Vous êtes sur la page 1sur 12

Burst and Collapse Responses of

Production Casing in Thermal Applications


D. DallAcqua, SPE, Noetic Engineering; M. Hodder, SPE, Shell Canada Energy; and
T.M.V. Kaiser, SPE, Noetic Engineering

Summary
American Petroleum Institute (API) design equations describing
burst and collapse limits of tubulars do not address pipe body
response when axial stress in the casing exceeds the material
yield strength. However, casing yielding commonly occurs in thermal operations in western Canada, where steam-assisted gravitydrainage (SAGD) and cyclic-steam-stimulation (CSS) operating
temperatures generally range from 200 to 350 C. Cemented production casing is subject to both passive and active loading conditions during operation: thermally induced strain-based cyclic axial
loading occurs in conjunction with net internal or external differential pressure. A sound engineering basis for selecting tubular
configurations that considers the combined loading state in this
situation and establishes an appropriate design margin does not
currently exist.
This paper describes numerical analyses for combined postyield loading conditions and provides a starting point for burst
and collapse design for thermal casing. Burst analysis of axially
constrained casing indicates that, contrary to what might be
inferred from elastic-strength calculations, an initial thermally
induced axial compressive strain does not substantially reduce the
burst (rupture) pressure. By contrast, even low net external pressures can lead to ovalization and loss of wellbore access when
combined with thermally induced axial strain if the cement sheath
does not offer adequate radial support. Sensitivity studies demonstrate the strong influence of pipe diameter-to-thickness ratio (D/
t) and pressure ratios and pipe material mechanical properties on
ovalization response. Analysis results are compared with API
burst and collapse predictions, thermal operating experience at
Shell Canadas Peace River project, and available physical testing
results for similar loading conditions.
Introduction
Pipe body collapse and burst limit characterizations have been the
focus of much effort and standardization. The work is leading to a
better understanding of limits: the impact of axial stresses, material properties, defects, and casing/cement/formation interaction;
statistical variations and probabilistic formulation; and comparison with physical testing results (primarily for collapse). However,
much of the focus of the past work and subsequent standardization
has been geared toward elastic designs and the quantification of
associated safety factors. Work described in this paper explores
deformation responses and sensitivities of cemented thermal casing strings that are axially loaded beyond material yield (by means
of thermally induced mechanical straining) but that are generally
operated at differential pressures that are substantially lower than
those required to satisfy elastic limits associated with burst and
collapse design. Rapid acceleration of the use of thermal
enhanced-oil-recovery (EOR) techniques and the lack of sound
burst and collapse design equations for such wells highlight the
need for advancement in this area. Stated simply, one might cast
the question this way: My pipe has already yielded; what are my
burst and collapse safety margins? Whereas the burst rupture
C 2013 Society of Petroleum Engineers
Copyright V

This paper (SPE 151810) was accepted for presentation at the IADC/SPE Drilling
Conference and Exhibition, San Diego, California, USA, 68 March 2012, and revised for
publication. Original manuscript received for review 24 August 2012. Revised manuscript
received for review 12 November 2012. Paper peer approved 19 November 2012.

March 2013 SPE Drilling & Completion

limit is relatively well-described in American National Standards


Institute (ANSI)/API Technical Report 5C3 (2008) for the range
of axial conditions, collapse limits in a similar state remain unaddressed, and little published work the authors have found enables
quantification of strain levels at intermediate yield conditions (i.e.,
beyond initial yield but far from burst).
Although many papers have been published on design and
selection of tubulars and connections for thermal EOR applications, not one includes an in-depth exploration of the impact of
pressure on the pipe-body design in the context of cross-sectional
stability. The experimental study conducted by Maruyama et al.
(1990) does provide some initial comparisons of the collapse
response of thermal tubulars in tension with API equations, an important starting point and appropriate complementary reference
for the work discussed here.
Pattillo and Huang (1982), Klever and Stewart (1998), and
Klever and Tamano (2006) indicate the relationship between
post-yield pipe pressure capacity and post-yield material properties for both burst and collapse, a dependency we consider highly
relevant to the present discussion, considering the yielding that
occurs during service. Collapse behavior is expected to be appreciably more complex than burst behavior and with different governing variables. We thus review burst as a precursor to the
collapse discussion.
The impact of cement on burst and collapse resistance is not
explicitly considered in this paper, other than in related assumptions made regarding overall pipe constraint and lateral buckling
support. For burst, good cement coverage (and load transfer to the
cement and formation) will add to burst resistance (Zinkham and
Goodwin 1962; Kalil and McSpadden 2011), but we ignore this
benefit here for design purposes. Similarly, good cement coverage
will resist progressive ovalization as external pressure acts on the
pipe (Rodriguez et al. 2003), but we ignore that improved support
so that local variations in cement quality are accommodated in the
design. We assume cement quality is adequate to restrain the casing from expanding axially along much of its length during heating, and to prevent lateral buckling of the pipe structure as
compressive forces are generated. Thus, the general intent is to
enable deformation-tolerant casing designs that do not implicitly
rely on the cement sheath and formation for burst and collapse
resistance. Connection behavior is not addressed in the present
discussion.
Application Details and History. Western Canadian thermal
EOR application started in earnest in the mid-1980s and has been
accelerating ever since. Applications range from low-temperature SAGD operations that operate at peak temperatures of 200
to 300 C to higher-temperature CSS operations that may reach
temperatures as high as 350 C and operating pressures of approximately 16.5 MPa. Whereas SAGD was initially considered a
kinder, gentler process than CSS because it was expected
to involve less thermal cycling, pilot and early commercial
SAGD experiences have made it clear that such wells can experience substantial cycling and should thus be designed with this in
mind.
Shell Canada Energys Peace River operation involves a combination of CSS and SAGD/steamdrive applications in northwestern Alberta. This field is currently composed of 9 pads consisting
of 7 to 16 wells per pad, and has produced up to 2200 m3/d of
93

500

800

50

100

150

200

250

300

350

700
600
API yield
strength 500

400

Stress (MPa)

Stress (MPa)

250

Slope = post-yield stiffness

Cycle 1
Cycle 2
Cycle 3
Cycle 4
Cycle 5
Cycle 6

400
300
200
Material A

100

Material B

250
0
0.0%

500

Temperature (C)

Fig. 1Axial stress development in constrained uniaxial material sample subjected to 330 C cyclic temperature change.

crude bitumen. Operating parameters include maximum temperatures and pressures of 340 C and 14.5 MPa, respectively. The
cyclic wells will generally experience approximately eight CSS
cycles during their life cycle of 15 to 20 years.
A typical modern configuration for these wells is either horizontal or deviated (inclined to near 45 ), and contains intermediate or production casing that penetrates the reservoir. Shell has
traditionally used pipe D/t ratios of between 22 and 24.
Tubular Loading. Thermal EOR tubulars that face the most
extreme service loading conditions are those that are axially confined during operation, but that enable local redistributions of mechanical strain from weak sections to stronger sections of pipe.
Cemented production casing typically endures the most severe
conditions. Axial conditions are largely strain based: In the absence of pressure or other secondary loading, pipe axial force is a
function of thermally imposed mechanical strain and thus the mechanical properties of the steel. By contrast, fluid-pressure conditions represent a sustained force on the pipe that will not be
relieved when the structure deforms. Combined effects of these
loads should be accommodated by the pipe-design basis and are
the key subjects of this paper.
The total, mechanical, and thermal axial strains are related by
etotal emech ethermal ; . . . . . . . . . . . . . . . . . . . . . . . 1
where one may consider the mechanical-strain term emech to be
associated with stress and the thermal-strain term ethermal to relate
to the free expansion and contraction that occur with imposed
temperature changes.
In an axially constrained production-casing string subjected to
thermal cycling, the average total axial strain etotal is zero (i.e., the
length of the cemented pipe does not change), making the
mechanical strain equal in magnitude and opposite in sign to the
thermal strain. In Shells Peace River application, the average
compressive axial mechanical strain applied to production casing
during heating is on the order of 0.5%, and subsequent cooling
(production) and heating (injection) cycles can be anticipated to
impose average cyclic axial strains of similar magnitude. Fig. 1
shows the thermo-mechanical response of a uniaxial microalloyed
L80 sample subject to multiple constrained heating and cooling
cycles of 20 to 350 C; a more detailed discussion of this response
is presented by Nowinka et al. (2008).
Hoop stresses in a nominal-geometry 177.8-mm, 34.2-kg/m
casing associated with Shells peak operating pressures are as
follows:
 For a maximum internal pressure differential of 14 MPa,
nominal tensile hoop stress of 155 MPa
 For a maximum external pressure differential of 8 MPa,
nominal compressive hoop stress of 88 MPa
94

0.5%

1.0%

1.5%
Strain

2.0%

2.5%

3.0%

Fig. 2Example stress/strain curves of two steels exhibiting


identical API yield strength and substantially different postyield stiffness.

