Vous êtes sur la page 1sur 0

SPE 160855

Comparisons and Contrasts of Shale Gas and Tight Gas Developments,


North American Experience and Trends
Robert L. Kennedy, SPE, William N. Knecht, SPE, and Daniel T. Georgi, SPE, Baker Hughes
Copyright 2012, Society of Petroleum Engineers

This paper was prepared for presentation at the SPE Saudi Arabia Section Technical Symposium and Exhibition held in Al-Khobar, Saudi Arabia, 811 April 2012.

This paper was selected for presentation by an SPE program committee following review of information contained in an abstract submitted by the author(s). Contents of the paper have not been
reviewed by the Society of Petroleum Engineers and are subject to correction by the author(s). The material, as presented, does not necessarily reflect any position of the Society of Petroleum
Engineers, its officers, or members. Papers presented at the SPE meetings are subject to publication review by Editorial Committee of Society of Petroleum Engineers. Electronic reproduction,
distribution, or storage of any part of this paper without the written consent of the Society of Petroleum Engineers is prohibited. Permission to reproduce in print is restricted to an abstract of not
more than 300 words; illustrations may not be copied. The abstract must contain conspicuous acknowledgment of where and whom the paper was presented. Write Librarian, SPE, P.O. Box
833836, Richardson, TX 75083-3836, U.S.A., fax 01-972-952-9435.


Abstract
All Shale Gas reservoirs are not the same. There are no typical Tight Gas reservoirs. These two statements can be
found numerous times in the literature on shale gas and tight gas reservoirs. The one common aspect of developing these
unconventional resources is that wells in both must be hydraulically fractured in order to produce commercial amounts of
gas. Operator challenges and objectives to be accomplished during each phase of the Asset Life Cycle (Exploration,
Appraisal, Development, Production, and Rejuvenation) of both shale gas and tight gas are similar. Drilling, well design,
completion methods and hydraulic fracturing are somewhat similar; but formation evaluation, reservoir analysis, and some of
the production techniques are quite different.

Much of the experience in shale and tight gas has been developed in the US and in Canada, to a lesser extent; and
most of the technologies that have been developed by operators and service companies are transferable to other parts of the
world. However, the infrastructure, including equipment and service company availability, governmental regulations,
logistics, processing, environmental considerations, and pricing are not the same as in the US. This may impact the rate of the
technology transfer as well as the selection of some of the technology. This paper is focused on operations challenges,
technologies, and experience associated with shale and tight gas projects. It is likely that environmental concerns and the drive
to reduce development costs of tight and shale gas reservoirs will drive new approaches to the development of these reservoirs
in China, Latin America, Middle East, North Africa, and other parts of the world.

Introduction
Unconventional shale and tight gas development in the US was sparked by the 1980 introduction of The Alternative
Fuel Production Credit of the Internal Revenue Code (an income tax credit). The 1980 WPT (windfall profit tax) included a
$3.00 (in 1979 dollars) production tax credit to stimulate the supply of selected unconventional fuels: oil from shale or tar
sands, gas produced from geo-pressurized brine, Devonian shale, tight formations, or coalbed methane, gas from biomass, and
synthetic fuels from coal. In current dollars this credit, which is still in effect for certain types of fuels, was $6.56 per barrel of
liquid fuels and about $1.16 per thousand cubic feet (mcf) of gas in 2004 (Lazzari 2006). Initially, the credit was set to run
until 1989; however, it was extended twice until the end of 1992 (Martin and Eid 2011).

Higher gas price was another reason for the continued development of tight gas and especially shale gas. Figure 1
shows Henry Hub spot prices from 2000 until J anuary of 2012. The spot price represents the price for natural gas sales
contracted for next day or weekend delivery and transfer at a given trading location. Henry Hub is the primary trading
location, centralized point, for natural gas trading in the United States, and is often a representative measure for wellhead
prices. Higher prices are reflected by the six year (2003-2009) run of gas prices over $6 per MMBtu after generally hovering
around $2 per MMBtu for the prior twenty-year period, 1980 to 2000. During this time, two significant peaks in gas prices
occurred. In the summer of 2005, hurricanes along the U.S. Gulf Coast caused more than 800 billion cubic feet (Bcf) of
natural gas production to be shut in between August 2005 and June 2006. As a result of these disruptions, natural gas spot
prices at times exceeded $15 per million Btu (MMBtu) in many spot market locations and fluctuated significantly over the
subsequent months, reflecting the uncertainty over supplies (Mastrangelo 2007). In 2008, due to physical and financial market
factors, spot prices broke from the $6-$8 per MMBtu range of the two previous years and peaked at $13.32 per MMBtu, but
ended the year at $5.63 per MMBtu. This was the beginning of the current fall in gas prices.
2 SPE 160855

USD / MMBtu

Figure 1Henry Hub spot prices for natural gas in the US www.tradingeconomics.com l NYMEX

The more recent growth in natural gas production from unconventional tight and especially shale reservoirs is a result
of technological advances (hydraulic fracturing and horizontal wells). Drilling in shale plays with high concentrations of
natural gas liquids and crude oil has shifted from drilling in dry gas plays. Figure 2 shows that by the end of 2010 shale gas
and tight gas comprised 23 and 26 percent, respectively, of the total US natural gas production. In the future (by 2035) tight
gas will decline slightly to 21 percent of the total, while shale gas will continue to increase to a level of 49 percent of the total
US natural gas production (Nitze and Gruenspecht 2012).


Figure 2US natural gas production, 1990-2035, EIA Annual Energy
Outlook 2012 Early Release (2012)

Beginning with a discussion of the location and size of shale and tight gas resources throughout the world, this paper
describes the characteristics of shale and tight gas showing how the two are different in a number of respects. The two
resources are different from play to play and in how the gas is evaluated, developed and produced. Shale and tight gas are
contrasted and compared as the operator challenges and objectives are enumerated at each phase of the Asset Life Cycle
(Exploration, Appraisal, Development, Production, and Rejuvenation). The single binding commonality is that both shale and
tight gas wells must be hydraulically fractured in order to produce commercial amounts of gas. Essentially all of the industry
experience with shale gas has been obtained and new technology developed in North America. Although the same is true for
tight gas, it is now being developed in other parts of the world, specifically the Middle East, North Africa, Argentina, and
Australia. This paper primarily documents the North American (with a focus on the US) experience and trends.

Shale and Tight Gas Resources Worldwide
Until April of 2011 when the US EIA published a study by Kuuskraa et al on World Shale Gas Resources the
industry has relied on 1996 data from Rogner for estimates of worldwide unconventional gas resources, gas in place (GIP). In
2001, Kawata and Fujita called Rogners gas resources rather optimistic assessment; however, the comment refers to the
numbers for Methane hydrate (which was also included in the estimates) and the ratio of unconventional gas (tight, shale, and
CBM) to the then assessed level of conventional gas resources. Table 1, which was published by Holditch (2006) shows the
distribution of worldwide unconventional gas resources, including CBM, Shale Gas, and Tight Gas. It should be noted that the
numbers are GIP and are the same as Rogners 1996 numbers.

The US EIA (Kuuskraa et al 2011) published the most recent estimates for worldwide unconventional gas resources
(albeit for shale gas only) since the Rogner estimates of 1996. However, Russia and Central Asia, Middle East, South East
Asia, and Central Africa were not addressed by the current report. This was primarily because there was either significant
quantities of conventional natural gas reserves noted to be in place (i.e., Russia and the Middle East), or because of a general
lack of information to carry out even an initial assessment (Kuuskraa et al 2011). Table 2 was developed by these authors in
order to provide a more complete comparison for worldwide total resources by region.
SPE 160855 3


Table 1Worldwide Unconventional Gas Resources, Table 2Worlwide Unconventional GIP Resources
GIP (Holditch, 2006 from Kawata, 2001, modified and adjusted from Kuuskraa et al
from Rogner, 1996) (2011) Study for EIA.

World Shale Gas Resources: An Initial Assessment of 14 Regions Outside the United States, the 2011 EIA study
referenced above, went beyond developing estimates of only GIP resources and looked at recovery by providing estimates of
Technically Recoverable Resources for individual countries within the regions of the world. The consultants approach relied
upon publically available data from technical literature and studies on each of the selected international shale gas basins to first
provide an estimate of the risked gas in place, and then to estimate the technically recoverable resource for that region. This
methodology is intended to make the best use of sometimes scant data in order to perform initial assessments of this type
(Kuuskraa et al 2011). Figure 3 shows the Technically Recoverable Resources (TRR) for the top countries of those evaluated
by the 2011 EIA Study. The total TRR of the 32 countries evaluated in the report is 6,622 Tcf. Although the Middle East was
not included in the study, Saudi Arabia is shown as fifth largest with 645 Tcf. These authors began with GIP numbers from
Rogner, added a recovery factor and increased the total by the average percentage of all other countries.


Figure 3Shale Gas Technically Recoverable Resourses in trillion cubic feet (Tcf); top countries from Kuuskraa
et al (2011) Study for EIA, total of 6,622 Tcf for 32 countries Saudi Arabia resource number added by authors
1275
862
774
681
645
485
388
0
200
400
600
800
1000
1200
1400
T
R
R
,

T
c
f
4 SPE 160855
As noted, the Kruuskraa et al (2011) EIA study strictly focused on shale gas, excluding other unconventional gas, tight
gas and CBM, as well as shale oil. From Table 2 it can be seen that shale gas GIP is approximately six times greater than tight
gas. If a more detailed evaluation were conducted for tight gas, such as was done for shale gas, these authors expect that tight
gas resources would also increase, primarily due to advances in technology.

Figure 4 is a map showing the location of the 23 significant shale gas basins in the US. The six currently most active
gas basins/plays are the Barnett, Woodford, Fayetteville, Haynesville, Marcellus, and Eagle Ford. Other active US shale
basins/plays are the Bakken, primarily a tight oil and not a true shale; the emerging Niobrara, primarily shale oil, and the
liquids-rich Utica. Figure 5 is a map showing the location of the 14 significant tight gas basins in the US. Four of these
basins, Pinedale Anticline, Anadarko, Piceance, and Deep Bossier, produce most of the US tight gas (Warlick 2010).


Figure 4US map showing the 23 significant shale gas Figure 5US map showing 14 significant tight gas basins.
Basins. Currently most active are Barnett, The four basins in green produce most of the US
Woodford, Fayetteville, Haynesville, tight gas (Warlick International 2010).
Marcellus, and Eagle Ford.

