation

The second function of the fracturing fluid is to extend or propagate the fracture after
it has been initiated. The less the fluid leaks off, the larger the fracture will be. Fluid
leakoff control is therefore the key to fracture propagation. The next function of the
fluid is to ensure that the created fracture is wide enough to accept the injected solid
propping agent, and to transport this proppant to the desired location. Since viscous
energy dissipation and created fracture width are interconnected, the fluid viscosity (its
rheological behavior) is of primary importance for creating fracture width.

Proppant Transport

Propping agents characteristically have a higher specific gravity than the carrier fluid,
and therefore tend to settle to the bottom of the fracture. Sufficiently high fluid
viscosities will enhance the fluids proppant-carrying capability. In considering
proppant transport, we should be aware of several factors affecting viscosity. In
particular, the temperature change as the fluid spends more and more time in the

583
fracture and the particular shear rate at a given location in the fracture both affect
viscosity and hence, proppant-carrying capacity.

Cleanup

Thus far, we have demonstrated that

the first fluid pumped should have penetrating properties

the main body of fracturing fluid should have a low leakoff rate (at least
during the pumping phase)

that portion of the fluid used to transport proppant should have high viscosity

With these three idealized characteristics, it should be possible to create a propped fracture with the
required geometry.

A fracture is satisfactory, however, only when we can easily remove the fracturing fluid
from the formation after closure. The fracturing fluid has to allow formation fluids to
flow into and through the induced fracture. Therefore, the final function of fracturing
fluid is to revert to a low-viscosity, non-damaging system that easily returns to the
wellbore and does not leave residue behind.

The presence of polymer residue in the proppant pack is called proppant-pack damage.
The filter cake at the fracture face causes fracture-face damage, and residue in the
polymer invaded zone causes formation damage. Formation damage can also be
caused by non-compatibility of the fracturing fluid or even its filtrate with the matrix
(i.e., clay swelling). In principle, we should keep in mind all the possible types of
damage that may occur during treatment; but in practice, the relative importance of
these elements varies, depending on the formation permeability. For low and medium
permeability fracturing, the major consideration should be to avoid proppant pack
damage. Due to the large area of fracture faces, the other damage constituents are of
less significance. In high permeability formations, however, the fracture face and
formation damage may become critical.

Fluid-Loss Control

Fluid loss results from several sub-processes that occur in the different zones shown in
Figure 1 (Fluid loss as a composed effect ).

584
Figure 1

Polymer-based fluids continually build up a filter-cake at the fracture face. The filter
cake represents a resistance to flow. The wider it is, the more pressure drop is used up
by the filter cake, and the less is left for driving the fluid into the formation. In the
near vicinity of the fracture face, we distinguish polymer-invaded zone and filtrate
zone. In the polymer-invaded zone, the flow is controlled by the non-Newtonian
polymer rheology. In the filtrate zone, the flow is Newtonian and the viscosity is
effectively that of water. Finally, there is the bulk of the reservoir containing the
original formation fluids. During the leakoff process, the reservoir has to accept fluid;
consequently, it reacts as any reservoir to injection. The reservoir response may be
complicated (transient linear flow, transient pseudo-radial flow; infinite acting or
boundary dominated) and might significantly effect the overall leakoff process.

It is important to understand that the engineering of the fracturing fluid can affect the
sub-processes associated with the filter cake and the polymer-invaded zone, but it may
have little or no influence on the other two sub-processes. Distributing a major portion
of treatment costs to fluid additives might be questionable if the overall leakoff
behavior is mainly controlled by the formation and remains unaffected by the
additives.

Fluid efficiency can be improved by adding gelling agents, special fluid-loss additives,
and specially formulated fluid systems. Viscosifiers added to base fluids increase the
fluid viscosity, thereby increasing pressure drop both through the filter cake and the

585
polymer invaded zone. Insoluble or slowly soluble fluid-loss additives creating a thin
skin of additional filter cake can also reduce the leakoff rate.

The terms that describe a given fluids leakoff characteristics are referred to as
fracturing fluid coefficients. These coefficients are expressed in the same units as the
overall leakoff coefficient, that is, length divided by square root of time. The three
coefficients commonly considered are viscosity controlled, wall-building, and
compressibility controlled fluid loss. For some time, it was believed that the overall
leakoff coefficient could be estimated if these three coefficients (measured in the
laboratory and calculated from formation permeability) and a typical pressure
difference were available. The rule for combining the individual coefficients was
reciprocal averaging or some variant thereof. Some sophisticated fracture design
programs do the calculation from layer to layer and use areal averaging to obtain the
overall leakoff coefficient.

A more practical approach is to consider the laboratory measurements ( Figure 2 ,


Laboratory determination of wall-building fluid loss coefficients) as a source of useful
information helping the fracturing engineer to compare fluid performance relative to
each other.

Figure 2

As far as the quantitative determination of the overall leakoff coefficient is considered,


prediction from laboratory measurements is of limited value. Instead, a fracture
calibration treatment (minifrac) should be carried out to obtain the overall leakoff
coefficient.

586
Rheology
Fluid viscosity affects a fluid systems in many ways. The rate of fluid leakoff, the solid
transporting capability of the fluid, the frictional losses in the tubulars and the cleanup
behavior are all strongly related to the rheological properties of the fluid system.
Modern fluid systems can meet multiple goals; the sufficient viscosity to carry the
proppant is only one of them.

Fracturing fluids typically exhibit non-Newtonian rheological behavior. This means that
instead of using a single viscosity value we have to specify a whole curve to describe
their flow behavior. The flow curve, relating the shear stress to shear rate is most
often described by a Power-Law model. This model is flexible enough to reflect the key
element: these fluids gradually loose their viscosity if the flow is more intense (i.e.,
they are shear thinning fluids).

A rotating viscometer, such as a Fann Model 50, is used to measure the viscosity
exhibited at different shear rates. Then the Power Law rheological model

= (/)v2 (1)

is applied to determine the fluids consistency index (K') and flow behavior index (n'). In Equation 1,
is the shear stress and is the shear rate. Shear rate has a dimension of reciprocal time and the
dimension of shear stress is similar to pressure. The flow behavior index is dimensionless, but the
consistency index has a rather odd dimension, depending on the value of n'. In the SI system, K'
has units of PaSn' while in field units it is expressed in lbf-secn'/ft2.

The apparent viscosity is defined as the ratio of shear stress to shear rate. In the case
of a Power Law fluid ( Figure 1 ,

Figure 1

587
Power Law flow curve), this reduces to:

a= K(d/dt)n-1 (2)

Apparent viscosity varies with the shear rate. Values for the apparent viscosity of the fluid at shear
rates corresponding to Fann Viscometer measurements at 100 rpm and 300 rpm (shear rates of
170 sec-1 and 511 sec-1, respectively) are typically reported and used by the industry. Their ready
availability makes them useful for comparing various fluids. Most printed references to apparent
viscosity are at one of these shear rates. These apparent viscosities are frequently taken to be
representative of the viscosity of a fluid in an open fracture (170 sec-1 for 100 rpm on the Fann
Viscometer) and in the tubulars (511 sec-1 for 300 rpm on the Fann Viscometer).

Of particular importance in hydraulic fracturing are the laminar flow regimes in two
limiting geometries ( Figure 2, Idealized cross sections for flow in fracture).

Figure 2

Slot flow occurs in a channel of rectangular cross section with an extremely large

588
aspect ratio (ratio of the longer side to the shorter side). Limiting ellipsoid flow occurs
in an elliptic cross section with an extremely large aspect ratio. The former
corresponds to the KGD geometry and the latter to the PKN geometry.

If the fluid is Newtonian, single expressions are available to calculate pressure loss in
terms of the viscosity. If the fluid is non-Newtonian, an equivalent viscosity can be
determined which substituted into the Newtonian expression gives the correct pressure
drop. The equivalent viscosity is not the same as the apparent viscosity, because it is
related to the particular flow geometry and reflects the specific shear rate profile in
that geometry. The solutions commonly used in hydraulic fracturing calculations are
summarized below, where uavg denotes volume flow rate divided by cross sectional
area and w0 stand for the maximum width of the elliptical cross section
(for details see Valk? and Economides, 1995).

Pressure Drop and Equivalent viscosity

Newtonian Model:

Slot Flow:

Flow in Limiting Elliptical Cross Section:

Power Law Model:

Slot Flow:

Flow in Limiting Elliptical Cross Section:

In calculating the friction pressure drop through the tubular goods, the well-known turbulent flow
correlations are less useful for fracturing fluids and special relations have to be applied because of
the drag reduction phenomena caused by the long polymer chains.

589
Proppant-Carrying Ability and Friction Loss
Fluid viscosity is of paramount importance in proppant transport. We can estimate the
settling velocity of an individual particle in a Power Law fluid (if the viscosity is large
enough to ensure laminar settling) by the following generalization of Stokes' law:

(1)

where dp is the proppant diameter, and (p-f) is the density difference between the proppant and the
fluid. Such a procedure provides only order-of-magnitude accuracy, because the settling shear rate
usually falls outside the validity range of the Power Law model. Still, it is often enough to answer the
main question:
Is the settling velocity low enough to consider the fluid to be of perfect transport type?
If the answer to this question is No, the fluid is of drop-out type.

When a low viscosity drop-out type fluid is used, the proppant continually settles
toward the bottom of the fracture as the fluid moves away ( Figure 1, Proppant
transport by a drop-out type fluid).

Figure 1

590
A bed of proppant is then deposited on the bottom of the fracture, gradually building
up in a dune-type depositional pattern. The pack width is equal to the dynamic
hydraulic width. The pack height is determined by the critical velocity required to move
proppant laterally over the top of the proppant pack. The propped length is usually
much less than the dynamically created length. The pack height continues to increase
after pumping has stopped, since those proppants still in suspension at that point
continue to settle. Settling continues until the fracture has closed on the proppant. The
resulting geometry of the proppant bed is difficult to predict, in fact most of the
proppant may be placed out of the pay layer.

When the more viscous sand-transport type of fluid is used, proppant settling is
negligible and the total created fracture height is propped. After closure, the width of
the proppant pack is more uniformly distributed, both in the vertical and lateral
directions, than it is in the case of a drop-out type fluid.

While thicker fluids generally have better proppant transport capability, a consideration
of other important fluid functions suggests using only as thick a fluid as is absolutely
necessary for reliable solid transport.

Friction Loss
Although high viscosity is generally desirable in a fracturing fluid, it can result in
extreme friction loss when pumping fluid through the wellbore tubulars. Fortunately,
this friction loss is less tightly related to viscosity in the turbulent flow regime than in
the laminar regime, especially for polymer-based fracturing fluids. The long linear
polymer chains have a tendency to suppress turbulence; fracturing fluids therefore
exhibit the so called drag reduction phenomenon and have less friction pressure at
high rates than predicted from correlations valid for Newtonian fluids. Nevertheless,
the friction losses in the tubulars may be responsible for a significant part of the cost
of the treatment because pumping costs are calculated from the horsepower
requirements.

Pressure/Temperature/pH Limitations

Most fracturing fluids are temperature-sensitive. We can overcome this sensitivity to a


certain extent by using special additives (such as temperature stabilizers) to delay
viscosity degradation at high temperatures, or by using additional breakers when
treating low-temperature wells. For very small treatments, the exposure to high
temperature may not last long enough for fluid degradation to become a serious
concern.

Other ways of overcoming temperature sensitivity include pumping larger fluid


volumes to cool down the fracture face, or pumping faster to reduce the exposure
time. An inexpensive cool-down prepad can be pumped ahead of a high-efficiency
fracturing fluid to better utilize the capabilities of the more expensive fluid. High-
temperature fluid systems and special cooldown techniques, which permit the effective
fracture stimulation of wells having a static bottomhole temperature in excess of 400
o
F, are also available.

591
It can be equally challenging to design fluid systems for use in shallow, low-
temperature wells. Conventional breakers used in crosslinked fluids have to reach a
certain triggering temperature before they can start to act. Thus, shallow wells
typically call for linear (non-crosslinked) fluids that are classified as drop-out type
fluids.

Crosslinked gels are especially sensitive to the pH and hence the chemical nature of
the rock. Therefore, the formation water composition should be taken into
consideration.

Types of Fluids
There are many different types of fluids used in fracture stimulation. Early fracture
treatments almost exclusively used crude oils or special refined oils to ensure complete
compatibility with the reservoir. Water-base systems, the safest and easiest to use,
are currently the most common. There are still several applications for oil-base and
other fluids, with the principal considerations being their compatibility with the
reservoir and their local availability.

Water-Base Fluids

Water may be used in a wide range of formation types, and over a wide range of
temperatures and pressures. It is generally available at low cost, and it is relatively
easy to modify its fluid properties for an additional moderate expense. We can increase
its viscosity, for example, by adding gelling agents. Some of the gelling agents used as
viscosifiers are

natural guar gum (Guar)

hydroxypropyl guar (HPG)

hydroxyethyl cellulose (HEC)

carboxymethyl hydroxyethyl cellulose (CMHEC)

Guar gum has been available in several forms for many years, and is still one of the most
commonly used viscosifiers. It is processed from the commercially grown guar bean. During the
rather crude refining process, some of the insoluble husk is ground and mixed in with the desired
end product. This insoluble material, accounting for about 9-13 percent of the total solids content,
serves to supplement the fluid-loss additives, thereby improving the frac fluids efficiency. There is
some concern, however, regarding the possible detrimental effect of these insoluble materials on
proppant conductivity. Ongoing studies are attempting to accurately quantify these negative effects.
In the meantime, many design engineers have requested the use of materials containing fewer
insoluble solids.

The desire to avoid solids-related problems has spurred the development of such
alternative gelling agents as Hydroxypropyl guar (HPG). This manufactured material
has a total insoluble solids content accounting for less than 3 percent of its total
weight. The viscosity development and friction loss properties of guar and HPG are

592
about the same. Most water-base fracturing fluid systems use one of these two
viscosifiers.

Cleaner viscosifiers, namely hydroxyethyl cellulose (HEC) and carboxymethyl


hydroxyethyl cellulose (CMHEC), contain less than 1% insoluble materials. They are
used infrequently, however, because of certain temperature limitations, higher cost,
and reduced fluid efficiency due to the lack of supplemental fluid-loss control that
insoluble materials usually provide. These "non wall-building" fluids have been
extensively used in high-permeability formations where fluid induced formation
damage has been of extreme concern. It is questionable, however, as to whether these
fluids are less damaging, because the polymer penetration distance is larger than for
wall-building fluids.

A linear gel (i.e., one that does not incorporate crosslinking chemistry) uses 10 to 100
pounds of gelling agent per 1000 gallons of water, with the usual concentration level
ranging from 25 to 60 pounds per 1000 gallons. A typical HPG gel of 40 lbm/1000 gal
concentration contains 0.48 percent-by-weight of polymer. This concentration provides
enough viscosity for the fluid to carry proppants through the surface equipment and
tubular goods when pumped at a nominal injection rate, although it is still classified as
a drop-out type fluid in terms of bottomhole performance.

Crosslinked gels use similar polymer loadings, but develop considerably higher
viscosities due to intermolecular bonding and the resulting formation of a complex,
three-dimensional network of polymer chains. Through crosslinking, we can increase a
fluids viscosity by several orders of magnitude while still maintaining its "pumpability."
Crosslinking agents include Borates and heavy metals such as zirconium and titanium.
The crosslinker is normally added to the slurry after the base liquid has been blended
with the proppant. Sometimes a delaying mechanism is incorporated using a
crosslinking activator to prevent the crosslinking reaction from taking place until the
fluid has reached the fracture. Most crosslinked fluids are extremely temperature, pH
and shear-sensitive. The application of these sensitive crosslinking and associated
delaying agents requires a precise control of the pH, accurate prediction of the
temperature changes and a knowledge of the shear history.

Shear sensitivity might be reversible or irreversible. If it is irreversible, the


deterioration of viscosity due to shear in the tubulars will be final. For such fluids,
delayed crosslinking might be advantageous (i.e., the crosslinking process takes place
when the fluid is already in the fracture). Table 1 lists some of the variety of
crosslinked fluids in current use.

Crosslinker Gelling Agent pH range Temperature range


o
F

B, non-delayed guar, HPG 8-12 70-300

B, delayed guar, HPG 8-12 70-300

Zr, delayed guar 7-10 150-300

Zr, delayed guar 5-8 70-250

593
Zr, delayed CMHPG, HPG 9-11 200-400

Zr-a, delayed CMHPG 3-6 70-275

Ti, non-delayed guar, HPG,

CMHPG 7-9 100-325

Ti, delayed guar, HPG,

CMHPG 7-9 100-325

Al, delayed CMHPG 4-6 70-175

Sb, non- guar, HPG 3-6 60-120


delayed

Table 1 Crosslinked fluid systems

Oil-Base Fluids

Lease crude oil is still used for some fracturing treatments because of its relatively low
cost and compatibility. It is a very inefficient proppant transport medium and has poor
fluid-loss control. But crude oil performance, like that of water, can be improved with
the use of additives. Fluid-loss additives can reduce the leakoff rate to reasonable
values, and new generation viscosifiers allow proppant transport capabilities
comparable with those of crosslinked water-based gels. The friction loss of gelled oil is
much lower than that of un-gelled oil, in some cases approaching that of gelled water.
Still, surface treating pressures for oils generally remain higher because of the lower
hydrostatic pressure of an oil column. Lease oil and gelled oils are used primarily in
formations that are extremely sensitive to water. Occasionally, Aluminum-phosphate-
esters are used in oil-base fluids as viscosifiers and friction loss reducers.

Acid-Base Fluids

In limestones or dolomitic formations, acids are sometimes used in conjunction with


fracturing fluids to etch flow channels on the formation face. The resulting fracture
conductivity is quite high, as it is proportional to the width of the etched fracture raised
to the third power.

Because of the rapid spending rate of hydrochloric acid (especially at elevated


temperatures), the use of acid-base fluids is limited to treatments in which only
nominal fracture penetration is required, or to treatments on wells with relatively cool
bottomhole temperatures. We can extend this techniques applicability by using acids
that have retarded spending rates, but the comparative cost of acid versus proppants
further limits its potential economic benefit.

Emulsions

594
Mixtures of oil and an aqueous material (either water or acid) are sometimes
emulsified and used as fracturing fluids. One such system, commonly referred to as K-l
emulsion, consists of two parts crude oil emulsified in one part of water. This system is
an economical alternative, particularly when the cost of crude oil is low. The high
viscosity of an emulsion creates wider fractures than does an aqueous linear gel, and
assists in reducing fluid leakoff and in transporting the proppants. Emulsions are
especially effective in controlling fluid loss because the fluid that leaks off from the
fracture is a multiphase mixture. The relative permeability to a multiphase system is
always lower than to either single phase. As discussed earlier, mixtures containing as
little as 5% by volume of a second discrete fluid phase have been effective in limiting
leakoff.

In order to minimize friction losses, those emulsions having water as the outside phase
are preferred, although emulsifiers that allow oil to form as the outside phase of the
emulsion are sometimes used for special situations such as in preparing a retarded
(liquid-encapsulated) acid system.

Gas/Foam Fluids

Specialized emulsions using nitrogen or carbon dioxide gas as the inner phase of an
aqueous mixture have become commercially available in recent years. These
emulsified foams typically contain 70 to 90 percent gas at bottomhole fracturing
conditions. This large volume of gas expands even more during cleanup to supplement
the reservoir energy and help with the recovery of injected fluids. The high viscosity of
the foam fluid allows it to transport proppants very efficiently and is especially
beneficial in reducing fluid loss. The multiphase composition of the leakoff fluid
satisfactorily improves the fluid efficiency. The use of foams is especially effective in
highly water-sensitive gas reservoirs where the use of oil is impractical. The relatively
small volume of water (or, in some cases, oil) included in the foam, coupled with the
normally rapid fluid recovery rate, minimizes the detrimental effect of using water in a
water-sensitive reservoir.

Due to the large volumes of expensive gas required at high pressures and the
exceptionally high horsepower required to pump this low-density fluid, this system is
generally not considered for high pressure or deep reservoirs. Exact limitations will-
depend greatly on the relative availability and cost of nitrogen and carbon dioxide, but
the use of high-quality foam as a fracturing fluid is generally limited to wells shallower
than 10,000 ft.

Fracturing Additives
Early fracturing treatments used petroleum-base fluids exclusively, primarily to ensure
the basic fluid functions and compatibility with the reservoir rock and fluids. In addition
to considering reservoir/fluid compatibility, the design engineer must also that all
essential additives are compatible with each other. The highly complex fluid systems
typically used todayespecially the crosslinked fracturing fluidsare very sensitive to
even minute changes in additive concentrations. Typical additives include

bacteria control agents

595
breakers

clay-stabilizing agents

demulsifying agents

dispersing agents

fluid loss additives

foaming agents

friction loss reducers

gypsum inhibitors

nitrogen and carbon dioxide gases

scale inhibitors

sequestering agents

sludge inhibitors

surfactants

temperature-stabilizing agents

water blockage-control agents

Fluid-Loss Additives

Extraneous solids have typically been added to fracturing fluids to reduce the amount
of leakoff into the formation and thereby improve fluid efficiency. The choice of
materials for this purpose has been quite extensive, with the primary requirement
being a mixture of widely ranged particle sizes that will form a relatively impervious
filter cake over the formation pores when the first amount of fluid leaks from the
fracture void into the matrix. The material most commonly used in fracturing
competent sandstone reservoirs is finely ground silica flour.

Larger-size silica particles (100/170 mesh, which is commonly referred to as "100


mesh") have proven quite effective in reducing fluid leakoff into natural fractures.
Other materials are also in use.

Breakers

Chemicalsusually enzymes, oxidizing agents or reducing agentsattack the


viscosifiers in the fracturing fluid, breaking the molecular chains and reducing the fluid

596
viscosity. This reduced viscosity eases the return of the fluid trapped in the reservoir
pores after the treatment and allows faster and more complete recovery of the
treatment fluids. The effectiveness of chemical breakers depends on their
concentration and temperature. The amount of breaker used in a large frac treatment
is sometimes tapered so as to provide longer stability in the leading portion and a
more rapid break in the last volume of fluid pumped. Breakers may start acting at the
moment of addition, or they may be delayed in action or triggered by another factor
such as temperature.

Other Additives

Potassium chloride (KCl) is commonly used as a special-function additive. This low-cost


salt is generally added at concentrations of about 2% by weight to prepare artificial
formation brine. It prevents the water from adversely affecting the clay particles
present in the formation, and ensures compatibility with the normally saline formation
fluids. In addition, it generally does not interfere with the performance of other
additives. In some cases additional clay stabilizers might be necessary.

Anti-foamers are sometimes required to prevent handling problems at the surface


during the pumping operation. High-rate fluid transfer frequently causes foaming,
particularly if surfactants are used in the fluid. Foaming is more severe if brines are
used instead of fresh water.

Sometimes colonies of sulfate-reducing bacteria start to grow in oil reservoirs as a


result of introducing untreated water into the reservoir during a workover operation.
The use of recycled fracturing fluid containing enzyme viscosity breakers aggravates
this situation. In order to cure and/or prevent this problem, small amounts of bacteria
control agents (biocides) are often included in the fluid.

Formation fines or minute particles such as fluid-loss additives tend to initiate and
stabilize emulsions. Demulsifiers may be required to prevent emulsions from forming
between frac fluids and the formation fluids.

Surfactants are used to reduce the surface and interstitial tension of fluids used in
completion and workover operations. Their use reduces the amount of reservoir energy
expended in removing the treatment fluids from the formation. They also act as friction
pressure drop reduction agents.

597
PROPPANT AGENTS

Introduction
The most important material used in hydraulic fracturing is the one that remains in the
well after the invoice has been paid: the propping agent, or proppant. And the most
important part of the fracturing operation is the placement of this proppant. All aspects
of treatment design should be considered from this viewpoint.

To optimize a treatments impact on the reservoirs long-term productivity, it is


essential to attain both deep fracture penetration and adequate fracture conductivity.
It is also essential to achieve a proper balance between these two parameters in order
to maximize their respective benefits. In reservoirs of very low permeability, we must
create very long fractures; at the same time, we must provide sufficient conductivity to
effectively utilize most of the created fracture length. In higher permeability reservoirs,
it is equally important to adequately prop the short fracture in order to realize the
maximum benefit from the created fracture width.

Many different propping agents are available. Each has an application or range of
applications for which it is best suited. It is the design engineers responsibility to
determine which of these materials will most economically provide the required
conductivity for a specific situation.

The factors that determine the extent of productivity improvement resulting from
fracture stimulation are

the extent of the fracture area contacting the pay layer

the dimensioned fracture conductivity (the product of the fracture width and
the proppant permeability)

Proppant Pack Permeability and Fracture Conductivity

The purpose of the propping agent is to prop open the fracture after it has been
created. The proppant must be capable of holding the fracture faces apart so that
formation fluids can flow through the fracture with a minimal loss of energy, and it
must be long lasting. From a practical standpoint, it should be capable of being placed
using pumping equipment and a fluid system that are currently available. It should
also be readily available, safe to handle, and relatively inexpensive. The most
important property of the created proppant pack is dimensioned fracture conductivity.

Dimensioned fracture conductivity is the product of the packed fractures in-situ


permeability and its effective propped width. Both permeability and width may vary
along the fracture. If there are less than three layers of the propping agent, the
permeability of the proppant "pack" increases. At one time, there were high hopes in
the fracturing community over the fact that in laboratory or in computer programs the
permeability of a partially filled monolayer might be an order of magnitude higher than
that of the densely packed pack. Now we understand that the attractiveness of the

598
monolayer concept is limited to laboratory and to computer programs. In the formation
there is no way to realize a partial monolayer which keeps the fracture faces apart
uniformly.

Proppant permeability is determined in the laboratory by measuring the flow rate


through a proppant-filled test cell of finite dimensions at several flowing pressure
differentials until steady state flow is achieved. The test cell is configured such that
elevated temperature and uniaxial loading to simulate closure stress may be applied.
The cross-sectional area of the test cell is then used in Darcy's linear flow equation to
determine the proppant permeability. The permeability is generally plotted versus the
applied stress loadings. Standardized test specifications have been accepted by the
industry, see API RP 56 "Recommended Practices for Testing Sand Used in Hydraulic
Fracturing Operations."

Fluid residue and proppant crushing can adversely affect in-situ proppant pack
permeability. Proppant crushing (mechanical damage) deteriorates permeability and to
a certain extent it may cause the delayed decrease of the effective width. For a given
type of proppant, in-situ stress will govern the level of permeability reduction due to
mechanical damage. Gas wells may exhibit an additional negative effect: the reduction
of apparent permeability due to non-Darcy flow.

The mechanical damage of proppant at the in situ closure stress is the key determinant
in proppant selection. However, the fact that sand loses part of its permeability at
higher closure stresses should not automatically mean that above a certain closure
stress, we must select a more expensive proppant. We may instead consider using
larger widths (i.e., larger areal proppant concentration) to compensate for the loss in
dimensioned conductivity. The correct decision should involve a detailed cost analysis
of the available options.

Properties Affecting Proppant Performance


Mechanical and Geometric Properties

The propping agent qualities that have consistently proven effective in achieving
proppant packs of high permeability and good integrity are:

small, rounded particles

uniform size (narrow mesh distribution)

high degree of sphericity

high compressive strength

high degree of roundness

consistent density

insolubility in reservoir fluids

599
stability at reservoir temperature

The API has established specifications and test procedures to ensure that essential proppant
properties are met.

Roundness and Sphericity

Particle roundness essentially refers to lack of angularity. In fracturing proppants, it is


determined visually and reported as a Krumbein Roundness factor on a scale of 0.1 to
1.0. A value of 0.1 indicates the presence of acute angles, as compared to 0.9
roundness indicating irregular but smooth grain curvature.

Sphericity is measured and reported on a similar basis, with 0.01 sphericity indicating
the presence of either or both highly convex and concave surface variances, and 1.0
sphericity indicating an almost perfect sphere.

A perfectly smooth sphere would have roundness of 1.0 and a sphericity of 1.0. No
naturally occurring proppant has this idealized characteristic. A manufactured proppant
having this characteristic is glass beads formed by dropping molten glass through a
cool atmosphere.

A well-formed proppant such as Northern White Sand, typically referred to as "Ottawa"


(which has a Krumbein roundness of 0.8 and sphericity of 0.7), has a permeability
versus closure stress profile as shown in Figure 1 (Sand permeability vs.

600
Figure 1

closure stress of Northern White Sand. Courtesy Dowell-Schlumberger). Texas Brown


Sand, although also generally considered to be well rounded (with a Krumbein
roundness of 0.6 and sphericity of 0.7) is less rounded than Northern White Sand. Its
permeability performance curve is shown in Figure 2 (Sand permeability vs.

Figure 2

closure stress of Texas Brown Sand. Courtesy Dowell-Schlumberger). The less rounded
sand has a slightly superior performance under conditions of low stress because the
irregular particles do not fit together quite as closely as the rounded particles.
However, as the stress is increased, the loss in permeability is more rapid in the less
rounded material because of a higher incidence of particle failure caused by the
angularity. After the amount of residue present in the fracture (from the crushed
proppant) reaches a critical level, the permeability drops off very quickly.

This particle failure by crushing is believed to be caused mainly by small imperfections


in the particle sphericity and the resultant random but highly concentrated point-to-
point loading that increases as the stress load is increased. These surrounded points

601
may be easily broken off or may actually "chisel" into adjacent particles, scratching the
surface. Like window glass, silica proppants, especially man-made glass bead
proppants, tend to fail at the point where the surface has been scratched.
Opportunities abound for the proppants to be scratched during the handling and
pumping operation. But the most likely time for this to happen is during the final
moments of fracture closure when the proppant pack is squeezed into a minimum
volume and forced to conform to any irregularities in the face of the fracture.

Size and Size Distribution

Proppant size is specified as a mesh range, such as 20/40 or 12/20. The diameters of
the largest and smallest particles in API standard mesh ranges are shown in Table 1.

API mesh Particle (in.)


size diameter

Largest Smallest

6/12 0.0661 8/16 0.0469 12/20


0.0331

0.1320 0.0937 0.0661

16/30 0.0469 20/40 0.0331 30/50


0.0232 0.0165 0.0232
0.0117

40/70 0.0165 70/140 0.0083


0.0083 0.0041

Table 1: Standard Mesh Sizes

The larger the proppant diameter, the greater the permeability will be up to a
certain point. Compare the permeabilities for 20/40 and larger sands, as shown in
Figure 1 and Figure 2 . Even though large particles provide high flow capacities at low
closure stress, they are more sensitive to increases in closure stress, and the
permeability of various sizes of proppants at high closure stresses is essentially the
same. However, this is not true for premium man-made proppants, which typically
have higher compressive strengths. They show similar proportionate losses in
permeability regardless of size.

Another important consideration in selecting the correct proppant size is the fracture
width required for proppant placement. A hydraulic width approximately equal to 3
times the nominal diameter of the proppant is considered the minimum width that will
allow free movement of the proppant through the fracture. For the 20/40 mesh, a
minimum fracture width of 0.08 in. is required. The larger 12/20-mesh proppant
requires a minimum fracture width that is twice that required for the 20/40 mesh.
Larger proppants are not only more difficult to inject but are more likely to be
damaged mechanically. More than 85% of all proppants used today are 20/40 mesh or
smaller.

602
Grain-size distribution is an important property. Using a narrow range of particle sizes
maximizes proppant permeability. By controlling the variance in particle size within a
mesh range, the particles within the pack are separated by a maximum distance, thus
providing the highest possible permeability. Proppant suppliers control mesh
distribution by passing proppant materials through stacks of vibrating sieves and
blending proportionate amounts of each cut. Table 2 lists the typical mesh range
distribution for a sample of 20/40 Ottawa Sand.

Sieve No Percent by
Weight Retained

16 0.1

20 1.4

30 37

35 34.3

40 21.7

50 5.3

pan 0.2

Table 2: Typical Sieve Analysis Result of Ottawa Sand

Manufactured proppants are typically sorted to more closely adhere to individual mesh-
distribution specification standards, with the discarded off-size material frequently
recycled through the manufacturing process.

Proppant Strength

Since resistance to mechanical damage under closure stress is known to have a major
impact on proppant performance, the effective compressive strength is one of the
prime determinants used in selecting the proppant to use. Figure 3 (Permeability of
sintered bauxite proppants. Courtesy Dowell-Schlumberger) shows the effect of closure
stress on the permeability of the proppant pack for sand and sintered bauxite.

603
Figure 3

Chemical/Temperature Stability

Extended exposure to high temperatures has been found to adversely affect the
performance of most proppants, including man-made materials.

The API standards include a quality control test to measure the solubility of a proppant
in hydrofluoric acid. This ensures that sands used as fracturing proppants have
minimal impurities, especially feldspars (alumina silicates). Feldspars not only reduce
the proppants compressive strength, but also are sensitive to some formation waters,
which causes additional strength degradation after prolonged exposure to high
temperatures. The maximum amount of feldspars allowed is two percent.

Brine has been found to be detrimental to the performance of proppants (including


manufactured materials) especially at extremely high temperatures. Long-term flow
studies using high-temperature reservoir fluids are currently underway in an attempt
to quantify the effects of temperature and time on proppant properties. Figure 4
(Effect of feldspars content on conductivity of sands at the same areal concentration.

604
Figure 4

Courtesy of Dowell-Schlumberger) shows the effects of feldspars on short-term


fracture conductivity.

Density

Knowledge of the proppant density is essential for predicting the most probable
depositional pattern of the proppant. Manufactured proppants have higher densities
than sand and therefore settle faster. This is not a serious concern unless proppants of
different densities are mixed together, in which case they would tend to segregate by
density. High-density proppants should be used with crosslinked fluids (i.e., those
having good proppant transport capabilities) to ensure proper placement of the
proppant.

Formation Properties Affecting Proppant Performance

While closure stress is by far the most important formation property affecting proppant
performance, other factors are also important. Hard rock will not allow the penetration
of proppant grains into the formation. In soft formations, however, a considerable part
of the width might be lost due to proppant embedment ( Figure 5 , Proppant

605
embedment.)

Figure 5

The compatibility of the proppant material with formation fluids is usually not a great
concern. In the case of resin coated sands, some caution might be necessary due to
possible long-term degradation of the resin coating.

Types of Proppants

The two major categories of propping agents are naturally occurring sand and
manufactured proppants. Manufactured proppants include sintered bauxite and fused
ceramic materials. Both naturally occurring sands and manufactured proppants can be
further coated with a plastic layer to improve mechanical and chemical properties.
Several names are used for large groups of propping agents.

White Sand

A special type of naturally occurring white sand, commonly referred to as "Ottawa"


sand, is an economically and mechanically attractive propping material. Ottawa sand

606
exceeds the minimum standards established by the API. Compared to other sands,
white sand has a higher apparent compressive strength, exceptional sphericity and
roundness properties, and high degree of purity.

Sintered Bauxite

Sintered bauxite has exceptionally high compressive strength ( Figure 1, Permeability


of sintered bauxite proppants.

Figure 1

Courtesy of Dowell-Schlumberger), good sphericity and roundness properties, and


relatively low density. Unfortunately, it is an order of magnitude more expensive than
white sand. Sintered bauxites are used in situations where the closure stress is very
high and the potential deterioration of the proppant pack permeability would be
difficult to compensate for with larger fracture width.

Intermediate Strength Proppants

607
This category includes all intermediate strength fused ceramic materials (often referred
to as ISP), which use a different manufacturing process as well as a lower specific
gravity raw material. ISP products, which cost less than sintered bauxite, were
developed primarily for use in the intermediate range of closure stress between 6,000
and 12,000 psi. The high degree of sphericity and the availability of narrow mesh
ranges of these manufactured proppants have broadened the application of these
materials to include some unique shallow applications in which exceptionally high
performance products are required ( Figure 2, Permeability of ISP proppants. Courtesy
of Dowell-Schlumberger).

Figure 2

Resin Coated Proppants

Resin-coated proppants are conventional proppants (commonly high-quality frac sand)


that have been coated with a thin film of plastic. The plastic film essentially
encapsulates many of the fines generated when crushing-type failure occurs. This
prevents the fines from moving about in the fracture and further reducing its flow
capacity. The film also reduces the proppants tendency to be scratched, and it reduces

608
the effect of concentrated point-to-point loading. Since this effectively increases the
closure stress limit at which the material may be used without causing significant
losses in conductivity, these products are frequently substituted for ISP materials.

A secondary effect of the plastic coating is higher friction between the individual
proppant particles while preventing several particles from being bonded together.
Therefore, resin coated materials are less subject to proppant flowback during the
flowback of the fracturing fluid after the treatment, and/or proppant production during
the first hours of hydrocarbon production from a treated well.

Another type of plastic-coated proppant uses partially cured plastic that continues to
harden after exposure to high bottom-hole temperatures.

Figure 3 shows a typical proppant selection guide (i.e., applicability ranges).

Figure 3

Table 1 gives selected permeability values for a given closure stress.

Proppant Proppant permeability


(md)

20/40 Ottawa Sand 150,000

12/20 Texas Brown 200,000


Sand

20/40 ISP, Ceramic 310,000

609
20/40 HS, Bauxite 370,000

Table 1: Laboratory-Determined Permeability of Selected Proppants at 6000 psi

610
ACID FRACTURING

Introduction

Acid fracturing, like propped fracturing, is based on the concept of inducing a fracture
hydraulically. Unlike propped fracturing, however, it does not involve using proppant to
hold the fracture open. Instead, it uses an acid-based fluid to dissolve portions of the
created fracture face. After consumption and leakoff of the fracturing acid and reaction
products, the fracture faces close. But the etched voids, which are connected to each
other, provide an increased conductivity conduit that is similar in its effect to a
propped fracture ( Figure 1, Acid etching of the fracture faces in a soluble formation).

Figure 1

Acid fracturing is conducted in the same manner as a conventional fracturing


treatment, except that a reactive fluid is substituted for the proppant and proppant
carrier fluid. The resulting geometry is determined by the pad fluid and the type and
concentration of acid used.

Hydrochloric acid (15 percent by volume), which accounts for over 90 percent of the
acid used in acid fracturing, reacts with pure limestone to dissolve approximately 1840
lbm of rock per 1000 gallons used. This is the equivalent of about 10.5 ft3 of zero-
porosity limestone. The same volume of acid would dissolve about 1680 lbm or 9.1 ft3
of dolomite having a density of 2.87 g/cm3.

611
Since corrosion of the producing well is of major concern, corrosion inhibitors have to
be added to the acid. This increases treatment costs significantly. Recently, organic
acids (formic and acetic acid) have been used successfully to replace HCl in acid
fracturing treatments. While organic acids are somewhat more expensive, they are
much less corrosive and need less or no corrosion inhibitor agents.

The impact of acid fracturing on a well's productivity is determined by the dimensioned


conductivity and length of the etched fracture. In principle, the dimensioned
conductivity of an acidized fracture is proportional to the cube of the width of the
etched fracture. This relation can be derived by combining Poisson's law for an empty
conduit with the definition of dimensioned conductivity.

The spent acid, however, does not create a real fracture-like continuous flow path
(which would not remain open, anyway). The created structure is more stochastic in
geometry and connectivity. The conductivity of the created "fracture" is much less
predictable than for propped fracturing and depends on the fine details of the etching
process, the created etched pattern, and the mechanical properties of the rock.

Because of this inherent uncertainty acid fracturing has lost significance relative to
propped fracturing, even in carbonate formations which have traditionally been treated
only by acids.

The most difficult aspect of acid fracturing is slowing and controlling the reaction to
allow it to take place where the maximum amount of benefit may be realized. Acid
spends very quickly, and special efforts are required in order to achieve deep lateral
penetration before the acid spends.

Acid fracturing involves a large number of physical and chemical phenomena, the
actual interaction of which will determine the overall success of the treatment.

Acid Fluid-Loss Control

Efficient control of acid fluid loss is the most effective method of achieving a deep-
penetrating etched fracture. Unfortunately, this control is difficult to attain, because
the acid either reacts with the fluid-loss additive (thereby destroying its effectiveness),
or it reacts with the formation. It is generally more cost-effective to attempt to control
the leakoff before the acid enters the fracture.

The most effective means of controlling acid leakoff is to use a large volume of high-
efficiency pad fluid. Leakoff control with the pad fluid works equally well in high or low
permeability formations, whereas acid fluid-loss additives are ineffective in higher-
permeability formations. This technique requires some caution, however, because the
subsequent acid stages may react with the pad's gelling agent.

A situation that is common in hard, dense, low-permeability limestone and dolomitic


reservoirs is the presence of natural hairline fractures. This type of formation typically
produces much of the reservoir fluid through these fractures rather than through the
low-permeability matrix. A hydraulic fracture may possibly intersect a large number of
these hairline fractures during propagation. Fluid leakoff into these fractures is
significant and extremely difficult to control, even when pumping nonreactive fluids.

612
When acid is pumped, the corrosive leakoff fluid quickly increases the width of the
hairline fractures. Since the amount of fluid flow through an open fracture is a function
of the cube of the width of the fracture, this increase in fracture width causes even
more acid to leak off. The use of alternating pads of acid and nonreactive pad fluid
containing high concentrations of large silica particles (70/140 mesh) has been found
to effectively achieve fracture propagation beyond natural fractures. Large volumes of
this solid material, frequently referred to as 100-mesh sand, have also proven effective
in reducing the fluid leakoff rate into natural fractures.

The types of fluid-loss additives most frequently used in acids are silica flour, organic
resins, and organic polymers, which are very slowly soluble in acid. These materials all
have some major limitations, either functional or economic. Several significant
research programs investigating alternative methods to control acid fluid loss are
currently underway.

An alternative technique for reducing the leakoff rate of acid includes the use of
viscosifiers, both polyacrylamides and celluloses, and occasionally gum Karaya, which
is similar to guar gum as used in gelling water. All of these materials have a limited life
when hydrated with acid, which particularly limits their applicability in high-
temperature treatments. One of the most effective ways of increasing acid viscosity is
the use of surfactants and oil to create high-viscosity emulsions. Normally, the acid is
mixed as the inner phase of the emulsion in an attempt to further retard the reaction
rate and minimize the corrosive nature of the leakoff fluid.

Acid Reaction Rate

One of the key issues in acid fracturing is controlling the reaction rate. The following is
a brief overview of the significant factors that affect reaction rate.

When acid reacts with rock, the reaction occurs at the rock surface. The amount of
surface area of the rock per unit volume (the area/volume ratio) is therefore an
important factor. The area/volume ratio of a porous formation is much higher than the
same ratio of an open hydraulic fracture.

Along with the area/volume ratio, we have to consider that the fresh acid has to be
transported to the rock surface and the reaction products have to be carried away.
These transport aspects of the overall process indicate that injection rate and fluid
viscosity have a large impact on the results.

The reaction rate between dolomite and acid is slower than that between pure
limestone (calcium carbonate) and acid. Because many carbonate formations are at
least partially dolomitized, the actual in-situ reaction rate is less than for a pure block
of limestone. This generally proves to be advantageous in that the slower reaction rate
effectively increases the penetration of live acid. In addition, a multi-component
formation experiences non-uniform etching patterns, reducing the chances of fracture
healing.

Temperature directly affects the reaction rate, and therefore the distance that live acid
can move through a fracture before spending. Since the rate of the chemical reaction
is sensitive to even small changes of temperature, this is one possible way to control

613
reaction rate. Cool-down effect can be reached relatively easily by the pumping of a
large volume of low-efficiency prepad fluid.

Acid viscosity affects the overall reaction rate by the transport processes. Increasing
the acid viscosity reduces the acid reaction rate by suppressing turbulence, and thus
minimizing the amount of forced convection. This holds true under both quiescent
conditions and during the movement of the acid through an open fracture. The acid
reaction product (calcium chloride brine) has a higher viscosity than does the original
acid, thereby hindering its replacement with live acid and serving to reduce the
reaction rate.

Emulsified acid systems have a retarding effect, due to the combined effects of high
viscosity and the presence of a separate, inert fluid phase. These mechanisms
effectively slow the forced convection. The presence of the separate fluid phase, which
in effect interferes with the transfer of live acid to the acid/ rock interface, apparently
is not affected by the external phase of the emulsion.

Another technique frequently employed to retard the reaction rate of acid is the
inclusion of an oil-wetting surfactant in the acid system to oil-wet the formation. This
oil film physically shields the rock, delaying acid contact.

The Changing Role of Acid Fracturing

As is true for propped fracturing, the success of an acid fracturing treatment depends
on the created fractures effectiveness at maintaining a high dimensioned conductivity.
In acid fracturing, the dependence of the final dimensioned fracture conductivity on the
materials used is less evident than it is for propped fracturing. The treatment response
depends more on the formation properties. Soft formations that undergo a high degree
of etching and have a limited surface area of support ridges may deform under the
high closure stress. Once the connectivity of the acid induced fracture deteriorates, the
effective dimensioned conductivity may be very low. This has been noted in the acid
fracturing of the soft chalk formation in the North Sea and in similar formations.

A specialized fracture acidizing technique using alternating stages of extremely viscous


pad fluid and low-viscosity acid has proven to be a viable treatment for some types of
soft formations. The low-viscosity acid "fingers" through the viscous pad, etching a
series of flow channels on the fracture face. The mostly undisturbed fracture face then
"heals" and cancels the induced stresses of fracturing, thereby allowing the etched flow
channels to remain open and provide high conductivity.

Another technique, closed fracture acidizing, targets already existing fractures in the
formation. The fractures can be natural, previously created fractures, or fractures
hydraulically induced just prior to the closed fracture acidizing treatment. The closed
fracture acidizing treatment involves pumping acid at low rates, improving the fracture
face etches.

The most damaging reaction by-products in fracture acidizing are emulsions and
scales, with emulsions causing the most concern. Most emulsion problems can be
easily prevented by pre-testing the acid system with the formation fluids it will contact.
Special scale inhibitors may be needed to prevent the precipitation of calcium sulfate

614
scale during the acidizing of formations containing anhydrite or highly sulfonated
connate water. Similarly, special inhibitors are available to prevent carbonate scales
from developing.

An additional consideration in acid fracturing is potential for corrosion of the wellhead


and wellbore equipment. This is of primary concern during fluid recovery operations if
the acid is not completely spent. Inhibitors are generally strongly cationic and have a
tendency to plate out on the formation. Using an adequate inhibitor concentration and
allowing sufficient shut-in time to completely spend the acid usually eliminates this
problem. If inhibitor costs become high, it might be necessary to consider less
corrosive acid systems (e.g., organic acids).

In deciding between an acid fracturing treatment and a propped fracture treatment, we


must first consider the relative formation solubility in acid. The most appropriate
candidates for acid fracturing are wells having fair permeability or deep and extensive
wellbore damage (i.e., those not requiring deep fracture penetration, but instead
requiring short, high-conductivity fractures). As a general rule, a propped fracture is
less expensive than an acid frac that achieves the same lateral fracture penetration, so
the next step is to determine if we can realistically achieve the required conductivity by
using high proppant concentrations to give exceptional propped fracture width. We
should examine well records to determine if there is a significant or historical problem
(e.g., lack of long-term response, many screenouts) resulting from the use of
conventional propped fracture treatments. We must address such logistical concerns as
location accessibility and the availability of sufficient fracturing equipment. Then, we
should perform comparative studies to predict the theoretical results from several
different treatments.

Because of the difficulties in predicting well response to acid fracturing, propped


fracturing has continuously gained as the technique of choice, even in formations
characterized by excellent acid solubility. Propped fracturing is the logical "next step"
in carbonate formations, where the response to acid fracturing treatments is not
satisfactory by any reason.

615
HIDRAULIC FRACTURE TREATMENT (DESIGN)

Well Completion Considerations


Casing

It is common to fracture-treat a well down casing and run tubing into the well after the
treatment. But even if it is necessary to perform the treatment through tubing, the
annulus is often "live". Therefore the casing has to tolerate the maximum pressure
occurring during the treatment. When designing a casing string or checking existing
casing, we have to consider the worst scenario, including possible screenout. Such
safety considerations determine the maximum allowable surface treating pressure.

Zonal isolation is of primary importance in modern fracturing. To place the fracture in


the target layer, the well must be cemented and the primary cement must provide a
satisfactory seal around the production casing or liner. Occasionally, flaws in the
primary cement job have to be corrected by "squeezing" additional cement behind the
casing.

Tubing

The diameter of the wells tubing string is a critical factor in fracturing operations as
well as in production. The production tubing should be specially designed to handle the
fluid injection if such a treatment is being considered in the well completion. Too small
a tubing diameter results in excessive friction loss, which increases horsepower
requirements or restricts injection rate. Either of these conditions could increase the
treatments total fluid volume requirements or could possibly even lead to a screenout.
The minimum recommended tubing size for a frac job is 2 7/8 in. EUE (2 3/8 in. EUE
for very shallow wells), while 3 1/2 in. OD tubing is a fairly standard size for a frac
string.

Normally, a tubing grade of N-80 or above is required for fracturing. Pressure is


sometimes held on the annulus above a packer to provide a tubing-burst safety factor.
Other factors to consider when using a packer on a fracturing treatment, especially on
deeper wells, are

the effects of fluid temperature acting on the tubing

the ballooning effect caused by the high internal pressure, causing shortening
of the tubing

the lengthening of the tubing caused by the frictional drag of the fluid velocity
as it is pumped through the tubulars

The ballooning and frictional lengthening effects are canceled as soon as the pumping is stopped,
but the temperature effects continue until the wellbore temperature returns to normal. The
calculations should determine the proper hook weight to minimize forces acting on the packer and
tubing. Alternative equipment, such as slip joints, can also be used to compensate for this

616
movement. The above considerations are particularly important on deep and hot wells, where the
effective tubing movement may be quite large, sometimes in excess of 20 feet.

Perforations

Both overbalanced and underbalanced perforation techniques are available. Perforation


diameter, density and phasing may significantly affect treating pressures and even
production rates. One important concern is shear degradation of the fracturing fluid
through perforations.

Multiple sets of perforations can be treated simultaneously by diverting techniques.


One diverting technique, known as limited-entry, implies a limited number of
perforations and uses a high injection rate. The perforation friction causes a pressure
increase inside the well and helps to utilize available perforations in all the individual
target zones. Ball sealers and packers are also used for treatment diversion.

Wellhead

The fracturing pressure will probably be the highest pressure to which the wellhead is
ever subjected. The wellhead equipment must have a pressure rating adequate to
accommodate the anticipated fracturing pressure plus a significant margin for error. It
may be necessary to install a special wellhead just for the frac job and change it later,
or use special high-pressure wellhead isolation tools (available for rental in most areas)
that can be run through the master valve to allow the use of standard wellheads. The
latter is generally the most economical alternative.

Basic Design
It is advantageous to start any fracturing treatment design from the point of view of
the reservoir. We should first resolve such important issues as the type and amount of
proppant placed into the pay layer and the corresponding optimal length and width,
without considering the technical details of inducing the fracture and placing the
proppant. Once we determine the amount of proppant required and the desired half
length, the next step is to select the fluid system and slurry injection rate.

(For a discussion of the equations used in determining these parameters, refer to the
IPIMS section titled "Quantitative Description of Fracture Growth," which appears
under the heading "Hydraulic Fracturing Fundamentals.")

At this point we assume that we have sufficient information to start the calculations.
That is, we assume that we know the fracture height hf, the plane strain modulus E,
the injection rate qi, the viscosity , the Carter leakoff coefficient CL and the spurt loss
coefficient Sp. We also assume a specified target length, xf. (Non-Newtonian rheology
will be considered later.)

Pumping timeThe first basic design step is to determine the pumping time, te, using
the combination of a width equation such as PKN and a simple material balance. This
part of a typical design procedure is summarized as follows:

617
1. Calculate the wellbore width at the end of pumping from the PKN (or any other) width
equation:

2.