The actual loading path of thermal tubulars is a function of


many design, pipe-material/manufacturing, and operating variables, and the design basis should ideally accommodate these paths
without being overly conservative. Furthermore, the path-dependent nature of plastic deformation highlights the need to consider
multiple loading paths, assessing the burst and collapse response
to path dependence. This is complicated substantially by changes
in material response with cyclic loading and by limitations of
analysis tools (Kaiser et al. 2008). Finally, results may enable
more-focused efforts into the remaining uncertainties associated
with the impact of plastic deformation on pipe-body sulfidestress-cracking (SSC) resistance and help us understand differences that might arise from path dependency.
Material Response. As a preface to the burst and collapse discussions presented in this paper, it is important that the reader
understand a few fundamental features of material mechanical
response and their relevance to both load-based and strain-based
designs. Fig. 2 shows two partial uniaxial stress/strain curves
from sample pipe materials that would, by API specifications, be
classified as having identical yield strengths of 552 MPa:
 Material A has an elastic (linear) range that extends to a
peak stress equal to the API yield strength, followed by a flat yield
plateau, and a rounded stress-strain response at higher strain
levels. This curve is representative of some materials at ambient
or low-temperature conditions.
 Material B has a smaller elastic range, followed by a gradual departure from the elastic curve. The slope of this curve
declines gradually with strain; this rounded stress-strain shape is
more typical of oil country tubular goods (OCTG) materials at
elevated temperature.
Both materials reach a stress value of 552 MPa at a total strain
of 0.5%, but the slope of the stress/strain curve of Material B at
that strain is much higher than that of Material A. The slope of
the stress/strain curve is known as the strain hardening modulus
or, as it is referred to here, the post-yield stiffness. It is the postyield stiffness and its progression with strain that generally govern
the deformation resistance of structures after yielding has begun.
For traditional load-based designs in which force levels are
known and traditional safety factors can be used, the amount of
plasticity and the dependence on post-yield properties are typically managed by keeping stress levels well below the yield
strength. Implicit in this design methodology is the acceptance of
a small amount of plasticity if low safety margins are used in conjunction with materials such as Material B.
For situations in which an imposed mechanical strain is anticipated (as in thermal production casing), it is appropriate to consider the material response at a particular strain level, rather than
referring to a stress level, which one might do in a load-based
design. If the imposed strain is adequately large (e.g., 0.5%), all
candidate materials will yield but will exhibit a range of postyield stiffness in the strain range that is relevant to stability.
March 2013 SPE Drilling & Completion

Alberta Energy Resources Conservation Board Directive 010


(2009) does provide suggested design factors for all tubulars
expected to enter sour service, but these are intended for installation loads; the document defers to IRP3 for specific design recommendations regarding casing design for thermal operation.

Relative collapse strength

1.2
API Equation E.8
Hencky-von Mises yield
relationship

1
0.8
0.6
0.4
Compression

Tension

0.2
0
1

0.8

0.6

0.4

0.2

0.2

0.4

0.6

0.8

Axial stress / yield stress

Fig. 3Relative collapse strength relationship with tensile axial


stress from API Technical Report 5C3 (2008) Eq. E.8 and extension to compressive axial stress by use of von Mises yieldsurface definition.

Goals of Investigation. The work described here has the following primary goals:
 Explore API treatment of burst and collapse loads to establish whether any material embedded in Technical Report 5C3
(ANSI/API 2008) or its Annexes, or any other literature, could be
used as part of the basis for thermal burst and collapse design.
 Demonstrate fundamental burst response with combined
loading conditions typical of thermal operations, with a particular
focus on the impact of initial axial stress state (i.e., how burst
response differs with load path).
 Demonstrate fundamental collapse response with representative combined loading conditions, and explore sensitivities to load
paths, pipe geometry, and material properties.
 Assess the burst and collapse responses of Shells existing
Peace River tubulars with representative conditions, and understand performance degradation associated with increases in pipe
D/t ratio.
 Compare modeling results to available testing and fieldperformance data, where possible.
 Set direction for subsequent work that will assist industry in
managing the risk of burst and collapse failures.
Existing Burst and Collapse References
The following discussion briefly reviews industry recommended
practices for thermal casing design and API formulations typically
used to assess burst and collapse limits.
DACC IRP3 (Heavy Oil). The primary thermal casing design
guideline available to operators in western Canada is the Heavy
Oil and Oil Sands Operations Industry Recommended Practice
(IRP), Volume 3 (2002), published by Enform and the Drilling and
Completions Committee and currently under revision. The existing document lists service loads that should be anticipated during
thermal operation, and recommends a basis for selecting pressures
that should be used for designing against burst and collapse:
 Burst pressure rating should be the maximum rated discharge
pressure of the steam generator.
 Collapse pressure rating should be the maximum fracture
pressure of any formation penetrated by the well.
The implementation of these loads in the casing design process
is left to the user. Generally, one can easily show that the anticipated loads will not exceed the pipes elastic pressure capacity in
the absence of thermally induced axial strain, but this is not representative of field-operating conditions. No publicly available prescriptive guidelines are available for a more complete treatment
that includes a combined loading response. The complexities
associated with predicting pipe response during plastic deformation mean there is value in further fundamental work on the topic
and ongoing comparison with field experience.
March 2013 SPE Drilling & Completion

API Formulations. API Technical Report 5C3 (2008) contains


descriptions of both deterministic and probabilistic (synthesis)
approaches to pipe-body performance design. Both approaches
use closed-form or empirically augmented limit states equations
(in conjunction with specific geometry details and material properties of a sample) to calculate resistance to specified loads.
Design equations derived from the limit states equations are then
created by use of conservative assumptions or statistical distributions of input parameters to provide lower limits of the load that
can be sustained.
API Burst Formulations. API Technical Report 5C3 (2008)
describes internal pressure limits of pipes in the context of both
the triaxial yield condition and the ductile rupture limit:
 The triaxial yield condition uses the von Mises effectivestress basis for determining the onset of yielding in combined
loading situations involving internal pressure. In a conventional
elastic design context in which superposition remains applicable,
this treatment is sufficiently rigorous to account for string weight,
bending, torsion, and even elastic changes in loads associated
with axial confinement (i.e., triaxial stress effects and modest temperature changes). Because the elastic von Mises envelope is constructed on the basis of the API yield strength, slight amounts of
plasticity should be anticipated but are generally ignored for the
purposes of design; this is particularly valid when a safety factor
is used. The basis does not consider fatigue loading.
 The ductile rupture limit predicts the pressure at which pipes
will burst (or be subject to necking and other failure modes), and
thus deals with large-scale plasticity situations. This relationship
is seen to depend strongly on diameter and wall thickness, material imperfections, and material mechanical properties (including
ultimate strength and strain hardening behavior). Axial tension is
explicitly addressed.
API Collapse Formulations. Collapse behavior is generally
acknowledged [both in API Technical Report 5C3 (2008) and the
literature] to be more complex than burst behavior, and a
theoretically rigorous equation for collapse that accounts for all
pertinent variables and imperfections does not exist. API Technical Report 5C3 (2008) thus provides both some deterministic- and
some synthesis-based methods for collapse prediction:
 A previously developed approach, which was finalized in
1968 and is currently within the main body of the technical report,
references four collapse relationships that correspond to applicable ranges of pipe D/t ratio. Some of the associated design equations are referenced to empirical constants from testing, and all
equations reference the specified minimal yield strength of the
material with the exception of the elastic collapse equation (used
for very high D/t ratios not generally considered for thermal applications). Tensile axial stresses are accounted for through a derating procedure that is based on the von Mises yield criterion, and
the specification indicates a range of applicability that ends when
the equivalent yield strength is less than 165 MPa (24 ksi) (i.e., at
high levels of axial tension). Compressive axial stresses are not
explicitly acknowledged, but if the von Mises effective stress is
used as a basis, a higher collapse capacity is implied in biaxial
compression. This indirectly suggests that thermal tubulars operating beyond yield in compression should pose no collapse concern,
and provides no basis for predicting collapse behavior at axial
tensile stresses approaching or beyond the API yield strength.
Fig. 3 shows the basic relationship between tensile axial stress and
the collapse derating factor used in API Technical Report 5C3
(2008).
 An updated synthesis basis, included in Annexes to API
Technical Report 5C3 (2008), refers to a more rigorous review of
a large number of physical collapse test results and describes the
API work groups efforts to refine the data to more accurate,
95

