You are on page 1of 13

SPE-189840-MS

Reinterpretation of Flow Patterns During DFITs Based on Dynamic Fracture


Geometry, Leakoff and Afterflow

Behnam Zanganeh and Christopher R. Clarkson, University of Calgary; Jack R. Jones, NSI Fracturing LLC

Copyright 2018, Society of Petroleum Engineers

This paper was prepared for presentation at the SPE Hydraulic Fracturing Technology Conference & Exhibition held in The Woodlands, Texas, USA, 23-25 January 2018.

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
The goal of this study is to explain the full spectrum of flow patterns observed before and after closure
during diagnostic fracture injection tests (DFITs) by considering the dynamic nature of fracture geometry,
variable leakoff rate and afterflow volume caused by wellbore storage.
A fit-for-purpose simulation model is used to simulate DFITs and generate pressure responses in low-
permeability (tight) reservoirs. The cohesive zone model in Abaqus® is used to simulate hydraulic fracture
propagation and closure. A customized leakoff model incorporated into the software accounts for variable
leakoff rate as a function of reservoir properties, fracture pressure, fracture surface area and exposure time.
The afterflow is modeled by including a wellbore volume and accounting for wellbore storage. Results
are compared to field data to explain the full spectrum of flow patterns and fracture dynamics observed in
pressure transient analysis of DFITs.
The overall falloff period is interpreted, using PTA diagnostic plots, for relative magnitudes of afterflow,
leakoff rate and fluid flow in the formation. Initially, afterflow is high, resulting in fracture expansion,
which is characterized by a unit slope on the Bourdet-derivative plot. The afterflow does not necessarily
end after the unit slope terminates; the end of fracture expansion is signaled by a characteristic hump on
the derivative plot. During fracture expansion, the afterflow decreases and the leakoff rate increases due to
the larger fracture area. When the leakoff rate dominates over afterflow, fracture closure mechanics can be
conceptualized as a moving hinge-closure, where the fracture volume reduces, and fracture tip extension
occurs as fluid is pushed to the tip of fracture (indicated by fluctuations in the derivative). The transition
from afterflow to leakoff dominance and the moving hinge-closure manifest as a semi-horizontal trend on
the Bourdet-derivative. Subsequently, a progressive fracture closure occurs gradually along the fracture,
identified by an increasing trend or a sharp decline on the primary pressure derivative, depending on the
conductivity. Different estimates of closure pressure will be obtained early and late in this process. The
pressure behavior immediately after full closure is observed to be affected by the residual leakoff and the
continuing afterflow. Once all of these fracture, wellbore and leakoff processes are abated, the reservoir
response is observed.
This study provides a clear understanding of the different mechanisms affecting pressure behavior during
DFITs for tight reservoirs in order to arrive at more reliable estimates of fracturing parameters and reservoir
2 SPE-189840-MS

properties. As an example, mechanisms leading to false before- and after-closure radial flow identification
are explained.