3. Convert wellbore width into average width:

4.

5. Assume a = 1.415 (or a similar value for other geometries, i.e., 1.478 for the
KGD model and 1.377 for the Radial model).

6. Solve the following equation for te:

Selecting as the new unknown, a simple quadratic equation has to be solved

1) Calculate injected volume

and fluid efficiency

We may refine the above simple design by considering several factors, such as deviation of
permeable and fracture heights and non-Newtonian rheology.

If the permeable height, hp is less than the fracture height, it is convenient to use
apparent leakoff and spurt loss coefficients. The apparent leakoff coefficient is the
"true" leakoff coefficient (the value with respect to the permeable layer) multiplied by
the factor rp shown in Table 1.

PKN KGD Radial ( Figure 1, Ratio of permeable to

fracture area: radial geometry)

618
Figure 1

Table 1: Ratio of permeable to total surface, rp

There are several ways to incorporate non-Newtonian behavior into the width
equations. A convenient procedure is to add one additional equation connecting the
equivalent Newtonian viscosity with the flow rate. Assuming Power Law behavior for
the fluid, we can calculate the equivalent Newtonian viscosity for the average cross
section. After substituting the equivalent Newtonian viscosity into the PKN width
equation we obtain

(1)

619
Proppant scheduleOnce we know the pumping time, we can establish a proppant schedule. Our
goal is to determine the pad volume and the particular curve of proppant concentration versus time
that we have to follow during pumping. To carry out the design suggested by Nolte (1986), we need
to specify just one additional parameter: ce, the maximum proppant concentration that is technically
possible for the injected slurry.

Ideally, the proppant schedule results in a uniform proppant concentration in the


fracture at the end of pumping, with the value of the concentration equal to ce. We
derive the schedule from the requirements that

the whole length created should be propped

at the end of pumping, the proppant distribution in the fracture should be


uniform

the schedule curve should be of the form of a delayed power law, with the
exponent and fraction of pad () being equal.

It is important to notice that once we know the maximum proppant concentration and the height,
length and width at the end of pumping, we can calculate the total mass of proppant that will be
placed into one wing by

(2)

We should use this equation to select the injection rate and fluid rheology corresponding to the
specified design goal of placing the proppant of mass 2M into the formation. At this stage, M and xf

are already specified and ce is usually constrained by technical limitations; is thus the only
parameter that we can adjust, which we do by changing the fluid rheology and the injection rate.

A general procedure for determining the proppant schedule is as follows:

1) Calculate the exponent of the proppant concentration curve :

2) Calculate the pad volume and the time needed to pump it:

and

3) Calculate the required proppant concentration (mass/unit injected slurry volume) curve,
which is given by

620
where ce is the maximum proppant concentration of the injected slurry.

4) Calculate the mass of proppant placed into one wing:

5) Calculate the propped width:

where p is the porosity of the proppant bed and p is the true density of the proppant
material.
Note that in the above procedures, the injection rate qi refers to the slurry (not clean fluid) injected
into one wing. The obtained proppant mass, M, also refers to one wing. The concentrations are
given in mass per unit volume of slurry, and any other type of concentration (e.g., added mass to
unit volume of "neat" fluid) has to be converted first.

More complex proppant schedule procedures may take into account proppant
movement (both in the lateral and the vertical directions), variations in the slurry
viscosity with time and location (due to temperature, shear rate and solid content
changes), width required for free proppant movement, etc.

If the resulting propped width and also the amount of proppant differ from the design
goal, we may consider using another type of fluid and/or consider using equipment
providing a higher maximum proppant concentration.

Other design considerations There are several other checks we have to conduct
during the initial treatment design. For instance, at the end of the pad injection, the
current hydraulic width should be large enough to accommodate proppant (a width of
three proppant diameters is considered sufficient).

A considerable part of a treatments costs relate to pump horsepower. The product of


surface treating pressure and injection rate provides the theoretically required
pumping power:

(3)

The theoretical energy requirement is the power multiplied by injection time:

(4)

621
To obtain the actual power and energy requirements, we have to account for the mechanical,
electrical and other efficiencies of the equipment.

The predicted surface treating pressure is the sum of the closure pressure plus the
friction losses in the tubulars and through the perforations, minus the hydrostatic
head:

(5)

Pumping costs should be a function of both the power and the energy requirements.

(All of the calculations outlined in this section can be easily programmed. Setting up a
customized fracture-design program is advantageous when we need to compare bids
from different service companies or make quick decisions at the location. It also helps
us to understand the output and underlying approximations of larger, more complex,
fracture simulator software packages.)

Material Selection

The proppant selected for a fracturing treatment must be capable of holding the
fracture faces apart so that formation fluids can flow through the fracture with a
minimal loss of energy, and it must be long lasting. From a practical standpoint, it
should be capable of being placed using pumping equipment and a fluid system that
are currently available. It should also be readily available, safe to handle, and
relatively inexpensive.

Sand is the least expensive proppant available. The superior-performing Northern


White Sand generally commands a slightly higher price than the Texas Brown Sand.
Intermediate Strength Proppant (ISP) products generally cost about three to eight
times as much as sand, and sintered bauxite 10 to 20 times more expensive than
sand. Additionally, service companies usually include a small surcharge for pumping
larger and premium proppants to defray the increased wear on their pumping
equipment.

The fact that sand looses part of its permeability at higher closure stresses does not
automatically mean that we must select a more expensive proppant above a certain
closure stress limit. We could always consider compensating for the loss in
dimensioned conductivity by designing for larger widths (i.e., larger areal proppant
concentration). The correct decision will involve a detailed cost analysis of the available
options. One possible compromise is to use a less expensive proppant for the major
part of the treatment, and then "tail in" with high-quality proppant at the end.

The fluid selection process involves careful consideration of the fluids primary
functions, which include fracture initiation, fracture propagation, proppant transport,
and post-treatment return to the wellbore. The most important concern should be
sufficient proppant transport and leakoff control. The use of complex and expensive
fluid systems is not always justifiable. For instance, while expensive additives might

622
provide excellent leakoff control in a homogeneous matrix, they would be much less
effective in formations where fluid loss is governed by a natural fracture network.

Equipment Selection
The basic equipment components required to perform a frac treatment are high-
pressure pump trucks, blender(s), and storage equipment. Most frac treatments also
employ a wide array of auxiliary support equipment. For an offshore situation in which
the equipment must be temporarily installed on a flat barge, the equipment should be
skid-mounted rather than mounted on trucks, to keep the center of gravity as low as
possible.

Pump Trucks The pump trucks used for fracturing have high-horsepower prime
movers, each of which drives one or more positive displacement, high-efficiency triplex
pumps mounted on a heavy-duty oilfield chassis. The fluid end of each pump is
designed to operate over a sizable pressure range, with the transmission system giving
a relatively constant horsepower performance. The fluid end of the pumps can easily
be changed to extend the performance range of the pumps.

Some pumping equipment is operated from remote control panels to facilitate overall
treatment control and improve safety conditions. The output performance of these
units is typically in the 800-1500 horsepower range, with some units having two prime
mover/pumps installed on the same truck chassis. Because of the extreme operating
conditions encountered when pumping proppant/fluid slurries at high pressures, at
least one extra unit should be available as a standby for most jobs. For some jobs with
long pumping times, as much as 100% excess horsepower should be kept in reserve,
ready for use.

Frac Boats A preferred solution for offshore operations is to use a special frac boat with
all the necessary equipment permanently installed and plumbed for efficient operation.
Several frac boats are currently in use in various locations around the world and have
already effectively extended the safe and economic application of fracture stimulation,
including acid fracturing, to the more difficult offshore environment. These boats
typically incorporate the latest technological advances in sophisticated monitoring
equipment, as well as the special automated blending operations control that is
essential to the use of continuous-mix fluid systems.

Storage and Mixing units typically include multiple compartments to allow a variety of
proppants to be used on each treatment. Because excellent fracture conductivity is the
ultimate objective of all frac treatments, correct handling and blending of the
proppants are among the most critical elements of on-site operations. Fluid storage
facilities have not changed considerably in the last few years, the industry mainstay
being a 500-barrel frac tank, sometimes with an axle to facilitate movement to the
location. These tanks can be manifolded together to allow high-rate fluid transfer with
a minimal labor requirement. Recent advancements in the associated monitoring
equipment allow the fluid level of each tank to be monitored continuously from a
central point as a cross-check of the volume of fluid pumped.

Most fracturing treatments use fluids that are specially mixed on location prior to the
actual pumping operation, and then stored in large frac tanks until needed. This is

623
impractical, however, for exceptionally large treatments or for offshore locations
having limited storage facilities. The viscosity of frac fluids may start to deteriorate
slowly within a few minutes after being mixed. Sometimes, due to unforeseen job
delays, it is necessary to discard considerable amounts of expensive fluids. The total
optimal surface storage time is highly dependent upon bacterial content and
temperature. During the summer, fluid properties may deteriorate after one or two
days unless a bactericide is added. A bactericide extends the surface life of the fluid as
well as limiting the buildup of bacteria in the reservoir. If there is doubt, the fluid
properties should be re-tested before use.

It is not advisable to add gelling agents to previously gelled fluids, because the
partially viscous fluid is very inefficient at hydrating the dry polymer gel particles. As a
result, the partially wetted polymer encapsulates other particles of completely dry
polymer, creating "fisheyes" of unhydrated polymer. These fisheyes are highly
undesirable, since some of them may be large enough to permanently plug the
formation and the propped fracture.

Many of the fluid systems used today, and others currently under development, are
designed to allow the continuous mixing of all the necessary additives as the fluid is
pumped downhole. Special mixing systems are usually used to provide quality control.
Alternatively, certain other fluids are designed for use in a modified continuous-mix
mode, where additives can be safely added continuously to the partially batch-mixed
system.

Blender The most critical piece of equipment in fracture stimulation is the blender. This
unit transfers the frac fluid from the storage tanks, blends the proportionate amount of
proppant and chemical additives with the fluid, and pressurizes the suction of the high-
pressure pumps with this slurry. Since all the fluid and proppant must go through this
single unit, its continuous operation is essential to the success of every frac treatment.
The modern blender includes the following components:

a suction pump to transfer fluid

control devices to meter the simultaneous addition of several liquid and solid
additives

a complete manifold system to allow fluid selections to be changed at will

a densiometer to determine the exact proppant concentration

a system to maintain a continuous discharge pressure at the suction end of


the triplex pumps

a monitoring system to ensure that everything is functioning properly

Operational Considerations

Operational restrictions may impose limits on the hypothetically optimum treatment.


Operating conditions encountered offshore frequently impose additional economical
and unique logistical considerations on the fracturing program design.

624
Surface Location

The size and shape of the surface location that is required for a fracturing treatment is
quite different from that needed to drill the well. Surface locations used for drilling
deep wells are generally large enough to accommodate all the frac pumps and
associated equipment. But the smaller sites that are characteristic of shallow wells
frequently impose severe space restrictions, especially if the rig is still on location.
Whenever practical, the rig should be moved off the location to make more room
during a frac treatment. Alternatively, it should be shut down during the pumping
operation for safety purposes.

Industry safety standards define the minimum distances to be maintained between the
wellhead and potential ignition sources. They specify that the storage facilities for
treating fluids should be located a safe distance from the wellhead, and from potential
ignition sources if the frac fluid is flammable. There must be sufficient space available
for spotting the blender, proppant storage facilities, frac pumps, pumping manifold,
and recording/command center, and still enough room for personnel to move easily
around the equipment. The equipment that will be in operation during the treatment
should be located upwind of the well to further minimize fire hazards. The site should
also have enough room for logging equipment or other specialty equipment involved in
the overall completion program.

Sometimes it is necessary to enlarge the location prior to a fracturing treatment, or to


use space adjacent to the hard pad. The cost of any special preparations, as well as
other eventualities (such as pulling heavy equipment out of the mud because the road
or location was not adequately prepared), must be taken into account when finalizing
the treatment costs.

The injection rate for most fracturing treatments today ranges from 15 to 40 barrels
per minute (bpm), with a maximum possible rate of about 150 bpm. Higher-rate
treatments are pumped down open casing, but most treatments are directed down
large-diameter tubing, or simultaneously through the casing and tubing. The minimum
pump rate for fracturing depends primarily on individual formation characteristics and
treating fluids. It is typically around five to eight barrels per minute.

Because large-diameter tubing is frequently impractical for production purposes,


special "frac" strings are often used. When fracturing sensitive formations, this practice
should be delayed until after cleanup in order to return the fracturing fluid as soon as
possible and thereby minimize damage to the formation.

Treatments may be pumped through either the tubing or casing annulus, or through
both simultaneously, to minimize friction losses. When pumping through both the
tubing and the annulus, the rates having the same friction losses would be combined
to determine the total injection rate possible. If the treatment is pumped down the
annulus, the pressure limit on the casing must be considered. When designing the
production casing for a new well that has even a slight chance of requiring fracture
stimulation, the pressure rating should be high enough to accommodate a typical frac
treatment to be pumped via the annulus.

625
Tubing that has been used in rod pumping wells should never be used for fracturing
because of possible rod wear, which would reduce tubing strength. New tubing should
be cleaned with acid. Any old production tubing must be thoroughly cleaned prior to
use as a frac string because of the possible presence of scale or paraffin deposits that
could come loose and interfere with the frac. Tubing may be conveniently cleaned in-
place by scouring it with sand. A small volume of sand-water slurry circulated down the
tubing and then reverse-circulated at a high rate has been used very successfully to
clean tubing.

Timing

A successful fracture treatment is the result of detailed planning and thorough


pretreatment organization. The key element is good communication between all
involved participants, including the service company, rig supervisor, trucking
contractors, downhole tool service company, suppliers of frac tanks, logging company,
and the company representative.

The time required for each facet of the operation depends on the jobs size and
complexity, and on local conditions. For example, transporting frac water to a remote
location may require up to two weeks, while the same size job in an active region with
many water trucks available may take only a day or two.

The freshly cleaned fluid storage tanks are the first pieces of equipment to be set on
location. They should be spotted by the fracturing supervisor in a level area of the
location that meets the company's safety standards and allows the proper setup for
the proppant and pumping equipment. They should be filled with the type of fluid
recommended by the fracturing service company representative, making sure there is
enough for an extra stage of flush or for other unplanned contingencies.

For treatments involving large amounts of proppants, the storage facilities will be large
portable tanks that are filled after being set on the location. These may take several
hours or days to load, depending on the job size and remoteness of the well location.
On small treatments, the proppant is usually transported to the location in trucks along
with the pumping equipment, and poured from these trucks directly into the blender
for mixing. The pumping equipment is normally brought to the location on the morning
of the job and hooked up in a few hours while the frac fluid premix is underway.
Exceptionally large treatments or critical operations may require an extra day to set up
all the equipment. In certain areas, it is recommended that the frac fluid be heated to
formation temperature to reduce paraffin formation and other damaging conditions.

The actual pumping operation may require only a few minutes or a few hours, but
preparing for a frac job may take several weeks.

Safety and Quality Control

Safety is of paramount importance throughout the fracture treatment. All service


companies have stringent standards for pumping operations that must be adhered to
by all site personnel. These standards may differ depending on the type of fluid being
pumped.

626
A fracturing treatment is normally pumped at high surface pressures. A maximum
treating pressure that will protect the tubular goods (and/or prevent fracture growth
through defined barriers) should be established prior to the treatment.

Because of the high complexity of the materials and equipment used in hydraulic
fracturing, quality control has emerged as a major issue. The role of quality control is
to ensure that all the fluids meet the design specifications. On-line quality control
involves, but is not restricted to, pH and viscosity measurements.

Fracture Calibration Treatment


The information obtained from a minifrac calibration treatment includes the closure
pressure, pc, the leakoff coefficient and possibly the condition of the perforations and
the near-wellbore region. The strategic locations registered on a typical pressure
response curve during a calibration treatment are shown in Figure 1 (Fracture-related
pressure points: (1)

Figure 1

breakdown pressure; (2) fracture propagation pressure; (3) instantaneous shut-in


pressure; (4) closure pressure; (5) fracture reopening pressure; (6) closure pressure
from flow-back; (7) asymptotic reservoir pressure; (8) rebound pressure). The falloff
part of the pressure curve is used to obtain the leakoff coefficient for a given fracture
geometry. The original concept of pressure decline analysis (Nolte, 1979) is based on
the observation that during the closure process, the rate of pressure falloff provides
useful information on the intensity of the leakoff process. By contrast, during the
pumping period, the pressure is affected by many other factors, and hence the
influence of leakoff is masked.

627
If we assume that the fracture area has evolved with a constant exponent and remains
constant after the pumps are stopped, at time (te+t) the volume of the fracture is
given by

(1)

where Ae = fracture face area (one wing, one face), SP = spurt loss coefficient, CL = leakoff
coefficient, the dimensionless shut-in time is defined as

(2)

and the two-variable function: introduced by Nolte is the generalization of the opening
time distribution factor. For computational purposes, we can use the following approximations of the
g-function for various exponents (d=tD):

Dividing Equation 1 by the surface area of one face, the fracture width at time t after
the end of pumping is given by

(3)

The first term on the right-hand-side is the would-be width. To obtain the actual width, we subtract
the spurt width and the leakoff width from the would-be width. The leakoff width increases even

628
after the pumps are stopped, and the g function is the mathematical description of this process. As
seen, the time variation of the width is determined by the g (t D,) function, the length of the injection
period and the leak-off coefficient, but is not affected by the fracture area.

Unfortunately, the decrease of average width cannot be observed directly; but


according to linear elasticity theory, the net pressure during closure is directly
proportional to the average width:

(4)

The coefficient Sf is the fracture stiffness, expressed in Pa/m (psi/ft). Its inverse, 1/Sf, is sometimes
called the fracture compliance. Table 1 gives expressions of the fracture stiffness for the PKN, KGD
and Radial fracture geometries.

PKN KGD Radial

= 4/5 2/3 8/9

Table 1: Proportionality constant, Sf and suggested for basic fracture geometries

Combining Equations 3 and 4 yields:

(5)

Equation 5, first derived by Nolte, shows that the pressure fall-off in the shut-in period will follow a
straight-line trend

(6)

if plotted against the g-function (i.e., transformed time). The g-function values should be generated
with the exponent, , considered valid for the given model. The slope of the straight line, mN, is
related to the unknown leak-off coefficient by:

(7)

Substituting the relevant expression for the fracture stiffness, we can estimate the leakoff coefficient
can shown below:

629
Minifrac Analysis by the Nolte-Shlyapobersky method

Leakoff coefficient, CL:

PKN geometry:

KGD geometry:

Radial geometry:

Fracture extent:

PKN geometry:

KGD geometry:

Radial geometry:

Fracture width, we:

PKN geometry: (Vi/xfhf) 2.830CL(te)1/2

KGD geometry: (Vi/xfhf) 2.830CL(te)1/2

630
Radial geometry: [Vi/(Rf2/2)] 2.754CL(te)1/2

Fluid efficiency:

PKN geometry:

KGD geometry:

Radial geometry:

Note that the estimated leakoff coefficient for the PKN geometry does not depend on unknown
quantities, since the pumping time, fracture height and plane strain modulus are assumed to be
known. For the other two geometries, the procedure results in an estimate of the leakoff coefficient
that is strongly dependent on the fracture extent (xf or Rf).

The effect of the spurt loss is concentrated in the intercept of the straight-line with the
g = 0 axis, therefore:

(8)

Unfortunately, we do not know the would-be width (Vi/Ae) because the fracture extent is not known.
As suggested by Shlyapobersky (1987), Equation 8 can be used in a reverse manner to obtain the
unknown fracture extent if we assume that the spurt loss is negligible. Refer to the "estimated
fracture extent" relationships shown above for the three basic models. Note that the no-spurt loss
assumption also results in an estimate of the fracture length for the PKN geometry, but this value is
not used for obtaining the leakoff coefficient. For the other two models, the fracture extent is
obtained first and then the value is used in interpreting the slope.

Once the fracture extent and the leakoff coefficient are known, the average width and
the fluid efficiency are easily calculated, as shown above.

It is not acceptable to take the fluid efficiency from a minifrac treatment and use it as
an input variable for designing the main treatment, because the fluid efficiencies in the
minifrac and main treatment are different. The only parameter that is transferable is

631
the leakoff coefficient itself, but we need to use caution in interpreting it. The bulk
leakoff coefficient determined from the above method is apparent with respect to the
whole fracture area. If we have information on the permeable height hp, and it
indicates that only part of the fracture area falls into the permeable layer, then we
should convert the apparent leakoff coefficient into a "true value" with respect to the
permeable area only. We can do this simply dividing the apparent value by rp , which is
given in Table 2 .

PKN KGD Radial ( Figure 1, Ratio of permeable to fracture area:


radial geometry)

Table 2: Ratio of permeable to total surface, rp

The conventional minifrac interpretation determines a single overall leakoff coefficient.


If the quality of data permits, more complex models and more than one leakoff related
parameters might be determined from a minifrac analysis. The discussion of such
procedures is out the scope of this treatise. The interested reader may consult the
special literature.

Main Treatment

Stage Design

The productivity ratio that results from a fracture treatment depends on the final
distribution pattern of the proppant. In a packed vertical fracture, this final distribution
pattern depends primarily on

the fracturing fluid viscosity and type (i.e., drop-out type versus transport-type)

the fluid flow velocity

the proppant size, density and concentration

We may calculate the settling velocity of an individual proppant particle using the generalization of
Stokes law for shear thinning fluids. If the settling velocity multiplied by the total time available for
settling is of the order of the fracture height, the fluid is dropout type. If this product is much less, the
fluid is transport-type.

For a dropout type fluid, it is very difficult to find a reasonable proppant schedule.
When the more viscous sand-transport type of fluid is used, the proppant particles
settle only slightly during pumping. Ideally the final proppant pack will be evenly
distributed along the whole fracture in both the lateral and vertical directions. To
achieve this, the proppant concentration in the injected fluid has to be gradually

632
increased throughout the pumping period, because proppant stages pumped early lose
more fluid than the ones pumped at the end of the treatment. In addition, a
considerable volume of fluid has to be pumped without any proppant before the
proppant carrying stages. That fluid volume is called pad. The subsequent step-by-step
increase of proppant concentration is called a ramped proppant schedule.

If the proppant schedule is designed correctly, the proppant concentration at the end
of pumping is uniform within the whole fracture and equal to the proppant
concentration of the slurry injected at the last moment. The width of the pack after
closure on the proppant will be more uniform, essentially being similar to the
hydraulically induced width profile. The proppant schedule can be continuous, but it is
reasonable to divide the treatment into several discrete stages (say from 5 to 15) and
approximate the continuous schedule by stairs or "ramps."

Screenout Problems

Regardless of the type of fluid used, fluid loss will cause the proppant concentration in
a small segment of the slurry to gradually increase as it moves away from the
wellbore.

When a particular segment of the slurry becomes very concentrated, or the fracture
width at a given lateral location is too small to allow the proppant to be displaced
further, a screenout will occur. A screenout condition may happen at any point within
the fracture or, in the worst case, even within the wellbore. A tip screenout, which
takes place at or near the extremity of the fracture, results in a much slower pressure
increase than a screenout near the wellbore. A tip screenout may be a desirable
condition, especially in high permeability fracturing. Near-wellbore or wellbore
screenouts , on the other hand, should be avoided by any means.

Excessive leakoff (i.e., higher than that used in the design model) is the most common
cause of screenouts. Other factors responsible for screenouts include sudden height
growth, perforation problems, multiple strands, near-wellbore tortuosity, and
insufficient proppant-carrying capacity of the fluid or insufficient injection rate.

Most of the time, we may prevent screenouts from occurring by using a sufficient
volume of pad fluid. We should pump enough pad fluid to create a fracture width at the
leading edge of the proppant slurry that is adequate to allow two or three proppant
particles to be carried side by side. Overdesigned pad volumes should be avoided,
however, because they increase the load on the formation and lead to additional fluid
and pumping costs.

When a tip screenout occurs, pumping may be continued. The additional proppant will
inflate the width of the already created fracture, at least in soft formations. Even in
hard rock, the pressure may increase slowly enough so that we can pump enough
solids-free flush fluid to clear the wellbore before reaching the maximum pressure
limit.

When a wellbore screenout occurs, we must shut down the pumping operation
immediately because of the surface limitation on pressure. If the fracture void that has
been formed at this time does not already contain enough proppant to satisfactorily

633
prop the fracture, the resulting well performance will be very poor. The subsequent
cleanup not to mention the transportation or disposal costs of the unused material
is expensive.

As a general rule, we should not overflush the proppant-carrying stages, nor should we
stop and restart proppant addition during a fracturing operation, especially when using
a frac fluid that has excellent proppant-suspension properties. This could cause the
creation of a proppant-free section within the fracture, which could "heal" in a fashion
similar to that shown in Figure 1 (Effect of overflushing proppant ) thereby effectively
forming a choke in the fracture and severely restricting flow.

Figure 1

Treatment Monitoring

Monitoring equipment is available that allows continuous, real-time recording of all


relevant treatment information in a mobile command post at the well site. The
standard monitoring equipment records the total injection rates and surface treating
pressures. Some specialty units continuously monitor and record the proppant
concentration and the rheological properties of the frac fluid in addition to the injection
rate and the surface pressures for both the tubing and casing.

Since on-line bottomhole pressures are rarely available, by far the most important
information during the treatment is the surface treating pressure. A log-log plot of the
treating pressure is often called a Nolte-Smith plot. A qualitative interpretation of the
log-log plot is based mostly on the log-log slope. A steady positive slope on the order
of 0.25 is interpreted as unrestricted (normal) fracture propagation. An abrupt increase
in the fracture surface (e.g., due to fast height growth into another layer) is diagnosed
if the slope changes to a negative value. An increasing slope approaching the value of
unity is considered a sign of restricted tip propagation. An even larger than unit slope
indicates the fast fill-up of the near-wellbore region and the wellbore itself. The
fracturing engineer uses this information to detect a screenout situation and to
determine if it is happening at the fracture tip or near the wellbore. Depending on the

634
location of the screenout, the type of formation (hard or soft), and the established
surface pressure limit, the engineer may intervene and, if necessary, prematurely stop
the treatment.

Shut-in and Clean-up

Following the treatment, the well is shut in for several hours to allow the fracture to
close and the fluid viscosity to break. Fractures, particularly in tight reservoirs, may
require long periods to close, and during the shut-in period, excessive near-wellbore
proppant settling may occur. Such proppant settling causes a "choke" effect and
should be avoided.

If proppant settling is a potential problem, a technique called forced closure is applied.


It consists of flowing back the well quickly and hence "trapping" the proppant near the
wellbore before settling may occur. As a consequence, the proppant bridges behind the
perforations, and a "reverse packing" occurs. Immediate flowback has the additional
benefit of producing back a substantial part of the gel, even in underpressured wells.
The buildup of pressure produced in the formation by the fracturing treatment also
helps to clean up the well.

Aggressively flowing back a well requires caution, because proppant may be carried
out through the perforations. Experience shows that overbreaking of the fracturing gel
can be more detrimental because of the near-wellbore settling of proppant. Aggressive
flow back is indicated for energized fluid and foam treatments to take advantage of the
energy stored in the compressed gas. Some CO2 treatments, however, are allowed to
soak for several days prior to flowback, in order to gain additional advantage from the
penetration of CO2 into the formation.

In individual cases, additional actions might be necessary to get the well on


production. For instance, it might be advantageous to blow high-pressure gas, if
available, through the created fracture.

Wells containing proppant after the treatment should be cleaned out. Coiled tubing
with nitrogen foams is often used for this purpose. A cheaper, but still effective method
is to break up proppant with a "notched" collar and reverse out with brine.

Treatment Evaluation

Post-job evaluation involves

routinely collecting and analyzing all information relating to the fracture treatment

periodically evaluating the wells post-frac performance in relation to those of


other wells in the field

comparing the predicted treatment results with the actual post-frac


performance

635
identifying problems that occurred and determining how they could have been
avoided

A thorough evaluation of each treatment enables us to incorporate improvements into subsequent


treatments in the same or similar fields, and prevent the recurrence of identical problems.

Post-Treatment Fracture Height Determination

Temperature surveys, which are used to determine the fracture height at the wellbore,
are among the most useful tools employed in post-treatment evaluation. A
temperature survey is conducted shortly after pumping is completed. It measures the
change in bottomhole temperature that has taken place because of the large volume of
fluid injected into the formation. More and more often, this method is being replaced or
supplemented by radioactive tracer logging. Both techniques have been successfully
used to determine fracture height, although they are both subject to errors in
interpretation, mostly because they provide information only on the fracture height at
the wellbore and not in the formation.

Post-Treatment Determination of Fracture Conductivity and Length

In low permeability formations, it is often not possible to run a pressure transient test
before the fracturing treatment, because the well will not produce without fracturing.
In such a case, a pressure buildup test of the fractured well is used to obtain the
permeability and the fracture extent simultaneously. Unfortunately, this is an ill-posed
problem in the sense that many different combinations of the unknown parameters
give a good fit. In high permeability formations, the permeability is usually known and
the primary goal of a post-treatment test is to evaluate the created fracture.

The transient behavior of a vertical well intersected by a finite conductivity fracture for
infinite- acting reservoir is well known due to the works of Cinco-Ley et al. (1978,
1981).

636
Figure 1

Figure 1 (Log-log plot of dimensionless pressure for a vertically fractured well. After
Cinco-Ley and Samaniego, 1978) shows the log-log plot of the dimensionless pressure
and the "time-log derivative" parameterized by the dimensionless fracture
conductivity. In the so called bilinear flow regime, where the flow is determined by
both the reservoir and fracture properties, the plot shows a quarter slope because in
this flow regime the dimensionless pressure can be expressed as

(1)

where tDxf is the dimensionless time based on the fracture half-length:

(2)

637
Once we identify such a regime, we can construct a specialized plot of the pressure versus the
quarter root of time. The slope, mbf, of the straight line fitted is a combination of the reservoir and
fracture properties:

(3)

It is obvious from the above equation that we cannot simultaneously determine the formation
permeability and the fracture conductivity from this regime. Knowing the formation permeability, we
can determine the dimensioned fracture conductivity (kfw) from the slope, but not the fracture
extent. Our suggestion is to assume CfD = 1.6, determine an equivalent dimensioned fracture
conductivity based on Equation 4,

(4)

and calculate an equivalent fracture length according to:

(5)

Comparing the equivalent fracture length to the design length may provide valuable information on
the success of the fracturing job.

The bilinear flow period ends before tDxf = 0.01, which may be a very short time for
higher permeability reservoirs. If we cannot identify the bilinear flow regime from a
well test, the best approach is to determine the Productivity Index from pseudo-radial
flow analysis, assuming either infinite-acting or pseudo-steady-state behavior. The
Productivity Index can be obtained from a build-up test, a multiple-rate test, or else
from production data.

We would hope that the post treatment Productivity Index indicates an improvement
with respect to the theoretically undamaged (zero-skin) behavior. If there is no such
improvement, we cannot evaluate the treatment in terms of having created a fracture.
Still, the treatment may prove economically successful simply by having reduced the
skin to a smaller positive value. Assuming that the Productivity Index indicates a
negative skin, and accepting proppant permeability and formation permeability values,
we can determine the theoretical volume of proppant that we have to place into the
formation to obtain the same Productivity Index. Comparing the theoretically
necessary volume of proppant with the volume of proppant actually injected, we obtain
an overall efficiency-type indicator. In addition, we can break down the theoretically
necessary volume into a length and width using the optimum dimensionless fracture
conductivity. We can then visualize the created effective fracture length, the same way
as we did above for the bilinear flow analysis.

638
HIDRAULIC FRACTURING EMERGING TECHNOLOGY

High Permeability Fracturing (Frac & Pack)


The rapid ascent of high permeability fracturing (commonly referred to as HPF or frac
& pack) from a few isolated treatments before 1993 to some 300 treatments per year
in the USA by 1996 (Tiner, et al., 1996) suggests that it is one of the major recent
developments in petroleum production.

There is no theoretical difference between low and high permeability fracturing. In both
cases, the same dimensionless fracture conductivity, CfD =1.6 is optimal. When
translated into real dimensions, however, the resulting length and width requirement
will substantially differ in the two cases. A given amount of proppant placed into a high
permeability formation will result in a shorter and wider fracture than would be
obtained if the same amount of proppant was placed in a low permeability formation.
While this is theoretically true, in practice the fracture should bypass near-well damage
to be able to fulfill its function. Bypassing near-well damage is a necessary condition
for appreciable production enhancements from high permeability fracturing.

The key to high permeability fracturing is the tip screenout (TSO) technique, which
arrests lateral fracture growth and allows for subsequent fracture width inflation and
packing. The result is shorter but exceptionally wide fractures. While in traditional,
unrestricted fracture growth an average fracture width of 0.25 in. would be considered
normal, in TSO treatments, widths of one inch or even larger are common.

While production enhancement is of primary importance, there are a number of


additional reasons to consider fracturing in high-permeability formations:

Improve communication between reservoir and wellbore

Bypass formation damage

Reduce pressure drawdown

Increase back-stress on the formation

Control sand deconsolidation

Reduce fines migration and asphaltene production

Reduce bottom water coning

The Evolution of High-Permeability Fracturing

Frac & Pack treatments did not necessarily originate as an extension of hydraulic
fracturing, although they borrowed heavily from established techniques. Rather, they
began as a means of sand production control. Downhole techniques for controlling

639
surface sand production are of two basic types: sand exclusion and sand
deconsolidation control.

Sand exclusion methods involve the use of filtering devices such as screens and gravel
packs. These techniques do not prevent sand migration in the reservoir, and so fines
migrate and lodge in the gravel pack and screen, causing large corresponding skin
effects. Well performance progressively deteriorates and often is not reversible with
matrix stimulation treatments. Attempts to stem the loss in well performance by
increasing the pressure drawdown often aggravate the problem further, and may even
lead to wellbore collapse.

A more robust approach is the control of sand deconsolidation (i.e. prevention of fines
migration at the source). It is widely perceived that frac & pack combines the sand
exclusion function of a traditional gravel pack with the more important function of sand
deconsolidation control, by mating with the formation in its relatively undisturbed state
and reducing fluid velocities or "fluxes" at the formation face.

The present trend in frac & pack indicates a marked departure from the heritage of
gravel packing, incorporating more and more from hydraulic fracturing technology.
This trend can be seen, for instance, in the selection of fluids and proppants.

Materials

When sizing proppant for a frac & pack treatment, we have to address the two
competing goals of optimizing proppant size to exclude sand, or maximizing proppant
size to ensure adequate fracture conductivity. Early frac & pack treatments used sand
sizes and "clean" fluids that were typical of those used in gravel packing. More
recently, proppant sizes that are characteristic of those used for hydraulic fracturing
(20/40 mesh) have become dominant.

Fluid selection for HPF has always been driven by concerns with damaging the high
permeability formation, either by filter cake buildup or (especially) by polymer
invasion. Most early treatments were carried out using HEC, the classic gravel pack
fluid, as it was perceived to be less damaging than guar-based fracturing fluids. This
perception continues to be a matter of debate, and while many operators continue to
use HEC fluids, the fluid of choice is increasingly borate-crosslinked HPG. The
increasing application of crosslinked fracturing fluids also illustrates the trend of
departing from gravel packing traditions and converging to hydraulic fracturing
practices.

Perforations

Common practice for perforating is to shoot the entire target interval with high shot
density (e.g., 12 shots per foot) and large-diameter holes. Some operators prefer to
perforate the middle of the target zone only (possibly modifying up or down based on
stress contrast) with a limited number of 0o or 180o phased perforations. Both
underbalanced and overbalanced perforating is in use.

Job Execution

640
The amount of proppant placed into the formation is usually considerably less in a frac
& pack treatment than in massive hydraulic fracturing.

Solvent or other scouring pills are commonly circulated to the bottom of the work
string, and then reversed out to remove scale, pipe dope or other contaminants prior
to pumping into the formation. Several hundred gallons (e.g., 10 to 25 gallons per
foot) of 10 to 15% HCl acid will then typically be circulated or bullheaded down to the
perforations and allowed to soak, thereby improving communication with the reservoir
by cleaning up the perforations and dissolving debris in the perforation tunnels.

The vast majority of HPF treatments have been performed with the mechanical sand
control equipment in place. Recently there is a trend toward screenless HPF.

While the overall volumes pumped are relatively small for a fracture treatment, the
high rates (20 bbl/min is common) and high proppant concentrations (up to 16 or 18
lbm added to one gallon base fluid) require high horsepower. Reliable mixing and
blending equipment is now available to handle the various fluid and additive
specifications of HPF, including very low to very high proppant concentrations and
slurry rates.

The typical tip screenout design involves a low pad volume. Often the proppant
addition starts in the very first minutes of pumping. A typical feature of these
treatments is the packing stage. Near the end of the treatment, when the high
proppant concentration hits the perforations, the slurry rate is gradually decreased.
Consequently, the proppant packs the near wellbore region and provides excellent
communication toward the fracture tip.

Fracturing Horizontal and Deviated Wells

It is now commonly accepted that good completion and stimulation practices are as
essential to the success of a horizontal (deviated) well as they are to that of a vertical
well. It is therefore no longer unusual to see hydraulic fracturing treatments performed
on horizontal wells. Also, it has become clearly understood that the relative positions of
the fracture and the well have a profound effect on the performance.

Fracture-Well Configurations

For vertical wells, there are only two basic well-fracture configurations: a horizontal
(pancake) or a vertical fracture. Horizontal wells, on the other hand, have three basic
configurations. In a shallow formation, a horizontal well can also be intersected by a
horizontal fracture. If a horizontal well is drilled in a deeper formation and
subsequently fracture simulated, a vertical fracture will be created. Depending on the
wells relative orientation to the minimum horizontal stress direction, this fracture will
be either longitudinal (lateral axis of the fracture coinciding with the wellbore axis) or
transverse (lateral axis of the fracture perpendicular to the wellbore axis). The number
of transverse fractures intersecting a well can be substantial; three to five transverse
fractures are not exceptional). ( Figure 1, Possible well-fracture configurations). A
longitudinal fracture provides much more fracture-well communication than a
transverse fracture. Multiple transverse fractures, however, may provide more fracture
surface.

641
Figure 1

(When we list these basic fracture-well configurations, we understand that these are
idealized limiting cases. In practice, intermediate configurations may also occur.)

Once a well is drilled, there is little or no way to influence the well-fracture


configuration. Therefore, knowing the minimum horizontal stress direction and
directing the well according to it should be of major concern prior to drilling.

To understand the performance of these well-fracture configurations, one has to keep


in mind that the horizontal anisotropy of the permeability field usually follows the
stress anisotropy. The created fracture will be not only perpendicular to the minimum
horizontal stress but also perpendicular to the minimum horizontal permeability.
Therefore, hydraulic fracturing cannot do too much to overcome the limitations posed
by a small horizontal permeability. On the other hand, hydraulic fractures are
extremely useful in overcoming limitations posed by small vertical permeability or
unusually large pay thickness.

The optimal configuration depends on the permeability contrast (both horizontal and
vertical), and on the pay thickness to well length ratio. For high permeability, the
longitudinal configuration is advantageous, especially because there is no need to
create large fracture width (in contrast to high-permeability fracturing of vertical

642
wells). For low permeability formations, multiple transverse fractures are likely to
perform better.

Turns and Twists

One of the main problems with a transverse fracture is that the fracture initiates in
parallel to the well axis, and then turns to be aligned to the far field stress. The turning
and twisting near- wellbore part of the fracture may cause excessive fracturing
pressures during job execution; even worse, it may cause a choking effect during
production. The phenomenon is often called near- well tortuosity.

Fracturing deviated wells is a great challenge because it involves several unresolved


issues and a variety of possible scenarios. A careful perforation strategy favoring the
far field stress direction might be useful if fracturing a deviated section is unavoidable.

Operational Issues

The uninvited effects of tortuosity can be partly cured by using a special technique
known as breakdown with proppant slugs. The technique involves injecting limited
volumes of proppant slugs during pumping of the pad. The proppant slugs act similar
to abrasive drilling, preparing a wider channel for the subsequent slurry.

Another suggestion for avoiding tortuosity problems for transverse fractures is to use
very short perforation intervals (1.5 times the well diameter.)

Multiple transverse fractures need zonal isolation and several sub-treatments. The
creation of a longitudinal fracture can be also accomplished in several isolated steps.

643
Units, Conversion Factors and Formulas

Multiplier Prefix name Symbol

109 giga G

106 mega M

103 kilo K

10-3 milli m

10-6 micro

( 0.1 deci d)

( 0.01 centi c)

Name, Basic Quantity Fundamental Basic Unit


Dimension

Length, L [L] m

Time, t [T] s

Mass, m [M] kg

Temperature, T [] K

Electric current, i [I] A

Quantity Definition Dimension Unit Other


Form

Velocity, v L/t [LT-1] ms-1

Area, A L2 [L2] m2

Volume, V L3 [L3] m3

Density, M/V [ML-3] kgm-3

Acceleration, a v/t [LT-2] ms-2

Force, F ma [MLT-2] kgms-2 N

Pressure, p F/A [ML-1T-2] kgm-1s-2 Pa

Compressibility, (-V/V) / p [M-1LT2] kg-1ms2 Pa-1

644
ct

Work, W LF [ML2T-2] kgm2s-2 J

Flow rate, Q V/t [L3T-1] m3s-1

Power, P W/t [ML2T-3] kgm2s-3 W

Heat capacity, cp W/(mT) [L2T-2-1] m2s-2K-1 J/(kgK)

Shear rate, v/L [T-1] s-1

Viscosity, (F/A) /
[ML-1T-1] kgm-1s-1 Pas

Permeability, k v/ [L2] m2
(p/L)

Conversion Factors

Inch x 0.0254 = m

Ft x 0.3048 = m

Mile x 1609.3 = m

Acre x 4046.9 = m2

Hectare x 1104 = m2

Gal x 3.785410- = m3
3

Bbl x 0.158987 = m3

Liter x 0.001 = m3

lbm x 0.45359 = kg

lbf x 4.4482 = N

psi x 6894.8 = Pa

bar x 1105 = Pa

Atmosphere x 101325 = Pa

mm_of_Hg x 133.32 = Pa

deg_R x 1.8 = K

hp x 745.70 = W

645
md x 9.869210- = m2
16

cp x 0.001 = Pas

BTU x 1055.1 = J

min x 60 = s

hour x 3600 = s

Day x 86400= s

Hydrostatic Pressure: p = g h

Darcy's Law: v = -(k/)(p /x)

Pipe Mechanical Energy Balance: p =g z + (/2)u2 +2fF u2L/D

(p: pressure, : density, h: height, v: (linear, darcy) velocity, k: permeability, :


viscosity, x: length coord; z: elevation; u: linear velocity, fF: Fanning friction factor, L:
pipe length, D: pipe diameter)

Hawkins formula, Cylindrical permeability damage:

Quantity

Universal Gas Constant, R 8.3144 J/(molK)

Acceleration of Gravity, g 9.8066 m/s2

Water Freezing Point (105 Pa) 273.14 K

Max Water Density 999.973 kg/m3

Air Density (273 K, 105 Pa) 1.293 kg/m3

Pressure Drawdown and Productivity Index

Vertical Well + Damage

Steady-state, oil

Pressure drawdown:

646
Productivity Index,

(field): J= kh/(141.2B)

Pseudosteady-state, oil

Pressure drawdown:

Productivity Index,

Infinite-acting transient, oil

Pressure drawdown:

Productivity Index,

where

647
(Field: )

( = 0.5772 )

Pseudosteady-state, gas (square approximation)

Pressure drawdown:

Productivity Index,

(Field: )

Vertical Well + Vertical Fracture


Pseudosteady-state, pseudo-radial approximation, oil

Pressure drawdown:

Productivity Index,

648
where

where

Infinite-acting transient, oil

Pressure drawdown:

Productivity Index,

where

(Field: )

(available in form of graphs and computational algorithms)

649
SAND CONTROL

SAND PRODUCTION IN OIL & GAS WELLS

Causes of Sand Production

Sand production occurs when the destabilizing stresses at the formation face exceed
the strength of the natural arching tendencies and/or grain-to-grain cementation
strength. Destabilizing factors include the following:

1. An increasing disparity between the maximum and minimum grain-to-grain


compressive stresses, which tend to cause grain slippage and, therefore,
collapse of the perforation tunnel, wellbore or cavity. This can result from:

high drawdowns;

pressure depletion;

wellbore deviation;

pressure surges during rod pumping.

2. A significant pressure gradient across the sandface, which tends to drag the
sand into the wellbore. This can result from:

high fluid viscosities (>50 cp,), especially in heavy oil wells


where viscosities may exceed 1000 cp;

turbulence in the pore throats in gas wells and high-rate or high


gas-liquid ratio (GLR) oil wells;

formation damage and plugging of perforations and pore throat


areas.

3. Increasing water saturation and water cuts, which destabilize the sand as a
result of:

reduction in the cohesive strength and intergranular friction

mobilization of the fines from around the grain-to-grain contacts

chemical reaction with the natural cementing materials,


especially if salinity changes occur with water breakthrough

650
operators increasing the gross production rate and draw-down
to compensate for decreasing relative permeability to oil and oil
production volumes

increasing drag forces on the grains due to movement of the


wetting phase (i.e., the connate water)

The factors tending to prevent sand production include:

1. The amount, strength, and nature of the rock cementation.

2. The shape and arrangement of the sand grains, which will determine the
intergranular friction.

3. The degree of compaction, which is generally proportional to the overburden


loading (or depth) and inversely proportional to the initial reservoir pressure.

4. The behavior of the rock after initial failure in terms of:

plastic deformation of the cavity wall;

natural arching of the unconsolidated or failed sand around the


perforation tunnel.

Most sand problems occur with young, poorly cemented rocks, especially at shallow
depths, in heavy oil wells; in highly overpressured and geopressured zones; and in
certain high-rate producers. While many companies have a depth limit below which
they rarely install sand control, this varies from field to field and especially from one
geological environment to another, and may be as little as 1500 ft (460 m) to as much
as 15,000 ft (4600 m).

It should also be apparent that sand production is often rate-sensitive, since The rate
determines the drawdown (or grain stress contact), drag forces, and, in some cases,
the water saturation. A maximum sand-free rate can often be established by slowly
varying the production rate. This changes with production conditions and with reservoir
depletion. However, offtake rate (or drawdown) control is often used as a temporary
means of sand control. In many cases, the required rate is economically unattractive,
or will rapidly become so as reservoir pressure depletion and/or water influx
necessitate further bean-back.

Classification of Sands and Their Failure Tendencies

The risks of sand production can be established qualitatively by a combination of


analogy to geologically similar areas, core inspection and testing, log interpretation,
well tests, and field experience. An understanding of the rock types susceptible to sand
production is a convenient starting point:

Rock Classification System

651
Term Brinell Geological Equivalent
Hardness
(BHN)
(kg/mm2)

unconsolidated <2 no cementing material

partially (or 2-5 pieces easily crushed with fingers


semi-)
consolidated

friable 5-10 pieces crushed when rubbed between fingers

consolidated 10-30 pieces can only be crushed with forceps

hard >30 pieces cannot be broken

medium 30-50 with forceps


hard

hard 50-125

very hard >125

The wellsite geologist has traditionally categorized rocks into these groups based on
visual and manual tests. To achieve a less subjective division for engineering design
work on well stimulations, Van der Vlis (1970) introduced the concept of classifying
rocks according to their Brinell Hardness (BHN).

Brinell hardness is a measure of a material's resistance to deformation, as determined


by forcing a 3.5 mm steel ball intoa polished suface of the material under a specified
static load!

Since serious sand production is predominantly associated with the unconsolidated,


partially consolidated, and friable sands, it is pertinent to discuss them in more detail.

Unconsolidated Sands (BHN kg/mm2)

Worldwide, the bulk of sand production problems are found in relatively shallow (i.e.,
less than 8000 ft [ 2400 m] ), young ( Miocene-Recent ), unconsolidated rocks.
Unconsolidated sands can conveniently be subdivided, in their degree of natural
cohesion, into quick-sands and competent uncemented sands.

Quicksands have very little cohesive force or compaction to hold them together. It is
difficult to drill through this type of formation as the sand readily collapses into the
wellbore. Sand production begins immediately with fluid production, and sand flows
readily with oil, water, or gas.

Competent uncemented sands result from the increase of in-situ stress with depth,
which provides the unconsolidated sand with a degree of natural cohesion, as a result
of internal friction. The wellbore will stand open through such an interval, even in

652
deviated wells, although the sand is easily washed away during drilling, completion, or
production operations. Conventional coring of such sands is difficult. Rubber sleeve
coring is sometimes effective, provided pump rates are controlled. A simple perforated
completion can sometimes be used in such a formation, although severe sand
production is often experienced if the produced fluid has a high viscosity (e.g., heavy
oils) or once water breakthrough occurs. Some operators use sand control in this type
of formation, at least to some predetermined depth, as a standard practice. Without
sand control, some degree of continual sand production can be expected, with sand
concentration being a function of production rate and drawdown.

Partially Consolidated Sands (BHN 2-5 kg/mm2)

Partially consolidated sands have some cementing agents, but only weak, unconfined
compressive strengths. A core can usually be taken from this type of formation with a
conventional core barrel, but it crumbles easily. An openhole completion is possible,
but the hole tends to collapse and sand itself off when production conditions change.
Similarly, perforated completions are initially stable, but as conditions change, cavities
or pockets cave in periodically, especially after sudden rate changes. This causes slugs
or clumps of sand to come into the wellbore that may fill the rathole or form bridges in
the tubing. Analyses of sand content in produced fluids may show large variations from
day to day as the sand is produced sporadically.

Friable Sand (BHN 5-10 kg/mm2)

The third type of potentially troublesome sand is a friable sand, which is well cemented
and easily cored. Cores appear strong and competent and do not look as though they
would give sand problems. However, under the combination of increasing grain-to-
grain stress, erosion, and changes in saturation, the cementation may break down and
permit sand production to occur. It is common for this type of formation to produce
sand for a few days or weeks after completion, and then for sand production to
diminish to only a trace, or cease. With significantly reduced pore pressure or water
influx into the well, sand production may recur, especially at high production rates,
where turbulence becomes significant in the near wellbore area. However, postfailure
stabilization effects make it extremely difficult to predict.

The behavioral divisions in the above classification are somewhat artificial, but
represent a decreasing probability of sand production. Consolidated and hard rocks
rarely produce sand, except during initial cleanup, and testing of poorly completed
wells. (Frac stimulation is often used in these harder rocks and frac sand is sometimes
produced back into the wellbore. This is the result of poor fracturing procedures rather
than a sand control problem.)

It is important to recognize that sand production during testing may not indicate a high
risk of continual sand problems, even when it results in sandoff of the DST (drillstem
test) tools. Natural wellbore cleanup and stabilization processes involve production of
perforation debris, fines, and minor quantities of sand. This process is exacerbated
where poor completion and well cleanup practices result in most of the production
evolving from a limited number of perforations. The resulting high initial drawdowns
cause cavity enlargement and sand influx. A similar effect is sometimes seen when
very high drawdowns are used for perforating.

653
Basic Rock Mechanics
An appreciation of rock mechanics principles is useful in trying to understand and
predict sand problems and for planning sand control procedures. Sand failure is a
formation collapse phenomenon, while fracturing and/or mud losses result from an
induced tensional failure. For failure prediction, we are concerned with collapse. But we
must also understand fracturing as it relates to sand control treatments. Geertsma
(1961), and Roegiers (1987) presented an excellent reviews of rock mechanics for
oilwell operations. The SPE Monographs of Acidizing (Williams, Gidley, and Schechter
1979) and Hydraulic Fracturing (Howard and Fast 1970) also address the subject.

