Sie sind auf Seite 1von 27

SPE-173378-MS

Applications of Geomechanics to Hydraulic Fracturing - Case Studies from


Coal Stimulations
Vibhas J. Pandey, ConocoPhillips and Thomas Flottmann, Origin Energy

Copyright 2015, Society of Petroleum Engineers


This paper was prepared for presentation at the SPE Hydraulic Fracturing Technology Conference held in The Woodlands, Texas, USA, 35 February 2015.
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 does not necessarily reflect
any position of the Society of Petroleum Engineers, its officers, or members. 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 SPE copyright.

Abstract
Modern hydraulic fracture treatments rely heavily on the implementation of formation property details
such as in-situ stresses and rock mechanical properties, in order to optimize stimulation designs for
specific reservoir targets. Log derived strain and strength calibrated in-situ properties provide critical
description of stress variations in different lithologies and at varying depths. From a practical standpoint
however, most of the hydraulic fracture simulators that are used for fracturing treatment design purposes
today can accommodate only a limited portion of a geologic-based rock mechanical property characterization which targets optimal data integration thus resulting in complexity. By using examples from
hydraulic fracture stimulations of coals in a complex but well characterized stress environment (Surat
Basin, Eastern Australia) we distil out the reservoir rock related input parameters that are determinants of
hydraulic fracture designs and identify those that are not immediately used. In order to understand the
impact on improving future fracture stimulation designs, the authors present workflows such as pressure
history matching of fracture stimulation treatments and the calibration process of key rock mechanical
parameters such as Poissons ratio, Youngs modulus, and fracture toughness. The authors also present
examples to discuss synergies, discrepancies and gaps that currently exist between geologic geomechanical concepts (i.e. variations in the geometry and magnitude of stress tensors and their interaction with
pre-existing anisotropies) in contrast to the geomechanical descriptions and concepts that are used and
implemented in hydraulic fracturing stimulations.
In the absence of a unifying hydraulic fracture design that honors well established geologic complexity,
various scenarios that allow assessing the criticality, usefulness and weighting of geologic/mechanical
property input parameters that reflect critical reservoir complexity, whilst maintaining applicability to
hydraulic fracturing theory, are presented in the paper. Ultimately it remains paramount to constrain as
many critical variables as realistically and uniquely possible. Significant emphasis is placed on reservoirspecific pre-job data acquisition and post-job analysis. The approach presented in this paper can be used
to refine hydraulic fracture treatment designs in similar complex reservoirs worldwide.

Introduction
The design and success of hydraulic fracture stimulations depends critically on the correct implementation
of models that account for in-situ stresses and physical properties of rocks such as Youngs modulus and

SPE-173378-MS

Poissons ratio. In particular, basin wide variations


of stress geometry can be challenging for fracture
stimulations as is seen to be the case in eastern
Australia. In the basin, the horizontal stresses are
commonly the maximum principal stress resulting
in dominance of reverse and/or strike slip stress
regime (HmaxV hmin and HMaxhminV,
respectively), whereas a normal stress regime
(VHmaxhmin) is less common. Variations in
stress geometry occur both laterally throughout the
east Australia Surat basin as well as with depth
(Flottman et al. 2013). Fig. 1 shows a schematic of
various stress regimes typically observed in the Surat basin.
Hydraulic fracturing treatments in coal seams of
Surat basin are influenced by the high Poissons
ratio (greater than 0.3) and low Youngs Modulus
combination which is commonly observed through- Figure 1Stress regime prevalent in Surat Basin, Eastern Australia.
out the basin. It is of primary importance to account
for all critical variables during treatment modeling
and design phase in order to ensure fracture containment in the target coal zones to prevent excessive
growth of hydraulic fractures out of the coals into the non-productive and often fluid-sensitive interburden. Using examples from the Jurassic Surat Basin in eastern Australia, the authors present workflow
case studies showing the need for detailed calibration of reservoir rock parameters and stresses and their
influence on design improvement loops in successive fracture stimulation treatments. Previous work
(Flottman et al. 2013) has shown a significant influence of horizontal differential stresses and fracturing
fluid selection on the hydraulic fracture geometry, where the influence of stress regime on hydraulic
fracturing geometry in light of their vertical and horizontal orientation is demonstrated.
Strain-derived stress models, also termed as Wellbore Stress Models (WSM), were developed to
characterize the stress variability over the region (Brooke Barnett et al. 2012, Flottman et al. 2013). Stress
or rock strength calibrated Hmin from WSMs are routinely used as key stress input in fracture designs.
The case studies presented in this paper show that log-derived and strain calibrated stress models have to
be further refined or shifted using pressure matched data from offset wells in order to obtain best-fit
realistic stress models for predictive fracturing treatment designs. The majority of results, in particular
from fracture diagnostic tests show that coals have the lowest bulk stress. Our observations suggest that
log based stress implementations, especially Poissons ratio and to a certain extent Youngs modulus,
must be viewed with caution in areas where the greatest principal stress is horizontal.

Hydraulic Fracture Modeling


Modeling of hydraulic fractures is a critical step in the fracture design process. Fracture treatment designs
aim at achieving optimized target fracture lengths, vertical coverage or fracture height, and fracture
conductivity. These main characteristics allow a design engineer to plan for total treatment volume and
execution pattern. However, the actual design of hydraulic fractures is far from this simplistic overview.
Depending on the application, the fracture designs are modeled using one or several of the commercially
available fracture simulators in the industry today. Most of these models are grid-based and rely on
numerical simulation to generate fracture geometry for a given set of inputs. The inputs primarily include
geomechanical attributes but are not limited to, vertical stress distribution in target and surrounding layers,
mechanical properties of rocks, critical fluid and proppant characteristics, and lastly a treatment schedule

SPE-173378-MS

that provides the pumping rates, volumes and proppant concentration. The output from the models may
vary depending on the underlying assumptions, but typically these models are designed to honor the
fundamental laws of conservation of mass, momentum and energy. They also follow the constitutive laws
related to stress-strain relationships and flux laws. The mathematics and computer codes involved in
simulations are rigorous but the user is seldom burdened with intricacies of calculations and internal
solution schemes, and is only required to have a general understanding of critical inputs to the simulator
that govern the output.
In the last decade, the industry has seen a rapid expansion of hydraulic fracturing activities in
unconventional reservoirs, mostly driven by shale exploration and development activity in USA (King
2010). Fracture initiation and propagation in these unique geological settings and rock fabrics has
purportedly resulted in increased awareness of the complexity that entails some of the hydraulic fracturing
treatments. Qualitative evidence from alternate measurements like microseimic surveys and mine-backs
show that in unconventional reservoirs rocks, fracturing can result in a highly complex network of
pathways (Fisher and Warpinksi 2011) which are not easy to predict. Though attempts to model these
complex fracture networks are currently ongoing industry wide, it is understood that solutions may not
be anywhere near owing to the heterogeneity and uncertainty involved.
Typical Simulator Inputs
The modeling of hydraulic fractures in coal seams was understood to be a fairly challenging task even
before the advent of shale fracturing. Most of the coal deposits worldwide that are potential candidates for
hydraulic fracturing occur in shallow depths often ranging from 800 to 2,500 ft [243.8 to 762 m]. Stress
regimes in these reservoirs could range from lithostatic conditions with near similar magnitudes of
principal stresses to complex stress regimes resulting from the influence of regional tectonics. In a
strike-slip or reverse-stress scenario, additional complexity related to orientation and rotation of hydraulic
fractures during the treatment can be expected. Hydraulically induced fractures open in the direction
perpendicular to minimum principal stress, 3, which if in a horizontal plane is denoted as hmin and
propagate in the direction of maximum horizontal principal stresses, 2 or Hmax which is the intermediate
principal stress. The implementation of stress regimes, is therefore of primary importance in fracture
designs. In strike slip and normal regimes, the hydraulic fractures propagate vertically whereas a reverse
stress regime tends to promote horizontal fracture propagation. Most commercially available fracture
simulators however, do not take into account for this 3 dimensionality of stresses.