Introduction
The DFIT, also known as a minifrac or fracture calibration test, has become the standard pre-stimulation
method for determining fracturing treatment properties, effective reservoir permeability and initial reservoir
pressure in unconventional resources. A DFIT is an injection-falloff test conducted before the main
hydraulic fracturing treatment. The DFIT procedure involves breakdown of the rock and creation of a short
fracture during the injection of high pressure fluid (usually water), and recording of the pressure response for
long enough that the created fracture closes, and the transient associated with this closure event is dissipated
before the onset of the reservoir flow regimes.
DFIT analysis was pioneered by Nolte (1979), who introduced the G-function based on material balance
and the Carter leakoff model (Howard and Fast 1957) for before closure analysis. Gu et al. (1993) and Nolte
et al. (1997) demonstrated the possibility of observing after-closure reservoir linear and pseudo-radial flow,
and their application in estimation of fracture geometry and reservoir properties.
Barree and Mukherjee (1996) and Barree (1998) recommended using the G-function and its diagnostic
derivatives to identify fracture closure. Barree (1998) and Barree et al. (2009) also presented
signatures for non-ideal leakoff behaviors such as tip extension, pressure-dependent leakoff and height
recession/transverse storage based on G-function combination plots.
Mohamed et al. (2011) and Marongiu-Porcu et al. (2011) presented a model to predict the falloff pressure
trend of an idealized DFIT (normal leakoff) using standard pressure transient (PTA) log-log diagnostic
plots. They identified normal leakoff as a 3/2 slope on the Bourdet-derivative (with respect to superposition
time), and picked fracture closure as the deviation from the 3/2 slope. Bachman et al. (2012) presented a
combination of conventional diagnostic PTA plots, to identify various flow regimes before and after fracture
closure. They demonstrated that a 3/2 on Bourdet-derivative corresponds to a straight line trend on G-
function derivatives, representing Carter leakoff (Nolte flow). Liu and Ehlig-Economides (2015) presented
analytical models to represent before-closure non-ideal behaviors. Van Den Hoek (2016) presented a PTA
approach for modeling pressure behavior in DFITs and waterflood-induced fractures based on simplified
numerical and analytical solutions. In the latter work, fracture growth rate was a predefined input into the
simulator, the Carter model was used as the leakoff model for the DFIT, and closure was modeled as a
gradual decline of fracture compliance.
While the current analytical methods in the literature provide insight into certain parameters, their
validity and accuracy are questionable due to fundamental assumptions being violated. A DFIT exhibits
very complex physical behavior, with various mechanisms active at the same time, including those related
to wellbore, fracture, leakoff and reservoir flow. Therefore, it is not surprising that the observed trends
in field data are not predicted using existing analytical methods, and some common signatures cannot be
interpreted. This underscores the need for a systematic simulation study of DFIT responses where all the
active mechanisms are captured simultaneously.
Recently, McClure et al. (2014), McClure et al. (2016) and Jung et al. (2016) conducted rigorous
simulation studies of before-closure DFIT responses and demonstrated the effect of variable fracture
compliance on pressure behavior. These studies are limited to before closure behavior. In a recent study by
the authors (Zanganeh et al. 2017), the main results of McClure, regarding the effect of variable compliance
on DFIT response, were duplicated. Further, progressive fracture closure was demonstrated, along with a
method to identify it on PTA plots for closure identification.
In this study, the full spectrum of falloff data during a DFIT is interpreted, including after-closure, and
the effects of wellbore storage, total leakoff rate, fracture dynamics and residual fracture conductivity on
the DFIT signature are evaluated.
SPE-189840-MS 3

Theory and Methods


A customized 2D plane strain model in a fully coupled stress-pore pressure simulator (Abaqus®) is used
herein to generate synthetic DFIT responses. The customized model is described in detail in Zanganeh at el.
(2017). The model is capable of simulating all the physical processes involved in a typical DFIT including:
porous media deformation; fluid flow inside the reservoir; hydraulic fracture initiation, propagation and
closure (based on the cohesive zone method); compliance change before and after closure; residual fracture
aperture and conductivity; and fluid flow inside the fracture and fluid interaction between the fracture and
reservoir (leakoff).
The generated synthetic results are compared against field data, and necessary modifications are made
to calibrate the model. The observations from the calibrated model are then used to explain field behaviors.

Model Setup and Description


Fig. 1 provides a schematic of the simulation model. Minimum and maximum horizontal stresses are acting
in the Y and X directions, respectively. The cohesive elements are embedded in the formation rock - they
are assumed non-existent in the model until fracture initiation and propagation criteria are reached, at which
time these elements act as the potential pathway for hydraulic fracture growth.

Figure 1—Plan view schematic of the simulation model setup.