In-Situ Stresses

In the ground, undisturbed rock materials are in an equilibrated state of triaxial


compression because of the weight of the overburden. This overburden creates vertical
and horizontal stresses, and it is these compressive stresses that hold the rock
together ( Figure 1 , in-situ stresses).

Figure 1

654
In tectonically relaxed areas, where normal faulting predominates, the weight of the
overburden material can be assumed to be the major principal stress. It can be
calculated by integrating the formation density log from the surface down to the point
of interest, or estimated by:

(1)

where:

v = total vertical stress, psi

= average rock density, lb/ft3

D = depth, ft

The overburden stress gradient is generally approximated as 1.1 psi/ft..

Effective stress is reduced when the formation has porosity and contains fluid. Part of
the overburden load is supported by the pressured fluid and, therefore,

(2)

where:

= density of overburden, lb/ft3

p = pore pressure at depth D, psi

= effective vertical stress, psi

D = depth, ft.

= 1- (Cr/Cb); this efficient is not included in some texts

Cr = rock matrix compressibility, psi-1

Cb = bulk compressibility, psi -1

This equation indicates that vertical matrix stress is increased by declining reservoir
pressures, and is reduced by abnormally high pore pressures resulting from rapid
burial or other geological activities.

Effective horizontal compressive stress can be calculated by

655
(3)

where:

= Poisson's ratio

v = total stress

Poisson's ratio may have a value of from 0.1 to 0.4, depending on the type of
formation. A U.S. Gulf Coast sandstone would typically have a value of 0.33, while a
midcontinent limestone may have a value of 0.27 (Allen and Roberts 1978). It follows
that in a relaxed tectonic area, the effective horizontal matrix stress will be about one-
half to one-third the effective vertical stress.

Thus it is one of the horizontal stresses shown in Figure 1 that is the least principal
stress. Under sufficient hydraulic pressure, a vertical fracture will be created
perpendicular to this stress. This fracture is held open by fluid pressure opposing this
horizontal stress. Thus the fluid pressure at which an induced fracture will open or
close (commonly termed the fracture [propagation] pressure) is a measure of the least
principal stress. This is also called the initial shut-in pressure (ISIP), and can be
determined from casing leakoff or formation breakdown tests. In tectonically relaxed
sand/shale sequences it can be approximated by

(4)

or

ISIP ~ 0.5D +0.5p (5)

It is important to note the gross generalizations used in deriving this expression and to
apply the results with appropriate caution. It is strongly recommended that the
minimum horizontal stress be established in each field from frac data or by conducting
formation breakdown tests. However, the above approach is useful in transferring the
results of such tests to other depths and pressure regimes.

Wellbore Area Stresses

The introduction of a borehole into the undisturbed rock distorts the stress field in the
general vicinity of the wellbore and affects the pressure required to initiate a fracture.
Hubert and Willis (1957) have shown that classic elastic theory can be applied to
estimate the magnitude of the resulting stress, provided corrections are made for pore
pressure. Most failure prediction programs are based on the use of this technique to
predict the radial, tangential, and axial stresses at the borehole, perforation, or cavity
wall for various geometrical configurations. The simplest form of such equations for a
vertical well with non-penetrating wellbore fluids is

656
radial stress = r = pw - p (7)

tangential stress = (8)

axial stress = (9)

where:

= Poisson's ratio

= 1 - (rock matrix compressibility/bulk compressibility)

h = minor horizontal total stress (least principal)

H = major horizontal total stress (intermediate)

v = vertical total stress (maximum principal)

p = resevoir (pore) pressure

pw = wellbore pressure

( is omitted in some texts)

From these equations, it can be seen that as the wellbore pressure decreases, (1) the
support of the grains in the radial direction decreases, (2) the tangential grain-to-grain
compressive stress increases, and (3) the difference between the maximum and
minimum principal stresses on the sand face therefore increases. These effects tend to
destabilize the sand and cause shear-type failures.

In reality, of course, the situation is more complex, since the pore pressure and
effective stresses in the wellbore area vary radially from the wellbore and are not
independent of the wellbore pressure, as assumed in this simplified case. Moreover, in
general, it is perforation or cavity stability that we are interested in, rather than the
entire wellbore. However, the simplified equations are useful for illustration of the
principles involved.

Failure Criteria

Although a number of failure criteria have been proposed for rocks under stress, by far
the most universally accepted is the Mohr-Coulomb concept that failure occurs
predominantly as a result of shear.

As indicated in Figure 2 (Stress analysis using Mohr's circle ), the Mohr's circle is a
convenient way of presenting how the stresses vary on planes, intersecting the
maximum and minimum principal stresses.

657
Figure 2

The radius of the Mohr's circle is equal to the maximum shear stress ( max):

max = = RM (10)

where:

= shear stress

1 = maximum principal stress

3 = minimum principal stress

RM = radius of Mohr's circle

If a series of Mohr's circles are constructed to represent the failure conditions of similar
materials under varying triaxial loads, then it should be possible to construct a failure

658
envelope tangential to these circles ( Figure 3 , Failure envelope based on Mohr's
circles).

Figure 3

The Mohr failure envelope is an empirically derived curvo-linear function that is not
represented by any general mathematical formula. Typically, it has a parabolic shape
that rapidly increases from a low tensional strength and levels off as the compressional
loading increases. The Coulomb failure criteria ( Figure 4 ,

659
Figure 4

Coulomb failure envelope) is just a special case of the Mohr's envelope that defines the
critical shear stress in a cohesive rock in terms of the angle of internal friction ( ) and
the natural cohesion (C):

crit = tan + C (11)

The failure envelope for a given reservoir can be determined by laboratory


measurements on core material or triaxially loaded formation sand. From such
measurements it has been found that the angle of internal friction () is normally very
close to 30, particularly at high confining stresses.

Correlations are available relating uniaxial compressive strength of a given lithology to


Young's modulus (e.g., Deeve and Miller 1966, Coates and Denoo 1981). Similarly,
several correlations are available for estimating Young's modulus based on sonic
transit time measurements, or Brinell hardness measurements on the core material.
Thus a first estimate of the failure envelope can be developed based on log data
and/or simple core tests, using these correlations and assumptions.

It should be recognized, however, that the failure envelope itself may vary with the
rock condition and be affected by the coring and handling process. In particular,
changes in the water saturation will affect the rock strength. If the core is allowed to
dry out, the cementation will often deteriorate and the rock may crumble into loose
sand. Increasing water saturation may also decrease the rock strength by reducing
internal friction coefficients and/or by dissolution of the cementing material. While this
is a problem for predicting sand failure (i.e., compressional loading) it is much less
serious for predicting fracturing, since all rocks have very low tensional strength,
particularly on the macro scale.

Application of These Rock Mechanics Concepts

In theory, stress analysis techniques can be used to predict the loading conditions at
the wellbore and/or perforation wall, and these can be compared with the rock failure
envelope. A set of stress conditions falling inside of this failure envelope can be
considered to be safe, and those outside to represent increasing risk of failure.

This theory is the basis of the Schlumberger MECPRO or sand-strength log. The sonic
and density logs are used to calculate the elastic moduli of the formation, and hence a
failure envelope. This is compared with the computed critical wellbore stresses to
develop a prediction of the maximum allowable draw-down prior to sand failure
(Coates and Denoo 1981).

Several other investigators have used similar concepts. However, as with other
strength-of-materials situations, there is confusion between the use of the yield,
ultimate strength, and actual failure points in defining the boundary of safe operating
conditions. From soil mechanics, it is well known that unconsolidated sand, under

660
triaxial compression, will deform plastically prior to failure. It is also commonly
accepted that consolidated rocks, under high confining stresses and temperatures,
show considerably increased ductility compared with surface conditions. It is therefore
not surprising that, as Geertsma points out (in EUR 38, 1961), in collapse situations,
"substantial yielding around the borehole is required before actual borehole failure
materializes." Thus, the theory of elasticity, as applied to porous rocks (poroelasticity),
is too conservative for defining sand failure. Even where failure does occur, the
production of the loose sand is controlled by natural bridging and arching tendencies
around the perforation tunnel and at the cavity wall.

In consolidated rocks, mining experience shows that arching not only occurs in the
loose material that forms the zone of disturbance at the inner boundary of the plastic
zone, but also within the plastic and elastic regions, thereby reducing the effective load
on the weakened rock. Thus, initial failure can be expected to be followed by a short
burst of sand production, which rapidly decreases as natural arching limits the growth
of the cavity. This phenomenon of postfailure stabilization is frequently observed in the
field. Although a joint industry project at the Colorado School of Mines has examined
the stability of arches in unconsolidated sands, we do not have an adequate theory for
predicting the postfailure stabilization of partially consolidated and friable rocks, or the
conditions at which the cavity reaches a state of incipient failure. This places a
limitation on the application of the rock mechanical theory in a quantitative manner,
especially for partially cemented sands.

The principles are, however, useful in developing other predictive tools:

1. Increasing drawdown and depletion will increase the shear loading, due to an increasing
difference between the maximum and minimum principal stresses.

2. Higher rock strengths and/or higher average initial stress conditions will
result in a greater region of stability (safe region) below the failure envelope.
The two effects are sometimes termed the cementation or formation strength
index, and are correlated with sonic transit time data in the formation and/or
surrounding shales (lower sonic transit times indicating stronger and/or more
compacted rock).

3. In addition to the safe and failure regions, there is likely to be a risk region in
which some zones produce sand while others do not (or show intermittent,
minor sanding).

4. Water production will probably increase the area of the risk and failure
regions (i.e., a downward displacement of the failure envelope).

Figure 5 shows the type of relationship that can be expected between cementation indicator (e.g.,
rock type, hardness, sonic transit time, log-derived strength index) and drawdown.

661
Figure 5

This concept was first proposed by N. Stein of Mobil in 1972 (SPE 3499) and has subsequently
been used by several other experts. It is particularly useful as a method of analyzing data from
existing perforated completions and extended well tests. It is used to develop empirical criteria for
determining where sand control may be required on new wells. In some cases, instantaneous
drawdown (pr - pwf) is found to be more important than cumulative drawdown (pi - pwf), especially for
higher-rate wells producing from partially cemented sands.

Sand Failure Prediction


The decision whether to use sand control in a given zone is based on the production
engineer's ability to predict the probability of unacceptable sand influx in a natural
completion.

In general, production engineers and sand control specialists rely on a combination of:

analogy;

core inspection and testing;

log interpretation;

662
special well tests;

field experience.

Analogy

Usually, the first step is to consider the field characteristics, including:

geological age;

depth of zones;

environment of deposition;

pressure regime;

development concept;

fluid type and viscosity;

porosity and permeability;

primary cementation material.

These are reviewed and compared with similar fields elsewhere in the region and/or the world.

It is well known that severe sand production can be expected from:

young, shallow, poorly cemented, dirty, deltaic sandstones, especially under depletion
drive conditions;

heavy oil wells;

highly overpressured, poorly cemented sands.

Core Inspection and Testing

Poor core recoveries are often an early indication of potential sand problems, provided
good coring techniques were used. Similarly, if the core falls apart as it is retrieved,
the risks of sand production should be assessed.

However, it is important to remember that destressing of the core material during the
trip out of the hole tends to destabilize the formation and reduce the cohesion obtained
from grain-to-grain friction. Moreover, weak cements, such as clay, may be affected by
the drilling mud and by subsequent handling and storage.

An important step commonly omitted by many production engineers is physical


inspection of the core. If it appears to be reasonably well cemented and grains are not

663
easily rubbed off or broken, the chances of sand production during the early life of the
field are quite low.

Engineering measurements should be made on marginal core material. Where the rock
is very weak, special coring techniques should be used, such as rubber sleeve or
plastic sleeve coring, and the samples should be preserved by onsite freezing. If this is
not possible, some understanding of in-situ behavior can be obtained by compacting
the loose sand in a triaxial soils testing cell.

Typically, core tests should include:

thin section, x-ray diffraction, and scanning electron microscope (SEN) studies on
cementing materials

core flow tests, to measure the stability of the sand at various water
saturations and flow rates, and its sensitivity to salinity changes

uniaxial and triaxial compressive strength measurements on dry and water-


saturated cores

Brinell hardness measurements, to relate tested samples to the rest of the


reservoir section

The objective of this program is to define:


the type and amount of natural cementing material and its sensitivity to water

the amount of mobile fines that are present and their destabilizing effects on
the rock strength

the effect of flow and stress changes on core stability

the relationship of the tested samples to the untested core, and, via log
response, to any zones from which no core was recovered and/or to other wells

This laboratory data can also be used as input to theoretical rock mechanical analyses to
qualitatively explain the sand production behavior observed in DSTs and production wells.

Log Analysis

A great deal of work has been done in this area, especially by the major companies,
some of whom have proprietary in-house techniques. Schlumberger has developed its
mechanical properties log (MPL) and MECPRO logs, or, more correctly, analysis
techniques, in an attempt to predict sand failure (Schlumberger, 1989). Various papers
have been published on the use of the borehole-compensated sonic, full wave train
sonic, and sonic/density combinations for this purpose. However, in the final analysis,
all of these techniques are semiempirical with the thresholds having application to one
particular area or basin.

The most effective log analysisi-related methods appear to be variations of the


technique (published by Stein and Hilchie of Mobil (1972)) of plotting all test and

664
production data as a plot of sonic velocity (BHCS) versus drawdown to identify the
safe, risk, and failure regions ( Figure 1 , relationship between cementation indicator
and drawdown).

Figure 1

This is sometimes reduced to a simple rule of thumb that sand production can be
expected from any zone with a sonic transit time in excess of some locally determined
threshold, usually ranging from 95 to 110 s/ft.

Of the other sand production indicators, the most commonly used is that developed by
Tixier (1975) using log-derived factors of shear modulus (G) divided by bulk
compressibility (Cb):

G = 1.34 x 1010 (psi) (12)

665
= 1.34 x 1010 (psi) (13)

where:
G = shear modulus (psi)

Cb = bulk compressibility (psi-1)

b = bulk density (gm/cm3)

tc = compressional transit time (s/ft)

The constants A and B are related to Poisson's ratio by the following equations:

(14)

(15)

where:
= Poisson's ratio
Several techniques have been proposed for estimating Poisson's ratio from log response. However,
G/Cb is not very sensitive to this factor.

= .2 .3 .4

A= .38 .29 .17

B= .5 .62 .78

AB = .19 .18 .13

For the U.S. Gulf Coast, the threshold above which sand production should not be
expected has been variously stated at between 0.7 and 0.8 x 1012 psi2 (Tixier 1975).
Initially, this was claimed to have fairly universal application; however, further
experience showed this to only be true if the depth versus sonic trend was similar to
the U.S. Gulf Coast.

Special Well Tests

Properly conducted well tests with effective surface, and possibly downhole, sand-
detection equipment are the best method for proving that continuous sand production
occurs at initial reservoir pressures. However, the test and completion equipment must
permit the critical draw-downs and rates to be achieved. The test must also be
continued beyond the point of initial sand production to prove that the sand influx rate
does not rapidly decline and/or that, even if it does, sporadic sand production will not

666
occur thereafter. Such sand tests should include both repeated step-rate tests and
extended flow periods under mild sanding conditions.

It is important that the well is properly completed with an effective perforation and
cleanup program so that localized skin effects and turbulence do not induce a sanding
condition that would not otherwise occur.

Some operators have undertaken high drawdown testing of a limited number of


perforations in an attempt to use geometric skin effects to reduce the wellbore
pressure in order to duplicate the stress conditions that can be expected after reservoir
pressure depletion. However, since this involves unrealistic flow rates, turbulence, and
near wellbore stress and pressure gradients, many experts question the validity of
such tests in proving that future sand production can be expected. Negative results
may, however, prove that the sand will remain stable.

Similarly, it is usually difficult to prove the effects of water breakthrough. However,


some operators include tests on water zones and backflow of well after injection tests
as part of their well evaluation programs.

On the other hand, the effects of depletion and/or future water breakthrough are not
usually critical development planning issues. It is generally more practical to perform
any future sand control measures via a workover, provided such operations can be
deferred for some years and are undertaken before excessive cavity development or
pressure depletion occurs.

Field Experience

In the final analysis, the decision on the sand control requirements has to be based on
experience in the area with natural, perforated completions. However, adjustments
must be made for differences in completion method and efficiency, hole deviation, well
depth, reservoir pressure, depletion and drawdown, water production expectations,
and development plans.

667
COMPLETION & PRODUCTION PRACTICES

Improved Completion Efficiency

Although not traditionally considered primarily as a sand control method, high-density,


underbalanced perforating (in clean, nondamaging fluids), along with properly installed
openhole completions, can reduce sanding tendencies in marginal sanding situations.
This is most clearly apparent in comparing well performance of openhole and
conventional cased-hole DSTs, where sanding is common, with the performance of
production wells and/or DST conducted with tubing-conveyed perforating (TCP) on the
same reservoirs.

The reason for this is that overbalanced perforating (and mud-cake removal in
openhole DSTs) generally results in less than 10% of the interval being properly
cleaned up, at least initially. The well is then subjected to high geometric and
mechanical skins, resulting in a rapid pressure and stress gradient into the wellbore
over the productive intervals. Moreover, these intervals may be subject to turbulence
as a result of the high production rate per unit area. These are all destabilizing effects
on the sand face. In some cases, this situation is exacerbated by ill-conceived, poorly
diverted mud acid jobs, aimed at removing the damage, but resulting in over-
acidization and removal of the clay cementing materials from the zones that are
already open. The resulting sand influx is both avoidable and nonrepresentative of the
formation's stability. While such sand influxes should only be temporary, the volumes
involved are often quite sufficient to cause a sandoff or a buildup of fill over the
perforations.

This is a common problem in multizone completions, as the upper intervals are


generally perforated in overbalanced conditions. As a result of the productivity
impairment, sanding tendencies, and other problems often incurred during workovers,
many operators eshew the standard use of this type of completion.

Rate Control and Arching Effects

A consideration of the mechanisms of sand failure should make it apparent that


reducing the rate and drawdown should reduce sand influx. This is verified in the field,
as illustrated in Figure 1 (Production test indicating relationship between rate and sand
influx).

668
Figure 1

This figure shows a well being slowly beaned-up over a six-week period to conditions
of unacceptable sand production (in this particular case the limit had been set at 10
lb/1000 bbl (28 gm/m3), or 0.002%). A slight reduction in choke or bean size from
40/64 in. to 32/64 in. stops the sand. In time, as the reservoir conditions change
(week 20/21), excessive sand production occurs again and the well has to be further
beaned-back to 24/64 in..

Rate control is therefore only a temporary sand control measure that results in
considerable deferment of production compared to mechanical sand retention. In
general, the present value of this lost production far outweighs any cost savings from
deferring a more effective sand control technique. This is no justification, however, for
prematurely gravel packing a well in which the risks of sand production cannot
adequately be demonstrated, or where sand production might not occur for several
years. This is true not only because gravel-packed completions take longer and
therefore cost more in up-front, undiscounted dollars, but also because the gravel pack
will impede remedial work and prevent cheap through-tubing modification of the
completion interval (additional perforations, plug backs, selective stimulations, etc.).
Changing bean sizes is therefore a useful method of handling sand production on a
limited number of wells late in a field's life, or of temporarily controlling well behavior
pending installation of a permanent sand control system. It is also appropriate where
multiple stacked reservoirs with limited reserves make other options impractical or
economically unsound.

Selective Perforation

Many thick reservoirs have occasional unconsolidated streaks of abnormally high-


porosity, poorly cemented sand that will give sand problems if perforated. If these
streaks are in communication with the surrounding rock, their reserves can be drained
indirectly by selectively perforating the stronger zones. The problem typically is that
these weak streaks are often the main productive zone, acting as "natural propped

669
fractures" in providing a large surface area and conductivity to collect and transport
the hydrocarbon from the mass of the reservoir rock to the wellbore. The critical issue
then becomes whether the well can reach either its production target or design rate
without perforation of these high-permeability streaks. This has to be established by
testing in delineation wells and/or by simulation.

The key to the application of the selective perforation technique is to be able to identify
the potential sand-producing zones from log measurements. As a rough rule of thumb,
many completion engineers often use a sonic log threshold of 100 s/ft (325 s/m) for
identifying potential sand-producing intervals.

670
LINERS & SCREENS

Plain Screens

The simplest and oldest sand control method is a slotted liner run into the openhole or
into a cased hole after perforating. The slots are cut small enough that the produced
sand is retained by a bridging process of grain interference, and does not pass
through. Bridging occurs when more than one sand grain enters a slot at the same
time. The particles interlock and form an arch, or bridge. Filter theory and field
experience have shown that bridging can occur with grains down to a third of the
diameter of the pore throat, although the stability is poor with grains less than half the
pore size.

A slot width just larger than the tenth percentile (d10) of the formation sand particles is
typical, although wider slot widths up to (2 x d10) can be used in nonuniform sands
(Coberley 1937), while narrower slots should be used in very uniform formations. By
the tenth percentile or d10 designation, we mean that sieve size that would retain 10%
of the formation sand and allow 90% to pass through. This is illustrated in the sand
grain analysis in Figure 1 (Grain-size analysis for screen sizing).

Figure 1

671
Here the tenth percentile (d10) size is 785 microns, or 0.030 in. (about 22 U.S. mesh).

The example in Figure 1 applies to sand grains down to 262 microns (0.01 in.) and
would retain 77% of the sand. Finer sands and silts would initially pass through the
screen, but as a rubble zone or sand pack develops, they would be trapped by a
combination of filtration and bridging on the pore throats of the coarse material (
Figure 2 , Effective screen control on a week zone).

Figure 2

This type of design is obviously going to be most effective with very weak, medium- to
coarse-grained, moderately uniform sands that will quickly build up a highly permeable
rubble zone around the screen. It is commonly used for protecting pumps in shallow oil
wells and especially for heavy oil wells, where technical problems and economics limit
the effectiveness of other methods of sand control.

672
Where the sands are somewhat stronger and more uniform, the bridging process
becomes increasingly inefficient. At low sand influx rates, most of the grains simply
pass through the slots; therefore, it takes a long time for the rubble zone or pack to
build up and the slots may become sand-cut or plugged with grains that are close to
the slot size. Moreover, plugging of part of the screen leads to increased velocities and
erosion in other slots ( Figure 3 , Ineffective screen control on a moderate- strength
zone).

Figure 3

To avoid these problems, it is common for screens used in deeper, more uniform sands
to be designed with a smaller slot size in order to increase the chances of bridging. If
minor or intermittent sand influx is expected, screen openings must be small enough
to stop the sand over the majority of the grain size range, so that screen erosion can
be avoided before a rubble zone accumulates. Since it is not possible to saw-cut slots
below 0.012 in. (305 microns), wire-wrapped screens (WWS), prepacked screens, or
filter screens are usually used in this service.

673
The plugging tendency of square-cut slots ( Figure 4 ) can be reduced by undercutting
the slots to give a shutterlike,

Figure 4

or "keystone," profile ( Figure 5 ), so that any grains that do not form a proper bridge
on the outside of the pipe will be carried through the screen. Undercut slots are
available down to 0.02 in. (508 microns); below this, square-cut profiles or wire-
wrapped screens must be used.

674
Figure 5

Figure 6 also illustrates the various slotting profiles commonly used.

675
Figure 6

Vertical slots are the most common, being stronger in tension and less prone to shape
distortion caused by liner buckling. However, horizontally slotted pipe is sometimes
used because it is easier to "fish."

Another problem with the use of slotted carbon steel pipe is the potential for corrosion
or corrosion-erosion around the slots. This not only occurs during production of waters
containing CO2 or H2S, but also due to rusting (oxygen corrosion) prior to or during
installation.

Wire-wrapped screens are often used in place of slotted pipe since they provide:

an ability to control finer sands (spacings down to 0.006 in. or 152 microns
are standard, and 0.003-in. or 76-micron spacings are available); however,
even a 6-thousandth screen (0.006 in., or 152 microns) cannot filter out very
fine sand (0.0025 to 0.005 in., or 63 to 125 microns);

a greater inflow area, and therefore reduced pressure losses after partial
plugging;

an ability to use stainless or special steels for corrosion resistance;

greater mechanical strength when mounted on a perforated pipe base;

676
a more pronounced keystone or "undercut" shape to avoid plugging; however,
very fine wire-wrapped screens have a tendency to become plugged with mud
particles, unless the workover fluid is very clean or the screen is protected
during installation with an acid-soluble coating.

Slotted liners and plain screens are a relatively low-cost sand control method, and are
reasonably successful in medium- to coarse-grained, low-strength sands, especially
when set across an openhole completion. Since they may have to be replaced in time,
the outside diameter should be selected to permit wash-over operations. The OD
should therefore be at least 1 in. (25.4 mm) or, preferably, more than 1.5 in. (38 mm)
smaller than the production casing drift ID; and centralizers should be collapsible.

There are two major problems with the use of plain screens inside perforated casing:

There is a tendency for the screen to be cut out by sand blasting from a few
highly productive perforations during buildup of the rubble zone ( Figure 1 ,
screen cut-out caused by sand erosion). This can be minimized by slowly
bringing the well through the critical sand production rate.

Figure 1

The perforation tunnels become filled with formation sand, which imposes an
additional pressure loss due to the limited cross-sectional area over which linear
flow occurs through the cement and casing. In the higher rate oil wells and in

677
gas wells, this not only increases the Darcy skin factor, but may also introduce
a rate-dependent skin due to turbulence. These effects can be minimized by
perforating at high shot densities with large diameter holes.

Both of these problems become increasingly serious as flow rates and GLRs increase.
Therefore, while plain screens are often effective in low-rate oil wells (and particularly
in heavy oil wells producing from highly permeable, coarse sands), their performance
in flowing oil and gas wells is often disappointing.

One nonstandard application for screens is for keeping marginal, minor sand influx out
of sensitive wells (e.g., subsea producers, high-rate gas wells). In this situation, the
perforation or openhole is essentially stable and the screen acts as a downhole filter,
forcing what little sand is produced to be deposited in the rathole. This seems to be the
logic behind the screens used in the Frigg field in the North Sea, for example.
However, once the rathole fills up, or sand production becomes significant, gravel
packing is often needed. Prepacked screens are more suited to this type of service
than plain screens, especially in cased hole, perforated completions.

In general, plain screens are not recommended for highly productive or high GLR wells,
especially if the sand is fine grained, partially consolidated, and has low permeabilities
and/or is highly laminated. Plain screens are therefore not recommended for U.S. Gulf
Coast-type deltaic environments. This is because:

sand movement into the wellbore tends to cause permeability impairment by


intermixing of sand sizes and/or of sands and shales;

formation sand within the perforation tunnels is very restrictive to


productivity;

fine sands tend to plug the slots in the screen, necessitating a workover;

at high flow rates and/or low or intermittent sand influx rates, the screen
often erodes before a sand pack can form outside;

poor support of the formation can cause shale layers to collapse and produce
fines problems, zonal isolation problems, plugging of the rubble zone, and/or
permit casing distortion.

However, slotted liners or screens have been used successfully in many conventional
and thermal heavy oil wells and in other shallow oil and water wells.

Prepacked Screens and Filters

Where filter-sized liners are required, prepacked screens, or porous metal filters (such
as sintered bronze) are often used. Prepacked screens ( Figure 1 ) usually consist of an
inner and outer screen with the annulus between them being filled with a filter
material, usually gravel pack sand.

678
Figure 1

This sand is often consolidated with plastic resin to prevent rough handling from
disturbing the pack. Alternatively, the plastic-coated sand may be simply stuck onto
the inner screen (and the outer screen omitted), although this type of equipment is
more prone to installation damage. Another version utilizes cylinders of consolidated
sand mounted directly onto the perforated backup pipe and held in place by end plates
fixed with retaining screws.

The sand or other filler material acts as a downhole filter. Since the pore throats will be
only 15 to 25% of the sand grain size, it is possible to retain much finer sand than
could be achieved with a wire-wrapped screen. Moreover, the multiple layers of gravel
pack sand provide a depth filter, which is much less susceptible to erosional failure
because of built-in redundancy.

There are essentially four applications in which prepacked screens and filters are used:

as an alternative to slotted pipe or wire-wrapped screen, especially where a


significant proportion of fine to very fine sand has to be controlled;

679
as a precautionary downhole filter in situations where significant sand
production is not anticipated, but where even small amounts of sand cannot be
tolerated (e.g., high-rate subsea gas well completions);

where gravel packing is economically unattractive (e.g., marginal reserves,


exploration well DSTs, water injection wells), and sporadic or low-rate sand
influx rates and/or very fine sands are likely to render plain screens ineffectual
(e.g., in the U.S. Gulf Coast);

as the screen for a gravel pack where conditions involve a high risk of
ineffectual packing of the entire zone, or of "holidays" (void zones) developing
in the pack at some future date; such conditions might include:

a) quicksands, where the formation may slump before a


gravel pack can be installed, especially if the screen has to
be washed into place;

b) closely spaced gravel packs, where there is very little


room to place a gravel reservoir above the perforations of
the lower zone;

c) highly deviated wells, where it may be difficult to


effectively pack the high side of the hole, and where voids
caused by duning or node buildup will not effectively fill up
through pack settlement;

d) wells with highly variable permeability profiles, where


node buildup may limit effective dehydration of the entire
pack;

The sand used for a prepacked screen should be sized in the same way as gravel pack
sand. The screen OD should, if possible, permit washover operations and be well
centralized at 15-ft (5-m) intervals with collapsible bow-spring centralizers or millable
fin-type centralizers.

Since pore throat sizes in the prepack are very fine, it is obviously essential that very
clean completion fluid and production casing be achieved if plugging is to be avoided.
Similarly, careful handling and running operations are required, especially when
passing obstructions (BOPs, wellheads, liner tops, etc.).

As with arching and plain screens, a slow bean-up rate has been shown to give the
best productivity and minimize the risks of screen cut-out, especially in formations that
produce fines.

680
GRAVEL PACKING

Gravel Packing

Effective gravel packing is a downhole filtration process in which coarse frac sand, or
"gravel," is spotted against the formation face and held in place by a screen, slotted
liner or prepacked screen. To maximize productivity, minimal mixing of the formation
sand and gravel must occur; therefore the gravel and screen are sized for filtration
rather than bridging. The majority of the sand particles will be larger than the gravel
pore throats, and all the gravel will be larger than the retaining screen slot size (
Figure 1 , Principles of gravel packing).

Figure 1

There are two basic types of gravel pack: (1) the external, or openhole gravel pack;
and (2) the internal, or cased-hole gravel pack (CHGP). In the external gravel pack,
the casing is set above the top of the pay zone and the completion interval is then
underreamed to remove drilling damage and maximize the inflow area ( Figure 2 and
Figure 3 , Openhole gravel pack-underreaming hole followed by gravel placement).

681
Figure 2

The gravel pack screen is hung as an uncemented liner and the space between the
screen and the formation face is packed with properly sized gravel.

682
Figure 3

A variation of this technique, used where completion intervals cannot be preselected or


for recompletions, is to set the production casing to TD and then mill out a window
over the completion interval prior to underreaming. In another variation, the
completion intervals are underreamed after logging and the screens are incorporated
in the production casing, isolated by openhole packers. The casing is then stage
cemented and the production zones gravel packed.

No special drilling preparation is required for an internal gravel pack. The casing is
cemented conventionally at TD. The completion intervals are then perforated with high
density, large diameter perforations capable of penetrating the damaged zone. The
perforation damage is cleaned out and a cavity created behind the pipe by
backflushing, underbalanced perforating, backsurging, or perforation washing. The
screen is run and the gravel pumped into the void outside the casing, the perforation
tunnels and casing/screen annular area. The key to success is the creation of a cavity
behind the casing to permit effective packing of a high-permeability path through the
casing, cement, and damaged zone, out against the native formation ( Figure 4 ,
Schematic of internal gravel pack).

683
Figure 4

Although external gravel packs can be used for multizone completions by the use of
external casing packers or milled casing sections, their use considerably complicates
the completion operation. Therefore, most operators with multizone completions use
internal gravel packs in the upper zones.

Conversely, for thick pay zones, long openhole packs have historically been more
successful than long internal gravel packs. However, the availability of reliable
differential pressure valves for use in the tailpipe during combined squeeze and
circulation packing operations is improving the success of long internal packs.
Nevertheless, some operators prefer to limit case-hole packs to 150-200 ft. (45-60 m),
especially in deviated wells; and to stack separately placed 100 ft. (30m) packs where
longer intervals are to be completed.

In theory, external gravel packs result in better productivity than internal gravel packs,
especially at high rates and/or with viscous crudes. In practice, they are often more
difficult to plan and install effectively. Moreover, improved squeeze-packing techniques
and perforation/cleanout methods have greatly improved the success of internal gravel
packs since the mid-1970s. Cased-hole or internal gravel packs are now the accepted
industry practice, since they offer flexibility, selectivity, planning time, and cost/rig-

684
time savings. Figure 5 is a schematic of a single-zone external gravel pack completion.

Figure 5

In either type of gravel pack, the gravel must be properly sized to control sand and
avoid permeability impairment. Various formulas have been proposed over the years,
of which Sauciers (1974) is perhaps the best known: the median grain size of the
gravel (dg 50%) should be 5 to 6 times the median grain size of the formation sand (ds
50%). However, some operators use 20/40 gravel (425 to 850 microns) as a standard
for all packs.

Screen slots are sized marginally smaller than the grain size of the smallest gravel
pack sand particle. Normally, 0.012-in. (0.3-mm) slotted screen is used for 20/40
gravel, with 40-mesh sand theoretically having a minimum size of 0.017 in. (0.4 mm).

Care must be taken when speaking of mesh sizes, as there is a significant difference
between various standards. The ASTM standard is the one commonly used by the oil
industry, as shown below:

685
U.S. mesh size (ASTM)

U.S. mesh Partcl. diam. Partcl. diam. Aprox.


size range size median
(ASTM) diameter

in. microns in. microns

40/60 0.010-0.017 250-425 0.014 335

20/40 0.017-0.033 425-850 0.025 637

12/20 0.033-0.066 850-1700 0.050 1275

There are several methods of gravel placement, of which the most common and
effective is that using a crossover tool and tailpipe ( Figure 2 and Figure 3 , Openhole
gravel pack-underreaming hole followed by gravel placement). There are seven
important things to remember with respect to placement:

1. Gravel must be squeezed out through the casing perforations; usually about 0.5 sack per
foot of perforations (about 1.65 sacks per meter) should be placed outside the casing.

2. The entire completion interval must be completely packed with gravel.

3. The fluids lost or squeezed into the formation must be clean and compatible
with the formation fluids and shales to avoid a damaged zone at the
sand/gravel interface.

4. Gravel placement becomes increasingly difficult with increasing hole angle in


deviated holes, and special procedures have been developed to minimize duning
and consequential incomplete packing.

5. Longer intervals, over 25 ft (8 m), require a combined circulation-squeeze


technique to ensure effective gravel placement.

6. The screen must be centralized at 15-ft (4.5-m) intervals, and should overlap
the perforations by at least 5 ft (1.5 m) on each side.

7. Filtering of the gravel pack carrying fluid is essential for successful gravel
packing. Stage filtering with 10-micron and 2-micron filters is recommended.

There must be a sufficient gravel reservoir above the top perforation to permit pack settlement. If
viscous fluids are used to carry gravel concentrations of 15 lb/gal (1800 kg/m3 ), a 40% reduction in
height can be expected as the slurry breaks. Pressures must be carefully controlled to avoid
fracturing, but if formation fracturing does occur, additional gravel should be pumped at below
fracture pressures to ensure that the screen is fully packed.

Cased-Hole (Internal) Gravel Packs

686
The cased hole, or internal gravel pack has been the industry standard for most light to
medium oil wells and gas wells since the mid-l970s.

A schematic view of an internal gravel pack is presented in Figure 1 .

Figure 1

The produced fluid must flow through the gravel-filled perforations, through the gravel
pack and screen, to reach the well-bore. Hence, the productivity is governed by the
flow resistance through each of these sections. The greatest potential for flow
restriction is at the perforations, where the flow pattern becomes linear instead of
radial. Moreover, flow must converge and diverge in the vicinity of the perforations.
The major challenge of the packing operations is to transport gravel through the
perforation to pack the entire area around the perforation tunnel with very high
permeability gravel. Prior to 1970, the importance of keeping the gravel/sand interface
well away from the linear flow area was not fully appreciated and, as a result, the
productivity of many of the older internal gravel packs was poor.

The perforations through the casing should have a large diameter (3/4 to 1 in., or 19
to 25 mm) and high density (8 to 12 shots per foot [spf], or 26 to 39 shots per meter
[spm]) in order to achieve maximum flow area. The perforation and drilling damage
must be removed and the cavity in the formation enlarged by either the perforating or

687
cleanup technique. In near hydrostatic reservoirs, the most common cleanup technique
is perforation washing.

The gravel size is optimized to achieve the maximum permeability that can completely
stop the formation material at the formation-gravel interface. Gravel should be placed
without incurring mixing of gravel with formation solids. Gravel should be tightly
packed in the liner-casing annulus and in the perforation tunnels, without any voids
(i.e., a packing efficiency of 100%). All completion fluids should be very clean and not
create formation damage.

The liner screen should be designed to achieve minimum flow resistance while still
preventing any gravel from entering the tubing string. The annular space between the
liner and casing should be adequate to prevent gravel bridging during placement and
to permit washover operations. The screen need be no larger than the production
tubing in a conventional single-zone completion. The screen should be centralized and
constructed from a material that will resist corrosion, erosion, and collapse.

An ideal completion is difficult to achieve, although improved methods and equipment


have been developed over the years and high success ratios have been reported. The
perforations are the source of the greatest number of problems, since they are the
most critical areas. Opening up the perforations and packing gravel efficiently inside
the tunnels without mixing gravel with formation solids is a challenge, especially in
heavily depleted zones, in deviated wells, and in shallow, poorly consolidated formation
sand, where sloughing and cave-ins often occur.

Another major challenge has been to place the gravel through the relatively narrow
annular gap between the screen and casing without forming bridges that result in
premature sandoff or voids. Bridges form due to gravel and fluid traveling at different
velocities otherwise known as phase slippage and due to node buildup opposite
highly permeable zones. With the trend toward more highly deviated wells, avoidance
of bridging in inclined wellbores has become a more prominent installation design
consideration.

Openhole (External) Gravel Pack

For the openhole or extend gravel pack, ( Figure 1

688
Figure 1

and Figure 2 , Openhole gravel pack-underreaming hole followed by gravel placement),


the production casing is set above the pay zone, or milled out over the pay zone.

689
Figure 2

The gravel is placed in the annular gap between a liner screen and the formation. The
hole is usually underreamed to remove drilling damage and to reduce flow restriction
by enlarging the wellbore radius of the sand/gravel interface, where some damage is
inevitable. Underreaming and enlarging the wellbore effectively stimulates the well and
results in a negative skin in the radial flow analysis.

As in the cased hole gravel pack, the produced fluid must flow through the gravel and
through the screen to reach the wellbore. Unlike the internal pack, however, the fluid
flow with the external gravel pack is purely radial. The greater surface area involved
results in a much lower flow resistance, especially in high-rate wells, where turbulence
is significant, and in heavy oil wells. Thus the external gravel pack provides potentially
higher productivity due to the absence of perforations.

Well Productivity

A wells productivity is usually discussed in terms of a productivity index (PI, or J). The
productivity index, is the ratio of the flow rate, q, to the difference between the

690
average reservoir pressure, pr, and the bottomhole flowing pressure, pwf,:

J= [b/d/psi or m3/d/kPa] (16)

The quotient, pr - pwf, is referred to as the drawdown.

For laminar, radial flow of a single-phase liquid with viscosity, , in a reservoir of


height, h, the relationship between drawdown and flow rate, q, is given by Darcy's law,
as follows:

J= [dimensionless]

J= [oilfield]

(17)

J= [metric units]

where:

B = formation volume factor (reservoir volume/stock tank volume)

= fluid viscosity at reservoir conditions (cp)

k = effective permeability of the reservoir to the fluid (md)

re = drainage radius (ft or m)

rw = wellbore radius (ft or m)

A deviation from the logarithmic pressure change with radial distance from the well can
be expected in the near wellbore region and is termed a skin effect (S).

Commonly an additional pressure drop, or positive skin, is seen. This is generally the
result of formation damage or ineffective or incomplete perforation of the pay zone.
Many effects can combine to cause this sort of damage; some of the most common
are:

completion/workover damage as a result of plugged perforations, solids deposition in the


pores, swollen shales, clay or scale precipitation, water blockage, etc. This results in a zone
of reduced permeability around the perforations.

691
perforation efficiency in terms of the length of perforation, shot density,
phasing, perforation diameter, and, most critically, the extent and permeability
of the crushed zone also affects the skin.

drilling damage; although commonly cited as the cause of skin, it is much less
critical now that high-powered perforating guns can fully penetrate the
damaged zone.

turbulent or rate-dependent nondarcy skin, which is common in gas wells and


high-rate oil wells. This effect is generally linear with flow rate and is best
determined by multirate testing. This effect dramatically increases with partial
penetration and is claimed proportional to (h/hp)2.

pressure losses in the completion, which appears as an additional skin effect


especially in internal gravel packs where both Darcy linear flow effects and
nondarcy effects can be expected.

Negative skin factor may be caused by:


a zone of increased permeability around the wellbore due to acidizing, underreaming,
sand production, or deep perforations;

hydraulic fractures, causing linear flow patterns in highly permeable fractures;

anisotropic reservoirs causing nonradial flow patterns;

deviated wells, resulting in an increased amount of sandface being exposed


compared to a vertical well; essentially this can be considered to be the inverse
of partial penetration.

Thus, the effect of a gravel pack (and/or the efficiency of the gravel-packing operation from a
production viewpoint) is seen in terms of a variation in skin factor. Alternatively, we can look at the
ratio or deviation between the well's actual productivity (Ja ) and the ideal productivity of an
equivalent fully penetrating openhole completion with zero skin (Jo ), which can be computed
theoretically.

A number of papers have been published illustrating methods of theoretically


calculating the various skin effects especially for gravel pack design and
performance prediction the simplest form being the computation of the additional
pressure drop over the perforation tunnel under Darcy and nondarcy conditions:

p1 - p2 = (18)

p1 - p2 = + (0.1 10-13 L )

(19)

692
where:

A = cross-sectional area (ft2)

B = formation volume factor (res. bb/STB)

= inertia (turbulence) coefficient (ft-1) ( see Figure 1 ,

Figure 1

Relationship between inertia coefficient and gravel permeability)

k = permeability of fill material (darcies)

L = length of perforation tunnel (ft)

q = production rate (b/d)

= fluid density (lb/ft3)

= fluid viscosity under downhole conditions

693
p1 = upstream pressure (psi)

p2 = downstream pressure (psi)

The value of the Reynolds number is used to determine whether the flow is laminar or
turbulent. The Reynolds number, Re, for Newtonian fluids is given by

Re = [dimensionless units] (20)

where:

V = fluid velocity

= fluid density

dp = pore throat diameter

= fluid viscosity

The transition from laminar flow to turbulent flow in porous media may occur for
Reynolds numbers between 1 and 10. For Re > 10 the effect on the pressure loss due
to inertial effects becomes significant. Turbulent flow in the reservoir is more
predominant for gas wells, but has been reported for high-rate oil wells. As seen in
Equation 19, the pressure loss due to inertial effects varies with the square of the fluid
flow rate.

Figure 2 (Effect of perforation size on pressure loss in high-rate wells.

694
Figure 2

).illustrates the use of these equations to demonstrate the value of using high shot
densities and large diameter perforating guns, and the importance of avoiding sand
influx into the perforation tunnels.

Table 1, below, provides some typical values for gravel porosity and permeabilities.

While the cased-hole gravel pack inevitably results in some loss in productivity, the
pressure losses across an openhole pack will probably be less than would have
occurred through the equivalent section of formation prior to underreaming. The
resulting improvement in production can be computed using the Muskat equation:

(21)

where:

695
kp = permeability of the pack

ko = formation permeability

re = drainage radius

rw = openhole wellbore radius

rp = radius to the outside of the pack

Gravel size Porosity Permeability


(U.S. mesh) (%) (darcies)

8-12 34-42 1500-2200

10-20 32-40 300-900

20-40 30-38 100-200

40-60 28-36 40-80

Table 1: Typical gravel properties.

Gravel Selection
Historically, many of the early problems with gravel packing resulted from improper
sizing of gravel. A common mistake was selecting of gravel sizes that were too large.
While this would maximize a packs initial permeability , this permeability would be
reduced, in ongoing production operations due to penetration of the pack by formation
sand. Mixing formation sand with gravel can severely reduce the permeability of the
mixture ( Figure 1 , Permeabilities of 20/40 mesh gravel mixed with Oklahoma No.1
sand), as demonstrated by Sparlin (1972, SPE 4033; 1974, SPE 4772) and Saucier
(1974). Therefore, the trend over the years has been to decrease the size of the
gravel.

696
Figure 1

Early criteria consisted of sizing gravel to induce arching or bridging of the formation
sand against the outer surface of the gravel. In his classic paper, Saucier (1974)
demonstrated that bridging was not a reliable criterion due to pressure fluctuations;
and that the best criterion to completely stop formation sand from penetrating the
gravel pack is the sizing for filter-type retention of more than 50% of the sand.

Mechanics of Sand Retention

In an unsupported well-bore, sand can move as (1) individual particles suspended in


liquid, (2) as a viscous fluidlike quicksand, or (3) in large chunks due to sloughing.
With a gravel pack in place, sand movement into the wellbore can only occur as
individual particles suspended in the liquid. Particles that are larger than the pore size
of the gravel do not penetrate the gravel pack. Since pore sizes are generally about
1/6 the grain size, a 6:1 ratio of gravel size to sand size is often used.

Particles smaller than the gravel pore size can be prevented from penetrating the
gravel pack by bridging or arching. Sand bridges form for certain flow conditions and
remain stable for a period of time. Changing the conditions can destroy the bridges

697
and allow sand to penetrate the gravel pack until new bridges form. Hence, the use of
bridging criteria for gravel sizing should be minimized.

Formation Grain-Size Distribution

The first step in determining the optimal gravel size is to perform a sieve analysis on
representative samples of formation material. Grain size is usually plotted as
cumulative weight percent of retained material on a linear, vertical scale versus sieve
screen hole size or mesh size on a logarithmic horizontal scale.

Figure 2 illustrates such an analysis for a fine-grained, moderately sorted sand. It also
illustrates the sand-size designations commonly used by geologists in accordance with
the Wentworth Scale.

Figure 2

Grain-size distribution can vary greatly among different reservoirs. Figure 3 shows
variations in formation grain-size distribution for different locations throughout the
world.

698
Figure 3

Note that the sizes are plotted from largest to smallest along the horizontal scale.

An important characteristic of the grain-size distribution is whether it is narrow or


wide, otherwise known as the sand's uniformity. Hence, as shown in Figure 3 , U.S.
Gulf Coast sands are much more uniform than California sands. A uniformity
coefficient, c, has been defined as the ratio of the 40 percentile diameter and the 90
percentile diameter as follows:

(22)

The 40 or 90 percentile diameter is the theoretical sieve size for which 40 or 90


percent by weight of the total formation sample will be retained. Using this approach,
sand is classified as follows:

c Sand Description

<3 highly uniform (very well sorted)

3-5 uniform (well sorted)

699
5-10 nonuniform (moderately to poorly
sorted)

>10 highly nonuniform (very poorly sorted)

It should be noted that Figure 2 (Characteristics of a grain size analysis for a fine-
grained, moderately sorted sand ) is based on the "standard" sieve opening
designation in microns, which is the ISO recommended procedure.

Formation Sampling

Collecting representative formation samples for determining grain-size distributions is


critical to designing gravel packs or other mechanical sand control devices. As
demonstrated by Maly (1971), grain-size distribution can vary greatly throughout the
formation and so the best source of samples is full-length and full-diameter cores.
Coring assembles with rubber plastic, fiberglass or aluminum inner liners are available
for recovering cores in unconsolidated sands.) Cores should be sampled on a rigorous,
uniform, close spacing usually at 1-ft (0.3-m) intervals across the entire
completion section. Additional samples should be taken in any very fine or coarse
streaks, especially if they appear weak. However, it is only necessary to analyze
material above the permeability cutoff; shales and fine silts can be ignored.

If full-length cores are not available, sidewall cores should be used. These samples,
which can be retrieved on wireline during openhole logging, are small and usually
contain some drilling solids and fines due to crushing, but the resulting distortion to
the low end of the analysis is easily identified and can be ignored in the design work. It
is common practice to take cores during delineation drilling (and/or in one well per
platform or per section [640 acres, or 256 ha]) for initial completion planning; and
then to use sidewall samples in subsequent wells to prove that the sand size
distribution conforms to the core data. If it does not, the gravel pack for that well is
designed using the sidewall cores, and additional coring is scheduled in subsequent
wells in that area. Spot coring may be used in a similar manner.

Samples from perforation cleaning or backsurging can be used if more appropriately


obtained samples are not available; however, these may be skewed due to gravity
segregation and/or perforation damage. Produced samples tend to be skewed to the
small grain sizes due to flow separation up the wellbore. For the same reason, bailed
samples tend to be skewed to the larger grain size. Caution should be taken in using
such samples.

It is useful to put all of the grain-size analyses together on a single plot ( Figure 4 ) to
create a "horsetail" showing the relative size and frequency of the various grain size
distributions.

700
Figure 4

For very long intervals, representative plots can be selected for the upper and lower
bounds, the median and various intermediate groups (largest 25%, smallest 25%;
smallest 10%, etc.).

Gravel Sizing

Coberly (1937) used bridging or arching of sand as a basis for establishing gravel
sizing criteria, which led to large gravel-sand ratios and low success rates.

The Schwartz (1969) method for sizing gravel accounts for sand uniformity and fluid
flow velocity Schwartz established the following sizing rules:

For sands with c ratios (see Equation 22) less than 5 (uniform sands) and entrance
velocities below 0.05 ft/s, use d10 as the critical grain size.

For sands with c ratios greater than 5 or entrance velocities above 0.05 ft/s,
use d40 as the critical grain size.

701
For sands with c ratios greater than 10 (extreme nonuniform sands) and
velocities near or above 0.1 ft/s, use d70 as the critical grain size.

Size the gravel so that:


dnn gravel = 6 x dnn formation sand
where dnn is the critical grain size from the above criteria. The flow velocity, v, is calculated as
follows:

(23)

However, it would probably be more appropriate to look at the velocity of the


sand/gravel interface assuming 50% of the interval or perforations are contributing.

Note that as the grain-size distribution becomes less uniform, the critical grain size
decreases, which ensures that the small grains will not enter the gravel pack.

The most commonly used gravel size criteria are those recommended by Saucier
(1974). Saucier performed experimental studies under multiphase and unstable flow
conditions to determine the effect of the gravel-sand (G-S) ratio on the gravel pack
permeability. The results from Saucier's work are shown in Figure 5 (Sauciers gravel
pack impairment data ),

Figure 5

702
where the ratio of effective to initial gravel pack permeability is plotted versus the G-S
ratio, which is the ratio of D50 of the gravel to d50 of the formation sand.

The experimental results show that for G-S ratios up to about 6, the formation sand
does not penetrate the gravel pack. For G-S ratios above 6, the gravel pack
permeability drops significantly with increasing G-S ratio. This behavior results from
sand invading the pack and partially blocking the pore spaces.

The gravel pack permeability improves for G-S ratios above 13, but this results from a
pore space so large that the formation sand can travel right through the pack, which
defeats the purpose of the pack.

Although the gravel pack permeability ratio is constant for G-S ratios less than 5, the
absolute permeability of the pack decreases with decreasing G-S ratio ( Table 1,
below).