800

1000
800

700

600
600

Hoop Stress (MPa)

Stress (MPa)

400
500

400

300

200
0
800

400

400

800

200
400

200
Engineering Stress-Strain
100

Plastic Strain:

600

0.2%

True Stress-Strain

800

4%

10%
0

0.02

0.04

0.06

0.08

0.1

0.12

Strain

(a) Stress/strain relationship

1000
Axial Stress (MPa)

(b) Yield surfaces at various equivalent plastic strains

Fig. 4Assumed material mechanical response characteristics for initial plane strain burst modeling.

ultimate collapse predictions. Combined loading is addressed in


Clause 4.2.2 of Annex F (Informative) of the technical report, but
it is based on von Mises and Tresca yield conditions; it does not
provide insight into behaviors that might be expected when axially yielded conditions exist.
Numerous other references discuss collapse behavior in combined loading. Pattillo and Huang (1982) acknowledge the impact
of stress/strain-curve shape on collapse capacity, and they formulate a model that incorporates post-yield behavior and provides
good agreement with experimental data for three different materials at subyield axial stress levels. The experimental work performed by Maruyama et al. (1990), which did provide roomtemperature test results at various axial stress conditions, indicates
N80 collapse performance at axial stresses less than yield that is
in good agreement with API equations. However, experimental
results with an as-rolled K55 material over a wide range of axial
tensile stresses (up to levels well beyond uniaxial yield stress)
indicated more collapse resistance than predicted by API equations, and the authors attributed this increase to the work-hardening behavior of the material. Klever and Tamano (2006) show
similar data for work-hardening K55 materials at axial stress levels exceeding initial yield in both tension and compression.
Discussion
On the basis of the preceding discussion, little guidance is available in API Technical Report 5C3 (2008) and in the available
literature regarding burst and collapse design with the combinedloadings condition typical to production casing in thermal applications. Numerical analysis investigating the fundamentals of
these behaviors has been conducted and is described here, along
with discussions of Shell Canadas pertinent field experience and
the relevance of available physical testing data. We note that the
work described here is a plasticity and stability study, and it does
not incorporate the effects of imperfections in the material.
Burst Assessment. Initial burst modeling was conducted with a
2D plane-strain finite element model with a large-deformation,
large-strain formulation to approximate the axially cement-constrained condition of the thermal production casing. The value in
a more rigorous treatment is discussed later in this paper. The key
behaviors of initial interest include the following:
 Plasticity response of axially yielded tubulars to net internal
pressure, both in the pressure region immediately beyond yield
and as the pipe approaches its rupture point
96

 Impact of initial axial stress state on the burst response,


because peak net internal pressure conditions could be incurred at
either minimal operating temperature (i.e., start of the steaming of
first or subsequent cycles) or maximum operating temperature
Material-Property and Pipe-Geometry Assumptions. For the
initial work described here, a relatively simple material model
was assumed.
 Fig. 4a shows the temperature-independent tensile curve that
was used to describe the (isotropic) material response, in both engineering stress/strain and true stress/strain spaces. The curve
approximates the room-temperature response of a minimal-yieldstrength L80 casing material, with a yield to tensile strength ratio
of 0.86 and an engineering strain of 10% at its ultimate strength
of 655 MPa.
 Fig. 4b shows corresponding current yield strength envelopes, assuming an isotropic hardening rule, at plastic strains of
0.2, 4, and 10% (the ultimate tensile strain of a uniaxial tension
coupon of this material).
With respect to pipe geometry, analysis was conducted using a
nominal-geometry 177.8-mm, 34.2-kg/m casing (nominal D/t
ratio of 22) with a worst-case wall-thickness variation (12.5%
less than API nominal). This extreme eccentricity was assumed
to ensure that the burst would start at a predictable location and
not be delayed by the perfect geometry of a finite element
analysis (FEA) model of a casing with no eccentricity. For the basic evaluation considered here, this treatment is considered
sufficient.
Combined Load Path. In field CSS operations, two internalpressure situations may arise in combination with axial strain:
 During steaming (i.e., in axial compression), the pipe is pressurized internally and may have less reservoir pressure acting on
it from the outside.
 After cooling (i.e., in axial tension), a pressure test may be
conducted to assess the integrity of the casing. This is typically
performed to an internal pressure of 7 MPa, with less reservoir
pressure acting on the pipe outside diameter (OD).
For initial exploratory burst analysis work, the casing was
assumed to undergo the axial mechanical straining of the magnitude experienced during steaming operations at Peace River
(0.5% strain in tension or compression) before internal pressurization. This initial axial strain condition is achieved in the model by
applying a uniform volumetric expansion or contraction while
maintaining a constant (zero) total axial strain. Internal pressure is
subsequently applied while maintaining a total axial strain of
March 2013 SPE Drilling & Completion

1000

1000
5

800

600
400
200

2
1

0
200
400

Initial von Mises yield


stress envelope

600
800
800

600

400

200
0
200
Axial Stress (MPa)

400

600

800

Fig. 5Predicted progression of hoop and axial stress to rupture point in a pipe sample subjected to initial axial compressive strain and subsequent internal pressure (with axial cement
constraint).

zero. Analysis of other load paths (e.g., internal pressure preceding temperature change) indicated fundamental responses similar
to those described in the subsequent paragraphs.
Burst Analysis Results. Initial analyses illustrate how dramatically the axial stress condition in the pipe changes when hoop
stress is applied. Fig. 5 shows the evolution of hoop and axial
stresses at the outer surface of the pipe in the 2D stress space,
with the initial von Mises envelope as a reference, when the casing is taken through five points in a combined axial compression/
pressure loading sequence:
1. Application of compressive axial mechanical strain of 0.5%,
generating an associated compressive stress of 552 MPa. The
axial stress increases along the x-axis of the chart until the yield
strength is reached (at a strain of approximately 0.28%). Because
the material being used for the analysis has a yield plateau characteristic, small amounts of incremental strain do not result in
appreciable material hardening or associated increments in axial
stress.
2. At an internal pressure of 14 MPa, which corresponds to the
peak differential that might be expected in operation at Peace
River.
3. At an internal pressure of 43.7 MPa, corresponding to the
limit defined by the API historical yield equation on the basis of
the initial yield strength and geometry of a 177.8-mm, 34.2-kg/m
L80 pipe with maximum allowable eccentricity.
4. At an internal pressure of 50 MPa, which corresponds to the
maximum pressure that might be elastically sustained by the same
pipe geometry in an optimal biaxial loading condition.
5. At the predicted rupture pressure of 63 MPa. This is an estimate of the peak pressure that can be sustained by the pipe.
It can be simply stated that the hoop stress imposed by internal
pressure causes plastic flow in the material and associated axial
stress changes that obey three conditions:
 Conservation of volume (i.e., incompressibility for steel
straining beyond yield)
 Plastic strain increments that occur in proportion to the
deviatoric stress components at any point in the loading sequence
(associated plastic flow law)
 Relationship between effective stress and plastic strain
described by the specified uniaxial material curve
The tendency for a free-ended pipe flowing plastically because
of internal pressure would be for hoop strain increments to be
counteracted by corresponding (and equal) decrements in radial
and axial plastic strain to conserve material volume. In the case of
an axially constrained pipe, the total axial strain remains zero
(i.e., length stays constant), and the radial and axial plastic strains
first accommodate this (according to the associated plastic flow
law) until the nose of the current von Mises ellipse is reached,
when tensile axial plastic strains can exactly correspond to onehalf of the tensile hoop strain imposed by the pressure increment
March 2013 SPE Drilling & Completion

initial compression, then pressure


initial tension, then pressure

600

Hoop Stress (MPa)

Hoop Stress (MPa)

800

400
200
0
200
Initial von Mises
yield stress
envelope

400
600
800
800

600

400

200
0
200
Axial Stress (MPa)

400

600

800

Fig. 6Comparison of hoop and axial stress responses during


pressurization for cemented pipes subjected to initial tensile
and initial compressive axial conditions.

