Sie sind auf Seite 1von 30

SPE-184865-MS

Analyzing ISIP Stage-by-Stage Escalation to Determine Fracture Height and


Horizontal-Stress Anisotropy

Nicolas P. Roussel, ConocoPhillips

Copyright 2017, Society of Petroleum Engineers

This paper was prepared for presentation at the SPE Hydraulic Fracturing Technology Conference and Exhibition held in The Woodlands, Texas, USA, 24-26 January
2017.

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
ISIP Analysis is a novel analytical method that calculates the hydraulic height of induced fractures and the
in-situ horizontal stress anisotropy from the evolution of instantaneous shut-in pressures during a multi-
stage horizontal completion. The fracture height calculated will be smaller than what is measured through
microseismic measurement, but larger than the propped and effective fracture height.
Because every frac stage contributes to increase minimum horizontal stress and reduce the formation's
horizontal stress anisotropy, ISIP Analysis may be a useful tool to guide the spacing design of perforation
clusters. Increased characterization of hydraulic fracture dimensions from ISIP Analysis also makes it a
useful addition to any workflow looking to optimize well spacing and stacking in unconventional plays.
Multiple formulations of the stress escalation equation were derived, as well as type-curves that relate
hydraulic-fracture height to the parameters of the equation. ISIP Analysis consists in finding the unique
combination of parameters of the stress escalation equation that best matches field ISIP data. Application
of ISIP Analysis is illustrated in the paper using field data taken from wells in various shale formations
across North America. It addresses key uncertainties in the design of unconventional field development and
is being proposed as an inexpensive alternative to other stress/fracture diagnostic techniques. In addition,
the method has been successfully used to design the number of perforation clusters and their spacing to
reach a desired magnitude of stress interference during horizontal-well stimulation.
While other techniques such as microseismic monitoring, tracers, downhole tiltmeters, pressure gauges,
may be utilized to characterize fracture dimensions, the main advantage of ISIP Analysis is the ability to
be applied to a vast majority of wells, without additional hardware, operational delay, or any modification
to the well or completion design. As a result of its simplicity, ISIP Analysis takes a trained completion
engineer only a few minutes to complete.

Introduction
ISIP, or Instantaneous Shut-In Pressure, is the pressure measured at the end of injection of hydraulic
stimulation, after friction forces in the wellbore, perforations and near-wellbore region dissipate (Fig. 1). In
plug-and-perf completions, shut-in pressures are recorded at the end of each stage and have been observed
to increase from one stage to the next (Dohmen et al 2013, Manchanda et al 2012, McClure and Zoback
2 SPE-184865-MS

2013, Roussel and Sharma 2011a, Roussel et al 2012, Vermylen and Zoback 2011). Following shut-in, fluid
inside the hydraulic fracture leaks off into the formation until closure on the proppant is achieved.

Figure 1Typical downhole pressure record during a fracturing stage (Yew and Weng 2015)

Escalation of shut-in pressures can be attributed to the mechanical stresses induced by previous
stages, often referred to in the literature as stress shadowing (Equation 1). Mechanical stress interference
phenomena have tremendous diagnostic value as they relate to the geometry of the induced fractures,
especially hydraulic-fracture height (Soliman et al 2008), as well as the in-situ state of stress (Roussel and
Sharma 2011b).
(1)

With n, the stage number


ISIP, the Instantaneous Shut-In Pressure
pnet, the net pressure in the hydraulic fractures at shut-in
hmin, the in-situ minimum horizontal stress (or closure stress)
shadow, the stress interference contribution, which increases which each new fracturing stage

In unconventional reservoirs, the dimensions of hydraulic fractures are key uncertainties in the design
of multi-well stacking and spacing, which are essential aspects of a field development strategy. Stress
interference mechanisms caused by the hydraulic fracturing process and production of reservoir fluid have
been shown to play an essential role in the performance of multi-well pads in shale plays (Gupta et al 2013,
Manchanda and Sharma 2013). The in-situ state of stress, and especially horizontal-stress anisotropy is of
particular importance, and may impact the spacing of perforations clusters, the sequencing of multi-well
fracturing operations, as well as the timing and design of infill and refracturing operations (Roussel et al
2013).

Diagnostic methods for estimating induced-fracture dimensions


Recent advancements in the field of diagnostics have helped to more accurately characterize the fracture
network, thus allowing for an improved assessment of the completion effectiveness. An induced fracture
may be divided into three different regions (hydraulic, propped, and effective). Out of the three fracture
dimensions, the last one is most relevant to a reservoir model and may be used to forecast future
production. Pressure transient testing (Lee et al 2003) and reservoir history matching (Gilman and Ozgen
2013) are widely used techniques to characterize effective fracture dimensions and fracture conductivity.
Unfortunately, as these methods analyze the combined contribution of all induced fractures and rely on
SPE-184865-MS 3

simplistic assumptions of the induced fracture system, they often lead to non-unique solutions and require
additional data to further constrain the range of potential outcomes.
Proppant tracer logs are sometimes used in vertical well tests, in conjunction with temperature logs, to
evaluate propped fracture height vs. hydraulic fracture height (Chernyshev et al 2016, Davis et al 1997,
Smith et al 1989). With the exception of proppant tracer technology, other fracture mapping technologies
generally evaluate hydraulic fracture dimensions, by interpreting one of its manifestations. Direct fluid
connection with an offset well may be interpreted through pressure and temperature variations, measured by
downhole gauges or fiber-optic cable (Daneshy and Pomeroy 2012, Downie et al 2015, Webster et al 2013).
The dilation of one or several induced fractures is known to cause a stress/strain perturbation in their
vicinity. This phenomenon is often referred to in the literature as "stress shadowing" and "mechanical stress
interference". The alteration of the distribution of stresses and strains during hydraulic fracture propagation
is the basis of some of the most recently-developed fracture-mapping technologies. In shallow tight-sand
plays, surface tiltmeter has historically been used to measure strain variations (Wolhart et al 2007). But
because of the large depth of most shale plays, the small magnitude of strain variations at the surface requires
the use of downhole tiltmeters (Mayerhofer et al 2006, Wright et al 1998). Another manifestation of stress
alterations created by propagating hydraulic fractures is shear slippage of surrounding natural fractures,
which can be captured through microseismic monitoring (Warpinki et al 2004, 2012). ISIP Analysis also
captures mechanical stress interference effects. Following the propagation of a given stage, the stresses in the
vicinity will be elevated, resulting in an increase of frac pressures during the subsequent fracturing stages.

Diagnostic methods for estimating magnitude of in-situ stresses


Well-proven techniques exist to measure two of the principal components, vertical and minimum horizontal
stress. As a result, they often are generally well characterized within an unconventional play, unlike the
maximum horizontal stress (also called the intermediate stress in a normal-faulting regime) which is often
unknown.
When it comes to stress magnitudes, there are large discrepancies on the method accuracy for each of
the three principal stresses. The overburden stress is by far the simplest to measure, by means of a Gamma
Ray density log. The measured density is integrated over the vertical depth, providing a good estimate of v.
The minimum horizontal stress (hmin) is also generally well characterized through fracture tests that may
be conducted during the drilling phase and prior to the well stimulation. Leak-off tests or extended leak-off
tests may be performed in the drilling phase by pressuring the open-hole section right below the casing shoe.
Normally designed to optimize mud density during drilling, the test results may be interpreted to calculate
the minimum horizontal stress (Addis et al 1998).
Another method called mini-frac or DFIT (Diagnostic Fracture Injection Test) consists in injecting a
relatively small volume of fluid at low rates through perforations in a cemented casing, in order to create a
small-scale hydraulic fracture. DFITs may be run at different depths of a vertical wellbore to obtain multiple
stress calibration points, or at the toe of a horizontal well, prior to stimulation operations. Along with closure
pressure, which is equal to the minimum horizontal stress, DFITs may be used to estimate leak-off rate,
reservoir pressure and permeability (Barree et al 2015).
The maximum horizontal stress has long been considered to be the most difficult component of the
principal stress tensor to measure, despite its significance in drilling and completions. It is of particular
importance in multi-stage horizontal fracturing of shale reservoirs where the horizontal stress anisotropy
(difference between maximum and minimum horizontal stress) plays a key role in the ability to stimulate
natural fractures and generate complexity (Manchanda and Sharma 2014, Weng et al 2014). Operationally,
it may impact the spacing of perforations clusters, the sequencing of multi-well fracturing operations, as
well as the timing and design of infill and refracturing operations.
Observations of drilling-induced tensile fractures and compressive wellbore breakouts through wellbore
imaging techniques may help constrain the magnitude of hmax (Zoback et al 2003). Knowing the rock
4 SPE-184865-MS

strength and hmin, hmax may be calculated from the measured breakout width. Application of frictional
faulting theory in areas of active faulting may also provide additional insight (Zoback 2007). Agharazi
(2016a, 2016b) recently proposed a more deterministic method based on stress inversion of microseismic
focal mechanisms induced by hydraulic fracturing, where each qualified event is treated as an independent
field experiment. It is important to note that all the methods cited above, as well as ISIP Analysis, require
a prior, independent characterization of hmin.