Gravel size (U.S. mesh) Porosity (%) Permeability (darcies)

8-12 34-42 1500-2200

10-20 32-40 300-900

20-40 30-38 100-200

40-60 28-36 40-80

Table 1: Typical gravel properties

Saucier recommends that the median grain size of the gravel be less than 6 times
larger than the median grain size of the formation sand, which is expressed as follows:

D50 gravel < 6 x d50 formation sand (24)

This corresponds to filter retention sizing, since the theoretical pore size for close-
packed spheres is around 15% of the sphere (gravel) diameter.

This results in a somewhat smaller gravel than does the Schwartz method. Schwartz's
criteria for uniform sand lead to G-S ratios of about 7 when corrected to a critical grain
size based on d50. According to Figure 5 (Sauciers gravel pack impairment data and
resulting design criteria ), this size would result in nearly a 60% reduction in the gravel
pack permeability.

Some experts recommend that the d50 size of the formation sand be multiplied by 4
and by 6 to obtain the gravel size range. If this range is not commercially available,
then the next smaller gravel size range is used.

703
Another approach is to plot gravel design lines over the sand size analyses ( Figure 4 ,
Gravel design using formation sand sieve analysis and Figure 6 , Overlay of gravel
design lines on the sand sieve size analyses(typical deltaic sand)).

Figure 6

These gravel design lines represent 1/6 the gravel grain size, based either on an actual
distribution, or a straight line between a sixth of the upper and lower size limits (i.e.,
for 20/40 U.S. mesh gravel the size range is 425 to 850 microns and the design line
would be 71 to 142 microns).

The sand grain analysis shown in Figure 2 (Moderately sorted, typical deltaic sand )
has the following characteristics:

d10 = 450 microns

d40 = 215 microns

704
d50 = 180 microns

C = 3.6

The resulting gravel pack sizing criteria would be as shown in Table 2, below.

Range of Gravel (in.) (microns)


Size U.S. Mesh

6-10 0.079-0.132 2000-


3350

8-12 0.066-0.094 1680-


2380

10-16 0.047-0.079 1190-


2000

12-20 0.033-0.066 840-1680

16-30 0.023-0.047 590-1190

20-40 0.017-0.033 420-840

40-60 0.010-0.017 250-420

Median (D50 micron)


Gravel Size
U.S. Mesh

6-10 2675

8-12 2030

10-16 1590

12-20 1275

16-30 890

20-40 637

40-60 337

Optimum Sand Range (1/6 gravel (d50) (=1/6


Control U.S. Mesh size) (microns) D50)

705
6-10 333-558 446

8-12 283-393 338

10-16 197-333 265

12-20 142-283 213

16-30 100-197 148

20-40 71-142 106

40-60 42-71 56

Optimum Size Gravel (in.) Screen (mm)


U.S. Mesh (gauge)

6-10 0.070 70 1.75

8-12 0.050 50 1.25

10-16 0.040 40 1.00

12-20 0.025 25 0.65

16-30 0.016 16 0.40

20-40 0.012 12 0.30

40-60 0.006 6 0.15

Slotted pipe is available down to 0.012 in. in straight-cut profiles and to 0.020 in. in
undercut profiles. Below these sizes, wire-wrapped screen must be used. Wire-
wrapped screen is usually available for the range 0.006 to 0.040 in.

Table 2: Gravel and screen sizing data.

The design line approach to this problem is illustrated in Figure 6 (overlay of gravel
design lines on the sand sieve size analyses( for typical deltaic sand )). This, too,
would suggest the selection of 16/30 mesh (600 to 1180 microns) gravel. However,
20/40 mesh (425 to 850 microns) might also be selected if:

the very fine sands occur as separate layers;

the downhole production conditions are very unstable due to turbulence, gas
breakout, or pump operations at high drawdown.

706
Some companies have a policy of using 20/40 U.S. mesh gravel (425 to 850 microns) for all fine
and very fine sands; or all completions in an area, such as the U.S. Gulf Coast. This is not only a
matter of operational efficiency and convenience, but is also based on good technical arguments:
Higher gravel permeabilities are rarely essential, unless the average formation
permeability exceeds 500 md or the produced fluid has a high viscosity.

Smaller gravels minimize the reliance on bridging to control the stringers of


very fine sand, and therefore maintain their permeability far longer under
conditions of increasing intake velocity (gas wells and/or increasing GLRs).

It is difficult to achieve a well system that is clean enough to avoid plugging of


40/60 gravel and its associated 0.006-in. (0.15-mm) screens during
installation.

Table 2 , above, is a summary of the commonly used gravel sizes and the associated sand ranges
based on Saucier's criteria.

In most gravel pack operations, silica-based gravel (i.e., sand) is used. However, for
special applications, bauxite or other materials may be used. In general, special
gravels are only used in thermal wells, where dissolution of the silica can result in
conventional packs becoming loose and ineffectual after three or four huff 'n' puff
cycles. Special pack materials are also used in deep wells, in which a combination of
fracturing and gravel packing is undertaken, and where sand is not strong enough to
resist the frac closing stresses.

Resin-coated gravel is often used in injection wells to prevent displacement of the


gravel into fractures and/or to minimize mixing at the sand/gravel interface.

Gravel Quality

Gravel quality can vary considerably depending on the source of gravel and how it is
handled. Construction-grade sand and gravel is totally unsuitable for oil industry
applications. Gravel can vary in purity, roundness, strength, and quartz content. The
maximum weight percent of material above and below the specified gravel size range
must be specified. Dirt can become mixed in with the gravel during transportation and
gravel can become crushed during transport or placement.

Gravel should be well rounded and spherical, since angular gravel tends to generate
fines during pumping and placement. A Krumbein roundness and sphericity of 0.6 or
better is often recommended. Figure 7 (Krumbein scale for roundness and sphericity)
shows the various classifications of roundness.

707
Figure 7

Gravel should be insoluble to reservoir brines and HCl, and less than 1% soluble in
12/3 HCl/HF mud acid. Quartz content of 98% or greater is usually recommended.

Sparlin (1974, SPE 4772) showed that as little as 10% by weight of fines can reduce
gravel permeability by 50%. It is therefore recommended that not more than 2% by
weight of gravel should be smaller, and not more than 5% by weight be larger than
the specified gravel size range. Not more than 0.5% fines should be smaller than 100
mesh (150 microns). More stringent specifications are desirable, but very difficult to
achieve from the supplier. However, many operators are prepared to pay a premium
for low-fines, received gravel to ensure high-quality packs.

On-site quality control should include steps to avoid contaminating or crushing the
gravel during placement.

Liner Selection

The liner is located on the inside of the gravel pack to prevent gravel from entering the
wellbore while maintaining sufficient flow area to prevent flow restriction. As shown in

708
Table 1, below, the slot in a gravel pack liner is selected to be lightly less than the
smallest gravel size. Thus, the liner exerts filter-type retention on the gravel.

Range of Gravel (in.) (microns)


Size U.S. Mesh

6-10 0.079-0.132 2000-3350

8-12 0.066-0.094 1680-2380

10-16 0.047-0.079 1190-2000

12-20 0.033-0.066 840-1680

16-30 0.023-0.047 590-1190

20-40 0.017-0.033 420-840

40-60 0.010-0.017 250-420

Median (D50 micron)


Gravel
Size U.S.
Mesh

6-10 2675

8-12 2030

10-16 1590

12-20 1275

16-30 890

20-40 637

40-60 337

Optimum Sand Range (1/6 gravel (d50)


Control U.S. Mesh size) (microns) (=1/6 D50)

6-10 333-558 446

8-12 283-393 338

10-16 197-333 265

12-20 142-283 213

16-30 100-197 148

709
20-40 71-142 106

40-60 42-71 56

Optimum Gravel Screen (mm)


Size (in.) (gauge)
U.S.
Mesh

6-10 0.070 70 1.75

8-12 0.050 50 1.25

10-16 0.040 40 1.00

12-20 0.025 25 0.65

16-30 0.016 16 0.40

20-40 0.012 12 0.30

40-60 0.006 6 0.15

Slotted pipe is available down to 0.012 in. in straight-cut profiles and to 0.020 in. in
undercut profiles. Below these sizes, wire-wrapped screen must be used. Wire-
wrapped screen is usually available for the range 0.006 to 0.040 in.

Table 1: Gravel and screen sizing data.

Both slotted pipe and wire-wrapped screens are used for gravel retention. Although
more expensive, the tendency in recent years has been for operators to use all-welded
stainless steel (or special steel) wire-wrapped screens whenever budgets permit.

Cost and performance can vary significantly with the type of screen. The engineer
must optimize screen selection to satisfy the completion objective while minimizing
overall costs, including deferment of future workovers and/or lost production due to
gravel pack failure.

Slot sizing or wire spacing should always be specified in inches (millimeters or microns)
and never in terms of a mesh size.

Slotted Liners

Slotted liners are relatively inexpensive, and good quality can easily be achieved by
local manufacturers.

This makes them an attractive option for low-cost operations, remote areas, and
locations where currency restrictions limit use of foreign-made goods. They are
generally much easier to recover from a well than a poorly made wire-wrapped screen.
However, their use is declining with the increased availability of moderate-cost, high-
quality wire-wrapped screens..

710
The main argument against the use of slotted pipe is that the pressure drop across
such a liner can be quite significant for large flow rates. As shown in Figure 1
(comparison of effective inflow area for screens and slotted liners), the effective inlet
area of slotted pipe is quite low, so that minor plugging can easily blind the flow.

Figure 1

Wire-wrapped screens have 10 to 30 times the open area of slotted liners. Slotted
liners are more subject to erosion and corrosion, being available mainly in carbon
steel, while wire-wrapped screens can be made in a corrosion-resistant alloy,
standardly 304 stainless steel.

Wire-Wrapped Screen Liners

There are many types of wire-wrapped screens available, ranging in method of


construction, materials used, and price. The types available include wrapped-on-pipe,
wrapped-on-grooved-pipe, ribbed, and all-welded ( Figure 2 ,

711
Figure 2

Figure 3 ,

712
Figure 3

Figure 4 ,and Figure 5 , Wire wrapped screen designs).

713
Figure 4

Earlier screens were constructed using solder, but these screens tend to unravel in a
"birdnest" fashion when being retrieved. Welding has solved this problem.

714
Figure 5

All screens are designed with undercut or V-shaped slots to reduce plugging of the
screens by the gravel ( Figure 6 and Figure 7 , Details of an all-weld screen).

715
Figure 6

Figure 7

The wrapped-on-pipe screen has the least flow capacity and tends to become plugged
with fine particles much more easily than the other screens. The screen performance is
somewhat improved by cutting grooves in the pipe. Adding ribs between the wire and
the pipe increases the flow area and substantially reduces plugging risks. However, the
OD of this screen type is considerably larger than that of the pipe. The most recent
development came with induction welding, which eliminates the weld beads and lugs
on the outside, thus increasing the open-flow area and allowing smaller slot widths,
greater slot-size control, and a stronger screen. As shown in Figure 6 and Figure 7 ,
each wire is welded at every contact with the support rods. The smallest slot opening
is 0.003 in. with all-welded screens, whereas the limit for the other screens is 0.006 to
0.008 in.

Screens can be constructed from carbon steel, 304 or 316 stain-less steel, Monel,
Inconel, or Hastelloy. The choice of material depends on the thermal/corrosive
environment in the wellbore. Although 300-series stainless steels are the industry
standard, they should not be used in wells where the bottomhole temperatures exceed

716
200 F (93 C) because of the risk of chloride stress cracking. Similarly, sour wells,
high CO2 levels, firefloods, and steamfloods all require the higher grade alloys.

All screens except the all-welded type have lugs impressed in the wire to maintain slot
size during construction. For highly corrosive wells, the lugs may be a site for corrosion
pitting due to work-hardening.

For the all-welded screen, the screen jacket is manufactured by welding the wire to the
ribs. The jacket can be welded to a perforated or slotted pipe, or used by itself if the
extra reinforcement is not required. All-welded screens are the most expensive, but
provide the highest flow area and best slot width control ( Figure 1 , Figure 6 , and
Figure 7 ).

It is important to remember that the OD of the screen is considerably larger than the
nominal diameter of the pipe by which it is designated, and is often larger than the
coupling (e.g., 4.5-in. [14-mm] screens have an OD of 5.15 in. [131 mm] and a
coupling of 5 in. [127 mm]).

The screen-to-casing radial clearance should be at least 0.75 in. (19 mm) and the
coupling radial clearance at least 0.5 in. (13 mm). For openhole situations, at least 2
in. (51 mm) and often 4 in. (102 mm) of pack is recommended around the screen.
These clearances are not only required to aid gravel packing but also to permit
washover operations during screen recovery. However, unless a multizone completion
is being installed, there is no advantage in using a screen that is larger than the
lowermost section of tubing. Moreover, larger screen-to-casing radial clearances tend
to result in better gravel packs and fewer bridging problems during installation. Thus,
while screens of up to 6.625 in. (168 mm) can be used inside 9.625-in. (244-mm)
casing, it is not uncommon for 3.5-in. (89-mm) screens to be used, giving a radial
clearance of 2.3 in. (58 mm).

The wire spacing, or slot width, of a screen should always be checked with a feeler
gauge prior to installation, and should be within 0.001 in. (0.025 mm) of the specified
size.

Screen or Slotted Liners

The minimum requirement on liner length is that the screen should be long enough to
overlap the upper and lower perforations (or the top of the underreamed section) in
order to compensate for any inaccuracies in the well depth measurement. An overlap
of at least 5 ft (1.5 m) at each end is usually recommended, but the overlap distance
depends on the accuracy of the depth control technique used to set the gravel pack
packer.

The screen should extend across the entire completion interval including any
unperforated or shale sections. Blank pipe should never be used in the middle of a
pack since the gravel may not dehydrate properly over this interval, leaving a "holiday"
that cannot be surged or washed for repacking.

A second small screen section is often used either above or below the main screen as a
"telltale" screen. This is used during installation operations to give a pressure

717
indication that the gravel buildup has reached a particular position. These telltale
screens are exactly the same as the main screen, except that they are only 1 to 3 ft
(0.3 to 1 m) in length. For conventional circulation packs they are located at the top of
the liner, while for squeeze packing they are located below the main screen just above
the isolation packer or plug.

Where packs are to be closely spaced, or where high deviations (>60) impair gravel
settlement, the screen may be continued above the completion interval across the
gravel reservoir up to the gravel pack port. Use of screen in place of blank pipe for this
permits positive placement and dehydration of the reserve gravel.

Blank Liner and Reserve Gravel

A section of liner is run above the gravel pack completion interval to create an annular
space to store a reservoir of gravel to compensate for any settlement. This is usually
made up with blank pipe from the top of the screen to the gravel pack port.

The length of blank pipe and/or a reserve gravel column is one of the most contentious
issues in gravel pack design:

Where space is not at a premium, some companies like to have sufficient


room to store the entire slurry volume in the annular space prior to sandout of
the telltale screen. This minimizes the risk of sanding-off the crossover tool, but
results in very long liners.

Other people believe that minimizing liner lengths reduces the risks of
bridging. Therefore they use only enough blank pipe to permit gravity
segregation of the gravel reservoir. In this case, the length of blank pipe is a
function of the gravel concentration (e.g., at 15 ppg, the free fluid occupies
some 40%, while the gravel and pore fluid occupies 60% of the interval when
settled). Companies using closer screen-to-casing clearances tend to favor
shorter liners because of the increased risk of bridging.

There is also little agreement over the optimum size of the gravel reservoir.
Recommendations range from less than 30 ft (<10 m) to more than 165 ft (>50 m),
and may be based on the following guidelines, according to past use and experience:

Lower bound Either one 30-ft (10-m) joint or 33% of the total pack volume,
whichever is less.

Upper bound 60 ft (20 m) minimum or 33% of the total pack length,


whichever is greater.

Most operators are prepared to compromise on their requirements where multiple


packs are to be placed in close proximity. To reduce the risks in these situations,
prepacked screens are often used to supplement the gravel pack and the screen is
continued across the entire gravel reservoir to aid placement and avoid any settlement
requirements. By use of only 10 ft of gravel reservoir, it is possible to complete two
separate zones within 25 ft (7.5 m) of each other.

718
Centralizers

All liners should have centralizers mounted on the outside to ensure that the liner is
completely surrounded by gravel and that all perforations can be accessed. Centralizer
spacings of 15 ft are usually recommended.

Spring- or bow-type centralizers are used in external (open-hole) gravel packs, since
they are adaptive to the varying wellbore diameters characteristic of unconsolidated
formations.

For internal gravel packs, the wellbore size is accurately known and rigid fin-type
centralizers are recommended. Fin-type centralizers can be welded on or clamped on,
and come in a variety of materials such as steel, aluminum and plastic ( Figure 8 ,

Figure 8

Figure 9 ,

719
Figure 9

Figure 10 and Figure 11 (Various types of gravel pack screen centralizers).

720
Figure 10

They must, however, be designed for easy milling or collapse during washover
operations.

721
Figure 11

In selecting the centralizer to be used, the engineer should address the risks of debris
being left on the sump packer during millover operations. For deviated wells, the visual
flow experiments by Elson et al. (1982) have shown that weld-on centralizers work the
best to reduce pocket formation around the centralizers.

Downhole Tools
As shown in Figure 1 (detail of internal gravel pack auxiliary equipment ), the auxiliary
gravel pack equipment consists of:

the sump packer or plug

722
Figure 1

the seal nipple (optional)

the shear-out safety joint

the gravel pack port or sleeve

the gravel pack packer

the gravel pack tools

the work string

The Sump Packer or Plug

The sump packer, bridge plug, or cement top acting as the bottom of the gravel-
packed interval should be located within 3 ft (1 m) of the lowermost section of screen.
As there is often a 1- to 3-ft (0.3- to 1-m) telltale screen located below the main

723
screen ( Figure 1 ), the "bottom" may be as much as 12 ft (3.6 m) below the
perforations. Longer sump intervals do not pack effectively and should be avoided.

Many operators prefer to use a sump packer as the "bottom" of the gravel pack, rather
than a bridge or cement plug, because a packer:

positively centers the screen;

provides a snap latch indication that the screen is on depth;

permits sloughing sand, or solids settling during packer running operations, to


fall through into the sump;

facilitates gravel collapse into the sump during screen recovery.

Since the sump packer or plug is the primary depth control device for positioning the liner with
respect to the perforation, most operators like to use a wireline set permanent packer (or bridge
plug) for this purpose.

The Seal Nipple

For squeeze packing operations, a seal or polished nipple is often used between the
lower telltale and the main screen. The tailpipe on the gravel packing assembly is
stung onto this nipple in the lower circulating position ( Figure 2 ).

724
Figure 2

This ensures that, during running operations, all circulation is taken through the
bottom of the liner and not through the main screen; and that, during initial circulation
operations, fluid travels across the entire completion interval before returning through
the lower telltale screen. Thus the circulation operation fills the entire annulus with
slurry, with minimal risk of bridging.

Sometimes this seal bore sub has an O-ring type seal mounted in a groove within the
nipple. In other systems, the seals are mounted on the tailpipe and the seal nipple is
simply a polished bore.

The Shear-Out Safety Joint

This is often run between the blank pipe and the gravel pack packer assembly to
facilitate washover operations. Quite often a screen must be washed over to remove
the gravel before it can be pulled from the well. The shear-out safety joint permits the
overlying packer to be pulled to facilitate washover operations without damaging the
screen.

725
While a chemical cut of the blank pipe would be an alternative, the tubing string and/or
packer is quite often sanded-up, or throughbore dimensions may make it difficult to
run a chemical cutter.

The shear rating of the safety joint should be sufficient to withstand both the tensional
and compressional loads associated with installation. Usually a unit that only shears
upwards is selected, and set for a total pull of 30,00 to 40,000 lb (135 to 180 kN).

Gravel Pack Ports

The gravel pack ports provide access from inside the liner to the liner-casing annulus
after the gravel pack packer has been set. Most operators prefer to iso-late these ports
during production to minimize the risk of pack fluidization.

Three options are available:

1. A perforated joint with seal bore subs above and below. This is isolated during
production operations by the production seal assembly that straddles the perforated
joint and engages the two seal subs ( Figure 2 , tailpipe stung onto nipple during
circulation).

2. A sleeve-type gravel packing port that is normally closed, but can be


shifted into the open position by the gravel packing tool. Once the sleeve
is in the open position, the flow area through the ports exceeds that of
the tubing.

3. Gravel pack ports that are an integral part of the gravel packing tool
and separate liner isolation system.

Gravel Pack Packer

A packer is required during placement of the gravel to ensure that circulation occurs
down around the screen, and, during production operations, to minimize the risk of
pack fluidization by upward flow through the pack. Commonly a single packer is used
for both purposes, and it may also serve as the production packer on flowing and
gaslifted wells ( Figure 1 , Figure 2 , and Figure 3 , packer used during gravel
placement, and production packer).

726
Figure 3

For this type of multifunctional service, a hydraulically set permanent-retrievable


packer is commonly used, especially in the deeper, deviated wells. This has all the
features of a permanent packer but can be released mechanically with a special pulling
tool during workover or repair operations. Double grip mechanical packers with an on-
off assembly are sometimes used in vertical wells.

Where throughbore is critical and/or if a repair is unlikely due to a limited reserve life,
permanent packers are sometimes used as a combination production/gravel pack
packer.

For shallower, lower-pressure wells, a simple, low-cost mechanical set or cup-type


service packer for gravel placement ( Figure 4 ) is commonly used.

727
Figure 4

This packer is hydraulically or mechanically released from the liner and recovered with
the other tools. The liner is then sealed either to the casing or to the tubing string with
a lead seal assembly ( Figure 5 ).

728
Figure 5

Gravel Pack Tools

Gravel pack tools vary somewhat in design and complexity, depending on the
placement method and conditions. However, the most common tools include:

a packer setting mechanism;

a crossover tool arrangement;

a tailpipe;

a mechanism for reverse circulation of excess gravel;

a mechanism for releasing the tool from the packer or liner.

The heart of the system is the crossover tool, which permits circulation via the inside of the work
string to go to the outside of the liner through the gravel pack ports, with returns being taken via the

729
tailpipe into the annulus above the packer. In general, this is a multiposition tool (used for running,
spotting, lower circulation, squeezing, and upper circulation) in which the circulation paths can be
adjusted by raising and lowering the work string. It is therefore important to check the effects of
hydraulic piston forces on the tool during pumping operations. Most operators prefer to have a
repeat collet, "J" or "W" path arrangement to aid positive location of the crossover tool in the desired
position.

The flow paths within the crossover tool should be carefully reviewed for tortuosity,
flow area, and sealing capability. Concerns include not only sandoff risks but also
erosion, sand impact, and ease of operation.

The tailpipe provides a means of spotting the gravel pack slurry and treatment fluids
across the entire completion interval. Relatively small clearances between the tailpipe
OD and liner ID are preferred to discourage flow through this area, especially if a seal
nipple and bottom telltale is not used.

The entire gravel pack assembly should be function and pressure tested both before
shipment to the location and prior to running. It is made up to the gravel pack liner
and packer on the rig floor as an integral assembly, which is run into the well on the
work string.

The Work String

The three critical requirements of the work string (which can be either tubing or
drillpipe) are that :

it must be clean and solids-free (note: the gravel slurry is highly erosive and must not
become contaminated with material picked up during pumping). A dedicated string of
special stainless steel pipe is often used. Alternatively, the pipe may be sandblasted, and
pickled with acid and corrosion inhibitor just prior to mobilization.

there must be sufficient pipe weight available for packer setting operations
and to control piston forces on the gravel pack tools.

the ID must be small enough to prevent the slurry from overrunning the
completion brine during pumping operations at normal circulation rates 3 to 5
BPM (0.5 to 0.8 m3/m). (This effect is called "roping" and is a function of the
relative fluid densities, viscosities, and flow velocity.) At the same time, the
pipe ID must be large enough to avoid excessive pump pressures, and/or to
permit the use of a variety of balls and/or darts for tool manipulation.

For most wells, 5.5-in. (140-mm) drillpipe is too large for effective gravel placement; therefore,
dedicated 2-7/8-in. (73-mm) or 3-1/2-in. (89-mm) work strings are commonly used for well cleanup
and gravel packing.

Through-tubing gravel pack systems have also been developed for use with coiled
tubing or snubbed macaroni work strings.

Perforation and Perforation Clearing Procedures

730
Pressure loss across the perforations is the major design consideration with cased-
hole, or internal, gravel packs. The largest possible flow area must be achieved by
high-density perforating with large entrance hole guns. These must be effectively filled
with clean, high-permeability gravel. Formation sand, sand-gravel mixtures, fines, mud
solids, and dirt must be kept out of the perforations to maximize productivity.

Perforation design must consider:

the perforation geometry;

effective removal of perforation damage;

avoidance of subsequent damage.

Gun Selection and Perforation Geometry

Perforating for gravel packs requires special considerations. The cross-sectional area of
the perforation is the point of greatest restriction to flow in the gravel pack. Hence,
large perforation diameters (0.75 to 1 in., or 19 to 25 mm) are required to facilitate
placement of gravel in the perforation tunnel outside of the casing and to minimize
pressure losses at the perforation during production. Figure 1

731
Figure 1

illustrates schematically the gravel-packed perforation tunnel and the effect of


perforation size on pressure losses (also see Figure 2 , effect of perforation size on
pressure loss in high-rate wells.). For the same reason, high-density perforations are
usually recommended, typically 8 to 12 shots per foot.

Figure 2

Maximizing the flow area is particularly important in gas wells and high-rate oil wells,
where turbulence can be expected, since, under these conditions, pressure losses
increase with the square of the velocity, and pack destabilization effects are much
more severe.

As in natural completions, the use of multidirectional shot phasing (60 to 180) and
perforation of the entire pay section reduces the geometric skin effects and maximizes

732
well productivity. However, in highly deviated wells, there has been a trend toward
avoiding shots on the topside of the hole, since these are difficult to pack.

For jet perforation, a tradeoff exists between perforation diameter and penetration.
Penetration depth is much less important in gravel pack completions than in natural
completions, especially when perforation washing is employed. In general, the
perforation need only penetrate a short distance beyond the casing, cement, and
drilling damage. Deeper perforation tunnels are unlikely to remain open until the
gravel is injected, and therefore have little benefit.

Small through-tubing wireline guns do not have the power and shot density required
for gravel packing. Large wireline casing guns may be used for vertical and moderately
(<60) deviated wells at relatively low cost.

Tubing-conveyed perforating guns (TCP) impose no limit on interval length or well


deviation, and allow for underbalanced perforating with large diameter guns. They are
also safer in high-pressure wells or when H2S is present. Furthermore, a tubing-
conveyed perforating system provides a means for immediate backsurging of the
perforations and flowing the well for a period of time to remove debris. However, for
short intervals,TCP costs are higher than those for wireline perforating. Misfires are
also more costly because of the extra time required to trip the tubing out of the hole,
and may be more of a problem since gun exposure to downhole temperature is longer.
Nevertheless, TCP

Both casing and TOP guns are designed to be centered and as near drift as possible.
Careful depth control is, of course, essential, especially with TOP systems. Ideally,
perforations should be round and free of burrs. Burrs can damage the rubber wash
cups on perforation washing tools.

Effective Removal of Perforation Damage

Perforating tends to compact the material around the perforation tunnel, causing
localized damage and reduced productivity. Cement, copper, lead, and carbon from the
jet charge penetrate the formation, adding to the damage ( Figure 3 ,

733
Figure 3

Figure 4 ,

Figure 4

Figure 5 , and Figure 6 , diagrams- perforating process and associated permeability


damage.

734
Figure 5

).

Figure 6

Furthermore, overbalanced perforating in unfiltered completion fluids can leave many


perforations plugged by fines, which are difficult to remove even with backflow.
Therefore, underbalanced perforating is recommended for all natural perforated
completions ( Figure 7 and Figure 8

735
Figure 8

, Cleanup of plugged perforations after overbalanced perforating,

736
Figure 7

and Figure 9 and Figure 10 , Cleanup after underbalanced perforating).

737
Figure 9

738
Figure 10

There is currently considerable debate over the relative merits of underbalanced


perforating versus perforation washing as the best method of preparing a well for
gravel packing. At the same time,there is universal agreement on the importance of
opening all perforations to remove perforation and drilling damage, and to create a
cavity to be filled by gravel if a highly productive internal gravel pack is to be achieved.
The major techniques for perforation cleanup include underbalanced perforating,
backsurging, washing, and acidizing.

Underbalanced Perforating

Underbalanced perforating involves reducing the fluid pressure inside the casing to
some value less than the reservoir pressure. When the perforation passes through the
casing, there is an immediate flow of liquid into the wellbore due to the pressure
differential. This high pressure surge effectively forces debris out of the perforation
tunnel into the wellbore. Unlike washing, this technique does not involve prolonging
leakoff of completion fluid into the formation.

The differential pressure required to clean the formation depends on the reservoir
characteristics, typically the reservoir fluid, formation permeability, and the rock
strength. Although industry opinion is not yet firm on specific drawdown requirements,

739
it is generally agreed that higher drawdowns are required for gas sands and low-
permeability zones. Underbalanced pressures of 500 to 1000 psi (3.4 to 6.9 MPa) are
usually recommended for gravel packing. There is a distinct tradeoff with reducing the
risk of sanding in the gun, especially in very shallow wells. Some proponents of TCP
recommend increasing the blank gun section to equal the gun length, especially in
high-permeability sands; this creates a large, atmospherically pressured chamber,
ensuring a reasonable pressure surge, which causes the sand to fail and creates
cavities behind the pipe. (All gun sections are atmospherically pressured before firing.)
However, this procedure requires extreme caution when pulling the guns because of
the risks of trapped pressure.

It is also important to flow the well immediately to transport the sand and fines into
the sump and/or to surface. At least one U.S. gallon (4.4 liters) of reservoir fluid per
perforation, and preferably 5 U.S. gallons (22 liters) per perforation should be
backflowed. This usually involves flowing the entire well to hydrocarbon, so some
operators take the opportunity to make a pregravel-pack test at the same time.
Moderate drawdowns should be used at this stage (200-500 psi, or 1.4 to 3.4 MPa).

There are conditions where other perforation cleaning methods are difficult because of
fluid placement and circulation problems, or because of a reservoir sand that is
extremely fluid sensitive. Underbalanced perforating is clearly superior in the following
situations:

very high permeability zones (>500 md) with hydrostatic or subhydrostatic pressures;

highly depleted reservoir pressures;

extremely water-sensitive sands;

sour wells;

operations where local conditions limit the available volumes of clean, filtered
fluids.

The main objection to only using underbalanced perforating prior to gravel packing is that in
unconsolidated sands, the perforation tunnels will probably be left at least partly filled with sand that
may be difficult to remove during gravel packing operations. Some operators, therefore, still wash
perforations even where TCP has been used to perforate a long interval.

Washing the Perforation

Historically, the most successful method of preparing a zone for gravel packing has
been perforation washing, and some operators claim that, where feasible, it is superior
to underbalanced perforating.

Washing involves running a temporary packoff system so that completion fluid can be
circulated behind the pipe through each of the perforations in turn ( Figure 11 ,

740
Figure 11

cup-type circulating washer and procedure for washing perforations) and ( Figure 12 ,
pressure- operated packer system and procedure for washing perforations). This
creates a void behind the pipe that can be filled with gravel.

741
Figure 12

The energy from the mud pumps is used to physically break down the rock over a
short distance. Then, circulation of 5 to 10 bbl (1 to 2 m3) of completion fluid at 2 to 4
bpm (0.3 to 0.6 m3/m) erodes a cavity behind the pipe. The resulting debris is
circulated or reversed out periodically.

Two types of washing systems are commonly used:

the opposing-cup washer ( Figure 11 );

the pressure-operated packer system ( Figure 12 ).

The cup systems are cheaper and allow for continuous pipe movement, but have limited head
capacity. The packer systems are better suited for deep wells, and some companies prefer them on
all wells because of the reduced risk of element damage and/or of losing an element in the well.
Inflatable packers are also sometimes used.

The element spacing is normally 1 ft (0.33 m) or less to ensure that the maximum
number of perforations are treated.

742
Normally, washing is done from bottom up, and the debris is reversed out of the sump
on a return downward trip. This may be repeated until the required volume of solids
has been removed or no further sand is seen in the returns. Most companies collect the
returns in a settling tank so that sand volumes can be estimated. The objective is to
remove 0.25 to 0.75 ft3/ft (0.02 to 0.07 m3/m), with 0.5 ft3/ft (0.05 m3/m) being the
norm.

Washing of the perforations should be performed with "super clean" fluids, filtered to 2
microns or less to minimize formation damage. Fluid leakoff could be high. Low-
viscosity fluids are preferable for washing, to easily penetrate the formation and create
the required jetting or turbulent action to remove perforation debris. In partially
depleted zones, foam can be used as a wash fluid, while in overpressured zones, heavy
brines can be used. Some companies wash overpressured zones under snubbing
conditions with the well pressurized. Washing perforations can result in very high fluid
losses, which may cause formation damage. It is sometimes necessary, therefore, to
use a fluid with leakoff control or increased viscosity either during the washing process
or while pulling the wash tools and running the gravel pack.

Backsurging

The traditional approach to reservoirs that were difficult to wash because of low
bottomhole pressures and/or fluid sensitivity was to backsurge the zone. This is done
by running a DST-type assembly by which the perforations could suddenly be exposed
to a very high drawdown but a limited volume of flow. The principle is similar to
underbalanced perforating.

Figure 13 shows a typical backsurge tool.

743
Figure 13

An atmospherically pressured chamber, with a volume of one U.S. gallon (4.4 lt) per
perforation, is trapped between two valves. The lower valve has a snap-acting
mechanism to create an effective surge. It may be a ball or shear disk and is usually
annulus pressure operated, although a drop bar system may also be used.

Long intervals can be backsurged in stages from bottom up with the lower perforations
being sanded-off after each run. However, the technique is best suited to short
intervals. While it is still used in remedial work, backsurging of initial completions has
generally been replaced by underbalanced perforating with TCP guns.

The main objection to backsurging, and to a lesser extent to underbalanced


perforating, is that it leaves the perforation tunnels filled with sand that is often
difficult to displace during the gravel placement process. Another objection is that the
process can lead to a well control problem when the packer is unset.

Cleanup Acidization

In discussing acidization, a clear distinction must be made between cleaning up the


perforations and the treatments used just prior to gravel placement to remove lost

744
circulation materials, to create a reasonably uniform injectivity profile, and/or remove
and stabilize mobile clays from around the wellbore.

Acidizing the perforations has had varying degrees of success for gravel packing. Good
results are often due to poor perforation techniques, which if improved, can eliminate
the need for acidizing. However, there are situations where acidization is required in
any event because of dirty sands and/or deep damage, and it can therefore be
combined with a perforation cleaning process.

As with other acid jobs, the key to successful perforation cleanup is effective diversion
of the acid. Traditional batch diverting agents, such as ball sealers or unibeads, may be
used to ensure that all perforations are broken down. Alternatively, a selective
acidizing tool may be used ( Figure 14 ) in a similar fashion to the perforation washer,
to treat the entire completion interval section by section.

Figure 14

745
One interesting approach, reported by McLeod, (1982) is to combine the acid job with
a prepack operation. The acid is diverted by slugs of viscous, gravel-laden fluid that
will also fill the perforation tunnels currently open.

The second concern with acidizing, especially when using HF acid to remove clays, is
effective removal of the spent acid from the wellbore area. The dangers of secondary
products and silica redeposition from spent mud acid is well documented, and it is
generally recommended that acid be produced back at once. This may be difficult to
achieve in a pregravel-pack situation. Therefore, many operators prefer to conduct
their acid treatments as part of the gravel packing operation, so that the acid is
immediately displaced away from the wellbore area by the gravel carrier fluid.

Avoidance of Subsequent Damage

Once the perforations are open, a degree of leakoff is inevitable while the perforating
or perforation washing assembly is pulled and the gravel packing equipment run. The
fluids in the well at this time must be super clean and filtered to less than 2 microns.

In higher-permeability zones, it is necessary to spot a fluid-loss agent to permit


adequate overbalance for safe tripping operations. This agent must be properly
designed in terms of:

selecting the optimum combination of viscosity and particulate material;

selecting the required viscosity break times and method;

ensuring thin filter cake, which does not block the perforation tunnels;

obtaining a filter cake that is properly sized to shut off losses at the sandface
but can still be backflowed through the pack;

selecting a fluid loss material that is easily removed prior to gravel placement.

Fluid-loss pills should depend primarily on viscosity to slow down the leakoff rate. Where solids
must be used, they should consist of an easily removed material, such as oil-soluble resins, calcium
carbonate, or salt.

Gravel Placement Techniques


Many different placement techniques have been developed over the years but only a
few remain popular. Full-scale laboratory studies, field experience, and the
development of new gravel pack tools have all led to improvements in gravel
placement techniques. Much of the present research work in gravel packing is directed
toward improving gravel placement techniques, especially for highly deviated wells.
Placement is critical since any voids in the pack or poor gravel compaction can result in
failure.

Gravel placement is more difficult for internal gravel packs due to the necessity of
packing the perforations. Also there is less margin for error with the internal gravel
pack.

746
Wash Down Method

One of the earliest methods, referred to as the wash down method, consists of
depositing the gravel in the wellbore and squeezing or fracturing it into the
perforations in stages. Then the screen is washed through the packed gravel across
the completion interval. This method is illustrated in ( Figure 1 ,

Figure 1

Figure 2 ,

747
Figure 2

Figure 3 , and Figure 4 ).

748
Figure 3

The wash fluid is injected down the tubing and the returns are carried up the annulus.

749
Figure 4

After the screen is in place the flow is stopped and the gravel settles down around the
annular gap between the screen and casing. This method is now rarely used on new
completions, since it does not achieve good compaction in the annulus and
perforations. However, it is still used for through-tubing packing operations.

Circulation Methods

In the circulation method, the gravel is carried into the screen-casing annulus by the
fluid, then the fluid leaks off through both the screen and formation, leaving behind the
gravel in a compacted state. The reverse circulation method, shown in Figure 5
(Reverse circulation, where the sand is pumped down the annulus and the carrying
fluid returned up the tubing) is the simplest of the circulation methods, but it is not
recommended.

750
Figure 5

The slurry travels down the tubing-casing annulus and potentially scrapes off the dirt,
cement, and pipe dope to be deposited in the gravel pack or formation.

The crossover tool circulation method, illustrated in Figure 6 has led to significant
improvements in gravel placement.

751
Figure 6

The slurry travels down the tubing, through a crossover tool located below a packer
and down into the screen-casing annulus. The returning fluid passes through the
screen and travels up a tailpipe, through the crossover tool, and then up the tubing-
casing annulus. In this technique, the tubing can be meticulously cleaned, and thereby
maintain a clean packing fluid. Sizing of the tailpipe is a key design parameter in this
application, since fluid flow through the screen prior to reaching the settled gravel
height can lead to void formation, especially in highly deviated wells. Usually the
tailpipe is sized to be as large as practically possible without risking its getting stuck in
the screen.

Squeeze (Bullhead) Method

The squeeze, or bullhead, method involves pumping the entire tubing volume and
following slurry fluid down the tubing. The pressure squeeze on the gravel, after it is in
place, forces the slurry into the perforations, where dehydration occurs. This method is
shown in Figure 7 , where the slurry is forced into the screen-casing annulus through
the crossover tool and the ball valve.

752
Figure 7

Usually a viscous fluid with a high gravel concentration is used with this technique.

The squeeze method is used primarily for short, uniform intervals of 30 ft (10 m) or
less at relatively shallow depths. In longer intervals of variable permeability, uniform
packing is difficult to achieve since the high-permeability streaks pack off first and
node buildup may obstruct the slurry flow.

For quicksands, a viscous squeeze technique can be used to force the formation sand
away from the wellbore while minimizing mixing of formation sand with gravel. The
squeeze rate should be very slow (0.25 b/m, or 0.04 m3/m) to prevent sand-gravel
mixing.

A squeeze pack is sometimes conducted on perforations alone as a pretreatment prior


to running the screen.

Combination Squeeze/Circulation

At present, a combination of circulation and squeeze is usually employed. A two-stage


operation involves an initial squeeze to gravel pack the perforations behind the casing,

753
which is usually carried out as part of the cleanup procedure; followed by running of
the liner and tools for a circulation pack of the screen-casing annulus ( Figure 8 and
Figure 9 ).

Figure 8

Proponents of the two-stage method believe gravel can be packed more efficiently in
the perforations if they are squeezed separately, and for longer intervals this is usually
carried out in stages.

754
Figure 9

This method requires a lot of rig time, and is therefore a more expensive technique.

The one-trip crossover circulation-squeeze technique combines circulation and squeeze


techniques in one trip by means of advanced tools. The method is illustrated in Figure
10 (One-step combination squeeze/circulation pack.

755
Figure 10

).

After the tools are lowered in place and the packers are set, gravel is circulated down
the tubing through the crossover tool. An increase in pressure indicates that the gravel
slurry at the lower telltale screen is dehydrated, indicating that the entire interval is
covered by slurry. The tool is then lowered into the squeeze position, which cuts off
access to the tubingcasing annulus above the upper packer. Additional gravel is
injected to squeeze the slurry through the perforations until a predetermined pressure
is reached. The gravel is thus screened out against the formation face and
progressively dehydrated back toward the liner. Screenout occurs once the entire
completion interval is covered with tightly packed gravel, since it is difficult to squeeze
the carrying fluid from behind the blank pipe vertically down through the sand pack.
The tool is then raised to the upper circulating position in an attempt to place
additional gravel. A second squeeze may be made before the excess gravel is reverse-
circulated out of the well. Finally, the setting tool crossover assembly and tailpipe are
removed from the well and the pack is ready for production.

Several variations on these basic procedures may be used, for example:

756
Partial losses may be induced during circulation by closing the annular preventer and
checking the returns. In some deviated well situations, this is claimed to better pack the
perforations and to prevent premature screenout from slurry dehydrating on the screen.

The lower telltale and seal nipple may be omitted and an oversized or baffled
tailpipe used instead, permitting some fluid flow through the screen for
simultaneous circulation/ squeeze packing of long intervals with slurry
dehydration occurring both radially inward from the formation and axially
upward from the bottom of the tailpipe.

An upper telltale may be used above the blank pipe in near-vertical wells that
are being packed at low slurry concentrations. This permits positive dehydration
of the reservoir gravel rather than gravity segregation.

Initial sand placement into the perforations may be conducted under fracture
conditions to create a short, wide fracture to bypass any deep damage.

For long zones with varying permeabilities, differential pressure valves may be
installed at several positions in the tailpipe and set to open progressively from
bottom upwards. The first stage occurs conventionally; the lowermost valve is
then opened, permitting circulation to an intermediate point in the screen and a
second phase of circulation and squeezing is then undertaken. This is repeated
with the second valve, and so on up the string.

Placing Openhole Gravel Packs

The same tools and techniques are used for openhole gravel packs as for cased-hole
packing. However, since there are no perforation tunnels to be filled, it is less critical
that dehydration takes place radially inward from the formation. Circulation packing
techniques are therefore reasonably successful with external gravel packs, and fluid
viscosities and sand concentrations need not be as high.

The major concerns with external gravel packing are:

effective removal of the drilling damage by underreaming with a clean,


nondamaging fluid;

fluid-loss control while underreaming and tripping for assembly changes;

maintaining stability of the underreamed section in the pre-gravel-packing


stage;

proper estimation of the gravel volumes required by openhole caliper logs;

effective removal of the fluid-loss control materials prior to gravel placement;

avoiding erosion and/or collapse of the sandface during gravel placement;

keeping any slumped sand away from the gravel pack screen;

757
control of duning in the deviated wells caused by the lower placement
velocities.

Considering these problems it should be apparent that the basic requirements of


openhole packing are similar to internal pack placement:

an ability to circulate fluids across the entire completion interval, for


displacement of fluid loss pills, spotting acid, and spotting slurry; this requires a
tailpipe arrangement, usually with a lower telltale or baffle;

use of low fluid volumes and velocities to minimize sandface destabilization;

use of squeeze packing to push back and tighten loose sand and maintain
borehole stability;

use of a liner that is hung from a gravel pack packer in the casing to permit a
small rathole (or sump) to collect any collapsed sand.

Therefore, the tendency in recent years has been to use similar combination circulation
squeeze packing procedures for both types of gravel pack.

Gravel Pack Slurry Requirements

As in fracturing, gravel pack slurries are designated in terms of weight of dry sand per
unit volume of carrier fluid. Therefore, while this has the units of density, concentration
does not represent the actual slurry density. The slurry density is a function of the
gravel pack fluid density and sand concentration.

vs = vf + vg (25)

v s = vf + (26)

ws = wf + wg (27)

ws = vf f + cgvf (28)

(29)

758
s (30)
gg = 21.93 lb/gal = 164 lb/cu ft
= 2630 kg/m3

l /gg = absolute volume of gravel = 0.0456 gal/lb


= 0.0004 kg/m3

In oilfield units:

s = lb/gal (31)
where:
vs = volume of slurry

vf = volume of fluid

vg = volume of gravel

cg = concentration of gravel in carrier fluid

gg = density of gravel grains (matrix density; not the dry sand density,
which includes air-filled pores)

ws = weight of slurry

wf = weight of fluid

wg = weight of gravel

f = density of fluid

Oil field units Metric units

cg = lb/gal cg = kg/m3

= lb/gal = kg/m3

Table 1, below, illustrates the effect of gravel concentration on slurry density.

Gravel Fluid Slurry Slurry Percent Settled


Concentration Density Density Head Gravel Fill
(Cg) (f) (s) Hs (vg/vs) (vgs/vs)

759
(lb/gal) (lb/gal) (lb/gal) (psi/ft) (%) (%)

1 8.5 9.08 0.473 4.4 6.7

10.0 10.52 0.547

12.0 12.43 0.646

5 8.5 10.99 0.572 18.6 28.6

10.0 12.22 0.635

12.0 13.84 0.720

10 8.5 12.71 0.661 31.3 48.2

10.0 13.74 0.714

12.0 15.11 0.786

15 8.5 13.95 0.726 40.6 62.5

10.0 14.85 0.772

12.0 16.03 0.834

Assumes gravel porosity is 35%.

Table 1: Typical slurry properties.

The percentage of gravel and free fluid in the slurry is also a function of the gravel
concentration:

(32)

(33)

In oilfield units:

(34)

760
Therefore, 15 ppg slurry contains 40.6% gravel. However, for settlement calculations, we must
recognize that the gravel is very porous (35%), and therefore part of the carrier fluid will be trapped
in the pore spaces. The settled volume of the gravel pack (vgs) will be

vgs = (35)
Therefore the settled volume as a proportion of the slurry volume is

(36)
In oilfield units and assuming a 35% packed gravel porosity:

(37)
Therefore, the proportion of gravel settling from a 15 ppg slurry is 62.5%. Table 1, above, gives
typical slurry concentrations and settling volumes.

The required gravel volume (vg) is equal to the volume of the annular space between
the screen and casing or open hole (vA), plus the volume of gravel outside the pipe
(vp), plus the volume of any rathole (vh), plus the volume of reserve gravel (vr) behind
the blank pipe.

vg = vA + vp + vh + vr (38)
Usually these are computed using the specific volume per unit length:
vg = vA x Ls + vp x Lp + vH x LH + vB x Lsg (39)
where:
vA = volume per unit length between the screen and hole or casing (in
openhole gravel packs it is usually determined based on a caliper log;
otherwise tables can be used); cu ft/ft, m3/m

vp = volume of cavity per unit length of perforated section,


depends on cleanout/working efficiency 0.5 cu ft/ft
(0.05 m3/m) (cross-check with volume of solids recovered during
washing)

vH = volume of rathole per unit length (from caliper or tables);


cu ft/ft, m3/m

vB = volume of annular space between blank pipe and casing per


unit length, based on tables; cu ft/ft, m3/m

Ls = length of screen; ft, m

Lp = length of perforated interval; ft, m

LH = length of rathole; ft, m

Lsg = length of settled gravel reservoir; ft, m

761
Lsg = LsB (40)
where:
LsB = length of slurry-filled blank pipe when pumping stops; ft, m
The weight of gravel required (wg) is therefore:
wg = vg (l - ) gg (41)
In oilfield units:
wg = vg (l - 0.35) 164 (42)

wg = 106.6 vg (43)
This is sometimes approximated to the dry gravel density of 100 lb/cu ft and since gravel is often
measured in l00-lb sacks, the approximation of 1 sx/cu ft is often used in determining gravel
requirements.

The fluid volume required (vf) depends on the gravel concentration (cg) to be used:

(44)

vs = vf (l + cg/gg) (45)
The total head during gravel placement should also be calculated based on the length of the work string that will be filled with
gravel and the corresponding density and net head of the fluids filling the remainder of the string. By deducting this head
from the fracture propagation pressure (FPP), the maximum allowable pump pressures can be determined for various stages
of a conventional gravel packing operation.

If no better data is available, it is normally assumed that the fracture propagation pressure is

FPP = 0.5 v + 0.5 pr (46)


where:
v = total vertical stress

pr = reservoir pressure

D = depth
In oilfield units, assuming a vertical stress gradient of 1 psi/ft:
FPP = 0.5 D + 0.5 pr psi (47)
So that for an undepleted, hydrostatically pressured zone (pr/D = 0.445 psi/ft) the approximate frac
pressure gradient is 0.72 psi/ft. Note how the fracture pressure decreases with pressure depletion
so that it is increasingly difficult to avoid fracturing partially depleted reservoirs.

When gravel packing low-pressure zones, additional gravel must be used to allow for
some initial fracturing, or else the job has to be conducted in a series of stages. Some
operators like to induce minor fracturing initially in all internal gravel packs to ensure
all perforations are gravel-filled.

There are three commonly used ranges of slurry concentration, which go under various
designations:

762
Circulation Conventional Modified Slurry Packing
Circulation Packing
Packing

low density med. density high density

Gravel-fluid 0.25-1 5-10 10-15

Ratio (ppg)

Fluid 0.5-20 40-200 350-1000


Viscosity

(Cp at 170
sec-1

and 80 F)

Pump Rate 5-10 1-5 0.2-1.5

(bbl/min)

Dehydration circulation circulation/induced squeeze

Method losses/squeeze

The conventional circulation systems are no longer commonly used except for some
long openhole gravel packs. Although slurry packing is now used for most gravel
packs, there is some evidence that in highly deviated wells, a better placement can be
achieved by simultaneous circulation and squeezing. The intermediate densities are
also used to reduce the head effects on very long zones in relatively shallow wells,
where fracturing problems are severe.

Outline Procedure for Placing an Internal Gravel Pack

The general procedure for placing a cased hole gravel pack is as follows:

Run bit and scraper.


Spot filtered perforating fluid.
Run cased-hole logs (CBL/VDL-GR-CCL).
Run gauge ring and set sump packer.
Perforate with high-shot-density, large-hole guns.
Clean up perforations by TCP and backflow, or perforation washing.
Spot lost circulation material if necessary.
Make up screen, tailpipe, blank, gravel pack assembly and packer.
Run slowly (one joint per minute) on work string with minimum pipe dope being used.
Tag sump packer.
Reverse-circulate.
Set packer.
Test packer by overpull.
Pressure test annulus.