Figure 6 is a map of Canada which shows the five significant shale gas basins/plays: Montney, Horn River, and
Colorado Group in Western Canada and the Utica and Horton Bluff Group in Eastern Canada. Another is the emerging
Duvernay in west-central Alberta (NEB-Canada, 2009). The three significant typical Canadian tight gas plays/basins are the
Shallow, J ean Marie, and Deep Basin (Dixon 2005). The Bakken tight oil play is also located in Canada.



Figure 6Map of Canada showing the five significant shale gas basins, Montney, Horn River, Colorado
Group, Utica, and Horton Bluff; and the three significant tight gas basins/plays, Shallow, Jean Marie,
and Deep Basin (Source: Modified from NEB Canada, 2009).

Total shale gas TRR from Figure 3 for the US and Canada are 862 Tcf and 388 Tcf, respectively. Currently the EIA
does not report reserve estimates for tight gas; as these are included with conventional gas (EIA 2010). Estimates of the
volume of recoverable gas in tight reservoirs in the U.S. range from 200 to 550 Tcf (Oil and Gas Investor, 2006). Meanwhile,
Warlick (2010) states that the total Reserves, Dec 2009 for the four leading US tight gas basins is equal to 92 Tcf (high of
ranges). There is also similar uncertainty with respect to tight gas TRR or reserves in Canada. The NEB even hedges on the
definition of tight gas; and does not offer any estimates of tight gas reserves or TRR. Therefore, for tight gas, we are left only
with Rogners GIP total for North America of 1,371 Tcf (Table 2). Assuming an increase for new technology and a
SPE 160855 5
conservative recovery factor of 20 percent, these authors estimate North Americas TRR tight gas to be approximately 430
Tcf.


The Unconventional Gas Resource Triangle
The concept of the resource triangle was used by Masters (1979) to find a large gas field and build a company in the
1970s. (Holditch 2006) Figure 7 illustrates the principle of the resource triangle. Conventional gas is located at the top of
the triangle with better reservoir characteristics/quality, uses conventional technology, easy to develop; but exists in small
volumes. As you go deeper down in the triangle passing tight gas and coalbed methane (CBM), shale gas (and gas hydrates)
are found at the bottom. Progression to the bottom of the triangle sees permeability and reservoir quality decreasing, level of
technology to develop increasing (becoming more complex), and difficulty of development increasing; however large volumes
of these resources can be found. The concept of the resource triangle applies to every hydrocarbon producing basin in the
world. Martin et al (2008) validated the resource triangle concept using a computer program, database and software they
developed. They also expanded the resource triangle to include liquid and solid hydrocarbons adding heavy oil and oil shale,
and referencing work from Gray (1977).




Figure 7The Unconventional Gas Resource Triangle The concept of the resource triangle applies
to every hydrocarbon producing basin in the world (Holditch from Masters, 1979).


Characteristics of Shale and Shale Gas
Shale gas is an unconventional gas reservoir contained in fine-grained, organic rich, sedimentary rocks, including
shale, but composed of mud containing other minerals like quartz and calcite (U.S. DOE 2009: Warlick 2010; U.S. EIA 2011).
A number of formations broadly referred to by the industry as shale, may contain very little shale lithology/mineralogy, but is
shale by grain size only. Passey et al (2010) describes shale as extremely fine-grained particles typically less than 4
microns in diameter, but may contain variable amounts of silt-sized particles (up to 62.5 microns). No two shales are alike;
they vary aerially and vertically within a trend and even along horizontal lateral wellbores (King 2010). Not only will shales
vary from basin to basin, but also within the same field (Economides and Martin 2007). These reservoirs are continuous gas
accumulations, and persist over very large geographic areas. The challenge in shale is not to find the gas, but to find the best
areas, or sweet spots, that will result in the best production and recovery (J enkins and Boyer 2008).

Shale reservoirs have no trap like conventional gas reservoirs, and do not contain a gas/water contact. They are the
source rock, which also now acts as the reservoir, where the total or partial volume of gas (hydrocarbon) remains. Shales have
been the source rock for much of the hydrocarbons in North America. In fact, these same shale source rocks are now being
exploited as shale reservoirs. The key is to find a shale play where the remaining hydrocarbon, that was not expulsed and
migrated into conventional formations, is now economically viable for development. Take caution Not all shales are source
rocks.

Natural matrix permeability of shales is extremely low, often in the nano Darcy range. It has been said that
measurement of shale permeability is difficult, and results are probably inaccurate. In this very low permeability environment,
gas (hydrocarbon) flow through the matrix is extremely limited and insufficient for commercial production. Various authors
have estimated that a gas molecule will move no more than 10 to 50 feet per year through shale matrix rock. Shale porosities
are also relatively low ranging from 6-12%. Therefore, shale reservoirs require hydraulic fracturing/induced fractures in order
to produce commercial amounts of gas.

A number of different reservoir parameters, that are not necessarily deemed important for conventional gas, are
significant for assessing economic viability, development, and well completion techniques for shale. Total Organic Carbon
(TOC) content, Kerogen type, Thermal Maturity, Mineralogy/Lithology, Brittleness, existence of Natural Fractures, Stress
6 SPE 160855
Regime, multiple locations and types of gas storage in the reservoir, characteristic production decline profile, Thermogenic or
Biogenic systems, as well as depositional environment, thickness, porosity and pressure are parameters than we must now
consider for shale reservoirs. The following discussion will briefly cover each of these parameters and the unique aspects of
shale to provide the reader with an understanding of their significance in play analysis and development.

Organic materials, microorganism fossils, and plant matter provide the required carbon, oxygen, and hydrogen atoms
needed to create natural gas and oil. TOC is the amount of material available to convert into hydrocarbons (depending on
kerogen type) and represents a qualitative measure of source rock potential (J arvie et al 2007). TOC is expressed as a percent
by weight; and is sometimes expressed as percent by volume (Volume % is approximately twice that of Weight %). Oil and
gas source rocks typically have greater than 1.0% TOC. TOC richness can range from Poor - <1%; Fair - 1-2%; and Good-
Excellent - 2-10% (PESGB 2008). TOC is not the same as kerogen content, as TOC is made up of both kerogen and bitumen.
Measurement of TOC in shale is determined from wireline logs and by direct measurement from cores and drill cuttings.

Kerogen is a solid mixture of organic chemical compounds that makes up a portion of the organic matter in
sedimentary rocks. It is insoluble in normal organic solvents because of the huge molecular weight of its component
compounds. The soluble portion of kerogen is known as bitumen (Wikipedia 2011). Understanding the kerogen type helps to
predict the hydrocarbon type (gas or liquid) in a play. There are basically four types of kerogen, three of which can generate
hydrocarbons. Type I generates oil, Type II wet gas, and Type III dry gas (Holditch 2011). Table 3 shows the types of
kerogen and their hydrocarbon generating potential.

Thermal Maturity measures the degree to which a formation has been exposed to high heat needed to breakdown
organic matter in hydrocarbons. As temperature increases with the increasing depth in the earths crust, the heat causes the
generation of hydrocarbons and can ultimately destroy them. Typical temperature ranges at which oil and gas are generated
are shown on Figure 8. The oil window is 60-175 C (140-350 F) and the gas window is 100-300 C (212-570 F). The
position of oil and gas windows within a basin is dependent on the type of organic matter and heating rate. Thermal maturity
is a function of both time and temperature (Holditch, 2011). Understanding the level of thermal maturity, or indeed whether
the shale is thermally mature at all, is key to understanding shale resource potential (PESGB 2008). Also, higher thermal
maturity leads to the presence of nanopores, which contribute to additional porosity in the shale matrix rock (Kuuskraa et al
2011).

Vitinite Reflectance, R
o
%, the most commonly used technique for source rock thermal maturity determination,
measures the intensity of the reflected light from polished vitrinite particles (a maceral group formed by lignified, higher-land
plant tissues, such as leaves, stems, and roots) in shale under a reflecting microscope. Increased reflectance is caused by
aromatization of kerogen and loss of hydrogen (J arvie et al 2007). Figure 9 is the Thermal Maturation Scale. Dry gas occurs
when R
o
is greater than 1.0%, wet gas when R
o
is between 0.5 and 1.0%, and oil when R
o
is between 0.5 and 1.3% (Kuuskraa
et al 2011).


Table 3Types of Kerogen and Their HC Figure 8Typical temperature Figure 9Themal Maturation
Potential (Holditch 2011) ranges which oil & gas are Scale (Kuuskra et al
generated (Holditch 2011) 2011)


SPE 160855 7
Mineralogy and lithology are important for 1. TOC quantification, 2. Reducing porosity uncertainty, 3. Identifying
shale lithofacies, 4. Indicating variations in mechanical rock properties including brittleness, and 5. Assisting in the planning
of well hydraulic fracturing and completion designs. Most shale reservoirs can be chemostratigraphically classified into three
primary lithofacies siliceous mudstone (such as the Barnett), calcareous mudstone lithofacies, and organic mudstone
lithofacies. Additional lithofacies have been identified in some reservoirs based on their unique characteristics.
Lithology/Mineralogy information is obtained from conventional and pulsed neuron log responses, laboratory analysis of cores
and cuttings, and mineral spectroscopy analyses. TOC is quantified by the amount and vertical distribution of kerogen,
kerogen type, level of maturity, and mineral spectroscopy plus core analysis. Log derived and computed geomechanical
properties include minimum horizontal stress (S
H, Min
), Poissons Ratio, Youngs Modulus, Fracture Migration, and static
mechanical properties. Brittleness indicators (for identifying best interval to initiate a fracture and location at the vertical from
which to drill horizontal laterals) are computed from mineralogy and geomechanical brittleness and hardness (J acobi et al
2009; LeCompte et al 2009; Pemper et al 2009; Mitra et al 2010).

Geomechanics is central to the development of shale gas resources. The stress regime in a basin must be considered
during well drilling, fracturing, and production. Well orientation is dictated by in-situ stress systems and wellbore stability
during drilling. In general, initiating a fracture depends on the stresses around the wellbore both from the geologic produced
tectonic effects and from changes in stresses produced by the growth of fractures. Fractures are difficult to initiate where total
rock stresses are very high. A major consideration during shale production is the stress evolution accompanying drawndown
and depletion. It is now well established that reservoir pressure changes have an effect on both the stress magnitudes and
direction in the sub-surface (Addis and Yassir 2010; King 2010).