Stress Escalation Model (SEM)


Analytical solutions of mechanical stress interference
Sneddon et al (1946) developed analytical solutions of the stress distribution in the vicinity of a dilated crack,
for two extreme cases: semi-infinite fracture (Lf >> hf) and penny-shaped fracture (Lf = hf). In hydraulic
fracturing applications, in particular for horizontal wells in shale plays, induced fractures tend to propagate
more transversely than vertically. Hence, the analytical solution of stress around a semi-infinite fracture
applies best (Equations 2 to 5), where the coordinate system, and fracture geometry are defined in Fig. 2:

(2)

(3)

(4)

(5)

Figure 2Coordinate system for the Sneddon analytical solution of stress around a semi-infinite crack

To quantify the increase in shut-in pressure from stress interference, we may simplify the problem and
focus on the stress increase on a line perpendicular to the fracture face and passing through its center.
Sneddon's equations 2, 3 and 5 can then be simplified to the following (Pollard and Segall 1987, Olson
2004, Equations 6 to 8):
SPE-184865-MS 5

(6)

(7)

(8)

The increase in minimum horizontal stress caused by a dilated crack (xx) was normalized by the net
pressure in the fracture and was plotted in Fig. 3. The quantity is generally referred to in the literature
as the 3D stress correction factor or 3D influence factor (Meyer et al 2013). An analytical solution is also
available for the stress along the axis perpendicular to the fracture face, in the particular case of a penny-
shaped crack (Manchanda and Sharma 2013, Equations 9 and 10):

(9)

(10)

Figure 3Comparison of analytical solutions for the stresses normal


(y=z=0) to a semi-infinite and penny-shaped fracture (Sneddon et al 1946)

Net pressure in the fracture (=pf hmin) changes dramatically following shut-in of a given frac stage.
The stress load (load) represents the net pressure in the hydraulic fracture(s) of one stage just prior to the
start of the subsequent stage, which is the source of induced stress interference (shadow). It thus takes into
account the decline in net pressure resulting from leak off of the frac fluid into the formation taking place
during closure. In practice, for the same net pressure at shut-in, the stress load will be lower as the lag time
6 SPE-184865-MS

(tlag) between subsequent frac stages is prolonged. This may be a strategy, albeit a costly one, to reduce
stress interference effects. Zipper-fracing across multiple wells has been shown to limit stress escalation
by increasing the time between subsequent well stages (Manchanda and Sharma 2013). Other parameters
affecting loadfor a given net pressure at shut-in include frac fluid viscosity, formation permeability, pumped
fluid volume, and natural fractures.

Superposition of mechanical stress interference for multiple consecutive frac stages


Based on the evolution of the minimum closure stress away from a transverse fracture, we may calculate
stress interference for multiple consecutive fractures. The concept is the following: we use equation 6 to
determine the new net pressure in the subsequent frac stage (say stage 2) assuming the stress load (load)
is equal for all stages. We then calculate the new stress interference profile corresponding to the new net
pressure, and assuming stage spacing is uniform, calculate the net pressure of the subsequent stage (stage 3).
We now illustrate the above strategy and assume the stage spacing is equal to half the fracture height, the
increase in closure stress for the second stage would then be equal to 65% of the stress load (Fig. 4a). This,
in turn, would cause the ISIP for the second stage to be elevated by the same quantity. The same process
should be repeated as we march toward the heel of the well. The closure stress for stage 3 will also be
elevated by 65% of the net pressure in stage 2, which is itself equal to 1.65 times the stress load, assuming
the stress load is identical for all stages (Equation 11):

(11)

Figure 4(a) Stress interference away from a transverse fracture at the onset time of the subsequent stage,
and (b) load-normalized magnitude of stress interference versus the number of frac stages for sf /2hf = 0.5

Following the same process for 20 stages, the stress interference (or stress shadowing) component has
been calculated and plotted on Fig. 4b. The superposition of mechanical stress interference for consecutive
frac stages can be summarized by the following recurrence equation (Equation 12):

(12)
SPE-184865-MS 7

Note that closure stress ramps up quickly in the first couple stages, before plateauing at a value
approximately equal to 183% of the stress load. The calculated profile resembles many ISIP trends observed
in the field, thus showing the value of such model for diagnostic purposes.
It is very important to note that while the absolute magnitude of stress interference is influenced by the
geology and the rock's mechanical properties, the way in which mechanical stress interference escalates for
a given stress load is independent of the formation's Young's modulus and Poisson's ratio. As a result, the
analysis of ISIP escalation does not require any geomechanical input.

Extending superposition model for multiple perforations clusters per stage


A similar calculation may be conducted when multiple perforation clusters are stimulated simultaneously.
In this case, the increase in the closure stress will differ from cluster to cluster. The most toe-ward clusters,
located closer to the previous stage, will experience more stress interference than its neighbors. Returning
to the example described in Fig. 4, we still assume that stage spacing is equal to half the fracture length,
except this time, we decide to double the number of perforation clusters per stage.
The toe-ward perforation cluster of stage 2 is now located at a distance equal to just 0.25 times the fracture
height and will experience an increased closure stress of 91% of the stress load (Fig. 5a). Similarly to the
previous example, the second perforation cluster will experience an increased normal stress equal to 65%
of the stress load. Assuming the increased stress interference is the average of both perforation clusters, its
value for the second stage equals 78% of the stress load. The same process takes place for stage 3, taking
into account the increased net pressures in stage 2 (Equation 13).
(13)

Figure 5(a) Stress interference in two perforation clusters away from a previous stage, and (b) load-normalized
magnitude of stress interference for different perforation-cluster designs versus the number of frac stages for sf /2hf = 0.5

After 20 stages, the stress interference in the two-perforation-cluster completion is increased by 90%,
compared to one perforation cluster, for the same stage spacing. The study was extended to completion
designs with up to five perforation clusters per stage. All the cases for one to five perforation clusters are
summarized on Fig. 5b, for a stage spacing equal to half the fracture height.
Note that stress interference does not increase proportionally with the number of perforation clusters. For
the specific ratio of stage spacing over height investigated, there is little difference in stress interference as
8 SPE-184865-MS

the number of perforation clusters per stage exceeds 4. While the magnitude at which stress interference
plateaus changes with the number of perforation clusters, so does the number of the stages needed to
approach this plateau. For the case of 1 perforation cluster, it takes barely more than 5 stages for stress
interference to plateau. For 5 perforation clusters/stage, stress interference is still increasing after 20 stages.

Stress Escalation Equation


Looking at Figs. 4b and 5b, it is hard not to see a general trend in the way closure stress escalates from one
stage to the next. This prompted the author to find a simple equation that would match the calculated stress
interference trends. The result of the investigation led to an exponential recovery equation function of just
two parameters. This form of equation is popular in many physics and engineering application, including
signal theory, and is often used to describe the response of a first-order, linear time-invariant (LTI) system
to a step input (Liptak, 2006). The key parameter characterizing the dynamic response is the time constant
, which is replaced by the Escalation number in Equation 14.

(14)

plateau represents the total value of stress interference that is produced by the stimulation. It may
theoretically be reached only at an infinite number of stages (Equation 15):

(15)

The escalation number represents how "quickly" stress interference approaches the plateau. More
specifically, it represents the number of stages necessary for induced stresses to reach 63.2% of the stress
plateau (Equation 16). This number comes from the form of the equation, and echoes applications of the
exponential recovery equation in other contexts.

