You are on page 1of 63

Fracture Evaluation Using


Pressure Diagnostics
Sunil N. Gulrajani and K. G. Nolte, Schlumberger Dowell

9-1. Introduction response after fracturing characterizes its hydrocar-


bon production potential. Finally, the versatility and
The hydraulic fracture design process requires sub- ease of use of these techniques lend their application
stantial information about the reservoir and the frac- to most field situations, enabling an assessment of
turing fluid. The effectiveness of a design is also the fracturing process, either on site and in real time
dependent on the quality of the required data. The or after completion of the treatment for the improve-
fracture design parameters can be inferred from wire- ment of future designs.
line logs and through laboratory fluid and core testing Figure 9-1 shows a recording of the bottomhole
procedures. The reliability of the data inferred from pressure measured during a fracture stimulation.
these methods is reduced by factors such as the scale Fluid and proppant were injected for approximately
of measurement, variability in the geologic environ- 3 hr, followed by an extended period of shut-in that
ment, assumption of an overly simplified fracture lasted for more than 18 hr. The pressure response
response and significant deviations of the test envi- was recorded for all facets of the fracture and reser-
ronment from in-situ reservoir conditions. Fracture voir response—the pumping period, pressure decline
parameter uncertainties can result in a suboptimal as the fracture closes, time at which the fracture
fracture, at best, or in a complete failure of the stimu- closes on the proppant and finally the asymptotic
lation procedure under the worst-case scenario. approach to the reservoir pressure. Of particular
An on-site procedure for predicting fracture interest to the fracture design process is the analysis
dimensions is desirable but extremely difficult to of pressure measured during a calibration treatment.
obtain. Numerous fracture-mapping techniques, such A calibration treatment, or a mini-fracture treatment,
as radioactive tracers, surface and bottomhole tilt- precedes the main fracture treatment and is per-
meters and various electromagnetic measurements, formed without the addition of proppant under con-
have been applied to infer fracture dimensions (see ditions that mimic the main treatment. The pressure
Section 12-1.1). The techniques, however, provide response for this test follows the sequence in Fig. 9-1.
only a limited amount of information (e.g., fracture Fracturing pressures during each stage of fracture
azimuth or wellbore height) that is available gener- evolution (i.e., growth, closing phase and after-
ally only after completion of the fracture treatment. closure period) provide complementary information
Although sophisticated microseismic measurements pertinent to the fracture design process. The frame-
have been developed to infer the created fracture
dimensions, their restricted spatial range of observa-
Bottomhole pressure, pw (psi)

Injection Shut-in
tion and expensive instrumentation currently limit 9000
Fracture Transient reservoir
widespread application. closing pressure near the wellbore

In contrast, pressure analysis during and after a 8000


Net fracture Fracture closes on
fracture treatment is recognized as a powerful tech- pressure proppant at well
7000 pw – pc
nique for developing a comprehensive understanding
Closure pressure
of the fracturing process. A recording of the well- 6000 pc = horizontal rock stress
bore pressures provides an inexpensive measurement Reservoir pressure

for fracture diagnostics. The specialized analysis of 5000


38 40 42 44 46 48 50 56 58
pressure provides a qualitative indicator of the frac-
Time (hr)
ture growth, as well as estimates of the primary frac-
ture parameters. An analysis of the reservoir Figure 9-1. Bottomhole fracture pressure history (Nolte,
1982, 1988c).

Reservoir Stimulation 9-1


work for analyzing fracturing pressure and its variation pressure response during this period loses its depen-
during a calibration test, along with example applica- dency on the mechanical response of an open fracture
tions, is the focus of this chapter. and is governed by the transient pressure response
Pressure analysis is based on the simultaneous con- within the reservoir. This transient results from fluid
sideration of three principles that are central to the loss during fracturing and can exhibit either linear
fracturing process: material balance, fracturing fluid flow or a long-term radial response. Each of these
flow and solid mechanics or the resulting rock defor- flow patterns can be addressed in a manner analogous
mation. Section 9-3 describes these principles. to conventional well test analysis for a fixed-length
Material balance, or the conservation of mass, is conductive fracture, as discussed in Chapter 12. The
important for analyzing the pumping and closing after-closure period characterizes the reservoir’s pro-
phases of fracturing. It also provides a framework for duction potential. Its analysis enhances the objectivity
determining the fluid volumes and proppant schedule of otherwise uncertain preclosure pressure interpreta-
for a basic fracture design. Fluid flow and solid tion. The theoretical, operational and application
mechanics define the interaction between the fluid aspects for describing the pressure following fracture
pressure and the formation. The fracture width and closure are outlined in Section 9-6.
pressure relations are derived by combining these two Each fracturing phase, from fracture creation
principles. In addition, solid mechanics also intro- through the after-closure period, provides a sequence
duces the concept of closure pressure, which is the of complementary information pertinent to the frac-
reference pressure for fracture behavior. Closure pres- turing process. A comprehensive assessment thus
sure is the most important parameter for fracturing requires an integrated procedure that combines the
pressure evaluation. information provided by each phase. Sections 9-7 and
The pressure measured during pumping provides 9-8 present a generalized pressure analysis method-
an indication of the fracture growth process. The pri- ology that unifies the various interpretative, analytical
mary diagnostic tool for this period is the slope of and numerical analysis techniques discussed in this
the log-log plot of net pressure (i.e., the fracturing chapter.
pressure above the reference closure pressure) versus
pumping time. The slope of the log-log plot is used
to characterize the fracture geometry by combining 9-2. Background
the fundamental relations, as outlined in Section 9-4.
The log-log plot is complemented by the pressure Injection pressure has been measured for safety con-
derivative to identify complex fracture growth pat- siderations since the inception of hydraulic fracturing.
terns and the effects of proppant injection. The pump- Its importance for characterizing a fracture was rec-
ing phase additionally characterizes the formation ognized as early as 1954 by Harrison et al., who
pressure capacity, which is the fracture pressure above developed a relation between pressure and the frac-
which only limited fracture propagation occurs. ture volume. The potential importance of understand-
The pressure response during fracture closure is ing the fracturing pressure response was noted by
governed largely by the rate of fluid loss. The analysis Godbey and Hodges (1958). They concluded, “by
of pressure during this period estimates the fluid effi- obtaining the actual pressure on the formation during
ciency and the leakoff coefficient. These parameters a fracture treatment, and if the inherent tectonic
are determined from a plot of the pressure decline ver- stresses are known, it should be possible to determine
sus a specialized function of time, commonly referred the type of fracture induced.” Furthermore, they
to as the G-plot. This specialized plot provides the stated, “the observation of both wellhead and bottom-
fracturing analog to the Horner plot (see Section 2-1) hole pressures is necessary to a complete understand-
for well testing. Theoretical principles and example ing and possible improvement of this process.”
applications of G-function analysis are outlined in Development of two-dimensional (2D) fracture
Section 9-5. This section also describes simple analyt- models by Khristianovich and Zheltov (1955),
ical corrections that extend the basic pressure decline Perkins and Kern (1961) and Geertsma and de Klerk
analysis to nonideal fracture behavior. (1969) provided a theoretical means for estimating
The final fracturing pressure analysis pertains to the the fracture width and its dependence on the net pres-
evaluation of pressure following fracture closure. The sure. The early 1970s were characterized by regulated

9-2 Fracture Evaluation Using Pressure Diagnostics


gas prices in the United States and significant oil dis- Schlottman et al., 1981; Elbel et al., 1984; Morris and
coveries in the Middle East, which caused a lull in Sinclair, 1984; Cleary et al., 1993) and fracturing for
fracturing activity and hence a reduced interest in specialized applications (e.g., Smith, 1985; Bale et al.,
fracturing research. Renewed activity was fostered 1992). It is most effective when used to characterize
during the mid-1970s when massive hydraulic frac- fracture growth behavior during the early stages of
turing led to an emphasis on improved fracturing eco- field development. Variations of the traditional cali-
nomics. The fluid and proppant volumes utilized for bration test have also been proposed for specialized
these fracturing operations significantly increased, purposes, such as the step rate/flowback test to deter-
which increased execution costs. Successful fracture mine closure pressure (e.g., Felsenthal, 1974; Nolte,
execution was recognized as critical for economic 1982; Singh et al., 1985; Plahn et al., 1997), step-
field development. These developments spurred a down test to identify near-wellbore effects (e.g.,
renewed interest in understanding fracture pressure Cleary et al., 1993) and short impulse injection test
behavior for the effective design and analysis of frac- to obtain reservoir permeability (e.g., Gu et al., 1993;
ture treatments. The Appendix to Chapter 5 provides Abousleiman et al., 1994). Additional studies
a detailed discussion of these activities. (Mayerhofer et al., 1993; Nolte et al., 1997) extend
A numerical simulation describing the pressure fracture pressure analysis to the realm of well testing,
decline response during fracture closure was pre- wherein reservoir information typically obtained from
sented by Novotny (1977). Nolte (1979) presented a conventional well tests can be inferred from calibra-
fundamental analysis to estimate the fluid-loss para- tion treatments.
meters and the fracture length for the PKN fracture
geometry. This analysis was subsequently extended
to the other 2D fracture geometry models (Nolte, 9-3. Fundamental principles
1986a). Simonson et al. (1978) characterized height
growth into stress barriers. Nolte and Smith (1981)
of hydraulic fracturing
related trends on the log-log net pressure plot to the Three basic relations govern the hydraulic fracturing
evolution of the fracture geometry, and Clifton and process: fluid flow in the fracture, material balance or
Abou-Sayed (1981) generalized fracture modeling conservation of mass, and rock elastic deformation.
and the related pressure response to three-dimensional These relations are reviewed in Chapter 5, presented
(3D) hydraulic fracture models. in the context of fracture modeling in Chapter 6 and
Later developments attempted to address the non- reformulated in this section to facilitate the develop-
ideal conditions that commonly occur during field ment of pressure analysis techniques.
practice but had been excluded in these early studies.
Soliman (1986a) developed an analysis to consider
the effects caused by fluid compressibility and tem- 9-3.1. Fluid flow in the fracture
perature change during shut-in. A correction to incor- The fracture essentially is a channel of varying width
porate pressure-dependent fluid-loss behavior was over its length and height. The local pressure gradient
proposed by Castillo (1987). Nolte (1991) addressed within the fracture is determined by the fracturing
several nonideal conditions during injection as well fluid rheology, fluid velocity and fracture width.
as shut-in and provided techniques to diagnose the Equations governing fluid flow within the fracture
related pressure responses. A semianalytical approach can be derived using the principle of conservation of
to account for variation of the fluid-leakoff coefficient momentum and lubrication theory applied to a fluid
following the end of treatment as well as the effects traveling in a narrow conduit. The rheology of frac-
resulting from fluid flowback was developed by turing fluids is generally represented by a power law
Meyer (1986a). A decline analysis methodology that model (see Chapter 8) that incorporates the parame-
addresses a comprehensive list of nonideal factors ters K and n. In recognition that fluid flow within a
during fracture closure was proposed by Nolte et al. fracture is laminar for most fracturing applications
(1993). (Perkins and Kern, 1961), the global pressure gradi-
Pressure analysis has been applied since the early ent along the length of a fracture can be expressed as
1980s to improve fracture performance in a variety
dp Kvxn
of applications, including routine fracture design (e.g., ∝ , (9-1)
dx w 1+n

Reservoir Stimulation 9-3


where vx is the average fluid velocity along the length conservation, or volume balance, to replace those
of the fracture and is defined in terms of the volumet- of mass conservation.
ric injection rate qi, fracture height hf and height-aver- Pressure analysis, irrespective of the propagation
aged fracture width w – . Material balance or conserva- model, is based on three expressions of material or
tion of mass suggests that vx is proportional to qi/w– h. volume balance. The first defines the treatment effi-
f
Equation 9-1 then becomes ciency η as the ratio of the volume of the fracture
n created at the end of pumping Vfp and the cumulative
dp K  qi 
∝ 1+2 n h  . (9-2) injected volume Vi:
dx w  f Vfp
η= . (9-4)
In the special case of a Newtonian fluid (n = 1 and Vi
K = µ, where µ is the fracturing fluid viscosity), The second expression states that at the end of
Eq. 9-2 reduces to pumping, Vi is equal to Vfp plus the cumulative vol-
dp µ  q  ume of fluid lost to the formation during pumping
∝ 2  i , (9-3) VLp:
dx w  wh f 
– h . is readily recognized as the aver- Vi = Vfp + VLp . (9-5)
where the term w f
age fracture cross-sectional area. Equation 9-3 is It follows from Eqs. 9-4 and 9-5 that
essentially Darcy’s law with the permeability propor-
– 2. VLp = (1 − η)Vi . (9-6)
tional to w
Equations 9-1 and 9-2 are formulated in terms of Finally, during any shut-in period ∆t, the volume of
the average velocity and implicitly ignore change in the fracture is
the fracture width over its height. The varying width Vf ( ∆t ) = Vfp − VLs ( ∆t ) , (9-7)
profile has an effect on the flow resistance relative to
the case of a constant-width channel, as discussed in where VLs(∆t) is the volume of fluid lost to the forma-
Chapter 6. The increase in the flow resistance is tion between the shut-in time and any time ∆t there-
accentuated during periods of fracture height growth after.
into barriers at higher stress. The varying width pro- At closure (i.e., ∆t = ∆tc) the volume of the fracture
file affects other physical phenomena that are highly is equal to the bulk volume of proppant Vprop that was
sensitive to the velocity (e.g., temperature profile and injected during pumping. At closure, Eq. 9-7
proppant distribution). becomes
Vfp = VLs ( ∆tc ) + Vprop . (9-8)
9-3.2. Material balance or Eliminating Vfp from Eqs. 9-5 and 9-8 gives
conservation of mass Vi − Vprop = VLp + VLs ( ∆tc ) . (9-9)
Fluid compressibility is neglected in this chapter for These material-balance relations are illustrated in
the purposes of simplicity and clarity. For water- and Fig. 9-2. Equation 9-9 simply states that for a cali-
oil-base fracturing fluids, fluid volume changes are of bration treatment in which no proppant is added
secondary importance to the elastic deformation of (i.e., Vprop = 0), all injected volume is lost at closure.
the fracture. The physical effects of pressure and tem- “Mathematical relations for fluid loss” in the
perature changes on the fracturing fluid in the well- Appendix to this chapter provides expressions for
bore could be significant for foamed fracturing fluids. VLp and VLs(∆t) that are based on the derivations pre-
In this case, using a direct measurement of the bot- sented by Nolte (1979, 1986a).
tomhole pressure and incorporating changes in the
wellbore volume during shut-in significantly reduce
errors that may be introduced by the assumption of 9-3.3. Rock elastic deformation
an incompressible fluid. The generally applicable
assumption of an incompressible fracturing fluid The principles of fluid flow and material balance are
therefore enables using simple expressions of volume coupled using the relation between the fracture width
and fluid pressure. The relation defines the fracture

9-4 Fracture Evaluation Using Pressure Diagnostics


The first relation predicts the width for a planar
Pumping, tp Closure, ∆t 2D crack (Sneddon and Elliot, 1946), with one
dimension infinite and the other dimension with a
Lost finite extent d. The second relation provides a similar
VLp
expression for a radial, or circular (also called penny-
Stored VLs(∆t) VLs(∆tc)
shaped), crack in an infinite elastic body (Sneddon,
Vfp
Vf(∆t) Vprop
1946). In both cases, the fracture width has an ellipti-
cal shape. The maximum width is proportional to the
product of the characteristic dimension (d for 2D
End of Pumping Vi = Vfp + VLp
cracks and R for radial cracks) and the net pressure
( p–f – σmin) and inversely proportional to the plane
Lost
VLp → CL, κ strain modulus E′ = E/(1 – ν2). The formation
Volume in
VLp
Poisson’s ratio ν and Young’s modulus E are typi-
Vi = qitp cally estimated from laboratory experiments using
Stored
Vfp → w, hf, L cored samples of the reservoir rock (see Chapter 3)
Vfp
or sonic logs (see Chapter 4), and E′ is preferably cal-
ibrated from pressure data (see Section 9-7.2). The
Figure 9-2. Fracture volume-balance relations. average width w – and maximum width w for the 2D
max
crack (i.e., fracture) are, respectively,
compliance. Linear elastic deformation of the reser-
voir rock is assumed for the fracturing process. The w=
(
π p f − σ min d ) (9-10)
linear elasticity assumption is justified because field- 2E′
scale fractures produce relatively small additional 4
stresses superimposed on the much larger in-situ wmax = w (9-11)
π
stresses, excluding possibly the more complex defor-
mation occurring in the fracture tip region (see and for a radial fracture:
Section 6-7.2). The rock deformation, or fracture
width, can be predicted using two classic relations for w=
(
16 p f − σ min R ) (9-12)
cracks in an elastic material of infinite extent that are 3πE ′
subjected to a constant internal pressure p–f, with an
external far-field confining stress σmin applied perpen-
dicular to the plane of the crack, as shown in Fig. 9-3.

2D Crack Plane View of


Radial Crack

wmax d
σmin wmax

σmin
pf

2(pf – σmin)d 8(pf – σmin)R


wmax = wmax =
E´ π E´
π 2
w= w w= w
4 max 3 max

Figure 9-3. Sneddon cracks for 2D and radial fractures.

Reservoir Stimulation 9-5


3
wmax = w. (9-13) 9A. What is closure pressure?
2
The fracture closure pressure pc is defined as the fluid pres-
These relations indicate that the fracture has a sure at which an existing fracture globally closes. Mathe-
width greater than zero only if p–f > σmin. The fluid matically, for a linear relation between the fracture width and
pressure at which an idealized unpropped fracture pressure (i.e., Eq. 9-21), pc equals σmin, the minimum princi-
pal in-situ stress in the reservoir. Ideally, the value of σmin is
effectively closes is globally invariant in homogeneous formations. Reservoirs,
however, are commonly characterized by lithology variations
pc = σ min (9-14) and natural fissures. These cause σmin to become a local,
directional quantity. In this case, the choice of pc depends
and is termed the fracture closure pressure. For com- on the scale and orientation of the representative fracture
geometry.
mercial fracturing applications, pc is distinguished Closure pressure thus is a fracture-geometry-dependent
from σmin, which is a local, directional quantity. The quantity. For example, a micro-hydraulic fracture treatment
closure pressure approximates the average stress over creates a fracture with a limited height (≈5 ft) and hence pro-
vides an estimate of σmin only at that scale. Fracture propa-
the scale and orientation of the initial fracture height. gation over this limited dimension is described by a value of
It represents the stress that governs the propagation of pc that equals the measured σmin. A large-scale, “commercial”
fracture that initiates from the same small interval, however,
a fracture over this scale of interest, as discussed in will grow well beyond the limited height dimension. During
Sidebar 9A. Field practices for estimating pc are this process, it certainly will cross various heterogeneities
prior to establishing coverage over the primary or gross inter-
described in “Estimating closure pressure” in the val, which is defined by meaningful stress barrier layers. Its
Appendix to this chapter. closure pressure is not represented by the measured σmin but
by a value averaged over the gross interval. For large-scale
• Basic fracture geometry models fracturing, pc is equal to σmin only in the unlikely event that
the gross reservoir height is devoid of any variations in the
The elastic relation for the 2D fracture (Eq. 9-10) magnitude or direction of the minimum stress.
is used in two fundamentally different ways to Direct measurement of the fracture width cannot be
achieved during routine field fracturing operations. The infer-
model a fracture. Using a more general form of ence of pc based on the pressure-width relation outlined in
this relation, the characteristic dimension d was Fig. 9A-1 is limited in field practice. An indirect approach that
estimates pc is based on formation tests that create fractures
assumed to be the total fracture length 2L by over the scale of interest. Such large-scale fractures are com-
Khristianovich and Zheltov (1955) and Geertsma monly achieved by the high injection rates used during step
rate and calibration tests. However, conventional shut-in diag-
and de Klerk (1969). The latter study also included nostic plots based on these tests (see “Estimating closure
the effect of fluid loss. This model of a fracture, pressure” in the Appendix to this chapter) may contain multi-
ple inflection points caused by fracture length change and
denoted the KGD model, implicitly assumes that height recession across the reservoir layers. Pressure inflec-
the fracture height is relatively large compared tions are also introduced by the after-closure reservoir
with its length. The other way to use this relation response, as discussed in Section 9-6. These additional
physical phenomena introduce uncertainty in identifying pc
is to assume that the characteristic dimension d is
the fracture height hf, as assumed by Perkins and
Kern (1961) and Nordgren (1972). Their approach,
denoted the PKN model, implicitly assumes that
the fracture length is the infinite dimension.
Each of these two basic, idealized models
assumes that one of the fracture dimensions, the
fracture length or its height, is relatively large in
Width

comparison with the other. Elastic coupling is gen-


erally ignored along the direction of the larger
dimension. Which 2D model is pertinent depends
on which physical dimension of the fracture more
closely replicates the assumptions of the corre-
sponding 2D fracture, as shown on Fig. 9-3. As σmin pc
indicated by Perkins (1973) and Geertsma and
Haafkens (1979), the KGD model is more appro- Pressure
priate when the fracture length is smaller than the
height, whereas the PKN model is more appropri- Figure 9A-1. Mathematical definition of closure
ate when the fracture length is much larger than pressure.

9-6 Fracture Evaluation Using Pressure Diagnostics


9A. What is closure pressure? (continued)

from a calibration treatment. In contrast, the step rate and


flowback tests discussed in “Estimating closure pressure” in 3800
the Appendix avoid the interpretation uncertainty that is intro-
duced by such phenomena. They provide a more objective

Bottomhole pressure (psi)


diagnostic procedure and should be the preferred field tech- 3700
nique for estimating pc.
The relationship between pc and σmin is best illustrated with
a field example adapted from the M-Site fracture experiments 3600
(Branagan et al., 1996). The reservoir is characterized by a
thin, clean layer (≈6 ft) near the top of a larger sandstone inter-
val (≈30 ft), as shown on the well log plotted in Fig. 9A-2. From 3500
a micro-fracture test conducted within this small interval, σmin
3440 psi
was potentially obtained for the thin layer as 3030 psi (Bran-
agan et al., 1996). A step rate test created a fracture that, 3400
because of the higher injection rates, quickly grew within the 0 1 2 3 4
entire 30-ft interval and indicated that pc was 3440 psi Rate (bbl/min)
(Fig. 9A-3). The calibration injection used downhole inclino-
meters (Branagan et al., 1996) to quantify the fracture width,
and its normalized measurement plotted as a function of the Figure 9A-3. Step rate test analysis.
bottomhole pressure on Fig. 9A-4 shows a trend similar to
that in Fig. 9A-1. Fracture opening is seen to occur at about
3030 psi, which equals approximately the measured σmin.
The linear portion of the curve, however, has an intercept of 1.0
3470 psi, which is near the value suggested by the step rate
test. A pc value of 3030 psi characterizes the micro-fracture
test, where the fracture is limited to the thin sandstone interval.
A pc value of 3470 psi is more appropriate for analyzing the
calibration and proppant injection treatments. 0.8
Normalized tiltmeter response

4520
0.6
4530

4540 Sandstone stringer


Gross
+

interval
4550 Micro-fracture 0.4
test
Depth (ft)

4560
Perforations
4570
0.2
4580

4590 3470 psi

4600 0
0 50 100 150 200 250 3000 3200 3400 3600
Gamma ray (API units) Bottomhole pressure (psi)

Figure 9A-2. Gamma ray log showing the micro- Figure 9A-4. Normalized inclinometer response versus
fracture test location and perforated interval. bottomhole pressure.

the height. Consequently, the 2D model is valid in as is the case for a horizontal fracture in a vertical
cases where the fracture length is either relatively wellbore, or as an intermediate condition between
small or large in comparison with the height. In the two limiting cases of the 2D models.
practice, these models are applicable when the • Correction for fluid pressure gradient
dimensions differ by a factor of about 3 or more.
The fundamental elastic relations (Eqs. 9-10
The radial model is most appropriate when the
through 9-13) assume that the pressure in the frac-
total length 2L (2R in Fig. 9-3) is approximately
ture is constant. The fluid flow relation (Eq. 9-2),
equal to the height. This condition occurs for frac-
however, indicates that a pressure gradient exists
ture propagation from a point source of injection,

Reservoir Stimulation 9-7


within a fracture. The fluid pressure varies from its horizontal well, large pressure gradients may occur
maximum value pw at the wellbore to the forma- near the wellbore. In such cases, the value of β is
tion closure pressure pc at a short distance behind relatively small. Following shut-in, the fluid pres-
the fracture tip. As discussed in Section 5.4-5, the sure is relatively constant near the wellbore, and a
fracturing fluid never quite penetrates the near-tip value of β nearer to unity can be expected.
region, referred to as the fluid lag zone (see also It is clear from the preceding discussion that the
Appendix Fig. 4 of the Appendix to Chapter 5). pressure gradients and hence the values of β dur-
The fluid lag zone is characterized by a pressure ing injection and shut-in differ. During injection,
that rapidly decreases from pc at the fluid front to β is selected as the value at the end of pumping βp.
a value at the fracture tip that is approximately Its value for a PKN fracture can be obtained from
equal to the reservoir pore pressure for permeable Nolte (1979, 1991) as
rock or the vapor pressure for relatively imperme-
β p = (n + 2) (n + 3 + a) . (9-19)
able rock.
The effect of the pressure gradient can be incor- The parameter a defines the degree of reduction
porated into the pressure-width relations of Eqs. in viscosity from the well to the fracture tip result-
9-10 through 9-13 by introducing the factor β ing from thermal and shear degradation. For a con-
(Nolte, 1979). β is defined as the ratio of the aver- stant-viscosity profile, a = 0; for a linearly varying
age net pressure in the fracture ∆p–f to the net pres- profile, a = 1 (i.e., effectively zero viscosity at the
sure at the wellbore pnet: tip).
∆p f No expressions similar to Eq. 9-19 have been
β= , (9-15) reported in literature for the KGD and radial mod-
pnet els. An estimate of βp = 0.85 can be inferred from
where Daneshy (1973) for the KGD model. For the radial
model where fluid enters from within a limited set
pnet = pw − pc . (9-16)
of perforations, βp can be much smaller than unity
The term ∆p–f is defined as the net pressure cor- because of the high entrance flow rate and the
responding to a constant internal pressure p–f that consequently enhanced pressure gradient.
would produce the same average width as where During the fracture closing phase, β is selected
a pressure gradient exists along the fracture length. as its value after shut-in βs, and Nolte (1979, 1986a)
Thus, showed that
∆p f = p f − pc . (9-17) (2 n + 2) (2 n + 3 + a) PKN

β s ≈ 0.9 KGD (9-20)
Equation 9-15 can then be expressed as
(3π 2 32) Radial .