(e.g., at Points 4 and 5 in Fig. 5, the ratio of hoop stress to axial


stress is almost exactly 2:1).
We note that when stresses reach the nose of the von Mises
ellipse, further increments in pressure can be sustained only if the
material hardens (i.e., if the material can support higher effective
stresses). Such hardening is possible only when the plastic strain
in the material increases, and the associated increment in strain
(and the eventual rupture limit) will be associated with the materials post-yield stiffness characteristic.
Fig. 6 shows the predicted 2D principal stress progression
from a second analysis, with the application of axial tensile strain
before internal pressurization, and compares it with the result
obtained with initial axial compressive strain. In the tensile case,
pressure loading initially reduces the effective stress (i.e., the material is no longer on the yield surface). This elastic state is sustained (and reflected by the ratio of axial stress to hoop stress of
0.3) until sufficient pressure causes the stress state to reach the
yield ellipse again at a hoop stress of approximately 300 MPa.
From this point on, plastic flow governs the change in axial stress
with imposed hoop stress until the same condition is approached
as in the initial axial compression case, in which incremental axial
stresses are approximately half the incremental hoop stress. The
rupture pressure in both cases is approximately 63 MPa. This reinforces Klever and Stewarts (1998) assertion that the initial axialstress condition is almost irrelevant to the ultimate burst point for
a cemented casing, but in this case for the situation in which the
initial axial loading is strain-based. From a safety factor standpoint, this preliminary, somewhat rudimentary analysis indicates
an ambient-temperature rupture limit more than four times the
maximum anticipated internal operating pressure at Peace River.
A more refined estimate of this ratio will use the elevated-temperature static yield strength of the material in use.
After a closer comparison of the two analysis results described
in the preceding paragraphs, one can note a difference in the
responses. The effective plastic strain incurred when the casing
starts in tension tends to be lower than that seen at the same pressure when the casing starts in compression, which is shown in
Fig. 7. This is not surprising because the effective stress initially
drops as the internal pressure is applied in the tensile case; the
pipe stays more elastic in biaxial tension. A similar comparison
could be conducted for other load paths that may mobilize more
plastic strain, including pressurization to operating pressure
before temperature change (axial strain), and multicycle analysis.
Attention to plastic strain development should be maintained
while more is learned about the plastic-strain sensitivity of sulfide-stress-cracking (SSC) behavior in thermal tubular materials.
Load Path and Strain Localization Considerations. A more
rigorous analysis could consider the effects of different loading
paths and of axial variations in both cross-sectional geometry and
material strength, primarily to refine the prediction for the rupture
pressure. However, we believe the simplified analyses described
97

1.0%
initial axial compression

Effective Plastic Strain

initial axial tension

0.8%

0.6%

0.4%

0.2%

0.0%

10

20
30
40
Internal Pressure (MPa)

50

60

Fig. 7Comparison of effective plastic strain during pressurization for cemented pipes subjected to initial tensile and compressive axial strain conditions.

in the preceding paragraphs characterize the burst behavior (to the


degree necessary at this point) for the range of D/t ratios, materials, and operating conditions. If more accurate predictions of local
plastic strains are required, a more complete analysis treatment
could be pursued.
Field Burst Performance. Thus far, Shell Canada Energys
experience at Peace River substantiates the findings shown here.
No burst issues have been observed during thermal operations in
compression at steaming operating pressures of 12 to 14 MPa and
in tension when pressure testing the casing with water and a surface pressure of 7 MPa.
Collapse Assessment. The collapse assessment conducted for
this work is somewhat more rigorous in its treatment of pipe
response than the preceding burst assessment, both with respect to
the completeness of the model and the extent of exploration. Very
little, if any, of the preceding burst discussion should be considered applicable for collapse. The assessment described in the following paragraphs provides examples of collapse sensitivities that
are intended both to provide some context for existing designs
and to safely enable design modifications in the event that anticipated well or operating parameters change. Parameters thought to
be relevant include loading path, pipe D/t ratio, initial ovality, material mechanical properties (yield and post-yield), wellbore curvature, and the amount of free pipe available to localize strain
into the ovalizing section. The effects of local wall thickness variations resulting from drilling wear and ovalization resistance
offered by cement were ignored for these analyses. The assumption regarding lack of cement support was made so that the analyses clearly demonstrate the pipe body collapse sensitivities and

focus any pipe body design basis development on deformationtolerant casing structures.
A 3D finite-element model with continuum elements was used
to predict collapse response with combined loading and its parametric dependencies. This model incorporates the initial stress
state from an imposed uniform curvature, and includes lateral support to maintain this curvature when axial strains are imposed.
Results presented here are generally monotonic in nature (in other
words, they do not consider cyclic response), but the modeling
strategy could be used in conjunction with cyclic material descriptions and loading paths to establish multicycle collapse response.
Material Property and Pipe Geometry. A simplified, temperature-dependent material formulation was used for the collapse
analyses described here. Materials were assumed to have temperature-sensitive yield strengths with constant post-yield stiffness
(i.e., bilinear stress/strain characteristics) at each temperature.
Past work by the authors (DallAcqua et al. 2005; Kaiser 2009)
acknowledges the temperature- and strain-rate-dependence of
these properties, but incorporation of more rigorous descriptions
is reserved for subsequent work, if deemed necessary. Engineering stress/strain curves and an associated tabular summary of the
baseline material used for analysis, intended to be roughly representative of an L80 material, are shown in Fig. 8.
Two pipe geometries were used for the demonstration analyses
presented herein. The first is a 177.8-mm, 34.2-kg/m casing typical of many of Shells Peace River wells; this configuration has a
nominal D/t ratio of 22. For comparison purposes, a 244.5-mm,
53.6-kg/m configuration was also analyzed as a higher-D/t-ratio
casing (nominal 27.3) that might be considered as an economical
choice for SAGD operation. The D/t range represented by these
configurations is indicative of the approximate range currently
being considered for use in western Canada, and enables clearer
sensitivity demonstrations than if only Shells casing configuration were analyzed. For brevity of the explanations that follow,
only the cases demonstrating key trends have been described.
For each analysis, cross-sectional geometry was assumed to
have a prescribed initial ovality (baseline value 1.5%, the maximum possible given API OD specifications), minimum linear
weight allowed by API (3.5% less than nominal), and maximum
eccentricity. For most of the cases described here, the pipe was
assumed to have a fixed total length (i.e., cement constraint at the
couplings), although some sensitivity to free length of stronger
pipe adjacent to the weaker ovalizing section is also included in
the subsequent Strain Localization section. Ovalization is triggered at the center plane of axial symmetry (at the middle of the
joint) by a slightly increased initial ovality relative to the adjacent
pipe material.
Load Paths. Load path dependencies are a key focus of the
collapse study described here, and four paths (reflecting heating
load paths) are initially identified to bracket the range of possible
responses and to provide a field-representative path:

600

Stress (MPa)

500
400
Temperature Yield Strength Post-Yield Stiffness
(C)
(MPa)
(MPa)

300
200
20C
180C
270C
350C

100
0
0.0%

0.5%

1.0%
Strain

1.5%

20
180
270
350

552
487
451
419

1000

2.0%

Fig. 8Baseline L80 engineering stress/strain curves used for collapse assessment.
98

March 2013 SPE Drilling & Completion

Effective Plastic Strain at Pipe OD

4.0%
Pipe Ovality

1.0%

Load Path 1: temperature, then


6.9 MPa pressure
Load Path 2: 6.9 MPa pressure,
then temperature
Load Path 3: saturation condition
Load Path 4: field-representative

3.0%
6.9 MPa

2.0%
0 MPa

1.0%
increasing compressive axial mechanical strain

0.0%

50

100

150

200

250

300

0.8%

4.0%

3.0%

0.6%

Full crosssection yield


temperature

0.4%

2.0%

1.0%

0.2%

350

Temperature (C)

Fig. 9Load-path dependence of ovalization response of


244.5-mm, 53.6-kg/m casing in 10 /30-m dogleg.