(16)

Going back to Fig. 5b, we can now match the analytical model of multi-stage stress interference with the
stress escalation equation (Table 1). For the case of one perforation cluster, the value of the stress plateau
is equal to the value reached to the second digit after stage 16, which is equal to 1.83 times the stress load
(load). As expected, the magnitude of the stress plateau increases with each new perforation cluster, although
unproportionally. For instance, total stress interference is multiplied by 2.7 when the number of perforation
clusters is multiplied by 5. The matching exercise also confirmed our initial suspicion that stress escalation
"slows down" relatively as the magnitude of the stress plateau is increased. Hence, it takes more than twice
the number of stages to reach 63.2% of the stress plateau for 5 perforation clusters, compared to one.

Table 1Matched values of load-normalized stress plateau and


escalation number for different perforation-cluster designs (sf /2hf = 0.5)

Number plateau / load Escalation


of Perf.
clusters

1 1.83 2.27
2 3.51 4.0
3 4.30 4.76
4 4.76 5.26
5 4.94 5.56
SPE-184865-MS 9

Impact of stage spacing and hydraulic fracture height


The stress-shadowing calculations are extended to different values of the sf/2hf ratio, which is the space
spacing distance over the total fracture height. As the stage spacing is reduced, or fracture height is increased,
the stress interference induced by the stimulation grows, as reflected by the higher stress plateau and
escalation values (Fig. 6, Table 2). Another interesting phenomenon is that stress interference does not
increase proportionally as spacing is reduced. For instance, while stress interference may not increase much
when stage spacing is reduced starting from a large value (i.e. from 300 to 250ft), even a small reduction
of 10ft or less may cause a significant increase in stress interference if stage spacing is already tight (i.e.
from 50 to 40ft).

Table 2Matched values of load-normalized stress plateau and escalation number for different sf /2hf ratios

sf / 2hf plateau / load Escalation

0.3 6.34 6.67


0.5 1.83 2.27
0.7 0.86 1.30
2 0.40 0.79

Figure 6Load-normalized magnitude of stress interference versus the number of frac stages for different sf/2hf ratios

By repeating the operation of building the multi-stage stress interference model for multiple stages and
matching it with the stress escalation equation for multiple combinations of perforation clusters per stage,
we can start plotting the relationship of the spacing-to-height ratio to the two matching parameters: the
induced stress plateau (plateau) and the escalation number (Escalation).
The resulting work is presented in Fig. 7. One important conclusion is the fact that there is only one ratio
of stage spacing over height that corresponds to a given value of the escalation number or the stress plateau.
10 SPE-184865-MS

Figure 7Type-curves of (a) the load-normalized stress plateau and (b) the
escalation number for different numbers of perforation clusters per stage

The established figures may thus serve as type-curves for the determination of hydraulic fracture height
for a given multi-stage stimulation, provided the corresponding escalation number is known.

The Interference Ratio


Going back to the simpler case of one perforation cluster per stage, one can plot together the corresponding
type-curves of escalation and load-normalized stress plateau (Fig. 8a). It appears immediately that the two
curves converge for low values of spacing-to-height ratio and tend to diverge as the spacing increases, or
fracture height decreases. Another way to represent this phenomenon is by plotting the ratio of the load-
normalized stress plateau over the escalation number (Fig. 8b). What the trend of Fig. 8b represents is the
relative magnitude of stress interference between subsequent stages, which is always comprised between 0
and 1. The tighter the stage spacing, the larger the induced stress plateau is for a given value of the escalation
number, which is why the quantity of Equation 17 is named Interference Ratio.

(17)

Figure 8(a) Type-curve of load-normalized stress plateau and escalation number for one perforation
cluster per stage, and (b) evolution of the interference ratio with the stage to fracture height ratio
SPE-184865-MS 11

Incidentally, the interference ratio may be graphically calculated directly from the plot of multi-stage
stress interference evolution (Fig. 9a), as it corresponds to the slope of the curve at the origin (Equation
18). The steeper the initial slope, the more frac stages are interfering in stress, with the limit being the slope
reaching a value of 1.

(18)

Figure 9(a) Graphical representation of the Interference Ratio and (b) type-curves
of the Interference Ratio for different numbers of perforation clusters per stage

Similarly to the escalation number and the load-normalized stress plateau, type-curves were developed
for the interference ratio that relate to the spacing-to-height ratio (Fig. 9b).

Analytical derivation of the Stress Escalation Equation


In the previous section, the stress escalation equation was inferred, and it was then verified to accurately
match the stress escalation curves. The parameters of the equation were chosen to represent physical
manifestations of the stress escalation, namely the total amount of stress interference (plateau) and how fast
stress shadowing converges toward the stress plateau (Escalation).
Another way is to derive analytically the solution of the recurrence relation of stress interference for
multiple consecutive stages (Equation 12). The stress correction factor , always comprised between 0 and
1, represents the relative decline in mechanical stress interference perpendicular to a dilated fracture, and is
plotted in Fig. 3 for two simplistic cases: penny-shaped and semi-infinite. Equation 12 is a 1st-order linear
non-homogeneous recurrence equation. To solve this equation, we first need to render it homogeneous by
elevating its order. Thus, for n = n+1, Equation 12 can be rewritten as Equation 19:

(19)

Subtracting Equation 12 from Equation 19 yields a 2nd-order homogenous linear recurrence equation
which can now be solved analytically (Equation 20):

(20)
12 SPE-184865-MS

Solutions to the Equation 21 are of the form shown in Equation 21, for n1, with r1 and r2 the solutions
of the characteristic equation (Equation 22):
(21)

(22)
To determine coefficients 1 and 2, we apply the two known initial conditions to Equation 22 (Equation
23):

(23)

The first stage does not suffer from stress interference from previous stages; hence, the stress shadowing
contribution is equal to 0 at stage 1. The second stage feels the contribution from the first stage, which
is equal to the stress load load times the stress-correction factor . Finally, we obtain a solution to the
recurrence equation (Equation 24), for any stage number n1:

(24)

We can now relate the parameters of the inferred Equation 14 to the stress load (load) and the stress
correction factor (). The stress plateau is the total stress interference induced as the number of stages tends
to infinity (Equation 25):

(25)

Equation 25 may be used to predict, or design the total stress interference induced by a horizontal
stimulation, for multiple cluster numbers and spacing values, and for typical values of the stress load and
hydraulic fracture height. The interference ratio is the slope of the load-normalized stress escalation curve
at the origin (n = 1, Equation 26):

(26)

Finally, the escalation number is related to the stress plateau and the interference ratio through Equation
17, and is consistent with both forms of the stress escalation equation (Equation 27):

(27)

Another advantage of this formulation is the possibility to express the stress escalation behavior for
multiple profiles of the stress correction factor , which represents the stress evolution perpendicular to
one or more dilated fractures. Two analytical expressions of the stress correction factor are given below,
all derived from Sneddon's work. One assumes a semi-infinite fracture (Equation 28), and the other one
assumes a penny-shaped fracture (Equation 29). More realistic stress decline profiles may be calculated
numerically and used in conjunction with Equation 24.

(28)
SPE-184865-MS 13

(29)

Assumptions and limitations


Probably the most notable advantage of the proposed method is its simplicity both in the number of
parameters involved (stage spacing, fracture height and number of perforation clusters per stage) and in its
implementation. The corresponding downside is that it carries a certain number of assumptions that may
limit the range of applications. Another potential issue is that applying such method requires additional
focus in conducting stimulation operations in the field, so that analyses of stimulation data may be relevant.
The Stress Escalation Model assumes uniform: (1) stage spacing, (2) perf. cluster spacing, (3) number of
perforation clusters per stage, (4) stimulation design (especially volume of fluid pumped per stage), (5) lag
time between successive stages, (6) hydraulic height, and (7) mechanical properties.
Generally, the first three assumptions are respected in the field, at the exception of a few pilot tests where
perf. cluster spacing and number/stage may be altered along the well. Stress interference in such completion
will be very difficult to analyze other than qualitatively. The fourth assumption is also generally respected.
Typical exceptions include screen-outs and mechanical failure of pumps or blender. Also the first couple
toe stages are sometimes designed differently than the rest of the stages to prevent screen-outs, altering
fluid type or volume. Since the most valuable ISIP data points are the ones corresponding to the first couple
stages, it is recommended that such designs be avoided in order for the SEM to be applied.
The fifth assumption is a little trickier to maintain during the entire completion, especially in the event
of a screen-out or equipment failure that may require well intervention or repairs, hence modifying the
time interval between successive stages. Again, it is important to keep in mind most of the diagnostic value
resides in the first five to ten stages. As a result, additional attention to maintaining consistency in frac
operations should be focused on the toe stages.
It is generally acceptable to assume small lateral variations of mechanical properties. Yet on the other
hand, Young's modulus and Poisson ratio may change drastically between layers. Hence, drilling operations
may have a profound impact on the applicability of the proposed model. It is very common that the heel part
of the well may be drilled in a higher section than the rest of the well. In that case, it may be important to
discard the shut-in pressure values of the later stages. More details on the impact of the model's assumptions
on the choice of data points to be considered for analysis are provided in the section dedicated to QA/QC.