763
Release gravel pack tool and function test. Set annulus pressure required to avoid pumping
out tools.
Establish circulation and squeeze pressures.
Spot and squeeze acid (25 to 100 gal/ft).
Displace with overflush (50 gal/ft).
Pump brine spacer (5 bbl).
Pump viscous pad (5 bbl).
Spot gravel slurry controlling U-tubing.
Pump viscous pad (5 bbl).
Pump brine chase.
Circulate gravel to crossover tool at 3 to 5 b/m, reduce rate to 1 to 1.5 b/m while placing
gravel over pay until lower telltale screens out at 1000 psi.
Squeeze gravel at 0.2 to 0.5 b/m (this may need a special low-capacity pumping unit) until
screenout occurs. Work pressure up in stages to maximum allowable pressure.
Check placement volumes.
Repeat squeeze and/or pick up to upper circulating position as programmed. (If gravel is
still in crossover tool, it should be reversed clean as soon as possible.)
Reverse out excess gravel and measure volume of gravel recovered.
Spot fluid-loss material, if required.
Pull work string.
Run gravel pack evaluation log.
Run completion and produce clean as soon as possible. (If two batches of slurry are used,
allow sufficient time for first slurry to break before repacking.)

Outline Procedure for Placing an External Gravel Pack

The general procedure for placing an open hole gravel pack is as follows:

Run bit and scraper and circulate clean.


Spot filtered, viscosified workover fluid.
Change bottomhole assembly and run cased-hole logs.
Drill out cement and drill pilot hole over the pay, or mill casing over the pay. Run logs if
required.
Run underreamer.
Circulate to clean fluid.
Underream pay section.
Spot lost circulation material and stabilization fluids.
Run openhole caliper log.
Calculate fluid and gravel requirements.
Run gravel pack assembly and liner to predetermined depth.
Set packer and place gravel as detailed in the procedure outline for cased hole gravel
packs, although the entire pack may be placed by circulation in some cases.

Fluids Used in Gravel Pack Operations

Functional Requirements

In designing a gravel pack, we must consider not only the fluids used for gravel
placement, but all of the fluids that have contact with the pay zone. As with all
completion and workover fluids, the concerns can be summarized in terms of:

764
control of reservoir pressure

of fluid losses
of borehole/cavity stability

cleanliness to avoid formation damage and screen plugging

compatibility with shales

with formation fluids

with reservoir permeability

carrying capacity for gravel placement

for cuttings/sand removal

cleanup of filtration control materials

of injected/lost fluids
Fluids are used in the following completion operations involving gravel-packed wells:
drilling the pay zone;

plugging back the pilot hole (External Gravel Pack) and cementing the casing;

perforating the casing (internal pack), or drill-out (external pack);

acidizing and/or washing the perforations (internal pack), or under-reaming


(external pack);

controlling the well and fluid losses during tripping;

re-inducing fluid losses and stimulating the pay zone;

transporting and placing gravel;

squeezing and compacting (dehydrating) the gravel;

reversing excess gravel out of the hole;

controlling the well and fluid losses during completion operations;

bringing the well on production and removing any damage or filter cake.

In wells that are to be gravel packed, preventing formation damage and borehole instability requires
special precautions in drilling, cementing and perforating operations. Borehole sloughing during
gravel placement may cause gravel to mix with formation sand, which reduces the gravel pack

765
permeability. When underreaming, for external gravel packs, fluid-loss additives are sometimes
used to deposit a thin filter cake on the borehole wall to induce stability, although a crosslinked
polymer pill often suffices. The resulting filter cake must be easily removable before the gravel is
placed to prevent formation damage.

Perforation damage must be effectively removed without the perforation fluid or wash
fluid causing formation damage. Since substantial fluid losses usually occur, all fluids
must be thoroughly cleaned and filtered to remove insoluble fines, even if a soluble
fluid loss additive is going to be used. Filtration control may be needed not only to
facilitate tripping and to reduce the total volume of lost fluid, but also to aid
stabilization of the borehole and cavities and uniform distribution of the gravel.

The gravel packing fluid must be viscous enough to transport the gravel to the pay
without sanding off the string or crossover tool, but have sufficient fluid loss for the
gravel to effectively dehydrate. This can be achieved by varying the viscosity, velocity,
and sand concentration.

Gravel must be tightly packed in the space between the liner and the formation; any
voids could jeopardize the completion. In a cased hole, this means effectively
displacing the slurry through the perforation to the formation face. Leakoff must occur
through both the formation and screens to allow the gravel to pack tightly. The
placement procedure and fluid velocities must also prevent mixing of gravel and
formation material. Finally, excess gravel must be reverse-circulated out of the hole.

Early gravel pack operations usually employed water or other low-viscosity fluids to
transport gravel. But because these fluids could not adequately suspend the gravel,
operators began to use more viscous carrying fluids and decreased fluid velocities in
gravel placement. The use of highly viscous carrying fluids with high gravel
concentrations is referred to as slurry packing after Sparlin (1972, SPE 4033). This
technique accounts for much of the gravel packing performed today. New water-base
polymeric fluids are continuously being developed to perform the multi-task functions
of gravel packing, drilling, and hydraulic fracturing. The chemistry and flow behavior of
these fluids are complex, and a good understanding of fluid characteristics, such as
leakoff and rheology, is essential to gravel pack design and evaluation.

Base Fluids

The primary consideration in selecting a base fluid for any completion operation is
reservoir control. Gravel packing is performed under overbalanced conditions (i.e., a
small positive differential from the wellbore into the pay). Ideally, this should be just
sufficient to overcome any swab and surge pressures during tool handling (about 100
to 300 psi, or 0.7 to 2 MPa). The density of the base fluid should therefore be tailored
to the reservoir pressure.

Fluid options available include foam, oil, water, brines, and workover fluids. A few
operators have experimented with the use of invert oil muds, but this was generally
found to be surprisingly damaging to productivity. Brines are by far the most
commonly used system with HEC or multiple polymers being used for developing
viscosity for fluid-loss control and for solids transport.

766
The base fluid must be tested for compatibility with the formation and formation fluids.
Fresh water should never be used because of its effect on shales. The minimum salt
concentrations for control of clay hydration are

potassium chloride (KCl) 2%

ammonium chloride (NH4Cl) 4%

sodium chloride (NaCl2) 8%

Viscosifiers and Breakers

Hydroxyethylcellulose (HEC) is the most popular gelling agent because it exhibits less
formation damage than other polymer systems. Poorly refined guar gums and guar
derivatives can leave residues in the formation. Carboxymethylhydroxyethylcellulose
(CMHEC) is sometimes used but is sensitive to shear, its viscosity decreasing
significantly when sheared through the pumps and tubular equipment. Biopolymers
(xanthum gum) have good gel strength but can leave residues in the formation. Mixing
problems can occur with all polymer systems if not properly handled, the most
common being the formation of "fish eyes" globs of undissolved polymer which
are extremely damaging.

There is some debate over the advisability of crosslinking polymers for additional gel
strength. In most cases it is not necessary and should be avoided in gravel placement,
since it substantially reduces leakoff and therefore makes dehydration difficult.
However, crosslinked polymers have application where very coarse gravel is being
used in long, openhole gravel packs (500 to 1000 ft, or 150 to 300 m). There may also
be application in highly deviated wells in weak formations, where it is difficult to place
gravel in the perforations on the high side of the well. In conventional packing
operations, a cross-linked pill is sometimes injected across the formation during
tripping operations to maintain borehole stability and prevent excessive losses.

For slurry packing operations, typically some 65 to 80 lb of HEC are added per 1000
gal of 2% KCl to achieve an apparent viscosity of about 350 to 450 centipoise at 100
sec-1. For highly deviated wells, 100 to 120 lb of HEC are used. It is important to
remember that it takes time for the HEC to build viscosity so gravel pack fluids are
often premixed. Viscosity decreases with temperature and should be designed for
conditions at the crossover tool.

The brine chemistry and density are important parameters affecting polymer
performance. Increasing brine density above 12 lbm/gal decreases the hydration rate
of HEC polymer sufficiently so that heat and/or extended mixing times are required.
For example, the time to yield full viscosity in a 12.5 lbm/gal CaCl2/CaBr2 brine can be
six times longer than that for an 11 lbm/gas CaCl2 brine. Damage can be caused by
adding excess polymer to speed viscosity buildup. Globs of partially hydrated polymer
will form, and cause formation damage that is difficult to remove.

The effect of temperature on the fluid viscosities and break times for HEC brines can
vary from brine to brine. Zinc brines can be three times less sensitive to temperature
than zinc-free brines. However, high zinc concentrations are detrimental to HEC

767
thermal stability. The fluid temperature increases as the fluid moves downhole, thus
reducing the viscosity and break time. Prediction of the bottomhole temperatures
during gravel placement is one of the problems in fluid design. The major service
companies use thermal simulation programs for this purpose, and there are packages
available on a timeshare or service basis.

Breakers are used to control viscosity break-back times. Ideally the viscosity should
decrease during squeezing operations, but remain viscous enough at the crossover tool
to avoid sandoff and to facilitate reverse-circulation. The optimum break time depends
on the tool/liner/blank-pipe configuration. If all the slurry is behind the blank pipe,
control of the viscosity reduction with time can be used to optimize gravel placement
and compaction operations. The break-back time for a given breaker concentration
depends on the temperature and the shear rate. Published break-back data is often
confusing because of the various instruments and shear rates used to measure
viscosity. Usually much higher shear rates are used than those experienced during
gravel packing, which leads to erroneous prediction of break-back times. Low-shear-
rate viscometers are more representative of the shear rate experienced in the field.

A variety of acid, oxidative, and enzyme polymer-degrading breakers are used to


reduce viscosity over a certain time period. Acid viscosity breakers such as HCl are
normally used because of their reliability and insensitivity to mixing errors and
contaminants. Acid breakers also provide additional stimulation. However, acid
breakers are not suitable in high-carbonate-content formations or when some resin-
coated gravel systems are used. Enzyme breakers can be used up to 130 F (54 C).
Oxidative breakers, such as sodium or ammonium per-sulfates, can be used for
intermediate temperatures. Breaker concentrations should be selected so that the fluid
viscosity approaches that of the base brine within 24 hours. Sometimes the best
breaker is not selected by the service company because of patents and royalties.

Fluid Preparation

Special care is required to prepare polymeric fluids. The hydration of the polymer in
the brine is greatly affected by the temperature, composition, and concentration of the
brine. In most polymer systems, increasing the pH can increase the polymer hydration
rate. Pumping down partially hydrated polymer should be avoided. The trick is to
slowly add the polymer with the correct degree of agitation.

For high-density brines (above 12 lbm/gal), heat and extended mixing is required. If
fluid heaters are not available, mechanical heating by high-rate flow through a choke
can be used. However, the polymer should not be added until the brine is heated, since
high shear will degrade the polymer.

Alcohol or diesel can be used to disperse the polymer into the brine as this does not
affect the hydration rate. Alternatively, a liquid HEC concentrate can be used. Once the
HEC is added to the near neutral pH fluid, the pH is raised to 8 and the fluid is mixed
for about 30 min to 1 hr. Adjusting pH in the field is difficult for some brines, such as
NH4Cl, but easier for NaCl and KCl. For brines containing magnesium or zinc,
increasing the pH can result in precipitate formation. Oxygen can degrade polymers, so
air entrainment during mixing should be minimized.

768
Viscosities can vary by as much as 50% from one batch to another because of different
mixing procedures. Fluid viscosities should be monitored on site and should be 150 to
300 cps at speeds of 6 and 30 rpm on the Fann viscometer at 80 F (27 C) for l0-ppg
slurry, and twice that for l5-ppg slurry.

Advances have been made in improving hydration and reducing "fish eyes" by shear
filtration. As reported by Ashton et al. (1986), shearing the fluid on surface in a special
shear device untangles the polymers and allows the "fish eyes" to be removed
effectively by cartridge filters or filter press systems. Precise control of shear rate and
shear times is the key to success. Fluid loss increases significantly while viscosity
decreases only slightly. The potential for reducing formation damage is significant.
Even if the fluid will not be filtered after hydration, controlled shearing of the hydrated
fluid has proven to be a valuable step in fluid preparation and is now used by many
operators.

The gravel is then added to the viscous carrier in a paddle blender. Gravel blenders
often have sufficient size to permit the entire treatment to be mixed as a single batch.
Blender volumes of up to 60 bbl (10 m3) are now available, although their application
offshore is a function of deck load capacity. For very large jobs, twin blenders are
usually used. The breaker is added just prior to pumping.

Fluid Rheology

The rheology of a fluid is the description of how a fluid deforms under stress. The fluid
rheology affects gravel-carrying capacity, erosion, gravel compaction, pumping
pressure, and leakoff.

Rheology is usually described graphically in terms of shear stress versus shear rate,
which are the parameters measured in a simple shear viscometer (e.g., Fann unit). A
few typical rheograms are shown in Figure 1 .

769
Figure 1

The Bingham fluid is characterized by a gel strength that is the shear stress required to
initiate flow. The Power law fluid is characterized by a shear-thinning (time-
independent) behavior. Some fluids exhibit both a gel strength and a shear-thinning
behavior. The rheology of a fluid can be adjusted by adding chemicals such as
polymers. Most gravel pack fluids behave in accordance with the Power law.

The viscosity is defined as the slope of the shear-stress versus shear-rate curve. For
the Newtonian fluids the viscosity is constant, while for Power law fluids the viscosity
decreases with shear rate. Hence, these are termed shear-thinning fluids, and shear
rate is an important parameter in estimating the fluid viscosity ( Figure 2 ), carrying
capacity, injectivity, and friction losses.

770
Figure 2

Many linear gels and all crosslinked fluids are viscoelastic, which means that they flow
like fluids but exhibit elastic behavior as well. Hence, not all of the deformation energy
is dissipated; some of the energy is elastically recovered, which provides drag
reduction and suppresses turbulence. The flow behavior of viscoelastic fluids cannot be
fully characterized using shear viscometer techniques; special oscillating viscometers
are required to obtain the elastic properties.

Fluids with high gel concentrations thin with time under shear, and recover their gel
strength when shearing ceases. Any stoppage of flow for these fluids requires
additional pumping pressure to reinitiate flow. Gel strength is an important parameter
for gravel suspension, since gravel settling involves low shear stresses; the higher the
gel strength the greater the suspension capacity. A degree of gel strength is desirable
to avoid a sandoff of the crossover tool when pumping stops at the end of the job or
during temporary interruptions.

Protection against tool plugging during low-rate squeeze operations is also provided by
the natural properties of polymer-based fluids.

771
As discussed by Scheuerman (1983), in shear-thinning fluids most of the viscosity
reduction occurs at low shear rates, as shown in Figure 2 . Hence, the fluid viscosity
increases as the shear rate is lowered. This is beneficial for suspending and
transporting gravel. If the flow rate drops, the viscosity increases to improve gravel
suspension. When the flow rate increases, the viscosity decreases, thus reducing the
pumping pressure.

It is important to remember that the apparent fluid viscosity also increases with solids
concentration (e.g., 22% solids [6.5 ppg slurry] has double the viscosity of the base
fluid, and 48% solids [20 ppg] has ten times the viscosity of the base fluid). This is
important in estimating sand settlement velocities.

Two important parameters to consider for suspending and transporting gravel are the
settling velocity of the gravel in the fluid and the velocity of the fluid transporting the
gravel. High fluid flow velocities and/or low gravel settling velocities promote good
transport and suspension of gravel. Separation or slippage of the gravel in the liquid
can lead to tool plugging and/or duning in the gravel pack, or bridging in the gravel
pack, all of which leads to poor compaction efficiency. Duning is predominant in highly
deviated wells where the gravel settles on the low side of the casing, creating a bank,
or dune, of gravel. Bridging is caused by phase separation at the perforations and
screen due to rapid leakoff of liquid and a localized increase in gravel concentration.

For Newtonian fluids with particle Reynolds numbers less than 2, the settling velocity,
vs, is given by

vs (48)

where:

p = gravel material density

= fluid density

dp = particle diameter

= fluid viscosity

The particle Reynolds number is given by

Rep = (49)

Note that the settling velocity is very sensitive to particle diameter. This sensitivity
leads to segregation of the gravel where larger particles settle much faster than
smaller particles. For 12/20 gravel in 10-centipoise fluid, the settling velocity for the

772
mean diameter particle is 15.6 cm/min. for 100-centipoise fluid the settling velocity is
10 times slower. For 40/60 gravel the settling velocity is over 12 times slower.

For non-Newtonian fluids, the expressions for vs are more complex, especially since
the solids concentration affects the fluid viscosity. Data can be obtained from Acharya
(1986). For the same viscosity, particles settle faster in viscoelastic fluids than in
Newtonian fluid because of drag reduction on the particle in the viscoelastic fluid.

For ungelled brine, the maximum carrying capacity for gravel placement is about 1.0
lb/gal, whereas for gelled brine with a viscosity of 500 cp, the carrying capacity can be
15 lb/gal or higher. A concentration of 15 lb/gal provides a slurry in which the free
fluid amounts to only 40% of the total volumes.

Higher gravel concentrations mean less fluid pumped, less damage risk, and shorter
placement times.

The settling velocity is also used to calculate the slip factor as well as the friction factor
for gravel-laden fluids (Subash et al. 1985, SPE 13836). The presence of gravel in the
fluid increases the friction factor for the fluid, especially at high concentrations. The
slip factor for a flowing fluid is the velocity of the particle divided by the velocity of the
surrounding fluid. The slip factor increases with decreasing viscosity and decreasing
flow velocity. This can lead to roping of the gravel through the brine and is the reason
for pumping a viscous pad ahead of the gravel-laden fluid.

Fluid Leakoff

Control of fluid leakoff is a major problem in gravel packing. Leakoff control is highly
desirable during perforation washing or underreaming, during tripping and completion
operations. However, a significant and relatively uniform leakoff rate is highly desirable
during gravel placement to create a tight pack against the sandface. But even during
squeezing, excessive leakoff into a limited number of perforations is undesirable, since
it can lead to node buildup and bridging.

In gravel packing, leakoff control is usually achieved using viscosified fluids, although
in very high permeability zones (>1 D) graded soluble fluid-loss additives may be
needed.

The leakoff properties of a clear Newtonian fluid are governed primarily by the fluid
viscosity. However, polymers provide additional resistance to flow in porous media
(Savins 1970). The converging-diverging flow path through porous media for polymeric
fluids gives rise to what is called shear thickening. The extensional viscosity, which is
the rate of change of normal stress to normal strain, can be much higher than the
shear viscosity, even for small concentrations of polymers (i.e., less than 100 parts per
million). This is the reason that multipolymer pills can be so effective at controlling
losses, their major advantages over the solids-based systems being their ability to be
circulated out and their high sensitivity to acid.

Fluid-loss material in the form of solid particles can be added to the fluid to control
leakoff. Extreme caution has to be used for internal gravel packs because of the
danger of plugging the perforations. However, a solid filter cake is often used in

773
external gravel packs to increase wellbore stability and reduce sloughing. It is
important that the material used can be completely removed before or after the gravel
pack is placed. This is achieved by selecting materials that are soluble in acid, brine, or
reservoir fluids, such as calcium carbonate, salt, or resins. Diverting agents may also
be used in the pretreatment to ensure uniformity in the fluid injectivity.

After the gravel has been placed, it is often necessary to spot a fluid-loss pill inside the
liner to permit tripping operations. Although designed to be temperature degradable, it
is often necessary to make a cleanout trip or acid job with coiled tubing before the full
well potential can be achieved. Alternatively, a knock-out reverse flapper valve or
shear-out check valve can be used to isolate the reservoir during completion
installation.

The greatest fluid loss occurs during perforation washing. Some companies do not
advocate perforation washing because of this, especially in highly permeable and/or
subhydrostatic reservoirs, where underbalanced perforation with tubing-conveyed guns
is preferred.

Gravel Placement Fluid Selection

Selection of the appropriate fluid viscosity for gravel packing requires the consideration
of many factors.

Historically, fluids with viscosities between 0.5 and 1000 centipoise have been used for
gravel packing. The trend now is to use medium-viscosity fluids, typically 500 cp, to
optimize the overall fluid performance. The advantages and disadvantages of using
higher-viscosity fluids are summarized here.

Advantages:

good gravel transportation;

higher gravel concentrations, thus less fluid volumes;

reduced gravel crushing;

reduced gravel/sand mixing;

reduced gravel size segregation;

enhanced borehole stability;

lower pumping cost/placement time.

Disadvantages:
lower leak-off rates at the formation face;

higher tubular pressures;

774
higher risk of fracturing the formation;

more expensive in terms of chemicals;

poorer compaction, especially in the gravel reservoir;

poorer cleanup if fluid does not break properly or is inadvertently cross-linked


by reservoir contaminants.

Formation Damage and Fluid Filtration

Formation Damage

Formation damage is the term used for an additional pressure drop in the wellbore
area ( Figure 1 ). The following are some causes of formation damage:

invasion of drilling mud solids into the formation (especially into fractures);

Figure 1

drilling mud filtrate invasion into the formation;

cement losses into fractures;

cement filtrate invasion into the formation;

775
plugged perforations (often due to overbalanced perforating);

inadequate perforations (size, number, or penetration);

partial penetration of the producing zone (i.e., not opening the total pay);

crushing and compaction of formation matrix surrounding a perforation;

invasion of solids in completion or workover fluids into the formation or


perforations;

invasion of completion or workover fluids into the formation;

plugging of the formation from the swelling of water-sensitive native clays;

asphaltene or paraffin precipitation in the formation or perforations;

scale precipitation in the formation or perforations;

creation of an emulsion in the formation;

injection of acids or solvents that contain solids or precipitate solids;

sand fill in the wellbore;

injection of an oil-wetting surfactant into the formation;

excessive drawdown that causes movement of formation fines, compaction of


a weak formation, or instigates water production.

From the work of Abrams (1975), Darley (1965), Tuttle and Barkman (1974), Nowak and Kruegar
(1951), and others, it can be concluded that solids entrained in the drilling or completion fluids are
the main cause of impairment. This is particularly apparent in gravel packing, which involves the
injection or loss of relatively large volumes of fluid into pay. In this situation, the formation is
behaving as a deep-bed sand filter, so that filter theory can be used to predict where solids will be
deposited. If dp = mean pore size (microns) of the formation, then the effects of invading materials
and the treatment for their removal can be characterized as follows:

Diameter of invading Effect Treatment


solid

Bridging Backflow
>33% (e.g., drilling mud filter cake)

Shallow invasion Acidize or reperforate


<33% and >10% (e.g., skin caused by solids in
completion fluid)

776
Probably not harmful
<10%
Two rules of thumb for estimating pore size of sandstones:

15 - 25% of mean grain size

where:
k = permeability in md

= pore size, microns

Completion fluid filtration policy should be aimed at avoiding shallow invasion where solids are
difficult to remove. Therefore, as permeabilities decrease, it is necessary to use finer filters:

k (md) optimum filter size (microns)


(microns)

1000 31 3

300 17 2

50 7 0.7
This is also illustrated in Figure 2 , which also shows how filter cakes should be designed to bridge-
off the pores.

777
Figure 2

Although the standard used by the industry today is 2-micron filters, which are suited to formation of
more than 300 md permeability, commercial filter units are available down to 0.5 microns.

Fluid Filtration

Stage filtration is most efficient when the fluid passes through a screen and coarse
filter before entering the main filtration unit.

The highest water quality is obtained using a deep-bed filtration process with either a
sand filter or, more commonly, a diatomaceous earth (DE) filter press. In this system,
a filter cloth supported on a screen is precoated with a porous permeable bed of
diatomaceous earth. The fluid is filtered as it passes through the filter cake, which is
continuously built up by a body feed. Periodically, the filter is backflushed and the
process repeated. In general, these filters not only give the clearest water, they also
give the highest filtration rates. Deep bed filters remove all particles down to the
designated size.

The other alternative is to use a replaceable cartridge filter. Again, two basic types are
available:

depth filter cartridge increasing in density towards the center of the element to give
longer filter life;

778
surface filter cartridge particles are trapped on the face of the filter
medium; these are less expensive, but clog faster.

Cartridge filters are adequate where the feed fluid is already relatively clean (<100 ppm suspended
solids). However, since they only remove a percentage of the solids in the feed, dirty fluid can only
be cleaned adequately by multistage series filtration at each filter size range.

A cartridge filter should always be placed downstream of any DE filters in case of filter
cloth breakage.

Multiple-Zone Completions
Where multiple pay zones occur and wells are IPR-limited, it may be economical to
produce more than one zone at a time. The zones may be commingled or sequentially
produced up one tubing string or, more commonly, multiple tubing strings may be
used. A long zone may be completed with several gravel packs, one above the other, if
gravel packing the entire zone would entail excessive difficulties or if future zonal
isolation is required.

An example of a multizone, single-string completion is given in Figure 1 where all


zones have internal gravel packs.

779
Figure 1

Figure 2 (Schematic of external gravel pack with lead seal adapter) shown earlier,
illustrates a dual-zone, dual-string completion where the lower zone is an external
pack and the upper zone is an internal pack.

Figure 2

Figure 3 illustrates dual-zone completions with two external packs.

780
Figure 3

One option is to include the screens (or slotted pipe) as part of the casing, isolating the
zones with external casing packers (ECPs) and cementing through port collars. This is
a complex and tedious operation and dependent upon getting a stable openhole
section and clean drilling fluids.

The other option for dual external packs involves milling the casing and underreaming
prior to gravel packing, at least for the upper zone, and possibly for both. This has
proven very successful, once drillers gain some experience in milling a window.
However, improvements in perforating and internal gravel packing techniques have
resulted in a decrease in the use of external packs for the upper zones.

Multiple internal packs and/or dual internal/external packs do not impose any special
equipment requirements except careful attention to the sizing of the packers and seal
nipples. These must be staggered so that the smallest nipples are on the bottom. Seal
assemblies must be carefully spaced and of adequate length to accommodate tubing
movement, especially when the zones extend over a considerable distance. The
geometry usually permits limited use of locator or anchor seal assemblies for positive
positioning of the completion, so space-out operations are critical.

781
To maximize the throughbore through the upper pack, permanent packers are
generally used and screen-to-casing annulus is minimized, although there is clearly a
trade off with achieving an efficient packing operation, especially on long zones. While
radial clearances down to 0.7 in. (18 mm) have been used, better results are obtained
if the radial clearance exceeds 1 in. (25 mm). Nodal analysis should be used to
determine the effect of reducing the flow areas in the tailpipe to the lower zone, and in
the screen-to-tubing annulus for the upper zone. Generally, additional area can be
given up to the gravel pack with little consequence to productivity.

However, this is not always true, in which case a basis is established to show that the
risk of a poorer upper pack can be accepted because of the resulting increase in
productivity from the lower zone. Ultimately, such analyses may show that the limited
benefit and increased risk of a multizone completion do not justify its installation on
the initial well completion, and that perforation of the upper zone should be deferred to
a workover.

In summary, multiple zone gravel packs for separately produced zones require special
packers and seals, and careful planning and execution. Efforts should be made to keep
the operation as simple as possible, and equipment should be easily retrievable in case
of failure. The lower zone is perforated and gravel packed first by standard methods.
The zone is then plugged temporarily while the upper zone is perforated and gravel
packed by standard methods. Finally the isolation plugs are recovered and the dual
string is run with the appropriate packer and seal configuration.

One of the critical issues with multizone gravel packing is how closely packs can
effectively be placed. This comes down to an issue of how much gravel reservoir is
required.

Another issue is how many zones can be completed in one well. Up to seven gravel
packs have been placed in some wells. However, this greatly increases the completion
time and risk. The tendency in recent years has been to reduce the number of packs to
two or three per completion operation, deferring completion of upper zones to
subsequent workover operations.

Some companies are completing multiple gravel packs in a single operation, especially
when the zones will subsequently be commingled and/or have limited differential. All
the zones are perforated at the same time and all equipment is run on the same string,
correctly spaced out. After the hanger/top packer is set, the isolation packers are set
one at a time, starting from the bottom upward. Each packer must be tested
separately. Sliding sleeves are used to gain access to each zone. The bottom zone is
gravel packed first, using standard procedures. After the gravel is placed and the
excess gravel circulated out of the hole, the placement tools are moved up to the next
zone and the operation is repeated. With this type of completion, failures are harder to
repair.

For long (>150 ft) internal packs in highly deviated wells, the zone is sometimes
completed as a series of gravel packs. However, breaking up a long interval into two or
more zones requires more equipment and rig time, and thus greater expense.
Moreover, part of the pay behind the packer and gravel reservoir cannot be perforated
and may not contribute its reserves. Hence, the engineer must select the optimum

782
number of intervals that will minimize cost and maximize productivity, while still
allowing effective gravel packing operations and complete drainage of the reserves.

Another problem for areas with stacked reservoirs is that it is often desirable to gravel
pack an upper interval while using natural perforated completions for the deeper
zones. Geometrical considerations usually preclude the use of the standard multiple
retrievable packers in this case, and several permanent packers or a preinstalled
bottomhole assembly must be used instead. Both of these configurations complicate
workovers.

It is very poor practice to use a natural perforated completion above a gravel pack.
The upper zone is invariably weaker and will produce sand that cannot be safely
washed out by through-tubing methods. The zone, therefore, is lost quickly, and
subsequent workover operations are unnecessarily complicated. However,
consideration could be given to consolidating the upper zone if casing size limitations
prohibit the use of multiple gravel packs.

As with other multizone completions, the designer must consider the trade-offs
involved with increased complexity, isolation problems, and less efficient gas-lift
operations.

Highly Deviated Wells


The percentage of highly deviated wells is increasing due to reduced economics for
multiple platforms and more advanced drilling techniques, especially in severe
environments and deep water. There are two major problems with gravel packing
highly deviated wells. One problem is incomplete packing of the perforations and the
annulus on the high side of the hole. The other is the tendency of the gravel to form
dunes, which bridge-off the screen-casing annulus, resulting in incomplete packing.
Laboratory studies by Maly et al. (1974) show that circulation packing efficiency
reduces significantly as the hole angle approaches 600 from the vertical, as illustrated
in Figure 1 .

783
Figure 1

As gravel settles to the bottom of the casing, gravel flow stops locally and a gravel
dune forms. As the dune grows, the flow becomes more and more restricted and the
flow velocity increases locally. This results in an equilibrium dune height. However, if
the flow velocities are low, resulting in a high equilibrium dune height, a premature
sandout could occur, leaving the downstream section poorly packed ( Figure 2 and
Figure 3 , Dune formation and its control by restricting flow in the liner).

784
Figure 2

Fluid leakoff into the screen or lower perforations can nucleate a sand dune. The
equilibrium velocity and dune height increases with inclination angle above about 45.
Below 45, settling gravel falls freely down the pipe due to gravity.

785
Figure 3

An additional problem in high-angle wells is that gravel settling during placement


results in a higher gravel density along the wellbore traveling along the bottom of the
pipe. Hence, the upper perforations receive mostly fluid. If the perforations are not
packed with gravel, formation sand fills the perforations and mixes with gravel in the
pack, reducing productivity or causing a sand control failure.

A number of approaches have been taken to improve packing efficiency in highly


deviated wells. Viscous slurry packing has led to improved efficiency, as discussed in
previous sections. In many cases, use of 80 to 120 lb/1000 gal HEC linear polymer
fluid with viscosities between 300 and 500 centipoise, and a lower telltale, is sufficient
to ensure a good pack in wells of up to 60, as reported by Elson (1982, SPE 11012),
although gravel concentrations are often reduced from the 12 to 15 lb/gal used on
near-vertical wells to 4 to 8 lb/gal in highly deviated wells.

For near-horizontal wells in very permeable, very weak formations, crosslinked


polymers are sometimes used to carry high concentrations of gravel into the
perforations. Crosslinked polymers are normally not recommended because of their
poor leakoff capabilities, but they do permit flushing of formation sand from the

786
perforation tunnels to create a behind-pipe cavity that has sufficient surface area to
permit leakoff.

Alternatively, higher fluid velocities can also be used with lower viscosity fluids to
increase packing efficiency in deviated wells with a combination of circulation and
induced losses. Moderate viscosities are used to minimize the risk of stoppages, which
can cause settling and, potentially, a premature sandout. The increased velocity
reduces duning tendencies and heights. The reduced viscosity gives better leakoff to
the formation. However, fluid leakoff at the screen must be controlled so that flow
along the hole and into perforations would be more predominant.

Gruesbeck (1979) has reported that the most important variable when gravel packing
inclined wellbores is the resistance to flow in the screen-tailpipe annulus. Fluid
escaping into the screen reduces flow velocities in the casing-screen annulus and
leaves gravel behind, thus forming a dune.

One technique that has increased packing efficiency by this principle is the baffled
wash pipe, or Unipack tool, reported by Maly (1974, SPE 4032). Deformable baffles are
placed along the tailpipe, as shown in Figure 3 , to reduce fluid flow between the
tailpipe and the screen. Baffle spacing and stiffness are the two critical parameters. If
the baffles are placed too far apart, duning can occur between them. If the baffles are
spaced too close together or are too stiff, excessive pressures can develop, resulting in
formation fracture, and making recovery of the tailpipe difficult. The baffles are flexible
to allow fluid flow through the screen-tailpipe annulus if a blockage occurs.

Nowadays the most popular method of increasing flow resistance in the screen-tailpipe
annulus is to enlarge the size of the tailpipe. Computer models are now being used to
optimize the ratio of the tailpipe OD to the screen ID. The wellbore geometry, fluid
properties, and pump rates are also optimized in the calculation. A ratio of 0.8 or
greater is a good general guideline. Enough clearance must be available to allow easy
movement of the tailpipe within the screen. Some researchers claim that by careful
sizing of the tailpipe, the lower telltale screen can be eliminated.

Epoxy-coated gravel has been used for prepacking of the perforations. This permits an
effective prepack of perforations on the high side of the hole. The coated gravel can be
squeezed into the perforations and drilled out of the wellbore prior to installing the
screen. This approach ensures a more complete packing of gravel into the perforations
but incurs the expense and problems associated with coated gravels. Prepacked
screens have been used to ensure sand control even if a uniform pack is not achieved.
However, to maximize productivity, gravel must still be placed around the prepacked
screen. This can present problems, but does provide insurance against a major failure
due to a void.

Some authors recommend perforating only the low side of the casing to avoid
problems with packing the upper perforations. This approach is fine if production
expectations are low, but in order to achieve maximum productivity, fully packed high-
density perforations on all sides of the pipe are required. The placement techniques
described above can achieve this with special care and proper design.

Quality Control

787
A successful gravel pack requires a great many things to go right at the same time. No
matter how brilliant the gravel pack design is, one error can lead to failure. When a
failure occurs it is very difficult to ascertain what has caused the problem. A superior
approach is problem prevention.

Quality control involves on-the-job inspection of equipment and materials to ensure


that everything is within design specifications. Experience has shown that materials
delivered to a site are not always within specifications and a good quality control
program can save time and money.

Gravel Quality Inspection

Gravel containing excess fines can severely impair well productivity. Gravel delivered
to the site can contain significant quantities of undersized gravel as reported by Boulet
(1979, SPE 7002). The gravel may have been undersized to begin with, or the fines
may have been generated by crushing or contamination. Oversized gravel also
damages the pack. Poor quality gravel may be partially dissolved by acid, or generate
additional fines going through the pump or traveling downhole. Gravel that contains
less than 96% quartz, is angular, or contains microcracks is considered poor-quality
gravel.

All gravel should, if possible, be resieved onsite prior to injection into the carrier fluid.
If this is not possible, then samples should be collected from the batch and tested on a
portable shaker and sieve unit. If the gravel is not within specification it should not be
used. This latter technique is not as good as resieving, since gravel quality can vary
within the batch and even small amounts of undersized material can cause damage.

Gravel solubility can be inspected on site using 12% HCl and 3% HF acid. A hand lens
or microscope can be used to check for hairline cracks and to evaluate gravel
roundness. As recommended by Zwolle (1983, SPE 10660), gravel crush resistance
can be measured by crushing the gravel under a predetermined triaxial load for two
minutes and noting the percentage increase in fines. Only minimal fines should be
generated under the loads expected during gravel packing. Zwolle recommends a
crushing stress of 13.8 MPa for the test. If the loads are high, stronger gravel, such as
bauxite, should be used. A more common onsite test is to attempt to crush grains by
hand between two coins.

Liner Inspection

The liner slot size must be within 0.001 in. (0.025 mm) tolerance of the specifications
to prevent gravel from penetrating the liner. This can be checked with a feeler gauge
or gap tester. The liner should be clean and efforts should be made to prevent the liner
from becoming plugged with dirt or grease, or being damaged during handling and
installation.

Fluid Cleanliness

One of the most challenging and critical tasks is to ensure that only clean fluids are
used during gravel packing. This task begins by inspecting and ensuring that all tanks,
surface lines, and tubulars are thoroughly cleaned to remove any solids that could end

788
up in the formation or gravel pack. The salts used to prepare brine often contain
insoluble materials which should be removed. Mixing salts with natural brines can
precipitate insoluble materials. Gelled brines can contain "fish eyes." All solids should
be removed prior to injection into the well.

As recommended by Sparlin (1985), the solids-removal process should include settling


tanks, shakers, desanders, desilters, and series filtration. There are a variety of filter
systems in use. For nominally rated cartridges, polypropylene filters have proved to
perform the best. For absolute-rated filters, diatomaceous earth is highly
recommended for filtering brines. For filtering gelled brines, a special shearing device
is available to break up fish eyes.

Pipe dope can cause permanent formation damage and thus should be applied
sparingly to the pin threads in making tubing connections. Pipe dope is often used
excessively; controlled dope application is a key quality-control consideration.

Probably the most contentious quality-control issue is determining when the well has
been "circulated clean" and whether the completion fluids are adequately clean. Most
operators recommend reducing the solids content to the range of 200 to 360 ppm
(mg/lt). This corresponds to 50 to 100 NTU (Nephelometer Turbidity Units) and cannot
be determined by a shake-out test. If a turbidity meter or Millipore filter tester is not
available, settling tests should be performed with large volume samples, or the fluid
should be passed through filter paper and the paper examined under a hand lens or
microscope.

The viscosity of the hydrated and sheared carrier fluid should be tested with a Fann
viscometer at low rotational speeds prior to adding the gravel.

789
FRAC-PACK TECHNICS

"Frac-and-Pack" Methods of Sand Control


Frac-and-pack completion methods, so-called because they incorporate hydraulic
fracturing and gravel packing into a single well treatment, are designed to create
relatively short, highly conductive fractures in reservoirs of moderate to high
permeability. They offer the advantages of fracture stimulation combined with the
benefits of sand control, and can be used to:

bypass near-wellbore damage that matrix stimulation treatments cannot remove;

increase formation support in compacting reservoirs, where there is potential


for severe casing failure in perforated intervals;

vertically connect productive intervals in thin, laminated sand-shale


sequences;

improve productivity in some low-permeability reservoirs;

alleviate problems, such as asphaltine deposition, that arise from pressure-


dependent processes;

control sand production and fines migration in poorly consolidated or


unconsolidated reservoirs.

A frac-and-pack treatment is a two-step process:


1. Fracture growth followed by tip screen-out (TSO): The pumping schedule is designed
such that the proppant slurry reaches the tip of the fracture early in the treatment, resulting
in a screenout that prevents additional fracture growth.

2. Fracture inflation and packing: Following TSO, continued proppant injection


causes the fracture to widen and the net pressure to increase.

In moderate-to-high permeabilitty reservoirs, the basic treatment objective is to maximize the


fracture's propped width, and thus its conductivity.

Frac-and-Pack as a Sand Control Technique

Sand control has traditionally involved placing some sort of filter (such as a gravel
pack) between the well and the formation. The problem with this approach is that a
gravel pack's effectivenesslike that of any other filterdeteriorates over time as it
collects solids. Fines migration and lodging cause progressive damage within the pack,
resulting in a production rate decline. Operators may try to remedy the problem by
increasing the pressure drawdown on the well; unfortunately, this increased drawdown
accelerates sand and fines migration, further damaging the gravel pack. Matrix
stimulation treatments are ineffective at removing such damage.

790
A properly designed and well-executed frac-and-pack does more than simply filter solid
particles. It increases the effective wellbore radius (Prats, 1961) and significantly
reduces near-wellbore flow velocities. Thus, a frac-and-pack completion can produce at
the same rate as that of a conventionally gravel-packed well, but at a much lower
drawdown, thereby avoiding the release of formation stresses that can cause sand
production.

Well Selection Criteria

Roodhart et al. (1994) outlined seven general considerations for determining whether
a well might be a frac-and-pack candidate, which are summarized as follows:

Frac-and-pack is not a good treatment option for wells in highly depleted reservoirs, wells
having high water/gas cuts, or wells producing in close proximity to the oil-water or gas-oil
contacts.

Frac-and pack may be considered in wells having a high positive skin factor (in
new wells, this criterion will be based on anticipated damage).

Frac-and-pack may be effective in bypassing near-wellbore skin, particularly


where previous acid treatments have not been effective.

Inflow performance simulation may be used to evaluate pre-treatment versus


post-treatment production rates, based on a comparison with equivalent wells
completed in the same formation. For example, frac-and-pack may be an option
for vertically connecting thin productive layers in laminated reservoirs, in a way
that would not be possible through perforating alone.

Mechanical properties of the formation (sand production tendencies, effect of


drawdown reduction) should be considered in evaluating frac-and-pack as an
alternative to gravel packing.

Frac-and-pack fluid and proppant requirements are based on reservoir


permeability and pay thickness.

A treatment's suitability depends largely on the well's mechanical integrity,


including its ability to withstand increased packer forces resulting from the
injection of cold fracturing fluids, the presence of a tubing anchor or seal
receptacle, the adequacy of the perforations, the wellhead and tubing pressure
limitations, and the effects of corrosion.

As part of the SPE Technology Today series, Meese et al. (1994) describe process applications
and operating parameters for frac-and-pack treatments being performed in the U.S. Gulf Coast
region, and list several concerns and possible limitations related to offshore frac-and-pack
techniques. Along with the problem of fracturing high water/gas-cut wells mentioned above, they
cite the following:
In small-diameter casings, erosional effects on downhole tools may limit the injection rate
available for creating the hydraulic fracture.

791
Injection rate restrictions may make it difficult to stimulate long intervals in a
single treatment, which could result in improper packing of the casing/screen
annulus.

At the time of their study, the demand for stimulation vessels equipped to
perform frac-and-pack treatments significantly exceeded their availability,
making for potential logistics problems and lost rig time while waiting on
equipment.

Basic Treatment Guidelines

The fundamental objective of a frac-and-pack treatment is to maximize the


dimensionless fracture conductivity (FCD), which is defined as:

FCD = (FC)/(RC) = (wfkf)/(Lfk)

where:

FC = fracture conductivity = wfkf

RC = "reservoir" conductivity = Lfk

wf = propped fracture width

kf = fracture permeability

Lf = fracture propped half-length

k = formation permeability

When FCD = 1, the fracture flow capacity is equal to that of the reservoir. FCD < 1 indicates that the
fracture is limiting well productivity, while
FCD > 1 signifies that the fracture is enhancing productivity. In formations having moderate-to-high
permeability, maximizing FCD is primarily a matter of maximizing the propped fracture width.

Wong et al. (1993) set forth requirements for optimizing frac-and-pack treatments
based on a study of ten jobs performed on wells in the Gulf of Mexico. The study
included a skin analysis of frac-and-pack completions, which showed that near-
wellbore skins (from perforation tunnels and a choked fracture) can be particularly
damaging when FCD values are low and perforations are not filled with permeable
proppant. The authors therefore recommend the following general guidelines:

The treatment should be designed to attain an FCD of greater than one, in order to reduce
near-wellbore skin resulting from a choked fracture.

To minimize perforation skin, the perforation should be packed with high-


permeability proppant (i.e., ratio of proppant permeability to formation
permeabilty greater than 200). (Note that the proppant sizes required to attain
this permeability may conflict with sand exclusion requirements for gravel

792
packing. However, as Roodhart et al. (1993) point out, the lower flow velocity
and drawdown resulting from a successful frac-and-pack may reduce the sand-
exclusion criteria that a gravel pack alone would require.)

The authors also recommend that steps be taken before and after the treatment to further reduce
the impact of near-wellbore skin effects. These include perforating with high shot density, large-
diameter guns, breaking down and initiating the fracture with high pump rates to reduce near-
wellbore friction, maintaining sufficient control over the pumping schedule to ensure development of
a tip screenout (TSO), packing the fracture with high concentration loading of proppant near the
wellbore, optimizing fluid design and quality control, reviewing the gravel packing operation with the
fracture operator, and accurately measuring slurry volumes.

Attaining a TSO is key to successful fracturing in moderate-to-high permeability


formations. Because this requires strict control over the pumping schedule, it is
essential to conduct pre-job step-rate and pressure decline or calibration tests to
determine the fracture extension, closure and instantaneous shut-in pressures, along
with the closure time and fluid-loss coefficient (Meese et al., 1994). Such tests can
help refine treatment designs obtained from computer simulations done during the
treatment planning stage.

793
CONSOLIDATION TECHNICS

Chemical Consolidation

An alternative approach to mechanical methods of sand control in poorly consolidated


sands is to strengthen or consolidate them by injecting chemicals into the formation.
The objective is to "cement" the grains together at the grain-to-grain contact points to
provide stable compressive strength, while maintaining a high fraction of the original
permeability.

Chemical consolidation has been performed for many years with various types of
chemicals and processes. It is the most complex sand control process. There is
significant risk involved of inadvertently plugging the pay and/or of ineffective chemical
placement. The consolidation technique usually involves multistage injection of several
different chemicals into the formation, each chemical carefully prepared to suit the
specific conditions. It involves a high level of design and supervision.

In most cases, quality control, job supervision, and cleanliness are more important
than the particular consolidation system used. Experience indicates that consolidation
is best suited for thin (<l2-ft, or <3.6-m), clean, homogeneous, undamaged sand
zones.

The following are some of the advantages of chemical consolidation:

Consolidation can be done through tubing as a remedial technique.

It can be performed without a service rig, which can reduce the overall cost.

It is more suitable than gravel packing for abnormally pressured zones (high
or low) because it is not as dependent on fluid circulation and can be done
under pressure.

In multiple reservoir completion, consolidation allows sand control on the


upper intervals without sand control equipment in the borehole, which would
hinder completion and workover operations on the lower zones.

Consolidation does not inhibit reservoir control by through-tubing techniques,


such as plug backs or squeezing perforations.

Consolidation should be more effective than gravel packing for formations with
wide grain-size distributions, which are difficult to control with gravel packing.

Grain-to-Grain Cementation

In consolidated formations, grain-to-grain cementation is provided by silica, clay,


calcite or other materials. In unconsolidated formations these cementing materials are
scarce or missing. The purpose of consolidation is to deposit chemicals at the grain-to-

794
grain contact points to provide adhesion between the grains while still keeping the pore
channels open. A schematic view of consolidation is shown in Figure 1 .

Figure 1

The consolidation material must adhere to the sand grains and form a strong bridge
between the grains. Since the formation is usually water-wet, the wettability must first
be converted to an oil-wet state and the formation water displaced. From
thermodynamic considerations, the resin material will preferentially deposit at the
contact points between grains. However, special precautions must be taken to keep
the pores open and in communication.

The degree of cementation is measured in terms of compressive strength. Usually the


consolidation technique is simulated in the laboratory under ideal conditions, and both
the compressive strength and permeability is measured. Compressive strengths from
2400 to 7000 psi (20 to 50 MPa) have been reported with retained permeability
between 50 and 90%. Normally there is a trade-off between compressive strength and
retained permeability, since more deposit results in more strength but less free pore
space. Also, most deposited material is oil-wet, which reduces relative permeability to
oil.

Great care must be taken in interpreting laboratory data. Laboratory tests are
performed under idealized conditions, whereas field conditions are often not ideal.
Also, since laboratory procedures have not been standardized, the data may be

795
misleading. The durability of the consolidation under exposure to produced fluids is
critical to the success of the completion. In many cases the consolidation was found to
deteriorate quickly with time after production began. Some laboratories have
investigated durability to produced fluids and this has led to improved consolidation
systems.

Chemicals Used

The materials used for cementing the grains are usually organic-based, and include
oxidized polymers and various polymerizable resins, such as phenolformaldehydes
(Bakelite), furans, and epoxies. Nickel plating has been used, but with limited success
and great expense. Coking of the oil in place by injecting warm air (SPE 1239) has
been attempted, but more research is required. A new technique using silicon
tetrachloride in a gaseous phase to form silica cement has been developed with some
success for gas wells (Davies et al., JPT Nov. 1983). The furan and epoxy resin
systems have achieved the most success over the years. The newly developed oxidized
polymer process (Burger et al. 1986) and the silica cement process require further
testing in order to become established. The chemicals involved in the
phenolformaldehyde and silica systems are quite toxic.

For the resin-based systems, the resin is injected in a liquid state and a curing agent
or catalyst (activator) is used to solidify the deposit. A multistage preflush is used to
displace reservoir fluids, dissolve fine particles, and to oil-wet the sand surface.
Treatment usually consists of an HCl preflush; mud acid, to remove clays; diesel, to
remove residual oil; alcohol or mutual solvent, to remove water; and a spacer fluid to
prevent contact between the resin and the alcohol. Isopropyl alcohol (IPA) acts as a
mutual solvent for brine and oil, but care must be exercised since IPA causes NaCl to
precipitate when the IPA-water ratio exceeds 4. The epoxy system requires an IPA
preflush to ensure all water is removed. Viscosifiers are often added to the preflush
fluids, especially mud acids, to prevent fingering and ensure a more uniform
displacement.

There are two types of resin-based systems:

The phase separation process, or one-step process, involves combining the resin and
activator at the surface and injecting the mixture containing about 15 to 25% resin
into the formation. The resin is attracted to the grain-to-grain surface by capillary
forces and separates out of solution, leaving the center portion of the pore volume
filled with the carrier fluid. This system is also referred to as an internally catalyzed
system, since the activator is premixed with the resin.

The overflush process, or two-step process, involves injecting a high concentration of


resin solution into the formation and then injecting an overflush solution to restore
permeability by displacing all but the film of resin solution at the grain-to-grain
boundaries. The overflush solution is usually hydrocarbon-base, but may be water-
base. The activator is usually contained in the overflush solution, but may be premixed
with the resin solution. This system is usually referred to as an externally catalyzed
system, since the activator is injected after the resin is in place.

796
The resins solution comprises resin, solvent, coupling or curing agent, and perhaps an
activator. In dirty sands with high-surface-area materials, such as clays, excess
concentrations are required because of adsorption onto the surface.

The phase separation process is less attractive, since there is a limited pumping time
before the resin hardens and the chemicals require mixing at the wellsite. For the
overflush process the activator can be quick-acting, thus speeding the curing time.
Also, the overflush process allows for higher-strength consolidation and higher
permeability. However, for the over-flush process, viscosity control on each solution is
essential to ensure that the resin solution is effectively displaced.

Most furan-based resin systems are overflush processes, while most phenol-based and
epoxy-based systems are phase separation processes. Recent developments with
epoxy-based systems (SPE 6803, Shaughnessy et al.) have produced a low-viscosity
system utilizing the overflush process.

For the silica cement system ("Silicalock"), the technique consists of vaporizing liquid
silicon tetrachloride, SiCl4, into a stream of high-pressure nitrogen gas and injecting
the mixture into the well through coiled tubing. The SiCl4 reacts with residual water to
form silica cement and HCl. This technique is attractive, since it involves only one
chemical, requires no fluids, doesn't require a rig, and the reaction is instantaneous.
The cement dissolves slightly in water, but this may not be a problem in gas wells.

For the oxidized polymer system ("Oxpol"), the formation is first saturated with
polymerizable organic material and catalysts and then oxygen is injected at controlled
rates. The oxygen attacks unsaturated bonds of the organic compound and forms
polymers that adhere to the sand surface. This method is similar to warm air coking,
since it involves an exothermic reaction. A preflush of alcohol is required to remove
any residual water.

Treatment Design Considerations

Since chemical consolidation processes are quite complex, strict quality control must
be exercised during all aspects of the well completion process. The guidelines provided
below are specific to resin-based systems, since these systems are much more
established than the other systems, which are still in the research and development
stages.