p f − pc = βpnet = β( pw − pc ) . (9-18)
The expressions for the KGD and radial frac-
Equation 9-18 includes the pressure gradient tures are approximate and based on the observa-
effect from flow and fluid rheology along the frac- tion that the pressure gradient is concentrated near
ture. In combination with Eqs. 9-10 through 9-14, the fracture tip for these models. In particular, the
it also relates the average fracture width to the value of 3π2/32 = 0.925 was selected for the radial
bottomhole wellbore pressure pw, which is a com- model to enable subsequent cancellation with its
monly available field measurement. inverse in Eq. 9-22.
The factor β incorporates the effects of fluid Figure 9-4 is an example of pressure and flow
pressure gradients in a fracture. Tip-dominated profiles during pumping and after shut-in based
fracturing behavior (see Sidebar 9B) is character- on a numerical simulation using the PKN model.
ized by a relatively constant pressure profile, and In addition to the associated values of β, the figure
the fractures exhibit a value of β that approaches shows that flow in the fracture continues until the
unity. For injection from a limited number of per- fracture closes because of the redistribution of the
forations, such as a horizontal fracture from a ver- stored volume from the larger width near the well
tical well or a transverse vertical fracture from a to the higher rate of fluid loss near the tip. This

9-8 Fracture Evaluation Using Pressure Diagnostics


9B. Pressure response of toughness- response and thus dominates the fracture width and fracturing
dominated fractures process. For example, for assumed values of n = 0.7 and
ptip = 1⁄2∆pµ, the final wellbore pressure is increased by only 3%
M. B. Smith, NSI Technologies, Inc. from the case where ptip is negligible. Alternatively, if ptip is 50%
greater than ∆pµ, then the final wellbore net pressure is just
In its broadest terms, the fracture net pressure that is mea- 6% greater than for the case where ∆pµ is ignored.
sured at the wellbore represents the contribution from fluid vis- The higher injection rates used with very viscous fracturing
cosity and the rock resistance to fracture propagation fluids during large-scale, high-volume fracture treatments
(Shlyapobersky, 1985). The net pressure associated with fluid result in both a relatively large and increasing magnitude of
viscosity results from the flow of the fracturing fluid within the ∆pµ, which changes the pressure response from its early-time
narrow fracture. The dependency of the viscous fracturing tip-dominated behavior to one that is viscous dominated. This
pressure on the rock mechanical and fracture geometry para- change in the relative contribution of ∆pµ compared with that of
meters is provided by Eq. 9-24, which is derived by assuming ptip leads to a corresponding change in the net pressure
that the net pressure is negligible at the fracture tip. behavior on a log-log plot.
Various mechanisms at the fracture tip have been postu- An example is exhibited in Fig. 9B-1 by the pressure
lated to explain the contribution from the rock resistance to the response monitored during a calibration injection into a reser-
fracturing pressure, and these are discussed in Section 6-7. voir with a relatively low Young’s modulus E. The value of pnet
From the perspective of fracture pressure analysis, their contri- is initially constant, signifying toughness-dominated behavior.
bution can be cumulatively represented as the tip-extension At approximately 2.5 min, the pressure response gradually
pressure ptip. The tip-extension pressure can be expressed in increases and eventually approaches a slope of 0.18 on the
terms of the apparent fracture toughness KIc-apparent, which is log-log plot, as expected for a PKN-type fracture with a negligi-
better rationalized from the basis of fracture mechanics. The ble tip pressure (Eq. 9-24). The transition from tip-dominated
relationship between these two quantities generally depends to viscous-dominated behavior is identified from this positive
on the fracture geometry, and for the assumption of a semi- slope on the net pressure plot. In contrast, tip-dominated frac-
circular fracture tip is (Shlyapobersky, 1985) ture behavior would continue to exhibit a relatively constant
net pressure response throughout the treatment.
2
K Ic −apparent = p tip h 2 , (9B-1) The nearly constant pressure response during the initial
π stage of the treatment provides an estimate of ptip, which in
conjunction with Eq. 9B-1 enables the inference of KIc-apparent.
where h is the fracture height in the tip region.
For the example in Fig. 9B-1, ptip is obtained as 260 psi.
Using the PKN fracture geometry model for the viscous
Substituting this value in Eq. 9B-1 with a fracture height
pressure contribution ∆pµ beyond the tip region, the wellbore
of 38 ft suggested by well logs obtains
net pressure can be given as (Nolte, 1991)

  p  2n + 2 
1 ( 2n + 2 )

K Ic −apparent =
2
260 psi
(38 ft × 12 ft / in.) = 4430 psi / in.1 2.
= ∆p µ    + 1
tip
p net , (9B-2) π 2
  ∆p µ  
   
This estimate for KIc-apparent is higher than the critical stress
where the fracture half-length L > h/2 and ∆pµ is the PKN pres- intensity factor KIc commonly measured with laboratory tests
sure contribution from the tip to the wellbore but without the (see Section 3-4.6). The larger value can be attributed to any
rock resistance to propagation. An analytical relation for ∆pµ of the tip mechanisms described in Section 6-7 and results in
for the PKN fracture model beyond the tip region is derived as a correspondingly higher resistance to fracture propagation.
This discussion applies exclusively to an elongated fracture
1 ( 2n + 2 )
E ′  Kh  q i   geometry, as approximated by the PKN fracture model. In con-
n

∆p µ ≅ p net ,PKN ≅ 1.5    (L − Lt ) , (9B-3) trast, both ptip and ∆pµ decrease with continued injection for the
h  E ′  h   radial fracture model. As a result, it is less clear which mecha-
nism dominates the pressure response for radial fractures.
where E ′ is the plane strain modulus, K is the fluid consistency
coefficient, qi is the fluid injection rate, and Lt is the length of
the tip region. Equation 5-19 provides a more specific relation- 1000
ship for pnet with a Newtonian fluid.
The relative contributions of ∆pµ and ptip determine whether Toughness dominated Viscous dominated
the fracture growth is viscosity dominated or tip dominated,
respectively. Usually one phenomena dominates; a nearly
equal contribution from both mechanisms is only rarely
observed. Although ptip generally shows little variation during
pnet

a treatment, Eq. 9B-3 shows that ∆pµ gradually increases with


continued extension. This indicates that a tip-toughness-domi- ptip 0.18
nated fracture response is most likely manifested during the 1
early stage of the fracture treatment. The pressure response
could continue to be toughness dominated if the pressure con-
tribution from ∆pµ remains modest, as is the case for low injec- 100
tion rates, low fluid viscosity (i.e., low K), short fracture lengths 0.1 1.0 10.0
or soft rocks (i.e., low E′). Most of these conditions are likely
Time (min)
during micro-fracture stress tests (see Chapter 3) and tip-
screenout treatments in unconsolidated formations.
The exponents for Eq. 9B-2 suggest that usually one mech- Figure 9B-1. Net pressure response for the transition
anism effectively dominates the wellbore net pressure from toughness-dominated to viscous-dominated frac-
ture growth.

Reservoir Stimulation 9-9


h f
= ∆pf
PKN
1.5
πβ 
1.3 cf = 2 L KGD (9-22)
Average net pressure
2E′ 
(32 3π ) R
1.0 Before shut-in
Net pressure

2
After shut-in
Radial,
0.5 β = ∆p f/pnet where L is the fracture half-length.
βp = 1/1.5 = 0.67 (before)
0 βs = 1/1.3 = 0.77 (after) Martins and Harper (1985) derived the compli-
ance for a fracture that grows as a series of con-
focal ellipses. In this case, the fracture width
Flow rate in fracture

1.0
depends on the elliptic integral of the fracture
Injection rate

Before shut-in
After shut-in aspect ratio. This analysis is applicable during
0.5
Decreasing during closure
fracture growth in an unbounded fashion following
0
initiation from a perforated interval that is shorter
Well 1/2 Tip than the fracture height.
Distance into fracture

Figure 9-4. Pressure and flow rate in a fracture before and


after shut-in (Nolte, 1986a). 9-4. Pressure during pumping
Equations for interpreting pressure during pumping
afterflow causes additional extension during shut-
are developed by combining the basic relations of
in (Perkins and Kern, 1961), which can be signifi-
material balance, fluid flow and rock elastic deforma-
cant for high fluid efficiencies and in reservoirs
tion. The relation between the fracture geometry and
that show a moderate to low resistance to fracture
pressure during pumping was initially proposed by
propagation. Furthermore, shortly after shut-in, the
Nolte and Smith (1981), with application to the PKN-
pressure gradient equilibrates in the near-wellbore
type fracture geometry. This analysis was subse-
region. The redistribution of pressure is accompa-
quently generalized for application to each of the
nied by a change in the wellbore width. The
basic fracture geometry models (Nolte, 1986b).
change in the value of β from injection to the shut-
Extensions were also proposed by Nolte (1991) to
in period reflects the fact that the average width
consider deviations in the fracture geometry from the
should remain relatively constant during this
idealized 2D fracture geometry conditions.
period.
The fundamental relation that defines fracture
• Fracture compliance behavior during pumping at conditions of nearly con-
The net pressure within the fracture compresses stant injection rate and rheology can be obtained by
the formation and results in the fracture width. The combining the fluid flow relation (Eq. 9-2) with that
relation between the net pressure and the fracture of the fracture width and compliance (Eq. 9-21):
width averaged over its length and height 〈w –〉 can
n
be expressed by combining Eqs. 9-10 through 9-13 dp K  qi 
∝ 1+ 2 n   .
( )
with Eq. 9-18: (9-23)
dx c f pnet  hf 
w = c f pnet . (9-21)
Integration over the length with the assumption that
Equation 9-21 indicates that the average fracture pc is a constant and that pnet is negligible at the frac-
width is linearly proportional to the wellbore net ture tip gives
pressure. The constant of proportionality cf is 1/( 2 n + 2 )
referred to as the fracture compliance. The use of  K  q n 
pnet ∝  1+2 n  i  L  . (9-24)
compliance to describe the deformation of solid  c f  hf  
materials under externally applied loads is analo-  
gous to the compressibility of fluid systems during The integration implicitly assumes that the flow
reservoir analysis. The fracture compliance profile along the fracture length has a constant shape.
depends on the formation plane strain modulus This is essentially the case for the three fracture mod-
E′, β coefficient and pertinent 2D fracture geome- els; e.g., the velocity is essentially constant for the
try model: PKN model (Nolte, 1991).

9-10 Fracture Evaluation Using Pressure Diagnostics


Equation 9-24 and the following equations assume qi ηt qi t *
Af = = , (9-28)
that pnet is dominated by the frictional effects from w w
flow of a viscous fluid within the fracture. This
assumption may not be valid under specific condi- where t* is referred to as the reduced time (Nolte,
tions, as discussed in Section 5-4.5. Sidebar 9B out- 1991). The fracture surface area Af for the basic frac-
lines a diagnostic procedure for comparing the net ture geometry models is
pressure contribution from the fracture tip to the net 2 Lh PKN
pressure from fluid flow (or the viscous net pressure) 
predicted by Eq. 9-24. A f = 2 Lh KGD (9-29)
πR 2
Introducing the appropriate compliance relation in  Radial.
Eq. 9-24 for the three models and using L = hf /2 = R Combining Eqs. 9-27 through 9-29, an expression
for the radial model: for the fracture width as a function of reduced time
e
 L  for the three basic models can be obtained:
PKN pnet ∝ ( E ′ Kq )  3n+1 
2 n +1 n e
i
1 ( 2 n +3 )
 h f   Kq n+1 
 1 
e
PKN w ∝ i n  (t *)1 ( 2 n+3)
KGD pnet ∝ ( E ′ Kqi )  n 2 n 
2 n +1 n e
(9-25)  E ′h f 
1 ( 2 n+4 )
 h f L   Kqin+2 
e
KGD w ∝ (t *)1 ( n+2 )
pnet ∝ ( E ′ 2 n+1 Kqin )  3n  ,
n+2 
1 (9-30)
 E ′h f 
e
Radial
R  1 ( 3 n+6 )
 K 2 qin+2 
where the exponent e represents Radial w ∝   (t *)( 2−n ) ( 3n+6 ) .
 E′2 
e = 1 (2 n + 2 ) . (9-26)
Substituting the relation between width and net
On the basis of Eq. 9-21, the fracture width w pressure in Eq. 9-21 into Eq. 9-30 gives the following
at the wellbore is proportional to cfpnet for the three expressions for pnet in terms of the reduced time:
fracture geometry models. Multiplying each side
1/ ( 2 n +3 )
of Eq. 9-25 by the appropriate definition of cf from  KE ′ 3n+1qin+1 
Eq. 9-22 results in PKN pnet ∝  3 n +3  (t *)1/( 2 n+3)
 hf 
pnet ∝ ( KE ′ n+1 )
e 1/ ( n + 2 )
 Kq n  (t *)− n/( n+2 )
[ ] KGD (9-31)
e
PKN w ∝  i  h1f −n L
 E′  Radial pnet ∝ ( KE ′ n+1 )
1/ ( n + 2 )
e
(t *)− n/( n+2 ) .
 Kq n   L2 
e

KGD w ∝  i   n  (9-27) Equation 9-31 indicates that for typical fracturing flu-
 E ′   h f 
e
ids (i.e., n ≈ 0.4–1.0), the fracture pressure during
 Kqin  2−n e
 [R ] .
Radial w ∝  injection is only nominally sensitive to the reduced
 E′  time t* or fluid efficiency η. Consequently, the effi-
These fracturing pressure and width relations indicate ciency, or alternatively the fluid-leakoff coefficient,
that their dependence on the fluid rheology parame- cannot be determined by analyzing pressure during
ters K and n, wellbore injection rate qi and plane fluid injection exclusively.
strain modulus E′ is the same for all the models. Equations 9-30 and 9-31 also show that the net
Their dependence on the fracture extension L or R pressure and fracture width for any efficiency η can
and height hf differs. In addition, Eqs. 9-25 and 9-27 be approximated by their values for the case of no
also show that pnet and w, respectively, have a weak fluid loss, if the time is scaled by ηt. This time scal-
dependence on qi and that for increasing penetration ing is illustrated for the PKN fracture model in
L or R, pnet increases for the PKN model but Fig. 9-5, which shows the net pressure corresponding
decreases for the KGD and radial models. to no fluid loss (i.e., η = 1) and to an efficiency η =
The time dependence of the fracture width and 0.2 at a time of 50 min. The latter case corresponds
pressure is developed using the definition of η from to a reduced time of t* = ηt = 50 × 0.2 = 10 min.
Eq. 9-4 at a constant injection rate qi (i.e., Vi = qit): Figure 9-5 illustrates that the net pressure at a time of
50 min for the fluid-loss case is equal to the net pres-

Reservoir Stimulation 9-11


R ∝ t 1/ 4 η→ 0
1000 Radial (9-34)
η=1 R ∝ t ( 2 n+2 ) ( 3 n+6 ) η → 1.
in
at 50 m
600 η = 0.2
Limiting expressions for the fracture net pressure

Efficiency, η
400 0.4 are similarly outlined in the Appendix:
pnet (psi)

pnet ∝ t 1/ 4 ( n+1) η→ 0
PKN (9-35)
200 0.2 pnet ∝ t 1 ( 2 n+3) η→1

pnet ∝ t − n /2 ( n+1)
Net pressure
100
Efficiency
0.1
η→ 0
KGD (9-36)
2 5 10 20 50 pnet ∝ t − n ( n+2 ) η→1
Injection time (min)
pnet ∝ t −3n /8( n+1) η→ 0
Figure 9-5. Reduced time illustrated for PKN fracture Radial (9-37)
geometry (Nolte, 1991). pnet ∝ t − n ( n+2 ) η → 1.
Each of these bounding expressions for the net
sure derived for the particular case of no fluid loss at pressure is a power law relation. Consequently, the
a time of 10 min. log-log graph of net pressure versus time should yield
This observation is significant because simple ana- a straight line with a slope equal to the respective
lytical expressions for the three basic models are exponent: positive for PKN behavior and negative for
readily available when η → 1. The various fracture KGD and radial behavior. In particular for PKN
parameters for any generalized value of η can then behavior, the log-log slope for commonly used frac-
be obtained from this limiting conditions merely by turing fluids (i.e., n ≅ 0.5) is typically less than 1⁄4 and
scaling the time by a factor of 1/η. decreases as the efficiency decreases. The log-log
plot of the net pressure versus time during injection,
commonly known as the Nolte-Smith plot, forms the
9-4.1. Time variation for limiting fundamental basis for the interpretation of pressure
fluid efficiencies profiles during fracturing and is analogous to the log-
log diagnostic plot for reservoir flow, as discussed in
Approximations for the time dependency of the frac- Chapter 2.
ture penetration and pressure can be derived from the
equations presented in the previous section for the
two extreme values of the fluid efficiency η. These 9-4.2. Inference of fracture geometry
limiting cases are for very high and low fluid effi- from pressure
ciencies, approaching 1 and 0, respectively. This sim-
plification provides bounding expressions for the The primary reservoir interval is bounded on both
fracture penetration and related pressure. A similar sides by shale formations in the majority of fracturing
approach is used in Section 9-5 to derive relations for applications. Shale zones are generally at higher
analyzing pressure decline during the shut-in period. stress and provide the primary barrier to fracture
Following the mathematical derivations outlined in height growth, particularly during the initial stage
“Mathematical relations for fluid loss” in the Appen- of fracture propagation. The restriction of fracture
dix to this chapter, it can be shown that the fracture height growth is important in low- to moderate-per-
penetration is bounded in the following fashion: meability formations, where relatively long fractures
are required for effective stimulation. Figure 9-6
shows the evolution of the fracture geometry and the
PKN L ∝ t 1/ 2 η→ 0 (9-32)
corresponding wellbore pressure for fracture propaga-
L ∝ t ( 2 n + 2 ) ( 2 n +3 ) η→1 tion under these conditions.
The initial character of fracture propagation, labeled
KGD L ∝ t 1/ 2 η→ 0 (9-33)
as stage 1 on the figure, depends on the length of the
L ∝ t ( n+1) ( n+2 ) η→1 perforation interval providing fluid entry into the frac-
ture relative to the reservoir thickness. Two limiting

9-12 Fracture Evaluation Using Pressure Diagnostics


source, and the fracture area evolves in an elliptical
Point Source Line Source
Stage 1 Stage 1 shape. The KGD geometry model best describes the
Well early phase of this fracture growth (Martins and
Harper, 1985).

+ + + ++ +
Radial
For either the radial or elliptical propagation mode
model KGD model during stage 1, the net pressure decreases with con-
+

tinued injection. It also exhibits a log-log slope


Barrier between –1⁄8 and –1⁄4 depending on the fluid rheology
∆σ exponent n and the efficiency η, as in Eqs. 9-36 and
9-37. The decreasing pressure reflects the fracture’s
PKN model pc = σmin
preference to grow with decreasing resistance and in
1 Stage 3
Stage 2 an unrestrained fashion as it gets larger. Stage 1 may
occur for only a short time for fracture initiation
within a relatively small interval or for the entire
treatment in a massive zone (Smith et al., 1987).
When barriers at higher stress exist above and
Linear Plot of Pressure
below the reservoir pay zone, fracture height growth
Barriers could be confined following stage 1. Under these
conditions, the fracture is prevented from expanding
pressure, pw
Bottomhole

1 3 in its preferred circular shape and fracture length


2
pc = σ
extension is promoted. This mode of propagation is
denoted as stage 2 in Fig. 9-6 and results in increas-
Time ing pressure as the fracture becomes long relative to
its vertical height. This type of fracture propagates
Log Plot of Net Pressure in a manner similar to the PKN model. For this stage,
the log-log slope of the fracturing pressure is between
log(pnet = pw – pc)

3
1
⁄4 and 1⁄8, once again depending on n and η (Eq. 9-35).
2
Confined fracture height with its characteristic
1
positive log-log slope can be expected until the frac-
turing net pressure approaches a value that is approxi-
log (time) mately one-half of the stress difference ∆σ to which-
ever stress barrier bounding the fracture has the lower
Figure 9-6. Evolution of fracture geometry and pressure stress value. At this magnitude of the net pressure,
during pumping.
the fracture begins to penetrate in a restricted, or con-
trolled, fashion into the adjacent barrier layer with the
cases are described: a limited fluid entry interval and lower stress value. The fracturing pressure continues
one where fluid entry occurs over the complete reser- to increase with penetration, although at a rate that is
voir thickness. Short fluid entry intervals (i.e., limited progressively less than for the PKN model. This con-
perforation intervals) may be desired in vertical well- dition of fracture propagation is indicated as stage 3
bores to mitigate the occurrence of near-wellbore in Fig. 9-6.
problems (see Section 11-3.2). They also occur in hor- If one of the formation barriers is absent (i.e., ∆σ
izontally oriented fractures, during the placement of = 0), height growth into the higher stress barrier is
transverse hydraulic fractures in a horizontal well or arrested. The fracture height, however, continues to
in wellbores that are inclined with respect to the plane grow essentially in a radial-like fashion along the
of σmin. The limited fluid entry into the fracture is direction where the barrier is absent and exhibits a
approximated by a point source. As shown in Fig. 9-6, continuously decreasing pressure (stage 1). This frac-
the fracture area increases in a circular shape for a ture height growth pattern could also occur when
point-source fluid entry and hence is best described fractures are deliberately initiated from zones at
by the radial geometry model. Fluid entry over the higher stress and propagated into bounding layers
complete reservoir thickness is approximated by a line at lower stress, as during an indirect vertical fracture
completion (IVFC; see Section 5-1.2).

Reservoir Stimulation 9-13


The magnitude of the net pressure during stage 2 Higher stress barriers normally have only a limited
(i.e., PKN-type fracture growth) can be used to infer extent. Growth through barriers could eventually be
the magnitude of the fracture compliance based on followed by uncontrolled height growth, resulting in
Eq. 9-31 and therefore the average width using adverse effects during fracturing (see Section 9-4.5).
Eq. 9-30. The net pressure exhibits a relatively small It thus is necessary to estimate the primary parame-
and decreasing value during stage 1, and it cannot be ters that govern fracture height growth: the magni-
effectively used to estimate fracture dimensions. In tude of the stress difference ∆σ between the reservoir
such cases, significant errors could be introduced and barrier and the bounding zone thickness. Figure
owing to uncertainties in the closure pressure and its 9-7, for idealized conditions such as Eq. 6-49,
change resulting from poroelastic effects (see Section addresses these requirements. The figure assumes that
3-5.4). the upper and lower barriers have the same stress
Decreasing net pressure during the initial growth value and an infinite extent. It also approximates
period indicates a radially evolving fracture in either height growth into bounding zones with unequal
the horizontal or vertical plane. Alternatively, an stress magnitudes. As shown, controlled height
increasing net pressure with a small log-log slope (i.e., growth depends on the ratio of pnet and ∆σ. For a
between 1⁄8 and 1⁄4) after the initial growth period is ratio of about 0.4, negligible height growth occurs;
indicative of a vertical fracture extending primarily for a ratio of about 0.65, the total fracture height is
in length with restricted height growth. Following this twice the initial fracture height hi and each barrier
period, if a reduction in the rate of pressure increase thickness must be at least one-half of the height of
is observed, fracture height growth into a barrier zone the reservoir to ensure continued controlled height
should be expected. The net pressure during the growth. Figure 9-7 implies that this condition
period of height growth is governed primarily by the requires a barrier thickness that is at least equal to
difference in stress between the primary reservoir and the height of the reservoir for a ratio of about 0.8.
penetrated zones. Consequently, the pressure response It follows from the previous discussion that the
during the height growth period can be used to esti- amount of height growth into the bounding zones
mate the stress difference, as discussed in the next depends on the thickness of the reservoir and the ratio
section. The stress of the bounding formation is an pnet /∆σ. The reservoir thickness is defined using stan-
important parameter for fracturing design, and it can dard well logs whereas the magnitude of the net pres-
also be used to calibrate log-inferred values of stress sure is estimated from a fracture simulator. The
(see Chapter 4). appropriate fracture model for assessing net pressure
In conclusion, this discussion indicates how the and growth into barriers is based on the PKN model
pressure response during pumping can provide infor-
mation on the state of stress, type of fracture created
and fracture geometry, or more generally the fracture
∆σ
volume term for the material-balance relation in
Eq. 9-5.
4 4
hi

9-4.3. Diagnosis of periods of controlled


Width compliance, cf/cf(hi)

fracture height growth


Fracture height, hf/hi

Fracture height growth into bounding barrier zones at


2 2
higher stress requires an increasing pressure response
prior to the height growth period into the barriers.
Height growth into a higher stress barrier thus cannot
occur during stage 1, where the decreasing pressure 1.1
characteristic of the KGD and radial models occurs. 1 1
0.1 0.2 0.4 0.6 0.8 1.0
Height growth into higher stress barriers (stage 3 in Net pressure, pnet/∆σ
Fig. 9-6), however, is a commonly occurring devia-
tion from the constant-height assumption of a PKN- Figure 9-7. Net pressure and compliance for idealized
type fracture. fracture height growth (Nolte, 1986a).

9-14 Fracture Evaluation Using Pressure Diagnostics


(Fig. 9-6). Equation 9-25 provides this relation and Table 9-1. Treatment parameters and rock
indicates the importance of the initial fracture height mechanical properties for controlled
(i.e., height of the reservoir) on the magnitude of the fracture height growth example.
net pressure and hence the tendency for growth into
E 5.3 × 105 psi n 0.44
barriers. The net pressure is approximately inversely
proportional to the height. Therefore, the net pressure ν 0.22 K 0.248 lbf-sn/ft2
approximately doubles when the initially fractured hL 24 ft a 0 (constant viscosity)
zone height is halved. The general conclusion is that
Calibration test
a smaller zone is more likely to experience height
growth and will require both higher stress differences Vi 95 bbl tp 12 min

and thicker barrier zones for controlled fracture


height growth. height hL of 24 ft and is bounded by higher stress
Following fracture height growth into a barrier, the shale barriers on both sides (Fig. 9-8a). As shown
efficiency is relatively constant because no fluid loss in Fig. 9-8b, the calibration injection lasted for
is expected to occur in the barrier zone. The pressure 12 min and was followed by an extended shut-in
response during this period deviates from its other- period of approximately 40 min. The shut-in
wise straight-line response on the log-log plot and is pressure approached the far-field reservoir pressure
characterized predominantly by height growth behav- (≈8100 psi) within a relatively short shut-in period
ior. It thus can be used to identify the onset of barrier because of the high formation permeability (≈250
penetration and to estimate the magnitude of the stress md). The closure pressure pc for the formation was
difference (Nolte, 1991; Ayoub et al., 1992a). It relies
on the use of a characteristic signature of the pressure (a)
derivative of the fracturing pressure during controlled
height growth. Because of its increased sensitivity,
the pressure derivative magnifies the deviation in the
fracturing pressure from its expected response and
therefore enhances the identification of fracture
height growth, as discussed in the following section.
The specialized pressure derivative diagnostic has
several applications in addition to quantifying frac-
ture height growth, as discussed in Sidebar 9C: to
validate the fracture geometry inferred from the log-
log plot and to objectively confirm fracture closure
pressure, as well as its capacity to identify the onset
of a screenout. (b)

9400 20
9-4.4. Examples of injection 9200
Calibration test 18 Step rate
Bottomhole pressure (psi)

pressure analysis 16
Injection rate (bbl/min)

9000 Bottomhole pressure 14


Two field examples are presented here. The first 8800 12
example describes log-log analysis for a reservoir 10
8600 8
bounded by shale barriers. The second example dis-
8400 6
cusses the pressure response for a radial fracture
4
geometry in a reservoir where higher stress barriers 8200
Injection rate 2
are absent. 8000 0
0 10 20 30 40 50 60 70 80 90 100
• Example of controlled fracture height growth Time (min)
Table 9-1 lists the parameters relevant to the analy-
sis of a calibration test in a gas-bearing sandstone Figure 9-8. Calibration test analysis for controlled height
growth. (a) Well logs. (b) Bottomhole pressure and rate
reservoir. The producing interval has a permeable record.