 Application of net external pressure after heating of the pipe


to 350 C. This situation could occur as the hot CSS well is drawn
down at the start of the production cycle (i.e., after soak period).
 Application of net external pressure before heating of the
pipe, with no internal pressure. This may be an overly conservative estimation of field conditions, but could be used as a conservative case for design.
 Application of net external pressure according to the pressure/temperature relationship for saturated steam, with no internal pressure. Although not necessarily representative of a field
condition, this is an intermediate loading path that we anticipate
will provide a different response than Paths 1 and 2, which
involve sequential application of loading.
 Field-representative heating path: Application of net external
pressure before substantial heating of the pipe, and subsequent internal pressure equalization. Shells collapse design loading path
for Peace River wells assumes the presence of a trapped pocket of
water on the outside of the casing, per the IRP3 (IRP 2002) recommendation. Thermal expansion of liquid water generates pressure
sufficient to fracture the formation with only modest changes in
temperature (e.g., 12 MPa pressure can be generated in a fully confined water pocket with a temperature change of less than 40 C).
This path thus assumes 12 MPa of pressure applied early in the
heating process and then increases internal pressure per the steam
saturation curve when higher steam temperatures are sustained
down the wellbore. It also includes a conservative provision for internal wellbore pressure (4 MPa) at the depths required to sustain
an external pressure of 12 MPa before formation fracture.
All four of these loading paths relate to casing collapse performance during heating, which imposes axial compressive mechanical strain (and thus stress) on the cemented casing according
to the relationship described in Eq. 1. The analysis of collapse
during cooling (i.e., in axial tension) is discussed subsequently.
To provide an optimal demonstration of sensitivities for this
analysis, net external pressure magnitudes for Load Paths 1
through 3 were different for the two casing sizes analyzed:
 For the 244.5-mm, 53.6-kg/m casing (high D/t configuration), a maximum net external pressure of 6.9 MPa was assumed,
corresponding to the steam saturation pressure at an operating
temperature of 285 C, which is representative of a high-temperature SAGD operation. For L80 material, this casing is within the
D/t range governed by the API transition collapse equation, which
predicts a collapse pressure (at nominal wall thickness) of 14.4
MPa at ambient temperature with no axial stress.
 For the 177.8-mm, 34.2-kg/m casing (field-representative
configuration), the maximum net external pressure of 8 MPa
anticipated in Peace River tubulars was assumed. For L80 material, this casing is within the D/t range governed by the API plastic collapse equation, which predicts a collapse pressure (at
nominal wall thickness) of 26.4 MPa at ambient temperature with
no axial stress.
Load Path 4 was identical for both casing configurations.
March 2013 SPE Drilling & Completion

5.0%

Strain at Pipe Intrados


Strain at Pipe Neutral Axis
Strain at Pipe Extrados
Ovality

Pipe Ovality

5.0%

0.0%

50

100
150
200
Temperature (C)

250

0.0%
300

Fig. 10Predicted plastic-strain development at various points


in the ovalizing cross section of 244.5-mm, 53.6-kg/m L80 casing in 10 /30-m dogleg with Load Path 2.

Collapse Sensitivity Analysis Results. The results presented


in the following paragraphs demonstrate both the fundamental
collapse response of constrained pipes in thermal conditions and
the sensitivities to certain design and operating parameters.
Load Path DependenceHeating. Basic characteristics of
thermal casing collapse response in axial compression are demonstrated here by comparing the deformation response of the 244.5mm, 53.6-kg/m L80 casing subjected to the four load paths
described in the preceding paragraphs. By definition, the final
deformation state in a conventional elastic design with external
pressure would not exhibit any substantial sensitivities to the
sequence of loading, because superposition is appropriate in the
elastic range. In the thermal casing, however, the sequence of
loading is seen to have a large influence on the response. Fig. 9
shows the progression of pipe ovality for the four candidate load
paths, and it indicates the following:
 Load Path 1 (temperature change, then pressure change) exhibits very little ovalization during heating (i.e., application of compressive axial strain) until the external pressure is applied. It results in
the lowest final pipe ovality of the four load paths considered.
 Load Path 2 (pressure change, then temperature change)
exhibits considerable ovalization. When the temperature reaches
approximately 230 C, the ovality increases dramatically; this is
the onset of full-section collapse.
 Load Path 3 (pressure change during temperature change)
results in an intermediate amount of ovalization and a full-section
collapse at approximately 290 C.
 Load Path 4 (field-representative pressure and temperature
changes) also indicates a full-section collapse, occurring at a
slightly lower temperature than for Load Path 2 because the fieldrepresentative pressure differential (8 MPa) is slightly higher than
the differential pressure for Load Path 2.
The collapse behavior demonstrated in Fig. 9 is related to plastic flow at zones in the pipe cross-section that are more highly
stressed when external pressure is applied. The initial pipe ovality, in combination with the uniformly applied external pressure,
serves to create zones of higher through-wall bending stress, particularly at the major and minor axes of the slightly ovalized pipe.
When the pipe is elastic and the ovality small, these stresses
remain small and do not have much effect. However, when the
pipe stress reaches the yield surface, its stiffness changes dramatically, and any increments in thermally induced axial strain or
pressure lead to plastic flow of the pipe material.
Fig. 10 shows the progression of plastic strain at three key
points in the ovalizing cross section of the pipe for Load Path 2
(pressure, then temperature), which resulted in a full-section collapse at 230 C. Material at the intrados, neutral axis, and extrados
of the pipe initially yields at temperatures of approximately 140,
180, and 230 C, respectively. When the extrados material yields
99

5.0%

800
initial yield envelopes:
20C
230C

4.0%

400
200
0
nominal hoop stress
Intrados OD
Intrados ID
Neutral Axis OD
Neutral Axis ID
Extrados OD
Extrados ID

200
400
600
800
800

600

400

200
0
200
Axial Stress (MPa)

400

600

800

Fig. 11Progression of hoop and axial stresses during ovalization at various points in the ovalizing cross section of 244.5mm, 53.6-kg/m L80 casing in 10 /30-m dogleg with Load Path 2.

(i.e., full cross-section yield), the rate of plastic straining at all


three points increases substantially and so does the ovality of the
pipe (which references the vertical axis on the right of the chart).
Results shown are extrapolations to the outer surface of the pipe
from integration points in the finite element model and are therefore approximate. Note that the local strain magnitudes in the
ovalizing section are relatively low (less than 1%) when the section begins to ovalize substantially; this indicates that the zone of
interest in the material stress/strain curve relevant to collapse is
also in the low-strain range.
Fig. 11 complements Fig. 10 by showing the progression of
approximate local biaxial stress components at the inner and outer
surfaces of the pipe during axial straining. Each curve indicates
the stress progression to a pipe ovality of approximately 5%, and
is referenced to the initial (unstrained) biaxial yield envelopes at
20 and 230 C.
The stress components at the intrados, extrados, and neutral
axis after application of pressure (and before axial straining,
shown in the figure as circular and diamond-shaped markers) are
elastic but display some important differences:
 Initial levels of axial stress differ from the intrados to extrados because of the imposed pipe curvature. Small variations are
seen from the inside to outside surface of the pipe at a given circumferential position; these arise during pressure application.
 Initial levels of hoop stress vary substantially through the
wall thickness and are a strong function of orientation relative to
the oval shape of the pipe.
 Nominal (calculated) hoop stresses, with the magnitude of
external pressure acting on the pipe and the nominal pipe geometry, are approximately 100 MPa.
 Near-zero hoop stresses are seen at the intrados and extrados
on the inner surface of the pipe and at the neutral axis on the outer
surface of the pipe.
 Similarly, much larger compressive hoop stresses than predicted are initially seen at the intrados and extrados on the outer
surface of the pipe and at the neutral axis on the inner surface of
the pipe.
These hoop stress variations are the result of local bending
stresses imposed on the pipe wall by the combination of external
pressure and pipe ovality, and would be diminished significantly
with lower initial ovality or higher pipe D/t ratio.
After application of compressive axial strain after pressurization, the stresses at each key location are seen to increase elastically
until the yield surface is reached. Once full cross-section yielding
occurs, through-wall hoop stress variations increase dramatically
to react with the net collapse load as the pipe continues to ovalize.
In the context of existing collapse calculation and design
bases, the overall result shown here is quite surprising: For some
loading paths, a relatively low external pressure is shown to be
sufficient to significantly deform the pipe cross section, in combination with axial compression. Recall that the API collapse for100

Pipe Ovality

Hoop Stress (MPa)

600

3.0%

2.0%
P, then T

P, then T

1.0%
Axial Compression (Heating from 5C)
Axial Tension (Cooling from 350C)

0.0%

50

100

150

200

250

300

350

Temperature (C)

Fig. 12Comparison of ovalization responses in tensile (cooling) and compressive (heating) scenarios for 244.5-mm, 53.6kg/m L80 casing in 10 /30-m dogleg, with 6.9-MPa external
pressure applied before temperature in both cases. Analysis
assumes no net axial load at start of heating or cooling.