The leakoff model used herein is based on the work of Sarvaramini and Garagash (2015):

(1)

where qleak(x, t) is leakoff rate at point x along fracture and time t, S is the storage coefficient, αl is the
hydraulic diffusivity, τ(x) is the time when fracture is created at point x (exposure time) and ΔP(x,t) is
the pressure difference between fracture and formation at point x. Unlike the Carter model (Howard and
Fast 1957), this model is pressure dependent, and it is directly related to formation properties such as
permeability, porosity and total compressibility.
Three different simulation models are presented in this study, each focusing on a certain scenario with
different input settings.Table 1 lists the main parameters that are constant among all three models. The main
differences of the three models are listed in Table 2. Simulation settings are chosen to represent different
possible scenarios observed in field examples.
4 SPE-189840-MS

Table 1—Base Case model simulation properties

Input parameter Value Input parameter Value


Permeability, md 0.025 Initial porosity, % 10
Young's modulus, GPa 20.7 Fracture fluid viscosity, cp 1
Pumping rate (per unit thickness),
Poisson's ratio 0.25 2×10-4
m3.sec-1
Total compressibility, KPa 4.68×10-6 Maximum tensile strength, MPa 1.25
Initial pore pressure, MPa 24.8

Table 2—Differences in model setup and simulation settings

Model 1 Model 2 Model 3


Wellbore storage Coefficient, m /MPa
3 3×10 -3 4×10 -6 4×10-5
Shmin, MPa 38.6 30.2 30.3
Shmax, MPa 44.1 38.3 38.3
Tip extension Significant present Negligible
Conductivity of closed fracture infinite infinite finite

Model 1 has a very high wellbore storage coefficient, requiring longer pumping time to pressurize the
wellbore. Also, cohesive zone properties (i.e. stiffness, initiation criteria and evolution law) are selected
to favor tip extension. In Model 2 and 3, the wellbore storage coefficient is reduced, and an exponential
softening law for the cohesive zone is chosen to reduce the possibility of tip extension and compliance
contrast during closure. In Model 3, advanced formulated cohesive elements, COD2D4P, are used (Abaqus
Analysis User's Guide 2016). This type of element support transitions from Poiseuille flow to Darcy flow
and the other way around, as fracture propagates during pumping or closes during falloff. This enables
modeling of the closure process where the closed portion of the fracture has finite conductivity. Each model
is initialized using the Geostatic Stress State step in Abaqus (Abaqus Analysis User's Guide 2016). This step
is used to verify that the input stress field is in equilibrium with the input properties and applied boundary
conditions, and to iterate, if necessary, to obtain equilibrium. This is the reason why stress values in Table
2 are different for the three models.

Results and Discussion


In the following, synthetic pressure responses are presented for Models 1 to 3 and interpreted for various
physical phenomena. Field cases are then analyzed in the context of the simulated results to understand the
signatures that commonly occur in the field.

Model 1
The main purpose of this model is to demonstrate the effect of variable afterflow and leakoff rates before
and during fracture closure on fracture dynamics and PTA plots. Fig. 2(b) presents the plot of afterflow and
total leakoff rate from the fracture surface during the shut-in period. Based on the trend of total leakoff rate,
and its relative magnitude with respect to afterflow rate, the overall falloff period on PTA plots (Fig. 2(a))
is divided into the following zones:
SPE-189840-MS 5

Figure 2—(a) PTA plots for Model 1; (b) Total leakoff rate and afterflow during falloff

Zone 1: wellbore storage dominance and fracture expansion. While afterflow is decreasing during
this period, its magnitude is much larger than the total leakoff rate. As a result, the fracture continues to
grow in all directions (aperture, length and height in case of a 3D model). A conceptual schematic of this
phenomenon is illustrated in Fig. 3(a). Because the total leakoff is a function of fracture surface area (Eq.
1), enhanced surface area due to fracture expansion increases the total leakoff rate during this period. This
zone follows a unit slope trend on PTA plots (Fig. 2(a)). It must be noted that afterflow does not end at the
end of this unit slope. The characteristic hump at the end of unit slope indicates that fractured expansion
has stopped.
6 SPE-189840-MS

Figure 3—Illustration of (a) fracture expansion during wellbore storage dominance


(Zone 1); (b) fracture tip extension during moving hinge-closure (Zone 2)