Reservoir Stimulation 9-15


9C. Pressure derivative analysis for the net pressure response for the simulated treatment in
diagnosing pumping pressure Fig. 9C-2. The pressure derivative registers a rapid increase at
50 min, identifying the occurrence of a TSO, whereas no signifi-
Joseph Ayoub, Schlumberger Dowell cant change in the net pressure response is visible until later.
Another observation from Fig. 9C-2 pertains to the long-
The fracture net pressure exhibits a power law variation with term pressure derivative response after a TSO. The log-log
respect to time, as demonstrated in Section 9-4. This can be slope becomes greater than 1 with continued injection. The
generalized as ratio of the pressure derivative and the net pressure eventu-
ally becomes larger than 1. Consequently, after a TSO the
p net = p w − p c = At b , (9C-1) pressure derivative eventually becomes larger than the net
presure, as shown on the figure.
where A is a constant and the exponent b is the slope of the Figure 9C-2 also shows an increase in the pressure deriva-
log-log plot of the net pressure versus time t. The slope b tive at 25 min. This response is attributed to increased viscos-
depends on the fracture geometry and the fluid rheology and ity caused by the introduction of proppant. The pressure deriv-
efficiency. ative can be used to assess the importance of the apparent
For pressure data measured during a fracturing treatment, rock toughness relative to the viscous pressure (Nolte, 1991).
the slope depends on the choice of the fracture initiation time
and closure pressure. The fracture initiation time is selected
by examining the pressure record during pumping, and it 10,000

Net pressure or pressure derivative (psi)


often coincides with the time when the fracturing fluid first Underestimated pc
reaches the perforations. The closure pressure is indepen-
dently estimated using one of the techniques discussed in
“Estimating closure pressure” in the Appendix to this chapter.
ct pc
Significant uncertainty is often associated with its determina- Corre
tion, which could result in an incorrect interpretation; e.g.,
treatments for which the pressure data exhibit a small 1000 pc
increase in pressure tend to exhibit a relatively constant net ated
estim
Over
pressure response if a lower closure pressure estimate is
e
used. rivativ
ure de
This sidebar introduces the pressure derivative to enhance Press
the fracturing injection pressure diagnosis and analysis
(Ayoub et al., 1992a). The pressure derivative was initially
introduced in well testing, for which it quickly became stan-
dard practice because it significantly enhances the identifica- 100
tion of various flow regimes during the analysis of transient 1 10 100
well test data (Bourdet et al., 1989). Similarly, when applied Time (min)
to the fracturing injection pressure, the derivative “magnifies”
and detects fracturing events earlier in time. It also assists in Figure 9C-1. Closure estimation using pressure deriv-
the determination of closure pressure. Differentiating Eq. 9C-1
ative analysis.
with respect to time gives
d (pw − p c ) dp w
= = Abt b −1, (9C-2)
dt dt
and multiplying both sides of Eq. 9C-2 by t gives 2000
Net pressure or pressure derivative (psi)

Tip screenout
dp w
t = Abt b . (9C-3) Begin proppant
dt
1000
pnet
Defining the left side of Eq. 9C-3 as the pressure deriva-
tive, it follows that
• Pressure derivative versus time exhibits the same log-log 500
slope as the net pressure.
• Net pressure and pressure derivative are separated by a
factor of 1/b on a log-log plot.
The pressure derivative is independent of the particular 200
choice of closure pressure and is thus unaffected by errors Pressure derivative
in its determination. For typical PKN, KGD and radial fracture
behavior, the closure pressure can thus be inferred from the
injection pressure by selecting a value that makes the net 100
pressure response parallel to the pressure derivative on a log- 5 10 20 50 100
log plot (Fig. 9C-1). This feature of pressure derivative analy- Injection time (min)
sis was applied to the calibration test in Fig. 9-8 to confirm the
closure pressure magnitude (Fig. 9-9).
The pressure derivative magnifies fracturing events Figure 9C-2. Pressure derivative analysis for a TSO
because of its enhanced sensitivity. This characteristic of the response (Nolte, 1991).
pressure derivative is used to quantify fracture height growth
into higher stress bounding zones, as for the examples in
Section 9-4.4. The occurrence of a tip screenout (TSO) is also
magnified and can be detected earlier in time. This is noted on

9-16 Fracture Evaluation Using Pressure Diagnostics


inferred to be 8910 psi from the step rate test (see Thus, the stress difference between the lower
“Estimating closure pressure” in the Appendix to stressed bounding zones and the perforated interval
this chapter) that followed the calibration treatment. is approximately 360 psi.
The log-log plot of the injection pressure in • Example of radial fracture growth
Fig. 9-9 is similar to the idealized example shown
The example of a radially propagating fracture was
in Fig. 9-6 and can be interpreted in the same
inferred during a calibration treatment performed
manner. The initially decreasing pressure response
in a high-permeability, heavy-oil-bearing sand-
is representative of either a KGD or radial mode
stone reservoir (Fig. 9-10 and Table 9-2). The cali-
of fracture propagation. The initial log-log slope
bration treatment was preceded by a short injection
of –0.18 is between the bounds suggested by
using completion fluids (also called a mini-falloff
Eqs. 9-36 and 9-37 for the KGD and radial models,
injection) and a step rate test. The mini-falloff test
respectively, for the value of n in Table 9-1. The
is used to characterize the reservoir producing
subsequent log-log slope of 0.16 indicates a period
parameters and is discussed in Section 9-6. The
of fracture extension in the PKN mode and a low
formation closure pressure was estimated to be
fluid efficiency.
4375 psi from the step rate test (see “Estimating
1000 (a)

Height
Radial PKN growth
pnet or pressure derivative (psi)

100

10

pnet
1 Pressure derivative
0 1 10
Time (min)

Figure 9-9. Log-log net pressure and pressure derivative


analysis for the calibration treatment in Fig. 9-8.
(b)
The reduced rate of pressure increase during the 5500 100
last 3.5 min of injection is attributed to fracture Mini- Step Cali-
Bottomhole pressure (psi)

5000 falloff rate bration


Injection rate (bbl/min)

growth into the higher stress bounding shales. This Bottomhole 75


pressure
diagnostic is supported by the constant 36-psi 4500
value of the pressure derivative during this period. 4000 3726 psi
50
The magnitude of the stress difference is approxi-
mately 10 times the constant pressure derivative 3500
25
value during this period (Nolte, 1991): 3000
Injection rate

1  d ( pnet )  2500 0
≈ 0.1
∆σ  dt 
t (9-38) 25 75 125 175
Time (min)

 d ( pnet )  Figure 9-10. Calibration test analysis for radial fracture


∆σ = t   0.1 = 36 0.1 = 360 psi. (9-39) growth. (a) Well logs. (b) Bottomhole pressure and rate
 dt  record.

Reservoir Stimulation 9-17


Table 9-2. Treatment parameters and rock 9-4.5. Diagnostics for nonideal
mechanical properties for radial fracture fracture propagation
growth example.
The log-log diagnostic for injection pressure, pre-
E 4.5 × 105 psi n 0.40 sented in the previous sections, is based on idealized
ν 0.25 K 0.084 lbf-sn/ft2 behaviors for fracture height growth and fluid leakoff.
This section discusses common conditions that cause
a 0 (constant viscosity)
deviation from the idealized behaviors and could
Mini-falloff test result in treatment failure from a premature near-
Vi 14.75 bbl tp 3 min wellbore screenout during the proppant treatment.
Calibration test • Rapid growth through a barrier—uncontrolled
Vi 3
107 bbl [600 ft ] tp 4.6 min fracture height growth
Moderate or controlled fracture height growth into
higher stress zones following a period of confined
closure pressure” in the Appendix to this chapter)
fracture extension is described in Section 9-4.3.
and an after-closure linear flow analysis (see
Uncontrolled or runaway fracture height growth
Section 9-6).
occurs when the higher stress zone is traversed and
The log-log net pressure plot for the calibration
the fracture extends into a lower stress zone. This
injection (Fig. 9-11) indicates a slope of –0.11.
is shown on Fig. 9-12, where fracture growth dur-
This slope is within the bounds of a radial fracture
ing stages a and b is the same as that in stages 2
suggested by Eq. 9-37 for the value of the fluid
and 3, respectively, on Fig. 9-6 (The initial stage 1
rheology exponent n in Table 2. A radially grow-
period on Fig. 9-6 of radial- or KGD-type fracture
ing fracture should also be expected in this case
growth is not shown on Fig. 9-12.) Stage b ends
because the reservoir lacks significant shale zones
when the fracture enters a lower stress zone. When
(Fig. 9-10a) that potentially could have constrained
this occurs, the fluid pressure is greater than the
the fracture height.
stress of the zone, initiating an accelerated rate of
growth that leads to stage c. This uncontrolled
growth begins at the well, where the pressure is
1000

b
log pnet

a c
Net pressure, pnet (psi)

log t
pnet

100
0 1 10
Time (min) hi a b c a b c

1 hf/hi
Figure 9-11. Log-log net pressure analysis for the calibra-
tion treatment in Fig. 9-10.

Figure 9-12. Pressure and width for height growth through


a pinch point.

9-18 Fracture Evaluation Using Pressure Diagnostics


greatest, and progresses farther along the fracture of the fracture geometry was inferred following
as pumping continues. In a similar manner, uncon- a consistent evaluation of the propped treatment
trolled height growth ceases progressively from using a pseudo-three-dimensional (P3D) fracture
the fracture tip to the wellbore when another simulator (see Section 6-3) and the methodology
higher stress barrier is reached. The fracturing outlined in Section 9-8.
pressure then increases again, indicating either As in the case of a radial fracture, decreasing net
height confinement or a controlled rate of barrier pressure generally indicates uncontrolled height
penetration. growth. The pinch point resulting from height
Figure 9-12 also indicates that the pressure growth to a lower stress zone can cause proppant
between stage b and stage c is relatively constant to bridge at its location during the proppant treat-
and is regulated by a pinch point. During the initial ment but will allow fluid to pass through freely.
penetration into the lower stress zone, the pinch Consequently, excessive dehydration of the slurry
caused by the higher stress barrier layer results in a and decreasing width in the primary reservoir zone
fracture width of nearly zero that closes if the pres- can result in a rapid screenout, even at low prop-
sure decreases. The reduced fracture width also pant concentrations.
causes additional pressure loss in the vertical • Horizontal fracture components—pressure greater
direction, which limits fluid flow vertically into the than overburden
lower stress zone. This pinching mechanism regu-
For the normal state of rock stress in moderate to
lates the pressure to a nearly constant value until
deep reservoir depths, the horizontal stress is less
the penetration becomes sufficiently large to main-
than the overburden or the vertical stress, as dis-
tain a reasonable open width within the barrier
cussed in Section 3-5. When the bottomhole treat-
layer. After the pinch point is overcome for stage c,
ing pressure is less than the overburden stress, a
the rate and extent of vertical growth increase sig-
fracture can propagate only in the vertical plane.
nificantly, accompanied by decreasing pressure
A vertical fracture can also contain a horizontal
and width in the primary reservoir.
component when the pressure exceeds the over-
Uncontrolled fracture height growth is charac-
burden or vertical stress component. This condi-
terized by declining pressure, as indicated on
tion may occur
Fig. 9-12. Figure 9-12 also shows that during the
phase of uncontrolled fracture growth (stage c), – at shallow depths where erosion has removed
a rapid increase in the fracture height hf is accom- some of the overburden to reduce the vertical
panied by a corresponding decrease in the frac- stress
turing net pressure pnet. In contrast, the onset of – in reservoirs in tectonically active thrusting
controlled fracture height growth results in a nomi- environments or in geopressured reservoirs.
nal change in the rate of the net pressure increase, Both conditions increase the horizontal stress.
as discussed in Section 9-4.3. Controlled height – in formations with low in-situ shear strength that
growth, therefore, is relatively difficult to identify undergo stress relaxation resulting in an increase
on a net pressure log-log plot, particularly during of the horizontal stress.
its early stage. The fracture geometry under these conditions
The net pressure response for uncontrolled could have both a vertical component and a hori-
height growth is shown by the field example zontal component. This geometry is called a T-
in Fig. 9-13. A pinch point above the perforated shaped fracture. Examples where T-shaped frac-
interval was identified on the stress log developed tures resulted were reported for the fracturing of
using sonic measurements (see Chapter 4). Frac- a shallow coal bed (Mahoney et al., 1981), shallow
ture height growth into the lower stress sandstone limestone formation (Wood et al., 1983) and lami-
above the pinch point commenced after approxi- nated sandstones (Fragachan et al., 1993). The cor-
mately 150 min of injection and was accompanied responding pressure response and vertical cross
by a steady decrease in the fracture pressure. sections of the width profile are illustrated in
Uncontrolled fracture height growth was also Fig. 9-14. The figure indicates that stage c has a
confirmed using radioactive isotopes that were nearly constant pressure response. Growth into the
injected during the treatment. The time evolution higher stress horizontal plane is similar to uncon-

Reservoir Stimulation 9-19


(a)
30
Calculated pressure
Simulated pressure
5500 60
Slurry rate
Proppant concentration

Proppant concentration (ppg)


50
Bottomhole pressure (psi)

20

Slurry rate (bbl/min)


4500 40

30

10
3500 20

10

2500 0
50 100 150
Treatment time (min)
(b)

5400

5600
Depth (ft)

5800

6000
3500 5000 –0.6 0.6 0 500 1000 1500
Stress (psi) Width (ft) Fracture half-length (ft)

Figure 9-13. Pressure response for accelerated height growth. (a) Bottomhole pressure match plot. Calculated pressure is
from surface pressure. (b) Fracture profile at the end of injection.

trolled vertical growth beyond a pinch point, as in al., 1988). The horizontal fracture component
Fig. 9-12. Although uncontrolled vertical growth increases the area available for fluid loss and
commences at a pressure less than that of the stress decreases the treatment efficiency. In addition, the
barrier, T-shaped fracture growth occurs at a pres- horizontal component readily accepts fluid but pre-
sure slightly larger than the vertical stress. vents proppant from entering because of its limited
The width of the horizontal fracture component width. Both effects can excessively dehydrate the
is narrow and has twin pinch points at the juncture slurry in the vertical component, which could lead
with the vertical component because of the elastic to premature screenout.
interaction of the two components (Vandamme et

9-20 Fracture Evaluation Using Pressure Diagnostics


The pressure magnitude provides a diagnostic basis
for determining whether the fracture plane is entirely 10,000
vertical or has a horizontal component as well. The
horizontal component occurs when the fracture pres-
sure is nearly constant and approximately equal to the
overburden stress of the formation, as illustrated in

Net pressure, pnet (psi)


Fig. 9-14. The magnitude of the overburden stress pw = 10,145 psi
can generally be estimated (see Section 3-5.1) and = overburden stress

should always be compared with the magnitude of 1000


the bottomhole fracturing pressure as part of fracture
pressure analysis.
PKN-type fracture T-shaped
fracture
log pnet

b c pw ≈ overburden
100
a
10 100
Time (min)
log t
Figure 9-15. Calibration treatment pressure response for a
Overburden T-shaped fracture.

of the rock matrix is negligible. Natural fissure


systems are highly directional and show a prefer-
σ
ential orientation along a single axis (Warpinski,
1991). The fissures can produce a complicated
fracture behavior because of fracture offsets,
σ < pw < overburden pw ≥ overburden pw ≈ overburden
enhanced fluid friction along the fracture length
(a) (b) (c)
and the creation of secondary fracture strands.
Figure 9-14. Pressure response for T-shaped fractures.
An important effect of natural fissures is enhanced
leakoff, which can lead to a premature screenout
during proppant injection.
This pressure diagnostic was used to infer the
Natural fissures have a negligible effect on the
occurrence of a T-shaped fracture during a calibra-
fluid leakoff process if the reservoir matrix perme-
tion treatment performed in a tectonically active
ability is high. In low-matrix-permeability condi-
reservoir at a depth of 9750 ft. The log-log net
tions, however, the transmissibility of natural fis-
pressure plot (Fig. 9-15) shows a positive slope
sures can be significantly higher than that of the
lasting approximately 75 min, which indicates a
reservoir matrix. The fracturing fluid can readily
confined mode of fracture extension. The pressure
penetrate into natural fissures during the fracturing
subsequently stabilized at approximately 10,145 psi.
process and maintain a pressure nearly equal to
This constant pressure value is the magnitude of
the pressure in the primary fracture. This process
the vertical stress component for an overburden
is illustrated in Fig. 9-16 in terms of the pressure
gradient of 1.04 psi/ft. During the remaining part
response and a horizontal cross section of the
of the treatment, the penetration of the vertical frac-
width profile. Stage a represents the normal frac-
ture component becomes less efficient because of
turing response in which the fluid pressure is
the propagating horizontal fracture component.
lower than the normal stress on the fissure σf, and
• Natural fissure opening—enhanced fluid loss hence there is a relatively small increase in the fis-
Natural fissures can be important for hydrocarbon sure transmissibility. A continued increase in the
production in the majority of low-permeability fluid pressure, however, reduces the effective
reservoirs, particularly where the permeability stress acting to close the fissure. This effect is

Reservoir Stimulation 9-21


Nolte and Smith showed that the wellbore net
pressure required for fissure opening is
c
a b σ H ,max − σ h,min
log pnet pnet , fo = , (9-40)
(1 − 2 ν)
log t
where σH,max and σh,min are the maximum and mini-
σ
mum far-field principal horizontal stresses, respec-
σf σf
tively. Equation 9-40 applies to PKN-type frac-
p < σf p ≈ σf p > σf
tures, and the fissure can have any orientation with
respect to the main fracture. An important observa-
(a) (b) (c)
tion is that a significant difference between the two
principal stresses in a horizontal plane is required
Figure 9-16. Pressure and width for opening natural to ensure effective fracturing in fissured reservoirs.
fissures. The second model, presented by Warpinski
(1991), is applicable to reservoirs where natural
fissures are the primary source of permeability.
partially compensated for by an increase in the
This model provides a more detailed description
normal stress across the fissure σf resulting from
of the effects resulting from natural fissures. It pre-
the effect of Poisson’s ratio from the main
dicts an enhanced rate of fluid loss throughout the
hydraulic fracture. The overall effect, however,
treatment, with an accelerating effect as the frac-
is a decrease in the effective stress on the fissure.
turing pressure increases (stage a of Fig. 9-16).
When the fluid pressure eventually exceeds σf,
The increase in fracture pressure reduces the effec-
the effective stress on the fissure becomes nega-
tive normal stress acting to close the fissures and
tive. The fissure mechanically opens at this stage,
hence increases their permeability, as described in
labeled stage b in Fig. 9-16. A significant portion
Sidebar 9D. For hydraulic fracturing purposes, the
of the injected fluid can be lost during this process
effect of the magnified permeability is reflected as
because of the large number of fissures that can
an increase in the fluid-leakoff coefficient. As dis-
open at this critical pressure. The conductivity of
cussed in Sidebar 9D, the fluid-leakoff coefficient
fissures is increased by orders of magnitude when
in the presence of natural fissures could be as high
the threshold value is exceeded. The corresponding
as 2 to 3 times that for normally occurring pres-
increase in fluid loss at an essentially constant net
sure-dependent leakoff behavior, even under the
pressure greatly reduces treatment efficiency and
net pressure conditions for stage a.
can lead to excessive slurry dehydration in the
Injection pressure during fracturing for these
main body of the fracture and premature screenout
conditions exhibits a continuously decreasing
(stage c).
slope on a log-log plot, indicating a progressively
Optimized fracturing of a fissured formation
increasing rate of fluid loss. Under continued fluid
requires a model to predict the effect of the fis-
injection, the negative effective stress condition
sures on fluid loss. Two fissure models have been
described by the Nolte and Smith model can occur
reported in literature. The first was proposed by
to open fissures and regulate the pressure to a con-
Nolte and Smith (1981). Their assessment pertains
stant threshold value (stage b of Fig. 9-16). Frac-
to fissured formations with only a slight fluid-loss
turing pressure during pumping in the presence
enhancement unless the threshold fluid pressure
of fissures, therefore, behaves in a manner similar
required for fissure opening is exceeded. The fis-
to the pressure response during periods of con-
sures enlarge when the fluid pressure within the
trolled fracture height growth. Because of this sim-
fracture—and the pressure inside the fissures—
ilarity, the primary diagnostic for distinguishing
becomes greater than the normal stress acting to
between height growth and natural fissures is pres-
close them. When this occurs, fluid loss becomes
sure decline data, as discussed subsequently in
significant and represents the worst-case situation
Sidebar 9F.
for fissure-related leakoff behavior.

9-22 Fracture Evaluation Using Pressure Diagnostics


9D. Fluid leakoff in natural fissures σ H ,max − σ h ,min (9D-1)
p net ,fo =
,
1− 2ν
Norman R. Warpinski, Sandia National Laboratories where pnet,fo is the wellbore net pressure for fissure opening,
σH,max and σh,min are the maximum and minimum horizontal
Control of fluid leakoff is one of the critical elements for opti-
stresses, respectively, and ν is Poisson’s ratio. Equation 9D-1
mizing hydraulic fracture treatments. The primary reasons for
applies to vertical fissures at any orientation and shows that
leakoff control are to ensure that the fracture acquires the
fissure opening is a likely possibility in formations where the
desired penetration, to keep the sand slurry sufficiently
difference in the horizontal stresses is low. When the fissure
hydrated to flow readily and to limit the invasion of potentially
opens, it behaves much like a hydraulic fracture, accepting
damaging fluids into the formation. In most fracture treat-
large amounts of fluid and resulting in leakoff coefficients that
ments, it is assumed that the leakoff coefficient is constant.
increase by orders of magnitude (thus the term accelerated
However, in naturally fissured formations the assumption of a
leakoff).
constant leakoff coefficient may cause considerable problems
For slightly elevated pressures, the conceptual model is
during the execution of the treatment and seriously hinder
more complicated. The fissure porosity begins to open as the
productivity.
pore pressure increases because the elevated pressure
The concept that natural fissures, or fractures, could alter
relieves some of the net stress on the asperity contacts.
leakoff has been examined or accounted for in some studies
Several models of this process have been developed, the
(Nolte and Smith, 1981; Castillo, 1987; Nolte, 1991; Warpinski,
most well known of which is by Walsh (1981). In his model,
1991; Barree and Mukherjee, 1996) through the use of fis-
the change in permeability of the fissure resulting from
sure-opening conditions or pressure-sensitive leakoff equa-
changes in stress and pressure is
tions. Figure 9D-1 shows conceptually the ways that elevated
pressure could affect natural fissures. Fissures with rough 3
  σ  
surfaces and minimal mineralization are most likely highly k = k o C ln ref   , (9D-2)
sensitive to the net stress pushing on them. Under virgin   σ − p  
reservoir conditions (i.e., the pressure p within the fissure
equals the initial reservoir pressure pi), the effective stress is where C and σref are constants determined from empirical
fairly high and the open slot pores are most likely deformed data, ko is the initial fissure permeability, σ is the stress on the
and closed. As the pressure in the fissure increases because fissure, and p is the pore pressure in the fissure.
of leakoff of the high-pressure fracturing fluid (p > pi), the net The Walsh model has been shown to reasonably represent
closure stress is reduced and the fissure porosity opens. In data from a fissured reservoir in the Mesaverde formation in
this regime, the leakoff coefficient is highly pressure depen- the Piceance basin of Colorado, USA (Warpinski, 1991).
dent. As the pressure exceeds the closure stress on the fis- Figure 9D-2 shows a fit of the Walsh model to data measured
sure (p > pfo), the entire fissure opens, yielding an accelerated during injection and drawdown conditions. The reservoir is
leakoff condition. Vuggy porosity, on the other hand, is gener- overpressured, giving a relatively high initial pressure (5400 psi)
ally insensitive to stress and remain unchanged until the pres- relative to the initial in-situ stress (~7000 psi). The initial reser-
sure exceeds the closure stress and opens the entire fissure voir permeability was measured using drawdown and buildup
(i.e., accelerated leakoff). tests when the interval was first completed. The permeability
The case of fissure opening when the pressure exceeds during injection conditions was measured using both injection
that closing the fissure is the simplest case to consider. Nolte and falloff behavior during nitrogen tracer testing to assess the
and Smith (1981) derived a relation for the critical pressure in interconnectability of the fissure system among nearby wells.
the hydraulic fracture for fissure opening to occur: The permeability collapse that occurred during full drawdown
was a common phenomenon in all intervals at the site. Inev-
itably, a hard drawdown would produce a reasonable rate of
No mineralization, Vuggy gas flow until the pressure reached a critical level at which the
rough surface porosity production would drop to an immeasurable level (but produc-
tion would be restored when the pressure built back up). On
the basis of the flow rates at the initial conditions, the perme-
ability was estimated to have dropped 2–3 orders of magni-
p = pi tude to achieve the observed decrease.

100
Injection
Permeabiltiy (normalized)

Measured reservoir permeability


during nitrogen injection tracer tests
10
pi < p < pfo Initial permeability
1
Estimated permeability
0.1 when production shuts
off from full drawdown
0.001
Initial In-situ
pressure stress
p > pfo 0.001
0 1000 2000 3000 4000 5000 6000 7000
Pore pressure (psi)

Figure 9D-2. Measured permeability variations of the


Figure 9D-1. The effects of pressure on fissure open- Mesaverde natural fissure system at the M-Site
ing and porosity. (Warpinski, 1991).