mulations do not explicitly reduce the pipe-collapse rating for


compression loads, but rather for tension loads.
A final comment regarding the fundamental response
described here is that the length of the ovalizing section predicted
by the analysis is actually quite short. For the example shown,
analysis results indicate a length of approximately 2 diameters is
substantially deformed in the ovalizing zone.
Cooling (Tension) Load Path Sensitivity. In field operations,
situations may arise in which the pipe is cooling (or cold) whereas
external pressure is acting on the pipe and the wellbore pressure is
low. Whereas pressures approaching the formation fracture point
are very unlikely to be retained in any trapped water pockets in
this situation, it is conceivable that the saturated steam pressure or
(at minimum) the formation pore pressure will remain. A single
analysis of pipe behavior during cooling was conducted not only
to enable a comparison between collapse responses when the casing is yielded in tension and compression but also to provide an
appropriate focus for future collapse design bases.
On the basis of API collapse derating factors for axial loading,
one might infer that the collapse susceptibility would be greater in
tension (i.e., when cooling) than in compression. Fig. 12 indicates
the ovalization response of the baseline 244.5-mm L80 pipe during
cooling as a function of temperature, in which the cooling path
involved initial external pressurization to 6.9 MPa at 350 C and
subsequent cooling. Little difference is seen in the character of the
cooling response and the heating response with the same loading
sequence (also shown in the figure). The overall temperature change
required to start substantial ovalization is roughly equivalent (220
to 230 C) and corresponds to the point of full cross-section yielding
in both cases. More importantly, the rate of ovalization after fullsection yielding is actually lower for the cooling path than for the
heating path. Furthermore, one specific modeling assumption made
here for the cooling scenario should be considered much more conservative than for the heating scenario: Whereas the pipe will
actually be at the yield point in axial compression when cooling
starts, the cooling analysis instead assumes no initial net axial load.
This means that the model provides a conservative (low) estimate
of the temperature change required to cause yielding (and accelerated ovalization) of the pipe in axial tension. Because of the results
of this analysis demonstration and the degree of conservatism used
in the model, all subsequent analyses shown here focus on the heating path as a key indicator of collapse stability.
Wellbore Curvature Dependence. Wellbore curvature imposes an initial stress state on the casing that has an influence on
when the pipe begins to yield, and when the full cross-section
yields. Onset of yield in a curved pipe can be expected earlier in
the heating process than for a straight pipe because a compressive
stress state already exists at the intrados of the pipe. Similarly,
full cross-section yielding occurs when the material at the
March 2013 SPE Drilling & Completion

5.0%

5.0%

straight pipe
10/30 m curvature

4.0%
Pipe Ovality

Pipe Ovality

4.0%

3.0%

2.0%

1.0%

0.0%

0.5% initial ovality


1.0% initial ovality
1.5% initial ovality

3.0%

2.0%

1.0%

50

100

150

200

250

300

350

Temperature (C)

0.0%

50

100

150

200

250

300

350

Temperature (C)

Fig. 13Curvature sensitivity of ovalization response of 244.5mm, 53.6-kg/m casing with Load Path 2 (pressure followed by
temperature).

Fig. 14Initial ovality sensitivity of ovalization response of


244.5-mm, 53.6-kg/m casing with Load Path 2 (pressure followed by temperature) in 10 /30-m dogleg.

extrados of the pipe is strained enough to yield in compression.


From a collapse performance standpoint, we are therefore interested in comparing the onset and acceleration of the ovalization
of curved pipes with those of tubulars in a straight section of wellbore. Fig. 13 shows the ovality progression for the 244.5-mm,
53.6-kg/m L80 pipe configuration in both straight and deviated
(10 /30-m) sections, and indicates that the collapse situation for
this demonstration is more severe for the pipe in the deviated section than for the straight pipe. We envision certain situations in
which this trend could be reversed, and we suggest that the range
of anticipated curvatures be considered for design purposes.
Initial-Pipe-Ovality Dependence. API Technical Report 5C3
(2008, p.140) acknowledges manufactured pipe ovality as one
potential factor in collapse performance. For the purposes of this
discussion, ovality is defined as

post-yield stiffness (at field-representative conditions) will improve the deformation-tolerance of a particular tubular configuration. The same is not always true for yield strength, in which the
increased strength can generate larger instability potential because
it effectively increases the axial force carried by the structure; an
example is thermal slotted liner (DallAcqua et al. 2005, 2010). A
small sensitivity study was executed to establish the impacts of
both yield strength and post-yield stiffness on ovalization performance of thermal production casing. We expect that materials
with higher post-yield stiffness will more readily harden to sustain
the bending moments incurred at the major and minor axes of the
oval-pipe geometry, and thus be more collapse-resistant. We
would therefore anticipate a range of suitable combinations of
yield strength, post-yield behavior, and D/t ratio to exist for a
given application, and suggest that any other relevant material
selection factors (such as SSC resistance) identify a subset of
these thermal-deformation-tolerant materials for field use.
The ovalization response of a single pipe configuration (244.5
mm in 10 /30-m dogleg, ovality 1.5%) was analyzed with five combinations of yield strength (Sy) and post-yield stiffness (Ep) and
the bilinear, temperature-sensitive constitutive model described
previously:
 Material 1 (baseline L80): Sy 552 MPa, Ep 1000 MPa
 Material 2 (perfectly plastic L80): Sy 552 MPa, Ep 0
MPa
 Material 3 (high-PYS L80): Sy 552 MPa, Ep 2000
MPa
 Material 4 (low-PYS K55): Sy 379 MPa, Ep 2000 MPa
 Material 5 (high-PYS K55): Sy 379 MPa, Ep 8000 MPa
Fig. 15 compares the ovalization response of the 244.5-mm,
53.6-kg/m casing configuration with the five material characteristics shown previously, and highlights the following key
responses:
 Pipes made from the baseline and perfectly plastic L80 materials (#1 and #2), which exhibit very little post-yield stiffness, are
predicted to collapse very soon after initial yielding occurs. This,
in conjunction with results for the other materials, means that
yield strength alone (i.e., grade) must not be used as the sole basis
for selecting candidate materials.
 The yield strength has a significant impact on the temperature
at which incremental ovalization becomes evident. For materials
with consistent post-yield stiffness and differing material yield
strengths, more collapse resistance is exhibited by the material with
higher yield strength. This can be attributed to the greater opportunity for plastic straining to occur for lower-yield-strength materials
in regions of the cross-section carrying through-wall bending
moment.
 Notwithstanding the preceding, the high-post-yield-stiffness
low-yield material (#5) exhibits a less sudden change in ovality
than all the L80 materials, and stays stable to higher temperatures
than both the perfectly plastic L80 and baseline L80 materials.

Ovality

2  Dmax  Dmin
: . . . . . . . . . . . . . . . . . . . 2
Dmax Dmin

On the basis of APIs allowable range of pipe ODs, the maximum permissible initial ovality for an as-manufactured pipe is
1.5%, and this would be expected only at the extreme end of the
statistical distribution. The predicted ovalization response of the
244.5-mm L80 casing configuration under collapse Load Path 2
(pressure, then temperature) in a 10 /30-m dogleg at initial ovalities of 0.5, 1.0, and 1.5% is shown in Fig. 14. A stark contrast is
evident in the progression of ovality for the three cases, with the
lowest-ovality pipe reaching a maximum temperature of 350
without collapsing and the highest-ovality pipe collapsing at a
temperature of 240 C. We attribute this difference to the magnitude of the through-wall bending moment (and the associated
hoop stresses) generated because of the net force ovalizing the
casing. For a consistent uniformly applied pressure load, elastic
hoop stresses associated with the net ovalization force will
increase approximately linearly with starting ovality at small
ovalities. As was shown in the preceding paragraphs, it is the evolution of the stress and plastic strain state in the most highly
stressed regions of the cross-section that results in progressive deformation and, ultimately, collapse. Lower initial ovality delays
the onset of this behavior, and the collapse tendency can thus be
expected to be greater for pipes with higher initial ovality. Similarly, higher-D/t pipes (which have the same geometric ovality
tolerance as low-D/t pipes) contain higher bending-related hoop
stresses than low-D/t pipes for the same applied pressure, and
they will be more susceptible to ovalization.
Material-Property Dependence. It is well understood that
the deformation-tolerance of thermal tubulars is a function of both
the yield and post-yield properties of the material, and that the
strain-based nature of the axial loading actually governs the forces
carried by the tubes. In general, one can expect that increased
March 2013 SPE Drilling & Completion

101

5.0%

4.0%
Pipe Ovality

4.0%
Pipe Ovality

5.0%

baseline L80
perfectly plastic L80
high-PYS L80
low-PYS K55
high-PYS K55

3.0%

2.0%

1.0%

3.0%

2.0%
177.8 mm Load Path 2: 8 MPa pressure, then temperature

1.0%

177.8 mm Load Path 4: field-representative


244.5 mm Load Path 2: 6.9 MPa pressure, then temperature

0.0%

50

100

150

200

250

300

350

Temperature (C)

0.0%

50

100

150

200

250

300

350

Temperature (C)

Fig. 15Material sensitivity of ovalization response of 244.5mm, 53.6-kg/m casing with Load Path 2 (pressure followed by
temperature) in 10 /30-m dogleg.

Fig. 16Load-path dependence of ovalization response of


177.8-mm, 34.2-kg/m casing in 10 /30-m dogleg and comparison
with response of 244.5-mm, 53.6-kg/m casing.