Presence, location, and orientation of natural fractures in shales are significant with respect to the hydraulic fracturing
process. In these naturally fractured reservoirs the well placement for initial development is dictated by two sub-surface
factors:
a. location and orientation of the natural fracture sets, the orientation of the most conductive natural fracture set,
and the in-situ stress magnitudes and directions
b. the propagation direction of the hydraulic fractures from the wellbore and the intersection of the natural fracture
system (Addis and Yassir 2010)
One of the purposes of hydraulic fracturing is to connect the existing natural fractures, intersecting them in a near
perpendicular or transverse manner to create a complex network of pathways to enable hydrocarbons to enter the wellbore
(King 2010; J enkins and Boyer 2008).

Gas is stored in three ways in a shale reservoir:
1. Free Gas
a. In the rock matrix porosity
b. In the natural fractures
2. Sorbed Gas
a. Adsorbed (chemically bound) to the organic matter (kerogen) and mineral surfaces within the natural
fractures
b. Absorbed (physically bound) to the organic matter (kerogen) and mineral surfaces within the matrix rock
3. Dissolved
In the hydrocarbon liquids present in the bitumen

To obtain the total amount gas in place (GIP), free gas, sorbed gas, and dissolved gas must be added together. Free gas is the
initial flush production that occurs early, during the first few years of the life of a well. The absorbed gas volume is often
significantly more than the free gas stored in the matrix porosity itself. Gas contents can exceed apparent free gas-filled
porosity by 6 to 8 times where organic content is high (Warlick 2010). However; sorbed gas is produced by
diffusion/desorption and does not occur until later in the field life after the reservoir pressure has declined. It is generally
accepted that sorbed gas does not have an appreciable effect on shale field economics.

Most shale gas wells produce only dry gas (90% methane) and essentially no water. A notable exception to this is the
Eagle Ford with part of the play producing dry gas, part wet gas, and another part producing shale oil. The Antrim and New
Albany shales, which are minor and somewhat inactive, do produce formation water (this is discussed later). Water production
that causes concerns (handling, treating, re-use or disposal) for shale gas wells is frac flowback water.

Shale gas wells (and hydraulically fractured tight sands) display a rather unique decline profile character. Shale and
tight gas wells typically exhibit gas storage and flow characteristics uniquely tied to geology and physics (Rushing et al 2008).
IP (Initial Productivity) rates are relatively low, in the 2-10
+
MMcfd range (horizontal wells), and these rates decline rather
rapidly. During the first year, the rates can decline by 65-80
+
%; while the second year the decline is 35-45%; and the third
year decline is around 20-30%. After that, production levels out to about 5% decline per year. This flat production or the tail,
8 SPE 160855
as it has been called, could last for 25 to 30 years (Nome and Johnston 2008; U.S. EIA 2011). However; it seems that a well
producing, say less than 100 Mcfd would be approaching the economic limit. Some examples of typical decline type curves
are shown as Figure 10 Barnett shale; Figure 11 Haynesville shale; Figure 12 Eagle Ford shale; and Figure 13 Fayetteville
shale. These type curves were taken from U.S. EIA (2011).



Figure 10Barnett shale type curve and other Figure 11Haynesville shale type curves of decline and
information (U.S. EIA 2011) cumulative production (U.S. EIA 2011)



Figure 12Eagle Ford shale type curves of decline Figure 13Fayetteville shale type curves of decline and
and cumulative production (U.S. EIA 2011) cumulative production (U.S. EIA 2011)

Shale natural gas is either biogenic in origin, formed by the action of biologic organisms breaking down organic
material within the shale, or of thermogenic origin formed at depth and high temperatures. Relatively few biogenic gas
systems are producing economic gas within the United States. The Antrim shale in Michigan is one of those systems.
Another is the New Albany shale of Illinois and Indiana. Wells producing from the biogenic Antrim and New Albany shales
have relatively low production rates, e.g., 135 Mcfd; however they will produce for a long time, 20
+
years. In many cases
large quantities of water are produced with or before any gas is produced. Gas production is closely tied to dewatering the
system (like CBM) to gain economic production. Geochemistry analysis indicates that the water is usually fairly fresh.

The majority of producing shale gas reservoirs in the US are thermogenic systems. Thermogenic gas occurs as a
result of primary thermal cracking of the organic matter into a gaseous phase. Secondary thermal cracking of remaining
liquids also occurs. Thermal maturity (which has been previously discussed) in these reservoirs determines the type of
hydrocarbon that will be generated. The gas produced in a thermogenic environment will be relatively dry, previously
mentioned (Economides and Martin 2007).

Reservoir pressure is the key parametrer to how conventional gas (and oil) reservoirs perform. Pressure controls
production rates and is used to predict recovery. Most shale reservoirs range from normally pressured, to slightly
overpressured, to highly overpressured. The higher pressured shale reservoirs, like the Haynesville, have higher IPs and
higher recovery than others (see Figure 11). Higher reservoir pressures do have an effect on the hydraulic fracturing designs;
especially selection of appropriate proppants, as higher reservoir pressure can crush some types of proppants.
SPE 160855 9
Depositional environment of shales is important, particularly whether it is marine or non-marine. Marine-deposited
shales tend to have lower clay content and be high in brittle materials, such as quartz, feldspar and carbonates. Because of this
mineralogy, they respond favorably to hydraulic fracturing. Non-marine deposited shale, i.e., lacustrine and fluvial, tend to be
higher in clay, more ductile, and less responsive to hydraulic fracturing. Transgressive systems are characterized by higher
TOC and quartz and less clay. Shales deposited during transgressive systems not only respond favorably to hydraulic
fracturing, but also have higher hydrocarbon recoveries. Regressive systems are characterized by lower TOC and quartz and
higher clay content. Shales deposited during this time are less responsive to hydraulic fracturing and have lower hydrocarbon
recoveries. Depositional environment for shales can be even more important than thickness.


North American Shale Gas Basins and Statistics Comparison
The shale gas story in the US is placed in front of the public almost daily. Based on the observations of these authors
and available statistics from a number of sources, the following conclusions are drawn and information offered:
All shale reservoirs are not the same.
Shale wells must be fracture stimulated to produce commercially.
The two key elements of shale gas development in the US are:
1. Multi-stage hydraulic fracturing
2. Horizontal wells
These two elements together maximize the reservoir volume that is connected to the well, w/optimum lateral length.
The effectiveness (design, placement, implementation, flowback) of hydraulic fracturing has a significant effect on
production rates, drainage area, and recovery (of course, reservoir characteristics are significant factors).
Vertical wells are required to gather data.
Horizontal wells with lateral lengths ranging from 3,000 to 6,500 feet are used for development.
Average well spacing is approximately 80
+
acres.
Formation thickness ranges from 20 to 600 feet.
Formation depth ranges from 6,000 to 13,500 feet.
Well IPs range from 2 to 10
+
MMcfd.
Primarily dry gas production (90% Methane) exception is the Eagle Ford which also produces wet gas.
Most shales do not produce significant amounts of water.
Wells exhibit high decline rates in first few years on production.
A high number of wells are required to develop shale (low per well EURs).

The six major shale gas plays in the US are the Barnett, Fayetteville, Woodford, Haynesville, Marcellus, and the
Eagle Ford (the two major shale gas plays in Canada are the Horn River and the Montney). Table 4 is a comparison of the six
major US shale plays with respect to some of the physical aspects of the play and the individual wells. Most of the
information included in the table is from the recent U.S. EIA (2011). The well cost data were obtained from various sources.



Table 4Comparison of the six major US shale plays, physical aspects of the plays and wells (U.S. EIA 2011)

The next 12 figures are plots of production history, numbers of producing wells, and numbers of rigs operating in
each of these six major US shale plays. There are two plots for each play. The plot on the left includes the historical
producing well counts and number of rigs operating in the play; rig counts are shown beginning in 2009. Historical
production, gas, oil (if applicable), water, and condensate (if applicable) are shown on the plots on the right. The production
and producing well counts data are from the Drill Information Database, and the operating rig data are from Baker Hughes. It
should be noted that all plots begin with the year 2005, the Barnett, discovered in 1981, certainly has more historical data. The
Marcellus and Woodford also have just a couple more years of data. The sharp, abrupt, drop in producing wells in the
10 SPE 160855
Marcellus is due to incomplete data; a result of some states not reporting information on a timely basis. Drops in gas-directed
rigs for 2010-2011 can be seen, most notably, in the Barnett, Fayetteville, and Haynesville plays.


Figure 14Barnett Shale producing well Figure 15Barnett Shale production
and rig counts


Figure 16Fayetteville Shale producing well Figure 17Fayetteville Shale production
and rig counts


Figure 18Woodford Shale producing well Figure 19Woodford Shale production
and rig counts

SPE 160855 11


Figure 20Eagle Ford Shale producing well Figure 21Eagle Ford Shale production
and rig counts




Figure 22Haynesville Shale producing well Figure 23Haynesville Shale production
and rig counts




Figure 24Marcellus Shale producing well Figure 25Marcellus Shale production
and rig counts

12 SPE 160855
Canadian shale gas has been somewhat slower in developing than the US. In the Horn River shale, Canadas largest,
only 55 wells were drilled in 2008-2009, and an estimated 210 more wells in 2010-2011. The B.C. Ministry of Energy and
Mines National Energy Board (2011) estimated the Horn River Shale to contain up to 96 Tcf TRR. The earlier comparable
number for 2009 was approximately 120 Tcf (20% recovery of 600 Tcf, see Table 5). Note the 10% CO
2
content of the Horn
river shale gas. This gas is also reported to contain 0.01% H
2
S (Reynolds et al 2010).



Table 5Comparison of the five significant Canadian Shale Basins (NEB 2009)


Characteristics of Tight Gas
In the 1970s the U.S. government decided that the definition of a tight gas reservoir is one in which the expected
value of permeability to gas flow would be less than 0.1 md. This definition was a political definition that has been used to
determine which wells would receive federal and/or state tax credits for producing gas from tight reservoirs (Holditch 2006).
Holditch goes on to say that the tight gas definition is a function of a number of physical and economic factors. The best
definition of a tight gas reservoir is a reservoir that cannot be produced at economic flow rates nor recover economic volumes
of natural gas unless the well is stimulated by a large hydraulic fracture treatment or produced by the use of a horizontal
wellbore or multilateral wellbores (Holditch 2006; Shrivastava and Lawatia 2011). Other authors say that perhaps flowrate
rather than permeability should be the measure of what is termed a tight gas reservoir. Certainly that has merit, as some
reservoirs in countries outside of the US with 10
+
md permeability are being fractured and increasing flowrates.