Reservoir Stimulation 9-23


9D. Fluid leakoff in natural fissures (continued)

The importance of Fig. 9D-2 is that it provides a mechanis-


tic model of understanding the behavior of fissured reservoirs 3.0
during and after stimulation. When the fracturing pressure is
low, the fissure permeability is near the initial value and the
leakoff coefficient is relatively low. As the fracturing pressure 2.5
increases, the leakoff coefficient increases rapidly, causing Pressure-sensitive fissures
problems with fracture size, slurry dehydration and screenout.
2.0
However, the situation during cleanup reverses because the

CL,fissure/CL
well is drawn down to extract stimulation fluids and the natural
fissure permeability decreases. Thus, stimulation fluids are 1.5
injected under wide-open fissure conditions but produced
under clamped fissure conditions, making it difficult to clean √∆p
up the reservoir. 1.0
The Walsh and other models can be incorporated into a
fluid-loss equation to represent changing leakoff conditions
0.5
(Warpinski, 1991). Figure 9D-3 shows the calculated pres-
sure-sensitive leakoff of a Mesaverde fissure system com-
pared with the normal pressure sensitivity of a conventional 0
pore space (recall that the filtrate leakoff coefficient is propor- 0 200 400 600 800 1000 1200 1400
tional to the square root of the fracturing pressure minus the
Net treatment pressure (psi)
reservoir pressure). Generally, the pressure sensitivity of con-
ventional reservoirs is ignored because the changes are
small; in this case, it varied from 1.0 to 1.4. However, the Figure 9D-3. Pressure-sensitive leakoff of a Mesa-
pressure sensitivity of the fissures greatly exceeded this
change, and the leakoff of the fissures reached about 3 times
verde fissure system.
the conventional leakoff. Even at low fracturing pressures, the
leakoff of the fissures was greater than conventional leakoff
decreasing slope with decreasing pressure. Fissure-sensitive
because the fracturing pressure was large relative to the
fluid loss can be recognized in the injection pressure behav-
reservoir pressure and the fissure pores had much less stress
ior, but it is difficult to identify because it looks much like
closing them. This example is based on a tight Mesaverde
height growth during the injection (i.e., a nearly constant pres-
coastal zone reservoir with a base leakoff coefficient CL of
sure derivative). Fissure dilation is usually followed by flatten-
0.0004–0.0006 ft/min1⁄2 (measured during pressure declines)
ing of the fracturing pressure, and screenout most likely
and a fracture-calibrated leakoff coefficient CL,fissure of
occurs relatively fast, depending on the injected proppant
0.0015–0.0019 ft/min1⁄2 during injections.
concentration (Fig. 9-16).
Pressure-sensitive fissure behavior is best recognized in
the pressure decline where the G-plot shows a continuously

9-4.6. Formation pressure capacity These three complicating mechanisms produce


excessive fluid loss from the main body of the frac-
As discussed previously, for a PKN-type fracture, a ture and potential treatment problems. They depend
constant pressure response can occur after a period on the stress differences along the three principal
of normal pressure increase resulting from controlled directions in the formation. The formation acts like
fracture height because the pressure a pressure vessel with a pressure capacity governed
• approaches the stress of a barrier and causes signif- by the pressure-limiting mechanism and the corre-
icant height growth into a lower stress zone (deter- sponding in-situ stress difference.
mined by the horizontal stress difference between When the net pressure reaches the formation pres-
the reservoir and adjacent barrier; Fig. 9-12) sure capacity, the fracturing process may exhibit one
• exceeds the overburden, and the initiation of a T- of the three growth patterns discussed in Section 9-4.5.
shaped fracture begins at a pressure slightly greater Each growth pattern makes additional fracture propa-
than the overburden (determined by the horizontal gation relatively inefficient. Under these conditions,
and vertical stress differences in the reservoir; an inaccurate diagnostic could lead to a design change
Fig. 9-14) that either compromises the effectiveness of the frac-
ture treatment or further aggravates the problem. For
• exceeds the normal stress acting on natural fissures
example, the pad volume or the fracturing fluid vis-
and causes them to open (determined by the stress
cosity may be increased to alleviate recurring screen-
difference between the two horizontal stresses in
outs. These approaches, however, could be ineffective
the reservoir; Fig. 9-16).
if the screenouts are not due to excessive fluid dehy-

9-24 Fracture Evaluation Using Pressure Diagnostics


dration or insufficient fracture width, respectively. of proppant convection and settling is not a limi-
An important function of calibration tests is to iden- tation with the commonly used polymer fluids
tify the formation pressure capacity and the associ- (Nolte, 1988a; Warpinski et al., 1998a; Shah and
ated complicating mechanism. The propped treatment Asadi, 1998). In addition, a reduction in the fluid
can then be planned to attain the desired fracture viscosity only nominally affects the treatment effi-
characteristics in a cost-effective manner. ciency. Therefore, a moderate decrease in the fluid
Numerous alternatives have been proposed to mod- viscosity should not significantly alter the propped
ify fracture treatments that are constrained by the for- fracture characteristics. Fracturing equipment is
mation pressure capacity. They range from the appli- capable of delivering a continuous decrease in fluid
cation of unconventional designs to reduce the frac- viscosity throughout the treatment (see Chapter 11).
turing pressure (e.g., Veatch and Crowell, 1982; Britt All fracture treatments benefit from a reduction in
et al., 1994) and the use of specialized additives (e.g., the polymer mass used during the treatment. This
Nolte, 1982; Nguyen and Larson, 1983; Warpinski, approach limits the gel residue in the proppant pack
1990) to even altering the reservoir stress state to prevent an excessive loss of fracture conduc-
through changes of either the reservoir pressure or tivity (see Chapter 8).
temperature (Cleary, 1980a). The particular choice of Figure 9-17a shows a pressure record during a
an alternative is driven primarily by the nature of the fracture treatment in a gas reservoir with micro-
problem and how quickly the formation pressure darcy permeability. The formation net pressure
capacity is reached during the propped treatment, capacity identified as 1180 psi is attributed to
as discussed in the following. uncontrolled height growth that occurred after
approximately 75% of the propped treatment had
• Pressure capacity approached during the
been injected. Fracture treatments on offset wells
late stage of the treatment
were successfully designed to maintain the pres-
The recommended approach when any of the three sure below the formation pressure capacity by con-
complicating mechanisms occurs is to engineer the tinuously reducing the fluid viscosity during the
propped treatment so that the fracturing pressure proppant treatment (Fig. 9-17b). The initially
remains below the formation pressure capacity. increasing fluid viscosity was required to over-
Limiting the fracturing pressure may require come enhanced degradation of the early treatment
reducing the treatment volume, injection rate or stage from longer exposure to the reservoir tem-
fluid viscosity. The pressure relation for the PKN perature (Nolte, 1982, 1988a).
fracture geometry in Eq. 9-25 provides an assess-
ment for determining which of these changes is
most effective at reducing the fracturing pressure. (a)
For example, using the typical fluid rheology 2000
Gel (lbm/1000 gal) 60
exponent n = 0.5 in Eq. 9-25 shows that a twofold
pnet capacity = 1180 psi
reduction in the injection rate reduces the net pres- 1000
Net pressure, pnet (psi)

sure by only 11%. A twofold reduction in the fluid


viscosity (or the fluid rheology coefficient K), 500
10 100 1000
however, provides a 22% reduction in the net pres-
sure. A reduction in the fluid viscosity thus will be (b)
2000
more effective in controlling the fracturing pres- Gel (lbm/1000 gal) 60 50 40 30
sure. Reducing the fluid viscosity, however, pnet capacity = 1180 psi
1000
requires additional considerations of the potentially
increased fluid loss, modifications to the breaker 500
and proppant scheduling, fluid degradation from 10 100 1000
exposure to the reservoir temperature and near- Time (min)
wellbore effects (Nolte, 1982, 1988a).
Viscosity requirements for efficient proppant Figure 9-17. Control of fracture height growth through vis-
transport are often overestimated because the effect cosity reduction (Nolte, 1982, 1988a).

Reservoir Stimulation 9-25


• Pressure capacity approached during the
1500
intermediate stage of the treatment

Net pressure, pnet (psi)


Fissure opening
The mitigating design for reaching the pressure pnet capacity = 1000 psi
1000
capacity during the intermediate stage of the treat-
ment depends on the specific mechanism associated Accelerated leakoff
with the formation pressure capacity. Relatively
fine solids such as silica flour or platelet-based
additives (Vinod et al., 1997) can control acceler- 10
ated fluid loss in formations where fissures provide 10 20 30 40 50 60
a meaningful increase in the overall permeability, Time (min)
even before they are open. The fine-size additive 1500

Net pressure, pnet (psi)


(e.g., 300-mesh particles) should be used through-
pnet capacity = 1000 psi
out the pad so that fluid leakoff into fissures, and 1000
hence the fluid pressure within them, is reduced
100-mesh sand
from the onset of the treatment.
Controlling accelerated leakoff can significantly
delay or even prevent the mechanical opening of
fissures during the propped treatment. The additive 10
10 20 30 40 50 60
will not bridge at the fracture tip because the far- Time (min)
thest it can travel is the boundary of the fracturing
fluid and the fluid lag region (see Fig. 5-16). The Figure 9-18. Control of accelerated leakoff into natural fis-
fracture width in this region is generally much sures using particulate-based additives (Warpinski, 1990).
larger than the extremely small diameter of the
fine-size additive.
growth (Mukherjee et al., 1994; Nolte, 1988a) or
A larger size solid additive (e.g., 100-mesh
the occurrence of T-shaped fractures. The additives
sand) should be used to control accelerated leakoff
bridge in the narrow pinch points associated with
resulting from mechanically opened fissures. This
both of these mechanisms and arrest the further
additive should be scheduled immediately prior to
growth of secondary features. The benefit of the
proppant addition. It will either bridge in the open
particle-size mixture described by Nolte (1988a)
fissure or block its opening, thereby impeding fluid
is exhibited in the field example in Fig. 9-19. The
entry and the associated increase in the rate of
calibration treatment shows that the critical net
fluid loss. The additive alone may be sufficient if
pressure of 800 psi was reached after approxi-
the fissures exhibit negligible permeability before
mately 20 min of fluid injection (Fig. 9-19a). The
they are mechanically opened. It can be used in
pressure subsequently declined, exhibiting a signa-
conjunction with the previously discussed finer
ture for uncontrolled height growth similar to that
300-mesh particles within the pad when fissures
in Fig. 9-12. Rapid height growth led to a prema-
are responsible for increased fluid loss before they
ture screenout soon after the introduction of prop-
are mechanically opened.
pant. A subsequent successfully placed propped
The effectiveness of 100-mesh sand in control-
treatment exhibited a higher net pressure once
ling fluid loss is demonstrated by the calibration
height growth activity was controlled after the
test shown in the field example in Fig. 9-18. Fis-
addition of a proppant mixture (Fig. 9-19b).
sure opening occurred at a net pressure of 1000 psi,
The use of solid-based additives requires caution
defining the pressure capacity of the formation.
and proper design procedures. Elbel et al. (1984)
The accelerated fluid loss was mitigated by the
showed that incorrect scheduling of the larger size
introduction of 100-mesh sand on a subsequent
additives (e.g., 100-mesh sand) within the pad
calibration test, which exhibited a net pressure well
could result in a premature screenout. Particulate-
in excess of the previous net pressure capacity of
based additives should not be scheduled in con-
1000 psi.
junction with proppant stages because their combi-
A 100-mesh sand or specialized proppant sys-
nation reduces proppant pack permeability. In
tems should be used to control excessive height

9-26 Fracture Evaluation Using Pressure Diagnostics


(a) horizontal stresses in the reservoir. Under the
Conventional Design altered state of stress, undesirable fracture growth
2000
10 MPa
patterns will occur at much higher net pressure
values. A subsequent, larger size fracture treatment
1000 can then be designed to obtain the desired fracture
Proppant
geometry with a fluid pressure that remains below
Net pressure (psi)

600
the magnitude triggering complex fracture growth.
400
5 10 20 50 100 200

(b) 9-4.7. Pressure response after a screenout


Height Control Design
1000 The pressure response in Fig. 9-1 is for a fracture that
Proppant mixture
600
5 MPa is free to extend until shut-in. A common occurrence
during proppant injection is a screenout. Screenouts
400
5 10 20 50 100 200
can occur in the vicinity of the wellbore when the
Time (min)
formation pressure capacity is reached (see Section
9-4.6) or at the fracture tip and around the complete
Figure 9-19. Control of height growth using a mixture of circumference of the fracture area (i.e., a tip screen-
particle sizes (Nolte, 1982, 1988a). out, or TSO) when the proppant slurry bridges
because of insufficient width or dehydrates as a result
general, only small concentrations of the additive of fluid loss. In addition to resulting from proppant
are sufficient to mitigate any of the complicating bridging, screenouts can also be caused by the exces-
mechanisms. Furthermore, even a moderate loss sive buildup of polymer filter cake at the tip or by
of fracture conductivity has a nominal effect on incorrect scheduling of particulate-based fluid-loss
poststimulation production in low-permeability control additives (Elbel et al., 1984). Either of these
reservoirs, where additive use is more likely. circumferential flow restrictions effectively stops
Fissured formations may benefit from the use of fracture propagation, and subsequent injection is
particulate additives because they effectively prop stored primarily by increasing the width. A screenout
the natural fissures and increase their permeability may be undesirable in a low-permeability reservoir
(Miller and Warembourg, 1975). where increasing the fracture length is the primary
objective, or it can be deliberately designed to occur
• Pressure capacity approached during the
to increase production in high-permeability environ-
early stage of the treatment
ments (see Chapter 5).
The approach of pressure capacity during the early As the fracture pressure and width increase during
stage of treatment presents the worst-case scenario the restricted growth period, the pressure gradient
for effective fracture stimulation. None of the mea- from fluid flow decreases. The value of ∆p–f approaches
sures discussed previously may be effective in that of pnet. The value of β approaches 1, as discussed
controlling the complicating mechanisms and their in Section 9-3.3 and illustrated on the log-log plot in
damaging consequences. Mitigating strategies Fig. 9-20. Figure 9-20a shows the variation of pnet
under these conditions should focus on altering the and ∆p–f, which defines the average width. Post-
stress state of the reservoir. Section 3-5.4 indicates screenout behavior is depicted in Fig. 9-20b. The
that any change in the reservoir pore pressure time when the wellbore pressure apparently responds
could change the horizontal stresses by 46% to to the restriction is tso /βp, where tso is the time when
80% of the pore pressure change. The stress state, the screenout occurs. From the values of βp deter-
therefore, is most practically reduced through an mined with Eq. 9-19, tso /βp is on the order of 1.5tso or
extended period of hydrocarbon production prior about 50% greater than at the time of injection before
to the fracture treatment. the restriction occurred. In contrast, Sidebar 9C
The best approach is to initially produce the shows that the pressure derivative is much more
reservoir for an extended period, possibly with a sensitive, with its change perceptible shortly after
much smaller fracture. The corresponding reduc- a screenout.
tion in the reservoir pore pressure will reduce the

Reservoir Stimulation 9-27


where ∆tDso is the dimensionless time after a screen-
(a) Unrestricted Extension out, defined as the ratio of the incremental time after
pnet the screenout and the screenout time, and ηso is the
efficiency when the screenout occurs.
βp < 1
Figure 9-20 shows that the log-log slope of the net
∆pf

Log (“averaged” net pressure, ∆pf)


pressure generally changes after a screenout. Its par-
ticular value at ∆tDso = 2 is shown in Fig. 9-21. An
Log (net pressure)

(b) Restricted Extension at Fracture Tip


important conclusion based on this figure is that the
postscreenout log-log slope approaches unity only
tso /βp
for a high treatment efficiency before the TSO. An
tso example showing the net pressure response in a low-
β≈1 permeability oil-bearing reservoir characterized by a
pnet >1 high fluid efficiency is presented in Fig. 9-22a. A unit
βp < 1
1 slope follows the TSO. In lower efficiency situations,
∆pf however, the fracture pressure exhibits a log-log
behavior with a slope that can be larger than 1.0. This
case is shown in Fig. 9-22b, where the TSO is char-
acterized by a log-log slope of 1.7. The pressure data
Log (time) are from a frac and pack treatment in a 1.5-darcy, oil-
bearing, unconsolidated reservoir that exhibited a low
Figure 9-20. Log-log relations of net pressure versus time treatment fluid efficiency.
for restricted and unrestricted extension. Proppant bridging between the wellbore and the tip
increases the rate of the pressure increase over that
The pressure response and fracture width relation described by Eq. 9-41. Successive packing of the
after the screenout can be derived using the material- fracture toward the wellbore also results in a progres-
balance relations from Section 9-3. For the assump- sively increasing log-log slope with continued slurry
tion of an ideal screenout occurring around the frac- injection (Martins et al., 1992c).
ture circumference, the fracture area remains constant
and equal to the area immediately before the screen-
out. The material-balance relations following a 2.5

screenout therefore are the same as those applied to


describe the pressure behavior following the end of
Log-log slope

2.0
pumping, as discussed in Section 9-5. The only dif- ∆tso
∆tDso = =2
tso
ference, however, is that although no additional fluid 1.5
volume is introduced during the decline, slurry injec-
tion continues after the onset of the screenout and
1.0
should be included in the fundamental relations pre- 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
sented in Section 9-3. Efficiency at screenout, ηso
Applying these modifications to the material-bal-
ance relations, it can be shown (see Appendix Eq. 51) Figure 9-21. Log-log slope after a TSO for ∆tDso = 2.
for the commonly used polymer fracturing fluids that
the log-log slope of the net pressure plot after a screen-
out and in the absence of spurt as β approaches 1 is
9-4.8. Fracture diagnostics from log-log
 π
− sin −1 (1 + ∆t Dso )
−1/ 2


plot slopes
2
π
t dpnet  2 − sin (1 + ∆t Dso ) − ∆t Dso (1 + ∆t Dso )
−1 −1/ 2 The previous sections provide conceptual analyses
= of the different types of log-log slopes of net pressure
pnet dt  ηso → 0 versus time. The slopes are characteristic of various

1.0 ηso → 1 , types of fracture geometries and modes of propaga-
 tion. Therefore, the log-log plot, its associated slopes
(9-41)

9-28 Fracture Evaluation Using Pressure Diagnostics


After the fracture is confined by barriers, the pres-
(a)
sure increases as predicted by the PKN model, with
1000 the range of the slope as given by Eq. 9-34. This
range is from about l⁄8 for low efficiency to about
Net pressure, pnet (psi)

1
⁄4 for high efficiency.
As the fracture pressure increases, it can reach the
1
pressure capacity of the formation. This leads to a
1 regulator effect, resulting in nearly constant pressure
because of accelerated fluid loss primarily near the
wellbore (see Section 9-4.5). A nearly constant pres-
100
sure measurement that is equal to the overburden
1 10 100 stress indicates a T-shaped fracture. Controlled frac-
Time (min) ture height growth into a barrier is characterized by
a gradually decreasing log-log net pressure slope and
(b) a constant pressure derivative. Net pressures steadily
Field Data
1000 decrease if uncontrolled fracture height growth
beyond a pinch point occurs. Fissure-dominated
fluid-loss behavior regulates the pressure to a con-
Net pressure, pnet (psi)

stant value when the fissures are mechanically


opened. If the fissure permeability increases before
100
the mechanical opening occurs, this response will be
1.7
preceded by a gradually decreasing log-log slope.
1 A significant pressure increase (i.e., log-log slope
between those suggested by Eq. 9-41) indicates
10 restricted extension or a screenout near the fracture
1 10 100
tip, whereas a significantly higher slope should be
Time (min)
expected for a restriction nearer the wellbore than the
Figure 9-22. Log-log plot of the postscreenout pressure
fracture tip (Nolte and Smith, 1981).
response for (a) low-permeability long fracture and (b) The types of slopes and associated interpretations
high-permeability short fracture. for vertical fractures are listed in Table 9-3. This
table, in conjunction with the interpretation plot
and the pressure derivative provide a diagnostic tool
Table 9-3. Interpretation of log-log plot
for interpreting the fracturing process. The analyses fracture pressure slopes.
presume that the pressure measurement represents the
actual fracturing behavior, corrected for near-well- Propagation Log-Log Slope Interpretation
Type
bore effects (see Section 9-4.9). Also, the pump rate
and fluid properties are assumed to be relatively con- Ia –1⁄6 to –1⁄5 KGD (Eq. 9-36)
stant during the treatment. Therefore, for a correct Ib –1⁄8 to –1⁄5 Radial (Eq. 9-37)
interpretation of the fracturing pressure, it is impor-
II ⁄6 to 1⁄4
1
PKN (Eq. 9-35)
tant to note variations in the pressure response that
occur when proppant is first injected into the forma- III Reduced from II Controlled height growth
Stress-sensitive fissure
tion or with significant changes in the pump rate or
fluid rheology (e.g., at about 17 min on Fig. 9-18). IV 0 Height growth through
pinch point
The basic net pressure interpretation includes the Fissure dilation
initially decreasing pressure before the fracture is T-shaped fracture
influenced by barriers (Fig. 9-6). This time is gener- V ≥1 Restricted extension
ally short, particularly for zones of relatively small
VI Negative following IV Uncontrolled height
height. For this initial behavior, the theoretical range growth
of the slope is given by the exponents in Eqs. 9-36 Note: n = 0.5
and 9-37.

Reservoir Stimulation 9-29


in Fig. 9-23, shows that the log-log plot with its char- contribution and the fracturing-related pressure
acteristic slopes provides a diagnostic tool analogous response are additive:
to the log-log plot for identifying flow regimes within
pmeas = pc + pnet + ∆pnear wellbore , (9-42)
a reservoir.
where pmeas is the measured bottomhole pressure and
∆pnear wellbore is the cumulative near-wellbore pressure
loss. The result of the near-wellbore effects is an ele-
vated pressure measurement during fluid injection
Log net pressure

and potential misinterpretation when the techniques


V described in the previous sections are used.
III IV
Insufficient perforations could result from improper
VI
perforating practices, poor perforation cleanup or
II ineffective formation breakdown procedures. It has
Ia, Ib
been experimentally shown (Crump and Conway,
1988) that the perforation diameter changes only
Log time
when proppant-laden slurry enters the perforation.
Figure 9-23. Log-log interpretation plot for various fracture For a constant injection rate, perforation friction is
propagation modes. reflected as a constant increase in the treatment pres-
sure during a calibration test and the pad stage of the
main fracture treatment. The introduction of proppant
erodes the perforations, increasing their diameter.
9-4.9. Near-wellbore effects
This results in a progressive decrease in the measured
The previous discussion of log-log plot diagnostics treatment pressure during proppant injection until the
assumes that the bottomhole pressure measurement pressure drop across the perforations is negligible, as
reflects the actual fracture response. The injected shown from 25 to 35 min on Fig. 9-24.
fluid may experience pressure loss prior to entering An apparently higher net pressure with a reduced
the main body of the fracture. Near-wellbore pressure log-log slope is indicated on net pressure plots that are
loss could occur as a result of restrictions arising not corrected for the presence of perforation friction.
from the well completion (e.g., perforations; see The bottomhole pressure measurement, however, will
Section 11-3) or the near-wellbore fracture geometry exhibit an instantaneous change during shut-in if sig-
(e.g., a convoluted fracture pathway; Cleary et al., nificant perforation friction is encountered. The
1993). The latter cause can be particularly damaging, instantaneous pressure change provides a measure of
because it reduces the fracture width close to the the excess pressure associated with perforation friction
wellbore and increases the likelihood of a premature and, in conjunction with Eq. 6-120, is used to esti-
screenout. Near-wellbore pressure loss is rate depen- mate the number of perforations accepting fluid.
dent and affects the pressure response exclusively Fracture tortuosity, or the existence of a convoluted
during the fluid injection. The occurrence of near- pathway from the perforations to the main fracture
wellbore problems thus masks the actual fracture body, causes the fracture to open against a normal
response and complicates the interpretation of injec- stress that is higher than the formation closure pres-
tion pressure. sure. Consequently, the fracture width is narrower in
Chapters 6 and 11 identify three mechanisms to the near-wellbore region, which restricts fluid flow
explain near-wellbore pressure loss: insufficient per- and increases the bottomhole pressure. The associated
forations, a convoluted fracture pathway in the vicin- pressure drop in the tortuous region can be approxi-
ity of the wellbore (or fracture tortuosity) and a mated using the methodology outlined in Section
hydraulic annulus between the cement and the rock 6-8.4. Because the fracture width is smallest at the
that connects the perforations with the main body of beginning of a treatment, the fracture reorientation
the fracture. The evolving fracture geometry, away pressure loss is largest during the early stage of the
from the wellbore, is completely independent of the treatment. As fluid injection proceeds, the pressure
near-wellbore effects. It is assumed that their pressure drop associated with fracture reorientation progres-

9-30 Fracture Evaluation Using Pressure Diagnostics


sively decreases. Figure 9-25, developed using than 10°; Behrmann and Elbel, 1991), the fracture
numerical simulation, illustrates this response for var- may not initiate at the perforations. Rather, the fluid
ious degrees of fracture turning. communicates with the fracture through a narrow
If the orientation of the perforation with the plane annulus around the casing (Fig. 9-26) (see Section
of the hydraulic fracture is large (typically greater 11-3 and Sidebar 11C). As discussed in Section 6-8,

4100
End of injection
Fluid injection 4000 Proppant
at perforations

Bottomhole pressure (psi)


3900

Casing 3800
3700
3600
Proppant injection
3500
3400
3300
0 50 100 150 200
Treatment time (min)

Figure 9-24. Perforation friction pressure response.

4100

4000
End of injection
Bottomhole pressure (psi)

3900
15° reorientation
Fracture 3800
Wellbore plane
at infinity 3700

3600 5° reorientation

3500
Fracture
reorientation 3400 Aligned fracture

3300
0 10 20 30 40 50
Treatment time (min)

Figure 9-25. Fracture reorientation pressure response.

Restriction area
Wellbore

Fracture Fracture

Channel to
fracture wings
Perforation

Figure 9-26. Pinch point caused by the rock–casing annulus.

Reservoir Stimulation 9-31


the formation elastic response (i.e., Poisson’s ratio Fracture pinch effects resulting from fracture reori-
effect) results in a pinch point at the intersection of entation or a hydraulic annulus can be surmounted
the channel with the fracture body. This annulus can during fluid injection. Consequently, an instantaneous
cause higher treating pressure because of the width pressure drop at shut-in, as seen for perforation fric-
restriction, as in the case of fracture reorientation. tion, may not be observed when either of these
However, unlike for fracture reorientation, the effect effects is present. An additional contrast to perfora-
of the annulus can increase with increasing fracture tion friction is that the associated pressure loss
width because of the Poisson’s ratio effect (see Fig. depends on the fracture width and hence exhibits
6-19). The annulus effect decreases with erosion of continuous variation during injection, rather than
the pinch point by fluid in soft rocks or following the being a discrete, time-related event. The step-down
injection of proppant in hard rocks. Section 6-8.5 dis- test, as discussed in Sidebar 9E, can be applied to
cusses the near-wellbore pressure loss associated with develop a more objective prediction of the magnitude
perforation phasing misalignment. of and the mechanism governing near-wellbore pres-
sure loss.