Critical parameters in Geomechanical models


Most fracturing simulators require formation in-situ stresses to be input along the depth axis and assume
a constant stress regime along a given horizontal plane. Some more recent versions of popular fracture
simulator available in the industry also allow the user to vary stresses, mechanical properties and other
petrophysical property inputs along the horizontal plane. This feature allows for better handling of
hydraulic fracturing modeling in horizontal wells where there may be significant variations of properties
along the lateral length of the wellbore, for e.g. depletion or residual stress from adjacent fractures.
In the workflow utilized to generate log based in-situ stresses, Poissons ratio v, is determined first
using incremental changes in shear and compressional wave travel times ts and tc as shown in Eq. 1
below:
(1)

Shear modulus G, which is the measure of a material response to deforming shear stresses, is calculated
next using both Poissons ratio and bulk density, b as inputs as shown in Eq. 2 below:

SPE-173378-MS

(2)
where, d is some known distance over which the measurements are made. Poissons ratio and Shear
modulus are then used to compute dynamic Youngs modulus, E using the following relation of Eq. 3:
(3)
Several correlations (Morales and Marcinew 2007) have been reported in various studies that allow
transforming dynamic E measurements to static values. The values obtained from static tests in the
laboratory experiments on representative rocks samples may also be used for calibration purposes and
correcting the log-derived values.
As a last step in log calculations, in-situ stresses are computed using the following poro-elastic stress
relation (Hubbert and Willis 1957) in Eq. 4 below that has been in traditionally use by the industry to
generate log based in-situ stress:
(4)
where, h is calculated minimum in-situ stress, v is Poissons ratio, v is vertical stress (overburden),
Pp is pore pressure, is Biots poro-elastic constant, E is static Youngs modulus, and h & H are strains
in minimum and maximum principal stress directions respectively. The first two terms on the right side
of the equation reflect stress conditions under a uniaxial environment where little stress anisotropy may
exist in horizontal plane which is generally the case for tectonically relaxed extensional environments. The
last two terms on the right side account for additional stresses due to regional tectonics. Initially they were
denoted by only tectonic stress t, which in the current form of equation is elaborated further. The
multicomponent equation provides the first approximation of formation stresses and thus a close estimate
of fracture closure pressures for given conditions.
Both, Poissons ratio and Youngs modulus are extensively used in fracture mechanics computations,
with the latter being more influential in fracture widths and net pressure calculations. The effects of
Poissons ratio are generally diluted in these mechanics calculations owing to the plane strain assumption
where the strain resulting from deforming forces existing in a 3D plane are simplified by neglecting strain
(assumed to be infinitesimally small) in length direction, thus assuming that the planes remain parallel
before and after deformation.
Fracture widths are calculated on the basis of a width equation (Sneddon and Elliot 1946) for mode I
fractures given by Eq. 5 in the text below.
(5)
where, pf is average fracture pressure, L is the length of the fracture, min is minimum principal stress
and is a dimensionless constant that depends on the shape. If net pressure given by the relation (pf
min) is substituted by relation for the critical stress intensity factor KIc, Eq. 5 can be modified to calculate
width at the wellbore given by Eq. 6 as:
(6)
KIc is also referred to as fracture toughness and is considered as a material property for a perfectly
elastic medium. When the width equation is combined with fluid equation, the following relation for net
pressure Pnet presented by Eq. 7 is arrived at (Nolte 1991):
(7)

SPE-173378-MS

where EE/(1-v2) is plane strain modulus and Ptip are fracture tip pressures that can be calculated from
the fracture toughness defined above. The above equation clearly shows the dependence of net pressure
on pumping rates and fluid viscosity. If it is solved for fracture width, one gets the relation ww
{qiL/E=}1/4.
For a case where the stresses in bounding layers (2), are higher than the ones in pay interval (1) from
where the fractures would initiate, and also higher (21), the following relation represents an
equilibrium condition as shown by Simonson et al. (1978) in Eq. 8 follows:
(8)
where, P is the net pressure in the fracture, h is the pay zone thickness and H is the total fracture height.
Eq. 8 takes a conservative approach in height prediction but with similar relations developed for
non-symmetric stress cases, the simulators account for flow resistances offered by in narrow width regions
in fractures and tend to limit fracture heights.

Fracture Complexity and Fracturing Simulators


The more simplified two dimensional (2D) analytical models assumed a constant fracture height and
calculated evolution of fracture widths with time. With the introduction of high-speed computers, these
models eventually made way for numerical models that heavily rely on the increased computational ability
to carry out complex numerical schemes, that additionally incorporate 2D fluid low effects in the fracture
and fluid leak-offs in the formation to account for material balance and net pressure calculations which
influences fracture geometry. Layered models were introduced and fracture heights were no longer
assumed but were calculated during the simulation. Four major parameters that have an effect on fracture
height are layered stresses, modulus, fracture toughness and slippage at weak interfaces. Amongst these,
the first 3 are generally included in calculations in most simulators but owing to the uncertainties
associated with the failure pattern and ensuing computational complexity, shear slippage is not currently
accounted for adequately in most simulators. This particular shortcoming prevents accurate fracture height
growth and/or containment prediction in heavily layered formations. Also, the individual influence of
these parameters on actual geometry will vary from region to region. For example in ultra-soft rock
formations of Youngs modulus of less than 5x105 psi [3,447.4 MPa] fracture toughness KIc plays a
critical role in net pressure computation and consequently fracture propagation; similarly in layered
formations with high contrast in modulus and weak interfaces, interfacial slippage may have more
influence on dictating the vertical propagation of fracture.
Complexity in fractures generally arises when conditions that favor fracture propagation in multiple
planes exist; this could range from presence of low stress anisotropy between principal stresses that affect
orientation of fractures, highly fractured reservoirs that could lead to ample interaction between hydraulic
and natural fractures, high in-situ stresses in comparison to overburden which is generally the case in
shallow formations, to regions that are heavily influenced by tectonic stresses or several other similar
factors. Microseismic mapping can aid in determining the complexity of network created by hydraulic
fracturing though this is often open to interpretations and many believe that most fractures are still planar
in nature (Fisher and Warpinski 2012); other source is fracture excavations via mine-backs. Fracture
diagnostic tests can also help in detecting non-ideal behavior prior to the main pumping often they are
the precursors to the pressure behaviors observed during the actual fracture treatment. However, interpretation and forecasting requires considerable expertise and experience for a given field. In an ideal
setup, to accurately predict fracture geometries from simulation, geomechancis should be intertwined with
fracture mechanics. In the sections that follow the authors address some of the limitations and gaps that
currently exist in fracture modeling and the potential areas that are relevant to hydraulic fracturing in coal
seams.

SPE-173378-MS

Figure 2Geology of Surat Basin, Australia.

Geological Background
The Jurassic age Surat Basin of south-east Queensland, Fig. 2, formed as part of an extensive intracontinental sag basin (Cook and Draper 2013). Surat Basin sediments are dominated by siltstones and
mudstones with minor sandstones all of which were deposited in a fluvio-lacustrine depositional
environment which was affected by intermittent volcanic activity. Inter-burden (non-coal) rocks have
characteristically low porosity and permeability which generally results in low fluid leak-off for typical
polymeric fracturing fluid systems. The Surat Basin contains multiple coal seams, on average 60, with an
average thickness of 0.4 m [1.3 ft]. Fig. 3 shows location map of Walloons fairway (Flottmann et al. their
Fig. 1).
Contrasting permeabilities in coals appear related to drape of the Walloons sequence over the
underlying Triassic structures. Dome-like drape over structural highs leads to regional curvature resulting
in multiple fracture orientations, typical for regions with high permeability. Poorer permeability regions
are characterized by a unidirectional single fracture orientation which appears to result from hinge-like
movement on a monoclinal flexure. These features are predominant where the Walloons fairway overlies
structural lows of the underlying Bowen Basin.

Stress Characterization
This paper focuses on low permeability coals that are generally less than 10 mD though higher
permeability coals seams may be seen in the various coal packages infrequently. Image log data from 30
wells (Flottman et al. 2013, see also Brooke-Barnett et al. 2012) were interpreted to identify stress
indicators. Three principal stress regimes were identified with depth based on one dimensional stress
profiles. Log-derived input data include bulk density, compressional sonic, shear sonic, and gamma ray.
Poissons ratio and Youngs modulus measurements were calculated using sonic and density wireline
data, and a dynamic to static conversion relationship was derived regionally from rock mechanics
laboratory measurements. Pore pressure was calculated based on a fresh water hydrostatic gradient of
0.433 psi/ft [9.795 kP/m], which is commonly observed in non-depleted reservoirs where coals have not
been produced at a large scale. Minimum and maximum horizontal stresses were calculated using

SPE-173378-MS

poro-elastic stress equations (Zoback 2007, BrookeBarnett et al. 2012), which incorporate static Poissons ratio, vertical stress, pore pressure, static
Youngs modulus, a Biots constant (equal to 1),
and tectonic strain in the minimum (min) and maximum (max) horizontal stress directions. More details on these are presented in the literature review
in the later sections of the paper.
Calibration of the stress profiles was undertaken
using either image logs (stress polygon method) or
fracture diagnostic injection test data from individual wells or rock strength calibrations. Whenever
available, fracture closure pressures were derived
from pressure data such as injection tests, Modular
Dynamic Testing (MDT) based small scale fractures
and pre-treatment injection tests in order to estimate
minimum principal stresses (Baree et al. 2007,
2009).
Stress
Figure 3Location map Walloons freeway. Color ramps shows depths
Overburden stress in the Surat Basin remains con- of top basement.
stant at close to 1.0 psi/ft [22.621 kPa/m]. However,
most regions display characteristic stress regime transitions with depth (Fig. 1). The stress regime
shallower than around 400 to 500 m [1,312.3 to 1,640.4 ft] is characterized by reverse stress field
described in the order, vhminHmax, where v vertical stress, hmin minimum horizontal stress;
Hmax maximum horizontal stress. At deeper depths, the stress field is typically of strike slip stress
regime described above (Johnson et al. 2002). In some areas a second transition from strike slip to a
normal stress regime (hminHmaxV) can occur around 650-800 m [1,968.5 to 2,624.7 ft] as shown
in Fig. 1. In particular, the transition between reverse and strike slip regimes is corroborated by tilt meter
data. In-situ stresses are the principal guiding parameter for hydraulic fracture geometry and propagation,
differential stress magnitude, as well as Hmax orientation.