There are no typical tight gas reservoirs. They can be (Holditch 2006):
Deep or shallow
High-pressure or low-pressure
High-temperature or low-temperature
Blanket or lenticular
Traps are usually stratigraphic
Homogeneous or naturally fractured
Single layer or multiple layers
Sandstone or carbonate
It is thought by some that gas shales and CBM are Tight Gas.

Tight gas, unlike shale gas, is sourced in another formation, migrates, and is trapped (like conventional gas) in the
formation where it is found. Discrete gas/water contacts are usually absent, but wells do produce water. Tight gas reservoirs
in the US Rocky Mountains can be grouped into four general geolocic and engineering categories: 1. marginal marine blanket,
2. lenticular, 3. chalk, and 4. marine blanket, shallow deposits (Spencer 1985). The Pinedale Anticline (Figure 5), the largest
tight gas reservoir in the US is a lenticular formation. Microscopic study of pore/permeability relationships indicates the
existence of two varieties of tight reservoirs. One variety is tight because of the fine grain size of the rock. The second variety
is tight because the rock is relatively tightly cemented and the pores are poorly connected by small pore throats and capillaries.
Most of the tight gas reservoirs of the Rocky Mountain region of the US are overpressured (Spencer 1985).
SPE 160855 13
North American Tight Gas Basins and Statistics Comparison
Based on the observations of these authors and available statistics from a number of sources, the following
conclusions are drawn and information offered:
There are no typical tight gas reservoirs.
Tight gas wells must be fracture stimulated to produce commercially.
Average well spacing is now 5 to 10acres in the lenticular formations, Pinedale Anticline and Piceance.
Formation thickness ranges from 600 to 6,000 feet.
Formation depth ranges from 4,700 to 20,000 feet.
Multi-wells pads wells are S-shaped, Directional or Vertical (Pinedale Anticline and Piceance).
Some horizontal and multilateral wells (Texas Panhandle, Anadarko Basin)
Well IPs range from 3 to 20 MMcfd.
Production is dry gas, some wet gas, and water.
Tight gas formations producing water require deliquification.
Wells exhibit high decline rates in first few years on production.
A high number of wells is required to develop shale (low per well EURs).


The four tight gas basins that produce most of the US tight gas are the Pinedale Anticline, Anadarko, Piceance, and
Deep Bossier. Table 6 is a comparison of these US tight gas basins.



Table 6Comparison of the four significant US tight gas basins (Source of information: Warlick 2010)


The Pinedale Anticline Field is the largest US tight gas play, holding 73 Tcf of TRR. It produces from stacked
lenticular sands and is typical of other lenticular US tight gas (Table 6). Figure 26 shows a plot of historical production and
well count of producing wells. Currently, Pinedale is producing over 1,500 MMcfd and 54 MBWPD from 1900 wells.
Currently there are 17 gas-directed rigs working in the Pinedale; that number has been about the average during 2011.



Figure 26Pinedale Field historical production and producing well counts


14 SPE 160855
Canadian tight gas information is limited; therefore only basic information is listed for each of the three typical plays
(Dixon 2005).
Shallow: Well Depth - 2,000 2,500 ft, Typically Vertical; Well Rate - 33 Mcfd Avg, some 1.0 MMcfd; Operators -
200; located - SE Alberta / SW Saskatchewan
J ean Marie: Well Depth - 4,000 4,500 ft TVD, Typically Horizontal, Lateral Length - 2,900 5,000 ft; Well Rate -
480 Mcfd Avg; located - NE British Columbia
Deep Basin: Well Depth 7,000 11,000 ft, Typically Vertical; Well Rate 400 Mcfd Avg; Operators 200;
located - West Central Alberta / parts of British Columbia

All tight gas wells display the unique decline curve profile similar to shale gas. Figure 27 shows several modeled
production profiles of various tight gas well scenarios compared to the profile for a conventional gas well plotted from actual
measurements obtained from a productive field. Plots show that the initial rates and EUR per well are significantly less than
those for conventional gas wells.


Figure 27Modeled typical tight gas production profiles compared
to a conventional gas well (Al Kindi et al 2011)

At this point it is deemed beneficial to make a total comparison of a number of the reservoir and producing
characteristics of shale gas, tight gas, and conventional gas. Table 7 is that comparison. The authors have assembled the data
for this table from various sources.




Table 7Comparison of Shale Gas, Tight Gas, and Conventional Gas
SPE 160855 15
The Asset Life Cycle
The authors have compiled the information included in the next section of this paper based on their experience in the
industry and with unconventional gas. It represents our thoughts, and it is the platform adopted by our company as the
suggested method for operators to use when analyzing, developing and producing unconventional shale and tight gas
reservoirs. The Asset Life Cycle, Figure 28, includes five phases A. Exploration, B. Appraisal, C. Development, D.
Production, and E. Rejuvenation. It is recognized that most of the terms have been around the industry for a number of years,
except for possibly Rejuvenation, a term coined by the authors. We are not implying that the phases of the Asset Life Cycle
are totally new, but only how they are implemented and the objectives to be accomplished during each phase. As the
discussion ensues, it will become obvious that the description of the objectives and challenges of the operator and technologies
required to implement each phase of the life cycle do address the uniqueness of both shale and tight gas.

Figure 28The Shale Gas and Tight Gas Asset Life Cycle

Choices made at every phase of the life cycle can affect ultimate recovery. We have seen that not all shale and tight
gas reservoirs are the same, and each may require different choices. Also, each choice can affect later options. Each phase of
the life cycle has a number of different objectives and challenges.

A. Exploration Phase Objectives
Conduct a basin/area screening study to identify core areas (sweet spots) and to determine an initial estimate
of gas in place (GIP).
Begin to characterize the reservoir.
Determine the initial economic value and reservoir potential.

A screening study is particularly important when entering a new basin or area. The primary purpose of the study is to
identify the core areas, i.e., locate the sweet spots. Well-by-well production data indicate that shale formations have small
spots of very productive wells (Sweet Spots), surrounded by large areas of wells that produce far less gas. Sweet spots are a
function of TOC, thermal maturity, thickness, GIP, natural fractures, mineralogy, and geomechanics stresses in the area.
Sweet geologic spots may not necessarily be sweet economic spots. Also, if an area possesses most of these attributes, but is
not a favorable area in which to frac (mineralogy or stresses), it is not a sweet spot. It may sound trite, but develop the sweet
spots first, then go back to the less attractive areas.

The basin screening study should involve gathering and analyzing data including:
Geology sedimentology, statigraphy, and depositional environment
Geochemistry TOC (initial reserve estimate), thermal maturity (type of hydrocarbon).
Is the shale a source rock?
Geomechanics stress regime for well drilling and fracturing design
Petrophysics rock type, lithology/mineralogy, porosity (from cores and logs)
Existing well data

To begin initial characterization of the reservoir, conduct geophysics - 3D seismic. From 3D seismic the usual
information on faults, formation thickness, depth, and lateral continuity can be obtained. However: 3D seismic can also:
Identify areas of highest TOC using acoustic impedance
Increase understanding of natural fractures using seismic attributes
Assist in identification of sweet spots using seismic cross-plots
16 SPE 160855
Seismic information is relevant through the Exploration, Appraisal, Development, and Rejuvenation phases of the life cycle.

Openhole logs (conventional, pulsed-neutron, and spectroscopy) and cores from exploratory wells provide the data
for petrophysical analysis for initial reservoir characterization for both shale and tight gas. Wellbore image logs and nuclear
magnetic resonance (NMR) logs provide necessary information for shales and useful information for tight gas (Holditch 2006).
An example of one of the special logs and analysis techniques is shown in Figure 29, an Integrated Shale Analysis plot that is
described below by Mitra et al (2010). This shale gas facies expert system provides operators with a quick and accurate
method of classifying shale gas reservoirs, identifying favorable zones for hydraulically fracturing, identifying frac barriers
and locating zones from which to drill horizontal laterals (J acobi et al 2009; LeCompte et al 2009; Pemper et al 2009; Mitra et
al 2010). Although this example is for shale, mineralogy/lithology is also being used for complex tight gas sands and
carbonates fracturing/lateral location and identifying reservoir layers. It should be noted that cores are a must, either whole
cores or sidewall cores for analysis and to calibrate logs.

An initial assessment of reservoir potential and economic value can be determined from all these data. Individual
operators have different drivers and specific financial and leasehold situations in the US. Gas price, regulatory, and
infrastructure are all different for countries outside the US. Martin and Eid (2010) cover these topics at length in their paper
on The Potential Pitfalls of Using North American Tight and Shale Gas Development Techniques in the North African and
Middle Eastern Environments.

B. Appraisal Phase Objectives
Drill the appraisal wells
Build reservoir model(s) for simulation
Generate a Field Development Plan
Validate the economics of the play

More wells are drilled during the Appraisal phase than in the Exploration phase; thus data from these additional wells
will continue to be used to further characterize the reservoir. Vertical wells are required to collect data; and some horizontal
appraisal wells are drilled to test hydraulic fracturing and mechanical well completion designs. Horizontal wells will also
provide information to assist in determining optimum lateral length, and to begin early drilling otimization.

Cox et al (2002) recognized that Tight gas reservoirs present unique challenges to the reservoir engineer. Applying
classic reservoir engineering techniques to these reservoirs is problematic due to the length of time to reach pseudo-steady
state flow and/or establish a constant drainage area. This leads to the inability to accurately estimate the recoverable reserves
in a timely and consistent manner. Both decline curve and material balance methods were found to have serious drawbacks
when applied to tight gas reservoirs that had not established a constant drainage area. Holditch (2006) concluded that the best
reserves evaluation techniques for tight gas were careful application of hyperbolic decline curves and reservoir modeling
simulations. Kupchenko et al (2008), upon recognizing that production performance from tight gas reservoirs displays steep
initial decline rates and long periods of transient flow, realized that inaccurate forecasts would result from using this transient
production data. Their work resulted in use of Arps original equation with certain exponent restrictions to obtain better
forecasts. Also in 2008, Ilk et al introduced the power law exponential decline (form of power law loss ratio) concluding
that it offered a better match to production rate than hyperbolic decline. Others, including Duong (2010), have also developed
and proposed decline curve analysis (DCA) methods, and some have offered new techniques for using the material balance
approach (Payne 1996; Engler 2000). Holditch (2006) concludes that the most accurate reservoir analysis technique for tight
gas is to build a reservoir model that includes layers, and these authors also suggest a dual porosity model.