Zone 2: transition to leakoff dominance with moving hinge-closure (tip extension). This period starts
when the value of the total leakoff rate exceeds the afterflow rate (Fig. 2(b)). At this point, fracture hinge
closure (reduction in fracture aperture) starts while afterflow is still present; the fracture is conductive and
there is a process zone at the tip of fracture. Therefore, as fluid is pushed toward the tip of fracture, it
can cause additional tip extension/fracture growth (Fig. 3(b)). The abrupt changes and fluctuations on PTA
plots represent this phenomenon (Fig. 2(a)). Again, total leakoff rate increases as fracture surface area is
extended. This zone can last for 1 to 2 log-cycles, and, depending on its duration, it can show up as a semi-
horizontal trend on Bourdet-derivatives accompanied with fluctuations.
Zone 3: leakoff dominance. During this period, the total leakoff rate is larger than the afterflow rate. Also,
it follows a decreasing trend due to reduced fracture pressure and higher shut-in time (pressure and time
terms in Eq.1). The overall trend on Bourdet-derivatives is smooth, but does not necessarily follow a 3/2
slope due to pressure dependency.
Zone 4: progressive fracture closure. This concept is illustrated in detail in Zanganeh et al. (2017). That
study noted that, for planar fractures, closure is a transient process, starting from the tip of the fracture to
the vicinity of the wellbore (perforations).
Fracture retains a residual aperture after mechanical closure. Depending on the value of residual aperture
and matrix permeability, the closed fracture can either have finite or infinite conductivity. In Model 1, the
closed portion of the fracture has infinite conductivity. As a result, progressive fracture closure causes an
upward deviation on the Bourdet-derivative plots and PPD curve (PPD violation). The start of the PPD
violation corresponds to tip closure, and the end of the violation indicates full mechanical closure near the
injection point. In contrast, for the finite conductivity closed fracture case, progressive closure causes a
downward deviation on PTA plots. This will be discussed in detail for Model 3.
Zone 5: residual leakoff with residual afterflow. As mentioned earlier, the closed fracture retains a residual
aperture even after mechanical closure. The fracture pressure is also slightly elevated above initial formation
pressure. Because we have accounted for pressure dependency of leakoff in our simulation model (Eq. 1),
the leakoff mechanism is active as long as fracture pressure is elevated above formation pressure, even
after mechanical closure. Further, as illustrated in Fig. 2(b), the afterflow continues after closure with its
value being comparable to the residual leakoff (same order of magnitude). During this period, a unit slope
trend again appears on the plot. The derivative plot deviates from the unit slope as residual afterflow
becomes negligible.

Model 2
This model is used to investigate long time pressure behavior (after-closure) and the corresponding flow
patterns and time zones developed during falloff. PTA plots and the afterflow/leakoff combination plot for
Model 2 are shown in Fig. 4(a) and Fig. 4(b), respectively.
SPE-189840-MS 7

Figure 4—(a) PTA plots for model 2; (b) Total leakoff rate and afterflow during falloff

As observed in Fig. 4(b), afterflow is much smaller than total leakoff rate from the beginning of the
falloff data, and Zone 1 (as described in Model 1) does not occur with this model. Moving hinge-closure
occurs briefly during Zone 2, causing an abrupt fluctuation on all derivative plots (Fig. 4(a)). Excpept for
the duration of tip extension, Bourdet-derivative plots follow a unit slope trend until Δt = 6 sec = 0.1 tp ,
where they deviate from this trend. The 3/2 slope trend appears at the end of leakoff dominance (Zone 3;
Fig. 4(a)), where Δt > 10 tp. Progressive closure starts at the tip of fracture (Δtc1=0.36 hr) and causes an
upward deviation on all derivative plots (including PPD violation) as the closed part of the fracture remains
infinitely conductive. Progressive closure ends at Δtc2= 0.58 hr. The rest of the falloff period can be divided
into 3 other zones as follows:
Zone 5: residual leakoff without afterflow. As illustrated in Fig. 4(b), residual leakoff continues until Δt =
104 seconds. Compared to Zone 5 with Model 1, there is no residual afterflow in this case. Residual leakoff
without considerable afterflow appears as a semi-horizontal trend (m=0) on curve, and negative unit
8 SPE-189840-MS