The Role of Horizontal-Stress Anisotropy


One fundamental assumption of the Stress Escalation Model (SEM) is that all induced fractures are perfectly
transverse to the horizontal wellbore. In reality, mechanical stress interference materializes not only as a
change of stress magnitude, but also of direction, causing hydraulic fractures to curve either toward or
away from previously placed fractures (Bunger et al 2012, Wu and Olson 2013). As identified by Roussel
and Sharma (2011a, 2011b), multi-stage stress interference and the resulting increase in shut-in pressures,
is hampered by the reorientation of hydraulic fractures as the minimum horizontal stress approaches the
maximum horizontal stress (in a normal-faulting stress regime). In the following section, we focus on the
reorientation of stresses produced by mechanical stress interference, its impact on the evolution of shut-in
pressures, and the key role played by horizontal-stress anisotropy.

Stress reorientation from mechanical stress interference


Mechanical stress reorientation can be better understood by looking at the distribution of maximum
horizontal stress orientation in the vicinity of a dilated fracture. Fig. 10a shows two zones of stress
reorientation: a repulsion zone, in which the subsequent fracture will tend to deviate away from the previous
stage, and a stress reversal region, which may cause the subsequent fracture to propagate longitudinally
14 SPE-184865-MS

to the wellbore. When a bi-wing fracture propagates in a non-transverse direction, another zone of stress
reorientation appears, called the attraction zone (Fig. 10b). If the subsequent fracture is initiated in this zone,
it will tend to propagate toward the previous stage. Two scenarios may arise: either the induced fracture
reaches the repulsion zone away from the wellbore and will deflect away from the previous stage, or it
may enter the stress reversal zone and be drawn to intersect the previous stage. The propensity for stress
reorientation is governed by three major factors: net pressure in the fracture (=pf hmin), hydraulic fracture
height (2hf), and horizontal-stress anisotropy (hmax hmin). The lower the in-situ stress anisotropy, the
more the direction of horizontal-stress anisotropy will deviate from its original direction during multi-stage
fracturing.

Figure 10Distribution of maximum horizontal stress in the vicinity of (a) a


transverse and (b) an inclined dilated fracture (Roussel and Sharma 2011b)

In summary, fracture reorientation tendencies, and deviation from the transverse fracture direction,
will increase as the stage spacing or the horizontal stress anisotropy decreases. The relationship between
fracture reorientation and fracture spacing is illustrated in a numerically simulated example of a multi-stage
horizontal well in the Barnett shale (Fig 11).
SPE-184865-MS 15

Figure 11Trajectory of multiple consecutive fractures spaced 400ft, 300ft, 250ft, and 150ft apart (Roussel and Sharma 2011b)

In a very simplified approach, it was assumed only one fracture is propagated for each stage, and
that the fracture exactly follows the direction of maximum horizontal stress. The stress distributions
were numerically calculated by means of a three-dimensional finite-difference geomechanical simulator
(FLAC3D), for an in-situ horizontal stress anisotropy equal to 100psi, and a hydraulic fracture height equal
to 300ft. For a 400-ft fracture spacing, the fractures deviate slightly away from the previous stage. As the
spacing is decreased, the angle of deviation increases (sf= 300ft).
When reduced further (sf= 250ft), the spacing causes fractures to be initiated in the attraction zone,
resulting in some of the fractures to intersect previous stages, alleviating some of the stress interference.
Finally, as the fracture spacing is reduced to 150ft, some of the stages are now initiated in the stress reversal
region. The consequence is that half of the stages are now propagating longitudinally to the wellbore.

Impact of fracture reorientation on shut-in pressure escalation


As fractures reorient, and finally propagate in a longitudinal direction, the stress exerted by the multiple
dilated fractures will relax. This phenomenon will be reflected in the evolution of shut-in pressures.
Roussel and Sharma (2011b) numerically calculated the shut-in pressure trends associated with the fracture
geometries shown in Fig. 11 (Fig. 12).
16 SPE-184865-MS

Figure 12Evolution of the net pressure in the fracture for multiple


stage spacing configurations (for a stress load load = 245psi)

As the spacing decreases from 400ft to 300ft, stress interference increases, causing the shut-in pressures
to escalate higher. In the first two cases, the shut-in pressures only increase with each new stage, as would be
predicted by the SEM. But when some of the fractures become attracted by the previous stages, it causes the
shut-in pressure of the subsequent stage to drop (sf = 250ft). When half of the fractures are longitudinal, the
shut-in pressure now oscillates around a plateau value (sf = 150ft). The shut-in pressures from the numerical
model were matched with the SEM, and the results are provided in Table 3.

Table 3Analytically calculated values of hydraulic height and horizontal-stress anisotropy using SEM

sf plateau
Escalation 2hf (ft) hmax - hmin (psi)
(ft) (psi)

150 118 0.20 51 118


200 98 0.28 87 98
250 80 0.41 146 80
300 82 0.92 322 >82
400 23 0.52 278 >23

As stage spacing decreases, the stress plateau increases, reflecting higher stress interference. For the cases
that exhibit a monotonic stage evolution of the net pressure (sf = 300ft and 400ft), it is possible to calculate
the hydraulic fracture height using the type-curves of escalation number. The average value of hydraulic
height for the two cases is 300ft, which is exactly equal to the height of the dilated fractures in the numerical
simulator. We also verify that stress plateau values are inferior to the in-situ horizontal-stress anisotropy.
As the spacing is reduced to 250ft, an oscillating behavior starts to be exhibited, indicating that stress
anisotropy has been overcome by stress interference, causing fracture reorientation. At that point the
SEM can no longer be employed to calculate hydraulic fracture height, as stress relaxation caused by
the reorientation of hydraulic fractures prevents shut-in pressures to further escalate. As a result, when
matching the stress escalation equation to shut-in pressures in the field, the escalation number will drop
substantially, causing the analytical type-curves to underestimate the value of the hydraulic fracture height.
This phenomenon can be seen in Table 3, where calculated heights drop well below the actual height as
spacing is reduced (sf = 250, 200, and 150ft).
Comparing the stabilized value of the stress plateau (when an up-and-down trend is exhibited) to the
in-situ value of horizontal-stress anisotropy, demonstrates the potential for ISIP Analysis to quantify the
in-situ difference between maximum and minimum horizontal stress. The average stress plateau for stage
SPE-184865-MS 17

spacing sf = 150, 200 and 250ft is 99psi, which is remarkably close to the 100psi stress anisotropy used
in the numerical simulator. By applying ISIP analysis to numerically calculated values of the evolution of
net pressure during multi-stage fracturing, we verify that the induced stress plateau is always inferior to the
in-situ horizontal stress anisotropy (in a normal faulting environment, Equation 30), and that by reducing
stage spacing, it is possible to use ISIP Analysis to calculate the in-situ horizontal stress anisotropy and thus
calculate hmax, if the closure stress (hmin) is known.
Other mechanisms than fracture reorientation, such as increased branching with natural fractures and
fracture complexity are likely to contribute in limiting ISIP escalation when approaching an isotropic stress
condition. It is also important to note that when analyzing field data, because of geological variability, it
is generally difficult to visually identify an up-and down behavior caused by fracture reorientation. In the
following section, a method will be presented to identify when horizontal-stress anisotropy is overcome.
(30)

Sensitivity to the magnitude of horizontal-stress anisotropy


To test the validity of Equation 30, ISIP Analysis was applied to additional numerical simulations from
Manchanda et al (2012). Results of three numerical simulations of shut-in pressure escalation representative
of Barnett shale are plotted in Fig. 13. All parameters are kept the same, except for the horizontal-stress
anisotropy, which is varied from 100 to 400psi. It should be noted that in all three cases, the stresses induced
during the completion are enough to offset in-situ stress anisotropy, and lead to fracture reorientation. The
stress escalation equation is used again to match the results of the numerical simulation, and the results are
shown in Table 4. In all three cases, the value of the induced stress plateau obtained by matching numerical
results is remarkably close to the value of the in-situ horizontal stress anisotropy used in the numerical
model, thus confirming the validity of Equation 30.