The industry has taken a traditional approach to developing shale gas; looking at these unconventional shale plays in
a statistical manner. The classic DCA approach is being applied, but the average curves that have been developed are not
truly representative of the physics of shale gas flow. Actual performance has been found to be quite dissimilar from these
average or type curves. Since operators do not understand the exact reasons for the deviation, they have been limited in their
ability to optimize the development and properly prioritize operations based on sound engineering and geological information.
A more reliable analysis and predictive approach was needed. According to Vassilellis et al (2010), conventional reservoir
engineering tools have been found to be inadequate for use with the change in reservoir characteristics after hydraulically
fracturing a shale well. This complex newly-altered reservoir (after fracturing) must be described and properly modeled in
order to reliably predict long term production and recovery. They introduced a multi-disciplinary integrated approach called
shale engineering. Shale engineering involves building three models - reservoir, well, and fracturing models. Data and
analysis techniques involve the disciplines of geology, petrophysics, geomechanics, geochemistry, seismology, and, of course,
engineering. Application of the shale engineering techniques are documented by Vassilellis et al (2011) and Moos et al 2011).
Cipolla et al (2209a, 2009b) also has introduced a new approach to more comphensive modeling of complex shales.

SPE 160855 17

1 2 3 4 5 6 7 8 9 10 11 12 13
Figure 29An Integrated Shale Analysis plot of a representative well from the Barnett shale Tracks 8 and 9 show the
results of the shale gas facies system applied to this well. Track 8 represents the lithofacies from the Barnett model.
The organic-rich shale lithofacies is shown in black, the non-siliceous organic-rich shale is shown in orange, the low
organic shale lithofacies is shown in gray, the siliceous mudstone lithofacies is shown in yellow, the calcareous
mudstone lithofacies is shown in blue, the phosphatic zone lithofacies is shown in light green, and the pyritic zone
lithofacies is shown in red. Track 9 represents the stop-light component of the shale gas facies expert system, where
favorable frac zones are marked in green and unfavorable frac zones are marked in red. The predominant lithofacies
in the Barnett shale is the siliceous mudstone followed by the organic-rich shale lithofacies. Track 10 provides
information on frac migration, Track 11 geomechanical properties and Tracks 12 and 13 information on anisotrophy.
(Mitra et al, 2010).

Field Development Plans (FDP) for both shale and tight gas include well type, placement, attitude, direction, and
spacing. Drilling wells in the direction of maximum principal stress maximizes access to existing natural fractures when
transverse-trending hydraulic fractures intersect these natural fractures (previously discussed). Therefore, it is important to
understand the stress regime in the field. Usually a full plan also includes completion and fracturing designs. The industry has
been successful in generating FDPs in the past for conventional reservoirs. Tight gas, and especially shale gas, have
introduced uncertainty in the traditional approach. From the earlier discussion we have seen that it requires a large number of
wells to develop either a tight gas or shale gas play. Tight gas well spacing in Pinedale and Piceance Basins is down to 5 and
10 acres. Although typical shale gas well spacing is somewhat larger, approximately 80 acres, the continuous shale formations
extend over large geographical areas. Figure 30 shows the number of existing wells in the six major US shale plays and the
total number of wells required to develop the TRR from EUR per well for each play (Table 4) using the typical number of 200-
300 wells required to recover 1 Tcf of gas. Most plays have not yet even approached the required number of wells.
18 SPE 160855


Figure 30Typically it takes 200-300 wells to develop 1 Tcf of gas. Based on the total TRR of a
play and the EUR per well, required number of wells to develop is shown in red.
Current (12-1-12) total number of producing wells is shown in black.

Finally, with all of the data collected from the drilling and analysis of the appraisal wells, and with an understanding
of the unique aspects of these unconventional reservoirs and characterization data for the particular play, operators can
complete the final step of the Appraisal phase - validating the economics of the play. The decision whether to proceed with
play development is then taken.

C. Development Phase Objectives
Implement the Field Development Plan
Install surface production and export facilities, including compression and pipelines
Design wells and optimize drilling costs
Refine and optimize the hydraulic fracturing and well completion designs

In implementing a Field Development Plan, that includes a large number of wells, an operator is cautioned not to
become complacent and continue to allow the rig schedule to totally drive the plan. Interim analyses should be undertaken to
ensure that drilling and completion programs are delivering wells with the expected IP producing rates. That is the purpose of
steps three and four of the Development Phase.

Surface facilities will not be discussed in this paper. They are only included in the life cycle for completeness and to
ensure construction schedules match timing of well completion and availability.

Tight and shale gas well designs are different. Historically, most tight gas wells have been drilled conventionally,
over-balanced and with conventional rotary rigs. Typically, Pinedale has proved technically challenging to drill and cement.
It has over 5,000 feet of stacked lenticular sands. Pore pressures ranging from normal to over 16.5 lbm/gal, and unplanned
circulation losses have negatively impacted drilling and cementing. In 2002 desired top of cement was being achieved only
31% of the time (Garcia et al 2002). Operators have solved these problems by using geomechanics and integrating
information from fracturing completion design. As late as 2007 operators were still having drilling problems and facing the
inability to get casing to bottom. The solution was to forego openhole logs, run casing immediately, then run cased hole
pulsed neutron logs for evaluation. These cased hole logs provided equal or better results than openhole logs. In the Pinedale
and Piceance, S-shaped and J type wells are being drilled from 16-20 well pads, such that the well vertically penetrates the
2,000 to 6,000 feet thick formations. J anwadkar et al (2006a) introduced new technology to overcome the directional drilling
challenges of the S and J type wells. Pilisi et al (2010) introduced the Tight Sand Advisory System for selecting the best
drilling method and technologies for any specific tight gas well. Horizontal wells, and some multilateral wells (Goodlow et
al 2009), are being drilled in the tight gas of the Texas Panhandle, Anadarko Basin, as the formation allows. Some recent
drilling methods include, under-balanced drilling (Cade et al 2003) in US and Middle East, casing drilling, managed pressure
drilling, and coil tubing drilling.

The early shale gas wells in the Barnett were mostly verticals; it was not until 2003 that there was a total shift to
horizontal wells for developing shale in the Barnett as well as all other US shale basins. Now that the template has been set for
shale gas drilling in the US, service companies have been successfully reducing drilling costs through optimized drilling and
SPE 160855 19
new technology. To date there have been approximately 55,000 shale wells drilled in the US. Some of the technological
advances include J anwadkar et al (2006b), BHA and drilling string modeling to optimize Barnett drilling performance:
J anwadkar et al (2007), advanced LWD and directional drilling technologies to overcome completion challenges in Barnett
horizontals; J anwadkar et al (2009), innovative rotary steerable system to overcome challenges of the Woodford complex well
profiles: Isbell et al (2010), new use of PDC bits to improve performance in shale plays; and J anwadkar et al (2010), using
electromagnetic MWD to improve Fayetteville drilling performance.

Observations on drilling shale wells in the US by these authors:
All development wells are horizontals.
Typical drilling time in Barnett and Marcellus =12 days, Eagle Ford =17 days
Directional drilling with motors or rotary steerable systems or combinations
Using mostly PDC bits
After drilling the vertical part of hole, most (90
+
%) of the wells convert the drilling fluid to some form of oil
based mud (OBM) to drill the curve and lateral.
Some wells being drilled with environmentally friendly water based mud (WBM)
Mud weights depend on formation, which ranges from normal to overpressured.
High bottomhole temperatures are experienced in the deeper Haynesville and parts of the Eagle Ford.
On a fair number of wells the curve and lateral are being drilled in one run (one BHA, one bit, one trip) in
the Barnett, Marcellus, and Eagle Ford.
Early preference for drilling wells in the toe up attitude is gradually changing to drilling the lateral as flat
as possible and perfectly horizontal.
Completion part of hole typically drilled out of 7 in. casing set either before or after the curve
Cased hole wells typically use either a 4 in. or 5 in. completion string.
Optimum (and shorter) lateral lengths are now preferred over longer lateral lengths.
Longer laterals face increased risk of encountering a geohazard, problem with initiating the frac at the well
toe, and possibly even losing the wellbore.
Wells drilled in a direction normal to maximum principal stress
Some pad drilling is being used for both logistics and environment; 4-10 wells per pad in US; larger 16-well
pads in Canada.
The drilling cost constitutes 40 50% of the total well cost.

The last step of the Development phase of the life cycle involves optimizing the hydraulic fracturing and completion
designs. First, we will cover the types of mechanical completions that are necessary when conducting multi-stage fracturing of
tight gas wells. In the S and J type wells of the Pinedale and Piceance tight gas, operators usually run cased and cemented
completions. Drillable plugs are used to separate stages, and perforations are in clusters throughout the stage length. This
method is commonly termed Plug-N-Perf. In the Texas Panhandle granite wash tight gas, some 70% of the wells are now
horizontals; and the completions are primarily cased and cemented with Plug-N-Perf multi-stage fracturing. Some wells are
also completed openhole and use the mechanical completion method of frac sleeves separated by packers for isolation of
individual stages. These same two methods of mechanical completions, cased and cemented with Plug-N-Perf and openhole
with frac sleeves separated by packers, are being used for shales. There are advantages and disadvantages to the two these
types of mechanical completions; however, operators tend to prefer the cased and cemented method for shale gas completions.
Studies have been conducted, and generally conclude that there is no appreciable difference in well IPs by using either
method. Certainly there is a time difference with the frac spread being on the well for a typical Plug-N-Perf operation of about
a week versus about a day for the openhole frac sleeve and packers completion. As a note, operators tend to prefer the
openhole completions for shale oil. Wells are also being fractured stimulated using coil tubing (Ravensbergen 2011).

Although the hydraulic fracturing process has been around the oil and gas industry for over 60 years, combining it
with horizontal drilling (in existence even longer) has resulted in the shale gas boom. To cover the topic of hydraulic fracturing
in depth would take a much longer discussion than allowed in this paper. However, we will provide an overview frequently
calling on the two very comprehensive papers by King (2010 and 2012). Starting with No two shales alike, King also states
that There are no optimum, one-size-fits-all completion or stimulation designs for shale wells.

The hydraulic fracturing process consists of five steps:
1. Pump the Pad, mainly fluid that cracks the rock creating fractures to accept the proppant.
2. Pump the Slurry, fluid and Proppant (size-graded sand particles or man-made) which props open the fractures.
3. Flush to clean equipment/tubulars of proppant, then shutdown pumps.
4. Bleed off well pressure to allow fractures to close on proppant.
5. Recover injected fluid by flowing or lifting well (typically <30% of frac fluid recovered).