slope trend (m=-1) on the plot (Fig. 4(a)). This period may be interpreted incorrectly as formation radial
flow, and therefore incorrectly used to estimate reservoir permeability and pressure.
Zone 6: reservoir flow dominated. During this period, the fracture is finally closed (static) and leakoff/
afterflow are negligible. The falloff period can now be analyzed similarly to buildup/falloff tests, where the
expected sequence is formation linear flow followed by formation radial flow. With Model 2, formation
linear flow is not observed, and the pressure transient approaches radial flow after 106 seconds (Fig. 4(a)).
Table 3 compares the model input reservoir permeability and initial pressure with values estimated using
conventional Horner analysis for both Zone 5 and Zone 6. Falloff analysis of Zone 6 (true formation radial
flow) provides an accurate estimate of reservoir permeability and initial pressure as compared to the input
data. Contrarily, analysis of Zone 5 (residual leakoff; false radial flow) results in a significant overestimate
of permeability and initial pressure.

Table 3—Comparison of input reservoir permeability and initial pore


pressure in Model 2 with estimated values using radial flow (Horner) analysis

Permeability, md Initial pore pressure, MPa


Model input 0.0250 24.800
Zone 5 0.1400 25.840
Zone 6 0.0255 24.812

Zone 7: Reservoir boundary and derivative effects. The pressure transient reaches the boundaries of the
simulation model after 2×106 seconds, causing a downward deviation on all derivative plots (Fig. 4(a)). At
the end of the falloff data, pressure changes become very small with values being beyond the accuracy of
the simulator (or pressure gauges in field data). This causes fluctuations in all derivative calculations.
In some of the field data (e.g. Field Example 2), where pressure is recorded using surface gauges, ambient
temperature changes can significantly affect pressure recordings and derivative calculations once pressure
changes become very small.
Depending on the size of formation/model and reservoir permeability, Zone 7 can occur much earlier.

Model 3
This model demonstrates the effect of closed fracture conductivity on pressure behavior during progressive
closure. As demonstrated earlier with Model 1 and Model 2, when the closed part of fracture remains
infinitely conductive, progressive closure causes a PPD violation (upward deviation on Bourdet-derivative
plots). Model 3 is designed so that, after closure, fracture permeability is of the same order as matrix
permeability (finite conductivity). In this case, progressive closure causes a sharp decline on the PPD curve
(or downward trend on Bourdet-derivative plots).
As illustrated in Fig. 5(a) using PTA plots of Model 3 output, progressive closure starts at the tip of
the fracture at Δt = 120 seconds and causes a sharp decline on PPD curve and downward deviation on the
Bourdet-derivatives. The fracture fully closes near the injection point at Δt = 300 seconds.
In Model 3, after full mechanical closure (Zone 5; Fig. 5(b)), closed fracture has finite conductivity and
both residual leakoff and residual afterflow are present for almost one log-cycle. This scenario appears as
an unusual trend on curve (almost 7/2 slope), and unit slope trend (m=-1) on the plot (Fig. 5(a)).
This period may be interpreted incorrectly as Carter leakoff (Nolte flow).
SPE-189840-MS 9

Figure 5—PTA plots for Model 3 demonstrating progressive closure when the closed
part of fracture has finite conductivity; b) Total leakoff rate and afterflow during falloff

The rest of falloff data is interpreted similarly to Models 1 and 2.