Figure 13Evolution of the net pressure in the fracture for multiple


values of the horizontal-stress anisotropy (and a stress load load= 370psi)

Table 4Matched values of the induced stress plateau and escalation number

Input stress
plateau (psi) Escalation
anisotropy (psi)

100 136 0.02


200 205 0.25
400 382 0.79
18 SPE-184865-MS

ISIP Analysis Workflow


In the previous section, the Stress Escalation Model (SEM) was introduced and highlighted for its potential
in calculating hydraulic fracture dimensions during a multi-stage horizontal completion, as well as the
magnitude of in-situ horizontal stress anisotropy (hmax - hmin). The following section details the several
steps to follow in a typical ISIP Analysis, from matching the ISIP data to extracting its diagnostic value.

Matching ISIP data


The goal of the first step is to match the stress escalation equation with the shut-in pressure data. The most
straightforward method consists in minimizing the sum of the squared deviations between the data and the
stress escalation equation (least-squares regression). A few solvers are available in Excel allowing an easily
implementation of ISIP Analysis. At the end of the matching process, we end up with a value of the stress
plateau and escalation representative of the well. This is illustrated in Fig. 14 using a field case from shale
formation A.

Figure 14Matching of shut-in pressures with the stress escalation equation for well A1

Determining what causes the stress plateau


We have identified two factors that may be responsible for the existence of a stress plateau. Because of
the decline of mechanical stress interference away from a dilated fracture, the superposition of multi-stage
stress interference always reaches a maximum value. We also identified that if the stress induced by the
completion reaches a value equal to the horizontal-stress anisotropy, fracture reorientation will take place
providing a stress-relieving mechanism, thus preventing the shut-in pressure to increase any higher.
Differentiating the two factors is crucial, as the implications for fracture/stress diagnostics change. In the
first case, the matching process will lead to the calculation of an average hydraulic height. In the second
case, the implication of fracture reorientation violates a fundamental assumption of the SEM, which renders
the hydraulic height estimate invalid, but will inform on the in-situ horizontal stress anisotropy.
Below is a proposed workflow to evaluate whether the plateau is naturally occurring, or caused by
overcoming the maximum horizontal stress (Fig. 15). It relies on calculating the stress load (load) and
comparing it to the net pressure at shut-in. When horizontal-stress anisotropy limits the escalation of shut-
in pressures, it causes the application of the SEM to result in low values of the hydraulic height and high
values of the stress load.
SPE-184865-MS 19

Figure 15Workflow to determine phenomenon responsible for stress plateauing

After matching ISIP data with the stress escalation equation, we now have a value for the stress plateau
and escalation. From there we can calculate the interference ratio: (1) use the escalation type-curve to
obtain a value of the spacing-to-height ratio (Fig. 7b) and (2) use the interference ratio type-curve to read
the interference ratio corresponding to the same spacing-to-height ratio (Fig. 9b). Finally using the stress-
interference equation, we are able to back calculate the stress load.
If the plateau is naturally occurring, the stress load will have a value that is much smaller than the net
pressure at shut-in. On the other hand, when maximum horizontal stress is overcome and stress escalation
is cut short, stress loads will take abnormally high values.
There is a clear indication that horizontal stress-anisotropy is responsible for the stress plateau if the
calculated stress load becomes higher than the net pressure at shut-in, which is physically impossible.
Generally, if the stress load is less than half the net-pressure at shut-in, this should give some confidence
that fracture reorientation is limited and that the height calculation may be trusted. There are many factors
that may affect the cut-off ratio for the stress load over shut-in net pressure. For instance, a small value of
the formation permeability, or of the lag time between stages, will tend to increase that ratio.
Returning back to the example of Fig. 14 and using the 5 perforation cluster/stage type curve (Fig. 7b),
we find that an escalation number of 3.0 corresponds to a spacing-to-height ratio of 0.72. Then from the
interference ratio type curve (Fig. 9b), we find that the interference ratio is equal to 0.84. The stress load
may now be calculated (Equation 31).

(31)
20 SPE-184865-MS

Based on DFITs, hmin may be calculated for the well in question, hence the net pressure at shut-in for
stage 1 (Equation 32):

(32)

The analysis indicates the stresses induced were not large enough to overcome horizontal-stress
anisotropy, thus hmax > hmin + plateau > 6168psi.

Calculating hydraulic fracture height or horizontal-stress anisotropy


After verifying that the plateauing of the ISIPs during multi-stage fracturing cannot be attributed to fracture
reorientation, we can now calculate the average hydraulic fracture height, which along with fracture spacing
is the only factor impacting stress escalation, according to the SEM. All we need to know is the escalation
number for a given well, number of perforation clusters/stage and stage spacing. Applying the escalation
type-curve to the example of Fig. 14, we find that for 5 perforation clusters/stage, the stage spacing-to-
height ratio is equal to 0.72, which results in a hydraulic fracture height of 242ft (35-ft perf. cluster spacing,
Equation 33).

(33)

Calculating hydraulic fracture height is fairly trivial in the case described above, but becomes trickier if
horizontal-stress anisotropy is overcome during fracturing. In that case, the escalation of stresses is being
halted, causing the escalation number to decrease substantially, which will result in a widely underestimated
value of the hydraulic fracture height.

Field Application Examples


In this section, we go through the multiple applications of ISIP Analysis. Considering the few methods
available to characterize maximum horizontal stress, we emphasize how applying ISIP Analysis to multiple
wells across a formation may narrow down considerably the uncertainty range. The other potential outcome
of the method is the evaluation of hydraulic fracture height, which may be used to calibrate hydraulic
fracturing models and influence stacking strategy in thick shale formations. In addition, as stress induced
by the completion is strongly influenced by stage and perf. cluster spacing, the method may be used in
conjunction with other diagnostic methods to help guide completion optimization, especially in the early
appraisal phase.

Narrowing down range of horizontal-stress anisotropy


Knowledge of the horizontal-stress anisotropy is critical for many aspects of the design of multi-stage
completions and field development. Its in-situ magnitude may influence the ability of the hydraulic fractures
to interact with the natural fractures and generate a complex fracture network. The impact may also be felt
when trying to stimulate infill wells following production of one or several parent wells. Depletion may not
only impact stress magnitude, but also reorient stresses in the field, such that the propagation direction of
fractures initiated from an infill well may differ from the preferable transverse direction (Gupta et al 2013,
Roussel et al 2013).
We already know from the field example of Fig. 14 that the in-situ horizontal-stress anisotropy in shale
formation A is higher than 810psi. We analyzed three additional wells, whose ISIP data and matches are
SPE-184865-MS 21

plotted in Fig. 16. Well A4 has a cluster spacing similar to well A1 but a smaller lag time as it was not
zipper-fraced. Perforations clusters in the other two wells (wells A2 and A3) are much more closely spaced
(17ft), and thus will contribute to induce higher stress interference.

Figure 16ISIP evolution and match of three wells in shale formation A

The results of the analysis of all three wells ISIPs are provided in Table 5. Following the workflow of
Fig. 15, we can clearly see that stress anisotropy has been overcome for wells A2 and A3, for the calculated
stress load is higher than the net pressure at shut-in. Even though stage spacing has been reduced by 50%, the
lower values of interference ratio compared to A4 provide another indication. Calculated hydraulic heights
are meaningless in this case, but in-situ horizontal-stress anisotropy may be extracted. Values of hmax hmin
for both wells are very close, with the average value being 1392psi.

Table 5Calculated values of hydraulic height and horizontal-stress anisotropy using SEM

Interference Calculated Calculated


Well sf (ft) plateau (psi) Escalation sf / 2hf load (psi)
Ratio 2hf (ft) h (psi)

A2 85 1440 0.92 1.72 0.56 2795 > 49 ~1440


A3 85 1344 1.63 1.13 0.72 1227 > 102 ~1344
A4 140 1108 6.0 0.49 0.91 264 292 >1108

On the other hand, Well A4 does not appear to overcome horizontal-stress anisotropy, as the calculated
stress load comes up well below the net pressure at shut-in. This is a strong indication that horizontal-
stress anisotropy is higher than 1108psi. This study demonstrates that analyzing ISIPs for multiple wells in
a similar area can tremendously narrow down the range of values for horizontal-stress anisotropy.