20 SPE 160855
The first fracs in the Barnett were gelled fracs until the successful slickwater fracs became the default design not only
in the Barnett, but also other US shales. Slickwater fracs are made up of 94% water (no polymer gelling agents) as the frac
fluid; 0.3% chemicals - friction reducers, surfactants, biocides, and clay stabilizers; and 5.6% sand proppant. These fracs are
pumped at very high rates. Slickwater fracs are less expensive than polymer gel fracs.

Hydraulic fracturing challenges in shale reservoirs:
Simple or complex geometry
Compatibility of fracturing fluid with reservoir
Proppant types and selection
Reservoir volume accessed
Number of stages, spacing; perforation clusters, spacing (Frac Design)
Geohazards faults, karsts, wet zones
Where did the frac go, and what did it touch (or not)?

Rules of Thumb for shale fracs (The Trend?):
Distance between frac stages =1 to 1.5 x zone height (250 350ft)
Distance between perf clusters =35 to 50+ft
Length of perf cluster =4 x well diameter (about 1 to 2 ft)
Number of perf clusters in each stage is 4 to 8 (roughly 1.5 bpm/perf 10-15 bpm per cluster
Number of stages depends on lateral length, normal 4 to 20+
Slickwater / Linear Gel / Hybrid / Cross-linked depends on type production:
- Slickwater or Linear Gel Fracs Dry gas shale or with little liquids
- Hybrid Fracs (Slickwater & Cross-linked fluids) Gas condensate/liquids/liquid bearing shales
- Crosslinked Frac Oil-bearing shale or with higher GORs

Table 8 shows typical fracture treatment parameters for some of the major US shale plays. Figure 31 shows the trend
in fracture treatment size.


Table 8Typical fracture treatment parameters for some major US shale plays (Various Sources)



Figure 31Trend in Shale fracture treatment size, fluid volumes (Various Sources)
SPE 160855 21
The fracturing procedure for the stacked lenticular sands of the Pinedale and Piceance involves selecting the best 30
from as many as 60 lenses (each 20 to 50 feet thick) to frac. The 30 lens are grouped in five to six to frac together as a stage.
Often post-frac well performance in tight gas reservoirs correlates more directly with fluid volume than proppant volume. For
the Piceance completions, several operators have improved well productivity by doubling fluid volume and maintaining the
same proppant volume by cutting the proppant concentration in half. Rules of Thumb like this and those above are being
used by operators, as there is a general lack of reliance in using complex hydraulic fracture simulators to design and optimize
fracture treatments (Cramer 2008). King (2010) states that there are many good fracture simulators developed primarily for
sand; however, these simulators may not be suited for shales. Table 8 and the preceding Rules of Thumb show that the
fracture treatments being used in shales today are what are called geometric fracs, i.e., a frac stage every 250 to 350 ft. with 4
to 8 perforation clusters per stage. This approach totally ignores the changing reservoir characteristics along the 4,000 to
5,000 ft. long lateral. Geometric fracs are being used, because those changing characteristics along the lateral are not known,
quantitatively at least. No logs or any characterization is being done for the laterals; which could provide information as to
where to place stages and perf clusters, and which places to avoid. These authors suggest running LWD imaging tools along
the lateral, as only a limited number of operators are doing. Costs are relatively inexpensive, and the process is transparent to
the drillers. Imaging tools can identify natural fractures, faults, bedding planes and even induced fractures from nearby offset
wells. One point is that resistivity imaging tools must be run in WBM or invert emulsion with part water in order to record
significant data. Other methods to further characterize the lateral are, analyzing drill cuttings for TOC and mineralogy, and
gas isotopes from mud logging.

Hydraulic Fracturing works for tight gas because we change the flow pattern in the reservoir. Fracturing can improve
the productivity of a well in a tight gas reservoir because a long conductive fracture transforms the flow path gas must take to
enter the wellbore (Holditch 2009). A number of operators are now opting to monitor fracturing treatments, shale and tight
gas (Warpinski et al 2010) in real-time using microseismic. Monitoring does require a nearby offset well in which to run
sondes for recording the data. Microseismic monitors the treatment as to the direction (azimuth) and height, and whether the
treatment is going out of zone, into a water zone, or being lost to a fault. This provides the operator with the ability to stop the
treatment if not going as planned. Microseismic does not indicate where the proppant or fluid actually goes, but where the
rock has slipped or cracked. The events being recorded are the faint sounds of slipping/cracking rock. Microseimic events
plotted as a cloud provide an approximation of the size and location of the stimulated reservoir volume (SRV).

D. Production Phase Objectives
Monitor and optimize producing rates
Manage the Water Cycle Sourcing for drilling and fracturing water, well flowback water, lifting, treating,
handling, and disposal of water
Reduce corrosion, scaling, and bacterial contamination in wells and facilities
Protect the environment

Managing and controlling well flowback rates are the first steps in optimizing production and ultimate recovery.
Multi-stage hydraulically fractured wells require a post-stimulation flow period to prepare a well for long-term production.
This is one of the most critical times in life of well; more so for shale gas wells as opposed to tight gas wells. Excessive
flowback rates are known to have caused proppant flowback or fracture collapse. Intensive management of flowback can
yield significant improvement in wells long-term performance (Crafton and Gunderson 2007; Crafton 2010). An operator in
the Haynesville reported in 2010 that Haynesville wells have been produced using restricted rate production practices.
Additionally, initial decline rates appear more gradual as a result of restricting production. This operator also said that the
decline curves modeled higher EURs from the restricted well rates. It appears that this is one instance of a technique that
could slow down the dramatic initial decline rates characteristic of shale gas wells. Total production from a multi-stage
hydraulically fractured well can be monitored, but there has not been a truly effective method of determining production rates
coming from individual perforated stages. Some operators have run production logging tools (PLT) in horizontal wells
(Heddleston 2009). It requires a tractor or coil tubing, and is somewhat problematic and with mixed success. These PLTs
have confirmed what operators have suspected that some 30 50% of the frac stages are not producing any gas at all. This
is good information to know, but the question begs as to whether any remedial work might be attempted aside from a possible
refrac. One other method of monitoring production from individual stages has been employed; that is DTS (Distributed
Temperature Sensing). A few wells have been equipped with DTS, but the fiber optic cable and equipment must be installed
as part of the original completion and the cost is difficult to justify based on the real benefit.

Most shale gas wells do not produce any significant amounts of water, and the water produced by tight gas wells is
handled with deliquification techniques plunger lift, foam sticks, gas lift, beam lift pumps and jet pumps. Water that is of
concern with shale wells is frac flowback water. Although not all of the water comes back, the amount that does brings with it
formation salts, scale, and sometimes low-level radiation (NORM). Frac flowback water must be treated, either for re-use or
disposal. With the water situation in many parts of the country and the world, re-use is strongly recommended. Service
22 SPE 160855
companies provide water treating services for flowback and produced water. Currently the most popular and effective
equipment uses electro-coagulation technology to remove suspended solids and heavy metals from flowback and produced
water. Fresh water is not required for fracturing wells. Formation brine water and seawater are other alternatives. Additional
frac additive chemicals are required due to salt content of these waters. Gaudlip et al (2008) describes the stringent regulatory
restrictions of the State of Pennsylvania for the Marcellus shale with respect water sourcing, handling and especially disposal.
There is more to managing the water cycle than treating and or disposal. Water sourcing for both drilling and fracturing has
become significant. In the Eagle Ford water is being sourced from shallow salt water formations and being lifted from wells
using large-volume electric submersible pumps (ESPs). In the Horn River Shale in Canada, Apache is taking brine source
water from a saltwater-containing formation just above the Horn River shale and using the water for fracturing (King 2012).
On the surface, produced and treated water must be handled and transported to central processing/treating facilities or removed
for disposal. This requires piping and surface pumps.

Preventing corrosion, scaling, and bacterial contamination in wells and facilities is handled much the same way as in
traditional oil and gas fields. Production chemical monitoring and treating programs must be developed and equipment
installed. Chemical automation systems can be used, especially for remote locations, wide-spread operations, and in low-
winter temperature operations. Protecting the environment should be included in every phase of the life cycle; however, it is
particularly critical during the Production phase.

E. Rejuvenation Phase Objectives
The Challenge for the Rejuvenation Phase is to remediate low-rate and sub-economic wells.
Evaluate wells for Re-Frac candidates
Analyze field for re-development potential (Infill Drilling)

It is the opinion of these authors that the most significant opportunity to accomplish rejuvenation lies with refracs. As
we have seen, unconventional wells decline rapidly reaching low unacceptable rates after only a few years on production. It
has not been proven that any form of production management or enhancement has been successful in arresting the rapid
decline or restoring original production rates. In his database of 100 published studies on refracs, Vincent (2010) attributed
refrac success to a number of mechanisms as listed below:
Enlarged fracture geometry, enhancing reservoir contact
Improved pay coverage through increased fracture height in vertical wells
More thorough lateral coverage in horizontal wells or initiation of more transverse fractures
Increased fracture conductivity compared to initial frac
Restoration of fracture conductivity loss due to embedment, cyclic stress, proppant degradation, gel damage,
scale, asphaltene precipitation, fines plugging, etc,
Increased conductivity in previously unpropped or inadequately propped portions of fracture
Improved production profile in well; preferentially stimulating lower permeability intervals [reservoir
management]
Use of more suitable fracture fluid
Re-energizing or re-inflating natural fractures
Reorientation due to stress field alterations, leading to contact of new rock

Production rates from refracs have matched, or sometimes exceeded, those from the original frac. Figure 32 is a
Barnett refrac with slickwater compared to the original gel frac (Cipolla 2005). Figure 33 is a refrac of a tight gas well, GRB
45-12 from Green River Basin, Wyoming (Reeves et al 1999).



Figure 32Refrac of Barnett shale well with Figure 33Refrac of tight gas well GRB 45-12
Slickwater (Cipolla 2005) (Reeves et al 1999)
SPE 160855 23
Re-development of a shale or tight gas field will more than likely involve infill drilling as a possible result from
downspacing. This has already been seen in the Pinedale and Piceance tight gas plays; where original spacing was on 160
acres and has now gone down to 5-10 acres.