Generating Log-Based Models


The stress curves calculated from logs are generally represented in a continuous format around the zones
of interest as seen in the Fig. 4 below. Data from Tracks 1 to 4 (L to R), including depth and the
perforation track on the extreme left, shows some of the measured properties, and tracks 5 onwards to the
right show calculated properties based on the equations mentioned in the earlier section. The data pertains
to Well A, drilled and completed across several of the coal seams in the Walloons coal measures in eastern
Australia. The lower most perforations are in a section termed as Tarooms that is generally associated with
low permeability coal seams with poorly developed cleat systems. Moving up in the well, higher
permeability coal seams with an enhanced fractured system are more commonly seen and are categorized
under zone named Juandah. These are further broken in to lower, middle and upper Juandah based on
depth.
It is a standard practice to calibrate log derived initial stresses and mechanical properties with measured
or analytically determined values in order to gain confidence in the geo-mechanical model. To accomplish
this important step, tests are conducted prior to the main fracturing treatment by first injecting a
pre-determined volume in the perforated coal interval at a planned rate and later, observing the pressure
response during the decline phase after the well is shut in. The primary objective of these injection tests

SPE-173378-MS

Figure 4 Log section of Well A showing measured and calculated log properties over gross interval from 400 to 1,050 m [1,312.34 to 3,444.88 ft].

Figure 5Closure pressure was determined to be 1,432psi based on GdP/dG analysis. Alternative possible late time closure pick is highlighted by
green line at 1,200psi.

is to estimate fracture closure pressures which provide an approximation of minimum horizontal principal
stress based on pressure fall-off diagnostic techniques (Nolte 1979, Castillo 1987).
Injection tests were pumped down a 7 inch [17.78 mm], 20 lbm/ft [29.76 kg/m] casing cemented at
3,638.45 ft [1,109 m] at 5 to 6 bbl/min [0.01325 to 0. 01589 m3/s] for an average of 14 bbl [2.241 m3]
per injection test. As is known, results from one injection test results in a single data point in comparison
to the continuous log curve; this approach is thus sometimes inadequate to characterize the stress and
mechanical property behavior over the entire zone of interest. Nevertheless, tests are conducted wherever
feasible so that the working model can be adjusted to actually observed results. Closure stresses derived
from diagnostic tests on 4 perforated test intervals are shown in the extreme right track in Fig. 4. Youngs
modulus calibrated on the basis of the 2D analytical model is also represented in the fifth track from left
as discrete data points.

SPE-173378-MS

Figure 6 Pressure History Match of Injection Test on Upper Juandah coal interval in Well A. Closure pressure, Pc 1,417 psi was used in the model
for simulation.
Table 1Summary of results from various diagnostic tests conducted on Well A.
Rate

Pc

Pr

F.G.

Modeled Pc

Modeled E

Perm.

Leak-off

psi

psi

psi/ft

psi

MMpsi

mD

ft/min0.5

2,485
2,340
1,432
1,782

2,368
1,441
972
1,476
848

0.746
0.719
0.636
0.844

2,487
2,085
1,417
1,780

1.39
1
0.65
2.1

Depth m

Rock Type

Data Source

bbl/min

1015.0
992.2
686.3
640.0
680.27
637.27

Silt
Coal
Coal
Sand
Coal
Coal

Inj. Test
Inj. Test
Inj. Test
Inj. Test
DST
DST

5.4
9.96
5.2
5

0.0044
0.004546
0.00692
0.0061
26.68
0.15

*Pc Closure pressure, Pr Reservoir pressure, Perm. Permeability, E Youngs modulus

Preliminary Data Calibration


In the initial phase of fracture modeling process in Walloons coal measures, during which the field-wide
stress characterization were still ongoing, some disagreement between initial estimates and observed
values was seen as would be normally expected. However, it was envisaged that over a period of time the
models would improve as additional data was being incorporated, ultimately leading to treatment designs
that could be placed precisely within the coal seams in order to maximize productivity.
The superposition derivative GdP/dG is often used in identifying the fracture closure pressure as seen
in Fig. 5; it is an accepted norm (Baree et.al. 2009) to consider the deviation from straight line behavior
as onset of the fracture closure event and the use of the derivative magnifies this effect. Correct closure
identification requires experience in the field. In complex rock structures such as coals, one may expect
a late closure of the main fracture shown by the late rise in derivative and an eventual recession, but it
is the higher stresses that tend to dominate the pressure behavior during the treatment. The plot shown in
Fig. 6 shows as pressure match using the values derived from closure analysis. Here the leak-off was
matched to 0.00692 ft/min0.5 [0.0002723 m/s0.5].
The early closure and associated hump seen in the early part of the GdP/dG curve also points to a
potential pressure dependent leak-off mechanism. The closure pressure thus identified also primarily
represents the closure of natural fracture system.

10

SPE-173378-MS

Figure 7Log section displaying Youngs modulus and in-situ stress curves after calibration process in Well A. Coal picks are based on Bulk Density
< 1.75 g/cm3.

Table 1 below summarizes the results from Drill stem Test (DST), Injection Test and main treatments
conducted at various depth intervals. Note that the tests were carried out in both coal seams and
inter-burden for stress estimation purposes.
During the calibration process of Well A, revised pore pressure estimates derived from injection tests
were used to apply base correction to log derived stresses. The revised curve is shown in Fig. 7 as Log
Stress and as shown the values are marginally different than the original estimates shown in Fig. 4 earlier.
Results from DST conducted during drilling phase were also used to validate the pore pressure data.
Further adjustments were then made to log stress curve by incorporating the effects of Youngs modulus
and strain. Static Youngs modulus curve was calibrated using the calibrated elastic moduli from injection
tests and shows a good agreement. Note that the regional stress curve generated to an offset field was
found to be inappropriate for use in this field as is shown under SHMIN_REG curve mentioned in Fig.
7.
Small fluid volumes pumped during injection tests allow for better initial estimates of in-situ properties
of the targeted intervals because by design, the fractures generated are confined to the zone of interest.
However, because of these smaller volumes, the near well region which is yet to fully clean up, can
potentially impose a choke effect and near-well tortuosity that can dominate the pressure response both
during the injection and also in the decline phase. Such a behavior often introduces uncertainty in the
analysis since it is difficult to quantify and account for accurately. As a practice thus, it was advised to
pump up to 20 bbl [3.18 m3] at at least 5 to 8 bbl/min [0.01325 to 0.0212 m3/s] in future injection tests
to ensure that most of near well entry restrictions are sufficiently minimized.
Additional Corrections to the Mechanical Property Model
Fracture treatments designed on the basis of initially developed stress and mechanical property models are
required to be corrected at least one more time after pumping of the main fracturing treatment. The extent
to which the correction is applied depends on how accurately the model has forecasted the pressure
behavior. The pressure behavior that is matched here is the calculated bottom hole pressure (BHP) using

SPE-173378-MS

11

Table 2Summary from Fracture Stimulation in Coal Seams.


Depth
Stage*
1
2
3
4
5

m
1,058.54
856.06
787.99
735.89
685.76

Rate

Pc

Pr

F.G.

E_Stat

Perm.