Field Example 1
The analysis for this dataset (conducted in the Montney Formation) is presented in Fig. 6. A small derivative
window of 0.001 is used in derivative calculations. The overall trend and signatures are very similar to
Model 1, except that the progressive closure period is longer in field data because the simulation model is
2D. All of the predicted falloff zones are present in this dataset. Fracture expansion and afterflow dominance
ends after 3.5 minutes (Zone 1). Additional tip extension occurs during Zone 2, with severe fluctuations
occurring in the derivatives (moving hinge-closure). Similar to Model 1, this transition period follows a
semi-horizontal trend. Progressive fracture closure (Zone 4) starts with a PPD violation (or upward deviation
on Bourdet-derivative plots) at 22.7 MPa and ends at 19.5 MPa. The progressive closure period lasts around
5 hours and the difference between closure pressure estimates is significant. Both residual afterflow and
residual leakoff are present after closure (Zone 5). The reservoir response finally appears when the residual
afterflow and leakoff are abated, and formation linear flow appears (m=0.5) at the end of the falloff period.
10 SPE-189840-MS

Figure 6—Interpretation of Field Example 1 based on afterflow, leakoff and fracture dynamics

Field Example 2
This DFIT is conducted in the Montney Formation with pressure values recorded using surface gauges. The
interpretation for this dataset is provided in Fig. 7. The derivative window used for smoothing is 0.001.
The overall trend and signatures are very similar to Model 2. Zones 1 to 6 are clearly present during the
falloff period. The early unit slope in Zone 1 indicates fracture expansion and afterflow dominance. All
derivative plots exhibit significant fluctuations in Zone 2, indicating moving hinge closure/tip extension.
The 3/2 slope is not clear during the leakoff dominance period. Progressive closure starts at the PPD violation
(Pc1=17.1MPa and Δt=0.95 hr) and ends at 15.3 MPa (Δt=2.1 hr).

Figure 7—Interpretation of Field Example 2 based on afterflow, leakoff and fracture dynamics
SPE-189840-MS 11

After closure residual leakoff (without afterflow) is clearly observed in this data set, demonstrated by
a semi-horizontal trend on or the negative unit slope on curve. As mentioned earlier, this can be
interpreted incorrectly as formation radial flow, which in turn would result in an overestimate permeability.
The reservoir flow dominated period (Zone 6) starts after 100 hours, with formation linear flow (m=0.5)
developing at the end of the falloff data. The late time pressure data are significantly affected by ambient
temperature changes during the day and night resulting in severe fluctuations on derivative plots.

Field Example 3
The interpretation for this dataset is provided in Fig. 8. The PTA diagnostic plots are modified from KAPPA
DDA Book (Houzé et al. 2017, p.563). The overall trend and signatures are very similar to Model 3
demonstrating the effect of closed fracture conductivity on pressure response.

Figure 8—Interpretation of Field Example 3 based on conductivity of closed fracture (modified from Houzé et al. 2017)

The PPD curve is not shown for this data set. However, it is expected to follow a similar trend as
Model 3 (Fig. 5). The fracture expansion and afterflow dominance ends shortly after shut-in (Zone 1). The
transition to leakoff dominance is smooth and no tip extension occurs in Zone 2. The 3/2 slope trend is not
observed during the leakoff dominance period (Zone3). Progressive closure starts at at Δt = 0.6 hr. Since
the closed portion of fracture has finite conductivity, it causes a sharp decline and downward deviation on
the derivative plots (Zone 4). The behavior after fracture closure (Zone 5) is affected by residual leakoff
and finite conductivity of the closed fracture.

Conclusions
In this work, for the first time, the full spectrum of flow patterns and signatures observed before and after
closure during DFITs are explained by considering the dynamic nature of fracture geometry, variable leakoff
rate and afterflow. The overall falloff period is divided into 7 zones that can be summarized as follows:

• Zone 1: Fracture expansion occurs due to wellbore storage dominance. Total leakoff rate follows
an increasing trend.
12 SPE-189840-MS