Cleanliness is critical, since any "dirt" may contaminate the chemicals and plug the
formation. Large volumes of fluid are injected into the formation and thus small
concentrations of solids can result in severe damage. All equipment should be
thoroughly cleaned and inspected before the treatment.

Drilling fluids cause formation damage, which, if not removed, will divert the treating
fluids, leaving parts of the formation untreated. Clean completion fluids should be
employed, and acidizing should precede the treatment to remove any damage.

Primary cementing is critical to confining the treating fluids to the perforated zone.
Poor cement jobs leave open channels for the treating fluids to escape, and the loss of

797
part of the treating fluid jeopardizes the entire completion. Hence, superior cementing
procedures are required.

Perforation density of at least four shots per foot is recommended in order to maximize
productivity. Open perforations are essential to ensure vertical distribution of the
treating fluids. Otherwise, untreated pockets of sand will move during production, thus
destroying the whole treatment. Significant improvements in productive life have been
reported by Schroeder (1974) by increasing the perforation density from 2 to 4 shots
per foot.

Perforation interval lengths between 6 and 12 feet (2 to 3 m) per treatment stage are
recommended, since fluid can be effectively distributed over small intervals by high
fluid pressures at the wellbore. Longer intervals would require diversion for effective
fluid placement and risk being unable to remove the diverter. Longer intervals should
be treated in several separate stages. Perforations should be deep-penetrating to pass
through any drilling damage. Perforation debris should be removed by any of the
various methods of underbalanced perforating, backsurging, backflowing, washing, or
stimulation. The backsurging or the tubing-conveyed perforating system is ideally
suited for this purpose. Perforating should be performed with very clean fluids, filtered
to less than 2 microns.

Rathole size should be minimized or the rathole fluid should be of high density to
prevent gravity segregation. Treatment fluids with densities higher than the rathole
fluid may displace the rathole fluid into the perforations and jeopardize the treatment.
Gravity segregation in the rathole due to the high density resin solution has been
reported by Hong (1979) as a major problem in resin sand consolidation.

Packer location should be as close as possible to the upper perforation interval to


prevent collection of low-density fluids in the dead space below the packer. Gravity
segregation of alcohol into the dead space must be accounted for in designing the
alcohol volume.

Fluid viscosities should be designed to minimize injection pressures while ensuring


uniform displacement. High-viscosity fluids require high pumping pressures and risk
fracturing the formation, which should be avoided. As demonstrated by Davies et al.
(1984), low-viscosity (high-mobility) fluids cause unstable displacement fronts and
gravity segregation both in the tubing and the reservoir ( Figure 2 , Effect of viscosity
on injected fluid stability).

798
Figure 2

Medium-viscosity fluids of about 30 cp have been recommended for uniform


displacement and acceptable injection pressures. Also, the mobility ratio, which is the
ratio of displacing fluid viscosity to displaced fluid viscosity, plays a key role in
displacement stability. Low mobility ratios (<1) are preferred. Furthermore, the
formation permeability must be considered in the design of the fluid viscosity, since
lower permeabilities dictate lower fluid viscosities. Plugs should be used to keep fluids
separate as they travel down the tubing.

Injection rates should be low to promote uniform displacement and to prevent


fracturing the formation. High injection rates cause unstable displacement fronts,
which result in fluid fingering.

Surface treating pressure should be monitored and recorded as a means of controlling


and evaluating the treatment. The injection rate of the preflush fluids is a good
indication of formation damage and provides a means to design the injection rate of
the treatment. Also, whether the formation was fractured or not can be determined
from the pressure response using techniques that are currently applied to hydraulic
fracturing.

799
Treatment volumes should be designed to consolidate the formation around a 3-ft (l-
m) radius from the wellbore. Recommended treatment volumes are about 100 gal/ft
(1.2 m3/m). For dirty sands or poor confinement, larger volumes should be used.
Careful control of the displacement volume is required in order to avoid displacing the
treatment fluid beyond the perforations. Overdisplacement is more detrimental to
phase separation systems than to overflush systems, but it should be avoided in both.

Preflush volumes should be about 100 gal/ft (1.2 m3/m) to remove water from
reservoir sand up to a radius of 3 ft, as discussed by Penberthy et al. (1978, SPE
6804).

Formation temperature should be measured before treatment in order to design the


fluid viscosities for bottomhole conditions. Also, the effectiveness of most catalysts is
affected by temperature.

Previous experience in the area is extremely valuable for designing a treatment. Since
there are still many unknown parameters affecting the treatment, results from
previous treatments can provide useful guidelines.

Laboratory studies should be conducted on cores from the formation, especially when
no previous experience has been obtained in the area. Care should be exercised in
interpreting the data, since laboratory tests are usually performed under ideal
conditions. Reasonable safety factors must be built into the design.

Treatment Problems

A major complaint of sand consolidation is that the well will suddenly start to produce
sand, without warning, after a few days or weeks. Attempts to reconsolidate after
failure have little success.

In the past, many problems in sand consolidation were resolved by using cleaner fluids
and by thoroughly cleaning the equipment. Although perforation and perforation
cleaning techniques have improved, plugged perforations is still a major problem.
Extra care should be taken to ensure all perforations are open, since plugged
perforations tend to open during flowback operations or subsequent production,
allowing sand production from the untreated sand.

Many problems occur on surface, since many chemicals have to be prepared and
blended. Contamination of chemicals is a major problem. An experienced crew is
essential for a smooth operation. There often is a tendency to skip key operations or
reduce quality control in order to get the job done in a reasonable time. Careful
preplanning and job simplification can reduce most problems.

Consolidated Gravel Packs

Resins are sometimes used to consolidate gravel. This can be done in two ways: either
the gravel is precoated with a temperature-dependent epoxy, or the gravel is coated
with a resin, blended with the carrier fluid as the slurry is pumped. The activator is
either mixed with the resin on surface or applied with the overflush technique. After
curing, the remaining material inside the casing can be drilled out. No liner is required,

800
but one is sometimes used for insurance. Consolidated gravel is also used for gravel
packing injection wells where the pack has a tendency to be displaced into the pay
and/ or dissolved by the injected fluids.

Applications

The technique is well suited for formations that have cavities due to previous sand
production. It is also sometimes used for the upper zones in multiple completions,
since it eliminates mechanical equipment in the well-bore. In cases where gravel tends
to move in an unstressed formation, or in an injection well, consolidated gravels give
greater pack life.

Consolidated gravel can also be used for repairing failed gravel pack or failed
consolidation completions by through-tubing methods, with coiled tubing being used
for placement and drill-out. Intervals of 30 ft and longer have been successfully
treated.

Design Considerations

The chemicals used in consolidated gravels are the same as the resin-based systems
used for chemical consolidation. Therefore, the same precautions and design
considerations apply. High viscosities are used for the resin to prevent it from leaching
out of the gravel into the formation.

The resins used for precoated gravel have to be specified for the formation
temperature to avoid premature curing and to ensure an effective in-situ consolidation.
In some shallow zones, hot fluid circulation or injection is needed to ensure an
effective cure and adequate compressive strength.

Gravel size should be determined in the same manner as for normal gravel pacing.
Larger gravel leads to fines invasion, less strength, and a greater tendency for the
resin to be washed away. Greater compaction and higher gravel angularity yield
stronger packs but reduced permeability.

Carrier fluid viscosity is usually high (500 to 1500 centipoise), to suspend high gravel
concentrations, control leakoff, and displace the formation sand without allowing it to
mix with the gravel. Care must be taken not to exceed fracture pressures, although
some companies recommend fracturing the formation with the gravel.

Gravel concentrations as high as 15 lb/gal have been used to reduce treating volumes.
Generally, enough slurry is squeezed through the perforations in one stage to obtain a
screenout. The technique is similar to the squeeze packing technique used for normal
gravel packing. Subsequent stages may be used if a screenout did not occur in the first
stage.

Fluid injection stages in sequence typically include acid, to clean perforations; isopropyl
alcohol, to remove water; diesel, to remove the alcohol; spacer, to isolate the resin
from the diesel; gravel slurry; spacer; diesel; and finally a displacement fluid.

801
Perforating practices established for normal gravel packing apply to consolidated gravel
packing. Large, round holes and high shot densities are recommended. Perforations
should be completely cleaned of debris, using backsurging and washing techniques.

Problems

Because of the consolidation chemicals, similar problems are experienced with


consolidated gravel and with consolidation without gravel. However, higher success
ratios are usually achieved with consolidated gravel, especially with the precoated
systems.

With the fluid consolidation systems, the chief concern is that the resin may leach out
of gravel and plug formation sand.

With the precoated systems, activation time is difficult to control downhole, and
premature setting of the resin on the gravel leads to poor packing efficiencies and low
consolidation strengths.

802
SAND CONTROL IN WELL PRODUCION HEAVY OIL

Sand Control for Heavy Oil Wells

Most heavy oil and tar sand formations are unconsolidated. The high crude viscosities
exert large flow forces on the formation sand, promoting high sand production rates.
Most heavy oil wells are pumped, and the cyclic stresses at the wellbore caused by the
pumping action reduce wellbore stability. Furthermore, many heavy oil wells are steam
stimulated or steam flooded, which significantly complicates sand control.

Conventional sand control techniques, especially internal gravel packing, often result in
loss of the already low productivity in heavy oil wells. Moreover, the projects are so
marginal and sensitive to capital costs that sand control may be difficult to justify
economically.

No Sand Control

In cases where sand production rates are low and steady, it is often more attractive to
produce the sand than to control it downhole. The viscosity of the crude or crude-water
emulsion is usually high enough to carry the sand to the surface without serious
erosion. Most heavy oil wells require artificial lift, such as tubing or screw pumps, and
the major problem with producing sand is erosion of the pump components. Sand
particles tend to erode the valves and seats. Recently, pumps have been designed to
handle sand production, but these pumps have limited head capacity and/or are still in
the development and field trial stage.

The major problem with no sand control is that unexpected shutdowns can cause a
sandoff of the string. Moreover, this occurs with changing fluid viscosities during
steaming, or with water breakthrough. Surface separation and disposal of the sand is
also a problem in some areas, although pretarred sand can be useful for road-building
projects.

Gravel Packing

For clean formation sands with narrow grain-size distribution, gravel packing is still
widely used. However, the pressure drop across the perforation tunnels precludes the
use of IGPs in many primary completions on high-viscosity oils. External, or openhole,
gravel packs (EGPs) are therefore primarily used in heavy oil areas, such as in
California and Venezuela. Internal packs can be used in thermal projects, although
they limit the duration of the production cycle.

Steam stimulation, or steam flooding, poses a significant problem to gravel packs.


Steam generator effluents are quite alkaline, with pH values ranging from 11 to 12.5.
This high alkalinity results from bicarbonate ions, present in most feedwaters, that
decompose during steam generation to produce carbonate ions and hydroxide ions.
Silicate materials, such as gravel, dissolve in hot alkaline water, so that gravel packs
frequently fail in the third or fourth production cycle. Also, the liner must withstand the

803
harsh corrosive environment of a huff 'n' puff steam project. It is also subject to
substantial mechanical loading due to the thermal cycling.

The weigh loss of siliceous material has been shown by Underdown et al. (1983) to
increase dramatically with temperature and pH. The higher the quartz content of
gravel, the longer the gravel lasts in the harsh field environment.

Sintered bauxite gravels have shown very little weight loss when exposed to high
temperatures and high pHs. Also, moderate compositional differences make very little
difference in weight loss for sintered bauxite gravels. Resin-coated gravels have
achieved good success in California at resisting steam dissolution. Improvements in
resin chemistry and manufacturing processes have increased the stability of resin-
coated gravels, but they are still unsuitable at high pHs.

Screens and Slotted Liners

Since many heavy oil wells produce from formations that are essentially gravel sized,
screens can be used for sand control. Wire-wrapped screens are the most popular
method of sand control in Canada's heavy oil wells, where economics and geological
conditions limit EGP applications. Slotted liners are still used, but rarely. However,
there are still many problems with wire-wrapped screens. Plugging of the screens is a
major problem for dirty sands. This has been demonstrated in laboratory tests where
clays and silts often collect at the screen.

Partial plugging eventually leads to erosion, due to the increased flow velocities in the
unplugged sections. Erosion caused by high flow velocities also occurs during steam
breakthrough or preferential steam injection into thief zones. This problem is
significantly exacerbated by the corrosive nature of the thermal fluids. The fluid
velocity appears to be the major parameter affecting screen erosion, and thus the use
of openhole completions to reduce flow velocities is often favored. Some screens last
for years while others last only for a few weeks. Steam breakthrough appears to be
another key factor influencing screen failure for steam floods.

Screen sizes are usually larger than prescribed by the standard formulas in order to
reduce flow velocities and, hence, erosion. The production of some sand is not so
serious, due to the high fluid viscosities and generally low production rates. Erosion of
the screen resulting from flow restriction is a more serious problem. Experience usually
dictates the specific screen size.

Damaged screens are difficult to remove, so screens should be designed to be easily


washed over during fishing operations. The screens must be kept clean prior to
installation.

Consolidation

Sand consolidation has not been very successful to date in heavy oil wells. The
problems of effective fluid displacement are much higher in heavy oil wells, especially
with dirty sands. There has been some consideration to using warm air coking of the
heavy oil to consolidate the sand, but more research is required to prove the viability
of this system.

804
Resin-coated gravel using furan resins and the overflush technique has achieved some
success in California. However, special laboratory tests are required to select a resin
coating that can withstand the harsh steam environment.

805
HORIZONTAL WELLS COMPLETION & EVALUATION

HORIZONTAL WELLS RESERVOIR STRATEGY

Horizontal Wells in a Reservoir Management Strategy

Horizontal wells are nothing new. Engineers in the old Soviet Union drilled a number of
them in the 1950s, although they ultimately abandoned the practice as uneconomical.
The general concept of horizontal drilling dates back even earlier (Ranney, 1939). It
wasn't until the early 1980s, however, that two western petroleum companies (Elf and
AGIP) established horizontal wells as viable substitutes for vertical wells in certain
types of reservoirs.

Since then, the industry has drilled thousands of horizontal wells worldwide, in a large
array of reservoirs. Horizontal sections of more than 3500 ft. [2700 m.] are
commonplace, with world records exceeding 35,000 ft. [~10700 m.] as of 1999.
Moreover, multilateral and multibranch completions, in which two or more horizontal
sections are drilled from a single vertical wellbore, have become increasingly popular.

As often happens in the petroleum industry, developments have been fast-moving in


some aspects of horizontal well technology, and very gradual in others. The most
spectacular growth has occurred in the areas of drilling, directional control, reservoir
engineering and formation evaluation:

Drilling and wellbore steering capabilities have matured such that drillers can maintain very
small thicknesses (e.g., 2-3 ft. [less than 1 m] ) for ever-increasing lengths. Thanks to real-
time formation evaluation technology, it is possible to steer within a desired layer even
when it is dipping.
Reservoir engineering applications have likewise matured. We have established, for
instance, that horizontal wells are generally attractive alternatives to vertical wells in
reservoirs having good vertical permeability (e.g., naturally fractured carbonates). For
relatively small reservoir thicknesses say, 50 ft. [15.2 m.] or less vertical permeability
is not as important.
We're also gaining a better understanding of formation evaluation criteria. We have
identified a number of variables as crucial for horizontal wells, including vertical
permeability, horizontal permeability, stress anisotropy and in long horizontal wells the
presence of lateral heterogeneities and discontinuities.

To determine these variables, the industry has developed new and revamped well
testing and logging methods. It has also made increasing use of borehole and 3D
seismic images and very importantly data integration techniques.

Horizontal wells have themselves proven invaluable formation evaluation tools. They
can provide data on lateral heterogeneities, the presence of faults and the position and
extent of fluid contacts, shale lenses and so forth, which would be difficult or
impossible to obtain from vertical wells.

806
Well completion and stimulation practices, by contrast, have evolved less dramatically.
For example, it's become clear only in relatively recent years that unstimulated or
poorly stimulated horizontal wells or those completed without regard to stimulation
needs may underperform vertical wells in the same reservoir.

During the early stages of the 1980s' horizontal drilling boom, several failures were
attributed to poor reservoir selection. Today, while this issue has been largely
resolved, we can blame many lackluster horizontal wells on inadequate completion or
stimulation practices.

Horizontal Well Deliverability

Under the right conditions, a horizontal well exhibits a much-improved productivity


index over that of a comparable vertical well. This can result either in a substantial
production rate increase at constant pressure drawdown, or a greatly reduced pressure
drawdown at constant production rates. Reduced pressure drawdown is particularly
beneficial in reservoirs subject to gas or water coning and, more recently, in reservoirs
with sand control problems.

Joshi (1988) has introduced, and Economides et al. (1991) have augmented, the
following expression for horizontal well deliverability:

(1.1)

where

q = flow rate (STB/D)

kH = horizontal permeability (md)

h = reservoir thickness (ft)

pe = reservoir pressure at outer flow boundary (psi)

pwf = flowing bottomhole pressure (psi)

B = formation volume factor (Bbl/STB)

= viscosity (cp)

L = length of well's horizontal section (ft)

807
rw = wellbore radius (ft)

Note: Where q is expressed in m3/d (with k in md, pe and pwf in kPa, B in m3/m3, in cp and h, L
and rw in m), 141.2 becomes 1867.

Equation 1.1 contains two other important variables: the index of horizontal-to-vertical
permeability anisotropy (Iani), defined as

Iani = (1.2)
where kv = vertical permeability (md)

and the large half-axis (a) of the drainage ellipse formed by a horizontal well, which
equals

(1.3)
where reH = drainage radius of the horizontal wellbore (ft)

Figure 1 (Horizontal well parameters )depicts two of the most intuitively obvious
differences between horizontal and vertical wells: the role of vertical permeability and
the elliptical shape of the formed drainage.

808
Figure 1

Example: Horizontal Well Production

Suppose that a reservoir has a horizontal permeability (kH) of 5 md and a thickness (h)
of 75 ft. Using the following data, plot production rate vs well length for Iani values of
5, 3 and 1. Compare the performance with a vertical well in the same reservoir.

pe = 5000 psi

pwf = 3000 psi

A = 640 acres = 27,878,400 ft2

B = 1.1 Bbl/STB

= 0.7 cp

rw = 0.328 ft.

809
Solution:

For the three values of Iani, and using Equation 1.2, the vertical permeabilities, kv are
0.2, 0.56 and 5 md, respectively. The drainage radius (reH) for A = 640 acres is 2980
ft.

If, for example, L = 2000 ft, then from Equation 1.3 above

= 3065 ft. (1.4)


Then, applying Equation 1.1, for Iani = 5

=2490 STB/D [396 m3/D] (1.5)

By comparison, a vertical well, whose state-steady production is given by the well-


known

(1.6)
would produce 757 STB/D [120 m3/D].

Permeability Anisotropy Effects

The term

in Equation 1.1 is of particular interest, since it is the only term affected by the vertical-to-horizontal
permeability anisotropy.

Note that Joshi's modified equation is limited to describing single horizontal wells, and
does not account fot area permeability anisotropy, which has been found to be an
important variable--particularly with respect to multilateral wells (see Smith et al.,
1995). For a more comprehensive model, we refer the reader to Economides et al.
(1994). Joshi's model, however, still serves as a good initial approximation of
horizontal well deliverability.

Referring back to Example 1.1, the first logarithmic term is equal to 1.79 in each case.
In contrast, for differing Iani values of 5, 3 and 1, the second term values become

810
0.984, 0.579 and 0.178, respectively. Thus, a well in a reservoir of Iani =1 would
produce 1.4 times the rate attainable from the reservoir where Iani = 5.

Figure 2 (Effect of permeability anisotropy (Iani) on horizontal well production) shows


production rates from horizontal wells in three reservoirs having different permeability
anisotropies.

Figure 2

The section length ranges up to 5000 ft [1524 m]. It also shows the production from a
vertical well, whose performance (in such relatively small thickness) does not depend
on the vertical permeability.

Reduced Pressure Drawdown

As we noted earlier, a horizontal well affords reduced pressure drawdown over a


comparable vertical well.

Referring again to Example 1.1, we see that the vertical well production rate of 757
STB/D [120 m3/D] corresponds to a pressure drawdown of 2000 psi [13789 kPa]. If we
maintained the same production rate for a horizontal well where L = 2000 feet [609.6
m], the pressure drawdown would be [(2000)(757/2490)], or 608 psi [4688 kPa]. The

811
flowing bottomhole pressure would be 4392 psi [30281 kPa], instead of 3000 psi
[20684 kPa].

In reservoirs with coning or sand production problems, the ability to reduce drawdown
while maintaining constant production can be a distinct advantage.

Effect of Reservoir Thickness

Figure 3 (Effects of permeability anisotropy and reservoir thickness on horizontal well


productivity),which plots productivity index ratios (horizontal well/vertical well) as a
function of lateral section length,

Figure 3

illustrates the effect of reservoir thickness on horizontal well deliverability.

In addition to the base case (h=75 ft [22.86 m]), it shows P.I. ratios for reservoir
thicknesses of 25 ft [7.62 m] and 225 ft [68.58 m].

We can see from Figure 3 that:

812
while permeability anisotropy is crucial for thick reservoirs (as shown by the distance
between the curves of the different Iani values), it is less important for thin reservoirs
note the relatively closer spacing of the three Iani curves.

for each Iani value, the thinner reservoirs exhibit the largest productivity
index ratios.

Effect of Damage on Horizontal Well Production

Up until now, our calculations have assumed an undamaged horizontal well producing
from its entire length. This assumption does not accurately describe real well behavior.
For now, it is useful to study the effect of near-wellbore damage on horizontal well
performance. We characterize this damage, as we do in vertical wells, by means of a
skin effect (seq).

We can incorporate this skin effect into Equation 1.1 as follows:

(1.7)

This skin effect may easily have a value as high as 10 or even 20. Although we
multiply it by the scaled aspect ratio (Ianih/L), it nevertheless can drastically reduce a
horizontal well's production rate. For instance, returning once again to Example 1.1,
remember that the calculated horizontal well production rate was 2490 STB/D. For an
seq value of 10 (where Iani = 5, h = 75 ft and L = 2,000 ft) the rate falls to 1483
STB/D, representing a 40% production loss.

Figure 4 (Effect of near-wellbore skin damage on horizontal well productivity) shows


productivity index ratios for a 2,000 ft [609.6 m] damaged horizontal well compared
with an undamaged vertical well over a range of skin effects,

813
Figure 4

again using permeability anisotropy ratios of 1, 3 and 5. It is evident that a damaged


horizontal well may not only fail to attain its expected productivity index increase, but
that it may even underperform a fully stimulated vertical well.

The reality of near-wellbore damage and its detrimental effects on production make it
clear that stimulation is important for both vertical and horizontal wells.

Horizontal Wells in Anisotropic Formations


In studying the effect of horizontal-to-vertical permeability on well performance (e.g.,
Figure 2 , Effect of permeability anisotropy (Iani) on horizontal well production), we
may assume for illustration purposes that the reservoir is isotropic in the horizontal
plane.

814
Figure 2

This is not always true. In fact, the same type of reservoirs that make good candidates
for horizontal wells (i.e., naturally fissured formations) are also likely to exhibit
horizontal permeability anisotropy. In such cases, we may identify two horizontal
permeabilities: kHmax and kHmin.

For a vertical well, this is a moot point, because in cylindrical flow, the effective

permeability ( ) is simply . For a horizontal well, however, the


well trajectory with respect to the permeability direction is critical. This is particularly
true for the case of lopsided permeability anisotropies (see Warpinski, 1991).
Therefore, we need an indication of the preferred permeability direction before drilling
a horizontal well. We also need to know the principal stress directions and the stress
anisotropy. We can obtain these measurements by first drilling a vertical pilot hole.

A hydraulic fracture follows the path of least resistance. In other words, its direction is
normal to the minimum horizontal stress ( ). This orientation is also likely to be
normal to the minimum permeability (Deimbacher et al., 1992). If the horizontal well
is drilled normal to the hydraulic fracture direction and, therefore, normal to the
maximum horizontal permeability, it would have a higher production rate. In fact, in
cases of large permeability anisotropy, a properly directed horizontal well will
outperform either a fractured vertical well or a longitudinally fractured horizontal well.

815
Figure 1 from Deimbacher et al.

Figure 1

(1992) shows a simulated example of the production from a fractured vertical well, a
longitudinally fractured horizontal well and an unfractured horizontal well drilled
normal to the maximum horizontal permeability (kHmax = 10 kHmin). The latter is by far
the most attractive option.

Horizontal Wells vs. Hydraulically Fractured Vertical Wells

In mature petroleum environments such as North and South America, an estimated


80% or more of all wells are hydraulically fractured (Willard, 1989). Hydraulic
fracturing is a long-established means of completing and stimulating wells in
moderate- to low-permeability reservoirs. Recently, there has been a tendency to
fracture higher permeability formations as well, either to bypass near-wellbore damage
or to control sand production by reducing pressure drawdown.

Reduced pressure drawdown, of course, also represents one of the main benefits of
horizontal wells. This raises the possibility of using horizontal wells in place of fractured

816
vertical wells. To evaluate this possibility, we must compare each alternative based on
expected well performance.

Mukherjee and Economides (1991) developed the following comparison for equal well
performance:

(1.8)

where variables are those used in Equation 1.1.

(1.1)

xf is the fracture half-length in the vertical well, and r'wD is the dimensionless effective
wellbore radius, which is a function of the relative fracture capacity (Prats, 1962):

(1.9)
This relative capacity parameter is related to the fracture conductivity (FCD) by

(1.10)
Figure 1 relates the relative capacity parameter and the dimensionless well-bore radius.

817
Figure 1

Using this relationship and Equation 1.8, we can determine the required section length for a
horizontal well to perform at the same level as a hydraulically fractured vertical well having a
fracture half-length xf and a dimensionless effective wellbore radius r'wD.

Figure 2 ( Brown and Economides, 1992) is such a comparison for five different
reservoir permeabilities, plotting the vertical well fracture half-length and the required
horizontal well length.

818
Figure 2

It also shows the optimum fracture lengths, based on a Net Present Value (NPV)
calculation. For example, for a reservoir permeability of 1 md, the optimum fracture
half-length is 1,800 ft [549 m]; for a horizontal well to produce at the same level, its
section length would have to be 2,800 ft [853 m]. The decision regarding a horizontal
versus a fractured vertical well thus becomes an issue of cost, and of whether we can
attain the optimum fracture length.

On the other hand, Equation 1.8 and Figure 2 assume no formation damage in the
horizontal well. A damage skin effect would require a longer section length (see Brown
and Economides (1992) for a comprehensive treatment). In general, we may conclude
that unfractured horizontal wells are not attractive in reservoirs where hydraulically
fractured vertical wells have traditionally been successful. This implies that in such
environments, horizontal wells themselves need to be fractured.

If a is approximately equal to reH (almost always true) and if reH reV, then Equation
1.8 has a much simpler approximation:

819
(1.11)

Example: Horizontal Well Length vs. Fractured Vertical Well

Suppose that a reservoir has a permeability of 1 md, a thickness of 75 ft, and an Iani
value of 3. Optimized hydraulic fracture design suggests that xf should be 1500 ft and
FCD = 1.8. If reH = 2980 and rw = 0.328 ft, calculate the minimum horizontal well
length required for equal well performance.

Solution:

The relative capacity parameter (), from Equation 1.10, is equal to 0.87. From Figure
1 , r'wD = 0.25.

Therefore, from Equation 1.11 and appropriate substitutions,

(1.12)
and by trial and error, L = 2400 ft [732 m].

Multilateral Wells

Multilateral wells, in which two or more horizontal sections are drilled from a single
vertical wellbore, offer a number of reservoir management benefits. Because
multilateral wells require only one vertical wellbore (as opposed to two or more for
equivalent single wells), location costs, drilling time, casing costs and wellhead costs
are significantly reduced. And on offshore platforms, multilateral wells can free
additional well slots for more extensive field development. Multiple horizontal sections
also increase formation exposure and improve recovery efficiency.

Horizontal sections may be drilled either as new wells or as re-entries, and may have
different configurations. These include the over-and-under, in which one section lies
directly above the other, the 180 opposed, in which two sections lie at opposite
angles to each other, and the Y-well multilateral, in which the sections are drilled "out
of phase" (i.e., the angle differences do not equal 180). These basic configurations
may also be used in combination with one another.

Multilateral wells may be a viable reservoir management option under any of the
following conditions (Gardner, 1994):

The reservoir contains vertical permeability barriers (this condition is a primary application
for the over-and-under well configuration).
The planned horizontal displacement is large.

820
The surface lease boundaries are irregular in shape.
The surface topography (or environmental considerations) preclude building multiple drilling
locations.
An existing wellbore is available for drilling re-entry holes.
Drilling is planned from an offshore platform having a limited number of slots.
The productive zones are laminated and have favorable reservoir characteristics (e.g., the
formation is thick enough and competent enough to allow for multiple hole sections).

Multilateral drilling has proven useful as a tool for enhanced oil recovery, particularly in
the area of steam-assisted gravity drainage. This technique, which is used in some
heavy oil and tar sand reservoirs, involves drilling an over-and-under multilateral well.
Steam is injected into the upper wellbore, while oil and condensed steam are produced
from the lower wellbore.

Disadvantages of multilateral drilling include the added complexity of the well control
process, and the potential difficulties involved in selectively re-entering a particular
horizontal wellbore. Often, these problems are related to the methods used to initiate
the horizontal section; improvements in these methods offer a means of overcoming
such problems (Knopczynski et al., 1996).

Formation Evaluation for Optimal Well Completions


Along with the usual variables that affect vertical well performance, we must
determine the following additional properties for horizontal wells:

vertical permeability;

horizontal permeability anisotropy and direction;

position of oil-gas and oil-water contacts (reflecting the current ability to


position the well away from a problematic contact);

lateral discontinuities;

stress anisotropy and direction (for hydraulic fracturing).

An emerging realization in completion design is that zonal isolation may be as critical for horizontal
wells as it is for vertical wells, either because of heterogeneous pressure depletion (which can be
natural or induced from an offset vertical well) or the need to properly distribute matrix stimulation
fluids. Zonal isolation is particularly important if the horizontal well is a candidate for hydraulic
fracturing. It is important to consider these factors when designing a formation evaluation program.

The horizontal wellbore shown in Figure 1 illustrates some of the critical variables
affecting well performance and completion design, including

pressure depletion, induced here by a vertical well

821
Figure 1

the presence of faults, which may separate regions of differing hydrocarbon


saturations and pressures

the well trajectory, which in this case is normal to the maximum horizontal
permeability and the maximum horizontal stress. If we perform a hydraulic
fracture stimulation, the fractures will be transverse (i.e., at right angles to the
wellbore axis).

(Note: There are instances where the well to be fractured should be drilled at
right angles to the direction shown here.)

the location of the gas-oil and oil-water contacts. Note the position of this well
in the vertical middle of the reservoir. This positioning is important in situations
where gas and/or water coning may be a problem. By knowing the location of
these fluid contacts, the operator can keep the well away from the troublesome
horizons.

Finally, whether the well should be drilled at all depends on the vertical permeability, the formation
thickness and the quality of the reservoir's lateral extent.

822
There are a number of formation evaluation technologies available for determining
reservoir variables, including:

seismic and logging measurements;

geosteering;

well testing;

stress measurements.

Seismic and Logging Measurements

Seismic measurements should be an integral part of horizontal well planning. For the
lengths currently being drilled, we can use modern seismic techniques to determine
lateral barriers and, consequently, the reservoir extent and the shape of the presumed
drainage. Surface 3D seismics can detect general trends and obvious discontinuities,
while borehole seismics from a vertical pilot hole help us focus on the area of
immediate interest.

The most obvious difference in logging a horizontal well as opposed to a vertical well is
the area of investigation ( Figure 2 ).

823
Figure 2

In a horizontal well drilled through a relatively thin formation, this area may traverse
overlying or underlying layers, allowing us to determine the distances to other
lithologies and, more importantly, to gas-oil or water-oil contacts. Logging along the
horizontal well also helps us detect the lateral dipping of these vertical boundaries or
their lateral extent.

We would obtain different information from a vertical well, the most important being
the reservoir thickness, the degree of lamination (which has a significant bearing on
vertical permeability) and the fluid contacts. Obviously, we should not drill a horizontal
well without first logging a pilot vertical hole.

Correlating and integrating 3D-seismic measurements with logs from vertical wells
provides valuable information on the proposed horizontal well trajectory and on its
suggested length. Such correlations may lead us to deliberately restrict a well's length.

Of particular interest in horizontal wells are tools that can detect fractures and faults
intersecting the wellbore. Figure 3 (Wellbore image of fault without associated drag)
shows a fault that was detected using Schlumberger's Formation Microimager
(FMI)&trade;.

824
Figure 3

Establishing the presence of fractures, and determining the manner with which they
strike the well can provide indications of the horizontal-to-horizontal permeability
anisotropy.

Geosteering

The term geosteering has been used to describe all real-time measurements that allow
the steering of the well trajectory within the target layer.

For example, Measurement-While-Drilling (M.W.D.) tools provide resistivity and


neutron density measurements, which indicate lithology, porosity and hydrocarbon
saturation. While extending a horizontal well, the operator can detect a dipping
reservoir or a dipping fluid contact, and adjust the well trajectory to become parallel to
the formation or fluid contact boundaries. Porosity and formation lithology can also
provide information on the quality of the reservoir and therefore help us decide the
optimum horizontal well length.

A highly successful technique for geosteering is paleostratigraphy, which involves the


micro-paleostratigraphic analysis of cuttings (Andersen et al., 1990). By examining

825
fossils within the cuttings, an expert micropaleontoligist can determine the depositional
environment and the well's relative position within the formation. This makes it
possible to establish the appropriate well course. The technique is particularly useful
where tectonic movements may have caused dipping or distortion of the layer.

Geosteering is generally an excellent means of establishing the appropriate well


direction and section length, and of using the well itself as a formation evaluation tool.
In horizontal wells, geosteering can greatly enhance reservoir delineation and
discovery of lateral heterogeneities and discontinuities. Such a deliberate use, although
rare at this time, has obvious applications.

Well Testing

Pressure transient testing has developed into a sophisticated, computer-aided means


of detecting and estimating reservoir geometries and reservoir parameters. Modern
well testing uses pressure, pressure derivative and convolution derivative plots to
diagnose the controlling flow regime and, through subsequent analysis of the models,
to determine reservoir or well variables.

Figure 4 from Eblig-Economides (1992) depicts some of the most common regimes
that are currently detectable from distinct patterns in the test data from a vertical well
(again, drilling and testing a vertical pilot hole should be a part of horizontal well
planning).

826
Figure 4

For example, testing a well that barely penetrates a formation causes the appearance
of pseudo-spherical flow ( Figure 4 (d)). Analyzing this flow yields a value for vertical
permeability. Drilling, completing and testing through the entire target interval causes
the appearance of full radial flow ( Figure 4 (g)) and allows us to calculate the
horizontal permeability. The vertical-to-horizontal permeability anisotropy is of
fundamental importance in deciding to drill a horizontal well, especially in thicker
formations.

Testing horizontal wells presents certain difficulties, but it may also provide otherwise
inaccessible information. Shah et al. (1990) provide an excellent analysis of the issues
involved in such tests.

One of the most obvious problems is the unavoidable presence of a large wellbore
storage effect. Even with a downhole shut-in, the length of the horizontal well would
impose wellbore storage effects which may mask early or even late-time flow regimes.

Goode and Thambynayagam (1987), Shah et al. (1990) and Domzalski and Yver
(1992) have identified the successive flow regimes likely to appear in a horizontal well
test after wellbore storage effects disappear ( Figure 5 , (a) early time radial flow; (b)
intermediate-time linear flow; (c) pseudo-radial flow):

Radial flow in the y-z plane should appear first.

827
Figure 5

Analyzing this flow regime should lead to an estimate of the controlling permeability
component (i.e., the average permeability in the vertical plane, perpendicular to the well).

This permeability is .

Shortly afterwards (if the top and bottom boundaries are no-flow) linear flow
into the well ensues, and we can estimate the ky permeability, normal to the
well trajectory.

Finally, pseudo-radial reservoir flow may emerge. Its controlling permeability

would be .

If all three of these regimes emerge (keep in mind that this will not happen if wellbore storage
effects mask the early-time data, or if constant pressure vertical boundaries exist), then we can
calculate all three permeabilities (kx, ky and kz). Such information is invaluable not only in
forecasting well performance, but also in determining the optimum well trajectory with respect to the
horizontal permeability direction.

Stresses and Stress Measurements

828
Every reservoir is subject to a stress field, which we can break down into three
principal stresses:

vertical stress ( );

minimum horizontal stress ( );

maximum horizontal stress ( ).

For a vertical well, there are no problems associated with a non-vanishing stress component. Nor is
there major concern with permeability variations that may be induced by stress anisotropy.

For example, a hydraulically fractured vertical well intersects the fracture plane
longitudinally. Theoretically, the well trajectory is a line on the plane formed by the
hydraulic fracture. The fracture itself is normal to the minimum stress direction, which
corresponds to the path of least resistance during fracture propagation. This path also
implies that the minimum permeability is normal to the fracture plane (Warpinski,
1991 and Deimbacher et al., 1992).

Highly deviated or horizontal wells that are not aligned with one of the principal
horizontal stresses are likely to experience shear stress components. These
components may affect wellbore stability, and especially in the case of hydraulic
fracturing may introduce near wellbore fracture geometry problems.

For an unfractured horizontal well, the principal horizontal stress direction has a large
bearing on the principal horizontal permeability direction. The drilling of the horizontal
well, its direction and (if it will be fractured) the type of fractures it will accept must
consider the stress direction. Obviously, these considerations must precede the drilling
of the horizontal well. An appropriate means of obtaining these data is to take
measurements from a vertical pilot hole.

There are several techniques for determining stress magnitude and directions:

A fracture injection test, as introduced by Nolte (1986), is the best technique for estimating
the closure pressure which, for a single layer, should be very near the minimum stress. (In
almost all cases, at the depth of petroleum interest, this corresponds to the minimum
horizontal stress.)

We can use downhole measurements to establish horizontal stress directions.


For example, if we use a four-arm caliper log to detect wellbore eccentricity,
knowing that stresses are compressive in nature, we may conclude that the
smaller axis of the ellipse should be normal to the minimum stress (i.e., a
hydraulic fracture will be along the minimum axis of the wellbore ellipse).

Laboratory analysis of directionally obtained cores is usually quite accurate and


very useful. Techniques include:
measurements of strain vs the disappearance or appearance of fissures,
such as the differential strain curve analysis (DSCA)

829
anelastic strain recovery (ASR)

an acoustic technique, the differential wave velocity analysis


(DWVA).

Roegiers (1989) presents these techniques, describes the details of their application and lists their
advantages and disadvantages.

830
HORIZONTAL WELLS COMPLETION

Horizontal Well Completions


We can divide horizontal wells into three general categories, based on their curvature
from vertical to horizontal:

short radius

medium radius

long radius

These categories are starting points for designing the well completion.

Short radius wells have curvature radii of less than 50 ft [15.2 m], and as low as 30 ft
[9.1 m]. Their buildup angles are consequently very large as much as 2 degrees per
foot [0.115 rad/m]. Horizontal sections are relatively small (typically 75 - 150 ft, or 23
- 46 m). With current technology, it is not possible to run casing or measurement-
while-drilling (MWD) tools in these sections. Hole diameters are limited to a maximum
of about 6 1/4 inches [158.75 mm].

Medium radius wells have curvature radii ranging from 300 to 700 ft [approx. 90 to
210 m], and buildup angles of between 8 and 20 degrees/100 ft [0.46 - 1.15 rad/100
m]. They require specialized, articulated drilling equipment, and commonly employ
compressive service pipe. These wells can be logged and cased. Hole diameters are
limited to approximately 12 3/4 inches [323.85 mm].

Long radius wells use standard drilling equipment to attain build angles of 3 to 5
degrees per 100 ft [0.17 - 0.29 rad/100 m]. Well diameters are of the same magnitude
as those in vertical wells. This configuration is becoming commonplace, with lengths of
3,500 ft [914 m] now considered routine,and sections approaching 20,00ft [~4100 m]
being reported as of 1996.

The decision to drill a short, medium or long radius well depends on a number of
factors, such as:

Geological discontinuities and well spacing: Although long radius and/or extended reach
wells make for spectacular drilling records, short lateral sections may be more suitable for
cases of severe reservoir heterogeneities along the projected well course.

Reservoir coverage: Short radius wells are often associated with multiple
completions emanating from the same vertical well. They are useful for
producing unswept zones in depleted reservoirs, where drilling an infill vertical
well may be uneconomical. The short radius configuration can also be applied in
enhanced oil recovery (e.g., steam injection).

Formation evaluation requirements: Long radius wells are useful for


measuring reservoir properties over extended intervals.

831
Cost: We must consider costs within the context of the overall reservoir
exploitation strategy.

Stimulation and completion needs, more often than not, point towards long-radius wells. Of the
three configurations, we should therefore consider long radius first (keeping in mind that we might
want to limit the length to better manage the well). We should consider short radius second, for
multiple horizontal completions and, in enhanced recovery applications, for injection/production
configurations. Medium radius wells represent an intermediate option, and are relatively less
common.

Hole diameter is an important consideration in selecting drilling and production


equipment, but it has relatively little impact on horizontal well performance. We can
illustrate this using the data from a previous example where we calculated a
theoretical production rate of 2,490 STB/D [396 m3/D] for a hole diameter of 7 7/8 in
[200 mm]. If we had made this calculation using a diameter of 9 3/4 in [248 mm], the
production rate would have been 2,527 STB/D [401 m3/D]. Or, if we had used a 5-in
[127-mm] diameter, the rate would be 2,417 STB/D [384 m3/D].

Hole size also has a relatively small effect on the pressure drop across the horizontal
section, as we see in Figure 1 (Lackner, 1992).

Figure 1

832
The standard tubing gradient curves at the top of this illustration show depth versus
bottomhole pressure, with the gas-liquid ratio (GLR) as the parameter. The bottom
plot shows the pressure drops in a 6,000 ft [1829 m]+ horizontal well versus the
estimated flowing pressures at the bottom of the vertical section. (Again, GLR is the
parameter.) The production rate is 10,000 STB/D [1590 m3/D].

It is easy to see that the pressure drops in the horizontal section are small and, as one
might expect, are smaller for higher pressures at the bottom of the vertical section.
The fluid is more liquid, and therefore the friction pressure drop which is the only
component of the pressure gradient is smaller.

Vertical wells have a limiting GLR above which the pressure gradient in the well
increases, marking the transition from hydrostatic to friction pressure domination. In a
horizontal section, there is no such division. As the GLR increases, so does the
pressure gradient, since all of the pressure loss is due to friction.

Note that Figure 1 (Vertical, connected to horizontal section gradient curves) is an


approximation, which assumes a sharp transition between vertical and horizontal
sections. We could incorporate the angle of curvature with a more rigorous calculation.
However, our general conclusions would not change.

Pressure Drop in a Horizontal Well

Assume a well depth of 9,000 feet, a GLR of 1,000 SCF/STB and a wellhead pressure
of 100 psi. From the top portion of Figure 1 , we see that the flowing bottomhole
pressure is 1,350 psi. Starting from this point in the bottom plot, the corresponding p
in horizontal section is 115 psi. Such a pressure drop is relatively small, unless the well
is intended to be run at reservoir drawdowns of this magnitude (note that Figure 1 is
for a flow rate of 10,000 STB/D).

The relationship between p and length is largely linear. Therefore, if we reduced the
length of the horizontal section to 2,000 ft,. the corresponding p would be
approximately 38 psi.

While well diameter is not a key factor in choosing a completion design, the type of
well completion does depend greatly on such variables as:

expected well service;

reservoir lithology;

lateral heterogeneities or differential pressure depletion.

Crouse (1991) has identified a number of possible uses for horizontal wells, each of which may
require different completion strategies. Figure 2 and Figure 3 illustrate some of these applications.

833
Figure 2

834
Figure 3

As an exploration or formation evaluation tool, a horizontal well (always preceded by a


vertical pilot hole) may be completed openhole. For draining multiple zones, a well
may contain multiple horizontal completions. The same holds true for the creative
possibility of using a single vertical well with two horizontal sections kicked off, where
one section is an injector and the other is a producer. Considerations like these form
the basis for understanding the features, benefits and limitations of various completion
types.

Types of Well Completions

Figure 1 illustrates the most common completion designs for horizontal wells:

Open hole (a);

835
Figure 1

Slotted, perforated or pre-packed liner in open hole (b);

Liner with External Casing Packers (ECPs) (c);

Cased, cemented and perforated (d).

In addition to these options, we may install pre-packed liners to alleviate sand production. We may
also employ combinations of these completion methods, such as using selective cementing to
isolate a gas cap or other overlying zone.

Open Hole

The open hole completion design was the one most commonly used in the early years
of horizontal drilling. It is still in wide use today, although there has been a gradual
shift to other options.

Of the four main completion types, the open hole configuration is the simplest and
cheapest to execute. In general, the vertical section is cased and cemented, and the
horizontal is left open to the formation. Frequently, a section of the horizontal well
nearest to the vertical is cemented off bottom (Reiley et al., 1987). This technique
involves installing a polymer plug (cross-linked when set; broken later with chemical
breakers) for up to 200 feet [60.96 m] into the horizontal section, filling the well

836
section in front of the plug with cement, and subsequently drilling it out. The ostensible
purpose is to isolate overlying layers or a gas cap.

The usual formation requirements for vertical open hole completions (i.e., competent
rock, no sand production) also apply to horizontal wells. In addition, we need to
understand that horizontal wells rarely traverse truly homogeneous formations one
reason for drilling a horizontal well in the first place is to cut across reservoir
heterogeneities.

For example, Italy's Rospo Mare field in the Adriatic Sea, where the modern era of
horizontal wells started, produces from a vugular formation. In this field, a well's success
depends upon its penetrating these vugs, which are separated by largely impermeable,
zero-porosity rock.

The Austin Chalk formation in Texas is another center of horizontal well


activity, where success likewise depends on wells intersecting massive natural
fracture networks.

In such environments, an open hole completion may not be an attractive option. Its inability to
provide zonal isolation may lead to unwanted comingling of reservoir fluids and, in the case of
differential pressure depletion, actual loss of production from one zone into another. Neither is it
possible to perform selective matrix stimulation treatments.

Slotted, Perforated or Prepacked Liner in Open Hole

In homogeneous, relatively less competent formations, we might run a slotted or pre-


perforated liner in open hole. Where sand production is a problem, we may instead use
a pre-packed liner. All of these designs have the same limitation that characterizes
open hole completions: lack of zonal isolation. An additional problem with installing
slotted liners is the inability to perform hydraulic fracture treatments. Unless we are
targeting homogeneous formations with relatively short horizontal wells, and
stimulation is unnecessary, neither open holes nor slotted liners are likely to be
attractive completion options.

Of course, in all cases, we need to balance the costs of completion against the costs of
potential subsequent problems. If the problems of zonal isolation and stimulation are
severe, then we should use one of the two remaining techniques, which afford
selective completion capabilities: slotted liners with ECPs, or cased and cemented
wells. (Note: Reiss (1985) and Lessi and Spreux (1988) present some of the first
applications of selective completion technologies and the difficulties involved in
applying them. Spreux et al. (1988) present relatively complete reviews of horizontal
well completion technology, especially as it applies to selective completions.)

Slotted Liners with External Casing Packers (ECPs)

Lessi and Spreux (1988), in publishing what is perhaps the first comprehensive
evaluation of horizontal completion technologies, described the testing of formation
packers under both laboratory and field conditions. They concluded that inflatable
packers cannot provide satisfactory seals, because horizontal wellbores are likely to be
oval-shaped and poorly gauged. As it turns out, however, even short ECPs with
common rubber elements have proven able to deform and take the shape of the hole,

837
and at the same time hold substantial pressure differentials. Figure 2 shows a typical
slotted liner/ECP configuration, and a blank liner installed and bracketed by ECPs to
shut off potentially unwanted influx or production loss.

Figure 2

In a number of wells, ECP-equipped slotted liners have provided adequate selective


completion capabilities, and also have lent themselves to zonal isolation for matrix
stimulation.

Of course, not all ECP applications have been successful. Moreover, they may be
somewhat less flexible for long-term production adjustments than our final completion
option: a cased, cemented and perforated well.

Cased, Cemented and Perforated Wells

There is no question that a cased, cemented and perforated well offers the highest
flexibility for zonal isolation and production options. Unfortunately, this configuration is
more expensive than the other completion designs, and has inherent problems of
execution.

838
But we should never exclude cemented and cased wells based on "up front" costs
alone. We must consider the long-term benefits. Coupled with the concept of partially
perforated wells, this type of completion should be the first one considered. We should
employ other types of completions only if the risks they present are minimal, or if the
incremental benefits from cemented wells are marginal.

Zonal Isolation Techniques


Zonal isolation contrary to frequent popular misconception is as essential for
horizontal wells as it is for vertical wells. The slotted liner/ECP design and the cased
and cemented well can offer selective completion capabilities adequate for overcoming
production-related problems and for performing stimulation treatments.

First, it is not necessary to complete the entire horizontal well length. Perforating only
a fraction (say, 60 percent) of a cemented and cased horizontal section, with
perforated sections evenly distributed, may result in a production rate that is more
than 90 percent of the theoretical, open-hole-equivalent production. We could attain
similar success using slotted liners, equipped with ECPs and with blanked-off sections.
This type of completion is not only conducive to excluding troublesome intervals, but it
is particularly applicable to stimulation.

To accomplish zonal isolation and selective fluid placement for matrix stimulation
treatments, we can use chemical diverters and coiled tubing. For temporary zonal
isolation during hydraulic fracturing, we might perform the following sequence ( Figure
1 , Temporary zonal isolation for hydraulic fracturing in horizontal wells using a sand
plug):

Starting from the far end of the well, perforate the first section.

839
Figure 1

Perform the fracture treatment on this section.

Place a temporary plug before performing the second treatment. This could be
a deliberate, inside-the-well screenout, using a multidensity, multimeshed
mixture of sand, or a cross-linked polymer plug that can be engineered to break
after the treatment. Take care in selecting the polymer, because certain
polymers are acid-sensitive and may decompose during an acid breakdown of
the next set of perforations. Section 4 contains a discussion of the optimal
distance between perforations (i.e., the number of treatments to perform) and
the length of the perforated interval.

Damgaard et al. (1989) and Andersen et al. (1990) have introduced a highly sophisticated zonal
isolation technique especially for horizontal wells. This method employs specially designed
hardware and procedures for perforating, stimulating and isolating the well, with the idea of
reducing the number of trips required for the job. Since stimulation treatment time constitutes a
large part of the total completion cost (especially offshore), the overall result is a significant cost
saving, even though the equipment itself is expensive. Application of this technology requires good
cement jobs. Furthermore, the published literature suggests long-radius horizontal wells as the
preferred candidates.

This technique has been applied in wells completed with 7 in. [177.8 mm] liners in a 8-
1/2 in. [215.9 mm] hole. The completion system described here remains in the well
permanently, and is especially useful for production management.

840
The system consists of three separate assemblies, as shown in Figure 2 (Completion
system for perforating, isolating and stimulating in one trip).

Figure 2

(Damgaard et al., 1989):

A sump packer, set permanently at the deepest point in the well.

A downhole assembly, consisting of:

a seal assembly that connects with the sump packer.


a length of 4-1/2 in. [114.3 mm] tubing (the length is equal to the designed
zone length).
a sliding sleeve that can be opened or closed with a coiled tubing conveyed
manipulation string.
a retrievable, hydraulically-set isolation packer.