We feel the material sensitivities shown here, whereas quite


basic in nature, highlight the need to have both an understanding
of the impact of yield strength and post-yield behavior on collapse
resistance and an assurance of the minimal associated mechanical
properties of produced pipe that will be run in the field, at representative operating conditions. Although an explicit comparison
is not described here, we would expect some of the materialrelated collapse dependencies shown in the literature and in past
full-scale testing will relate to some of the same mechanical properties that influence thermal casing collapse performance, even
though at different service conditions. For thermal operations, further exploration could provide a more comprehensive indication
of the impact of both cyclic material properties and material anisotropy (i.e., differences between axial and circumferential properties) on casing performance.
Shell Field-Configuration Response. The analysis results
described in the preceding paragraphs demonstrated some of the
parameters influencing collapse response. Selected analyses were
also performed on the 177.8-mm, 34.2- kg/m casing configuration
to assess the collapse resistance of Shells field casing configuration. For the cases described next, maximum external differential
pressures of 8 MPa were assumed.
Initial analyses indicate the stability of 177.8-mm, 34.2-kg/m
casing with the baseline L80 material properties described previously (1000-MPa post-yield stiffness) and maximum ovality with
two of the most severe loading paths. In each case, the response
remains stable to 350 C, even though some ovalization is displayed
for the worst-case load path (#2, pressure then temperature), which
is shown in Fig. 16. The field-representative load path (#4) exhibits
an improved response because of reductions in assumed differential pressure at higher temperature. For comparison, the ovalization

response of the 244.5-mm, 53.6-kg/m casing configuration with the


same material properties and worst-case load path is also shown in
Fig. 16.
Material sensitivities shown earlier for 244.5-mm casing suggested that higher-yield-strength materials would generally enable
collapse resistance to higher temperatures, but higher post-yield
stiffness was also shown to reduce the rate of ovalization (and prevent instability) after full cross-section yielding occurred. This comparison was repeated for the lower-D/t-ratio (177.8-mm) casing to
establish differences in material sensitivities for the more robust
pipe body configuration used at Peace River. Fig. 17 indicates the
ovalization responses for the five materials described previously,
and indicates minimal sensitivity of ovalization response to yield
strength and almost no differences with post-yield stiffness. This is
perhaps not surprising because the lower D/t ratio will substantially
reduce initial through-wall (local) bending stresses imposed by pressurizing the slightly ovalized pipe. These findings indicate that a
broader range of material properties will be acceptable for low-D/t
pipes from a collapse perspective, and that material properties
should be managed with more care if high-D/t pipes are used.
Strain Localization. An area demanding attention in thermalrecovery applications is the nature of pipe/cement interactions in
the context of strain localization. For the analyses described here,
we have assumed that the couplings at either end of the joint of interest are fixed axially, and that local redistribution may happen
within the interval. Production variations in pipe cross-sectional
characteristics and material mechanical properties will exist and,
because of associated strength contrasts, will lead to variability in
axial strain along the string length; see Nowinka et al. (2008) for a
more in-depth discussion. If significant variations occur between
adjacent joints, pipe/cement interactions will limit the lengths
along which strain may redistribute. The valid length (here, termed
free length of localization) for such redistributions has not been
quantified, but we include an example of the impact of such redistribution on collapse response here.
Fig. 18 shows the impact of the free length of localization on the
ovalization of the 244.5-mm, 53.6-kg/m casing subjected to 6.9MPa external pressure along Load Path 2 in a 10 /30-m dogleg,
along with an indication of the average amount of axial strain
imposed on the weaker (ovalizing) section of pipe. The adjacent
pipes are conservatively assumed to have a yield strength 15%
higher than the weaker pipe, and will thus stay elastic until the
weaker pipe hardens enough to cause yielding of the strong pipe.
The clear result is that the longer free length greatly reduces the
pipes ovalization resistance. We anticipate this example is conservative, but indicative of a variable that should be more fully characterized and understood as thermal casing analysis and design bases
continue to evolve.
Field Collapse Performance. No collapse issues have been
observed in Shell Canada Energys operating experience at Peace
River during thermal operations.

5.0%

baseline L80
perfectly plastic L80
high-PYS L80
low-PYS K55
high-PYS K55

Pipe Ovality

4.0%

3.0%

2.0%

1.0%

0.0%

50

100

150

200

250

300

350

Temperature (C)

Fig. 17Material sensitivity of ovalization response of 177.8mm, 34.2-kg/m casing with Load Path 2 (pressure followed by
temperature) in 10 /30-m dogleg.
102

March 2013 SPE Drilling & Completion

5.0%

One joint free length


Three joint free length

Pipe Ovality

4.0%

3.0%

2.0%

1.0%

0.0%

50

100

150

200

250

300

350

Temperature (C)

Fig. 18Impact of free length of localization on collapse


response of 244.5-mm, 53.6-kg/m L80 casing in 10 /30-m dogleg
with Load Path 2 (pressure, then temperature).

Discussion of Available Collapse-Testing Data. Very little


published information is available that can be directly compared
with the collapse analysis results described here. This is, in large
part, because the relevant combined-loading condition is unique
to thermal operating conditions, and much of the available testing
data are at subyield axial stress levels. Furthermore, the load-path
dependence demonstrated in Fig. 9 indicates that the testing
sequence is likely to have a large impact on the observed collapse
behavior.
Maruyama et al. (1990) conducted room-temperature collapse
tests on quenched-and-tempered N80 and as-rolled (and subsequently annealed) K55 casings in a biaxial collapse testing
machine. The N80 material exhibited mechanical response with a
yield plateau and high Y/T ratio, whereas the K55 material had a
relatively low yield strength and considerable post-yield stiffness.
Pipes that were tested were also selected or machined to have
minimal initial ovality. The paper indicates axial load was applied
to the test sample when external pressure was increased, but an
exact path is not described. Regardless, the authors concluded that
the N80 test results correlated well with API predictions at axial
tension levels approaching 75% of the uniaxial yield point. The
K55 collapse results (performed at D/t ratios of 18.5 and 28) correlated well with API calculation bases at low axial tensions, but
much more collapse resistance was shown than API calculations
would suggest at higher axial tension levels (well beyond uniaxial
yield stress). The authors attributed this improved performance
to the substantial work hardening (post-yield stiffness) of the
material.
Klever and Tamano (2006) also briefly reference biaxial collapse tests on 177.8-mm, 38.7-kg/m K55 material at axial stresses
beyond yield in both compression and tension, and provide comparisons with shell theory and a new collapse equation. The test
data shown in the paper indicate collapse capacities similar to
those of Maruyama et al. (1990) at axial stress levels outside the
elastic range of the material.
Even though not enough details are presented in the references
to provide a rigorous comparison with the analysis results shown
herein, we generally concur with Maruyama et al.s (1990) suggestion that for the K55 material is related to its post-yield-stiffness characteristics. Because of the load-path sensitivities shown
herein, we suggest that the load path be considered a second important feature, and encourage the ongoing use of a field-representative worst-case load path for design purposes. Furthermore, the
analysis results presented here indicate that the starting pipe geometry (in particular, ovality) affects hoop-stress variations in the
cross section, and will have significant effect on collapse
tendency.
Conclusions and Design Implications
The analysis results described in this paper demonstrate that thermal production casing has a substantially different set of burst and
March 2013 SPE Drilling & Completion