Leak-off

Formation Name

bbl/min

psi

psi

psi/ft

MMpsi

mD

ft/min0.5

Tarooms
Lower Juandah
Lower Juandah
Mid. Juandah
Upper Juandah

20.5
20.0
19.75
10 to 13
20.25

2102.9
1968.0
1795.9
1766.4
1415.6

1,410.0
1,067.3
982.4
917.4
855.0

0.606
0.701
0.695
0.732
0.793

0.494
0.43
0.383
0.444
0.377

0.10
1
5

0.001562
0.01177
0.002756
0.005376
0.01177

26

Figure 8 Plots shows Pressure History Match of fracture stimulation treatment in stage 5 in Upper Juandah.

surface treating pressures, hydrostatic and friction pressures as inputs. Pressure history matching from all
stages was carried out using commercially available third party fracture simulator. Original stress models
developed from preliminary injection data were used as a baseline and adjustments were made wherever
necessary. In most cases only nominal changes had to be made in the stress values and the fluid leak-offs
in order to obtain a good match. A summary of results from all the matches is shown in Table 2. Stages
1 and 3 screened out at the tail end of the job; in stage 4 the operations were curtailed due fluid gelling
issues.
In Table 2 above, mechanical properties and stress data are averaged across the perforations only. Fig.
8 shows the pressure history match plot for stage 5 of Well A, and the corresponding fracture geometries
from pressure matches on stage 3 and 5 are shown in Fig. 9 and 10 respectively. The fractures are shown
to be propagating in a vertical plane and planar in nature as would be expected as outputs from typical
simulators. The complexities associated with fracture stimulation in coal seams is not truly captured here
though the equivalent geometries resulting from pressure history matches are presented after processing
the material balance and proppant placement.
The case history in Well A was mostly used to present the work flow that is traditionally practiced in
fracture design process in coal seams stimulation designs. The fracture mechanics and rock geomechanics
interaction along with existing gaps in the simulators will be addressed in few other case histories where
additional data were collected to characterize and estimate fracture dimensions.
Stress and mechanical property data averaged for all layers that were constructed for the simulator was
collected from 5 simulation loadcases and merged in a single data file. This provided a continuous data

12

set representing the critical properties on which the


pressure history matches were carried out. When
included in log file display, they appeared to agree
well with the constructed model, indicating that
initial model provided good estimation of in-situ
properties. As seen in the log section of Fig. 11
above, the maximum departure of predicted versus
the actual properties was seen in the upper Juandah.
Youngs modulus was also revised, but there is little
influence of this parameter on simulation results,
especially on treating pressures when the rock
strength is less than 500,000 psi [3,447.38MPa].

SPE-173378-MS

Figure 9 Pressure History Match of fracture treatment in lower Juandah (stage 3) of Well A.

Case Histories
A number of wells in the Surat Basin were fracture
stimulated in 2011 to unlock gas reserves in low
permeability Walloons coals. During the initial
phase of the well stimulation campaign a number of
pilot wells were completed with various stimulation Figure 10 Pressure History Match of fracture stimulation treatment in
techniques to aid in understanding how fracture upper Juandah (stage 5) of Well A.
stimulation would influence well productivity. In
general, it was known that with higher Poissons
ratio reaching up to 0.4 in static measurements and Youngs modulus traditionally below 1.0E6 psi
[6.8948 MPa], the coals were softer and more ductile in comparison to their counterparts worldwide, thus
making them unique. The cleat structure in these Surat Basin low permeability coals has not developed
extensively though vitrine reflectance and gas content point to ample presence of exploitable gas.
Case History: Well B
Well B was the first well to be hydraulically fractured in the low permeability coal seam fracturing
campaign and was also a part of the pilot program where apart from gathering extensive log data, other
measurements such as microseismic survey and tiltmeter surveys were done. The well was drilled and
cemented with 5- inch [139.7 mm] N-80 casing to 758.0 m [2,486.9 ft]. The well lies in a region which
is characterized by a unidirectional fracture system ubiquitously seen in all of the wells drilled and
completed in this area. From the in-situ stress standpoint, there is a presence of high differential stress
which is evident in the image logs where bore hole breakouts in long vertical sections were seen to be
occurring commonly. The regional stresses also show a distinct presence of strike slip environment around
430 m [1,410.8 ft] where hmin is almost same as v. Calculations based on mechanical property logs
indicated a low Youngs modulus and higher Poissons ratio which was consistent with general observations in the Surat Basin coals.
The microseismic and tiltmeter data obtained at the end of the 7-stage treatment were very helpful in
independently verifying the ability of simulator to model fractures. Treatments were pumped to target the
Tarrooms, Lower Juandah, McAlister and Upper Juandah coal seams in the order mentioned. All 7 stage
of Well B were treated with a 20 lbm/Mgal [2.4 kg/m3] Borate cross-linked gel and 16/30 U.S.Mesh size
[1.194 0.584 mm] sand. Fine 100 U.S. Mesh size [0.043mm] was used occasionally in pre-pad and pad
stages for near well clean up and also to serve as a bridging agent. The observation well for microseismic
measurements was located around 100 m [328.1 ft] south and with the maximum principal stress direction
being N17W, some bias in fracture-wing dimensions may be expected. As a first step, based on the log
data, a preliminary model was constructed and treatments for various stages were designed on that basis.

SPE-173378-MS

13

Figure 11Log section displaying Youngs modulus and in-situ stress curves at the end of calibration process concluding after pumping of
treatments.
Table 3Well B Summary from Injection Tests and Treatment Analysis compared to results from MS Survey.
From Logs

Stg
1
2
3
4
5
6
7

Treatment Analysis

Net
m

Perm
mD

Perf Leak-off
m ft/min0.5

3.58
1.38
1.68
1.06
1.91
3.51
3.81

2.24
5.89
0.13

752.9
711.8
686.2
615.7
567
463.3
427.1

2.32
11.9
3.48

0.0032
0.00177
0.00175
0.00175
0.00255
0.0035
0.0041

Fracture Stimulation Details

Xf, m

hf, m

Hz

Stress
psi

DFIT
psi

F.G.
psi/ft

Rate
bpm

Fluid
bbls

Prop
klbm

Max
PPA

PM

MS

PM

MS

1,963
1,950
1,650
1,770
1,517
1,470
1,536

2,071
1,934
1,707

0.795
0.836
0.734
0.877
0.816
0.968
1.097

40
35
36
35
35
35
35

1,160
1,182
1,042
770
1,661
1,312
1,160

66
39
32
19
87
83
66

4
3
3.35
3.5
4.5
5
5

33.4
91.5
73.2
61
129
58.8
39.8

61
52.5
74
63.5
94
70
90

99.4
105.5
96.1
61.6
94.2
47.2
81

110
47
77
52
89
55
50

37
23
17
31
15
35
14

1,421
1,338
1,552

Proppant concentrations were designed up to a maximum of 5 lbm/gal [599 kg/m3] to provide optimal
fracture conductivity in low moduli coals.
Fracture diagnostic tests were conducted prior to some of the treatments for determining fracture
closure and fluid efficiency. Pressure history matching was carried out on all treatments and the results
of fracture geometry were compared with the ones obtained from microseismic interpreted geometry.
Table 3 below summarizes the pressure history match and microseismic survey derived fracture geometry.
In the Table 3, pressure-match and Microseismic are abbreviated as PM and MS respectively.
Proppant concentrations are shown in lbm/gal [kg/m3] or PPA and Hz denote the percent horizontal
component observed from tiltmeter survey data. Contrary to what maybe be expected, some of the
shallower stages had lower horizontal fracture component especially in last 3 stages located at the depths
from 427 to 567 m [1,401 to 1,860 ft]. Permeability and Net Coal data was obtained as an input from
subsurface team.

14

SPE-173378-MS

Figure 12Pressure history match of treatment data in Well B shows a good match but inaccurate representation of geometry. The fluid section in
the fracture is shown in blue shade and the proppant is depicted with shades of yellow and brown with increasing proppant concentrations inside
the fractures.