Estimating hydraulic-fracture height


One of the strengths of the proposed method is its ability to repeat the analysis across multiple wells and
gain more confidence on the hydraulic height of induced fractures as consistent estimates are seen. For the
shale formation A, hydraulic fracture height was calculated for a multiplicity of completion designs, and
a total of 7 wells (Table 6).
22 SPE-184865-MS

Table 6Calculated values of hydraulic height for 7 wells in shale formation A

Well scluster (ft) #clusters Fluid type 2hf (ft)

A1 35 5 Slickwater 242
A4 35 4 X-linked gel 292
A5 35 5 Slickwater 243
A6 35 5 Slickwater 235
A7 35 4 X-linked gel 192
A8 35 4 Slickwater 231
A9 35 5 Slickwater 298

The average value of the calculated hydraulic height for the seven wells analyzed above is 248ft, with a
standard deviation of 33ft. Applying ISIP analysis on a just few wells has provided confidence that hydraulic
fractures propagate vertically most likely between 215ft and 281ft. It is important to note here that these
estimates may differ substantially from propped and effective heights. Vertical compartmentalization has
been often observed as the shale laminations create pinch points that may prevent the transport of proppant
across the entire height, or limit its effectiveness. Hydraulic-height estimates still remain valuable insights
as they can be used to calibrate hydraulic fracturing simulations, and evaluate the probability of frac hits
with wells situated at higher/lower intervals. In the early appraisal phase, they may also shed light on the
potential of a given play, as low hydraulic heights may be an indication of potentially lower productivity.

Optimizing fracture spacing in multi-stage completions


The analysis of ISIPs may guide the process of decreasing stage and perf. cluster spacing and shed light
on the amount of stress needed to overcome the in-situ horizontal stress anisotropy, therefore favor the
propagation of complex fracture networks especially in very low permeability rocks. In naturally-fractured
formations, spacing the perforation clusters so that a near-isotropic condition is reached may considerably
increase the surface area stimulated, hence improving the well productivity. The completion design needed
to achieve such goal will depend on the magnitude of in-situ horizontal stress anisotropy, hydraulic fracture
height, the spacing between perforation clusters, and the number of perf. clusters per stage.
In other formations that experience a strike-slip stress regime, meaning that the overburden stress is the
intermediate principal stress (hmin < v < hmax), stress escalation may lead to the formation of horizontal
fractures. Contrary to a normal-faulting regime where reaching the intermediate stress tends to improve
well productivity, a tendency for horizontal propagation may severely constrain height growth, the vertical
effectiveness of the stimulation treatment, limit proppant concentration, or worse cause screen-outs. In this
context, the goal will be to design the completion to avoid "bumping" into the intermediate stress.
Starting from the results of ISIP analysis on well A, we can start to evaluate how changing the fracture
spacing would impact the stress induced by the completion. Now that we know the hydraulic fracture height
(2hf = 242ft), it is possible to calculate the stress correction factor for various combinations of perf. cluster
spacing and number of perf. cluster spacing / stage using Equation 28. For instance in a 25-ft cluster spacing
and 5 perf. clusters / stage scenario, the correction factor would be equal to 0.992, 0.944, 0.854, 0.741, and
0.629 for each of the perf. clusters. The calculated average stress correction factor across the five perforation
clusters is equal to 0.832.
Assuming an unchanged stress load (load = 320psi) and using Equation 25, the total stress induced by the
completion would be equal to 1585psi. The same process was repeated for many different perf. cluster and
stage spacing combinations (Fig. 17). The initial configuration of well A is indicated by the grey point. We
can imagine a hypothetical scenario where the stress induced by the completion should stay below 2000psi
in order not to overcome the overburden stress. With 3 perf. clusters/stage, the perf. cluster spacing would
SPE-184865-MS 23

have to be 35ft or higher. With 4 perf. clusters/stage, cluster spacing could be reduced down to 27ft, and to
22ft and 19ft respectively for 5 and 6 perf. clusters.

Figure 17Optimization of perf. cluster and stage spacing for shale formation A

QA/QC and Data Acquisition Guidelines


Going through the ISIP Analysis workflow takes little time, especially if the method has been implemented
in an Excel spreadsheet. More often than not, QA/QC of the ISIP data is the most time-consuming phase
of the analysis and a careful review is needed to avoid misinterpretation. Simple recommendations can
be provided on how to execute completions in order to facilitate the analysis and ensure that maximum
information is being extracted through ISIP Analysis.

Analysis of non-instantaneous shut-in pressure data


Sometimes the values of wellhead pressure at 3, 5, or sometimes 10 minutes past shut-in are provided along
with ISIPs. These pressure measurements have the advantage of being more consistent than ISIP, as they
generally occur after the water hammer period following shut-in, which is indicative of a good connection
between the well and the induced fractures. The question then arises on whether the method described above
yields similar results for all shut-in pressure curves (i.e. instantaneous, 3-min, 5-min) and consequently
which curve should be analyzed. An example of a well located in the shale formation B is shown below,
for which we have 3 shut-in pressure curves (Fig. 18a).
24 SPE-184865-MS

Figure 18(a) Instantaneous, 3-min, and 5-min shut-in pressure data for well B1 and (b) their respective SEM match

As expected the pressure in the induced-fracture system drops quite rapidly in the first few minutes
following shut-in, as the fluid in the fractures leaks off in the formation and the fractures gradually close.
The shut-in pressure curves thus shift down as more time elapses between shut-in and when the pressure is
being recorded. The three shut-in pressure escalation curves are then analyzed using the SEM.
It is clear from Fig. 18b and Table 7 that the total stress induced by the completion decreases as the time
after shut-in increases. There is also an impact on the stress escalation, which is considerably "slowed-down"
for higher shut-in times. Consequently, the analysis yields very different height estimates depending on
which curve is used, with the hydraulic height being more than doubled when using 5-min shut-in pressures
versus instantaneous shut-in pressures.

Table 7Calculated hydraulic heights for instantaneous, 3-min, and 5-min shut-in pressures for well B1

Time after Calculated


plateau (psi) Escalation
shut-in hydraulic height (ft)

0 701 1.38 154


3 min 612 2.78 263
5 min 583 4.0 325

This example shows that frac fluid leak-off is highly stress-dependent. Leak-off accelerates with each
new frac stage as stress interference builds up, and the normal stress exerted on the fractures increases. As
a result, the stress load in the latter stages will be less than the stage load at earlier stages, when looking at
non-instantaneous shut-in pressures, which violates a fundamental assumption of the SEM. For this reason,
analyses of non-instantaneous shut-in pressure data will result in erroneous evaluations of the total stress
induced by the completion and the hydraulic fracture height. Hence, it is very important that the ISIP analysis
workflow be only applied to instantaneous shut-in pressure data.

Weeding out outlier points


There are many reasons why ISIPs may deviate from the trend characteristic of the stress escalation equation.
Some of the factors may be operational in nature while others may be related to the geology: stage screen-
out, inconsistent slurry volumes or fluid type, inconsistent lag time between stages, well trajectory, vertical/
lateral heterogeneity in mechanical properties, faults.
SPE-184865-MS 25

In the event of a screen-out or equipment failure during a frac stage, it should be fairly straightforward
to identify the ISIP data point and exclude it from the match. This is one of the reasons why completion
engineers are well suited to be running the ISIP Analysis, as they are most knowledgeable in operational
factors.
Because outlier ISIP values may impact the quality of the ISIP match, it is imperative to take them out
of the analysis. It is sometimes easier said than done as detecting outlier ISIPs is especially tricky in earlier
stages, which are unfortunately most meaningful. An example encountered in shale formation C highlights
that difficulty (Fig. 19). The evolution of ISIP in the first two stages of well C1 is clearly inconsistent with
a typical stress-escalation behavior (ISIP(2) < ISIP(1)). Now the difficulty is to determine which point(s)
may be problematic. This was achieved by comparing ISIP data in well C1 to a nearby well in a similar
formation (C2).