9E. Rate step-down test analysis—a diagnostic ∆p pf ≅ k pf q i2 , (9E-1)


for fracture entry
where the proportionality constant kpf is determined in oilfield
Chris Wright, Pinnacle Technologies units by the fluid density ρ, perforation diameter Dp, number of
perforations N and discharge coefficient C:
The step-down test identifies and quantifies near-wellbore tor-
tuosity. It also quantifies perforation effectiveness and can pro- ρ
vide a rough estimate of the number of perforations accepting k pf = 0.237 . (9E-2)
N 2D p4C 2
fluid. With the step-down test, on-site evaluation can be made
of potential problems resulting from near-wellbore fracture tor- In contrast, ∆pnear wellbore is roughly proportional to the injection
tuosity and perforating. Careful analysis of near-wellbore pres- rate raised to an exponent that is less than unity because it is
sure losses, or the fracture entry friction, enables identifying due to laminar flow through a narrow channel in the pressure-
the problem for resolution and subsequently evaluating the sensitive near-wellbore region:
effectiveness of the remedial measures employed.
Rate step-down test concept ∆p near wellbore ≅ k near wellbore q iβ , (9E-3)
Fracture entry friction (Fig. 9E-1) typically is the combination
of the perforation friction ∆ppf and the tortuosity or near-wellbore where knear wellbore is a proportionality constant and the power
friction ∆pnear wellbore. Perforation friction is simply dissipation of law exponent β is between 0.25 and 1.0, with a value of 0.5
the kinetic energy imparted on the fluid as it flows through a appropriate for most engineering applications. This marked
small orifice at high velocity. Perforation friction, therefore, is difference in the injection rate dependence of the two compo-
proportional to the injection rate qi squared times a proportion- nents of fracture entry friction clearly distinguishes them in
ality constant: a rate step-down test.

Test procedure and analysis


1. Measure the surface pressure and slurry rate at a sampling
Near-Wellbore Friction Perforation Friction rate of every 1–3 s. In general, the measurement of the bot-
tomhole pressure assists the analysis by avoiding uncer-
tainties regarding the hydrostatic head (e.g., in foam treat-
Narrow pressure-
dependent opening
ments) or large values of wellbore friction (e.g., high-rate
tubing treatments).
2. After a breakdown injection or calibration treatment, reduce
Wide far-field fracture the injection rate in steps of 1⁄5 to 1⁄3 of the full rate until the
rate becomes zero. Hold the rate constant at each step to
High-velocity kinetic obtain a stabilized pressure measurement for about 15 to
energy dissipation 20 s. The exact rate at each step is not important; rather,
the goal is rapid change between the constant rates. The
easiest way to achieve this is to take one or two pumps off
Cased line at a time. Record the stabilized surface pressure at
borehole each rate step for input to the step-down test analysis.
with
perforations 3. Determine the change in the bottomhole pressure for each
change in the injection rate. Then plot the fracture entry fric-
tion ∆pentry versus the injection rate. Fit the points obtained
from the rate step-down test with two functions using

Figure 9E-1. Near-wellbore friction and perforation fric-


( ) ( )
∆p entry (q i ) = k pf × q i2 + k near wellbore × q i1/ 2 . (9E-4)
tion exhibit different dependencies on the injection rate.

9-32 Fracture Evaluation Using Pressure Diagnostics


9E. Rate step-down test analysis—a diagnostic for fracture entry (continued)

A dual-curve-fit algorithm can be used to determine the best- result, several proppant slugs were planned for as early
fit values of kpf and knear wellbore, thus defining the respective as possible during the pad period of the propped treatment.
values of ∆ppf and ∆pnear wellbore for any given rate (Fig. 9E-2). Before the slugs reached the perforations, the tortuosity
continued to increase, which limited the injection rate so
Limitations using surface pressure measurements that the surface pressure could be maintained below the
Rate step-down analysis using the surface treating pressure acceptable value of 6000 psi. However, when the proppant
can be difficult if the tubular friction is large compared with slugs arrived at the perforations, the tortuosity was signifi-
∆pentry because the wellbore friction can vary unpredictably cantly reduced, which enabled increasing the injection rate.
from published or expected values. The friction of water can The propped fracture treatment was successfully placed
be significantly reduced by small amounts of gel contamination with 6-ppg maximum proppant loading.
in the wellbore fluids or from a gel hydration unit. Fluid friction • Poor perforation effectiveness and excessive pad volume
with crosslinked gel may vary with small variations in the fluid
composition. Foam friction behavior is extremely unpredictable The rate step-down test in Fig. 9E-4 was performed after
and variable, and the analysis is further complicated by the first KCl injection. The test clearly shows the dominance
changes in hydrostatic pressure. With turbulent flow (i.e., of ∆ppf (i.e., near-wellbore pressure losses that relate to the
essentially all water injections), the friction is functionally closer injection rate to the 1.94 exponent). The estimated value of
to ∆ppf; with laminar flow (i.e., low injection rates with viscosi- ∆ppf was about 4500 psi at 18 bbl/min, the equivalent of
fied fluids), the friction is functionally closer to ∆pnear wellbore. only 4 of 60 holes open. This condition would not allow fluid
injection at the planned rate of 30 bbl/min. The near-well-
Field example
bore fracture tortuosity was low, at about 50 psi at 18 bbl/min.
The following example illustrates the usefulness of rate step- Additional KCl breakdown injections and surging did not
down test analysis resulting from the ability to understand and improve the low injectivity.
apply the information contained in fracture pressure behavior. The well was reperforated with larger holes, which
reduced ∆ppf to 1500 psi at 18 bbl/min. Following a
• Severe near-wellbore fracture tortuosity crosslinked gel calibration treatment (including a 20-bbl,
The rate step-down test following a second injection of 4-ppg proppant slug), the equivalent of 20 holes was open,
potassium chloride (KCl) water showed that the near-well- which provided an acceptable value of ∆ppf of 500 psi at
bore fracture tortuosity in a naturally fractured dolomite for- 30 bbl/min. The value of ∆pnear wellbore remained at less than
mation was extremely high, at 1900 psi (Fig. 9E-3). As a 50 psi, with C equal to approximately 0.6 for fluid injection.

High Near-Wellbore Friction:


High Premature Screenout Potential
Rate and bottomhole pressure

∆p1
∆p1 Total entry friction
Fracture entry friction

∆p2
∆p2

∆p3 ∆p3
∆pnear wellbore ≅ qi1⁄2
∆qi1
∆qi2 ∆p4
ISIP ∆p4
∆qi3
∆qi4 ∆ppf ≅ qi2

Time ∆qi4 ∆qi3 ∆qi2 ∆qi1


Rate
High Perforation Friction:
Only a Small Number of Perforations
Effectively Taking Fluid
Rate and bottomhole pressure

∆p1
Fracture entry friction

Total entry friction


∆p1

∆p2
∆p2
∆qi1 ∆p3
∆qi2 ∆p3 ∆ppf ≅ qi2
∆p4
∆qi3 ISIP
∆p4
∆qi4 ∆pnear wellbore ≅ qi1⁄2

Time ∆qi4 ∆qi3 ∆qi2 ∆qi1


Rate

Figure 9E-2. The step-down test is conducted to measure ∆ppf and ∆pnear wellbore. ISIP = instantaneous shut-in pressure.

Reservoir Stimulation 9-33


9E. Rate step-down test analysis—a diagnostic for fracture entry (continued)

Conclusions
Rate step-down tests are simple to implement and can provide
key insights into the nature of the near-wellbore connection
between the wellbore and the far-field hydraulic fracture.
Although rate step-down tests have limitations, they can pro-
vide the rare combination of critical information at minimal
additional cost.

6000 20
1300-psi reduction Surface pressure limitation = 6000 psi
(first slug)
Injection rate (bbl/min)

Injection rate (bbl/min) and bottomhole proppant (ppg)


Surface pressure (psi)
Bottomhole proppant (ppg)
4800 16
Step-down 1: 1900-psi
tortuosity and small
perforation friction
Surface pressure (psi)

3600 12
Step-down 2: Zero tortuosity at
300-psi tortuosity end of pumping

2400 8
Increased maximum
proppant loading
from 4 to 6 ppg

1200 4

0 0
0 30 60 90 120 150
Time (min)

Figure 9E-3. Propped treatment example with severe near-wellbore friction (i.e., fracture tortuosity) that was mitigated
by pumping two proppant slugs.

9-5. Analysis during fracture closure ties as those observed with the interpretation of con-
ventional well test data.
Fracture behavior during shut-in and prior to closure Basic decline analysis, outlined initially in this sec-
is governed by the fluid-loss characteristics and the tion, follows derivations presented by Nolte (1979,
material-balance relation (Fig. 9-2). A mathematical 1986b). A generalization of the technique using ana-
description of the pressure during the fracture closing lytical extensions to address nonideal conditions
period can be developed by also incorporating the (Nolte et al., 1993) is also presented.
fracture compliance relation (Eq. 9-21). These two
relations and that describing fluid loss are combined
to develop the specialized G-plot, which describes the 9-5.1. Fluid efficiency
pressure response during shut-in. Application of the
“Mathematical relations for fluid loss” in the
G-plot is analogous to the Horner analysis used for
Appendix to this chapter derives the fundamental
conventional well tests. The selection of an applica-
relations for fluid loss at the end of pumping VLp
ble slope for the G-plot also has the same uncertain-
(Appendix Eq. 23) and during the subsequent shut-in

9-34 Fracture Evaluation Using Pressure Diagnostics


9E. Rate step-down test analysis—a diagnostic for fracture entry (continued)

(a)
12,000 50 6000 600
Injection rate (bbl/min) Perforation friction (psi)

Near-wellbore friction (psi)


Bottomhole pressure (psi)

Perforation friction (psi)


Bottomhole pressure (psi) Near-wellbore friction (psi)

Injection rate (bbl/min)


10,000 40 4800 480

8000 30 3600 360


Entry friction
6000 20 2400 240
Points used for step-down
analysis (~end of each rate step)
4000 10 1200 120

2000 0 0 0
5.0 5.5 6.0 6.5 7.0 7.5 0 4 8 12 16 20
Time (min) Pumping rate (bbl/min)
(b)
50 10,000 100
Surface pressure limitation = 10,000 psi
Step-downs 1 and 2 show severe Proppant loading (ppg)
perforation friction (only 4 out of 60 Surface pressure (psi)
perforations open) and low tortuosity Slurry rate (bbl/min)
40 8000 80
Step-downs 3 and 4 show
lower perforation friction and
small tortuosity; No reaction
Proppant loading (ppg)

Surface pressure (psi)

on proppant slug

Slurry rate (bbl/min)


30 6000 60

20 4000 40
Cut pad percentage
from 50% to 25%
Reperforate
10 2000 20

0 0 0
0 60 120 180 240 300
Time (min)

Figure 9E-4. The step-down test (a) diagnosed high ∆ppf that was remedied by reperforating (b).

period VLs (Appendix Eq. 27). These equations are are not routinely used during field practice. Simple
in terms of the dimensionless volume-loss function analytical approximations, however, can be derived
g(∆tD) and its value at shut-in g0. In this context, ∆tD for certain values of α. These values of α are valid for
is referred to as the dimensionless time and is defined the commonly used crosslinked fluids that develop
in Appendix Eq. 15 as the ratio of the shut-in time ∆t a polymer filter cake along the fracture walls. These
to the injection (or pumping) time tp. fluids are the focus of the remainder of this section.
The general expressions for g(∆tD) and g0 in Corresponding relations for the non-wall-building
Appendix Eqs. 17 and 20, respectively, are based on fluids, such as linear gels or viscoelastic surfactant–
the assumption of a monotonically increasing fracture based fluids, are also discussed in “Mathematical rela-
area that is defined by a power law expression with tions for fluid loss” in the Appendix to this chapter.
an exponent α (Appendix Eq. 2). The generalized The value of the area exponent α can be explicitly
expressions are relatively complicated and therefore determined for two limiting cases of fracture growth.

Reservoir Stimulation 9-35


The lower bound α0 is for the case in which fracture
5
behavior is dominated by fluid leakoff (η → 0) and Bounds
the fracture area grows as the square root of time. Upper (α1 = 1)
Lower (α0 = 1/2)
Treatments in which fluid leakoff is negligible (η → 4
1) represent the upper bound of the area exponent α1,
4/3[(1 + ∆tD)3/2 – ∆tD3/2]

g(∆tD)
and in this case the area increases approximately lin- 3
early with time. (1 + ∆tD)sin (1 + ∆tD)–1/2 + ∆tD1/2
–1

“Mathematical relations for fluid loss” in the 2


Appendix to this chapter provides expressions for α0 π/2
and α1. The upper bound α1 is less than 1. It depends 4/3
1
on the fluid rheology exponent n and is given by 0 0.05 0.1 0.2 0.5 1.0 2.0 5.0
Appendix Eq. 40 for the three fracture geometry Dimensionless time, ∆tD = ∆t/tp
models. The lower bound α0 is independent of the
fracture geometry and is defined by the fluid filtrate Figure 9-27. Bounding values of the dimensionless vol-
rheology (Appendix Eq. 41). For wall-building fluids, ume function g(∆tD) for fracture closure (Nolte, 1986a).
the following bounding values are used to provide
simple analytical expressions: implies that an approximation of g(∆tD) based on a
1 suitable value for α, such as that suggested by Eq. 9-44,
< α < 1. (9-43) should be sufficient for field applications. The
2
approximation is best developed by interpolation
Generally the fluid efficiency varies throughout the using its bounding values and a particular value of α:
treatment. The efficiency is at its maximum value
g( ∆t D , α ) = g ∆t D , α =  + η(2α1 − 1)
during the early stage of fracture propagation, and it 1
gradually reduces with fracture propagation because  2
of the increasing surface area available for fluid
 1 
leakoff. As a result, the fracture area extends with a × g( ∆t D , α = 1) − g ∆t D , α =  , (9-45)
decreasing exponent. The time variation of the expo-   2 
nent affects the fluid loss because the rate of fluid loss where the values of g(∆tD) at α = 1⁄2 and α = 1 are
at any time depends on the prior history of the area from Appendix Eq. 29.
evolution. However, the effect is small, and to main- The efficiency of a calibration treatment can be
tain tractable fluid-loss expressions, it is assumed that derived from the relation that the fracture volume at
the exponent remains essentially constant and is the end of injection equals the total volume of fluid
defined by the value of efficiency at the end of injec- lost during shut-in (Eq. 9-8):
tion. This value of the exponent can be found by inter-
polation between its lower bound α0 of 1⁄2 and upper Vfp VLs ( ∆tcD )
bound α1: η= = . (9-46)
Vi VLp + VLs ( ∆tcD )

α = α 0 + η(α1 − α 0 ) = + η α1 −  ,
1 1 Substituting Appendix Eqs. 22 and 26 into Eq. 9-46
(9-44)
2  2 produces the following expression for the efficiency:
where α1 is from Appendix Eq. 40. For typical effi- g( ∆tcD ) − g0
ciency values of 0.3 < η < 0.6, α is equal to approxi- η= , (9-47)
mately 0.6. g( ∆tcD ) + (κ − 1)g0
A much simplified expression for the fluid-loss where the spurt factor κ is defined in Appendix Eq. 24.
volume function g(∆tD) and its initial value g0 can κ denotes the ratio of fluid loss for a case with spurt
then be developed by using the bounding values of to that without spurt. For the propagation period, κ =
α (Eq. 9-43). These are given by Appendix Eqs. 29 1 for no spurt; more generally,
and 30, respectively, and graphically presented in
κ −1 spurt volume loss
Fig. 9-27. Throughout this development, the differ- = . (9-48)
ence between the upper and lower bounds of g(∆tD) is κ total leakoff volume loss
nominal, as illustrated on the figure. This observation

9-36 Fracture Evaluation Using Pressure Diagnostics


Figure 9-28 shows the relation between the treat-
ment efficiency η and the dimensionless shut-in time 1.0
κ =1
when the fracture closes ∆tcD for no spurt (i.e., κ = 1). Bounds
0.8 Upper (α1 = 1)
The plot was generated in terms of the two bounding Lower (α0 = 1⁄ 2)
values of α from Eq. 9-43. The equation and figure

Fluid efficiency, η
are independent of the geometry model and enable 0.6
defining the efficiency for proppant scheduling (see
Sidebar 6L). 0.4
The fracture penetration is determined using the
following equation, which is obtained by combining 0.2
Eq. 9-6 and Appendix Eq. 22:

Af =
(1 − η)Vi , (9-49)
0
0 0.05 0.1 0.2 0.5 1.0 2.0 5.0 10.0
2 rp κCL t p g0 Dimensionless closure time, ∆tcD

where the fracture surface area Af for the three basic Figure 9-28. Relationship between efficiency and closure
models is from Eq. 9-29, rp is the ratio of permeable time (Nolte, 1986a).
(or fluid-loss) area to fracture area, and CL is the
fluid-loss coefficient. The average fracture width
is then obtained by using Eqs. 9-4 and 9-49: 9-5.2. Basic pressure decline analysis
2 ηrp κCL t p g0 The previous section presents relations for the fracture
V
w = fp = . (9-50) geometry parameters from simple considerations of
Af 1− η material balance and the assumption of power-law-
Finally, the maximum fracture width immediately based fracture area growth. The basic decline analysis
after shut-in for the three basic models is obtained as assumes that the end of injection marks the termina-
tion of additional fracture extension and that the
4 ( πβ s ) PKN change in the fracture volume during shut-in is attrib-

wmax,si = w 4 π KGD (9-51) uted entirely to the change in the average fracture
3 2 Radial, width during this period. The latter assumption also
 implies that the fracture area is invariant throughout
where 〈w– 〉 is from Eq. 9-50 and β is from Eq. 9-20.
s the shut-in period. The fracture geometry models
The maximum width at the end of pumping is relate the fracture width to the net pressure through
obtained similarly: their compliance. Consequently, the combination of
the compliance and fracture geometry equations
β s β p PKN
 enables determination of the fluid-leakoff coefficient
wmax,p = wmax,si ≈ 1 KGD (9-52) on the basis of the rate of pressure decline during
≈ 1
 Radial. shut-in.
Using the material-balance relation during shut-in
The values for the PKN fracture geometry model from Eq. 9-7, it follows for a constant area that
are from Nolte (1979), whereas those for the KGD
and radial models assume an elliptical width profile dVf ( ∆t ) d w
− = − Af = qL . (9-53)
(Eqs. 9-11 and 9-13, respectively) and no change in d∆t d∆t
the fracture volume immediately before and after For an assumed constant fracture compliance that
shut-in. The value of wmax,p for the radial model from is ensured by a constant area, differentiation of
Eq. 9-52 is valid only for a line-source fluid entry Eq. 9-21 and substitution in Eq. 9-53, with the
condition (Fig. 9-6). The high entry velocity and expression for the fluid leakoff rate qL from Appendix
pressure gradient during pumping for the point- Eq. 11, results in
source case produce a nonelliptical width profile
dpnet 2 rp CL A f
(Geertsma and de Klerk, 1969). − Af c f = f ( ∆t D ) , (9-54)
d∆t tp

Reservoir Stimulation 9-37


where the fluid-loss rate function f(∆tD) is from
L
(
( )
1 2 h 2f PKN
Appendix Eq. 16 and equals the derivative of g(∆tD). 2 − η) ′ 
Equation 9-54 can be integrated between ∆tD = 0 and L =
1 V i E
( )
2 g0 κβ s p * 
1 4h f KGD (9-59)
3
∆tD (assuming a constant value of pc): R  32 3π 3
Radial.

πrp CL t p
pws − pw ( ∆t D ) = G( ∆t D ) ,
This equation provides the appropriate value of pene-
(9-55)
2c f tration for determining CL from Eq. 9-58.
The fluid efficiency can be obtained by substituting
where pws is the bottomhole pressure at shut-in. The the definition of the G-function from Eq. 9-56 into
function G(∆tD) was introduced by Nolte (1979): Eq. 9-47 with the approximation g0 = π/2 for α = 1⁄2:
G( ∆t D ) =
4
π
[
g( ∆t D ) − g0 . (9-56) ] η≈
Gc
, (9-60)
2κ + Gc
This function shares the same bounds for α as g
(Eq. 9-45). where Gc is the value of G(∆tD) at closure (i.e., ∆tD
Castillo (1987) recognized that under ideal condi- = ∆tcD). The approximation provided by Eq. 9-60 is
tions, Eq. 9-55 linearly relates the pressure versus the exact at η → 0.
G-function defined in Eq. 9-56 (Fig. 9-29) with a Martins and Harper (1985) extended the applica-
negative slope p*: tion of pressure decline analysis to the case of
expanding confocal ellipses. This propagation model
πrp CL t p is relevant to the early stage of propagation, before
p* = . (9-57)
the radial model is applicable, and includes the KGD
2c f
and radial models as limiting cases. Its consideration
Combining Eq. 9-57 with expressions for the frac- requires including the appropriate definition of the
ture compliance cf from Eq. 9-22 yields the following fracture compliance cf in Eq. 9-57.
equation for determining the fluid-leakoff coefficient:
h f PKN
p * βs  9-5.3. Decline analysis during
CL = 2 L KGD (9-58) nonideal conditions
rp t p E ′ 
(32 3π ) R
2
Radial.
The basic pressure decline analysis (see Section 9-5.2)
This relation provides a direct solution for the PKN implicitly assumes that
model (assuming that hf is known). For the other
• Fluid loss is based on Carter’s (1957) formulation
models, the penetration is required and must be deter-
of the square root of exposure time (see “Mathe-
mined first. This is achieved by substituting CL from
matical relations for fluid loss” in the Appendix to
Eq. 9-58 into Eq. 9-49:
this chapter) and is characterized by a constant
leakoff coefficient that is independent of pressure.
• Fracture area evolution with time is described by
Pressure
Derivative
a power law area relation during injection.
Derivative, dp/dG (psi)

• Fracture area and compliance are constant during


the closing phase.
pw (psi)

Slope, p*
• Fracturing fluid is incompressible.
pnet,si
• Formation closure pressure is constant.
pc All these assumptions are seldom met in routine
G(∆tcD) field practice. A departure from any of them produces
a G-plot with a continuous curve (i.e., not a straight
G(∆tD) line with a constant slope). In such cases, the applica-
tion of basic pressure decline analysis generally pre-
Figure 9-29. G-plot of the G-function response approxi- dicts optimistic estimates of the fluid-leakoff coeffi-
mating idealized fracture propagation conditions. cient and treatment efficiency. However, a rigorous

9-38 Fracture Evaluation Using Pressure Diagnostics


characterization of the deviations introduced to the – Change in fracture penetration
analysis when its assumptions are violated results in Following shut-in, the initial extension of the
an overly complex analysis. Fortunately, most of the fracture penetration increases the area exposed
deviations from nonideal conditions are amenable to to fluid loss. The correspondingly increased
relatively straightforward analytical modifications. fluid-loss rate relative to the case of constant
Corrections to the basic pressure decline analysis area results in a correspondingly steeper initial
are presented in this section. Change in the fracture slope of the G-plot. The subsequently reduced
area during shut-in and a pressure-dependent fluid- rate of pressure decline during recession results
loss coefficient are discussed because of their com- from a decrease in the fracture area and elimina-
mon occurrence. Considerations for fluid compress- tion of the region of relatively higher fluid loss
ibility, thermal effects and varying formation closure near the fracture tip. The consequences of frac-
stress (i.e., poroelasticity) were presented by Nolte ture penetration change during shut-in are a
et al. (1993) and are not repeated here. Their effects concave-upward profile on the G-plot (Fig. 9-31a)
were found to be minimal when the procedure out- and a continuous change in its slope mG.
lined in Section 9-5.4 was applied for the decline The corrected slope p* that defines the fluid-
analysis. loss coefficient based on Eqs. 9-54 and 9-58
• Fracture geometry change during shut-in represents an instantaneously unchanged prod-
uct of compliance (e.g., dependent on βs) and
Both the fracture length and height can change
area (i.e., d(cfAf)/dt = 0). For the three fracture
during the shut-in period. The fracture penetration
geometry models, this condition occurs near the
initially increases before eventually receding back
transition between extension and recession
toward the wellbore. Continued fracture extension
(Fig. 9-30). Extensive numerical simulation
occurs because of the redistribution of stored vol-
shows that this condition occurs when the well-
ume from the larger width region near the wellbore
bore net pressure reaches about three-quarters
to the fracture tip region. Simultaneously, the
of its value at the shut-in net pressure pnet,si:
height recedes from any higher stress barriers
because of the reducing fluid pressure (Fig. 9-30). pnet 3
= . (9-61)
This variation in fracture geometry changes the pnet ,si 4
character of the otherwise straight-line G-plot and
calls for additional considerations in using basic Thus, the slope of the G-plot can be desig-
pressure decline analysis to estimate the fluid-loss nated as m3⁄4 and is evaluated at the 3⁄4 point.
parameters. This slope is used in conjunction with Eq. 9-58
to eliminate the effect of fracture penetration
changes during shut-in. The validity of selecting
the slope at the 3⁄4 point has been experimentally
Net pressure
established by de Pater et al. (1996).
fracture penetration and height
Dimensionless net pressure,

– Fracture height recession


Fracture Fracture height growth into higher stress bound-
penetration, L ing zones results in an increase in the fracture
compliance, as shown in Fig. 9-7. Equation 9-57
also indicates that the rate of pressure decline is
Height inversely proportional to the average fracture
growth, h
compliance. Therefore, substantial height
tp
growth during injection leads to a decreased
compliance during the initial shut-in period and
Dimensionless time, tD
reduces the rate of pressure decline until the
height recedes from the bounding zones. Figure
Figure 9-30. Time variation of the net pressure and frac-
ture geometry for the PKN fracture geometry.
9-31b depicts this initial period of reduced slope
on the G-plot.

Reservoir Stimulation 9-39


(a) (b)

Length extension

Length recession
pnet

pnet
pnet,si pnet,si
Height recession

Length recession
pnet ≈ 3/4pnet,si
pnet ≈ 0.4∆σ

Ideal response Ideal response

G(∆tD) G(∆tD)

Figure 9-31. Conceptual G-plot response for nonideal fracture behavior during shut-in for (a) fracture extension and (b)
height growth.