841
The same downhole assembly is repeated as many times as the desired number of zones in the
horizontal well. The seal of each new downhole assembly simply connects with the isolation packer
of the previous one.

A service assembly for perforating and stimulating each zone, which consists of:

a tubing-conveyed perforating gun (TCP), which can retract after firing.


a circulation port, which allows flow of fluids around the service assembly and into the
annular space of the casing.
a length of 2-3/8 in. [60.325 mm] tubing.
a retrievable packer.
The complete procedure of perforating, stimulating and isolating takes place in a single trip. The
operator:
positions the service assembly (which carries a downhole assembly) with the TCP guns
opposite the perforation interval.

sets the packer.

fires and then retracts the guns.

performs the stimulation treatment.

Following the treatment, the downhole assembly is set in place (sealed in the previous assembly's
packer, with its own packer set), and the service assembly's packer is released and retracted for the
next zonal completion.

Casing Design
The loads imposed on casing in horizontal wells are similar to those encountered in
vertical wells. The magnitude of these loads, however, may vary considerably at
different points in the well. We can identify three distinct regions of loading:

the horizontal, or reach section;

the angle build interval;

the vertical section.

Of course, we can break down these areas further (especially in the angle build interval) for the
purpose of calculating the load (Greenip, 1989).

Casing design criteria in a vertical well typically include:

tension load;

collapse resistance to external hydrostatic pressure;

burst pressure (an important consideration for fracturing jobs or for gas
wells).

842
We may apply the same considerations to horizontal wells, keeping in mind that the vertical stress
(which is normal to the horizontal section) is often much larger than the two horizontal stresses
(which are normal to a vertical well).

A significant difference between the highly deviated and the vertical well sections is the
bending load on the angle build interval. In certain cases, bending load may be the
major design criterion (Greenip, 1989).

The bending stress (b) in psi is equal to

(2.1)
where
dc is the outside diameter in inches

is the build angle in degrees.

Where b is expressed in kPa (with dc given in millimeters and in radians) 211 becomes 3282.

For example if = 30 degrees and dc = 5 inches, then b = 31,650 psi (157,000 lb).

This bending stress is additional to the axial load, and for certain cases it can account for 60 to 80 percent of the total load.

In the horizontal section, the pipe may be subjected to considerably smaller loads. To reach this part of the well, however, it must
pass through the angle build section. We must take this into consideration when designing the horizontal casing.

Cementing
While the same reasons for cementing casing in vertical wells are also valid for
horizontal wells (i.e., zonal isolation, casing support, corrosion control), a horizontal
well's geometry gives rise to two problems:

1. Poor drilling mud displacement, resulting from gravity/flow incompatibility with the cement
slurry, and aggravated by improper pipe centralization.

2. Free water accumulation at the upper side of the well as the slurry sets,
resulting in an inadequate seal.

Mud Displacement and Free Water Accumulation

Efficient drilling mud displacement is essential to forming a good cement seal. This is
true for any well, whether it is vertical or horizontal. Undisplaced mud or drilled
cuttings can result in channeling and zonal communication. In vertical wells, gas
migration is a severe and prevalent problem. While gas migration may be less of an
issue in horizontal wells, we still have to worry about heterogeneous pressures causing
intrazonal flow and loss of production.

We also have to consider gravity and flow direction. In a vertical well, the mud returns
flow up the hole, opposite the direction of gravity. Thus, if the flow rate is high

843
enough, it may slow or even prevent particles from settling. In a horizontal hole
section, however, fluid flows at right angles to gravity, and so flow rate does not
prevent particle settling. We need to take this into account when we design the mud
rheology. Also, we should plan for frequent cleanout trips during drilling to remove
mud solids and cuttings, and to reduce the risk of stuck pipe (Griffin, 1992).

Gravity may also work with the horizontal well geometry to produce an eccentric liner-
to-well configuration. This is likely to result in a buildup of drilled cuttings in the lower
part of the well and, after cementing, a free water accumulation at the top of the well (
Figure 1 , Gravity-related mud displacement problems in horizontal wells.

Figure 1

Drilled cuttings settle in the bottom and liner non-centralization causes a much smaller
cement slurry flow rate in the bottom of the well. Free water after cement setting
results in a channel at the top of the hole).

The smaller liner-to-hole clearance in the lower hole section leads to a considerably
reduced flow rate (and far less efficient mud displacement) relative to the upper hole
section.

We refer to this clearance as the standoff, where

844
% standoff = (2.2)
where
w is the minimum clearance, and

rw and rc are the radii of the well and casing, respectively, as shown in
Figure 1 .

Crook et al. (1985) studied the issue of mud displacement in highly deviated wells. Of particular
importance is drilled cuttings transport in wells with deviations from 50 degrees to 90 degrees [0.87-
1.57 rad].

In these situations, we may improve displacement efficiency by:

keeping the thixotropic behavior of the fluid at a minimum;

decreasing the mud viscosity;

increasing the mud density;

maintaining turbulent flow (this will be adversely affected by standoff


problems).

To prevent particles from settling when the mud is static (i.e., before pumping the cement slurry), a
minimum threshold point is required, which increases with angle of deviation. The mud's rheological
properties are therefore critical. Also, pumping should begin as soon as possible.

While the preceding guidelines are general, the mud design that facilitates cuttings
transport and allows for good mud displacement is specific to each well, and must
incorporate the appropriate job parameters. For critical wells, extensive laboratory
testing is recommended.

Griffin (1992) has set forth guidelines for mud circulation before cementing:

Circulation to break the gel strength should continue until at least 95 percent of the mud is
moving.

To avoid settling of solids due to the broken gel strength, avoid excessive
circulation two hole volumes should be adequate.

Pipe movement helps to break the gel strength an improve mud displacement.
Both rotation and reciprocation have been used in horizontal wells, and either is
recommended. Rotation should be 10 to 20 RPM and reciprocation should be in
10 - 20 ft [ 3 - 6 m] strokes with 1 to 2 strokes every 1 to 2 minutes.

Centralization

Field and experimental evidence suggests that mud is easy to remove in well-
centralized casings, where the standoff is more than 67 percent. Poor mud removal

845
occurs at standoffs of less than 30 percent (Parcevaux, 1987). Casing centralization,
however, can be difficult to achieve in highly deviated or horizontal wells because of
the increasing load on the centralizers.

The obvious solution to this problem would be to simply run more centralizers.
Unfortunately, this increases the risk of sticking the pipe in the bent hole section. An
alternative solution is to decrease standoff by reducing the density difference between
cement, spacer and drilling mud, and by employing the lightest possible displacement
fluid (Parcevaux, 1987, Griffin 1992). Parcevaux also suggests that in going from mud
to spacer to cement to displacing fluid, it may be desirable to sequentially reduce the
density. Of course, any density reduction is subject to well control considerations. Note
also that while these decreasing densities might work in the horizontal section, they
may be insufficient for adequate mud displacement in the vertical section.

Liner Sizing

The liner size relative to the hole diameter is important in determining the standoff and
number of centralizers required, and if the number of centralizers is relatively large
in reducing the risk of sticking. Hardman (1986) concludes that the ratio of casing to
well diameter should be much smaller than in conventional wells. He recommends 5-
1/2 in. [139.7 mm] liners in 8-1/2 in. [215.9 mm] holes. Others have suggested that
larger ratios of casing-to-well diameters are possible (Griffin, 1992). However, current
practice frequently uses the 5-1/2 and 8-1/2 in. configuration.

Cement Slurry Design

Potential problems, such as poor mud displacement efficiency and free water
accumulation, determine the desired slurry and set cement characteristics.

To offset casing centralization problems, the cement density must be at least as low as
the drilling mud density. Furthermore, casing eccentricity suggests using a slurry with
as low a yield point as possible, so that it can flow in the lower part of the annulus. The
problem with this approach is that low-yield point slurries can cause free water
breakout and cement particle sedimentation, resulting in a poor seal. Thixotropic
cements can prevent free water formation, but they exhibit high yield stresses
(Parcevaux, 1987).

The only cements that combine low density, low settling and low yield stress are
foamed cements or cements with microspheres. Moreover, the vertical stress on a
horizontal well is usually greater than the two horizontal stresses acting on a vertical
well, and thus requires a cement with relatively higher compressive strength. Foamed
or microsphere cements can provide this compressive strength.

A number of free water inhibitors, which also act as particle dispersants, have been
introduced as cement slurry additives, and are recommended for use on horizontal
wells.

Sand Control

846
In some formations, sand production is a serious problem, affecting surface production
facilities and, in extreme cases, causing wellbore instability. The magnitude of the
pressure drawdown is one of the key variables controlling sand production. Horizontal
wells, by affording lower drawdowns, may be ideal for these reservoirs.

In some formations where vertical wells require sand control measures, drilling
horizontal wells may in itself alleviate the problem. In other cases, however, drilling
horizontal wells may not be enough, and sand control measures may be necessary.
Gravel packing of horizontal wells has been attempted with relatively successful
results.

Long horizontal intervals that need to be gravel packed may generate considerable
friction pressure loss in the particle-laden slurry. This pressure loss reduces the fluid's
horizontal velocity, thus requiring particular attention to the particle settling velocity.

Roodhart (1985) has presented a modification of Stokes' Law for single particle
terminal settling velocity for power-law fluids.

(2.3)
(Equation 2.3 is in SI units).

where:

g = acceleration of gravity, m/s2

= density difference between that of the particle and the carrier fluid, kg/m3

dp = particle diameter, m

0 = zero shear viscosity, Pa-s

n' and K' describe the rheological properties of the fluid.

It is apparent from Equation 2.3 that the settling velocity is a minimum if is small (i.e., using
lower density particles) and if dp is small.

Example: Calculation of Settling Velocity

Assume that:

gr = 2650 kg/m3 [165.4 lb/ft3]

fluid = 1000 kg/m3 [62.4 lb/ft3]

n' = 0.37

847
K' = 4 Pa-sn' [0.0835 lbf/ft2-sn']

0 = 0.2 Pa-s [200 cp]

dp = 2.5 x 10-4 m [8.2 x 10-4 ft]

g = 9.78 m/s2 [32.2 ft/s2].

Calculate the single particle settling velocity.

Solution

From Equation 2.3:

(2.4)

and therefore:

ut = 2.8 x 10-4 + 2.6 x 10-3 ut0.63 (2.5)


From Equation 2.5, ut = 2.95 x 10-4 m/s [1 x 10-3 ft/s]

The results of Example 2.2 illustrate the contrasting factors affecting particle transport
in a horizontal well. Higher viscosity may result in lower settling velocity (if 0 were
2000 cp, then ut would be 3.18 x 10-5 m/s [1 x 10-4 ft/s]).

Higher settling velocity causes the particles to dune, which reduces the clearance,
increasing the velocity above a critical value. This in turn prevents further particle
settling and bridging in the well.

In gravel packing, however, the situation is more complicated. The fluid must flow
normal to the screen, and fast enough to force the particle-laden slurry to pack off
against the formation. Lower-viscosity fluid with reduced particle concentrations may
cause duning at the screen, with carrier fluid leaking off and causing inadequate gravel
transport at the formation interface.

There are several models available that can account for these phenomena. In general,
a model must balance the fluid rheological properties, gravel properties, temperature,
screen size and the length of the interval to be packed.

Milne et al. (1991) have offered the following practical guidelines. Keep in mind,
though, that job design is extremely well-specific, and that critical wells require
sophisticated, customized models.

848
Optimize the slurry by reducing particle concentration, taking into account the increased
leakoff during packing.

Reduce carrier viscosity by taking into account particle transport and settling
velocity.

Minimize the density differential between the particles and carrier fluid.

Reduce screen diameter to lower the friction pressure drop in the annulus and
prevent bridging at the screen.

849
MATRIX STIMULATION OF HORIZONTAL WELLS

Matrix Stimulation Of Horizontal Wells

Matrix stimulation is a standard procedure for removing near-wellbore damage.


Although acids cannot remove all types of damage, the vast majority of matrix
stimulation treatments involve the use of aqueous HCl solutions for carbonates, and
mixtures of HCI-HF for sandstones. These treatments also employ various chemical
and mechanical diversion techniques for placing the stimulation fluids at the desired
sections of the wellbore.

For conventional wells, common industry practice is to perform staged treatments,


using 50 to 150 gallons (U.S.) of stimulation fluid per linear foot of interval [0.621 -
1.86 m3 per meter]. This approach may not be feasible for long horizontal wells,
however, due to the large fluid volumes required. For example, to attain 50 - 150
gal/ft coverage across a 4000 ft [1219 m] horizontal section, we would need 200,000
to 600,000 gallons [757 - 2271 m3] of fluid.

Operational and economic constraints frequently limit treatment volumes. Formation


fracture pressure, for instance, places an upper limit on how fast we can pump the
fluid. Extremely long pumping times (i.e., 100 hours or more, depending on the flow
rate) can make a job prohibitively expensive, especially offshore, and can cause
corrosion problems.

Moreover, in a horizontal well, deeper damage penetrations may occur because of


longer exposure to drilling and completion fluids. An acid wash may leave a "collar" of
damage around the wellbore, resulting in substantial production loss. Thus, there is an
obvious need for optimizing matrix treatments in horizontal wells.

Damage Characterization
Just as hole geometry affects a well's deliverability, it significantly influences the
damage profile and the relative contribution of the skin effect. Consequently, it has a
bearing on how we design and execute the stimulation treatment.

Horizontal Well Damage Profiles

The simple cylindrical model that characterizes damage in vertical wells ( Figure 1 )
does not apply to horizontal wells.

850
Figure 1

In a horizontal well, damage penetration is likely to be deeper near the vertical section
due to its longer exposure to drilling and completion fluids. Also, because friction
pressure losses result in larger reservoir pressure gradients in this part of the well, the
damage profile forms a cone, with the larger cross section at the shallower part of the
horizontal section.

The damage penetration and profile also depend on the preferential flow directions
(i.e., on the different permeabilities in the horizontal and vertical plane). Therefore, in
an anisotropic environment, the cross section of the damage profile is elliptical, with
the larger axis in the direction of the highest permeability. The higher the anisotropy,
the more lopsided the ellipse. Thus, the damage profile along and around a horizontal
well in an anisotropic formation is likely to be a truncated cone with an elliptical cross
section (Frick and Economides, 1991; Figure 1 ).

Simulations of mud filtrate loss during the drilling process support this truncated cone
model. Figure 2 shows the simulated damage profile along a 1000-ft horizontal
trajectory.

851
Figure 2

The simulations were performed for three different values of the anisotropy variable
(Iani = 1, 3, and 5). Figure 3 shows the cross sections of the mud filtrate invasion
profiles.

852
Figure 3

This proposed model reflects horizontal well conditions more realistically than the
cylindrical damage profile. It is still simplistic, however, because it does not account for
heterogeneities along the well trajectory, which may influence the damage distribution.
Bearing this limitation in mind, the truncated elliptical cone geometry allows us to
develop easy-to-handle analytical expressions for pre- and post-treatment skin effects.

More important, the truncated cone image has implications for stimulation treatment
design:

It offers insights into optimizing treatments by suggesting the placement of more


stimulation fluids near the vertical section of the well, where the damage penetration is
deeper.

On the other hand, it indicates that distributing fluids uniformly along the well
would result in wasteful overtreatment in the deeper sections, and insufficient
damage removal elsewhere.

Horizontal Well Skin Effect

The term skin effect describes the additional pressure drop that occurs through a near-
wellbore area of altered permeability. Hawkins (1956) presented the classic
formulation for skin effect in a vertical well, based on the cylindrical damage model
illustrated in Figure 1 .

The variables that determine skin effect are:

damaged zone permeability (ks);

undamaged reservoir permeability (k);

wellbore radius (rw);

damage penetration (rs).

The skin effect for a vertical well (sv) is defined as

(3.1)
For the truncated elliptical cone geometry, taking into account horizontal well damage distribution
and anisotropy, we can develop an analytical expression for an equivalent skin effect analogous to
the Hawkins formulation

853
(3.2)

where

asHmax = the horizontal half-axis of the damage cone near the vertical section (
Figure 1 ).

Iani = (kH/kv)1/2 (3.3)


where
kH and kv are the vertical and horizontal permeabilities.
Frick and Economides (1991) present a detailed derivation of Equation 3.2. Muskat (1937) elegantly
outlines the concept of skin in a vertical well.

Figure 4 illustrates the effects of skin damage,

Figure 4

854
plotting horizontal well skin versus maximum damage penetration for 3 different
permeability impairments (k/ks = 5, 10 and 20), and two different horizontal-to-
vertical permeability anisotropies. The wellbore radius is 0.4 ft.

Example: Estimating Damage Distribution

Analysis of a test in a horizontal well (rw = 0.4 ft) gives a damage skin factor of 18.
The horizontal permeability (kH) is 60 md, and the vertical permeability (kv) is 15 md.
A core test reveals a damage permeability of 3 md.

Estimate the damage profile.

Solution

The anisotropy variable is

Iani = =2
The derivation for Equation 3.2 implies the same anisotropy in the damaged area as in the virgin
reservoir. To determine the ratio k/ks, we must calculate the average permeabilities

k=
for both the damaged area and the virgin reservoir. Assuming the damage permeability from the
core test to be the average value, then

Using Figure 4 , and interpolating between Iani = 1 and Iani = 3, the maximum horizontal damage
penetration (asHmax) is approximately 7.4 ft [2.26 m].

As we see from Figure 4 , the relationship between skin and damage penetration is
sensitive to the permeability impairment. To illustrate this point, doubling the ratio
k/ks would reduce the expected maximum damage penetration near the vertical
section by 66% (asHmax = 2.5 ft [0.762 m]).

Performance of Damaged Wells

The skin effect implies a pressure drop, which we add to the reservoir pressure drops
for vertical and horizontal wells in similar ways.

The steady-state formulation for fluid flow into (or, in the case of injection, from) a
vertical well, including the skin effect (Muskat, 1937, van Everdingen, 1953) is

(3.4)
where
kH = permeability in the horizontal plane (md)

855
h = formation thickness (ft)

pe and pwf = outer boundary and wellbore pressures (psi)

B = formation volume factor (Bbl/STB)

= viscosity (cp)

re = radius of drainage area (ft)

rw = wellbore radius (ft)

Note: Where q is expressed in m3/d (with k in md, pe and pwf in kPa, B in m3/m3, in cp and h, L,
re and rw in m), 141.2 becomes 1867. This applies also to Equation 3.5.

The skin effect in Equation 3.4 describes the total skin, which includes all near-
wellbore features such as perforation skin, partial completion, and so forth. The skin
due to damage and stimulation is just one of these components. Several authors have
addressed this topic, including Pucknell and Clifford (1991).

The expression for horizontal well production rate (developed by Joshi, 1988, and
augmented by Economides et al., 1991) including skin, can be written as

(3.5)

where

a = half-axis of the drainage, ft

L = length of the horizontal section (ft)

Iani = permeability anisotropy variable defined in Equation 3.3

seq = equivalent skin effect defined by Equation 3.2.

Equation 3.5 applies to a horizontal well under steady-state flow conditions forming an elliptical
drainage area. For these conditions, the equivalent skin has to be multiplied by the scaled aspect
ratio (Ianih/L). There are, however, other flow equations for horizontal wells, which account for
different drainage areas, boundary conditions and so forth. Economides and Frick (1992) describe
how to incorporate skin damage into the best-known of these equations.

Comparing the equations for vertical and horizontal flow, it is obvious that their
differences are in the logarithmic terms to which we add the skin effects. The
logarithmic expression in Equation 3.4 usually ranges between 8 and 9, whereas the
sum of the logarithmic groups in Equation 3.5 is in the range of 2 to 3. Therefore,

856
horizontal well skin effects even when multiplied by the scaled aspect ratio may
be more detrimental than they would be in vertical wells.

Example: Damaged Horizontal Well Performance

Using the data of Example 3.1, calculate the productivity index (PI) of a 1000 foot
horizontal well. How much potential production is lost because of damage? At what
skin are the productivity indexes of the horizontal well and a fully stimulated vertical
well in the same reservoir equal?

Use p = 2000 psi, A = 160 acres, B = 1.1 Bbl/STB, = 1 cp and h = 100 ft.

Solution

The drainage radius for A = 160 acres is 1490 ft. The half-axis of the drainage ellipse
(a), from Equation 1.4, is 1532 feet, and the PI of the damaged well (from the
Equation 3.5) is

PIH 6

A fully stimulated vertical well in the same reservoir would have a PIv of 4.7, which is
three fourths of the horizontal PI. Thus, the production improvement by drilling a
horizontal well is rather poor. (A rule of thumb is a successful horizontal well should
have at least 2 to 3 times the productivity index of a vertical well.)

The undamaged PI is 13.8. Thus, the production impairment due to damage is 56


percent. This well would be an attractive candidate for a matrix stimulation treatment.

To calculate the horizontal well skin effect (sH) at which horizontal and vertical
Productivity Indexes are equal, we combine Equations 3.4 and 3.5 into

(3.6)

Introducing the given variables into Equation 3.6

857
which is the skin value at which PIv and PIH are equal.

Figure 3 shows the impact of damage on both a vertical and a horizontal well, with
data taken from Example 3.2. Note that the slope of the horizontal well productivity
index decline is steeper, indicating the relatively larger detrimental impact of skin.
Another conclusion we can draw from this graph is that fully stimulated vertical wells
may outperform damaged horizontal wells at reasonable skin effect values (in the
previous example, at a skin of 27), and hence damaged horizontal wells ought to be
stimulated properly.

Matrix Stimulation
There is no difference between vertical and horizontal wells with respect to the type of
stimulation fluid used (McLeod, 1984). Formulations depend primarily on the types of
damage being removed (Piot and Leonard, 1989). Fluid volume and placement,
however, are different matters. Long stimulation intervals, together with the limited
acid volumes that can be pumped through coiled tubing (which is recommended),
imply that horizontal wells are more sensitive to the amount and distribution of
injected stimulation fluids. Proper job design is therefore all the more essential.

The major difference between vertical and horizontal wells is the way in which damage
is distributed along the wellbore and, consequently, the manner in which we place
stimulation fluids. Chemical diverting agents, mechanical isolation devices and even
deliberate blanking of well portions are ways we can selectively distribute fluids.

Stimulation Techniques

Coiled tubing is a recommended means of providing additional mechanical isolation


and diversion for horizontal well stimulation treatments. One technique, proposed by
Economides et. al. (1991), involves pumping inert fluids through the annulus formed
between the coiled tubing and the well, while injecting acid through the coiled tubing.
The inert fluids provide back pressure ( Figure 1 , Coiled tubing stimulation technique).

858
Figure 1

In this procedure, we run the coiled tubing to the end of the well, and then gradually
withdraw it while pumping the fluids. We calculate the tubing withdrawal rate to
achieve maximum diversion, based on the injection rate and the average volumetric
coverage. The interface between the acid and inert fluids should move at the same
rate as the coiled tubing, so that it stays close to the injection point at the end of the
string.

This technique is applicable to all types of completions, including open holes, slotted
liners and cemented and cased wells. Zonal isolation requirements are less stringent
for matrix stimulations than for hydraulic fracture treatments. Coupled with anticipated
productivity index increases, this might make a horizontal well matrix stimulation
treatment an attractive alternative to a hydraulically fractured vertical well.

Bullheading the stimulation fluids is not a recommended substitute for using coiled
tubing. It too often results in the acid being spent near the vertical section, which then
accepts even more stimulation fluids. Hence, massive leakoff zones develop. This is
especially true for carbonates.

However, Tambini (1992) does report a successful field application of a technique that
combines coiled tubing and bullheading. This procedure consists of two injection
stages:

First, the fluids are diverted using coiled tubing at low injection rates (1-2 Bbl/min [0.00265
- 0.0053 m3/s]), and withdrawal rates of 10-20 ft/min [0.05 - 0.10 m/s].

Subsequently, the acid is bullheaded at the maximum injection rates (slightly


below fracturing pressures).

859
Throughout the injection, the operator evaluates the skin effect's development in real time by
monitoring injection rates and pressures. The principle is fairly simple: if the skin decreases, the
injectivity of the well increases and the formation accepts more acid. Thus, it is possible to calculate
the skin evolution over time.

Figure 2 shows the rate and pressure development during a matrix stimulation job on
a 564 foot [172 m] open hole section.

Figure 2

Several stages of 28 percent HCl with two different viscosities (1 and 5 cp) were
pumped into a microfissured limestone reservoir. The total acid volume was 17,000
gallons [64.35 m3].

The advantage of this technique is that it combines optimal diversion (using coiled
tubing) and reduced treatment times (due to the high rates attainable through
bullheading). Moreover, the high injection rates provide an additional diversion effect,
as has been used successfully in vertical well treatments (Paccaloni, 1992).

Partial Stimulation

860
Considering the large fluid volumes sometimes required for full damage removal, it
may be worth considering the alternative of partial stimulation using less fluid than
would be needed to remove all damage. Although partial stimulation leaves a "collar"
of damage in the reservoir, it can be a viable treatment option.

To evaluate the effectiveness of a partial stimulation treatment, we must calculate its


post-treatment skin effects. Figure 1 illustrates the geometric model that applies to the
problem.

Figure 1

Three regions of different permeabilities surround the wellbore:

a stimulated area, with radius ri and permeability ki

a damaged area (rs, ks)

the undamaged reservoir (re, k).

861
Assuming an isotropic environment and uniformly distributed damage (i.e., vertical well conditions),
we may then define the post-treatment skin effect as follows (Frick and Economides 1991):

(3.7)

Figure 2 (Partial stimulation of a vertical well) is a graphical depiction of Equation 3.7,


showing the skin improvement in a vertical well (h =100 ft [30.48 m]) versus the
penetration of injection fluids.

Figure 2

The three curves represent three different permeability impairments (k/ks = 5, 10 and
20). The treatment of the most severe permeability reduction (k/ks = 20) is the most
beneficial. (The y-axis on the right indicates the stimulated bulk volume, which is equal
to Vbulk = (ri2-rw2)h).

We can convert Equation 3.7 to the horizontal damage distribution, including


anisotropy. However, different mechanisms of acid attack may result in different
configurations of the stimulated zone, and thus different skin effects. This is likely the
difference between sandstones and carbonate reservoirs.

862
Stimulation of Sandstone Reservoirs

In sandstones, stimulation fluids attack the particles plugging the pore space around
the wellbore. Hence, the original an isotropy does not change, and Equation 3.7
transforms to

(3.8)

where

asH,max and aiH,max, respectively, are the maximum horizontal half-axes of damage and
stimulation ellipses in ft.

Analogous to the vertical well, we can write the stimulated bulk volume as

(3.9)

Figure 1 (Partial stimulation of a horizontal well in sandstone)

863
Figure 1

and Figure 2 (Partial improvement by partial stimulation) illustrate the results of a


partial matrix treatment in a sandstone formation. These plots are based on the well
and reservoir data of Table 1, below.

864
Figure 2

Figure 1 shows skin improvement and the volumetric percentage of damage removed
as a function of maximum acid penetration near the vertical section. The stimulated
permeability (ki) is assumed equal to the virgin reservoir permeability (k). The
permeability impairment (k/ks) is 10, and the initial damage penetration near the
vertical section (asH,max) is 4 feet.

Figure 2 shows the production improvement (q = qstim - qdamaged) versus the


injected acid volume, and demonstrates that even moderate acid volumes (like those
used in a wellbore clean-up) can lead to significant productivity improvements.

On the other hand, we can also see from Figures 1 and 2 that only a full matrix
treatment delivers maximum productivity (or injectivity). The treatment design thus
becomes an issue of optimization.

L = 1000 ft kH = 60 md

865
rw = 0.4 ft Iani = 2

A = 160 acres h = 100 ft

pe = 5000 psi = 1 cp

pwf = 3000 psi B = 1.1 Bbl/STB

= 20%

Table 1: Well and reservoir data for Figure 1 .

Example: Partial Stimulation of a Sandstone Well

Suppose that you are contemplating a partial stimulation of a sandstone formation.


The damage penetration near the vertical section (asH,max) is an estimated 2 feet and
k/ks = 10. What productivity increase would result if the acid penetration were 50
percent and 80 percent, respectively, of the damage penetration?

Solution

For a sandstone reservoir, recovering the virgin permeability in the stimulated area is a
good target and a reasonable assumption. In this case, k = ki, and Equation 3.8
reduces to

(3.10)

Acid penetrations of 50 percent and 80 percent of the damage penetrations are then
equal to 1 and 1.6 feet, respectively. We then calculate the partial stimulation skin (for
aiH,max = 1 ft) as:

For the 1.6 ft acid penetration, the post-treatment skin would be 1.75.

The Acidizing Process

In sandstone acidizing, the sharpness of the reaction front increases with the
concentration of fast-reacting minerals (e.g., clays) in the pore matrix. If we assume a
sharp reaction front, the volumetric coverage of acid (in gal/ft) required to penetrate a
certain distance into the formation at a dimensionless distance x/L from the vertical
section is equal to (Schechter, 1992)

866
(3.11)

where NAC is the acid capacity number, defined as

NAC = (3.12)

where

Cr = concentration of the reactant in the liquid phase in mole/liter

= stoichiometric coefficient for the mineral (in this case, we assume that only one
mineral is reacting with the acid)

= porosity

Cm = concentration of the mineral, mole/liter

In SI metric units, the constant 23.5 becomes .

The second dimensionless group in Equation 3.11 is the Damkohler number (NDa),

which is the ratio of reaction rate to convection rate. The quantity , which is the
front's dimensionless position normal to the wellbore at distance x/L, is given for radial
flow as follows (from Schechter, 1992, and transformed to horizontal well conditions):

(3.13)

where

x = distance from the vertical section

aiH,max = maximum fluid penetration near the vertical section.

If we divide the horizontal section into several intervals, with the midpoint of each
interval (j) being at xj, we can calculate the rate of coiled tubing withdrawal at position
xj/L as

(3.14)

867
where

vCTj is in ft/min, qinj is in BPM and Vxj/L is in gal/ft (In SI metric unit, the constant 42
reduces to 1).

Example : Rate of Coiled Tubing Withdrawal

Using the data of Example 3.3, and dividing the horizontal well into ten sequentially
stimulated sections, calculate the amount of acid to be injected into each section to
remove the damage completely. What rate of coiled tubing withdrawal is required for
optimum distribution of stimulation fluids?

Use asH,max = 2 ft, NAC = 0.01, NDa = 0.02, qinj = 1 BPM and = 20 percent.

Solution

The length of each sections is

= 100 ft

The midpoint of each interval from the vertical section is 50, 150, 250 ft, etc. For a
given maximum penetration, with aiH,max being 1, 1.6 and 2 feet, respectively, we
calculate the dimensionless position of the front using Equation 3.13. For example, in
the first interval (x = 50 ft) for aiH,max = 1.6, the dimensionless position of the front is

We can then obtain the volumeric coverage from Equation 3.11. For the first interval, it
would be

= 527.5 gal/ft

Finally, we use Equation 3.14 to compute the coiled tubing withdrawal velocity for each
interval:

= 0.08 ft/min

868
Table 2, below, lists the results of the calculations for this example. The total volumes
of acid injected for the partial stimulation are 77,000 gal (aiH,max = 1 ft) and 238,000
gal (aiH,max = 1.6 ft). Complete damage removal (aiH,max = 2 ft) would require
387,000 gallons of acid. Figure 3 shows the rates of coiled tubing withdrawal in each
section for a partial damage removal of aiH,max equal to 1.6 ft.

Figure 3

aiH,max = 1 ft

x, Coverage, Vinj, vCT,

ft gal/ft gal. ft/min

50 155.03 15503 0.27

150 135.91 13591 0.31

250 117.47 11747 0.36

869
350 99.70 9970 0.42

450 82.59 8259 0.51

550 66.14 6614 0.63

650 50.34 5033 0.83

750 35.17 3516 1.19

850 20.63 2063 2.03

950 6.72 672 6.25

(a)

aiH,max = 1.6
ft

x, Coverage, Vinj, vCT,

ft gal/ft gal. ft/min

50 527.53 52753 0.08

150 447.06 44706 0.09

250 373.08 37308 0.11

350 305.29 30529 0.13

450 243.43 24343 0.17

550 187.28 18728 0.22

650 136.62 13662 0.30

750 91.26 9126 0.46

850 51.03 5103 0.82

950 15.79 1579 2.65

(b)

aiH,max = 2 ft

x, Coverage, Vinj, vCT,

ft gal/ft gal. ft/min

50 899.64 89964 0.05

870
150 749.21 74921 0.06

250 614.14 61414 0.07

350 493.32 49332 0.09

450 385.81 38581 0.11

550 290.76 29076 0.14

650 207.45 20745 0.20

750 135.25 13525 0.31

850 73.61 7361.1 0.57

950 22.09 2208.9 1.90

(c)

Stimulation of Carbonate Reservoirs

Carbonate formations differ fundamentally from sandstones with respect to the


acidizing process and its objectives:

While sandstone treatments remove plugging materials from the pore space, carbonate
acidizing primarily involves a reaction between the acid and the rock itself.
Whereas in sandstones, the goal is to restore reservoir permeability, carbonate acidizing is
designed to create secondary permeability, which is characterized by highly conductive
paths called wormholes.

Carbonate stimulation treatments may lead to extremely high permeabilities in the stimulated area,
resulting in negative post-treatment skin effects.

Because carbonate acidizing is based on creating secondary permeability, proper


diversion of stimulation fluids is even more important than it is for sandstones
(Tambini, 1992). Bullheading the acid is likely to generate large leakoff zones, which
take all of the stimulation fluids. This leakoff potential, combined with the long
intervals to be stimulated, suggests the need to divide the well into several
sequentially treated sections. This allows for optimum diversion. The treatment's
effectiveness thus depends strongly on the quality of zonal isolation, which would be
most effective with a cased and cemented completion.

Daccord et al. (1989) proposed the following equation for the skin effect resulting from
a partial stimulation treatment in a vertical well:

(3.15)

For horizontal well conditions (Kurmayr et al., 1992),

871
(3.16)

where ri,max= maximum radius of the stimulated area in feet.

The main difference with the sandstone skin model is that we assume the stimulated
area described in Equation 3.15 to be radial, irrespective of anisotropy. This is because
carbonate stimulation relies more on reaction kinetics than on fluid flow.

We can calculate the volumetric coverage of acid injected into section j as follows
(after Kurmayr et al., 1992):

(3.17)

where

Vacid,j is in gal/ft,

NAC = dimensionless acid capacity number defined in Equation 3.12,

b is an experimentally determined constant (104 ft )

df = fractal dimension (normally 1.65)

ri,j = radius of acid penetration for segment j (ft)

Npe = Peclet number describing the ratio of convection and diffusion velocity. For radial
flow it is defined as

(3.18)

where

qinj is the injection rate, D is the diffusivity constant and Linj is the length of the
horizontal section.

(In SI metric units, the constant 7.48 reduces to 1.)

872
Applying the concept of a section-by-section treatment, the total volume of acid
injected is simply the sum of Vacid,j times the length of each interval over the number of
stimulated sections, n.

We can then compute the rate of coiled tubing withdrawal for each section j:

(3.19)

where

VCT,j is expressed in ft/min

(In SI metric units, the constant 42 reduces to 1).

These formulations assume that within each segment, the coiled tubing withdrawal
rate is constant. In reality, however this rate changes from segment to segment to
allow more stimulation fluids in the shallower segments.

Optimization
Operators have various options for designing matrix stimulation treatments. Those
already mentioned include coiled tubing applications (perhaps combined later with
bullheading), partial stimulation and sequential stimulation. Another attractive option is
to partially complete the well by deliberately blanking certain intervals.

To select the best design, we must compare our options based on economics. Using
the previously presented equations for forecasting post-treatment well performance,
we can introduce a Net Present Value (NPV) criterion (Economides and Frick, 1992).
For matrix acidizing, the NPV concept involves balancing the discounted incremental
revenue against the treatment cost. The optimal treatment design is the one with the
highest NPV.

We can express the NPV for a matrix treatment as

NPV = (3.20)

where

n = total number of years for the period

i = discount rate

C1 = wellhead value of oil (e.g., $/STB)

873
C2 = value of injection time (e.g., $/hr)

C3 = value of the required stimulation fluids (e.g., $/gal)

Cfix = fixed costs (e.g., $)

= difference in incremental production between the stimulated and the


unstimulated well in the nth year.

Figure 1 (Impact of permeability impairment on NPV),

Figure 1

Figure 2 (Impact of oil price on NPV)

874
Figure 2

and Figure 3 (Impact of stimulation cost on NPV) indicate that the NPV is rather
insensitive to treatment variables such as permeability impairment, oil price and
stimulation fluid cost.

875
Figure 3

(Table 1, below, lists the well and reservoir data used to generate these plots).

L = 1500 ft rw = 0.4 ft

h = 100 ft reH = 3000 ft

kH = 20 md Iani = 3

= 1.5 cp
= 30%

p = 1500 psi k/ks = 20

Price of oil = $15/STB

Discount rate = 25%

Cost of inj. fluids = $5/gal

Cost of inj. time = $10,000/hr

876
Fixed costs = $100,000

Table 1: Well and reservoir data for Figures 1, 2 and 3.

All three of these figures show a continuous increase in NPV with injected volumes until
the damage is totally removed. Wasteful overtreatment forces the NPV to decline. This
decline is more severe for sandstones, where no additional permeability is created (ki
=k).

Two important conclusions that we can draw from Figure 1 and Figure 3 are that:

a full stimulation treatment (assuming it is physically feasible) generally delivers the


highest NPV;

overtreatment is more desirable than partial stimulation note that the NPV
curves are rather flat if the damage penetration is exceeded, and compare this
with the steep incline of the partial stimulation curves.

These conclusions are particularly valid for wells producing at very high rates.

Of course, we should always consider stimulation treatments over the long term; that
is, for the life of the well. We therefore need to design completion and stimulation
programs during the well planning stage.

If we are planning a cased hole configuration, we should consider partial completion as


a design option. Goode and Wilkinson (1991) have shown that a 60 percent perforated
well can deliver 90 percent of the theoretical possible production. This possibility
becomes even more attractive if stimulation is considered (Frick and Economides,
1992).

Figure 4 is a plot of expected production rate versus the number of open segments.
The total open length of this 1600 ft [488 m] horizontal section is held constant at
1000 ft [305 m] the well is completed across 62.5 percent of the total section
length.

877
Figure 4

We can clearly see that there is a minimum number of segments required for
optimizing production. In this case, 5 segments, evenly distributed along the wellbore,
deliver 87.5 percent of the maximum possible production of 2570 STB/D [409 m3/D].
However, if the partially completed well is damaged (seq = 34), we attain only 60
percent of the totally perforated (and damaged) production.

We can readily adapt the concept of Net Present Value to partially completed wells as a
means of selecting the optimal completion and stimulation design.

The NPV equation for a partial completion design is

NPV = (3.21)

where

(3.22)

878
with Lperf,k being the perforated length of segment k and C3 being the perforation
costs in $/ft. Np,n is the difference between the sum of the incremental cumulative
productions of the stimulated well's perforated segments and the cumulative
production of the fully open and unstimulated (damaged) well in year n.

L = 1600 ft rw = 0.417 ft

h = 30 ft k/ks = 15

kH = 5.1 md Iani = 1.5

= 1.5 cp
= 38%

Price of oil = $15/STB

Discount rate = 25%

Inj. fluids cost = $7/gal

Const of inj. time = $10,000/hr

Perforation costs = $200,000 per 100 ft

Fixed costs = $100,000

Option Number of Total open Design option


open segments length, ft. (open-blank-
open)

A 1 1600 1600

B 5 320 64-320-64

C 5 650 130-238-130

D 5 1000 200-150-200

E 5 1280 256-80-256

F 7 700 100-150-100

G 11 1100 100-50-100

Table 2: Data for NPV calculations of Figures 5-8.

Figure 5 ,

879
Figure 5

Figure 6 ,

880
Figure 6

Figure 7 and Figure 8 compare several completion and stimulation design options.

881
Figure 7

Note that among these options, there are significant NPV variations ( Table 2 contains
the data for this figure).

882
Figure 8

The type of economic analysis described above is recommended for every horizontal
well stimulation treatment. Combined with careful pressure and rate monitoring during
the treatment and the experience gained from post-treatment job evaluation (e.g. by
running a production log), the NPV analysis is a useful tool for enhancing stimulation
design.

Acidizing Physics

Until recently, the industry regarded studies on the physics of matrix acidizing
primarily with academic interest. The emergence of long-radius horizontal wells
prompted a re-examination of the issue because of the massive volumes of stimulation
fluids required for these wells, and the potential benefits (and losses) associated with
stimulation and completion designs. Ongoing acidizing physics research is attempting
to define the processes beyond the "rules of thumb" presently considered adequate for
vertical wells (Daccord et al., 1989, Schechter, 1992, Kurmayr et al., 1992, Pichler et
al., 1992 and da Motta et al., 1992).

In carbonate acidizing, research activities focus on quantitative methods for describing


complicated reaction patterns (Daccord et al., 1989 and Kurmayr et al., 1992), and on
the development of stochastic simulators (Pichler et al., 1992).

883
For sandstones, researchers are developing comprehensive three-mineral models to
account for more heterogeneous mineralogies (da Motta et al., 1992), and working on
techniques for calculating acid penetration profiles for different geometries.

The purpose of all of these studies is to allow more precise job designs, based on a
more detailed understanding of the physics involved in matrix stimulation.

884
HIDRAULIC FRACTURING OF HORIZONTAL WELLS

Hydraulic Fracturing Of Horizontal Wells

In the types of reservoirs that benefit from hydraulic fracture stimulation, unfractured
horizontal wells are generally not attractive substitutes for fractured vertical wells. Like
many rules, however, this one has exceptions:

In formations with massive natural fracture networks, proppant screenouts


during treatment execution may preclude the placement of long hydraulic
fractures.

We also have to consider that hydraulic fractures are usually vertical and
(away from the well) normal to the minimum horizontal stress. This means that
they are likely to be normal to the minimum horizontal permeability. Such an
orientation, while unavoidable, is the least favorable direction for a hydraulic
fracture. In formations with large horizontal permeability anisotropy, a
horizontal well drilled normal to the maximum permeability (i.e., at right angles
to the expected hydraulic fracture trajectory) may be quite attractive.

Figure 1 (Hydraulic fracture and possible horizontal well directions in a stress and
permeability anisotropic formation: (a)

885
Figure 1

fractured vertical well; (b) unfractured horizontal well) shows possible directions for
vertical hydraulic fractures and horizontal wells in a reservoir having significant stress
and permeability anisotropy.

These exceptions aside, an unfractured horizontal well in a relatively low permeability


formation is not likely to be economically preferable to a fractured vertical well, unless
we envision the horizontal well itself as having fracture stimulation potential.

The ability to direct a horizontal well that is a candidate for hydraulic fracturing allows
for two distinct, mutually exclusive configurations:

We can drill along the maximum horizontal stress, and therefore along the
hydraulic fracture trajectory. This is known as the longitudinal configuration.

We can drill along the minimum horizontal stress, resulting in a transverse


hydraulic fracture direction relative to the horizontal well trajectory. With
adequate zonal isolation, this configuration may entail multiple, parallel
fractures, intersecting the well at predetermined intervals.

General Considerations
Hydraulic fracturing is an established method for improving the productivity index of a
well that would not be economically attractive even after matrix stimulation. This
would be the case, for instance, in low permeability oil or gas wells (i.e., where k 2
md).

Contrary to frequent popular misconception, hydraulic fracturing does not alter


reservoir permeability. Instead, it superimposes a negative skin effect on the well.
Thus, the post-treatment productivity index of a fractured oil well, operating under
steady-state conditions, is

(4.1)

where s would be a large, negative number (e.g., -6). Equation 4.1 suggests that for a
given pressure gradient (pe - pwf), the production rate (q) would be considerably larger
than the pre-treatment value (all other flow types would have similar expressions). We
could write a similar expression for the injectivity index (pe - pwf becomes pinj - pe);
many fractured wells are put into injection service.

Because a production engineer can "allocate" the new productivity index in a manner
that he or she deems appropriate, an acceptably reduced production rate can result in
a lower pressure gradient, and therefore in a lower bottomhole pressure (pwf). This is

886
particularly interesting because drawdown-sensitive formations (i.e., those that
produce sand or have water or gas coning tendencies) can be good candidates for
hydraulic fracturing, even if they do not meet the traditional "low permeability"
requirements for fracturing. In fact, many wells today undergo combination treatments
of hydraulic fracturing and gravel packing ("frac-and-pack").

The skin effect in Equation 4.1 has no relation to any pre-treatment skin effect
influencing the radial flow into the well. Hydraulic fracturing bypasses this damage
zone, and so we can ignore this zone's influence on post-treatment well performance.

Hydraulic fracturing involves the high-pressure injection of polymer solutions in water,


oil or two-phase fluids (i.e., gas and water, gas and oil, oil and water). The idea is to
first break down the formation (fracture initiation) and then propagate a fracture to an
appropriate, designed length. Proppant is added to the fluid, so that when the pressure
subsides, the created fracture closes on the proppant, forming a highly conductive path
from the well to the reservoir.

Resisting the applied hydraulic pressure is the stress distribution around the well. Away
from the well, the fracture will be normal to the minimum resistance (i.e., the
minimum stress). In almost all petroleum engineering applications, this corresponds to
the minimum horizontal stress. Thus, hydraulic fractures are usually vertical. Their
trajectory is then controlled by the direction of the principal horizontal stresses.

For a vertical well, coinciding with one of the principal stresses, the fracture initiation
or breakdown pressure (pbd), is given by the well-known Terzaghi (1923) formula.

(4.2)

where

and = minimum and maximum horizontal stresses

T0 = tensile stress of the rock

p = pore pressure

For a horizontal well drilled along the minimum or maximum horizontal stress directions, we can
write an analogy for Equation 4.2, substituting the two stresses with their counterparts.

In all cases, the near-well stress concentration is likely to result in a fracture initiating
longitudinally from an open hole, and then turning, after a few well diameters, to
become normal to the minimum stress. This occurs irrespective of the well direction
(Yew and Li, 1987, McLennan et al., 1989).

The definition of principal stresses implies that along these three directions, shear
stresses vanish. It follows that any well trajectory not coinciding with one of the
principal stresses will give rise to a non-vanishing shear stress component, greatly
complicating the calculation of breakdown pressure.

887
Thus, while we have established the need to fracture horizontal wells in reservoirs
where vertical wells are fractured, the actual treatment execution involves both a
blessing and a curse. The blessing, of course, is the increased productivity index
resulting from a successful job. The curse is that a horizontal well that is drilled
indiscriminately and then fractured may develop a very complex near-wellbore
geometry, as shown in Figure 1 (Vertical fracture initiated from arbitrarily oriented
horizontal well).

Figure 1

This complex geometry may result in a tortuous fracture path, causing an elevated
fracture propagation pressure, and in certain cases, proppant screenouts during
execution (Owens et al., 1992). The same tortuous path may restrict production and
considerably reduce well performance (Deimbacher et al., 1993).

Therefore, where drilling is not affected by logistical considerations (e.g., offshore


platforms) we can direct horizontal wells in two limiting modes, as mentioned earlier in
this section: longitudinal or transverse to the fracture trajectory.

888
We must emphasize that even if we choose the transverse configuration and properly
drill the well to receive such a fracture, the length of the perforated interval cannot be
too large. Otherwise, the fracture will initiate longitudinally and then turn transverse (
Figure 2 ), with obviously undesirable consequences. Deimbacher et al. (1992 and
1993) have shown that this interval should not exceed 1.5 times the well diameter.

Figure 2

Once we initiate the fracture and it begins propagating along its final trajectory, and
after accounting for the additional pressure drop due to near-well complexities, the
rest of the operation is subject to the same considerations as the fracturing of vertical
wells.

Researchers have developed several models to describe vertical well fracture


geometry, including two-dimensional (2D), pseudo-three-dimensional (p-3D) and fully
three-dimensional (3D) formulations. Of these, the closed form, 2D, constant-height
approximations known as the PKN model, developed by Perkins and Kern (1961) and
augmented by Nordgren (1972), have been widely used for fractures whose length is
considerably larger than the height. (Fracture height is controlled primarily by the

889
horizontal stress contrast between the target and overlain or underlain formation. See
Simonson et al. (1978) for an explanation.)

For fractures where the length is smaller than the height, the KGD model is a good
approximation (Khristianovich and Zheltov, 1959; Geertsma and de Klerk, 1969).

While these 2D models have proven adequate for vertical well fracture modeling
approximations, the complex geometry of fracturing highly deviated and non-aligned
horizontal wells requires the out-of-plane capabilities of full 3D models. In the absence
of such complications however, we may use the PKN model. For a Newtonian fluid, we
can obtain the average width ( ) in inches, of a fracture of half-length xf, from

(4.3)

where

qi = injection rate (Bbl/min)

= fracturing fluid viscosity (cp)

v = Poisson ratio

G = elastic shear modulus (psi)

= a geometric factor equal to 0.75

Note: Where is expressed in mm and q, , xf and G are given in SI metric units, 0.3 becomes
16.216.

The bracketed expression on the right hand side of Equation 4.3 gives the maximum
width value (at the wellbore), whereas the multiplier in the parenthesis provides a
geometrically averaged width.

If we assume a fracture height hf, we can then write a material balance that combines
the created fracture volume, leaked-off fluid and total injected volume:

(4.4)

where

qi = injection rate

ti = total injection time

890
Af = total area of both fracture faces (= 4xfhf )

CL = leakoff coefficient ( ), introduced by Nolte (1979)

rp = ratio of the permeability to the fracture height

KL is given by (4.5)

where is the efficiency (i.e., the ratio of the created fracture volume to the total injected
volume).
Because Equations 4.4 and 4.5 are related ( is simply the ratio of the first term on the right-hand-
side of Equation 4.4 divided by the left-hand-side), we can solve the two equations by trial and
error. In fact, KL varies only between 2.7 and 3.14, and thus we can readily obtain an approximate
solution of Equation 4.4 for the injection time ti. This is known as an inverse solution (i.e., what total
time would be required to generate a fracture of interconnected dimensions w and xf (Equation 4.3)
while it undergoes leakoff, described by the second term of Equation 4.4).

After calculating ti, we can approximate the pad volume (Vpad), which is the volume of
fluid before proppant is injected (Nolte, 1986):

(4.6)

Nolte (1986) also developed a technique for generating a continuous proppant addition
schedule, which depends on the efficiency and the end-of-job slurry concentration.
Thus, we can calculate the total mass of proppant (Mp) by integrating the slurry
concentration profile from tpad to ti. The total proppant mass divided by the created
fracture area gives us the proppant concentration in the fracture, Cp (lb/ft2) which
then is related to the propped width of the fracture (in inches) by

(4.7)

where

p and p are the porosity and density of the proppant pack.

In SI metric units, with wprop expressed in mm, 12 becomes 1000.

The propped width (Equation 4.7) and the hydraulic width (Equation 4.3) are thus only indirectly
connected. Henceforth, we will use the variable w to denote propped width, which is the width
affecting fractured well performance. This factor appears in the dimensionless fracture conductivity
(FCD) relationship, along with the fracture permeability (kf) the reservoir permeability (k) and the
fracture half-length (xf).

891
(4.8)

Cinco et al. (1978) presented the first important work describing the performance of
"finite conductivity" fractures. Needed are the FCD and xf. In general, very low-
permeability reservoirs require long fractures, whereas higher permeability reservoirs
require fractures with as high a conductivity as possible, with the length requirement
being of secondary importance.

In designing the optimal fracture size, we must balance the expected post-treatment
benefits against the job costs (fluids, proppant and service). One selection criterion is
the Net Present Value (NPV), introduced in a systematic manner by Meng and Brown
(1987).