Figure 19ISIP gradient for two wells in shale formation C

Comparing the two wells, it becomes clear the ISIP of stage 1 for well C1 is too high, and the stage
2 ISIP may also be a little low. Looking at the completion operations in more detail, it was found that a
couple hours prior to the first stage a toe DFIT was conducted on the same well, explaining the abnormally
high value of ISIP for the first stage, as extra pressure and stress were present in the near-wellbore region
following the DFIT. As a result, to complete the analysis of well C1, stage-1 ISIP was decreased to the
same gradient value as well C2.
Because the stress escalation equation yields a relatively flat behavior of ISIP after a dozen stages, ISIP
variations occurring in the heel stages may generally be excluded from the analysis, as they can only be
explained by geological/operational factors. Three different matches were conducted on a well in shale
formation B (Fig. 20) including: (1) all ISIP data points, (2) solely the 12 first stages at the toe of the well
(to exclude variations in the heel section due to geological factors) and (3) same toe stages at the exception
of stage 6, which exhibits an abnormally high ISIP value, possibly caused by a nearby fault.
26 SPE-184865-MS

Figure 20Three matches of ISIP data using stress escalation equation for different value sets (well B2)

In this particular case, the three different ISIP data sets yielded relatively similar results (Table 8).
Nevertheless, the third match including only the 12 toe-most stages at the exception of stage 6 would be the
recommended choice. A similar process should be conducted in most ISIP analyses to exclude points that
may not be relevant and check for the stability of the ISIP matches for different sets of ISIP values.

Table 8Results of the ISIP match for different value sets (Well B2)

Match plateau (psi) Escalation

All points 753 0.77


12 toe stages 779 0.85
12 toe stages w/o outlier 731 0.73

Evaluating data and match quality


Generally, when spacing between perforation clusters is small and the stresses induced by the completion
are high, differences between ISIP data and the stress escalation equation will be small, as the impact of
geology on ISIP values will be minimal compared to stress-shadowing phenomena. For instance, for the
well of Fig. 14, absolute variance between data and match is 140psi, which corresponds to a relative variance
of 19%. When excluding the 8 heel-most stages from the analysis, which feature significant ISIP swings
from geological variations, absolute variance drops to 71psi and relative variance to 11%.
On the other hand, when the spacing between perforation clusters is significant (>100ft), the stresses
induced by the completion may be in the same magnitude or even smaller than geology-driven ISIP
variations. In those circumstances, ISIP swings will appear random, and the relative variance will be high,
as previously observed by Daneshy (2015a, 2015b, 2015c) and Daneshy et al (2015). Variations in the
completion design between the different frac stages may also contribute to hiding the stress-escalation trend.
Hence in some instances, it may not be possible to match ISIP data with the stress escalation equation.

Recommendations for ISIP data acquisition


Because QA/QC is a time-consuming phase and ISIP data quality is key to obtaining relevant information
from ISIP Analysis, focus should be brought on simple operational guidelines to improve data quality. The
most important consideration for completion operations is to maintain a consistent design across all stages.
A consistent design includes maintaining the same fluid type, slurry volume, proppant mass, pumping rate,
lag time between frac stages, number of perf. clusters per stage, and perf. cluster spacing, especially in early
frac stages since they have the most significant diagnostic value.
SPE-184865-MS 27

It is not uncommon to have the first toe stage be designed differently to avoid a potential screen-out.
Such practice may unfortunately lead to erroneous results, as accuracy in the first is essential to a successful
ISIP Analysis. Conducting a toe DFIT prior to the first frac stage may also significantly impact the first-
stage ISIP.
Another source of error is the method used to pick instantaneous shut-in pressures. Picking ISIPs has
been and still is a contentious topic in the literature and generally in the industry (Barree et al 2015). First
and foremost, one should be consistent in the methodology used for picking ISIP for each stage. In many
instances, a phenomenon called water hammer will be observed following shut-in, which results from a
sudden change in flow velocity in the confined system. This transient phenomenon may complicate the
identification of the right ISIP. One method consists in extrapolating the slope of the pressure decline at
the end of the water hammer, and pick the pressure value as the straight line intercepts the time when the
well was shut-in (Mondal 2010).

Summary
A new analytical method has been introduced to calculate hydraulic fracture dimensions and in-situ
horizontal stress anisotropy, from the escalation of instantaneous shut-in pressures in a multi-stage
horizontal completion. The two formulations of the stress escalation equation, the type-curves of the load-
normalized stress plateau, escalation, and interference ratio, as well as the fracture-reorientation criterion,
altogether form a complete and consistent framework that we refer to as ISIP Analysis.
Application of ISIP Analysis was illustrated by field examples taken from different shale formations. It
addresses key uncertainties in the design of unconventional field development, is proposed as an inexpensive
alternative to other stress/fracture diagnostic techniques, and shows potential in guiding the spacing of
perforation clusters. Because it relies on simple analytical equations and readily available data for most plug-
and-perf wells, ISIP Analysis may be applied in a systematic fashion in all unconventional shale reservoirs.

Acknowledgement
The author would like to thank the management of ConocoPhillips for supporting the work presented in
this paper and their permission to publish.

Nomenclature
ISIP Instantaneous shut-in pressure
ISIP(1) of toe-most stage
SEM Stress Escalation Model
pf Fracturing pressure
pnet Net fracturing pressure (=pf - hmin)
shadow(n) Stress interference at n stage
plateau Induced stress plateau
load Stress load (<pnet)
Escalation Escalation number
Interference Interference Ratio
hmin Minimum stress
hmax Maximum stress
v Overburden stress
h Horizontal stress anisotropy (hmax hmin)
hf Fracture half-height
Lf Fracture half-length
28 SPE-184865-MS

sf Spacing between fracturing stages


scluster Spacing between perforation clusters
ncluster Number of perforation clusters per stage
tlag Time between consecutive frac stages
Poisson ratio

References
Addis, M.A., Hanssen, T.H., Yassir, N., Willoughby, D.R., and Enever, J. 1998. SPE 47235 presented at the SPE/ISRM
Eurock 98, Trondheim, Norway.
Agharazi, A. 2016a. Determination of Maximum Horizontal Field Stress from Microseismic Focal Mechanisms A
Deterministic Approach. ARMA 16-691 presented at 50th US Rock Mechanics /Geomechanics Symposium, Houston,
Texas, USA.
Agharazi, A. 2016b. Determining Maximum Horizontal Stress With Microseismic Focal Mechanisms Case Studies
in the Marcellus, Eagle Ford, Wolfcamp. URTeC 2461621 presented at the Unconventional Resources Technology
Conference in San Antonio, Texas, USA.
Barree, R.D., Miskimins, J.L., and Gilbert, J.V. 2015. Diagnostic Fracture Injection Test: Common Mistakes, Misfires,
and Misdiagnoses. SPE Prod & Oper 30 (2): 8498. SPE-169539-PA.
Bunger, A.P., Zhang, X., and Jeffrey, R.G. 2012. Parameters Affecting the Interaction Among Closely Spaced Hydraulic
Fractures. SPE Journal 17 (1): 292306. SPE-140426-PA.
Chernyshev, A., Podberezhny, M., Astafyev, V., Timakov, E., and Koplik, A. 2016. SPE 182117 presented at the SPE
Russian Petroleum Technology Conference and Exhibition, Moscow, Russia.
Daneshy, A. and Pomeroy, M.D. 2012. In-situ Measurement of Fracturing Parameters from Communication Between
Horizontal Wells. SPE 160480 presented at the SPE Annual Technical Conference and Exhibition, San Antonio, Texas,
USA.
Daneshy, A. 2015a. Dynamic Interaction with Limited Entry Fractures in Horizontal Wells: Theory, Implications and
Field Verification. SPE 173344 presented at the SPE Hydraulic Fracturing Technology Conference, The Woodlands,
Texas, USA.
Daneshy, A. 2015b. Stress Shadowing: The Compelling Evidence? Hydraulic Fracturing Journal 1 (1): 9698.
Daneshy, A. 2015c. Dynamic Active Fracture Interaction (DAFI) in Horizontal Wells. Hydraulic Fracturing Journal 2
(2): 819.
Daneshy, A., Touchet, C., Hoffman, F., and McKown, M. 2015. Field Determination of Fracture Propagation Mode
Using Downhole Pressure Data. SPE 173345 presented at the SPE Hydraulic Fracturing Technology Conference, The
Woodlands, Texas, USA.
Davis, E.R., Zhu, D., and Hill, A.D. 1997. Interpretation of Fracture Height From Temperature Logs The Effect of
Wellbore/Fracture Separation. SPE Formation Evaluation 12 (2): 119124. SPE-29588-PA.
Dohmen, T., Zhang, J., and Blangy, J.P. 2014. Measurement and Analysis of 3D Stress Shadowing Related to the Spacing
of Hydraulic Fracturing in Unconventional Reservoirs. SPE 170924 presented at the SPE Annual Technical Conference
and Exhibition, Amsterdam, The Netherlands.
Downie, R.C., le Calvez, J., Dean, B.K., and Rutledge, J. 2015. Correlation and Interpretation of Microseismic Responses
Using Pressure Measurements in Offset Observation Wells. SPE 173386 presented at the SPE Hydraulic Fracturing
Technology Conference, The Woodlands, Texas, USA.
Gilman, J.R. and Ozgen, C. 2013. Reservoir Simulation: History Matching and Forecasting, Society of Petroleum
Engineers.
Gupta, J.K., Albert, R.A., Zielonka, M.G., Yao, Y., Templeton-Barrett, E., Jackson, S.K. et al 2013. SPE 164018 presented
at the SPE Middle East Unconventional Gas Conference and Exhibition, Muscat, Oman.
Itasca Consulting Group, Inc. FLAC3D Fast Lagrangian Analysis of Continua in Three-Dimension. Minneapolis:
Itasca.
Lee, J., Rollins, J.B., and Spivey, J.P. 2003. Pressure Transient Testing. SPE Textbook Series, Vol. 9, Society of Petroleum
Engineers.
Liptak, B.G. 2006. Instrument Engineer's Handbook, Volume Two: Process Control and Optimization, CRC Press.
Manchanda, R., Roussel, N.P., and Sharma, M.M. 2012. Factors Influencing Fracture Trajectories and Fracturing Pressure
Data in a Horizontal Completion. ARMA 12-633 presented at the 46th US Rock Mechanics/Geomechanics Symposium
held in Chicago, Illinois, USA.
Manchanda, R. and Sharma, M.M. 2013. Time Dependent Fracture Interference Effects in Pad Wells. SPE 164534
presented at the SPE Unconventional Resource Conference, The Woodlands, Texas, USA.
SPE-184865-MS 29