Complete height recession from the bounding Height growth into a higher stress barrier
zones occurs when the wellbore net pressure requires a previously increasing net pressure
equals about 0.4 times the stress difference ∆σ. response. It therefore is not consistent with the
Also from Fig. 9-7, the net pressure at shut-in is basic requirements of the radial and KGD mod-
about 0.8∆σ when significant height growth els, both of which exhibit decreasing net pres-
occurs. Therefore, complete height recession sures. Height growth should be anticipated only
from the bounding zones occurs when the well- for PKN-type behavior, with an increasing and
bore net pressure reduces to a value that is relatively large net pressure during injection. As
approximately one-half of that at shut-in. a result, for the PKN model with significant
Following fracture withdrawal into the primary height growth, the correction mGc is required to
reservoir zone, the pressure subsequently the slope of the G-plot following the termination
declines faster than the initially reduced rate and of height recession and prior to fracture closure.
as for a fracture geometry where no height The corrected slope mG′ that accounts for length
growth occurs. Thus, fracture height growth into recession during this latter phase can be inferred
the bounding zones changes the otherwise from numerical simulations and the material-
straight-line G-plot into a convex-upward curve balance relation during shut-in (Nolte, 1991):
(Fig. 9-31b).
(1 + ∆t ) f (∆t ) ,
1
From this discussion, it is clear that when mG′ ≈ mGc (9-62)
βs
cD D cD
height growth occurs, the fracture is still reced-
ing from its bounding zones when without where fD(∆tD) is the dimensionless fluid-loss rate
height recession the fracture area is momentarily function and is given in Appendix Eq. 28. The
stationary (i.e., pnet/pnet,si = 3⁄4). Therefore, the correction in Eq. 9-62 for fracture height reces-
fluid-leakoff coefficient is underestimated by sion complements Eq. 9-61 for length change to
using the slope at the 3⁄4 point in Eq. 9-58. The account for PKN-type behavior during the shut-
equation also assumes constant compliance and in process.
hence is valid only after complete height reces-
sion into the primary reservoir zone. Conse- • Variable fluid-loss coefficient
quently, the decline analysis requires that the Basic pressure decline analysis assumes that the
G-function slope be evaluated after the period fluid-loss volume is defined by a constant leakoff
of height recession. coefficient. This assumption has been shown to

9-40 Fracture Evaluation Using Pressure Diagnostics


be generally true (Mayerhofer et al., 1991; see
1.1
Chapter 8) for fluid leakoff governed by a polymer
wall cake (i.e., CL ≈ Cw). Fluid loss, however, is
pressure dependent when it is controlled by either 1.0
the invasion of fracturing fluid filtrate into the
reservoir (i.e., filtrate-controlled fluid-loss coeffi-

Cc correction factor, Kc
0.9
cient Cv) or reservoir properties (i.e., compressibil- η = 0.2

ity control leakoff coefficient Cc), as discussed in


Section 6-4. 0.8

η = 0.3
– Reservoir-controlled leakoff
0.7
Reservoir-controlled leakoff commonly occurs η = 0.4
when reservoirs with highly viscous oil are frac- mǴ = γ Kcm3/4
0.6
tured. It can also occur in low-permeability
η = 0.6
reservoirs that exhibit a high water saturation.
In the case of reservoir-controlled leakoff, the 0.5
slope of the G-plot at the 3⁄4 point can be modi- 0 1 2 3
Dimensionless fluid-loss pressure, pDLs
fied to include the pressure dependency on fluid
leakoff (Nolte et al., 1993):
Figure 9-32. G-plot slope correction factor Kc for reservoir-
mG′ = γKc m3/ 4 , (9-63) controlled leakoff (Nolte et al., 1993).

where the product γKc is applied in the same


manner as the correction for mitigating height be substituted into Eqs. 9-49 and 9-50 for calcu-
recession (i.e., Eq. 9-62) and lating the fracture area and width, respectively.

γ = 1 + β s pDLs , (9-64) – Fracturing-fluid-filtrate-controlled leakoff


Filtrate-dominated leakoff is more characteristic
where the dimensionless pressure difference pDLs
of high-permeability reservoirs where the fractur-
for pressure-dependent leakoff is defined as
ing fluid itself (e.g., viscoelastic surfactant
pws − pc or crosslinked polymer) invades the reservoir.
pDLs = , (9-65)
pc − pi It can also dominate in low-permeability forma-
tions at irreducible water saturation, with small
where pi is the reservoir pressure. values of relative permeability to the fracturing
The correction factor Kc in Eq. 9-63 is defined fluid filtrate. This fluid-loss mechanism depends
in Fig. 9-32 for various values of η and pDLs. on the square root of the pressure difference
This correction factor enables evaluation of the between the fracture and the reservoir (see
slope of the G-plot at a common reference point Section 6-4.2). The nonlinear dependence pre-
(i.e., when pnet/pnet,si = 3⁄4) for the decline analysis. cludes a derivation similar to that for reservoir-
For typical values of efficiency η < 0.5 and pDLs controlled leakoff (Nolte et al., 1993). However,
< 0.5, the figure indicates that Kc is equal essen- Eqs. 9-63 and 9-66 provide an acceptable
tially to 1. approximation if they are modified to reflect the
The corrected G-function slope from Eq. 9-63 square-root pressure behavior. The corrected 3⁄4
can be used with Eq. 9-58 for estimating the slope can be obtained in a manner analogous to
effective fluid-leakoff coefficient Cce during Eq. 9-63:
injection. Cce is related to the reservoir-controlled
fluid-loss coefficient Cc (see Section 6-4.3) eval- mG′ = ( γKc ) m3/ 4 .
1/ 2
(9-67)
uated at the pressure difference ∆p = pc – pi:
The equivalent filtrate-controlled leakoff coef-
Cce = γCc . (9-66) ficient Cve can be defined similarly to Eq. 9-66:
Cce is equivalent to the conventional pressure-
Cve = γ 1/ 2 Cv , (9-68)
independent fluid-leakoff coefficient CL and can

Reservoir Stimulation 9-41


where γ is defined by Eq. 9-64 and Cv is the vis- 4. Gc is corrected to include effects resulting from
cosity control leakoff coefficient (see Section nonideal conditions on the G-plot by defining a
6-4.3) evaluated at ∆p = pc – pi. Like for corrected value of G at closure:
Eq. 9-63, the correction factor Kc is obtained pnet ,si
from Fig. 9-32. G* = , (9-70)
In summary, for fluid loss governed by a poly- p*
mer filter cake, the fluid-loss behavior is inde- where pnet,si is the net pressure at shut-in.
pendent of pressure, and the analysis presented 5. The fluid efficiency is determined by modifying
for Eq. 9-61 or 9-62 is used. Closed-form cor- Eq. 9-60:
rections of the type presented in Eq. 9-63 extend
the basic decline analysis for the reservoir-con- G*
η= . (9-71)
trolled fluid-loss mechanism, such as for heavy- 2κ + G *
oil-bearing reservoirs. Filtrate-controlled behav- 6. Finally, CL is estimated by applying Eq. 9-58 with
ior, which is more common in the stimulation of p* determined using Eq. 9-69.
high-permeability reservoirs, shows a nonlinear
dependence on pressure. The corrections devel- In the absence of height growth and pressure-
oped for the reservoir-controlled leakoff case dependent fluid loss, m3⁄4 and mG′ are expected to be
can be adapted to quantify these effects. equal. With height growth but no pressure-dependent
fluid loss, mG′ is expected to be greater than m3⁄4 and
provide a better estimate for p*. Alternatively, for no
9-5.4. Generalized pressure decline analysis significant height growth but with pressure-dependent
fluid loss, m3⁄4 is expected to be greater than mG′
As discussed in the preceding section, commonly (Eq. 9-62) and provide a better estimate for p*.
occurring nonideal conditions modify the otherwise The relative magnitudes for m3⁄4 and mG′ provide
straight-line behavior of the G-plot. The 3⁄4 point (i.e., a diagnostic to infer the occurrence of pressure-
pnet /pnet,si = 3⁄4) provides a reference to extend basic dependent leakoff. An alternative diagnostic is the G-
pressure decline analysis for effects resulting from function semilog derivative, discussed in Sidebar 9F.
change in the fracture length. In addition, the correc- If pressure-dependent leakoff conditions occur,
tion for recession following fracture height closure or Eqs. 9-63 through 9-68 should be used to determine
pressure-dependent leakoff complements the 3⁄4-point the effective fluid-leakoff coefficient. The primary
slope with the following unified methodology for the deficiency of this method is the selection of an appro-
analysis of shut-in pressure decline before fracture priate slope when both significant height growth and
closure: pressure-dependent fluid loss occur.
1. For all geometry models, find the slope at pnet /pnet,si The 3⁄4 point provides a common reference point,
= 3⁄4 (i.e., m3⁄4). where the slope of the G-plot can be used to account
2. If the PKN model is applicable (determined by the for nonideal effects. Accordingly, this generalized
analysis of pumping pressure) and pressure-depen- evaluation methodology is commonly referred to as
dent leakoff is not expected, determine the slope at the 3⁄4 rule for pressure decline analysis.
closure mGc. Obtain the corrected slope at closure • Example field application of PKN fracture
mG′ using Eq. 9-62. decline analysis
3. The value of p* is selected from the following Shut-in pressures monitored during the calibration
equation on the basis of the applicable fracture treatment described in Section 9-4.4 are analyzed
geometry model: in this example to illustrate the application of
decline analysis to a PKN-type fracture (Fig. 9-8).
max( m3/ 4 , mG′ ) PKN The relevant formation mechanical parameters are

p* = m3/ 4 KGD (9-69) listed in Table 9-1.
m Radial. The plane strain modulus E′ is derived from E
 3/ 4 and ν:

9-42 Fracture Evaluation Using Pressure Diagnostics


9F. G-function derivative analysis Fissure-dominated leakoff
Fissure-dominated leakoff is illustrated by the analysis of a
R. D. Barree, Marathon Oil Company prefracture injection test (Fig. 9F-2) conducted in a low-per-
Ideal fracture closing behavior is characterized by a straight- meability sandstone formation. The data were acquired using
line pressure response on the G-plot, as described in Section a surface pressure gauge and show pressure-dependent
9-5.2. Extensions of this analysis use the derivative of pres- leakoff. The end of pressure-dependent leakoff is clearly indi-
sure with respect to the G-function (dp/dG; Castillo, 1987) and cated by the derivative becoming constant. This point corre-
the semilog or superposition derivative (Gdp/dG; Barree and sponds to the end of a hump on the superposition Gdp/dG
Mukherjee, 1996). The characteristics of these added diag- curve, following which the curve becomes linear. This early-
nostic curves provide a qualitative indication of the change in time hump above the extrapolated straight line on the super-
fracture geometry during shut-in. They also describe the dom- position curve, along with the sharply curving derivative, is a
inant leakoff mechanisms. This information can be reconciled signature of pressure-dependent leakoff resulting from fis-
with the observed pressure behavior during pumping, knowl- sures. Fissure-dominated leakoff presists up to a G-function
edge of in-situ stresses and rock properties, and postfractur- value of about 0.75, as shown by the onset of the straight-line
ing measurements (e.g., tracer and temperature logs) for an section of the superposition derivative. After fissure closure
improved understanding of the fracturing process. the derivative is constant and the superposition curve is linear
The diagnostics and analysis can be performed on site in a (i.e., constant slope), both indicating a constant leakoff coeffi-
simple and straightforward manner using the measured pres- cient. The pressure data alone provide a less clear indication
sure during shut-in. Values of the pressure p and the deriva- of the end of pressure-dependent leakoff. Closure of the pri-
tives dp/dG and Gdp/dG are plotted on Cartesian axes mary hydraulic fracture is apparent on the derivative and
against the G-function. The plot is interpreted primarily superposition curves at a G-function value of about 2.3, corre-
by inspection. As with a type-curve analysis, characteristic sponding to a pressure of approximately 5100 psi. The instan-
shapes in the curves are identified. Once the general nature taneous shut-in pressure (ISIP) is nearly 1500 psi higher than
of the falloff behavior is developed, specific numerical values the closure pressure.
(e.g., closure pressure and fissure dilation pressure) can be
determined.
The following field examples illustrate the application of
these diagnostics to characterize nonideal fracture and fluid- 2000 7000
loss behavior and the associated characteristic curve shapes. 6500
1600
Derivatives (psi)

Pressure (psi)
Fracture height recession Gdp/dG 6000
Pressure
1200
Height recession during shut-in from high-stress bounding lay- 5500
ers results in changes in the fracture compliance and total Fracture
800 closure
fracture surface area relative to the leakoff (i.e., permeable) 5000
area (Figs. 9-7 and 9-36). This behavior causes several obvi-
ous signatures on the pressure and derivative plots (Fig. 9F-1). 400 dp/dG 4500
The pressure versus G-function curve shows a distinct down- Fissure closure
ward bend as height recession progresses, as discussed in 0 4000
Section 9-5.3. This behavior results in an increasing magni- 0 0.5 1.0 1.5 2.0 2.5 3.0
tude of the dp/dG curve and the superposition Gdp/dG curve. G-function
Figure 9F-1 elaborates this diagnostic using shut-in pres-
sure measurements acquired during a water injection test in
a carbonate formation. Injection was confined to a relatively Figure 9F-2. Pressure derivative analysis for fissure-
small perforated interval surrounded by several hundred feet dominated leakoff.
of similar lithology. The continuously increasing values of
dp/dG and Gdp/dG indicate continuous height recession dur-
ing closure. The figure also suggests that complete fracture Pressure-dependent leakoff
closure has not occurred by the end of the shut-in period.
As discussed in Section 9-5.3, the value of dp/dG can contin-
uously decrease during the preclosure shut-in period. This
behavior results because of change in the fracture penetration
500 1500
during shut-in, and it may occur in spite of a constant value of
Bottomhole pressure (psi)

1450 the fluid-loss coefficient CL. The diminishing rate of change on


400 the G-plot is magnified if the fluid leakoff is predominantly
1400
Derivatives (psi)

pressure dependent (i.e., filtrate or reservoir controlled). In


Pressure Gdp/dG
300 1350 this case, the superposition analysis provides a distinct signa-
ture that identifies the occurrence of pressure-dependent fluid
1300
loss, despite its similarity to fracture length change during the
200 1250 shut-in period.
1200 Figure 9F-3 shows that pressure-dependent fluid loss is
100 dp/dG characterized by a superposition derivative that approaches
1150 a straight line after a short shut-in period. The straight-line
0 1100 period is unaffected by change in the fracture penetration dur-
0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 ing shut-in, and its slope is proportional to the total fluid leak-
off. The early-time Gdp/dG curve has a much steeper slope,
G-function and it falls below the straight-line extrapolation (i.e., the hump
is below the extrapolated Gdp/dG curve). This pattern differ-
Figure 9F-1. Pressure derivative analysis for fracture entiates fracture extension or recession from true pressure-
height growth.

Reservoir Stimulation 9-43


The steps outlined previously in this section are
9F. G-function derivative analysis (continued)
then applied to determine the efficiency and fluid-
leakoff coefficient.
The 3⁄4-point pressure p3⁄4 is derived from an
2000 10,500
instantaneous shut-in pressure pISI of 9012 psi and
1600 10,100 the estimated 8910-psi pc:
Derivatives (psi)

Pressure (psi)
( pISI − pc ) = pc + pnet ,si
3 3
1200 9700 p3/ 4 = pc +
Pressure
4 4
9300 3
= 8910 + (9012 − 8910) = 8987 psi. (9-74)
800 Gdp/dG

400
dp/dG
8900
4

8500
The G-function slope m3⁄4 at this 3⁄4 point is esti-
0
0 0.5 1.0 1.5 2.0 2.5 mated to be 136 psi in Fig. 9-33. For the PKN
G-function fracture geometry model, the slope prior to closure
mGc is inferred to be 123 psi. This slope is cor-
Figure 9F-3. Pressure derivative analysis for filtrate- or rected for the length recession following the termi-
reservoir-controlled leakoff.
nation of height growth using Eq. 9-62:
dependent leakoff. One ambiguity exists, however. The same 1
pattern can be generated by extension of a preexisting nat- mG′ = 123 1 + 0.58 × 0.93 = 194 .
ural fissure set oriented nearly parallel to the hydraulic frac- 0.74
ture or leakoff into a swarm of parallel fractures created by
the fracturing process.
9040

pISI = 9012 psi


E
E′ = 9000
1 − ν2
5.3 × 10 5
3/4 point = 8987 psi
pw (psi)

= = 5.6 × 10 5 psi. (9-72) 8960 m3/4 = 136


1 − 0.22 2 3/4p
net, si

Similarly, βs is determined from Eq. 9-20 using 8920


pc = 8910 psi
the fluid rheology exponent n = 0.44 and viscosity
mGc = 123
profile parameter a = 0: 8880
0 0.2 0.4 0.6 0.8 1.0 1.2
(2)(0.44) + 2
βs = = 0.74 . G (∆tD)
(2)(0.44) + 3 + 0
The fracture height hf selected as the zone of pri- Figure 9-33. Pressure decline analysis for a PKN fracture.
mary initial confinement is determined from well
logs (Fig. 9-8a) to be 24 ft. Because separation The value of mG′ greater than that of m3⁄4 signi-
occurs between the resistivity curves throughout fies fracture height growth, which agrees with the
this pay interval, fluid leakoff is inferred to occur pressure derivative analysis for the injection pres-
over the entire initial fracture height. Thus, the sure in Section 9-4.4.
ratio rp of the leakoff height to the initial fracture The value of p* is determined using Eq. 9-69 to
height is unity. be 194 psi, and G* is determined from Eq. 9-70:
The fracture closure pressure is determined from
9012 − 8910
a step rate test (see “Estimating closure pressure” G* = = 0.53 .
in the Appendix to this chapter) and pressure 194
derivative analysis (Fig. 9-9) to be 8910 psi. The Assuming negligible spurt (i.e., κ = 1), the effi-
dimensionless time at closure is ciency is determined from Eq. 9-71:
∆tc 7.0
∆tcD = = = 0.58 . (9-73) η=
0.53
= 0.21 .
t p 12.0 2 + 0.53

9-44 Fracture Evaluation Using Pressure Diagnostics


Equation 9-58 is then applied to obtain the fluid- 4555 − 4375
G* = = 0.4 .
loss coefficient: 450
194 × 0.74 × 24
CL = = 1.77 × 10 −3 ft min1/2 . As subsequently discussed in Section 9-6.8, a
1 × 12 × 5.6 × 10 5 nominal spurt with a spurt factor κ =1.02 is esti-
mated for this example. From Eq. 9-71 the effi-
• Example field application of radial fracture
ciency is predicted:
decline analysis
0.4
Application of the generalized pressure decline η= = 0.16.
analysis to a radial fracture is illustrated by analyz- (2 × 1.02) + 0.4
ing the shut-in period during the calibration treat-
The fracture radius R is determined from Eq. 9-59:
ment depicted in Fig. 9-10b. The value of rp is
equal to approximately unity on the basis of open-
(1 − 0.16) × 600 × 4.8 × 10 5 32
hole well logs. The plane strain modulus is deter- R=3 = 37 ft.
mined using Eq. 9-72 with the mechanical parame- 2 × ( π 2) × 1.02 × (32 3π ) × 450 3π 3
2

ters listed in Table 9-2:


Finally, CL is determined by applying Eq. 9-58:
4.5 × 10 5
E′ = = 4.8 × 10 5 psi . 450(3π 2 32) 32
1 − 0.252 CL = × 37
1 × 4.6 × 4.8 × 10 3π 2
5

The value of βs for the radial model is 0.925


from Eq. 9-20. = 1.62 × 10 −2 ft / min1/2 .
As in the previous example for PKN fracture
decline analysis, the steps outlined in this section
are applied to determine the fluid-loss parameters.
The 3⁄4 point is calculated as 9-6. Pressure interpretation after
3
fracture closure
p3/ 4 = 4375 + (4555 − 4375) = 4510 psi , Another application of pressure evaluation pertains
4
to the pressure response following fracture closure.
and m3⁄4 is obtained as 450 psi, as shown on
The pressure during this period reflects the transient
Fig. 9-34.
reservoir response to fracturing and is independent
The fracture is inferred to propagate in a radial
of the mechanisms governing fracture propagation.
fashion, so the correction for height growth sug-
Its character is determined entirely by the response
gested by Eq. 9-62 is not required. For the radial
of a reservoir disturbed by the fluid-leakoff process.
model, p* = m3⁄4 = 450 psi.
During this period, the reservoir may initially exhibit
The corrected value of G at closure is derived
formation linear flow followed by transitional behav-
from Eq. 9-70 as
ior and finally long-term pseudoradial flow (see
Section 12-2). Formation linear flow and pseudoradial
flow are hereafter referred to simply as linear and
pISI = 4555 psi
radial flow, respectively.
4500
3⁄4 point = 4510 psi The after-closure response is similar to the behavior
3⁄4 pnet,si
observed during a conventional well test of a propped
pw (psi)

fracture. It therefore supports an evaluation method-


m3⁄4 = 450
4400 ology analogous to the established principles of pres-
sure transient evaluation. However, one important
pc = 4375 psi
aspect differentiates after-closure evaluation. A propa-
gating or receding fracture exposes the reservoir to an
4300
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 unequal distribution of fluid-loss flux over its length.
G(∆tD) For example, fluid loss dominated by the leakoff coef-
ficient exhibits an elliptical profile of the specific vol-
Figure 9-34. Pressure decline analysis for a radial fracture. ume (i.e., volume lost per unit area) (Fig. 9-35a). The

Reservoir Stimulation 9-45


(a) Leakoff rate (b)
Specific volume
Spurt

Physical length, L
Apparent length, xfa

Figure 9-35. Fluid-loss volume and rate distribution (a) in the absence of spurt and (b) for a spurt-dominated treatment.

after-closure response can then be characterized in 9-6.1. Why linear and radial flow
terms of an equivalent and spatially uniform fluid-loss after fracture closure?
flux over an apparent half-length xfa that is generally
less than L, the physical half-length of a propagating A reservoir is disturbed by fluid-loss invasion during
fracture. Consequently, the value of dimensionless fracture propagation and closing. The fluid-loss rate
time T that is expressed in terms of L for after-closure changes with time as well as over the fracture length
analysis can be different from that of tD, the standard (i.e., has temporal and spatial components). This con-
dimensionless time (see Chapter 12) based on the cept is further explained on Fig. 9-36, which illus-
apparent stationary fracture half-length xfa inferred trates the change in fracture length during a calibra-
from the reservoir response. This distinction between tion test. At a time t = 0, the fracture length is zero.
T and tD is consistent with the notation for the fracture The fracture length increases over the injection time
lengths L and xf employed in this volume. tp, and it finally reaches a maximum value shortly
The after-closure period provides information that after the end of injection. Fluid leakoff at a hypotheti-
is traditionally determined by a standard well test cal location in the reservoir (labeled point a) is initi-
(i.e., transmissibility and reservoir pressure). It com- ated at the time ta when the fracture arrives at this
pletes a chain of fracture pressure analysis that pro-
vides a continuum of increasing data for developing
a unique characterization of the fracturing process.
At optimum conditions, it objectively determines L
parameters that either cannot be otherwise obtained
(e.g., spurt) or exhibit considerable uncertainty when
estimated with conventional pressure decline analyses
Distance

(e.g., closure pressure). a τ


Special attention is also required to address the
interpretation aspects for application of the after-
closure reservoir response. These considerations,
along with an overview of related theoretical frame-
work, are discussed in this section. A physical
description of the reservoir response to fluid loss is
presented initially to enhance understanding of the
ta tp tr tc
after-closure interpretation methodology. The syn-
ergy attained by combining the after-closure analysis Time, t

with information derived from the fracture injection


and preclosure periods is also discussed. Figure 9-36. Rock exposure to fluid loss during fracture
propagation.

9-46 Fracture Evaluation Using Pressure Diagnostics


location. The fluid-loss behavior can be interpreted fracture propagation, which is equivalent to a diffu-
as an injection source at point a with a strength (or sional rate of L2/t. The relative magnitude or ratio of
intensity) equal to the rate of fluid loss that perturbs the diffusional rates determines the shape of the pres-
the reservoir from its initially undisturbed state. Fluid sure patterns and defines the dimensionless time T:
leakoff continues while the fracture remains open for
pressure diffusional rate
the time interval τ and ends when the fracture recedes T=
to close at the reference point time tr. fracture growth diffusional rate
The calibration test creates a fracture that has negli-
=
(k φµc ) =
t kt
. (9-75)
gible conductivity upon its closure at time t = tc,
following which additional fluid loss is terminated.
(L t)
2
φµct L2
Furthermore, following closure, the reservoir loses all Several pressure propagation patterns are possible
recollection of the mechanical aspects (e.g., elastic depending on the value of T. Figure 9-37a, based on
deformation) of fracturing. The pressure disturbances a reservoir simulation for a low value of T, illustrates
created by leakoff, however, are preserved and dissi- the pressure profile shortly after fracture closure in a
pate with time. The wellbore pressure measurements reservoir characterized by low diffusivity and rapid
following fracture closure reflect the dissipation fracture extension, or a low-permeability reservoir.
process. The after-closure reservoir response is the The pressure disturbances near the wellbore have trav-
cumulative effect (i.e., superposition) from injection eled a small distance into the reservoir. They are gen-
sources distributed along the mechanical length of the erally transmitted perpendicular to the previously frac-
fracture and over a time interval representing their tured surface, indicating a one-dimensional (1D)
exposure to fluid loss. pattern or linear flow. The pressure profile also reflects
The pressure disturbances induced by fluid loss dis- the fracture length attained during fluid injection.
perse or diffuse into the reservoir at a rate governed A similar, and potentially unanticipated, pressure
by the ratio of the mobility k/µ and storage φct (i.e., profile is depicted in Fig. 9-37b for the same fracture
the reservoir diffusivity k/φµct) expressed in terms of propagation rate as in Fig. 9-37a but for a reservoir
the permeability k, porosity φ, viscosity µ and total with a much larger diffusional speed and therefore a
compressibility ct. The resulting pressure patterns cre- relatively large dimensionless time. These conditions
ated by the disturbances depend also on the rate of can be anticipated in a high-permeability reservoir

pi 0.25(pw – pi) 0.5(pw – pi) 0.75(pw – pi) pw

(a) (b)
Fracture Fracture
1 0.5re re 1 0.5re re
0 0

0.5re 0.5re

re re

Figure 9-37. Pressure distribution in the reservoir shortly after closure for (a) small and (b) large values of dimensionless
time. re = reservoir radius.