Pressure Match Analysis


Stage 1 of Well B was treated with 166,000 lbm [75 t] of 16/30 U.S. Mesh size proppant at 35.0 bb/min
[0.0927 m3/s] at a maximum wellhead pressure of 1,990 psi [13,720.6 kPa]. A limited entry approach was
used to treat 3 of the Tarooms Coal seams perforated at depths ranging from 749.5 to 757.6 m [2,459 to
2,485.6 ft] at 3.0 to 6.0 spf at 60 to 120o phasing. Fracture pressure history matching of surface data was
obtained using conventional pseudo-3D models that simulate 2D fluid flow inside the fracture. The
pressure plot on the left in Fig. 12 shows the pressure match of surface data and an overburden pressure
line across the plot indicating that a large portion of the treatment the effective BHP was above the
overburden pressure; this may not be an accurate representation of pressure state inside the fracture
however, because the near well bore related friction pressure of nearly 380 psi [2,620 kPa] is included in
the simulated bottomhole (BHP) pressure curve. If this excess pressure is removed, only a small portion
of treatment prior to onset of proppant appears to treat above overburden pressure. Resultant fracture
geometry is shown on the right, where the brown dots represent microseismic events recorded during the
treatment. The simulated fracture lengths are only 55% of observed activity whereas the height is within
10% of the observed values.
From a material balance standpoint if the fractures were contained vertically, the half-lengths would
have been longer and possibly result in geometries that could match fracture lengths inferred from
microseismic data. In all simulation runs, a symmetrical bi-wing fracture was assumed for simplicity.
Hydraulic fracturing in coal is difficult to model, but in this case it is apparent that the thin coal seams
do not dominate the fracture growth mechanism in either direction. The fracture complexity is introduced
only by the presence of a horizontal fracture component up to 37% measured by tiltmeters, which is not
traditionally accounted for in most fracture simulators. Simulated BHP clearly shows that there was ample
energy in the system to raise the overburden and re-orient the fractures in the horizontal plane, especially
during early stages of injection.
The pressure history match of Stage 5 on the other hand resulted in fracture geometries that were
extremely close to the observed values. One plausible explanation of this is the limited presence of small
horizontal fractures (15%) and lower in-situ stresses. A diagnostic test was pumped in the zone prior to
the main treatment where a fracture closure pressure of 1,421 psi [9,797.5 kPa] was identified. This was
close to the log derived value of 1,410 psi [9,721.6 kPa] but lower than averaged matched value of 1,517
psi [10,459.2 kPa]. Fracture pressure history match and associated geometry are shown in Fig. 13.
The fact that most fracture simulators are designed to handle relatively simple cases of homogenous
rock fabric is evident here. The reasonably good pressure history match of this treatment is not a
coincidence but a proof that most simulators can provide a reasonably good and agreeable results if they
are used in applications they are designed for. The pay interval that is fracture stimulated in this stage lies
in a depth range that corresponds to normal stress regime where in-situ stresses can be easily defined by
the uniaxial stress equation presented in the section above. There is only a 6% difference between

SPE-173378-MS

15

Figure 13Pressure history match of stage 5 in Well B is shown above. Both pressure match and corresponding fracture geometry show a favorable
simulator performance.

Figure 14 Pressure history match and corresponding geometry of stage 6 in Well B is shown above. Note that the simulated BHP shown in the
pressure plot in the left even after removing near well choke effects are considerably higher than the overburden and falls marginally below that only
after the treatments ends, as seen in ISIP.

calculated and measured stresses and the overburden line drawn on pressure plot above clearly shows that
after initial near wellbore clean up, the remaining BHP was considerably lower than overburden
corroborating the findings from tiltmeter survey that showed a mere 15% horizontal fracture component.
Diagnostic injection tests pumped prior to the treatment in stage 6 showed a usually high fracture
gradient of 0.961 psi/ft [21.738kP/m]. To accommodate this, the stress profile in the geomechanical model
had to be changed considerably in order to obtain a match as shown in Fig. 14 below. The stage for
creating and propagating horizontal fractures was thus already set and much of it was observed in the
mircoseismic survey during the treatment.
As seen in the pressure plot on the left of Fig. 14, the start of the treatment is marked by high surface
pressures mainly due to near well restrictions which are eventually removed with 100 mesh sand sweeps.
In the later part of the treatment as the higher proppant concentrations are admitted in the perforations,
the pressure resurges, indicating restrictions in the entryway to the hydraulic fractures created during
pumping. The restricted width is also a characteristic of horizontal fracture component that readily accepts
fluid but due to low width may not accept proppant easily (Daneshy 2003). The horizontal component
determined from tiltmeter survey is around 35% which puts most of the microseismic events close in a
horizontal plane around the perforated depth interval. Fracture height from microseismic survey is around
55 m [180.45 ft] which is close to the matched value of 47.2 m [154.85ft]. However, the simulated fracture
lengths are shorter by 16%. Under these conditions where the minimum horizontal stresses are close to
vertical stresses, a small gain in net pressure has a potential to re-orient the fractures in the horizontal
plane.
The last stage of Well B was treated with similar design as first stage but with higher maximum sand
concentration of 5.0 PPA [599.13 kg/m3] and nearly 40.0 bbl/min [0.10599 m3/s] with cross-linked gel.

16

SPE-173378-MS

Figure 15Pressure history match and corresponding geometry of stage 7 in Well B is shown above. The microseismic events show a vertically
contained fracture.

Figure 16 Log section shows the summary rock mechanical properties and stresses used in modeling after calibrating them with measure data. The
departure from predicted stresses is evident in shallower depths.

The treatment was placed successfully in the shallowest upper Juandah coal seam. Prior to the main
injection, a fluid efficiency test was performed with 2% w/w Potassium Chloride (KCl) as is shown in Fig.
15 below. Fracture closure pressure was identified at 1,552 psi [10,700.7 kPa] resulting in a fracture
gradient of 1.08 psi/ft [25.064 kPa/m] which is higher than the known 1.0 psi/ft overburden pressure
gradient. For majority of the treatment, the effective BHP exceeded overburden pressure which is also
evident from end of treatment ISIP. This remains the case even if friction pressures of 560 psi [3,861 kPa]
are removed from simulated BHP. The fracture geometry indicates a well contained vertical fracture
within the bounds defined by microseismic events and in sharp contrast to observed fracture gradients and
BH injection pressures, the tiltmeter survey indicates only 14% horizontal component. The fracture
containment may also be because of the limited pad volume used despite at higher leak-off observed in
the diagnostic tests.
The log section shown in Fig. 16 compares the measured and calibrated properties used in fracture
models for Well B. As was noted in the discussion above, the log-based stress model predicts the stresses
closely and depths greater than 550 m [1804.46 ft] but in shallower regions close to 400 m [1,312.34 ft]
higher fracture gradients are observed. This is observation is consistent throughout the field as shown by
schematic in Fig. 17 (courtesy Paul, P. 2011).

SPE-173378-MS

17

Case History: Well C


Well C was drilled to 1,032 m [3,385.8 ft] and a 7
inch [177. 8 mm], 20 lbm/ft [29.76 kg/m] casing
was set and cemented. The region is characterized
by intermediate differential stresses that fall in the
range of 600 to 750 psi [4,136.85 to 5,171.1 kPa].
The maximum horizontal principal stresses and coal
fracture orientation are mostly aligned and trending
in WNW-ESE, N60W direction. However coal fracture analysis from image logs indicates the presence
of a pronounced fracture sub-set which trends WSW
ENE, at N80E. The 1D Stress profile shows a
possible transition from reverse to strike slip stress
regime below 400 m. All fracture treatments were at
depth with a dominant strike-slip stress regime.
All the 3 stages in the hydraulic fracturing treatment were pumped with only water treated with 2%
17Field calibrated stress map for Well B shows general charw/w KCl solution for clay control. The objective Figure
acteristics and stress regimes. Plot was constructed on basis of various
was to study the effect of treatment type on well data sources mentioned in the legend.
performance. 16/30 U.S. Mesh size proppant was
used in concentrations to provide conductivity. As
in previous case, 100 U.S.Mesh size sand was pumped in initial stages for removing any near well
restrictions. The treatments were pumped at various rates ranging from 40 to 50 bbl/min [0.10599 to
0.1325 m3/s]. Table 4 below summarizes the critical data from Well C.

Pressure Match Analysis


Fracture diagnostic tests were conducted prior to the main treatment on the first stage and the results from
analysis showed the fracture gradients to be nearly 0.864 psi/ft [19.554 kPa/m]. During the treatment, the
pumping rates were increased up to 45 bbl/min [0.11924 m3/s] in an attempt to pump the proppant away
but ultimately the treatment was prematurely flushed because of continuously increasing surface pressures. A total of 50,000 lbm [22.7 t] sand was pumped using nearly 6,021 bbls [957.3 m3] treated fluid.
Microseimic survey in the stage shows dense cluster of events and despite a known presence of strong
stress anisotropy and a well-defined maximum principal stress and coal fracture orientation, the induced
fractures do not show a particular fracture azimuth. A cloud of microseismic events roughly shows the
fractures trending in E-W direction. As in traditional slickwater treatments, pressure matching is not easy
owing to several undulations in pressures which are responses from fluid/rock interaction and also to
proppant admittance in the formation via semi-open perforations. Fracture widths are a strong function of
fluid viscosity as shown earlier, the lack of which results in smaller fracture widths that hinders the entry
of proppant in the fractures. This results in abrupt increases in treating pressures which was also seen in
most of the treatment stages in Well C.
Fracture pressure history match and associated geometry for stage 1 is shown in (Fig. 18). Simulated
fracture half lengths are close to the geometry inferred through microseismic events. Despite several
attempts a good surface pressure history match could not be obtained owing to the inability of fracture
simulator to accurately predict the pressure responses, especially for slickwater treatments. The BHP data
presented in pressure plot of Fig. 18 suggests that for majority of treatment the pressures were above the
overburden pressure though it must be noted that the data is calculated and not from downhole pressure
gage. Despite complexity in fractures as is evident with layout of microseismic events, vertical component

18

SPE-173378-MS

Table 4 Well C Summary of Results and comparison with microseismic survey.