collapse performance dependencies than those readily derived


from existing API equations and the prior literature. They highlight an opportunity to assess risk and to develop new guidelines
for thermal casing design that will bring more rigor to the process.
We see this as particularly useful when changes to existing well
designs or to operating processes are being considered, and this
will also enable operators to define specifications for pipe materials and geometries that are customized to the needs of their applications. In addition to the pipe-body design aspects discussed
here, operators must use rigor in assuring environmental cracking
resistance of the tubular material and connection integrity and
sealability during thermal-service loading.
Following are the authors key conclusions from the work
described here:
 Burst. Analysis indicates that a typical thermal well-casing
design (D/t of 22) subjected to thermally induced axial strain in
combination with internal pressure is very robust from a burst perspective. The casing rupture point in cemented casings is independent of the initial axial stress and the thermally induced axial
mechanical strain.
 Collapse. Analysis shows that thermal well-casing designs
may be susceptible to casing collapse at higher D/t ratios. FEA of
244.5-mm, 53.6-kg/m L80 casing (D/t of 27) indicated a collapse
concern at thermal operating temperatures with a net external
pressure of approximately 7 MPa, whereas similar analysis and
field performance of 177.8-mm, 34.2-kg/m L80 casing (D/t of 22)
indicate good collapse resistance at Peace River operating conditions. Collapse behavior clearly depends on the assumed load path
(i.e., sequence of temperature and external pressure), and the use
of a field-representative load path is strongly encouraged in the
design process. Material mechanical properties at field-representative conditions can also play a large role in collapse resistance of
thermal tubulars.
Burst Design. The pipe-body burst behaviors shown here and
past literature pertaining to rupture limits indicate that thermal tubulars operating in an axially yielded condition will generally have
substantial incremental burst capacity beyond the operating pressures that are typical of western Canadian thermal applications.
We suggest that a reasonable reference pressure for safety factor
calculations is the internal yield pressure for a free-ended pipe
(i.e., no axial tension or compression) on the basis of the minimum
anticipated static strength properties of the steel at maximum operating temperature, because this represents a conservative estimate
of the pressure that would cause the pipe to burst. Thermal well
operators could base their burst design decisions on an appropriate
safety factor relative to the maximum differential pressure that can
be expected in the casing during operation. Our suggestion is that,
in addition to the simple burst pressure rating requirements outlined in IRP3 (IRP 2002), the ratio of the reference pressure to the
maximum anticipated differential pressure during operation is a
factor of at least 1.5. If it is shown that small amounts of plastic
strain will substantially influence the SSC potential in candidate
materials, this recommendation should be reconsidered, and a
more detailed analysis of the behavior undertaken.
The impacts of internal pressure on structural and sealability
performance of thermal-casing connections should be considered
separately through an appropriate evaluation basis.
Collapse Design. Collapse design for thermal casing is clearly
more complex than burst design, and will demand more rigor than
a simple safety factor on the basis of material strength at operating
conditions and calculations of elastic limits. We believe it appropriate for the casing design basis to rely on the cement only for
global support (i.e., global buckling prevention), and that the
roundness of the pipe during operation be ensured by its deformation-tolerance when exposed to representative design loading
conditions.
For Shells application, the 177.8-mm, 34.2-kg/m L80 casing
configuration (with nominal D/t of 22) is predicted to provide
relatively good collapse resistance during field-representative
103

loading, and field experience thus far has substantiated these predictions. Given the sensitivity to post-yield material properties
that has been demonstrated here, we attribute some aspects of this
success to the materials used.
Less severe temperature- and pressure-loading conditions might
enable the use of pipes with higher D/t ratios, and other material
formulations may provide similar or improved collapse resistance.
Some rigor should be used in the tubular selection process to
ensure that reduced collapse resistance offered by higher-D/t pipes
does not offset any associated economic gains. To accommodate
production variability in pipes and to enable qualitative comparison of candidate products, operators could establish suitable
combinations of D/t ratio, pipe ovality tolerance, and elevatedtemperature material mechanical properties (yield strength and
post-yield characteristics), and define associated minimal product
specifications.
Nomenclature
D pipe diameter
Dmax maximum pipe OD
Dmin minimum pipe OD
Ep steel post-yield stiffness
Sy yield strength
t pipe thickness
emech mechanical strain
ethermal thermal strain
etotal total strain

Acknowledgments
The authors wish to thank their respective organizations for permission to publish this paper. In addition, we thank Bruce Lepper
(formerly with Shell Canada Energy) for his input and substantial
contribution to this work.
References
ANSI/API. 2008. Technical Report on Equations and Calculations for
Casing, Tubing, and Line Pipe Used as Casing or Tubing; and Performance Properties Tables for Casing and Tubing. Technical Report
5C3, First Edition, December.
DallAcqua, D., Lopez-Turconi, G., Monterrubio, I. et al. 2010. Development of an Optimized Tubular Material for Thermal Slotted Liner
Completions. J. Cdn. Pet. Tech. 49 (2): 1522. http://dx.doi.org/
10.2118/132640-PA.
DallAcqua, D., Smith, D.T., and Kaiser, T.M.V. 2005. Thermo-Plastic
Properties of OCTG in a SAGD Application. Paper SPE/PS-CIM/
CHOA 97776 presented at the International Thermal Operations and
Heavy Oil Symposium, Calgary, Alberta, Canada, 13 November.
http://dx.doi.org/10.2118/97776-MS.
Directive 010: Minimum Casing Design Requirements. 2009. Calgary,
Alberta, Canada: Energy Resources Conservation Board (ERCB).
IRP. 2002. Volume #3Heavy Oil and Oil Sands Operations. Calgary,
Alberta, Canada: DACC/Enform.
Kaiser, T.M.V. 2009. Post-Yield Material Characterization for StrainBased Design. SPE J. 14 (1): 128134. http://dx.doi.org/10.2118/
97730-PA.
Kaiser, T.M.V., Yung, V.Y.B., and Bacon, R.M. 2008. Cyclic Mechanical
and Fatigue Properties for Oil-Country Tubular-Goods Materials. SPE
J. 13 (4): 480486. http://dx.doi.org/10.2118/97775-PA.
Kalil, I.A. and McSpadden, A.R. 2011. Casing Burst Stresses in Particulate-Filled Annuli: Wheres the Cement? Paper SPE 139766 presented
at the SPE/IADC Drilling Conference and Exhibition, Amsterdam,
The Netherlands, 13 March. http://dx.doi.org/10.2118/139766-MS.

104

Klever, F.J. and Stewart, G. 1998. Analytical Burst Strength Prediction of


OCTG With and Without Defects. Paper SPE 48329 presented at the
SPE Applied Technology Workshop on Risk-Based Design of Well
Casing and Tubing, The Woodlands, Texas, 78 May. http://
dx.doi.org/10.2118/48329-MS.
Klever, F.J. and Tamano, T. 2006. A New OCTG Strength Equation for
Collapse Under Combined Loads. SPE Drill & Compl 21 (3):
164179. http://dx.doi.org/10.2118/90904-PA.
Maruyama, K., Tsuru, E., Ogasawara, M. et al.1990. An Experimental
Study of Casing Performance Under Thermal Recovery Conditions.
SPE Drill Eng 5 (2): 156164. http://dx.doi.org/10.2118/18776-PA.
Nowinka, J., Kaiser, T.M.V., and Lepper, B. 2008. Strain-Based Design of
Tubulars for Extreme-Service Wells. SPE Drill & Compl 23 (4):
353360. http://dx.doi.org/10.2118/105717-PA.
Pattillo, P.D. and Huang, N.C. 1982. The Effect of Axial Load on Casing Collapse. J. Pet Tech 34 (1): 159164. http://dx.doi.org/10.2118/9327-PA.
Rodriguez, W.J., Fleckenstein, W.W., and Eustes, A.W. 2003. Simulation
of Collapse Loads on Cemented Casing Using Finite Element Analysis. Paper SPE 84566 presented at the SPE Annual Technical Conference and Exhibition, Denver, Colorado, 58 October. http://dx.doi.org/
10.2118/84566-MS.
Zinkham, R.E. and Goodwin, R.J. 1962. Burst Resistance of Pipe
Cemented into the Earth. J. Pet Tech 14 (9): 10331040. http://
dx.doi.org/10.2118/291-PA.

SI Metric Conversion Factors


in.  2.54*
E00 cm
ft  3.048*
E01 m
ksi  6.894 757
E03 kPa
lbf  4.448 222
E00 N
psi  6.894 757
E00 kPa
*Conversion factor is exact.

Dan DallAcqua is a senior consultant and engineering manager at Noetic Engineering in Edmonton, Alberta, Canada.
He has worked for more than a decade establishing and
evolving mechanical- engineering design bases for a wide variety of applications (including thermal-well tubulars) through
analytical and experimental investigation. DallAcqua holds
BSc and MSc degrees from the University of Alberta and is a
member of SPE, the Canadian Heavy Oil Association (CHOA),
the IRP3 Well Design Group, and a number of API working
groups focused on extreme-service equipment.
Mark Hodder has worked as a drilling engineer for Shell Canada in Calgary, for 3 years. For the last 2 years, he has been
involved in thermal (CSS) infill development within Shells
Peace River operations. Hodder holds bachelor degrees in
mechanical engineering and chemistry from Dalhousie University, Halifax, Nova Scotia.
Trent Kaiser is a principal consultant at Noetic Engineering in
Edmonton. He holds BSc, MSc, and PhD degrees in mechanical engineering. Kaiser has made many contributions to the
development of strain-based design methodologies for
unconventional resources, including thermal wells, tight gas
and oil, and compacting formations. His publishing record
includes seven peer-reviewed papers and numerous conference papers, and he is also the holder of more than 30 patent
grants. Kaiser served in various roles in CHOA over the past
12 years, including that of president in 20082009. He is a
lead developer and presenter in industry training courses for
the unconventional-resource industry on topics ranging from
well structures to strain-based design to flow control and
optimization.

March 2013 SPE Drilling & Completion

Vous aimerez peut-être aussi