Reservoir Stimulation 9-47


where the pressure disturbance caused by fluid 9-6.2. Linear, transitional and radial flow
leakoff travels much faster than the rate of fracture pressure responses
growth. Figure 9-37b clearly shows that the pressure
distribution is completely devoid of any linear flow The characteristic pressure response at the wellbore
character and is represented largely as a radial flow and beyond the filtrate region following the closure
pattern. The pressure distribution also does not reflect of an unpropped fracture is illustrated in Fig. 9-38.
the fracture length. Its magnitude at the wellbore is The pressure difference ∆p = p(t) – pi, normalized
independent of the nature of fluid injection into the with respect to its value at closure p(tc) – pi, was sim-
reservoir but is characterized by the reservoir mobil- ulated for a dimensionless time T = 0.0001 at fracture
ity and injected volume. closure, or for a fracture that propagated much faster
A final aspect of the volume distribution pertinent than the pressure disturbance in the reservoir. Figure
to the after-closure response must be described. In 9-38 also shows the pressure derivative and variation
contrast to the elliptical profile in the absence of spurt of the characteristic log-log slope defined as d(ln ∆p)/
(Fig. 9-35a), spurt loss acts as a moving injection d(ln t), where ∆p is the pressure above the initial
source and tends to distribute the fluid-loss volume reservoir pressure pi. This slope is used to identify the
equally over the fracture length (Fig. 9-35b). As indi- time intervals over which the various flow regimes
cated previously, radial flow behavior depends pri- prevail. The pressure response is illustrated for time
marily on the volume of fluid injection and is unaf- intervals that extend beyond 10 times the closure
fected by the leakoff profile. However, during linear period. For this limiting situation, known as the
flow, the reservoir remembers this difference in the impulse condition (see Sidebar 9G), the pressure
fluid-loss behavior because it affects the reservoir variation is described by the ratio of the closure time
pressure distribution shortly after closure. Spurt- to the current time (Gu et al., 1993; Abousleiman et
dominated fluid-loss behavior thus is manifested as al., 1994).
a change in the attributes of after-closure linear flow. (a)
Several conclusions can be derived from the pre- 1.0
ceding qualitative descriptions of the reservoir Pressure
Normalized pressure difference

difference
response:
and pressure derivative

• Fluid injection tests that achieve efficient fracture 0.1


Transitional knee time
extension (as in most calibration treatments) are T = 0.2
likely to achieve a low dimensionless time at the Pressure
derivative
end of pumping. Therefore, they could display 0.01
well-defined periods only of linear flow and long-
term radial flow after fracture closure.
• The pressure distribution during linear flow reflects 0.001
0.001 0.01 0.1 1.0 10.0
the fracture geometry attained during injection. The
(b)
related analysis can provide an estimate of the frac-
1.0
ture length.
• Linear flow is also affected by the relative contri- 0.9
bution of spurt, which can be distinguished by Extended Radial
radial
Log-log slope

evaluating the pressure data from this period. 0.8

• Fluid injection tests characterized by a low effi- T = 0.005 T = 0.05 T = 0.5 T = 5.0
0.7
ciency can be devoid of linear flow in spite of the Linear Extended
creation of a fracture. 0.6
linear

• Radial flow is independent of the nature of fluid


leakoff. This period is, however, governed by the 0.5
0.001 0.01 0.1 1.0 10.0
reservoir mobility and can be used to estimate the
Dimensionless time, T
reservoir transmissibility.
Figure 9-38. After-closure pressure difference and deriva-
tive (a) and log-log slope (b) plots for an impulse injection.

9-48 Fracture Evaluation Using Pressure Diagnostics


9G. Impulse testing where pi is the reservoir pressure, µ is the reservoir fluid vis-
cosity, k is the reservoir permeability, and h is the reservoir
~
Joseph Ayoub, Schlumberger Dowell height. The term q defines the average injection rate into the
reservoir. Equation 9G-3 can be reformulated on the basis
The impulse test is a specialized well testing procedure that that the injected fluid volume is the product of the average
enables analysis of the reservoir response following a rela- ~
injection rate and closure time (i.e., Vi = q tc):
tively short duration of fluid injection or production (Ayoub et
al., 1988). The pressure response to an ideal-rate impulse Vi µ 1  1 tc 
pw ( ∆t ) − p i = 1− 
(i.e., to an instantaneous source) is given by the derivative of 4 πkh ∆t  2 ∆t 
the pressure response to a step rate change. The impulse pro- Vµ 1  1 
cedure can also be applied to a short injection that creates a = i 1− ε , (9G-4)
hydraulic fracture within the reservoir (Abousleiman et al., 1994). 4 πkh ∆t  2 
The dimensionless wellbore pressure response pwD follow-
ing fracture closure resulting from a constant injection rate where ε = tc /∆t. It follows that for ∆t ≥ 10tc, pw(∆t) – pi is pro-
can be expressed by the principle of superposition as portional to 1/∆t within an error of less than 5%.
The pressure derivative ∆tdpw/d∆t can be obtained from
Eq. 9G-4:
p wD (t D > t cD ) = p wD (t cD + ∆t D ) − p wD ( ∆t D ) , (9G-1)
dp w ( ∆t ) Vi µ 1
where tD is the conventional dimensionless time from Eq. 12-1 ∆t
d∆t
=−
4 πkh ∆t
[1− ε] . (9G-5)
and tcD and ∆tD represent the dimensionless times for fracture
closure and any subsequent interval, respectively (Fig. 9G-1). The term ε can be neglected when the shut-in period is rel-
Equation 9G-1 can be reformulated using the standard atively long compared with the closure time. For ∆t ≥ 10tc,
Taylor series expansion: Eqs. 9G-2 and 9G-5 become equal and proportional to 1/∆t
2
within a 10% error. It follows that
p wD (t D > t cD ) = t cD
dp wD 2 1 d p wD
+ t cD + higher order terms .
d∆t D 2 d∆t D2 dp w ( ∆t ) Vi µ 1
p w ( ∆t ) − p i = ∆t = , (9G-6)
(9G-2) d∆t 4 πkh ∆t
Neglecting the higher order terms in Eq. 9G-2 and substi- where ∆t >> tc.
tuting for pwD with its expression during pseudoradial flow from
Eq. 12-15 gives Practical considerations
q˜µ t c  1 tc  In theory, impulse analysis is applicable only for an instanta-
p w ( ∆t ) − p i = 1− , (9G-3) neous fluid injection. In practice, however, a finite injection
2πkh 2∆t  2 ∆t 
time is required to create a pressure disturbance that is mea-
surable for a long shut-in duration. Consequently, the analysis
is valid only if the shut-in time is long in comparison with the
End of pumping finite injection time, as previously discussed.
Equation 9G-6 shows that during reservoir radial flow, a
pc log-log plot of the pressure difference and pressure derivative
versus 1/∆t exhibits a unit slope. Multiplying the pressure dif-
Pressure (psi)

ference and pressure derivative by ∆t provides a horizontal


line (i.e., zero slope) similar to the classic pressure response
for transient radial flow. The two lines also overlie each other.
tc ∆t This behavior provides a diagnostic method not only for iden-
tifying reservoir pseudoradial flow but also for determining the
initial pressure of the reservoir. In addition, the pressure
response during this period can be used in conjunction with
Eq. 9G-6 to estimate the reservoir transmissibility kh/µ. This
Time, t provides the basic parameters required for evaluating the
reservoir production performance.

Figure 9G-1. Reference times for the impulse injec-


tion. pc = closure pressure.

As shown on Fig. 9-38b, linear flow appears ini- also emphasizes the “knee” formed by the intersec-
tially and lasts until T exceeds 0.005 and exhibits the tion of the slopes for the linear and radial flow peri-
anticipated half-slope for the pressure difference and ods. This intersection defines a unique dimensionless
derivative. Reservoir radial flow, characterized by a time that can be used to infer the fracture length
unit log-log slope, is the terminal flow regime that achieved during propagation, as subsequently dis-
occurs only for T > 5. The intermediate transitional cussed in Section 9-6.7.
period lasts over a time period spanning a factor of Figure 9-38 also illustrates a significant shortcom-
approximately 1000 times beyond the end of linear ing for extracting the reservoir transmissibility from
flow. The plot of the pressure difference in Fig. 9-38a the after-closure response in low-permeability reser-

Reservoir Stimulation 9-49


voirs that have extensive, efficient fracture propaga- period of after-closure pressure data (Nolte et al.,
tion. Such conditions are characterized by a small 1997). As discussed in Chapter 8, polymer fluid inva-
value of Tp and require a long after-closure period to sion into a moderate- to high-permeability reservoir
attain radial flow. For the example in Fig. 9-38, a pro- can result in an internal cake of a highly concentrated
hibitively long shut-in period that equals 50,000 polymer gel. This internal cake experiences a sus-
times the injection time would be required prior to tained period of creeping afterflow following fracture
obtaining radial flow. However, from an engineering closure. It supports a fraction of the total pressure
perspective (i.e., an allowable error of 10%), the tran- drop between the fracture and the reservoir, masking
sitional period can be shortened by extending the lin- the reservoir response. After a time that equals
ear and radial flow periods, as shown in Fig. 9-38b. approximately the time before closure, the internal
This approximation considerably reduces the transi- cake begins to flow more freely. The pressure then
tional period from a factor of 1000 to 10, which quickly drops to a value that reflects the pressure dif-
greatly increases the likelihood of obtaining a linear ference in the reservoir beyond the polymer material.
or radial flow analysis with an acceptable accuracy Finally, the increasing effective stress during the
and under the normal constraints of field operations. after-closure period continues to consolidate the filter
cake and fracture faces and further reduce the near-
wellbore permeability. This can eventually result in
9-6.3. Mini-falloff test a loss of communication between the wellbore and
A robust strategy for estimating the reservoir parame- the reservoir that limits the availability of valid long-
ters from radial flow as well as for characterizing term pressure data.
fracture behavior and obtaining linear flow is to apply “Guidelines for the field application of after-closure
two separate tests, with the first test a short injection analysis” in the Appendix to this chapter discusses
test that may or may not create a fracture. This spe- operational considerations for the mini-falloff test.
cialized calibration test is referred to as the mini- Appendix Eq. 1 provides a guideline for determining
falloff test. The mini-falloff test should be performed the injection rate during the test. It follows from this
in an undisturbed reservoir. Except in very tight for- equation that early radial flow can be achieved in a
mations, the test can be engineered using inefficient low-permeability gas reservoir (i.e., in tenths of a mil-
fluids and a low injection rate such that radial flow lidarcy) with an operationally permissible injection
occurs during injection or shortly after closure (i.e., rate (≈ 1 bbl/min). Lower injection rates are required
early radial flow). The subsequent fracture calibration to attain radial flow within the shut-in period in a
test is performed with much higher injection rates reservoir with microdarcy permeability. Under these
and a more efficient fracturing fluid to characterize extreme reservoir conditions, only linear or transi-
the fracture behavior as well as attain after-closure tional flow may occur because of field operational
linear flow. The long-term radial flow behavior that constraints. However, transitional flow can also be
normally occurs only after a prohibitively long shut- analyzed to estimate the reservoir properties, as dis-
in period can be anticipated from the reservoir infor- cussed in “Comparison of fixed-length and propagat-
mation derived with the mini-falloff test. The mini- ing fractures” in the Appendix to this chapter.
falloff test, therefore, also facilitates the integration
of pre- and after-closure analyses for a calibration
test, as discussed in Section 9-6.4, under the time
9-6.4. Integration of after-closure
constraints of normal field fracturing operations.
and preclosure analyses
A nonpolymer fluid (e.g., well completion fluid) The after-closure analysis complements the preclosure
must be injected during the mini-falloff test for two analysis by providing parameter estimates that cannot
reasons. First, this type of fluid generally exhibits a otherwise be derived from pretreatment calibration
larger fluid-loss rate than that of conventional poly- tests. After-closure reservoir radial flow determines
mer fracturing fluids. Consequently, it promotes a the formation transmissibility and initial pore pres-
larger dimensionless time and the earlier emergence sure. These parameters are of paramount importance
of radial flow during the shut-in period. Second, for optimizing a fracture stimulation (see Chapters 5
physical effects resulting from the viscoelastic nature and 10). In addition, spurt can be a significant consid-
of polymer fluids can potentially corrupt an extended eration for wall-building fluids in high-permeability

9-50 Fracture Evaluation Using Pressure Diagnostics


fracturing. However, it cannot be estimated from the
standard pressure decline analysis described in
Pumping Shut-in After-closure
Section 9-5. As previously mentioned, spurt influ-
• Fracture model • cf • G* • p* •κ • xfa
ences the linear flow response. Consequently, after-
closure linear flow analysis can be used to obtain an
improved on-site assessment of spurt over conven-
tional core-based laboratory tests.
A significant enhancement provided by after- Fluid loss
closure analysis is its potential to validate preclosure •η • CL
analysis. The reservoir linear flow regime retains two
critical pieces of information concerning the fracturing
process: the time when fluid loss terminated (i.e., frac-
ture closure time) and the fracture length. Cessation of
Fracture geometry Analyses Reservoir
fluid loss provides the linear flow period with a more agree
•L •L
distinct marker for fracture closure than that obtained
from conventional shut-in procedures (see “Estimating
closure pressure” in the Appendix to this chapter) that Figure 9-39. Interrelation of pre- and after-closure pres-
may exhibit multiple inflection points or a continuous, sure analyses.
gradual pressure variation during closing. Fluid-loss
discontinuity provides the fundamental basis for using duction parameters. The synergistic, complementary
after-closure analysis to determine the fracture closure character of the pre- and after-closure pressure analy-
pressure (see Section 9-6.8). ses provides a comprehensive suite of crossvalidated
The transitional period exhibits a unique dimen- fracture parameters for the optimized design of frac-
sionless knee time (see Fig. 9-38a) that provides ture treatments based on the economic guidelines
information related to the fracture length through provided in Chapter 10.
Eq. 9-75. The fracture length can also be estimated
from the preclosure pressure decline analysis (see
Section 9-5.2). The after-closure linear flow charac- 9-6.5. Physical and mathematical
teristics are derived from the principles of transient descriptions
reservoir flow, which are fundamentally different
The total pressure difference between the fracture and
from the principles of linear elasticity and material
the reservoir ∆pT can be divided into three compo-
balance for a fracture that govern the preclosure pres-
nents, as shown in Fig. 5-17. Of relevance to after-
sure analyses. A consistent calibration evaluation is
closure analysis is the pressure difference ∆pR in the
indicated by agreement between the fracture lengths
reservoir beyond the filter cake and filtrate regions.
estimated by these divergent analyses.
This pressure difference represents the added contri-
The preceding discussion introduces the interrela-
butions of the two sources of fluid leakoff: a contri-
tion between the pre- and after-closure flow periods,
bution from the Carter-based leakoff loss (i.e., the CL
which is summarized in Fig. 9-39. The four inputs of
component of fluid loss) ∆pRC and that resulting from
cf, G*, p* and κ are required to determine the fluid
spurt ∆pRS:
leakoff coefficient (Eq. 9-57) and the fluid efficiency
(Eq. 9-71). The value of cf is provided by the pressure ∆pR = ∆pRC + ∆pRS . (9-76)
response during injection (see Section 9-7). The pre-
closure shut-in period (see Section 9-5) estimates of As discussed in Section 9-6.3, the pressure changes
G* and p* and the previously discussed after-closure across the filter cake and filtrate disappear within a
analysis can be used to predict the spurt factor κ. The short time after fracture closure. The subsequent bot-
fracture length is independently obtained by these tomhole pressure measurements p(t) then reflect the
two approaches, and a consistent fracture analysis is reservoir response to the pressure changes and fluid-
indicated by agreement of the two length estimates. loss distribution induced during the propagation and
Finally, either the after-closure transitional or radial closing periods:
flow period is also used to assess the reservoir pro- ∆p(t ) = ∆pR (t ) = p(t ) − pi t > tc . (9-77)

Reservoir Stimulation 9-51


A physical and mathematical description for the It follows from Eqs. 9-75, 9-78 and 12-1 that the
after-closure period is in this section. The focus is on dimensionless time tD, which corresponds to the
∆pRC, the reservoir response to Carter-based leakoff. apparent length xfa, is related to T as
The effects resulting from spurt loss are addressed
separately in Section 9-6.6. T = faL2 t D . (9-79)
The two primary changes that the reservoir experi- A general expression for faL in terms of the fluid-
ences following fracture closure are change in its loss parameters was given by Nolte et al. (1997):
boundary condition and change in its representation
in reference to the response during production of a g0  κ 
fixed-length fracture. First, as discussed in “Pressure faL =  . (9-80)
characterization for a propagating fracture” in the
(1 − η)  g0 (κ − 1) + 2 tc t p 
Appendix to this chapter, the pressure difference ∆pRC For the case of inefficient fracture propagation
in the reservoir is constant during fracture propaga- (η → 0) and negligible spurt (κ → 1), g0 = π/2
tion. In contrast, the after-closure period is character- and tc → tp. Equation 9-80 then indicates that
ized by cessation of the fluid-loss flux. Consequently, faL = π/4 (Fig. 9-35a). For spurt-dominated treat-
the boundary condition along the closed fracture ments (κ >> 1), the specific volume-loss distri-
changes from one of constant, uniform pressure to bution over the fracture length is approximately
zero or constant flux. The change in the boundary constant, as illustrated in Fig. 9-35b. As expected
condition is referred to as a mixed-boundary condition. for this case, Eq. 9-80 indicates that faL approaches
The second difference between the preclosure and unity.
after-closure periods pertains to the representation of
• Linear flow
the after-closure period, which is analogous to the
problem of a fixed-length fracture used in pressure The similarity between reservoir transient flow and
transient analysis. The leakoff distribution following heat transfer arises because of the underlying “dif-
fracture closure is shown in Fig. 9-35. The illustrated fusional” process that governs these two physical
volume profile closely reflects the distribution phenomena. The physical concept of heat conduc-
expected for a fixed-length, uniform-flux fracture and tivity is similar to that of reservoir mobility,
can be used to represent the after-closure pressure whereas the heat capacity of a solid body is equiv-
response. This observation contrasts with the pre- alent to the reservoir storage. An expression for
closure conditions of a propagating fracture, for ∆pRC during the after-closure linear flow period for
which the fluid-loss flux distribution is better repre- the mixed-boundary condition can be adapted from
sented by a fixed-length, uniform-pressure fracture. a similar condition presented by Carslaw and
This distinction is elaborated in “Comparison of Jaegar (1959) for heat transfer (Nolte et al., 1997):
fixed-length and propagating fractures” in the πµ
Appendix to this chapter. ∆pRC (t ) = CL t = tc
kφct
• Apparent fracture length πµ  2 −1  tc  
= CL  sin    t > tc
The previous discussion indicates that the after- kφct  π  t  
closure reservoir response can be analyzed by con-
πµ  2 tc 
sidering the leakoff flux to be uniformly distrib- ≈ CL   t > 3tc ,
uted over an equivalent fixed-length fracture. The kφct  π t 
apparent length of the equivalent fracture xfa is (9-81)
generally smaller than the physical length L
attained by the fracture during propagation. Its where t is the time since fracture initiation.
ratio with respect to the physical length is defined Equation 9-81 is based on 1D heat transfer
as the apparent length fraction: through the surface for an infinitely long fracture.
It also applies to the wellbore pressure of a finite-
x fa length fracture in linear flow because the reservoir
faL = . (9-78)
L surrounding the wellbore is experiencing 1D reser-
voir flow (Fig. 9-37a).

9-52 Fracture Evaluation Using Pressure Diagnostics


• Radial flow alent to the standard Horner function used during
No relation analogous to Eq. 9-81 for a finite- pressure transient testing.
length fracture with a mixed-boundary condition
applicable to radial flow has been presented in the
literature. This reservoir response can alternatively 9-6.6. Influence of spurt loss
be described by a superposition relation with the The preceding discussion of the reservoir response
logarithmic time function that is commonly used does not consider the effect of spurt. Its considera-
to describe the radial flow response for a fixed- tion, however, is important because spurt can domi-
length fracture (Eq. 12-15): nate the wall-building fluid-loss behavior of moder-
ate- to high-permeability formations.
µ Vi  1  λtc + (t − tc )   The role of spurt during either the pre- or after-
∆p(t ) =  ln 
2 πkh tc  2 λ  t − tc   closure period is difficult to describe with simple ana-
µ Vi lytical concepts. However, characterizing its contribu-
≈ t – tc >> λtc . (9-82) tion to the reservoir pressure difference ∆pRS was
4 πkh t − tc
found to be relatively simple, based on the results
The terms preceding the brackets on the right- of numerical simulations (Nolte et al., 1997). These
hand side of Eq. 9-82 are incorporated to ensure indicate that during the linear flow period spurt
that its long-term asymptotic behavior provides the causes a time-dependent increase in the pressure. Its
well-known response for an impulse test (see pressure response ∆pRS is added to the pressure con-
Sidebar 9G). tribution from Carter-based leakoff behavior
Equation 9-82 also incorporates the apparent (Eq. 9-81). The total reservoir pressure difference
closure time λtc. The apparent production time ∆p during the after-closure linear flow period that
used for the mixed-boundary condition in standard includes the spurt contribution can be expressed in
pressure transient analysis (Horner, 1951) is terms of ∆pRC and the spurt factor κ as
defined as the ratio of the cumulative volume and
2  t  κ − 1 tp 
the final production rate. This definition of appar- ∆p(t ) = ∆pRC (tc ) sin −1  c  + t ≥ tc .
ent time is not applicable for a propagating frac- π  t 2 t 
ture. Rather, the multiplier λ was selected from
(9-84)
numerical simulations as the value that provides
the shortest after-closure time for the application In particular, the pressure difference during an
of Eq. 9-82. For the mini-falloff test and its related extended period of linear flow can be approximated
conditions of negligible efficiency and spurt, the from Eqs. 9-81 and 9-84 as
optimized value of λ can be characterized as
πµ  2 tc 
∆p(t ) ≈ CL (1 + fκ )   t > 3tc , (9-85)
λ = 1+
0.14
η → 0, κ → 1. (9-83) kφct  π t 
Tp
where the spurt fraction fκ is defined as
Equations 9-82 and 9-83 provide the reservoir
π tp
transmissibility with acceptable accuracy signifi- fκ = (κ − 1) . (9-86)
cantly before the actual occurrence of radial flow. 4 tc
For example, for a dimensionless pumping time Tp The distinctive time behavior exhibited by spurt
= 0.1, the optimized multiplier expression enables during the after-closure period is illustrated in
the application of Eq. 9-82 at a dimensionless time Fig. 9-40, which is based on the linear flow time
T = 0.5 as opposed to T = 2 required for a standard function FL(t/tc) that is subsequently defined in
radial flow analysis. Furthermore, Eq. 9-83 shows Eq. 9-88. As discussed in Section 9-6.8, this presenta-
that λ is equal essentially to unity for well-devel- tion is consistent with the diagnostic log-log plot for
oped radial flow (i.e., Tp > 5). For these conditions, flow regime identification. The reservoir pressure dif-
the effect of the mixed-boundary condition dimin- ference is normalized with respect to ∆pRC(tc) and is
ishes. The time function in Eq. 9-82 then is equiv- compared for the conditions of negligible spurt (κ =

Reservoir Stimulation 9-53


• increase in the apparent length that delays the knee
time
η = 0.20
Tp = 0.01 • deviation from the expected pressure response dur-
1 ing linear flow that lasts over a shut-in time period
difference and derivative

Knee time
equal to approximately 3tc.
Normalized pressure

κ=5 9-6.7. Consistent after-closure


0.1 diagnostic framework
κ=1 The theoretical relations outlined in the previous sec-
Difference tions can be readily distilled into an elementary set of
Derivative
equations. These simplified relations provide a con-
1 10 100
sistent framework for analyzing after-closure behav-
1/{FL(t/t c )} 2
ior and are summarized in this section.
• Linear flow
Figure 9-40. Simulated pressure for spurt and nonspurt
cases (Nolte et al., 1997). The relation in Eq. 9-84 is used to describe the lin-
ear flow behavior because it provides a general
1) and dominant spurt (κ = 5). An efficiency of 0.2 mathematical representation for this flow regime.
is prescribed for both cases. Under these conditions, The equation can be reformulated as
Eq. 9-48 implies that 5 times more fluid is lost during ∆pR (t ) = mlf FL (t tc ) , (9-87)
propagation for this particular value of κ than for the
condition of negligible spurt. where FL(t/tc) is the linear flow time function and
Figure 9-40 shows that spurt results in a pressure mlf is the corresponding slope on a Cartesian plot:
difference ∆p and derivative that fall below the char-
2 −1  tc 
acteristic half-slope for linear flow over a time period FL (t tc ) = sin   (9-88)
of about t < 3tc (Eq. 9-85). The figure also indicates π  t
that spurt apparently decreases the knee time, which
is the time defined by the intersection of the asymp- πµ
totic pressure responses during linear and radial flow. mlf = CL (1 + fκ ) κ > 1, t > 3tc
kφct
The apparently earlier time of the knee in the pres-
πµ
ence of spurt occurs because the time function used = CL κ = 1, t > tc , (9-89)
in Fig. 9-40 is referenced to the closure time. Spurt, kφct
however, has a significant impact on the closure time
itself for a given value of efficiency (Eq. 9-60). For where the spurt fraction fκ is as defined in Eq. 9-86.
the spurt parameters used to develop Fig. 9-40, the The dependency of fκ on the spurt factor κ, as well
closure time for spurt is 7.5 times larger than its value as the closure time tc, implies that both parameters
in the absence of spurt. In reality, the dimensional cannot be simultaneously determined from the
knee time for the spurt case occurs later than for the reservoir linear flow response. An accurate value
nonspurt case (i.e., by a factor of 1.4 relative to the of tc is therefore required to properly characterize
beginning of pumping for these spurt parameters). κ for spurt-dominated conditions.
The later knee time for the spurt case results because In the absence of spurt, κ = 1, fκ = 0, and the last
of its larger apparent fracture length, as indicated by relation in Eq. 9-89 provides an expression analo-
Eq. 9-80. The relation between the apparent length gous to Eq. 2-50 for estimating the reservoir per-
and knee time is further discussed in Section 9-6.7. meability from the linear flow response, as sug-
In summary, spurt results in gested by Mayerhofer et al. (1993). The square-
root dependence of Eq. 9-89, however, makes this
• an added component to the reservoir pressure dif- procedure sensitive to the general uncertainty in
ference during the pre- and after-closure periods the parameters that define the fluid-loss coefficient