Conclusions
Unconventional shale gas and tight gas are different in many respects, while similar in other respects.
All Shale Gas reservoirs are not the same; there are no typical Tight Gas reservoirs.
Shale reservoirs are source rocks where the gas (some or all) has remained; gas has migrated and is trapped
in tight gas reservoirs.
Organic-rich shale lithology/mineralogy is quite different from tight gas sands or carbonates.
Both shale and tight gas have low permeability and relatively low porosity.
Gas is stored as free, sorbed in matrix and natural fractures, and dissolved in bitumen in shales; gas is
stored only in the pores of tight gas reservoirs.
Reservoir and flow mechanisms are different.
Formation evaluation and reservoir analysis methods are quite different for each.
Drilling methods are different and dictated by type of formation.
Development well types are different; horizontals for shale and vertical, S-shape, directional, and some
horizontals for tight gas.
Most significant common link; hydraulic fracturing require for both to attain commercial gas rates.
Fracturing techniques are similar with different designs for each; there is no one-size-fits all fracturing
designs for either tight or shale gas.
Completion techniques are similar for both.
Production decline character is similar for both.
Levels of production rates and EUR are generally similar for both.
Both produce mostly dry gas; no water from shale but water from tight gas wells requiring deliquification.
Water management is similar for both.
Refracs have been successful in both.

Can all of the technologies and the current development model employed in North America for shale gas and tight gas
be easily transferred to China, Latin America, the Middle East, North Africa and other parts of the world? It is the considered
opinion of these authors, that most of the technology, although with sound application considerations, can be transferred to
other parts of the world. The development model, especially for shale gas, established in North America cannot be easily
adopted by other countries for the obvious reasons of differences in infrastructure, including equipment and service company
availability, governmental regulations, logistics, processing, environmental considerations, and gas pricing. We believe that it
is likely that environmental concerns and the drive to reduce development costs of unconventional gas will drive new
approaches to the development in China, Latin America, Middle East, North Africa, and other parts of the world. Our North
American industry experience can assist in accelerating the process by providing our neighbors around the world with
information and data to help shorten their learning curve.


Acknowledgements
The authors would like to thank the management of Baker Hughes for allowing us to publish this paper. We would also like to
thank Lucy Luo for her work with the databases to generate a number of the plots in this paper.

References
Addis, M.A. and Yassir, M., 2010, An Overview of Geomechanical Engineering Aspects of Tight Gas Sand Developments,
Paper SPE 136919, presented at the 2010 SPE/DGS Annual Technical Symposium and Exhibition, Al-Khabor, Saudi
Arabia, 04-07 April.

Al Kindi, S., Weissenback, M., Mahruqi, S., Curtino, J ., and Al Siyabi, H., 2011, Appraisal Strategy for a Tight Gas
Discovery, Paper SPE 142735, presented at the SPE Middle East Unconventional Gas Conference and Exhibition, Muscat,
Oman, 30 January-2 February.

British Columbia (B.C.) Ministry of Energy and Mines National Energy Board, 2011, Ultimate Potential for Unconventional
Natural Gas in Northeastern British Columbias Horn River Basin, (May 2011).

Cade, R., Kirvelis, R., Nafta, M. and J ennings, J ., 2003, Does Underbalanced Drilling Really Add Reserves?, Paper SPE
81626, presented at the IADC/SPE Underbalenced Technology Conference and Exhibition, Houston, Texas, U.S.A., 25-26
March.
24 SPE 160855
Cipolla, C.L., 2005, The Truth About Hydraulic Fracturing Its More Complicated Than We Would Like to Admit, Paper
SPE 108817, SPE Distinguished Lecture Series, 2005 -2006.

Cipolla, C.L., Lolon, E.P., Erdle, J.C., and Rubin B., 2009, Reservoir Modeling in Shale-Gas Reservoirs, Paper SPE 125530,
presented at the SPE Eastern Regional Meeting, Charleston, West Virginia, USA, 23-25 September.

Cipolla, C.L. Lolon, E.P., Erdle, J .C., and Tathed, V., 2009, Modeling Well Performance in Shale Gas Reservoirs, Paper SPE
125532, presented at the SPE/EAGE Reservoir Characterization and Simulation Conference, Abu Dhabi, UAE, 19-21
October.

Cox, S.A., Gilbert, J .V., Sutton, R.P. and Stoltz, R.P., 2002, Reserve Analysis for Tight Gas, Paper SPE 78695, presented at
the SPE Eastern Regional Meeting, Lexington, Kentucky, U.S.A., 23-25 October.

Crafton, J .W., and Gunderson, D., 2007, Stimulation Flowback Management: Keeping a Good Completion Good, Paper SPE
110851, presented at the 2007 SPE Annual Technical Conference and Exhibition, Anaheim, California, U.S.A., 11-14
November.

Crafton, J .W., 2010, Flowback Performance in Intensely Naturally Fractured Shale Gas Reservoirs, Paper SPE 131785,
presented at the 2010 SPE Unconventional Gas Conference. Pittsburgh, Pennsylvania, USA, 23-25 February.

Cramer, D.D., 2008, Stimulating Unconventional Reservoirs: Lessons Learned, Successful Practices, Areas for Improvement,
Paper SPE 114172, presented at the 2008 SPE Unconventional Resources Conference, Keystone, Colorado, U.S.A., 10-12
February.

Dixon, R.K., 2005, Tight Gas in Western Canada: An Important and Continuing Component of Overall Supply, presented at
The Unconventional Gas Conference, Houston, Texas, 26 J uly.

Duong, A.N., 2010, An Unconventional Rate Decline Approach for Tight and Fracture-Dominated Gas Wells, Paper SPE
137748, presented at the Canadian Unconventional Resources & International petroleum Conference, Calgary, Alberta,
Canada, 19-21 October.

Economides, M.J ., and Martin, T., 2007, Modern Fracturing Enhancing Natural Gas Production, BJ Services Company,
Houston, Texas, 2007.

Engler, T.W., 2000, A New Approach to Gas Material Balance in Tight Gas Reservoirs, Paper SPE 62883, presented at the
2000 SPE Annual Technical Conference and Exhibition, Dallas, Texas, 1-4 October.

Garcia, J ., Huckabee, P., Hailey, B. and Foreman, J ., 2004, Integrating Completion and Drilling Knowledge Reduces Trouble
Time and Costs on the Pinedale Anticline, Paper SPE 9467, presented at the SPE Technical Conference and Exhibition,
Houston, Texas, U.S.A., 26-29 September.

Gaudlip, A.W., Paugh, L.O. and Hayes, T.D., 2008, Marcellus Shale Water Management Challenges in Pennsylvania, Paper
SPE 119898, presented at the 2008 SPE Shale Gas Production Conference, Fort Worth, Texas, U.S.A., 16-18 November.

Goodlow, K., Huizenga, R., McCasland, M., Clift, D. and Neisen, C, 2009, Multilateral Completions in the Granite Wash:
Two Case Studies, Paper SPE 120478, presented at the 2009 SPE Production and Operation Symposium, Oklahoma City,
Oklahoma, USA, 4-8 April.

Heddleston, D., 2009, Horizontal Well Production Logging Deployment and Measurement Techniques for US Land Shale
Hydrocarbon Plays, Paper SPE 120591, present at the 2009 SPE Production and Operations Symposium, Oklahoma City,
Oklahoma, USA, 4-8 April.

Holditch, S.A., 2006, Tight Gas Sands, Paper SPE 103356, Distinguished Author Series, J PT, J une, 2006.

Holditch, S.A., 2011, Unconventional Oil and Gas go for the Source, presentation, Texas A&M University, 2011.

Holditch, S.A., 2009, Stimulation of Tight Gas Reservoirs Worldwide, Paper SPE 20267, presented at the 2009 Offshore
Technology Conference, Houston, Texas, USA, 4-7 May.


SPE 160855 25
Ilk, D., Rushing, J .A., Perego, J .D., and Blasingame, T.A., 2008, Paper SPE Paper 116731, Exponential vs. Hyperbolic
Decline in Tight Gas Sands Understanding the Origin and Implications for Reserve Estimates Using Arps Decline
Curves, presented at the SPE Annual Technical Conference and Exhibition, Denver, Colorado, USA, 21-24
September.

Isbell, S., Scott, D. and Freeman, M., 2010, Application-Specific Bit Technology leads to Improved Performance in
Unconventional Gas Shale Plays, Paper SPE 128950, presented at the 2010 IADC/SPE Drilling Conference and Exhibition,
New Orleans, Louisiana, USA, 2-4 February.

J acobi, D., Breig, J ., LeCompte, B., Kopal, M., Hursan, G., Mendez, F., Bliven, S. and Longo, J ., 2009. Effective Geochemical
and Geomechanical Characteristics of Shale Gas Reservoirs from the Wellbore Environment: Caney and Woodford Shale,
Paper SPE 124231, presented at the 2009 SPE Annual Technical Conference and Exhibition, New Orleans, Louisiana,
USA, 4-7 October.

J anwadkar, S., Belavadi, M., Fortenberry, D., Dawkins, B., Kramer, M., Devon, S., Privott, S. and Rogers, T., 2006,
Innovative Advanced Technologies Overcome Directional Drilling Challenges of S and J Type Wells in North America,
Paper SPE 103198, presented at the 2006 SPE Annual Technical Conference and Exhibition, San Antonio, Texas, U.S.A.,
24-27 September.

J anwadkar, S.S., Fortenberry, D.G., Roberts, G.K., Kramer, M., Devon, S., Trichel, D.K., Rogers, T., Privott, S.A., Welch, B.
and Isbell, M.R., 2006, BHA and Drillstring Modeling Maximizes Drilling Performance of Lateral Wells of Barnett Shale
Gas Field of N. Texas, Paper SPE 100589, presented at the 2006 SPE Gas Technology Symposium, Calgary, Alberta,
Canada, 15-17 May.

J anwadkar, S., Morris, S., Potts, M., Kelley, J., Fortenberry, D., Roberts, G., Kramer, M., Devon, S., Privott, S. and Rogers,
T., 2007, Advanced LWD and Directional Drilling Technologies Overcome Completion Challenges of Lateral Wells in the
Barnett Shale, Paper SPE 110837, presented at the 2007 SPE Annual Technical Conference and Exhibition, Anaheim,
California, U.S.A., 11-14 November.

J anwadkar, S., Hummes, O., Fryer, C., Rogers, T., Simonton, S. and Black, D., 2009, Innovative Design Rotary Steerable
Technologies Overcome Challenges of Complex Well Profiles in Fast-Growing unconventional Resource Woodford
Shale, 2009, Paper SPE 119959, presented at the SPE/IADC Drilling Conference and Exhibition, Amsterdam, The
Netherlands, 17-19 March.