• Zone 2: Early transition from wellbore storage to leakoff dominance can be conceptualized as
a moving hinge-closure causing tip extension. Again, total leakoff rate follows an increasing
trend. This zone may result in a semi-horizontal trend on Bourdet-derivatives (false before-closure
signature of radial flow)
• Zone 3: Total leakoff rate starts to decrease as fracture pressure is reduced and shut-in time
increases.
• Zone 4: Mechanical fracture closure starts at the tip of fracture, moving towards the vicinity of the
injection point (progressive closure). Depending on the conductivity of the closed part of fracture,
this causes an upward deviation (PPD violation) or downward deviation (sharp decline in PPD)
on Bourdet-derivatives.
• Zone 5: It is possible for leakoff and afterflow to continue after closure (residual afterflow and
leakoff). Depending on the conductivity of the closed fracture, this period can be misinterpreted as
Carter leakoff or reservoir radial flow causing an overestimate of permeability and initial pressure
(false after-closure signature of radial flow).
• Zone 6: Reservoir dominated behavior is observed when the fracture is static, and leakoff and
afterflow are negligible compared to flow inside the reservoir.
• Zone 7: Boundary dominated effects may be observed if the boundaries of the reservoir are reached.
Also, as pressure changes become very small, derivative calculations are affected.

Acknowledgements
The authors would like to thank Dassault Systemes Simulia Corporation for providing Abaqus®, and Seven
Generations Energy Corporation and Robert Hawkes for providing field data. The sponsors of Tight Oil
Consortium at the University of Calgary are acknowledged for their support.

Nomenclature

Field Variables
G G-function time, dimensionless
m Slope of straight line, dimensionless
P Pressure, MPa
Pc1 Closure pressure at the start of progressive closure, MPa
Pc2 Closure pressure at the end of progressive closure, MPa
qleak Leakoff rate, m3/sec
S Storage coefficient, Pa-1
Shmax Maximum horizontal stress, Pa
Shmin Minimum horizontal stress, Pa
ta Agarwal's time, dimensionless
tp Pumping time, sec

Greek Variables
αl Hydraulic diffusivity, m2/sec
Δt Shut-in time, sec
τ(x) Exposure time, sec
μ Fracturing fluid viscosity, cp

References
Abaqus Analysis User's Guide 2016. Dassault Systemes Simulia Corp., Providence, RI.
SPE-189840-MS 13