Manchanda, R. and Sharma, M.M. 2014. Impact of Completion Design on Fracture Complexity in Horizontal Shale Wells.
SPE Drilling & Completion 29 (1): 7887. SPE-159899-PA.
Mayerhofer, M.J., Stutz, H.L., Davis, E.J., and Wolhart, S.L. 2006. Optimizing Fracture Stimulation Using Treatment-
Well Tiltmeters and Integrated Fracture Modeling. SPE Prod & Oper 21 (2): 222229. SPE-84490-PA.
McClure, M.W. and Zoback, M.D. 2013. Computational Investigation of Trends in Initial Shut-in Pressure during
Multi-Stage Hydraulic Stimulation in the Barnett Shale. ARMA 13-368 presented at the 47th US Rock Mechanics/
Geomechanics Symposium, San Francisco, California, USA.
Meyer, B.R., Bazan, L.W., Brown, E.K., and Brinzer, B.C. 2013. Key Parameters Affecting Successful Hydraulic Fracture
Design and Optimized Production in Unconventional Wells. SPE 165702 presented at the SPE Eastern Regional
Meeting, Pittsburgh, Pennsylvania, USA.
Mondal, S. 2010. Pressure Transients in Wellbores: Water Hammer Effects and Implications for Fracture Diagnostics.
MS Thesis. The University of Texas at Austin.
Olson, J.E. 2004. Predicting Fracture Swarms the Influence of Subcritical Crack Growth and the Crack-Tip Process
Zone on Joint Spacing in Rock, Geological Society, London, Special Publications, 231: 73-88.
Pollard, D.D. and Segall, P. 1987. Theoretical displacements and stresses near fractures in rock: with applications to faults,
joints, veins, dikes and solution surfaces. In: Atkinson, B.K., Fracture Mechanics of Rock, Academic Press, London,
277-350.
Roussel, N.P., Manchanda, R., and Sharma, M.M. 2012. Implications of Fracturing Pressure Data Recorded during a
Horizontal Completion on Stage Spacing Design. SPE 152631 presented at the SPE Hydraulic Fracturing Technology
Conference, The Woodlands, Texas, USA.
Roussel, N.P. and Sharma, M.M. 2011a. Optimizing Fracture Spacing and Sequencing in Horizontal-Well Fracturing. SPE
Prod & Oper 26 (2): 173-184. SPE-127986-PA.
Roussel, N.P. and Sharma, M.M. 2011b. Strategies to Minimize Frac Spacing and Stimulate Natural Fractures in Horizontal
Completions. SPE 146104 presented at the SPE Annual Technical Conference and Exhibition, Denver, Colorado, USA.
Roussel, N.P., Florez, H., and Rodriguez, A.A. 2013. Hydraulic Fracture Propagation from Infill Horizontal Wells. SPE
166503 presented at the SPE Annual Technical Conference and Exhibition, New Orleans, Louisiana, USA.
Smith, M.B., Reeves, T.L., and Miller, W.K. 1989. Multiple Fracture Height Measurements: A Case History. SPE 19092
presented at the SPE Gas Technology Conference, Dallas, Texas, USA.
Sneddon, I.N. and Elliot, H.A. 1946. The Opening of a Griffith Crack under Internal Pressure. Quarterly of Applied
Mathematics 4 (3): 262267.
Sneddon, I.N. 1946. The Distribution of Stress in the Neighborhood of a Crack in an Elastic Solid. Proc. R. Soc. Lond.
A 187 (1009): 229260.
Soliman, M.Y., East, L., and Adams, D. 2008. Geomechanics Aspects of Multiple Fracturing of Horizontal and Vertical
Wells. SPE Drilling and Completion 23 (3): 217228. SPE 86992-PA.
Vermylen, J. and Zoback, M.D. 2011. Hydraulic Fracturing, Microseismic Magnitudes, and Stress Evolution in the Barnett
Shale, Texas, USA. SPE 140507 presented at the SPE Hydraulic Fracturing Technology Conference, The Woodlands,
Texas, USA.
Warpinki, N.R., Mayerhofer, M.J., Agarwal, K., and Du, J. 2012. Hydraulic Fracture Geomechanics and Microseismic
Source Mechanisms. SPE 158935 presented at the SPE Annual Technical Conference and Exhibition, San Antonio,
Texas, USA.
Warpinki, N.R., Wolhart, S.L., and Wright, S.L. 2004. Analysis and Prediction of Microseismicity Induced by Hydraulic
Fracturing. SPE Journal 9 (1): 24 33. SPE-87673-PA.
Webster, P., Cox, B., and Molenaar, M. 2013. Developments in Diagnostic Tools for Hydraulic Fracture Geometry
Analysis. URTeC 1619968 presented at the Unconventional Resources Technology Conference, Denver, Colorado,
USA.
Weng, X., Kresse, O., Chuprakov, D., Cohen, C.-E., Prioul, R., and Ganguly, U. 2014. Applying Complex Fracture Model
and Integrated Workflow in Unconventional Reservoirs. Journal of Petroleum Science and Engineering 124: 468483.
Wolhart, S., McIntosh, G.E., Zoll, M.B., and Weijers, L. 2007. Surface Tiltmeter Mapping Shows Hydraulic Fracture
Reorientation in the Codell Formation, Wattenberg Field, Colorado. SPE 110034 presented at the Annual Technology
Conference and Exhibition, Anaheim, California, USA.
Wright, C.A., Davis, E.J., Golich, G.M., Ward, J.F., Demetrius, S.L., Minner, W.A. et al 1998. Downhole Tiltmeter Fracture
Mapping: Finally Measuring Hydraulic Fracture Dimensions. SPE 46194 presented at the SPE Western Regional
Conference, Bakersfield, California, USA.
Wu, K, and Olson, J.E. 2013. Investigation of Critical In Situ and Injection Factors in Multi-Frac Treatments: Guidelines
for Controlling Fracture Complexity, SPE 163821 presented at the SPE Hydraulic Fracturing Technology Conference,
The Woodlands, Texas, USA.
Yew, C.H. and Weng, X. 2015. Mechanics of Hydraulic Fracturing, Second Edition, Elsevier.
30 SPE-184865-MS

Zoback, M.D. 2007. Reservoir Geomechanics, First Edition, Cambridge University Press.
Zoback, M.D., Barton, C.A., Brudy, M., Castillo, D.A., Finkbeiner, T., Grollimund, B.R. et al 2003. Determination of
Stress Orientation and Magnitude in Deep Wells. International Journal of Rock Mechanics and Mining Sciences 40:
1049 1076.

Das könnte Ihnen auch gefallen