J anwadkar, S., Klotz, C., Welch, B. and Finegan, S., 2010, Electromagetic MWD Technology Improves Drilling
Performance in Fayetteville Shale of North America, Paper SPE 128905, presented at the 2010 IADC/SPE Drilling
Conference and Exhibition, New Orleans, Louisiana, USA, 2-4 February.

J arvie, D.M., Hill, R.J ., Ruble, T.E. and Pollastro, R.M., 2007, Unconventional Shale Gas Systems: The Mississippian Barnett
Shale of north-central Texas as one model for thermogenic shale-gas assessment, the American Association of Petroleum
Geologists, AAPG Bulletin, V. 91, No. 4 (April 2007), PP 475-499.

J enkins, C.D., and Boyer, C.M., 2008, Coalbed- and Shale-Gas Reservoirs, Paper SPE 103514, Distinguished Author Series,
J PT, February, 2008.

Kawata, Y. and Fujita, K., 2001, Some Predictions of Possible Unconventional Hydrocarbon Availability until 2100, Paper
SPE 68755, presented at the SPE Pacific Oil and Gas Conference and Exhibition, J akarta, Indonesia, 17-19 April.

King, G.E., 2010, Thirty years of Gas Shale Fracturing: What Have We learned?, Paper SPE 133456, presented at the SPE
Annual Technical Conference and Exhibition, Florence, Italy, 19-22 September.

King, G.E., 2012, Hydraulic Fracturing 101: What Every Representative, Environmentalist, Regulator, Reporter, Investor,
University Researcher, Neighbor and Engineer Should Know About Estimating Frac Risk and Improving Frac Performance
In Unconventional Gas and Oil Wells, Paper SPE 152596, presented at the SPE Hydraulic Fracturing Technology
Conference, The Woodlands, Texas, USA, 6-8 February.

Kuuskraa, V., Stevens, S., Van Leeuwen, T. and Moodhe, K., 2011, World Shale Gas Resources: An Initial Assessment of 14
Regions Outside the United States, prepared by Advanced Resources International Inc (February 17, 2011) for the U.S.
Energy Information Administration, U.S. Department of Energy, Washington, DC (April 2011).

26 SPE 160855
Kupchenko, C.L., Gault, B.W. and Mattar, L., 2008, Paper SPE 114991, Tight Gas Production Performance Uning Decline
Curves, presented at the CIPC/SPE Gas Technology Symposium 2008 Joint Conference, Calgary, Alberta, Canada, 16-19
J une.

Lazzari, S., 2006, Energy Tax Policy: History and Current Issues, Congressional Research Services Report for Congress, CRS
RL 33578, Resources, Science, and Industry Division, Congressional Research Services (J uly 2006).

LeCompte, B., Franquet, J .A., and J acobi, D., 2009, Evaluation of Haynesville Shale Vertical Well Completions with a
Mineralogy Based Approach to Reservoir Geomechanics, Paper SPE 124227, presented at 2009 SPE Annual Technical
Conference and Exhibition, New Orleans, Louisiana, USA, 4-7 October.

Martin, A.N. and Eid, R., 2011, The Potential Pitfalls of Using North American Tight and Shale Gas Development Techniques
In the North African and Middle Eastern Environments, Paper SPE 141104, presented at the 2011 SPE Middle East
Oil and Gas Show and Conference, Manama, Bahrain, 25-28 September.

Martin, S.O., Holditch, S.A., Ayers, W.B. and McVay, D.A., 2008, PRISE: Petroleum Resource Investigation Summary and
Evaluation, Paper SPE 117703, presented at the SPE Eastern Regional/AAPG Eastern Section J oint Meeting, Pittsburgh,
Pennsylvania, USA, 11-15 October.

Mastrangelo, E., An Analysis of Price Volatility in Natural Gas Markets, Energy Information Administration, Office of Oil
and Gas (August 2007).

Mitra, A., Warington, D., and Sommer, A., 2010, Application of Lithofacies Models to Characterize Unconventional Shale
Gas Reservoirs and Identify Optimal Completion Intervals, Paper SPE 132513, presented at the SPE Western Regional
Meeting, Anaheim, California, USA, 27-29 May.

Moos, D., Vassilellis, G.D., Cade, R., Franquet, J ., Lacazette, A., Bourtenbourg, E. and Daniel, G., 2011, Predicting Shale
Reservoir Response to Stimulation in the Upper Devonian of West Virginia, Paper SPE 145849, presented at the SPE
Annual Technical Conference and Exhibition, Denver, Colorado, USA, 30 October-2 November.

National Energy Board (NEB - Canada), 2009, A Primer for Understanding Canadian Shale Gas, Calgary, Alberta (November
2009).

Nitze, P.H. and Gruenspecht, H., 2012, Annual Energy Outlook 2012 Early Release, Energy Information Administration, U.S.
Department of Energy, Washington, DC (J anuary 23, 2012).

Nome, S. and J onston, P., 2008, From Shale to Shining Shale A Primer on North American Shale Gas Plays, Duetsch Bank
(July 22, 2008).

Oil and Gas Investor, 2006, Tight Gas (March 2006).

Passey, Q.R., Bohacs, K.M., Esch, W.L., Klimentidis, R. and Sinha, S., 2010, From Oil-Prone Source Rock to Gas-Producing
Shale Reservoir Geologic and Petrophysical Characterization of Unconventional Shale Gas Reservoirs, Paper
SPE131350, presented at CPS/SPE International Oil & Gas Conference Exhibition in China, Beijing, China, 8-10 J une.

Payne, D.A., 1996, Material-Balance Calculations in Tight-Gas Reservoirs: The Pitfalls of p/z Plots and a More Accurate
Technique, Paper SPE 36702, presented at the 1996 SPE Annual Technical conference and Exhibition, Denver, Colorado,
6-9 October.

Petroleum Exploration Society of Great Britain (PESGB), 2008, Exploration and Production in a Mature Basin: North Sea
Petroleum Geology, presentation, Aberdeen, Scotland, April.

Pemper, R., Han, X., Mendez, F., J acobi, D., LeCompte, B., Bratovich, M., Feuerbacher, G., Bruner, M. and Bliven, S., 2009,
The Direct Measurement of Carbon in Wells Containing Oil and Natural Gas Using Pulsed Neutron Mineralogy Tool,
Paper SPE 124234, presented at 2009 SPE Annual Technical Conference and Exhibition, New Orleans, Louisiana, USA, 4-
7 October.

Pilisi, N., Wei, Y. and Holditch, S.A., 2010, Selecting Drilling Technologies and Methods for Tight Gas Sand Reservoirs,
Paper SPE 128191, presented at the 2010 IADC/SPE Drilling Conference, New Orleans, Louisiana, USA, 2-4 February.

SPE 160855 27
Ravensbergen, J .E., 2011, Cased-Hole Multistage Fracturing: A New Coil-Tubing-Enabled Completion, Paper SPE 143250,
presented at the SPE/CoTA Coiled Tubing and Well intervention Conference and Exhibition, The Woodlands, Texas, USA,
5-6 April.

Reeves, S.R., Hill, D.G., Hopkins, C.W., Conway, M.W., Tiner, R.L. and Mohaghegh, S., 1999, Paper SPE 56482,
Restimulation Technology for Tight Gas Wells, presented at the 1999 SPE Annual Technical Conference and Exhibition,
Houston, Texas, 3-6 October.

Reynolds, M., Burke, L. Woitt, M. and Carle, D., 2010, Completions and Stimulation Techniques that Work in HPHT and
Unconventional Gas Reservoirs Focus on Shale Gas, 1QPC Shale Gas Drilling and Completion Post Conference
Workshop, Houston, Texas (May 2010).

Rushing, J.A., Newsham, K.E. and Blasingame, T.A., 2008, Rock Typing Keys to understanding Productivity in Tight Gas
Sands, Paper SPE 114164, presented at the 2008 SPE Unconventional Reservoirs Conference, Keystone, Colorado, 10-12
February.

Shrivastava, C. and Lawatia, R., 2011, Tight Gas Reservoirs: Geological Evaluation of the Building Blocks, Paper SPE
142713, presented at the SPE Middle East Unconventional Gas Conference, and Exhibition, Muscat, Oman, 31 J anuary-2
February.

Spencer, C.W., 1985, Geologic Aspects of Tight Gas Reservoirs of the Rocky Mountain Region, Paper SPE 11647, presented
At the 1983 SPE Symposium on Low Permeability Gas Reservoirs, Denver, Colorado, 14-16 March.

U.S. Department of Energy, Office of Fossil Energy and National Energy Technology Laboratory, 2009, Modern Shale Gas
Development in the United States: A Primer, prepared by Ground Water Protection Council and ALL Consulting (April
2009).

U.S. Energy Information Administration (EIA), 2010, U.S. Crude Oil, Natural Gas, and Natural Gas Liquids Reserves, U.S.
Department of Energy, Washington, DC (November 30, 2010).

U.S. Energy Information Administration (EIA), 2011, Review of Emerging Resources: U.S. Shale Gas and Shale Oil Plays,
prepared by INTEK Inc. for the EIA, U.S. Department of Energy, Washington, DC (July 2011).

Vassilellis, G.D., Li, C., Seager, R. and Moos, D., 2010, Investigating the Expected Long-Term Production Performance of
Shale Reservoirs, Paper SPE 138134, presented at the Canadian Unconventional Resources & International Petroleum
Conference, Calgary, Alberta, Canada, 19-21 October.

Vassilellis, G.D., Li, C., Bust, V.K., Moos, D. and Cade, R., 2011, Shale Engineering Application: The MAL-145 Project in
West Virginia, Paper SPE 146912, presented at the Canadian Unconventional Resources Conference, Calgary, Alberta,
Canada, 15-17 November.

Vincent, M.C., 2010, Refracs - Why Do They Work, and Why Do They Fail? In 100 Published Field Studies?, Paper SPE
134330, presented at the SPE Annual Technical Conference and Exhibition, Florence, Italy, 19-22 September.

Warpinski, N.R., Waltman, C.K. and Weijers, L., 2010, Paper SPE 131776, An Evaluation of Microseismic Monitoring of
Lenticular Tight Sandstone Stimulations, presented at the SPE Unconventional Gas Conference, Pittsburgh, Pennsylvania,
USA, 23-25 February.

Warlick International, North American Unconventional Gas Market Report 2010, Edition 1 of 2, (2010).

Wikipedia, 2011, http://en.wikipedia.org/wiki/Kerogen.

Vous aimerez peut-être aussi