Agarwal, R.G. 1980. A New Method to Account for Producing Time Effects when Drawdown Type Curves are Used to
Analyze Pressure Buildup and Other Test Data. Presented at the SPE Annual Technical Conference and Exhibition,
Dallas, Texas, 21-24 September. SPE-9289-MS. http://dx.doi.org/10.2118/9289-MS
Bachman, R.C., Walters, D.A., Hawkes, R.A. et al. 2012. Reappraisal of the G Time Concept in Mini-Frac Analysis.
Presented at the SPE Annual Technical Conference and Exhibition, San Antonio, Texas, 8-10 October. SPE-160169-
MS. http://dx.doi.org/10.2118/160169-MS
Barree, R. D. and Mukherjee, H. 1996. Determination of Pressure Dependent Leakoff and Its Effect on Fracture Geometry.
Presented at the SPE Annual Technical Conference and Exhibition, Denver, Colorado, 6-9 October. SPE-36424-MS.
http://dx.doi.org/10.2118/36424-MS.
Barree, R. D. 1998. Applications of Pre-Frac Injection/Falloff Tests in Fissured Reservoirs-Field Examples. Presented
at the SPE Rocky Mountain Regional/Low Permeability Reservoirs Symposium, Denver, Colorado, 5-8 April.
SPE-39932-MS. http://dx.doi.org/10.2118/39932-MS.
Barree, R.D., Baree, V.L., and Craig, D.P. 2009. Holistic Fracture Diagnostics: Consistent Interpretation of Prefrac
Injection Tests Using Multiple Analysis Methods. SPE Production & Operations 24(3): 396-406. SPE-107877-PA.
http://dx.doi.org/10.2118/107877-PA
Gu, H., Elbel, J.L., Nolte, K.G. et al. 1993. Formation Permeability Determination Using Impulse-Fracture Injection.
Presented at the SPE Production Operations Symposium, Oklahoma City, Oklahoma, 21-23 March. SPE-25425-MS.
https://doi.org/10.2118/25425-MS
Howard, C.C., and Fast, C.R. 1957. Optimum Fluid Characteristics for Fracture Extension. API Drilling and Production
Practice 24(1): 261-270. API-57-261.
Houzé, O., Viturat, D., Fjaere, O.S. et al. 2017. Dynamic Data Analysis, V5.12.01, p. 563. https://www.kappaeng.com/
papers
Jung, H., Sharma, M.M., Cramer, D.D et al. 2016. Re-examining interpretations of non-ideal behavior during
diagnostic fracture injection tests. Petroleum Science and Engineering 145(1): 114-136. https://doi.org/10.1016/
j.petrol.2016.03.016
Marongiu-Porcu, M., Ehlig-Economides, C.A. and Economides, M.J. 2011. Global Model for Fracture Falloff Analysis.
Presented at the SPE North American Unconventional Gas Conference and Exhibition, The Woodlands, Texas, 14-16
June. SPE-144028-MS. http://dx.doi.org/10.2118/144028-MS
Nolte, K.G. 1979. Determination of Fracture Parameters From Fracturing Pressure Decline. Presented at the SPE
Annual Technical Conference and Exhibition, Las Vegas, Nevada, 23-26 September. SPE-8341-MS. http://
dx.doi.org/10.2118/8341-MS
Nolte, K.G., Maniere, J.L., and Owens, K.A. 1997. After-Closure Analysis of Fracture Calibration Tests. Presented at
the SPE Annual Technical Conference and Exhibition, San Antonio, Texas, 5-8 October. SPE-38676-MS. https://
doi.org/10.2118/38676-MS
McClure, M.W., Blyton, C.A.J., Jung, H. et al. 2014. The Effect of Changing Fracture Compliance on Pressure Transient
Behavior During Diagnostic Fracture Injection Tests. Presented at the SPE Annual Technical Conference and
Exhibition, Amsterdam, The Netherlands, 27–29 October. SPE-170956-MS. http://dx.doi.org/10.2118/170956-MS
McClure, M.W., Jung, H., Cramer, D.D. et al. 2016. The Fracture-Compliance Method for Picking Closure
Pressure From Diagnostic Fracture-Injection Tests. Soc. Pet. Eng. 21(4): 1321-1339. SPE-179725-PA. http://
dx.doi.org/10.2118/179725-PA
Mohamed, I.M., Nasralla, R.A., Sayad, M.A. et al. 2011. Evaluation of After-closure Analysis Techniques for Tight
and Shale Gas Formations. Presented at the SPE Hydraulic Fracturing Technology Conference and Exhibition, The
Woodlands, Texas, 24-26 January. SPE-140136-MS. http://dx.doi.org/10.2118/140136-MS
Liu, G., and Ehlig-Economides, C.A. 2015. Comprehensive Global Model for Before-Closure Analysis of an Injection
Falloff Fracture Calibration Test. Presented at the SPE Annual Technical Conference and Exhibition, Houston, Texas,
28-30 September. SPE-174906-MS. https://doi.org/10.2118/174906-MS
Sarvaramini, E., and Garagash, D.I. 2015. Breakdown of a Pressurized Fingerlike Crack in a Permeable Solid. J. Appl.
Mech 82(6). http://dx.doi.org/10.1115/1.4030172
Van Den Hoek, P. 2016. A Simple Unified Pressure Transient Analysis Method for Fractured Waterflood Injectors and
Minifracs in Hydraulic Fracture Stimulation. Presented at the SPE Annual Technical Conference and Exhibition,
Dubai, UAE, 26-28 September. SPE-181593-MS. http://dx.doi.org/10.2118/181593-MS
Zanganeh, B., Clarkson C.R., and Hawkes, R.V. 2017. Reinterpretation of Fracture Closure Dynamics During Diagnostic
Fracture Injection Tests. Presented at the SPE Western Regional Meeting, Bakersfield, California, 23-27 April.
SPE-185649-MS. https://doi.org/10.2118/185649-MS

You might also like