From Logs

Stg
1
2
3

Treatment Analysis

Net
m

Perm
mD

Perf Leakoff
m ft/min0.5

7.74
3.67
1.41

0.45
0.028
0.3

982.4
806.6
672.7

Fracture Stimulation Details

Xf, m

hf, m Stg

Stress
psi

DFIT
psi

F.G.
psi/ft

Rate
bpm

Fluid
bbls

Prop
1bm

Max
PPA

PM

MS

PM

MS

0.0053 2,784
0.0105 2,975
0.02626 3,074

2,778
3,163
3,074

0.864
1.125
1.393

40
51.5
50

6,021
2,820
4,323

50,000
31,500
34,400

0.75
1
1

97.5
171.5
131

110
300
185

106.7
85.7
69.7

45
40
48

Figure 18 Pressure history match and corresponding geometry of stage 1 in Well C is shown above. Simulated fracture half lengths agree with
observed values, however simulated fracture height is nearly twice of observed values.

Figure 19 Well C Fracture geometry details from pressure match on stage 1. Stress profile shown in the left picture was altered to obtain the match
on geometry. Most modern-day simulators offer limited options for height-growth control.

of fracture would tend to dominate the fracture composition due to fracture gradient that remains below
the overburden pressure if near well related pressures are removed from simulated BHP.
Pressure history matches based on conventional models build for stages 2 and 3 were far from
representative of the geometries obtained from microseismic survey. This was mostly expected, based on
the knowledge of stress regime in the region which puts the upper 2 stages in the strike-slip region where
complex fracturing with a large horizontal component would occur readily.
The stress profile picture on the right in Fig. 19 below shows the increase in the upper barrier stress
carried out on the model to achieve the fracture geometry that resembles closely with the geometry from
microseismic survey. It is unrealistic to have stress differences of this magnitude in bounding layers as is
shown in the schematic on right; however most simulators are sensitive to these and offer stress contrast
as the only dominant mechanism to contain vertical fracture growth. Most modern day fracture simulators
offer limited options for height growth control other than stress manipulation. Those that do offer such a
feature, handle relatively simple layering formats that were found to be inapplicable here.
In addition to the data presented in the paper, analysis of several other wells that were fracture
stimulated in the campaign was carried out to improve the understanding of stress environment in the
Surat Basin. This analysis also led to changes in legacy pump schedules that were mostly adapted from
coal bed methane fracturing treatments worldwide but were eventually altered after realizing the unique-

SPE-173378-MS

19

Table 5Comparison of Fracture Geometries utilizing various fracture simulators.


Fracture Half Lengths, Xf (m)
Stg
1
2
3
4
5
6
7
1
2
3

Well
B

Fracture Height, hf (m)

Perf m

MS

Sim A

Diff. %

Sim B

Diff. %

MS

Sim A

Diff. %

Sim B

Diff. %

752.9
711.8
686.2
615.7
567
463.3
427.1
982.4
806.6
672.7

61
52.5
74
63.5
94
70
90
110
300
185

33.4
91.5
73.2
61
129
58.8
39.8
97.5
171.5
131

-45%
74%
-1%
-4%
37%
-16%
-56%
-11%
-43%
-29%

67.1
91.5
73.2
61
122
54.9
61
140.9
47.3
61

10%
74%
-1%
-4%
30%
-22%
-32%
28%
-84%
-67%

110
47
77
52
89
55
50
45
40
48

99.4
105.5
96.1
61.6
94.2
47.2
81
106.7
85.7
69.7

-10%
124%
25%
18%
6%
-14%
62%
137%
114%
45%

114.3
105.5
96.1
61.6
119.8
115.9
122
120.8
100.6
60.4

3.9%
124.5%
24.8%
18.5%
34.6%
110.7%
144.0%
168%
152%
26%

ness of Surat basin coals. In order to create a baseline for fracture geometries and for purpose of comparison, the above exercise was repeated using a
different simulator commercially available in the
industry.
The results from the findings are presented in
Table 5 above and show that in most cases the
differences in the geometry persisted. In the table,
column headers, Sim A and Sim B represent
fracture simulators A and B.

Fracture Vertical Growth


In complex stress regimes such as seen in the examples presented in the paper, the role of geomechanics should be integrated in fracture simulators.
Due to shallow depths and presence of strike-slip
and reverse-slip stress regimes, the induced fractures are multi-component and are not wholly represented by planar fracture models both in geometry and pressure responses. However, as noted by
previous researchers (Palmer and Sparks 1991) the Figure 20 Microseismic events on all treatments on Well B are proalongside the Youngs Modulus and Poissons ratio curve to see
fracture behavior in thin coals measures will be jected
possible effect of property changes on vertical growth containment.
dominated by the properties of interburden and
hence the fracture growth in vertical direction will
be influenced by the mechanical properties of the surrounding layers as well. This however does not
prevent the fractures from bifurcating into horizontal and vertical components, as was seen amply in
various tiltmeter surveys.
Observations from Microseismic Survey
Microseismic survey was carried out in some of the pilot wells and the results offered a great insight into
fracture behavior under various treatment types. When microseismic events from survey data were
arranged next to formation mechanical property data obtained from logs, as shown in Fig. 20 and 21 for
Well B (cross-linked gel) and C (slickwater) respectively, it was seen that the events tended to stay within
the bounds created by modulus layer contrast. This is highlighted in both plots by blue circles. The effects

20

SPE-173378-MS

Figure 21Microseismic events on all treatments on Well C are projected alongside the Youngs Modulus and Poissons ratio curve from logs. These
treatments were pumped with only treated water in a region that is characterized by strike-slip and reverse stress regimes.

of modulus layer in vertical growth containment have been studied extensively in past (Cleary 1980, Gu
and Siebrits 2008).
The effect is however not uniform throughout because of the variation in response. This only indicates
that the fracture vertical growth is controlled by a multitude of factors and the data can be interpreted in
many ways. Note that the fracturing treatments pumped with cross-linked fluid as in Well B exhibited
higher vertical growth in comparison to the jobs pumped with only 2% KCl water in Well C.
Rock Mechanical Properties
Higher Poissons ratio in coals (see PR track in Fig. 20 and Fig. 21 above) usually results in normally
pressured formations to have higher stresses than the bounding inter-burden sandstones as is also reflected
in Eq. (4). Thus a fracture initiating from higher stressed coals would be attracted to relatively lower
stressed surrounding layers of inter-burden because it requires relatively less energy to propagate in
low-stress environment. This can potentially result in an uncontrolled height growth in lower stressed
barriers. It must be noted here that in majority of diagnostics test, the coals seams exhibited a lower bulk
stress which is somewhat anomalous when strictly following Hubbert and Willis relation for stress
calculations shown in Eq. 4 above.
Fracture propagation across interfaces of varying mechanical properties also affects the fracture growth
and may abruptly blunt the fracture (Simonson et al. 1978). Gu and Siebrits (2008) also note that
depending on whether the fractures are crossing from low to high Youngs Modulus or vice-versa, some
hindrance to growth may be expected although it is governed by different mechanisms. In the former case,
both, the resulting fluid pressure inside smaller fracture widths in higher Youngs Modulus rock, and
higher stress intensity factor associated, as shown in Eq. (6), dominates fracture propagation. In the latter
however, the growth is mostly influenced by the smaller stress intensity factors associated with low
Youngs Modulus rock that prevents it from matching the formation fracture toughness, thus limiting
fracture propagation. In their simulations during parametric study, Gu and Siebrits showed higher growth
when the fractures traversed form low modulus to higher modulus rocks, which is relevant to Surat coal
seams where interburden has higher Youngs Modulus than coals by a factor of 6 or more in some cases.
In another mechanism influencing height growth, when fracturing across layered formations, any
interface with low cohesive strength and friction coefficient, can act as a plane of weakness and result in

SPE-173378-MS

21

Figure 22Injection in Well B prior to main treatment on lower Juandah shows increasing pressure response with subsequent injections. Red arrow
shows the increase in ISIP without using proppants.