9-54 Fracture Evaluation Using Pressure Diagnostics


and reservoir storage. Consequently, the technique Equations 9-87 and 9-92 suggest that a log-log
could introduce significant uncertainty in the diagnostic plot of the pressure difference and pres-
inferred permeability (Ispas et al., 1998). sure derivative based on {FL(t/tc)}2 would provide
• Radial flow a consistent basis for diagnosing both linear and
radial flow. The plot also would provide the half-
Equation 9-82 provides the general expression
slope and unit slope behavior during linear flow
of the logarithmic time function that describes the
and radial flow, respectively, as is expected for
reservoir pressure difference during radial flow.
after-closure pressure analysis.
This function can be approximated by the square
of its linear flow counterpart (Eq. 9-88) by intro- • Spurt
ducing a coefficient of π2/8: Spurt is estimated from the linear flow slope mlf
2 using Eqs. 9-86 and 9-89:
1  λt  π 2  2 −1  tc  
ln1 + c  ≈  sin    (1 + ε ) , 4  mlf kφct  t
2 λ  t − tc  8  π  t   κ = 1+  − 1 c . (9-94)
π  CL πµ  tp
(9-90)
The spurt factor κ can be related to the spurt
where ε is the error introduced by the approxi- coefficient Sp with Appendix Eq. 24.
mation. The relative error can be based on the
• Fracture length
pressure derivative because it is the primary mea-
sure for after-closure analysis. Thus, The fracture length is obtained from the dimension-
less time corresponding to the knee formed by the
ε = λ −  c 
4 t intersection of the asymptotic log-log slopes during
∆t > 3tc . (9-91)
 3   ∆t  linear and radial flow, in a plot similar to Fig. 9-38a.
A general relation for the fracture length can be
It follows from Eq. 9-91 that the error introduced
derived from the pressure response for a fixed-
by the approximation in Eq. 9-90 is at a minimum
length fracture and relating its behavior to that of
when the apparent time multiplier λ = 4⁄3. As indi-
the propagating fracture using the apparent length
cated by Eq. 9-83, this value of λ corresponds to a
fraction faL, as discussed in Section 9-6.5.
dimensionless injection time Tp = 0.4, which meets
The fracture length is required to verify the pre-
the desired condition of a mini-falloff test for early
closure analysis for a calibration test, where the
radial flow analysis. The approximation, therefore,
use of a more efficient fracturing fluid generally
is well suited for application to the mini-falloff test.
leads to a well-defined linear flow response. The
More generally, the relative error ε can be shown to
knee subsequently occurs only after a significantly
be less than 5% for the after-closure period recom-
longer after-closure shut-in period (Fig. 9-38).
mended in “Guidelines for the field application of
Under this extended shut-in condition, the after-
after-closure analysis” in the Appendix to this
closure response can be approximated by the
chapter and using the expression for λ provided
impulse assumption (see Sidebar 9G). The asymp-
by Eq. 9-83.
totic expressions for a fixed-length fracture for lin-
The radial flow period can then be represented
ear and radial flow, in terms of the classic dimen-
by substituting the approximation provided by
sionless time tD, follow from Eqs. 12-12 and 12-15,
Eq. 9-90 into Eq. 9-82:
respectively. The derivatives during these periods
{
∆p(t ) = mrf FL (t tc ) , }
2
(9-92) are
 π
dpD  2 t D
where the function FL(t/tc) is defined in Eq. 9-88 Linear flow
and mrf is the corresponding slope on a Cartesian = (9-95)
plot: dt D  1
Radial flow,
 2t D
π Vi µ
mrf = . (9-93) where pD is the standard dimensionless pressure,
16 khtc defined in Eq. 12-2. The knee time tD,knee, or the
dimensionless time for the crossing of the two

Reservoir Stimulation 9-55


derivatives, is determined by equating the two reservoir’s perspective of the fracture length, as
expressions: derived from the knee time concept of Eq. 9-100:
1 κCL t p
t D,knee =
. (9-96) 4 faL mlf tc k φct
π = . (9-101)
(1 − η) π µ π
Equation 9-96 can be expressed in terms of the
dimensionless time T for a propagating fracture by The derivation of Eq. 9-101 also uses the defini-
substituting Eq. 9-79: tion of mrf from Eq. 9-93. The application of
Eq. 9-101 to a calibration test serves as a valida-
faL2
Tknee = faL2 t D,knee = . (9-97) tion check for its evaluation and provides the fun-
π damental relation to verify the parameters pre-
As discussed in Section 9-6.5, for negligible dicted by the pre- and after-closure analyses.
efficiency and spurt, faL = π/4. It follows from
Eq. 9-97 that Tknee = 0.2, as shown for these con-
ditions assumed by Fig. 9-38. An estimate of the 9-6.8. Application of after-closure analysis
actual fracture length L can be obtained in terms A systematic procedure for the application of after-
of the dimensional knee time tknee by combining closure analysis is presented in this section. “Guide-
Eqs. 9-75 and 9-97: lines for the field application of after-closure analy-
1 πktknee sis” in the Appendix to this chapter discusses the
L= . (9-98) operational considerations required to engineer an
faL φµct
injection testing sequence that provides the pressure
The value of tknee can be defined by mlf and mrf, data required for an objective, comprehensive appli-
which are more readily estimated during field cation of the after-closure analysis.
practice. This definition of tknee also uses the
• Background information
extended time approximation for FL (Eq. 9-81):
2
A reliable estimate of the reservoir pressure should
4t  m  be established prior to the evaluation process. This
tknee = 2c  rf  . (9-99)
π  mlf  can be achieved from previous analyses (e.g., well
tests or production evaluation) or data measured
Combining this expression with Eq. 9-98 gives prior to fluid injection. The estimate suitably con-
the fracture length as strains the pressure difference within the flow
regime identification process, which ensures objec-
1 4 ktc mrf
L= , (9-100) tivity in the after-closure analysis. The fluid leak-
faL π φµct mlf off can be determined using the shut-in analysis
where the apparent length fraction faL is from described in Section 9-5. The formation storage
Eq. 9-80. φct is established using standard well logs and
It is unlikely that both linear and radial flow will pressure-volume-temperature (PVT) data or
occur during the same shut-in period. Consequently, correlations.
applying the length relation in Eq. 9-100 for a cali- • Flow regime identification
bration test requires a combination of its linear The occurrence of after-closure linear or radial
flow response and the radial flow analysis from a flow is identified using the diagnostic log-log plot
preceding mini-falloff test. Equation 9-93 can be of the pressure difference and pressure derivative
used to anticipate mrf for the calibration test from based on {FL(t/tc)}2, as discussed in Section 9-6.7.
the injected volume and fracture closure time of The two quantities are plotted as a function of
the calibration test and transmissibility estimate 1/{FL(t/tc)}2 so that the shut-in time conventionally
obtained from a preceding mini-falloff test. increases from left to right along the x-axis
• Validation of pre- and after-closure analyses (Fig. 9-41).
As a final step, the fracture length estimated by During linear flow, the curves for the two quan-
decline analysis in Eq. 9-59 is compared with the tities show straight-line behavior separated by a

9-56 Fracture Evaluation Using Pressure Diagnostics


• Linear flow analysis
1.0
Pressure difference Generally, linear flow analysis is applicable only
Pressure derivative
to a calibration test (i.e., because a mini-falloff test
is designed to achieve early radial flow). During
Normalized pressure differernce

Linear flow Radial


flow the calibration test, the linear flow period is identi-
and pressure derivative

0.1
fied from the diagnostic log-log plot (Fig. 9-41)

for estimating either the closure time or spurt.
1 Obtaining representative values of these param-
2
eters generally requires a good prior estimate of
0.01 the initial, undisturbed reservoir pressure.
In the absence of spurt, the closure time can
be estimated as the smallest value of tc that, when
used with the independently derived reservoir
1
1
pressure estimate, provides the linear flow diag-
0.001
1 10 100 nostic discussed in the preceding “Flow regime
1/ {FL(t/tc)}2
identification” step. If spurt is expected, the value
of tc should be independently determined using the
Figure 9-41. Identification of linear and radial flow using procedures outlined in “Estimating closure pres-
the diagnostic log-log plot. sure” in the Appendix to this chapter. The after-
closure linear flow analysis can then be used to
estimate the spurt, with the linear flow period ini-
factor of 2 and each has a half-slope. Radial flow
tially identified from the diagnostic log-log plot.
is characterized by a log-log slope that approaches
If linear flow is identified, a Cartesian plot of the
unity as well as an approximate overlying of both
pressure versus FL(t/tc) is constructed to determine
the pressure difference and pressure derivative
the slope mlf. Equation 9-94 is used to obtain the
curves.
spurt factor κ. The reservoir mobility k/µ required
• Reservoir parameters determination by this equation is derived from transitional or
The testing sequence should preferentially include radial flow analysis and CL from the decline analy-
a mini-falloff test to characterize the reservoir sis. The spurt coefficient Sp can then be obtained
parameters. Shut-in pressures monitored during the from Appendix Eq. 24.
mini-falloff test should initially be investigated for • Crossvalidation
radial flow using the log-log diagnostic described
Equation 9-101 is applied to compare the fracture
in the preceding step. If radial flow can be identi-
length predicted by the reservoir response to that
fied, a Cartesian plot of the pressure response dur-
derived from decline analysis (see Section 9-5) for
ing radial flow versus {FL(t/tc)}2 is constructed.
the calibration test. The radial flow slope mrf is
The slope of the straight-line portion on this graph,
anticipated from Eq. 9-93, with the reservoir trans-
or mrf, can be used with Eq. 9-93 to estimate the
missibility as previously determined from the
reservoir transmissibility. The pressure intercept of
mini-falloff test and Vi and tc defined as the vol-
the straight-line portion is an estimate of the initial
ume injected and closure time for the calibration
reservoir pressure. This procedure is analogous to
test, respectively. Agreement between the two
the conventional Horner analysis used for well
independently inferred fracture lengths indicates
testing.
correct evaluation of the calibration treatment.
In the absence of radial flow, the transitional
flow-based analysis described in “Comparison
of fixed-length and propagating fractures” in the 9-6.9. Field example
Appendix to this chapter should be applied to the
mini-falloff test. This type-curve matching proce- Application of the after-closure analysis methodology
dure estimates the initial reservoir pressure and its is illustrated by analyzing the pressure monitored
transmissibility based on the transitional flow during the field calibration tests in Fig. 9-10. Analysis
period. of the injection pressure for the calibration test

Reservoir Stimulation 9-57


described in “Example of radial fracture growth”
1000
in Section 9-4.4 determined that the fracture is best

Pressure difference and pressure derivative (psi)


Pressure difference
described by the radial fracture geometry model. The Pressure derivative
fracture is approximated as having a rectangular shape
of equal area to facilitate the after-closure analysis.
The shut-in pressure is analyzed in “Example field
application of radial fracture decline analysis” in
Section 9-5.4, from which the leakoff coefficient CL is
calculated to be 1.62 × 10–2 ft/min1/2. A closure pres-
sure of 4375 psi was inferred from the step rate test
(see “Estimating closure pressure” in the Appendix
to this chapter).
1
• Background information
1
The stabilized pressure on the bottomhole gauge 100
is 3726 psi in Fig. 9-10. The stabilized pressure 1 10
measurement provides an independent, objective 1/{FL(t/tc)}2
assessment of the reservoir pressure. A log-deter-
mined porosity of 19% and a saturation-weighted Figure 9-42. Identification of after-closure radial flow.
formation compressibility of 8.0 × 10–5 psi–1 were
also obtained. PVT analysis indicated a relatively
4400
high oil viscosity of 4 cp.
4300
• Radial flow identification 4200 660
Pressure (psi)

After-closure pseudoradial flow is observed in the 4100


reservoir for the initial mini-falloff test (Fig. 9-10), 4000
as shown by the diagnostic log-log plot in Fig. 9-42. 3900
A Cartesian plot of the pressure versus {FL(t/tc)}2 3800
during the after-closure period is in Fig. 9-43. An 3700 3736 psi

initial reservoir pressure estimate of 3736 psi is 3600


obtained as the y-axis intercept of the straight-line 0 0.2 0.4 0.6 0.8 1.0
period on the plot in Fig. 9-43. This value indi- {FL(t/tc)}2
cates consistency with the previously inferred esti-
mate of 3726 psi from the bottomhole gauge. The Figure 9-43. After-closure radial flow analysis.
slope mrf of the straight line is 660 psi and the clo-
sure time tc is 3.75 min, based on a closure pres- • After-closure linear flow determination
sure of 4375 psi (see “Estimating closure pressure”
in the Appendix to this chapter). A total volume of The shut-in pressure measured following the cali-
14.75 bbl was injected during the test. The trans- bration test is investigated for after-closure linear
missibility is obtained by using this information flow behavior. The observed potential period of
with Eq. 9-93 reformulated in dimensional form: linear flow is confirmed by pressure derivative
analysis (Fig. 9-44). The initial pressure pi of
kh Vi 1 3724 psi required for the analysis is in excellent
= 2.5 × 10 5
µ tc mrf agreement with the 3736-psi reservoir pressure
14 .75 1 derived from the radial flow analysis and its value
= 2.5 × 10 5 (9-102) of 3726 psi inferred as the stabilized pressure mea-
3.75 (660)
= 1490 md - ft / cp. surement on the bottomhole gauge prior to injec-
tion (Fig. 9-10). The linear flow slope mlf is
The transmissibility estimate implies a reservoir deduced to be 815 psi from the corresponding
permeability of 80 md based on the reservoir para- Cartesian plot.
meters.

9-58 Fracture Evaluation Using Pressure Diagnostics


the conditions of high viscosity and low compress-
Pressure difference and pressure derivative (psi) 1000 ibility, as in this example.
Pressure difference
Pressure derivative • Reconciliation
Equation 9-80 is used to obtain the apparent frac-
ture length faL:

π2  1.02 
faL =   = 0.86.
100

(1 − 0.16)  π 2 × (1.02 − 1) + 2 5.6 4.6 
1

2 The slope mrf for the calibration injection is pre-


dicted using Eq. 9-102, the dimensional form of
Eq. 9-93:
Vi  µ 
mrf = 2.5 × 10 5 ×
10 tc  kh 
1 10 100
= 2.5 × 10 5 ×
107  1 
1/{FL(t/tc)}2
5.6  1490 
= 3205 psi.
Figure 9-44. Identification of after-closure linear flow.
The fracture length is estimated using the
dimensional form of Eq. 9-100:
• Spurt estimation
The previously determined linear and radial flow 1 ktc m rf
L = 2.37 × 10 −3
slopes are used with the estimates of formation faL φµct mlf
porosity and compressibility to estimate spurt with 1 80 × 5.6 3205
Eq. 9-94, which is reformulated in oilfield units as = 2.37 × 10 −3
0.86 0.19 × 4 × 8.0 × 10 −5 815
4 mlf kφct  t = 33 ft.
κ = 1+ 1.18 × 10
−3
− 1 c
π CL µ  tp The length estimate is in good agreement with
the 37-ft fracture radius determined from the shut-
1.18 × 10 −3 815  in analysis, as described in Section 9-5.4.
4  1.62 × 10 −2  5.6
= 1+  
π  80 × 0.19 × 8 × 10 −5  4.6
× − 1 9-7. Numerical simulation of
 4 
pressure: combined analysis
= 1.02 ,
of pumping and closing
where the closure time of 5.6 min was based on
The pressure analysis techniques presented in this
the closure pressure of 4375 psi determined with
chapter are based on analytical derivations. They
the step rate test (see “Estimating closure pressure”
assume primarily a simplified fracture behavior to
in the Appendix to this chapter).
provide an engineering approximation of the fracture
The spurt coefficient Sp is determined with
parameters. A more comprehensive analysis is
Appendix Eq. 24:
required to determine rational changes in the propped
Sp = 7.48 × 10 2 g0 CL t p (κ − 1) treatment design (e.g., fluid viscosity, leakoff coeffi-
cient changes because of additives) without the nega-
= 7.48 × 10 2 × ( π 2) × 1.62 × 10 −2 4.6 × (1.02 − 1)
tive impact of trial and error adjustments during com-
= 0.8 gal / 100 ft 2 . mercial field development. This understanding is
particularly important for reservoirs where formation
The nominal value of Sp, in spite of the high for- pressure capacity considerations (see Section 9-4.6)
mation permeability, reflects that the fluid leakoff require unconventional designs (Nolte, 1982, 1988a;
is governed by the reservoir. This can occur for

Reservoir Stimulation 9-59


Warpinski and Branagan, 1988) to produce the with any inverse problem is the occurrence of
desired fracture geometry. nonunique solutions. With respect to hydraulic frac-
These objectives are best achieved with an appro- turing, nonuniqueness arises when more than one
priately calibrated numerical simulator that adequately combination of the unknown parameters can be used
replicates the important processes that occur during a to approximate the observed pressure.
fracturing application. The simulator that best meets A multitude of parameters govern fracture behavior
these objectives is most likely not one of the compre- in a nonlinear and coupled fashion. Uncertainty in
hensive, computationally intensive simulators that are some of the parameters leads to erroneous prediction
better suited to research environments (e.g., Plahn et of the unknown parameters inferred by the pressure-
al., 1997). The description of the overall fracture matching process. A nonunique parameter set that
behavior required for routine engineering applications produces an incorrect prediction of the fracture geom-
can generally be satisfied by P3D fracture geometry etry results in an inappropriate redesign of the
models (see Section 6-3). propped treatment. This effect is readily apparent
in Eq. 9-24, in which the pressure response during
pumping exhibits an approximately first-order inverse
9-7.1. Pressure matching dependency on the fracture compliance cf. For the
A typical approach for calibrating a numerical frac- commonly occurring PKN-type fracture, cf ≈ hf /E′
ture simulator is through iterative pressure matching. (Eq. 9-22). Consequently, an inaccuracy in the value
The fracture simulator is executed with varying para- of the plane strain modulus E′ produces a correspond-
meter inputs to visually match the simulated pressure ing error in the predicted fracture height hf and there-
with the field measurements. The quality of the pres- fore the fracture penetration, without affecting the
sure match is assessed by balancing objectivity, phys- quality of the pressure match.
ical insight and constraints from other data sources. Nonuniqueness in pressure matching may also
However, given the limited number of formation occur during periods of controlled and uncontrolled
parameters that can be accurately characterized using height growth. During both these conditions, the net
wireline logs or laboratory tests, the procedure can pressure response is determined largely by the stress
require significant effort when several parameters are difference between the bounding zones and the pri-
not known. A large number of unknown parameters mary interval. The pressure response exhibits a much
may also compromise the objectivity of the matching reduced sensitivity to other fracture parameters, such
procedure. as E′ or CL. Consequently, a reasonable pressure
An alternative approach for pressure matching match during these periods may be attained in spite
relies on an algorithm that repeatedly executes the of errors in the other fracture parameters.
fracture simulator to improve the estimates of the Nonuniqueness is avoided by preventing arbitrari-
fracture parameters within specified bounds until the ness during the pressure-matching process. This is
best match to the field-measured response is obtained. best achieved by initially applying the diagnostic pro-
This semiautomated approach, discussed in Section cedures outlined in this chapter to obtain a first-order
6-12, is based on conventional principles of nonlinear estimate of the unknown parameters and an approxi-
regression analysis for system identification. The mate understanding of the nature of the fracture
technique, initially introduced by Gringarten et al. growth. Only moderate variation of the unknown
(1979) for well test analysis, has also been applied parameters from these initial estimates should be
to log interpretation and reservoir characterization. allowed during the pressure-matching process. In
addition, the numerical simulator must be adequately
constrained so that the simulated fracture geometry
9-7.2. Nonuniqueness conforms to the growth patterns inferred by the pres-
sure diagnostics; e.g., the simulator must predict con-
Pressure matching is an inverse problem for which trolled height growth if this behavior was concluded
the input (i.e., fracture treatment schedule) and output from pressure derivative analysis.
(i.e., fracture pressure) are known and are used to The pressure-matching process should also include
determine the unknown model parameters (e.g., zone all independently derived information. For example,
stresses, leakoff coefficient). An inherent limitation nonuniqueness resulting from cf could be overcome

9-60 Fracture Evaluation Using Pressure Diagnostics


by a rational selection of the primary zone height
using well logs. Young’s modulus E can then be cali-
brated by matching pressures during the stage of the

Pressure/injection rate
Step rate/
treatment that exhibits PKN behavior, before penetra- Mini-falloff flowback/rebound Calibration
tion of the bounding layers occurs. Alternatively, the
value of E could conform to core tests or sonic logs
of the reservoir (see Chapter 4), and the primary zone
thickness could be calibrated with the pressure match
and verified against well logs. Similarly, variation in
the zone stresses must be based on their trends as pre-
dicted by stress logs. In the absence of stress logs, Time
lithology logs can be used to develop an approxima-
tion for particular geologic environments (Smith et
Figure 9-45. Formation calibration testing sequence.
al., 1989). Furthermore, the simulated fracture
dimensions could be compared against independent
measurements, such as radioactive tracers or micro- process based on these tests.
seismic measurements, to validate the outcome of the 1. Review existing information pertinent to fracturing
pressure-matching exercise (Gulrajani et al., 1998). A properly calibrated sonic log (see Chapter 4)
The inversion solution is confirmed by applying the provides an excellent basis for assessing the for-
numerical simulator based on these calibrated fracture mation stress profile. It can also be used to identify
parameters to predict the fracture behavior during the lithologic aggregates for developing an under-
propped treatment or for offset wells. An objective standing of the zone layers. In its absence, a stan-
pressure evaluation is indicated by agreement between dard gamma ray or spontaneous potential log
the simulated pressure response and actual field mea- should be applied to define the depth and thickness
surements. Conformity also establishes the ability of of the layers. The permeable zone thickness (i.e.,
the calibrated numerical simulator to develop the thickness over which fluid loss occurs) can be
improved designs for future fracture treatments. identified using specialized magnetic resonance
logs that directly measure permeability. This infor-
mation can also be obtained from the separation
9-8. Comprehensive calibration of various resistivity measurements, which indi-
test sequence cates drilling fluid invasion and therefore a finite
permeability.
An optimized hydraulic fracturing program integrates
An estimate of the formation stresses can be
the analyses and diagnostic methods discussed in this
derived by history matching prior fracture treat-
chapter with the proper design process (see Chapters
ments in offset wells. This information can be used
5 and 10) and execution aspects (see Chapter 11) of
to characterize the fluid leakoff behavior as well
the stimulation procedure. These considerations
as calibrate either sonic or gamma ray logs (Smith
include identifying candidate wells that will benefit
et al., 1989) to obtain the stress variation over the
from the stimulation treatment, determining the
depth of interest. Where available, laboratory core
appropriate injection rate and proppant addition
tests should be used to obtain related information
sequence as well as assessing the risks involved with
(e.g., mechanical properties, existence of natural
treatment execution. The techniques described in the
fissures, fluid-loss behavior).
previous sections provide a basis for estimating the
fracture and reservoir parameters required during the 2. Determine reservoir production parameters
optimization process. To ensure completeness and The reservoir pressure and permeability are best
consistency in the evaluation process, fracturing pres- determined from a well test. In its absence, the
sure data should be obtained through a planned information can be obtained using the mini-falloff
sequence of calibration tests, as shown in Fig. 9-45. test. The mini-falloff test attempts to propagate a
The following steps describe the fracture evaluation short, inefficient fracture to attain the required

Reservoir Stimulation 9-61


reservoir transitional or pseudoradial flow (see 6. Characterize fluid leakoff
Section 9-6) within an acceptable data-recording A pressure decline analysis performed during the
period. This objective is best achieved by applying calibration test closing period and the previously
the operational guidelines provided in “Guidelines calibrated fracture compliance produce the most
for the field application of after-closure analysis” reliable estimate of the fluid-leakoff parameters
in the Appendix to this chapter. Using an a priori (see Section 9-5). They also determine the treat-
estimate of the reservoir pressure reduces uncer- ment efficiency, using an estimate of spurt from
tainty in the after-closure analysis. the after-closure linear flow analysis. The fluid-
Pore pressure and permeability are the bases of loss parameters are used to provide a measure of
the reservoir’s producing potential and are required the fracture length (see Section 9-5.2).
for an economics-based design (see Chapters 5 and The shut-in pressure during the closing of a cali-
10). bration test can also be used to infer the closure
3. Define closure pressure pressure (see “Estimating closure pressure” in the
The preferred method for objectively determining Appendix to this chapter). This analysis, however,
the fracture closure pressure is a step rate/flowback is highly subjective, and caution should be exer-
test (see “Estimating closure pressure” in the cised during its application.
Appendix to this chapter). For low to moderate 7. Assess after-closure response
permeabilities, the step rate test typically follows Calibration test shut-in pressure should be moni-
the mini-falloff test (Fig. 9-45). In higher perme- tored for a duration that is at least twice that of the
ability formations, it follows the calibration test closure time tc. This enhances the prospect that a
because the subsequent loss of completion fluid majority of the recorded after-closure response
with which the step rate test is performed is signif- period is not masked by complications arising
icantly reduced relative to fluid loss before the cal- from polymer injection and that a well-established
ibration test. linear flow response is obtained. The pressure dur-
4. Characterize fracture geometry ing the linear flow period is used to estimate spurt
The character of fracture growth is identified dur- (see Section 9-6.8), which in conjunction with the
ing a calibration test by the analysis of pressure shut-in analysis provides the treatment efficiency.
during pumping (see Section 9-4). The analysis Linear flow may be completely absent in a high-
also defines the formation pressure capacity (see mobility reservoir.
Section 9-4.6) and can be used to calibrate the The operational guidelines outlined in “Guide-
fracture compliance through pressure matching. lines for the field application of after-closure
The injection pressure is used to determine the analysis” in the Appendix to this chapter should be
bounding stresses (Eq. 9-39). This additional stress applied to prevent additional contaminating effects
measure, in conjunction with an estimate of the on the after-closure pressure data, which could
closure pressure, provides the basis for a linear lead to an inconclusive linear flow analysis.
calibration for correcting a sonic-based stress log 8. Crossvalidate evaluation results
for the particular lithologic setting (Nolte and Information derived from the after-closure linear
Smith, 1981; Nolte, 1988a). flow response, in combination with the previous
5. Identify near-wellbore problems estimate of reservoir transmissibility, provides an
A short step-down test (see Sidebar 9E) at the end independent perspective for the fracture length
of the injection period can be used to identify the (see Section 9-6.8). Agreement between this length
magnitude and severity of near-wellbore problems. estimate and that determined from the preclosure
Net pressure analyses (e.g., log-log plots) should analysis indicates consistency in the pressure
be based on the bottomhole pressure corrected to analysis.
reflect measurements of near-wellbore pressure
that is not associated with the fracture behavior.

9-62 Fracture Evaluation Using Pressure Diagnostics


9. Conduct pressure history matching 10. Verify against propped treatment pressure
The previously obtained fracture parameter esti- response
mates (i.e., zone mechanical parameters and fluid The evaluation can be further refined for future
properties) can be refined through history matching treatments by matching pressures during the
the simulated pressure with the pressure measured propped treatment and comparing the required
during the calibration test. Care should be taken, parameter set for this treatment with that obtained
however, to prevent the occurrence of nonunique- from the calibration treatment. Conformity
ness, an inherent limitation in pressure matching between these two parameter sets provides confi-
(see Section 9-7.2). This shortcoming can be pre- dence in the fracture parameters. In the case of
vented by performing a comprehensive evaluation disagreement, a reanalysis should be conducted
and testing procedure to limit the number of with emphasis on deviations from ideal behavior.
unknowns as well as eliminate unrealistic solutions. A more authoritative measure of the correctness
of the predicted parameters is agreement among
parameter sets obtained in offset wells, with suit-
able corrections made to account for differences
in the formation thicknesses, well completions,
fracturing schedules or fracturing fluid systems.

Reservoir Stimulation 9-63

You might also like