Performance of Fractured Horizontal Wells


The justification for fracturing a horizontal well-and for drilling it in the first place is a
favorable economic evaluation, in which the project's anticipated incremental benefits
more than cover its potential incremental costs. We may conduct such an evaluation
using the NPV criterion.

The two limiting fracture configurations (longitudinal and transverse) have different
production characteristics. A horizontal well that is drilled in an arbitrary direction has
a more complex response, and warrants a special discussion. In all cases, we can
assume that production from the hydraulic fractures into the well is overwhelmingly
dominant (as it is for vertical wells), and thus we can neglect radial inflow.

To evaluate the desirability of a fractured horizontal well, we have to compare it to


vertical well fracture performance. We assume transient flow for almost all cases.
Boundary effects generally do not appear for a substantial amount of time (usually 2 to
3 years).

Thus, to forecast the well performance at any time, we must first calculate the
dimensionless time tDxf:

(4.9)

where

t = time, hours

= porosity

= viscosity, cp

ct = total system compressibility, psi-1

892
xf = fracture half-length, feet.

(Where k is given in md, t in hours, in cp, ct in kPa-1 and xf in m, 0.000264 becomes 3.557 x 10-6.)

Then, we must calculate the dimensionless fracture conductivity (FCD), using Equation
4.8. Once we determine tDxf and FCD, we can use Figure 1 (Finite conductivity
fracture solution) from Agarwal et al.

Figure 1

(1979) to obtain the corresponding dimensionless pressure (pD), defined for oil by

(4.10)

and approximated for gas by

893
(4.11)

where

k = reservoir permeability, md

h = reservoir thickness, ft

pi = initial reservoir pressure, psi

pwf = flowing bottomhole pressure, psi

q = flow rate, STB/d (for oil) and MSCF/d for gas

= oil or gas viscosity, cp

Z = gas deviation factor

T = gas temperature, R

(Where k is given in md, pe and pwf in kPa, q in m3/D, in Pa-s and T in K, the constants 141.2
and 1424 become 1867 and 1.312 x 106, respectively.)

This calculated pD value leads readily to a value for q if pwf is known, or if pwf is not
set, to a transient IPR curve.

Example: Forecast of Fractured Well Performance

Assume the following reservoir data:

k = 1 md h = 50 ft

= 0.1 = 0.7 cp

B = 1.1 Bbl/STB ct = 2 x 10-5 psi-1

xf = 1000 ft kfw = 1000 md-ft

Calculate the fractured well production rate after 3 months if pi = 4500 psi and pwf = 2000 psi.

Solution:

From Equation 4.9

894
(4.12)

From Equation 4.8, FCD = 1, and therefore, from Equation 4.4, pD =1.7.

Thus, from Equation 4.10,

= 675 STB/D[107 m3/D (4.13)

Longitudinal Fractures

Figure 2 illustrates the comparison of a longitudinally fractured horizontal well with a


fractured vertical well.

Figure 2

The horizontal well has been drilled deliberately along the expected fracture azimuth.
To assess the relative attractiveness of the horizontal well, we assume that the lateral
section length is equal to the tip-to-tip fracture length (2xf). In practice, we may
perform the fracture treatment in stages, along the horizontal well axis.

895
In comparing the Productivity Index ratios of the two configurations in Figure 2 ,
Economides et al. (1991) have presented the generalized curves of Figure 3
(Productivity Index Ratio between longitudinally fractured horizontal well and fractured
vertical well).

Figure 3

On the horizontal axis is the dimensionless fracture conductivity FCD. Dimensionless


time (tDxf) is a parameter (Equation. 4.9). The productivity index ratio on the vertical
axis is that of the longitudinally fractured horizontal well divided by that of a fractured
vertical well.

From Figure 3 , we may draw two conclusions, which dictate the need for a fractured
horizontal well:

As FCD increases, the PI ratio decreases. This should not be surprising. For large FCD
values, the fracture is already efficient. The outer reaches contribute substantially, and thus
there is no particular benefit in turning the well from vertical to horizontal to access them.
For low FCD values (meaning relatively inaccessible parts of the fracture length), turning
the well could substantially increase the PI ratio.

896
Thus, the longitudinal configuration would not be attractive in low permeability
formations (i.e., high FCD). It may be attractive, subject to economics, in
higher permeability reservoirs that are candidates for hydraulic fracturing.

The PI ratio decreases with time, as the importance of fracture/well


configuration in pseudo-radial or emerging boundary-controlled flow diminishes.
In these cases, the fracture simply impacts the effective wellbore radius.

Transverse Fractures

The next possibility is to generate transverse fractures ( Figure 4 , A horizontal well


drilled along the minimum horizontal stress intended to accept transverse fractures).

Figure 4

The idea is to create a number of appropriately spaced parallel fractures, as shown in


Figure 5 (Multiple transverse fractures, appropriately spaced, intersecting a horizontal
well).

897
Figure 5

Zonal isolation considerations apply in this situation; a cased, cemented and


appropriately perforated completion is a precondition for transverse fracturing.

Mukherjee and Economides (1991) have defined a choke skin effect to account for the
flow rate reduction resulting from the convergent linear-to-radial flow inside the
fracture:

(4.14)

The skin effect is additive to the dimensionless pressure of Equations 4.10 and 4.11.
Thus, for each perfectly transverse fracture in an oil reservoir,

(4.15)

(Note: Where q is expressed in m3/d (with k in md, pe and pwf in kPa, B in m3/m3, in
cp and h, L and rw in m), 141.2 becomes 1867.)

898
If the reservoir permeability (k) is small, then so is sc, which results in a relatively
lesser production rate drop. Therefore, the transverse fracture configuration in
contrast to the longitudinal configuration is a likely scheme for a low-permeability
reservoir. The transverse and longitudinal configurations are thus mutually exclusive in
almost all applications.

Example: Production Rate Drop from Transverse Fracture

Calculate the 3-month rate reduction if the fracture described in Example 4.1 was
transverse to a horizontal well. The well radius is 0.35 ft.

Solution:

The skin effect (Equation 4.14) is

(4.16)

Then, from Equation 4.15 (and remembering that pD = 1.7), q = 624 STB/D [99
m3/D]. This represents a reduction of 50 STB/D [7.95 m3/D]. If k was larger (or kfw
smaller), than the skin effect could be substantially larger, with a correspondingly
greater production loss.

Turning Fracture Effect

Deimbacher et al. (1992) presented an approximate expression for the width reduction
of a fracture turning from longitudinal to perfectly transverse. This reduction can be
expressed as a ratio:

(4.17)

where

d = well diameter

L = interval from which the fracture initiates. (i. the perforated interval)

Thus, if L >1.5 d, a width reduction occurs.

For a well at angle with the expected well trajectory, Deimbacher et al. (1993)
presented a similar approximation for width reduction:

(4.18)

899
We can incorporate the results of Equations 4.17 and 4.18 in the expression for the
skin effect, adjusting the fracture width downward. This would result in an increase in
the skin effect and a drop in the production rate.

Example: Production Rate Drop from a Longitudinal to a Transverse Fracture

Suppose that the transverse fracture described in Example 4.2 initiates from a 100-ft
interval. What would be the 3-month production rate?

Solution:

From Equation 4.17,

(4.19)

From Equation 4.15, q = 76.5 STB/D [12.2 m3/D], compared to 624 STB/D [99.2
m3/D] for a perfectly transverse fracture. This calculation shows the importance of
minimizing the length of the perforated interval if longitudinal fractures are not
intended.

Fracturing Pressures

The shape of an arbitrarily deviating well and fracture ( Figure 1 ) suggests a non-ideal
fracturing pressure response: one in which the fracture initiation and propagation
pressures differ from and in fact are much higher than those for a vertical well.

900
Figure 1

For fracture initiation, and with the idea of elastic breakdown (analogous to Terzaghi's
criterion of Equation 4.2), the non-vanishing shear stress component would lead to a
frequently substantial additional pressure. Owens et al. (1992) described this
condition. They also addressed the relationship between fracture initiation pressure
and fracture propagation pressure:

which is model-dependent, and is affected in turn by a number of other


factors such as the interlayer stress contrast ... the viscosity of the
fracturing fluid ... injection pressure and near-wellbore tortuosity.

They did not attempt to model the fracture propagation pressure from a horizontal
well. However, they did state that:

for relatively small tensile stress values and moderate fracture height
migration, fracture propagation pressure has been observed ... to
remain largely constant and very near the fracture initiation pressure.

They matched their model with a large number of treatments.

901
Weijers et al. (1992) interpreted the same data and attributed their values to the
added friction pressure caused by the near-wellbore tortuosity. These two studies are
not in contradiction; they simply address different facets of the fracturing process. In
reality, the fracture geometry and the fracture/well contact that cause the elevated
initiation pressures also cause the additional fracture propagation pressures attributed
to friction. Reducing the tortuosity, either by injecting acid (in carbonate formations) or
"sandblasting" with high injection rates and high-mesh solids, should reduce the
propagation pressure. This has been observed in the field (Owens, 1992).

Owens et al. (1992) published equations describing the elastic breakdown pressure for
an arbitrarily oriented horizontal well at any angle of deviation with the final hydraulic
fracture azimuth (these equations are outside the scope of this text, and so we refer
the interested reader to the original publication). If the two horizontal stresses and
their directions are known, and we determine the Poisson ratio and tensile stress of the
rock, we can use these equations to predict the fracture initiation pressure from any
horizontal well. Figure 2 shows the authors' match between predicted and observed
fracturing pressures from a North Sea Chalk formation.

Figure 2

We see a substantial increase in fracturing pressure as the horizontal well deviation

902
from the longitudinal increases; the difference in the fracturing pressure between zero
and 90 degrees deviation is approximately 2000 psi [~13800 kPa].

Fracture Design
The first step in deciding whether to fracture a horizontal well (or whether to drill
laterally at all) is to determine whether the reservoir is a candidate for a fractured
vertical well. The NPV criterion (Meng and Brown, 1987) is appropriate in this regard.

NPV depends heavily on expected post-treatment well performance. On this basis,


many low-to-moderate permeability reservoirs become candidates for fracture
stimulation.

In the NPV analysis, we must account for expenses associated with well completion
and zonal isolation before attempting the treatment, since these costs are important to
the job's success.

A horizontal well drilled in a formation where the fracture NPV of a vertical well
appears attractive must demonstrate a positive incremental NPV. Earlier, we identified
the limiting cases and their performance characteristics. If these limiting cases are not
possible (as may be the case in offshore or arctic locations, where many wells are
drilled at various angles), turning fracture effects will have a major impact on well
performance. Thus, the calculated NPV of a potential option must not only be higher
than that of the base case (i.e., a fractured vertical well), but it must be the best
among the options. The study results may limit lateral wells within certain angles from
the expected hydraulic fracture azimuth.

Once we calculate the NPV and decide on the well trajectory, we begin detailed work
on the drilling program. This program obviously has to take into account in-situ stress
directions and their magnitude. We obtain these measurements from a pilot vertical
hole before drilling the horizontal section. Furthermore, cementing, perforating and the
length of the perforated interval(s) must be properly planned and executed.

Decision on Well Trajectory

The following is a step-by-step procedure for deciding the trajectory of a horizontal


well that may be a candidate for hydraulic fracturing:

1. Perform an NPV calculation for a fractured vertical well. If the economics are not
attractive when compared with those of an unfractured vertical well, a horizontal well would
not be a likely fracture candidate (on the other hand, an unfractured horizontal well may
be).

2. If the fracture NPV of a vertical well is attractive, and if the reservoir


permeability is relatively high (k 2 md) then consider the longitudinal fracture
configuration.

The incremental NPV must take into account the incremental post-treatment
performance, balanced by any incremental costs to drill and complete the well.

903
For the longitudinal configuration, an open hole completion may be
feasible if there is no need for lateral control caused by heterogeneities
(e.g., faults, natural fractures or differential pressure depletion).

3. If the reservoir permeability is smaller (k < 2 md) then consider multiple transverse
fractures. (In the next section, we discuss the number and size of these fractures.)

4. For multiple horizontal wells drilled at various angles from the expected
fracture azimuth, estimate their post-treatment performance. Certain angles
may result in considerably lower incremental NPV (or none at all), and thus a
limiting angle may need to be established. This analysis should be done for all
potential fracture candidates drilled offshore or from a drilling pad.

Number of Fractures and Size Optimization

Based on NPV analysis, Mukherjee and Economides (1991) have presented a technique
for selecting the appropriate number and size of transverse hydraulic fractures. This
method applies to the multiple fracture configuration shown in Figure 1 .

Figure 1

904
The technique involves first making the NPV calculation for a vertical well, as described
in Figure 2 from McLennan et al.

Figure 2

(1989). In this example, the optimal fracture half-length is 1100 ft [335 m],
corresponding to a 3-year NPV of $4.5 million.

For the optimal number of transverse fractures, the horizontal well length is divided by
two, three, four or more evenly spaced entry points, as shown in Figure 1 . For each
configuration, the NPV of an individual fracture is calculated. The main features
include:

reduction to the post-treatment production;

allocation per fracture of the incremental drilling and well completion costs,
beyond what a vertical well would cost. This includes the costs of execution and
zonal isolation for each fracture;

performance reduction as a result of interference from closely spaced


fractures. (For 1 to 2 years' NPV calculations, for low-permeability formations
and for fractures spaced 400 to 500 ft [122 -152 m] apart, this reduction is
likely to be small.)

Figure 3 from McLennan et al.

905
Figure 3

(1989) shows the individual NPV of one of four transverse fractures intersecting the horizontal well.

The optimal fracture half-length is 500 ft (always considerably smaller than the
optimum size for the fractured vertical well) and the corresponding individual 3-year
NPV is $1.35 million. This results in (4 $1.35 million =) $5.4 million total NPV, which
is larger than the $4.5 million for the base case ( Figure 2 ). Remember that the
calculated total NPV for the transverse fractures already includes incremental costs
associated with the horizontal well.

The four transverse fractures whose individual NPV is shown in Figure 3 do not
represent the optimum number of fractures, but rather the minimum three fractures
would result in a total NPV of less than $4.5 million. A construction as shown in Figure
4 is necessary for identifying the optimum number of fractures.

906
Figure 4

Thus, Figure 3 must be repeated for any number of fractures until we reach the
maximum total NPV. For almost all applications, this number is likely to be between 5
and 9.

Fracture Execution and Perforating Considerations

The zonal isolation techniques are essential for placing multiple fractures. For
transverse or angled fractures, the perforated length and perhaps the phasing of the
perforations become important.

For fully transverse fractures, the length of the perforated interval should be limited to
less than 1.5 times the well diameter to avoid turning fracture effects and reduction of
the fracture width. Thus, Brown and Economides (1992) have recommended that
instead of perforating, the horizontal well could be notched with an abrasive jet. The
width of such a notch (cut to only 270 degrees to avoid the shifting of the casing)
would be sufficient to pump the treatment.

For horizontal wells drilled at an arbitrary angle to the fracture azimuth, one means of
avoiding tortuosity effects would be to perforate such that the holes form an angle with
the well that coincides with the expected fracture direction. The technology is

907
available, although such applications would require both excellent gun control and a
very accurate knowledge of the fracture azimuth.

908
MULTILATERAL, MULTIBRANCHED & MULTILEVEL WELLS

Introduction
It was a natural leap for the petroleum industry to go from highly deviated and
horizontal wells to far more complex well architecture, where branches can emanate
from one or more "mother" or "parent" holes.

Although the nomenclature is simply one of convenience and has been used
interchangeably by several authors, some of the more common definitions are
described below. To begin with, the main part of the well from the surface is often
called the vertical trunk. At times a main horizontal branch may be called a horizontal
trunk.

Multilateral Wells: This term refers to several largely horizontal wells that are drilled from
the same vertical trunk and generally intended for the same reservoir. A large number of
these branches can be drilled at angles to one another, including a configuration where
several wells, drilled in largely the same plane, resemble a bicycle wheel. A limiting but
common configuration is the opposing lateral, where two wells are drilled at nearly a 180o
angle to each other. A variant is the cross configuration, where the second well is drilled at
a 90o angle to the first.

Multibranched Wells: Several authors have used this term to describe


branches drilled transversely to a horizontal trunk. Although often these
branches are themselves horizontal, at least two authors have suggested uses
for vertical branches drilled from horizontal wells.

Multilevel Wells: These wells consist of horizontal branches drilled from the
same vertical trunk, but at different depths. They can target different
formations, or they can be drilled within a single formation at different depths.
The same vertical trunk can sometimes be used to inject into and produce from
two different multilevel horizontals drilled into the same formation, thus
creating an injection/production couplet. Finally, a variant of this
injection/production scheme allows injection (e.g. steam) in all multilevel wells
and subsequent production from the same.

The section titled "Reservoirs & Well Architecture" outlines applications of these configurations and
rationalizes their uses in modern production strategies. (For additional configurations see Ehlig-
Economides et al., 1996.)

Reservoirs and Well Architecture


Multilateral Wells

A single horizontal well is normally intended to drain a single layer, and to do so in one
lateral direction. Envision, however, a situation where a reservoir is well defined by
modern formation characterization (e.g., 3-D and 4-D seismics), and found to consist

909
of lenticular lenses or to be interspersed by shale lenses. Other formations may consist
of braided channels. In such marginal flow units, it may not be economical to complete
dedicated wells from individual surface locations. Instead, multilateral wells may be
directed to intersect various flow units, with all branches drilled from the same vertical
trunk.

Part (a) of Figure 1 (Options for multilateral wells) represents a likely multilateral well
architecture targeting different geological flow units.

Figure 1

The design of such a configuration is based not only upon the appropriate knowledge
of geology, but also of fluids (i.e., commingled versus separated production). Ideally,
formation pressures should be taken into consideration in deciding the point in the
vertical trunk from where each lateral should be drilled.

Most prediction models do not consider the pressure drop along the horizontal well.
And for most horizontal wells, this is not a bad assumption. However, if the pressure
drop in the well (all from friction) is not negligible in relation to the reservoir pressure
drawdown, a single long horizontal well may not be desirable. In such a case, it may

910
be better to drill two opposing laterals that offer minimum interference to each other
(part b of Figure 1 , Options for multilateral wells). Such configurations are already
widely used in underpressured reservoirs or reservoirs that are susceptible to pressure
drawdown. This would be the case in strong bottom-water-drive reservoirs or naturally
fractured formations.

In the case of pronounced areal permeability anisotropy, Smith et al. (1995) have
proposed the "cross" configuration consisting of two wells drilled at 90o angles to each
other (part c of Figure 1 , Options for multilateral wells). This configuration has been
suggested as a means of reducing the risk from single well misorientation in highly
anisotropic reservoirs. Smith et al. (1995) have shown that the cross configuration is
superior to any single horizontal well (if the extra cost of drilling can be justified), and
clearly superior to a single well drilled in the wrong direction.

Retnanto and Economides (1996) investigated the time-dependent effects of drilling


multiple well branches (e.g., more than four) from the same vertical parent hole to
drain the same reservoir. In high-permeability reservoirs, cross-branch interference
quickly diminishes any incremental benefits over a single horizontal. But in low- to
moderate-permeability reservoirs (e.g., k < 10 md), incremental benefits can be
sustained and the net present value (NPV) of such projects can be positive and
substantial.

An interesting variant with a potentially major application is the tightly spaced parallel
laterals suggested by Ehlig-Economides et al. (1996). The authors propose the drilling
of several wells in a configuration resembling a "pitchfork," and the alternate opening
and closing of these wells to control water cresting. As the water crest reaches one
branch, the branch is shut in, allowing the crest to subside because of gravity. While
the first branch is shut-in, a neighbor branch is open and produces until the water
crest reaches its level; then the well is closed. After the crest subsides to a reasonable
level, the first branch is re-opened and the cycle is repeated. Several branches can
thus be envisioned with alternating open/close cycles. The authors calculate that with
parallel horizontal branches at a distance of 2hIani the oil recovery may exceed 50

percent. (h is the reservoir thickness and Iani = , is the index of horizontal-


to-vertical permeability anisotropy.)

Multibranched Wells

Multibranched wells generally are those in which either transverse horizontal or vertical
wells are drilled from a horizontal trunk (e.g., Figure 2 , Plane view of multibranched
well architecture with all branches largely at the same level).

911
Figure 2

Retnanto et al. (1996) developed a procedure for the optimum spacing length and
number of sidetracks, taking into account reservoir characteristics such as permeability
and permeability anisotropy. The same well construction technology can be applied to
a number of other reservoir geometries. For example, the horizontal well branches can
be used to intersect braided channels that are largely parallel but limited in length.

Another possibility is a river channel (Figure 3 , Multibranched wells draining a river


channel, plane view).

912
Figure 3

Our first instinct for the reservoir of Figure 3 might be to drill a single well along the
channel. With the help of seismic images, a good directional driller can trace the path
of the river even if it meanders. The problem is that frequently, the permeability along
the channel path may be several times the transverse permeability. In such a case, it
would be better to drill a horizontal mother hole (even if it is outside the channel but
tracing its path), and then to drill several short-radius branches normal to the path.
Such a configuration would surely result in higher production. If the incremental
drilling costs can be justified, this type of well architecture can have a wide application.

Regarding the transverse fracturing of horizontal wells, it is generally true that multiple
transverse fracturing, with proper zonal isolation and with limited fracture-to-well
contact (i.e., short perforated intervals or notched pipe), may be attractive in lower-
permeability reservoirs. Longitudinal fracturing of horizontal wells has been found to be
more appropriate in higher-permeability formations.

An alternative to these configurations is to drill vertical wells from a horizontal branch


and then hydraulically fracture them in sequence (Figure 4 , Fractured vertical
multibranched wells vs.

913
Figure 4

a transversely fractured horizontal well. After Ehlig-Economides et al., 1996) Two of


the benefits of this configuration are that

it is not burdened by the tortuosity problems that are likely to appear when fracturing a
horizontal well;

it provides a way of overcoming the stability problems that horizontal wells


may encounter in loosely consolidated formations. A horizontal well, drilled
above the formation but in a stable layer, can readily form the mother hole
from where much more stable vertical wells can be drilled.

Multilevel Wells

Single horizontal wells are attractive in reservoirs having a favorable index of vertical-
to-horizontal permeability anisotropy that is, where there is good vertical
permeability relative to the horizontal permeability. The greater the formation
thickness, the more important this requirement becomes. While it is not the primary
consideration in thin reservoirs (e.g., <40 ft), a favorable anisotropy index is critical to
the economic attractiveness of a horizontal well in a very thick reservoir. This is why
massively naturally fractured carbonates have made such good candidates for
horizontal wells; thick reservoirs often have excellent vertical permeabilities.

Conversely, thick sandstone formations (especially laminated reservoirs) have been


less attractive candidates for horizontal wells. This situation can be changed if one
envisions multilevel wells ( Figure 5 , Stacked multilateral wells for heavy crude
reservoirs). The distance and number of these branches is, of course, subject to
optimization based on reservoir characteristics.

914
Figure 5

An offshoot of the architecture of Figure 5 is the possibility of using the wells in cyclic
injection and production ("huff & puff"), which can be readily applied for steam
injection in thick, heavy crude oil, reservoirs. A variant would be to use the top branch
for steam injection through tubing while production is from the bottom branch(es)
through the annulus, thus using the same vertical well. Similarly, the bottom may be
used for hot water injection, with the higher branch(es) used for production.

A Production Prediction Model


We cannot readily use single-horizontal-well models to predict the performance of
more complex well configurations. Economides et al. (1996) have introduced a
comprehensive multi- and single-well productivity model that allows the prediction of
any plausible well configuration within a drainage area. In turn, these predictions can
be used to assess the economic attractiveness of complex well architecture. The model
is general and allows the calculation of both transient and pseudosteady-state flow.
The latter is particularly useful for long-term performance predictions.

Central to the model is the calculation of the productivity index given by

915
(S.1)

where is the reservoir pressure, is the calculated dimensionless pressure and is the

average reservoir permeability . s is the summation of all damage and pseudoskin


factors. Dimensioned calculations are done on the basis of the reservoir length, xe; is the horizontal
well length.

The generalized solution to the dimensionless pressure, , starts with early-time


transient behavior and ends with pseudosteady state if all drainage boundaries are felt.
At that moment, the three-dimensional is decomposed into one two-dimensional
and one one-dimensional part,

(S.2)

where is a "shape" factor, characteristic of well and reservoir configurations in the horizontal
plane, and sx is the skin accounting for vertical effects.

The expression for this skin effect (after Kuchuk et al., 1988) is

(S.3)

and , describing eccentricity effects in the vertical direction, is

(S.4)

which is negligible if the well is placed near the vertical middle of the reservoir.

Well Shape Factors

916
For the calculation of long-term productivity, we can introduce an appropriate shape
factor, . Table 1a, Table 1b

Table 1b

, and Table 1c list shape factors for several well configurations, as presented by
Economides et al.

917
Table 1a

(1996) and Retnanto et al. (1996).

918
Table 1c

If the well is not located in the middle of the drainage area but is off-centered, the
shape factors proposed in Economides et al. (1996) can be applied with the use of a
correction factor. The correction factor can be derived from the equations proposed by
Babu and Odeh (1989) :

(S.5)

where xm is the midpoint of the well. The off-centered well location can also be accounted for by an
analogous skin effect:

(S.6)

To account for permeability anisotropy with kx, ky, and kz in the x-, y- and z-directions, the following
transformation, presented by Besson (1990), must be applied.

919
Transformations for the Well

Length

(S.7)

Wellbore radius

(S.8)

with

(S.9)

(S.10)

Transformations for the Reservoir Dimensions

(S.11)

(S.12)

(S.13)

(S.14)

920
Using these transformations and adjusting the well and reservoir variables, the pD calculation can
then proceed in the normal fashion.

Example Application: Single Horizontal vs. 6-branch


Multilateral Well Performance
Suppose that a reservoir drainage can be considered as a square drainage with xe =
3,000 ft and ye = 3,000 ft. (See Table 1a , Table 1b , and Table 1c for schematic
nomenclature). Also suppose for simplicity that B = 1 cp-resbbl/STB. The well radius,
rw is 0.328 ft.

Several calculations are presented below. The first considers a single horizontal well
with L = 1,200 ft and parallel to the x-axis.

I. Permeability Isotropic Cases

Consider = kx = ky = kz = 5 md.
A. Impact of reservoir height

Assume a well in the vertical middle but consider two reservoir heights: 50 ft and
(extraordinary) 500 ft.

From Eq. S.3,

sx = 3.2 - 0.007 = 3.2 (for h = 50 ft);


sx = 5.5 - 0.07 = 5.4 (for h = 500 ft).
From Table 1a, Table 1b , and Table 1c and noting that xe = ye and L/xe = 1,200/3,000 = 0.4 the
shape factor CH can be obtained and it is equal to 2.64.

Thus, from Eq. S.2

= 12.6 + 1.27 = 13.88 (for h = 50 ft).


For h = 500 ft
pD = 1.26 + 2.16 = 3.42.
The productivity indexes can be obtained from Eq. S.1. Assuming that all skin effects are zero, then

(for h = 50 ft)
and

(for h = 500 ft).


This exercise illustrates one of the best-known conclusions regarding horizontal wells: they are
particularly attractive in thinner reservoirs. While for vertical wells the productivity index ratio
between the h = 500 ft and h = 50 ft would be equal to 10 (since all other variables are presumed
the same), for horizontal wells the productivity index ratio is only 4.

921
B. Impact of an oblong drainage area

Suppose that xe x ye = 9 x 106 ft2 is constant but xe = 4ye. Thus, ye = 1,500 ft and xe
= 6,000 ft.

The ratio L/xe = 1,200/6,000 = 0.2 and from Table 1a, Table 1b , Table 1c (and by
extrapolation), CH = 4.11.

The dimensionless pressure pD can be obtained from Eq. S.2 (for h = 50 ft)

= 39.3 + 2.55 = 41.85


thus,

which represents a 34 percent decline from the square reservoir. This shows the impact of
boundary effects as they are nearer to the well.

(For h = 500 ft, pD = 9.33 and J = 3.62 STB/d/psi, representing a 26 percent decline.)

C. Effect of vertical eccentricity

Suppose that to retard water cresting, the well is drilled at the top 0.9 of the reservoir,
i.e., zw/h = 0.9. (Note: the same result would for zw/h = 0.1 because of symmetry
around the vertical middle, zw/h = 0.5.)

From Eq. S.4, h = 50 ft and L = 1,200 ft

= -0.01 + 1.17 = 1.16


(For h = 500 ft and Eq. S.4, se = 1.04.)

From Eq. S.3

sx = 3.2 + 1.16 = 4.36 (for h = 50 ft)


sx = 5.4 + 1.04 = 6.44 (for h = 500 ft)
The dimensionless pressure for h = 50 ft (see first term in the same calculation in Case A.)

thus,

922
representing only a 3 percent reduction from Case A where the well is assumed to be in the vertical
middle.

However, for h = 500 ft,

and

representing a 10 percent reduction.

The latter again shows the impact of the larger thickness. This result and the notion
that thinner reservoirs are relatively more attractive, can be readily attributed from a
reservoir engineering point of view to the lateral depth of the distortion of the flow
lines into the well.

II. Permeability Anisotropic Cases

D. Vertical-to-horizontal permeability anisotropy

Assume that kH = kx = ky = 5 md, but kz = 0.1 kH = 0.5 md.

There is a need here to use Eqs. S.7 to S.13 to adjust the radius and all lengths to
reflect this permeability anisotropy.

From Eq. S.14 the average reservoir permeability is

.
From Eqs. S.9 and S.10
(since kx = ky).
From Eq. S.7 the equivalent well length is

What this adjustment means is that in a permeability anisotropic reservoir, the apparent well length
shrinks. From Eq. S.8, the equivalent well radius is

Finally, from Eqs. S.11 to S.13 (h = 50 ft)

The ratios xe /ye and L/xe do not change (from Case A) and they are equal to 1 and 0.4, respectively.
Thus, again from Table 1a, Table 1b , and Table 1c , CH = 2.64.

The skin effect, sx is slightly different and equal to 3.6, reflecting the equivalent height
of 108 ft.

923
Thus, using the adjusted variables

= 3.97 + 1.43 = 5.4


and, thus,

a 19 percent reduction from the isotropic case.

For the thicker reservoir (h = 500 ft) the effect is far more pronounced. The ratios xe
/ye and L/xe still do not change leading to CH = 2.64, while the skin effect sx = 5.9
(using the adjusted thickness h = 1,080 ft.)

Thus,

= 0.4 + 2.34 = 2.74


and

representing a very large 54 percent reduction from the isotropic Case A.

This exercise illustrates the second best known conclusion of horizontal well
performance: in relatively thin reservoirs the impact of vertical-to-horizontal
permeability anisotropy is relatively forgiving, while in thick reservoirs it can be
devastating. In comparing Cases A and D and assuming p = 1,000 psi, the mere
vertical permeability difference (i.e., a sandstone vs. a carbonate reservoir) would lead
to a rate reduction of over 2,600 STB/d.

E. Areal permeability anisotropy

Lets again suppose that kz = 0.5 md and kH = 5 md but the latter is the result of kx =
5 ky (i.e., ky = 2.23 md and kx = 11.18 md).

The well is drilled parallel to the x-axis ( = 0o), i.e., along the maximum horizontal
permeability and normal to the minimum horizontal permeability.

While = 3.16 (from Eq. S.9), is different and it is equal to 0.67 (Eq. S.10.)

Thus, new equivalent lengths and radius must be calculated from Eqs. S.7, S.8 and
S.11 to S.13. The average permeability remains the same as in Case D.

Therefore :

924
.

.
and h = 108 ft (intact).

The skin sx is slightly changed (new rw) and is equal to 3.5.

The ratios are now

and (same as for Case A).


From Table 1a, Table 1b , and Table 1c , and (double) interpolation, CH = 4.

Thus,

= 4.03 + 1.39 = 5.42


and

representing a 46 percent reduction from the complete isotropic Case A and a 33 percent reduction
from the simple vertical-to-horizontal permeability anisotropy Case D. This shows the effect of well
misorientation.

For the thicker reservoir (h = 500 ft, h = 1,080 ft) the results are:

sx = 5.8, pD = 2.7 and J = 1.32 STB/d/psi representing a 73 percent reduction from


the complete isotropic Case A and a 42 percent reduction from the vertical-to-
horizontal permeability Case D.

F. Proper well orientation

Now, lets assume ky = 5 kx leading to ky = 11.18 md and kx = 2.23 md. From Eqs.
S.9 and S.10 = 3.16 and =1.5.

From Eq. S.7, L = 1,227 ft and from Eq. S.8, rw = 0.43 ft.

925
From Eq. S.11 = 3,060 ft and from Eq. S.12, = 1,365 ft leading to .

The ratio (intact).

From Table 1a, Table 1b , and Table 1c , and (double) interpolation CH = 2.4.

For h = 108 ft , sx = 3.7, thus

= 5.41 + 1.47 = 6.88


and

representing a 76 percent increase over the misoriented well (Case E.)

For h = 500 ft, sx = 6 and pD = 2.93 and J = 2.73 STB/d/psi representing 107 percent
increase over the misoriented well (Case E.)

These results show the clear importance of proper well orientation in the all too
common areal permeability anisotropic reservoirs.

Finally, a well can be allowed to rotate in the reservoir, using the angle in Eq. S.10,
which is the angle formed between the well and the x-axis, parallel to the xe side of
drainage.

III. 6-Branch Multilateral Well

Again, using Table 1a, Table 1b , and Table 1c, and assuming Ly = Lx, i.e., the
orthogonal branches are 0.5 of Lx, the shape factor CH = 1.33.

Assuming areal permeability isotropy (otherwise Lx and Ly must be adjusted) but


vertical to horizontal anisotropy, following Case D

= 2 + 0.36 = 2.36
and

or 128 percent increase over the single horizontal case. (Note that in Eq. S.2 the total length of the
branches must be inserted.)

Although slightly more cumbersome to show in these calculations (published


extensively by Smith et al., 1995), the multilateral well can be particularly useful in
reducing the risk in areally anisotropic reservoirs. Thus, the potential reduction from

926
misorientation shown in Case E will not be risked if a multilateral well (in its simplest
manifestation, the "cross" configuration) is drilled.

Reservoir Performance with Selected Complex Well Configurations


One of the oldest rationales for using multilateral wells is to reduce the economic risk
in areally anisotropic reservoirs by drilling a second branch at right angles to a first
branch.

Buchsteiner et al. (1993) have shown that areal permeability anisotropy is a real issue
in almost all reservoirs. Permeability anisotropy ratios are far more likely to be greater
than 3:1 than they are to be 3:1 or less. This is true not only in the obvious cases such
as massively naturally fractured formations (where permeability anisotropies of 50:1 or
larger should be expected), but also in presumably homogeneous sands or channels.
Yet in dealing with vertical wells, areal permeability anisotropy does not affect well

performance because the controlling permeability of radial flow is simply . In


horizontal wells, however, the well orientation is crucial (Smith et al., 1995).

Figure 1 (Incremental net present value ratio of single horizontal well compared to the
single horizontal well base case.

927
Figure 1

From Smith et al., 1995) shows the normalized 3-year NPV for a case study of a
misaligned single horizontal well in areally permeability-anisotropic reservoirs in
comparison to the isotropic case. Properly oriented wells (i.e., normal to the maximum
permeability) can take advantage of permeability anisotropy and can outperform
similar wells in less anisotropic formations. Conversely, the normalized NPV drops
sharply as the misorientation becomes more severe. The more severe the anisotropy,
the larger the impact of misorientation.

Clearly, if the permeability anisotropy and its direction are known, properly orienting a
single horizontal well is the most desirable solution. However, as Smith et al. (1995)
point out, if the permeability anisotropy and direction are unknown, a second branch,
at 90o can reduce the risk defined by Figure 1 .

These results are shown in Figure 2 (Incremental net present value ratio compared to
the single horizontal well base casecross configuration.

Figure 2

928
From Smith et al., 1995). The cross configuration leads to normalized Net Present
Values larger than 1 in all cases. More important, the worst ratio of maximum to
minimum NPV is only 1.13, whereas the ratio of maximum to minimum NPV for single
horizontal well is over 4 ( Figure 1. ) In Figure 2 , the extra cost of drilling the second
branch was incorporated in the calculation of the NPV.

Sustained incremental benefits from multilateral wells are likely to be more


pronounced in low- to medium-permeability reservoirs. In all cases, incremental
benefits derive from the far larger well-to-reservoir exposure afforded by the additional
branches. As pseudosteady state ensues and outer boundaries are felt, the productivity
indexes of all configurations converge. We can see this effect for the single horizontal
configuration, the "cross" configuration and the "star" configuration in Figure 3
(Dimensionless production rate vs.

Figure 3

dimensionless time for horizontal-to-horizontal permeability isotropy, Cases A through


C. From Retnanto and Economides, 1996).

929
However, as shown in Figure 4 (Cumulative production ratio for real time for the
"cross" configuration.

Figure 4

From Retnanto and Economides, 1996), sustained incremental cumulative recovery


(over a single horizontal well) is maintained for considerable time in lower permeability
formations. Of course, the drilling of the multilaterals must be justified not just on the
basis of incremental production, but rather on incremental NPV. As drilling and
completion costs are reduced, these configurations are likely to be very desirable
accelerators of production and, thus, a form of stimulation.

Multibranched wells also can be justified by incremental early- and, to a lesser extent,
late-time production enhancements.

Figure 5 (Normalized productivity index for the Lx=Ly configuration (kx=ky).

930
Figure 5

From Retnanto et al., 1996) shows normalized productivity indexes for several
multibranched well configurations. At very early time, the incremental transient PI is
exactly proportional to the total exposed well length. At later time, the normalized PIs
are reduced as the reservoir at-large and outer boundaries are felt.

In the case of severe permeability anisotropy (with the mother hole normal to the
maximum horizontal permeability), the incremental benefits are somewhat reduced,
although they are still substantial ( Figure 6 , Normalized productivity index for the
Lx=Ly configuration (kx = 50 ky). From Retnanto et al., 1996).

931
Figure 6

Partially Completed Wells


A potentially very important new type of well configuration is the deliberate
undercompletion of a well.

For a variety of reasons, it may not be desirable to fully expose a well to the formation.
For example, it may be important to be able to isolate zones; or to easily conduct
remedial operations; or, in the case of slotted liners, to readily detect points at which
problems arise. Thus, we can envision the need for partial completions, as depicted in
Figure 1 (Horizontal well completion options).

932
Figure 1

These can include slotted and blank liners separated by external casing packers
(ECPs), or cemented, selectively perforated liner strings.

Retnanto et al. (1997) used the model of Economides et al. (1996), which is described
in the Section titled "A Production Prediction Model," and first, corroborated the results
published earlier by Goode and Wilkinson (1991). They extended these results to show
that for a well separated in 4 to 5 segments depending on the dimensionless
variable, hD (=hIani /L) the well performance is highly and disproportionately
favorable to the fraction of the well open to flow.

Figure 2 (Normalized productivity index vs.

933
Figure 2

number of open segments for hD = 0.01. From Retnanto et al., 1997) shows that in a
relatively thin reservoir (i.e., for Iani = 1 and L = 2,000 ft, h=20 ft), a well that is
distributed into 5 segments and is 40 percent open to flow would produce at 90
percent of the "ideal" open-hole rate that would be attained if the entire well was
exposed to the reservoir. In a reservoir five times as thick, the normalized productivity
index for the same 40 percent would still be a very favorable 77 percent ( Figure 3 ,
Normalized productivity index vs.

934
Figure 3

number of open segments for hD = 0.05. From Retnanto et al., 1997). Figure 4
(Normalized productivity index vs.

935
Figure 4

open fraction for 4 segments. From Retnanto et al., 1997) shows a generalized
normalized productivity index for a four-segment configuration for different reservoir
thicknesses. One of the main benefits of a partial completion is the relative ease of
stimulation, which means appropriate fluid distribution and reduced costs.

The partial completion concept suggests that a partially open but fully stimulated well
can outperform a fully open well that is even mildly damaged. This is because the
partial completion provides de facto zonal isolation that allows a much better
distribution of stimulation fluids. For example, Figure 5 (Comparison of fully completed
but damaged well with a partially completed (four segments) well with zero skin.

936
Figure 5

From Retnanto et al., 1997) shows a well that is only 40 percent open to flow actually
outperforming a well that is fully open but has a skin effect equal to 10.

In light of these findings, savings of stimulation and completion costs may lead to
deliberate undercompletion and an optimum open portion that is far below 100
percent. Figure 6 (The 3-year net present value (NPV) for the case study.

937
Figure 6

From Retnanto et al., 1997) shows a North Sea case study where the maximum NPV is
for a 30 percent open fraction.

Completions for Complex Well Architecture


As is true for all types of wells, multilateral configurations cover the entire spectrum of
completion options. Open hole completions are certainly the cheapest and simplest,
but they are the least flexible with respect to zonal isolation or future intervention.

Hard rocks that are not susceptible to wellbore collapse are candidates for open hole
completions. Such wells should often be viewed as disposable, intended for rapid early
production without regard to future problems. If the drilling costs are minimal or if
early production is large (e.g. in massively naturally fractured formations) such
completions may be attractive.

An open hole completion starts with a vertical mother hole that is drilled to a point
above the target. Laterals are then kicked off, in succession and from the bottom up,
through windows drilled in the parent hole. Because no casing is run in these

938
horizontal wells, they can be short radius and can target relatively small geological
flow units.

In the simplest multilateral configuration, the multilaterals are connected to the


mother hole without any provisions for shut-off or isolation among the holes. Figure 1
(Multilateral well completion with isolation. Courtesy of Halliburton Energy Services)
shows one way to isolate (i.e., shut off) a problematic hole.

Figure 1

First, a vertical parent hole is drilled, cased and cemented below the desired exit point (it
is also possible, as shown in the figure, to extend the bottom of the well into a lower
reservoir as a horizontal section).

Then, a retrievable whipstock is set at the appropriate spot, a window is milled


and the drilling assembly is diverted into the formation.

After this lateral is drilled, the whipstock is retrieved, thus allowing access to
the lower lateral (or vertical hole).

939
This process can be repeated with several laterals (again, starting from the bottom up). Completion
involves installing production packers above and below each exit window, and tubing strings to
allow comingled production. Access sleeves and sliding sleeves can be installed to allow thru-
tubing access for selective intervention or the shut off of the lateral, if problems emerge.

In cementing a lateral (or, at least, the curved part of the well, which has frequently
been the location of wellbore collapse), the first consideration is the lateral hole size.
Slim hole cementing has been discussed and even attempted in some cases. In
general, however, laterals to be cemented are of relatively larger diameters. A typical
configuration is a 7-in. lateral out of a 9-3/4 in. vertical parent.

As shown in Figure 2 (Cementing a liner in lateral hole.

Figure 2

Courtesy of Halliburton Energy Services) the liner casing string consists of a float, 150
to 300 feet of casing extended into the parent hole, and a liner hanger.

Once the casing string is cemented, it can then be perforated across the entire
formation interval, or across selected intervals ( Figure 3 , Cemented and perforated

940
lateral, courtesy of Halliburton Energy Services). Remedial cementing techniques can
be applied in case of future problems. As is the case with open hole completions,
several multilaterals can be installed.

Figure 3

Finally, slotted liners can be installed in a manner similar to that of the installation of
the cemented casing string.

A far more advanced well completion technology is the "Level 5" which allows full
hydraulic isolation of the junction between the vertical trunk and the lateral. This is
depicted in Figure 4 (Cased hole trunk with hydraulically isolated junction.

941
Figure 4

Courtesy of Baker Oil Tools), where a lateral entry nipple allows production isolation
from either lateral and thru-tubing re-entry capability in either lateral. While
production can be commingled, dual-string completions are also possible, as depicted
in Figure 4 . A dual packer is installed and, again, thru-tubing re-entry is possible in
either lateral.

The concept can be extended to a true multilateral application as shown in Figure 5


(Multilateral completion: lower junction is Level 4, mechanical support and comingled
production; upper junction is Level 5, hydraulic isolation, dual- string production.

942
Figure 5

Courtesy of Baker Oil Tools). In this illustration, the lower junction is "Level 4"
mechanical support while the upper junction is "Level 5", with complete hydraulic
isolation. The lower junction is commingled flow, while the upper junction is dual-
packer, dual-string, isolated production. A Level 6 completion is basically the same as
a Level 5, the difference being that hydraulic isolation is achieved at the casing and not
through the use of additional completion equipment.

Sand Exclusion
Although high-permeability hydraulic fracturing is one of the most effective sand
control techniques, sand exclusion is often accomplished with gravel packing and
screens.

Figure 6 (Open hole completion for sand exclusion.

943
Figure 6

Courtesy of Baker Oil Tools.) is a schematic of an open hole completion designed to


exclude sand. In this process, low-viscosity brine is used to carry gravel to pack
unconsolidated formations below the fracturing pressure. New erosion-resistant
screens are now installed, allowing considerable longevity to the well completion.

An interesting sand exclusion technique that combines both fracturing and gravel
packing principles has been applied to cased hole completions Figure 7 , Cased hole
completion combining fracturing and gravel pack techniques.

944
Figure 7

Courtesy of Baker Oil Tools). In this technique, low-viscosity brine is injected above
the fracturing pressure, creating a short fracture. Ideally, the perforation should be at
180o-phasing and along the fracture path. The perforations and the created fracture
are then filled with gravel, providing a very effective flow path while largely eliminating
sand production.

945
Nomenclature

A = area, ft2 [m2]

Af = fracture area, ft2 [m2]

a = large half-axis of drainage ellipse formed by a horizontal well, ft [m]

aiH,max = maximum horizontal half-axes of stimulation ellipse, ft [m]

asH,max = maximum horizontal half-axis of damage ellipse, ft [m]

B = formation volume factor, Bbl/STB, CF/SCF; resm3/m3

b = constant relating to volumetric coverage of stimulation fluid

C1 = wellhead price of oil per unit volume

C2 = injection costs per unit time

C3 = perforation costs per unit length of interval

C3 = cost of stimulation fluids per unit volume

Cfix = fixed costs

CL = leakoff coefficient, ft/(min)1/2, m/(s)1/2

Cm = concentration of mineral, mole/liter

Cp = proppant concentration in fracture, lb/ft2 [kg/m2]

Cr = concentration of reactant in liquid phase, mole/liter

ct = total system compressibility, psi-1 [kPa-1]

D = diffusivity constant

dc = casing outside diameter, in. [mm]

df = fractal dimension (normally 1.65)

dp = particle diameter, in. [mm]

FCD = dimensionless fracture conductivity

G = elastic shear modulus, psi [kPa]

g = acceleration of gravity, ft/s2 [m/s2]

GLR = gas-liquid ratio, SCF/STB [m3/m3]

h = formation thickness, ft [m]

hf = fracture height, ft [m]

946
i = discount rate

Iani = index of horizontal-to-vertical permeability anisotropy

k = reservoir permeability, md

K' = consistency coefficient in power law fluids, lbf-secn'/ft2

KL = multiplier to fluid loss coefficient

kf = fracture permeability

kH = effective horizontal permeability, md

kHmax = maximum horizontal permeability, md

kHmin = minimum horizontal permeability, md

ks = damaged zone permeability, md

kv = vertical permeability, md

L = length of horizontal section, ft [m]

Lperf,k = perforated length of segment, ft [m]

= viscosity, cp [Pa-s]

0 = zero shear viscosity, cp [Pa-s]

Mp = total mass of proppant, lbm [kg]

NAC = acid capacity number

NDa = Damkohler number (ratio of reaction rate to convection rate)

NPe = Peclet number describing the ratio of convection and diffusion velocity.

NPV = Net Present Value

Np,n = difference in incremental production with and without stimulation in the nth year,
STB [m3]

Np,n = difference between the sum of the incremental cumulative productions of a


stimulated well's perforated segments and the cumulative production of the fully
open and unstimulated (damaged) well in year n, STB [m3]

n' = power law exponent

p = pore pressure, psi [kPa]

pbd = breakdown pressure, psi [kPa]

PD = dimensionless pressure

pe = outer boundary pressure, psi [kPa]

947
pe = reservoir pressure at outer flow boundary, psi [kPa]

pi = initial reservoir pressure, psi

PIH = Productivity Index, horizontal well, B/D/psi [m3/D/kPa]

pinj = injection pressure, psi [kPa]

pwf = flowing bottomhole pressure, psi [kPa]

p = differential pressure, psi [kPa]

q = flow rate, STB/D [m3/D] for oil, MSCF/D [m3/D] for gas

qi = injection rate

rwD = dimensionless effective wellbore radius

rc = outside radius of casing, in. [mm]

re = radius of drainage area, ft [m]

reH = drainage radius of horizontal wellbore, ft [m]

ri,j = radius of acid penetration for segment j

ri,max = maximum radius of stimulated area, ft

rp = ratio of permeability to fracture height

rw = damage penetration

rw = wellbore radius, ft [m]

seq = equivalent skin effect

sc = choke skin effect

seq = skin effect

sV = skin effect for a vertical well

T = temperature, R [K]

t = time, hr

T0 = tensile stress of rock, psi [kPa]

tDxf = dimensionless time

ti = total injection time, h

ut = single particle terminal settling velocity for power-law fluids

Vacid, j = volume of stimulation fluid per unit length opf interval, gal/ft

948
Vbulk = bulk volume

vCTj = coiled tubing withdrawal rate, ft/min

Vpad = pad volume

v = Poisson ratio

w = minimum clearance between borehole and outside wall of casing, in. [mm]

w, wp = propped fracture width, in. [mm]

= average fracture width, in. [mm]

x = distance from vertical section, ft [m]

xf = fracture half-length, ft [m]

xj/L = relative position of coiled tubing

Z = gas deviation factor

= relative fracture capacity

= stoichiometric coefficient for a mineral reacting with acid

= geometric factor for fracture shape

x/1 = position of stimulation fluid injection front normal to the wellbore at distance

= efficiency (ratio of created fracture volume to total injected volume)

= build angle, degrees (radians)

= viscosity, cp [Pa-s]

p = density of proppant pack, lb/ft3 [kg/m3]

= density difference between particle and carrier fluid, lb/ft3 [kg/m3]

b = bending stress, psi [kPa] dc

H,max = maximum horizontal stress, psi [kPa]

H,min = minimum horizontal stress, psi [kPa]

v = vertical stress, psi [kPa]

= porosity

p = porosity of fracture proppant pack

949
SI METRIC CONVERSION FACTORS
FOR COMMON FIELD UNITS

(From SPE Metric Standard (1984))

Quantity & Traditional Conversion SI


dimensions "Field" Unit factor metric
unit

Length (L) foot (ft) x 3.048 x 10-1 = meter (m)

inch (in) x 2.54 x 10-2 =m

Area (L2) square foot x 9.290 304 x 10-2 = sq. meter


(ft2) (m2)

Volume (L3) cubic foot x 2.831 685 x 10-2 = cu. meter


(ft3) (m3)

gallon x 3.785 412 x 10-3 = m3

barrel (bbl) x 1.589 873 x 10-1 = m3

Mass (M) pound-mass x 4.535 924 x 10-1 = kilogram


(lbm) (kg)

Density lbm/gal x 1.198 264 x 102 = kg/m3


(M/L3)

lbm/ft3 x 1.601 846 x 101 = kg/m3

Speed foot/second x 3.048 x 10-1 = m/s


(L/T) (ft/s)

ft/min x 5.08 x 10-3 = m/s

Force pound-force x 4.448 222 x 100 = Newton


(ML/T2) (lbf) (N)

Pressure lbf/in2 (psi) x 6.894 757 x 103 = N/m2 (Pa)


(M/LT2)

psi x 6.894 757 x 100 = kPa

Viscosity centipoise x1 x 10-3 = Pa-s

950
(M/LT) (cp)

951

Vous aimerez peut-être aussi