slippage which can cause a fracture to thus terminate at the plane of weakness (Warpinski 1981). A mix
of horizontal and vertical fractures has been routinely seen in mine back operations in coals (Jeffery et al.
1992); however, this phenomenon cannot be modeled very effectively though some researchers (Gu et al.
2008) have recently made an attempt by creating an arbitrary variable of shear stiffness that ultimately
calculates the amount of deformation between formation layers and thereby determines the fracture width
profile where a slip could potentially occur. A high shear stiffness value implies perfect bonding and
no-slip conditions whereas low values can result in slippage.
Fracture Plane Orientation and Back Stress
As noted earlier, for most of the completions in Surat Basin where the in-situ stresses start transitioning
to strike-slip regime around 500 to 600 m depths, the fractures no longer remain in vertical plane but are
generally a mix of vertical and horizontal component to almost 100% horizontal in shallower depths
starting from 400 m and less, where the reverse stress regime sets in. Even in the deeper depths seen so
far, there was always a presence of horizontal component in the fracture system mapped using tiltmeters.
This can be attributed to the structural make up of coal that is generally complex, layered and with a cleat
system that provides ample sites for shear failures related to slippage. Such effects are not adequately
addressed in most of the commercially available fracture simulators. Furthermore, at depths where the
minimum principal horizontal stresses are only marginally lower than vertical stresses, any excess net
pressure that is generated during fracturing process, can cause the fractures to re-orient in horizontal plane
by lifting the overburden.
In a heavily fractured or cleated rock system such as coals, injection of low viscosity fluids in
combination with other factors can sometimes result in a pressure signature that mostly points to an
apparent net pressure gain or what is generally termed as Back Stress. Also, higher effective pressures
and vertical fractures in coal seams can result in cleat-compression and subsequent loss of permeability
(Palmer 1993) which can further exacerbate the situation if the pumping continues. This apparent stress
generated under these conditions has little influence on fracture geometry as it is not directly applied to
the fracture. In these cases, the BHP generated in excess of in-situ stress is not pure net pressure but also
reflects fractured systems inability to effectively transmit the pressures to surroundings thus creating a
pressure trap that enhances the local stresses in wellbore area and gives a false sense of net pressure gain.

22

SPE-173378-MS

Figure 23Simulated fracture geometry when using existing pump schedules on calibrated geomechanical model. Note the excessive height growth
and proppant placement in non-pay.

If these excessive pressures are not dissipated efficiently in the formation the resulting surface pressures would continue to rise as more volume is
pumped, which would ultimately lead to a premature screen out. The effect is more pronounced in
high permeability softer rocks such as the coal
seams in Surat Basin and also where the formation
fluid has low compressibility which again is true for
Figure 24 Fracture geometry simulated after reduced the rate and
most coal seams where free gas is not present.
job volume. The fracture vertical growths are contained and
This behavior was seen during diagnostic tests overall
proppant distribution inside the fracture shows improvement.
carried out just prior to the treatment in stage 1 of
Well B, as shown in Fig. 22, where two subsequent
injection tests that were carried out after a brief shut-down, resulted in higher fracture closures. The initial
injection was carried out with 100 bbls [15.89 m3] of 2% KCl water pumped at 20 bbl/min [3.18 m3/min]
which was later on followed by a mini-frac of 440 bbls [69.95m3] of 20 lbm/Mgal [2.38 kg/m3] Borate
cross-linked fluid. While some gain in net pressure is expected when higher viscosity fluid enters the
fracture, it is generally not expected that these injections would re-stress the coal in excess of 650 psi [4.48
MPa] as was observed here. It is believed that the initial injection of low viscosity fluid aided in enhancing
the complexity in the near well region which was also corroborated by pressure responses seen in early
portion of pressure decline after shut-in. The fully-coupled mode in some of the fracture simulators allows
for poro-elastic affects to be included in the simulation. These generally result in higher net pressure gains
when back stresses induced by rock poro-elasticity and thermal conductivity are taken into account.
Attempts to simulate the behavior using these features were moderately successful as the quality of
pressure-match is average. In all the simulations shown in Fig. 22, the observed net pressure increase of
650 psi could not be obtained the simulated net pressures as the end of second injection are nearly 350
psi less. However, the late period pressure decline simulated by the simulator under pore-elastic or
coupled model, matches well with the actual decline observed.

Fracture Design Optimization


Despite their limitations seen during the fracture modeling and analysis process in coals, it was generally
observed that in the depth ranges where the vertical fractures dominated, the planar mode fracture
simulators provided a good estimate of fracture parameters. A typical simulation run in the simulator often
suggested a vertical growth for the given conditions which was often viewed as over prediction of fracture
height; however after modeling efforts and comparison with results from microseismic survey, it was
apparent that vertical growth does occur and that apart from mechanical properties of relatively thin coal
seams, even the inter-burden properties played a significant role in fracture propagation.

SPE-173378-MS

23

Figure 25Plot indicates improvement in well performance with lowering of pumping rates. Gel loadings were also lowered in subsequent treatments
(not shown here).

Improving Existing Models


The ultimate aim of the modeling exercise is to develop representative earth models that are able to predict
plausible fracture geometries when applied in fracture simulators even outside the ranges that were used
to calibrate them. These models are also crucial in generating optimized treatment designs. In case of
Walloons coals, as noted earlier, it was most beneficial if the fractures remained contained and, that
majority of the treatment was placed in the targeted coal seams.
Using the techniques described above, the continuous log based model was first calibrated using field
and injection diagnostics data to generate a calibrated geomechanical model. The model was then used in
pressure history matching exercise and the resultant geometries were taken as baseline for the given
conditions. In the next step aimed at optimizing fracture designs, with the an added objective of limiting
fracture vertical growth and maximizing conductivity, basic net pressure principals described in Eq. 7
above were utilized. Pumping rates and fluid viscosity were lowered and an improved pump schedule to
maximize fracture conductivity was employed similar to the approach (Pandey and Agreda 2010) adopted
during fracturing treatment campaign carried on low permeability shallow sandstone reservoirs in Chittim
Field, south Texas, USA.
The output from simulator runs showed that for the same base fracture model, the new designs
provided a better placement with improved proppant distribution up to 3.0 lbm/ft2 [14.65 kg/m2] instead
of 2.0 lbm/ft2 [9.76 kg/m2]earlier and better containment of fracture height. The designs not only used
lower fluid and proppant volumes but also reduced equipment and horsepower requirements. Fig. 23
shows fracture geometry simulated using current designs and Fig. 24 shows fracture geometry resulting
from an altered pump schedule and fluid selection when applied on identical model.
Well Performance
The above exercise was a continuous process and helped in understanding the limitations of fracture
stimulators used for modeling, but nevertheless the improved understanding also allowed the engineers to
modify some of the existing legacy pumping schedules to the ones that were perceived as more
engineering solutions. As the campaign progressed, it was increasingly clear that substantial portion of

24

SPE-173378-MS

fracture treatment was ending up in non-productive interburden which resulted in concerted efforts to
avoid that. It was felt that by reducing treatment rates and fluid viscosity, and by modifying the pumping
schedules to allow large volumes of higher concentration slurry, the treatment would be more successful.
This resulted in modification of fracture treatment designs with the major limitation only being the extent
to which the rates could be reduced given the larger diameter (7.0 inch) casing.
The changes made in the schedule ultimately resulted in a favorable well performance as can be seen
in the gas rate vs. fracture conductivity plot shown in Fig. 25. In some cases however, premature screen
outs were also witnessed. In-depth analysis showed that most of these were related to fracturing fluid
instability arising out of source water inconsistency. The issues were addressed and majority of treatments
were placed as designed.

Conclusions
During the current study geomechanical models were generated for several wells under the pilot well
program. Pertinent data obtained from multiple sources was used to carry out the analysis mainly to
evaluate the ability of various fracture simulators to accurately model the observed fracture geometry
trends based on external measurement. At the end of the study, the authors made the following
conclusions:
1. Use of geomechanical models is necessary to accurately model hydraulic fractures in coal seams.
2. At the end of each step in geomechanical model generation, necessary shifts in stress data had to be
applied to match observed values. Adjustments were made locally as opposed to shifting the stresses
across entire section.
3. During the calibration and model building phase, reservoir pressures obtained from extended
post-injection declines and from DST with downhole pressure gages were found to be more useful.
Accurate measurement of pore pressures is strongly recommended in model building instead of
assuming a pore pressure gradient.
4. It was observed that depth greater than 650 m [1,968.5 ft] where low fracture gradients existed and
majority of the hydraulically created fractures were vertical, most simulators could match the pressures
and geometry with fairly good accuracy, provided the treatments were pumped with viscous fluids. In
a nutshell, 2 of the fracture simulators used in the study could predict fracture half lengths within /20% of observed values provided the depths exceed 615 m [2,017.7 ft].
5. In shallower depths, with increasing fracture gradients and higher reported horizontal components,
the variation in the geometry increased considerably.
6. The treatments pumped with treated water were difficult to pressure match and there was a large
discrepancy between the predicted geometry and the observed values mostly due to the complexities
in the fractured systems that the fracture simulators do not address efficiently.
7. The presence of non-traditional stress regimes contribute to fracture complexity which can range
from re-orientation of fractures in horizontal planes due to lifting of overburden to abrupt truncation
of fracture growth due to rapidly changing geomechanical properties. If these effects are effectively
accounted in the fracture simulators, a more representative fracture geometry and azimuth prediction
can be made.
8. For depth ranges where predictions were reasonably good, the fracture simulators can be constructively used in improving placement designs.
The learnings from this study can be applied to similar shallow coal bed methane or other unconventional reservoirs that face a similar challenge from rock mechanics and stress regime perspective.

SPE-173378-MS

25

Acknowledgements
The authors wish to thank the management of ConocoPhillips and Origin Energy for their permission to
publish this work.
Nomenclature
d
known distance, L, ft [m]
E
Youngs Modulus, M/Lt2, MMpsi [MPa]
E
Plane strain Modulus, M/Lt2, MMpsi [MPa]
G
Shear Modulus, M/Lt2, MMpsi [GPa]
Total Fracture Height, L, ft [m]
H,hf
h
Height of Pay, L, ft [m]
Critical Stress Intensity Factor, M/L0.5t2
KIc
L
Fracture Length, ft [m]
P
Net pressure, M/Lt2, psi [kPa]
Fracture Closure Pressure, M/Lt2, psi [kPa]
Pc
Fracture Pressure, M/Lt2, psi [kPa]
Pf
Pore Pressure, M/Lt2, psi [kPa]
Pp
Reservoir Pressure, M/Lt2, psi [kPa]
Pr
Pressure at Fracture Tip, M/Lt2, psi [kPa]
Ptip
Injection Rate, L3/t, bbl/min [m3/s]
qi
w
Fracture Width, L, inch [mm]
Fracture width at wellbore, L, inch [mm]
ww
Incremental travel time compressional wave, t/L, s/ft [s/m]
tc
Incremental travel time shear wave, t/L, s/ft [s/m]
ts

Biots Constant, dimensionless


Strain in minimum principal horizontal stress direction, dimensionless
h
Strain in maximum principal horizontal stress direction, dimensionless
H
v
Poissons ratio, dimensionless

Equation Constant

Fluid Viscosity, m/Lt, cP [Pa.s]


b
Bulk Density, M/L3, lbm/gal [kg/m3]
1
Maximum Principal Stress & Payzone Stress in Eq. (8), M/Lt2, psi [kPa]
2
Intermediate Principal Stress & Bounding Layer Stress in Eq. (8), M/Lt2, psi [kPa]
3,min Minimum Principal Stress, M/Lt2, psi [kPa]
Hmax Maximum Principal Horizontal Stress, M/Lt2, psi [kPa]
hmin
Minimum Principal Horizontal Stress, M/Lt2, psi [kPa]
t
Tectonic Stress, M/Lt2, psi [kPa]
v
Vertical Principal Stress, M/Lt2, psi [kPa]

References
Barree, R. D., Barree, V. L., & Craig, D. 2009. Holistic Fracture Diagnostics: Consistent Interpretation of Prefrac Injection Tests Using Multiple Analysis Methods. SPE. doi: 10.2118/107877-PA.
Barree, R. D., Gilbert, J. V., & Conway, M. 2009. Stress and Rock Property Profiling for Unconventional Reservoir Stimulation. SPE. doi: 10.2118/118703-MS.
Castillo, J. L. 1987. Modified Fracture Pressure Decline Analysis Including Pressure-Dependent
Leakoff. SPE. doi: 10.2118/16417-MS.

26

SPE-173378-MS

Cleary, M.P. 1980. Analysis of Mechanisms and Procedures for Producing Favourable Shapes of
Hydraulic Fractures. Paper SPE 9260 presented at the 55th Annual Fall Technical Conference and
Exhibition, Dallas, Texas, USA, 2124 September. doi: 10.2118/9260-MS.
Cook, A.G. and Draper, J.J. 2013, in: Jell, P.A Geology of Queensland, Geological Survey of
Queensland, Australia.
Daneshy, A. 2003: Off-Balance Growth: A New Concept in Hydraulic Fracturing. J. Pet Tech (Apr.,
2003) 78 85. http://dx.doi.org/10.2118/80992-JPT.
Draper, J.J. 2013, in: Jell, P.A. (ed.) Geology of Queensland, Geological Survey of Queensland,
Australia.
Brooke-Barnett et al 2012, PESA East Australian Basins Symposium, Brisbane September 11-14
extended abstracts.
Fisher, M. K., & Warpinski, N. R. 2012. Hydraulic-Fracture-Height Growth: Real Data. SPE. doi:
10.2118/145949-PA.
Brooke-Barnett, S., Naidu, S. K., Paul, P. K., Flottman, T., Kirk-Burnnand, E., Busetti, S., Hennings,
P., Trubshaw, R. L. 2013. Influence of in Situ Stresses on Fracture Stimulation in the Surat Basin,
Southeast Queensland. SPE. doi: 10.2118/167064-MS.
Gu, H. and Siebrits, E. 2008. Effect of Formation Modulus Contrast on Hydraulic Fracture Height
Containment. SPE Production and Operations 23 (2): 170 176. SPE-103822-PA. doi: 10.2118/
103822-PA.
Gu, H., Siebrits, E. and Sabourov, A. 2008. Hydraulic-Fracture Modeling with Bedding Plane
Interfacial Slip. Paper SPE 117445 presented at the Eastern Regional/AAPG Eastern Section Joint
Meeting, Pittsburgh, Pennsylvania, USA, 1115 October. doi: 10.2118/117445-MS.
Hubbert, M.K. and Willis, D.G. 1957. Mechanics Of Hydraulic Fracturing, 210. Petroleum Transactions, AIME.
Jeffrey, R. G., Brynes, R. P., Lynch, P. J., & Ling, D. J. 1992. An Analysis of Hydraulic Fracture and
Mineback Data for a Treatment in the German Creek Coal Seam. SPE. doi: 10.2118/24362-MS.
Johnson, R. L., Glassborow, B., Datey, A., Pallikathekathil, Z. J., & Meyer, J. J. 2010. Utilizing
Current Technologies to Understand Permeability, Stress Azimuths and Magnitudes and their
Impact on Hydraulic Fracturing Success in a Coal Seam Gas Reservoir. SPE. doi: 10.2118/
133066-MS.
King, G. E. 2010. Thirty Years of Gas Shale Fracturing: What Have We Learned? SPE. doi:
10.2118/133456-MS.
Morales, R. H., & Marcinew, R. P. 1993. Fracturing of High-Permeability Formations: Mechanical
Properties Correlations. SPE. doi: 10.2118/26561-MS.
Nolte, K.G. 1991: Fracturing-Pressure Analysis for Non-Ideal Behavior. J. Pet Tech (Feb., 1991)
210 218. http://dx.doi.org/10.2118/20704-PA
Nolte, K. G. 1986. Determination of Proppant and Fluid Schedules From Fracturing-Pressure
Decline. SPE. doi: 10.2118/13278-PA.
Palmer, I.D. and Sparks, D.P. 1991. Measurement of Induced Fractures by Downhole TV Camera in
Black Warrior Basin Coalbeds. JPT 43 (3): 270 275; 326-328. SPE-20660-PA. doi: 10.2118/
20660-PA.
Palmer, I. D. 1993. Induced Stresses Due to Propped Hydraulic Fracture in Coalbed Methane Wells.
SPE. doi: 10.2118/25861-MS.
Pandey, V. J., & Agreda, A. J. 2014. New Fracture-Stimulation Designs and Completion Techniques
Result In Better Performance of Shallow Chittim Ranch Wells. SPE. doi: 10.2118/147389-PA.
Simonson, E. R., Abou-Sayed, A. S., & Clifton, R. J. 1978. Containment of Massive Hydraulic
Fractures. JPT 18 (1): 2732. SPE. doi: 10.2118/6089-PAent.

SPE-173378-MS

27

Sneddon I.N. Elliot H.A. 1946. The opening of a Griffith crack under internal pressure. Applied Math
1946;4:2627.
Warpinski, N.R., Tyler, L.D., Vollendorf, W.C., and Northrop, D.A. 1981. Direct Observation of a
Sand Propped Hydraulic Fracture. Sandia National Laboratories Report No. SAND81 0225.
Livermore, CA (May).
Zoback, M.D. 2007. Reservoir Geomechanics, Cambridge University Press. 449pp.

SI Metric Conversion Factors


cP 1.0*
in 2.54*
ft 3.048*
ft2 9.290 304*
ft3 2.831 685
gal 3.785 412
lbm 4.535 924
psi 6.894 757
bbl 1.589 873
bbl/min 2.6497 884

E-03 Pa.s
E-01 cm
E-01 m
E-02 m2
E-02 m3
E-03 m3
E-01 kg
E00 kPa
E-01 m3
E-03 m3/s

Das könnte Ihnen auch gefallen