Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Appendix: Background for


Hydraulic Fracturing
Pressure Analysis Techniques
Sunil N. Gulrajani and K. G. Nolte, Schlumberger Dowell

Estimating closure pressure is created, and the shut-in pressure is commonly


used as a first-order approximation of the stress.
The distinction between the formation closure pres- Significantly higher net pressure, however, occurs
sure and the local stress is discussed in Sidebar 9A. with the procedures employed to determine pc. Thus,
The closure pressure pc is distinguished from the pc can be considerably different from the instanta-
minimum stress σmin because σmin generally varies in neous shut-in pressure (ISIP), and it must be esti-
magnitude and direction over the gross pay interval mated with alternative procedures.
(i.e., the zone between meaningful stress barriers) Techniques commonly used to determine pc are
whereas pc provides a global average for the interval. the step rate, shut-in decline and flowback tests
Field procedures for estimating these two stress (Appendix Fig. 1) (see Section 3.6-2). Step rate and
measures differ in two primary ways. First, estimat- flowback tests are conducted exclusively to deter-
ing the magnitude of the local stress requires the cre- mine pc. The shut-in decline test is used in conjunc-
ation of a small fracture by using a relatively small tion with a calibration test, which is usually con-
fluid injection rate and volume. Determining pc ducted to quantify fluid-loss behavior (see Section
requires the creation of a fracture over the entire 9-5). The value of pc can also be estimated from
thickness of the gross pay interval, which requires injection pressure derivative analysis (see Sidebar
a much larger injection rate and volume. Second, a 9C) or after-closure linear flow analysis (see Section
smaller net pressure occurs when a smaller fracture 9-6).

(a) (b)
Bottomhole pressure
Injection rate

at step end

pc

Extension
pressure

Time Injection rate


(c) (d)

Pump-in
Bottomhole pressure

Bottomhole pressure

Flowback

a pc

pc
b
Possibilities
c

Time √∆t or √∆t + tp

Appendix Figure 1. Tests to determine closure pressure (after Nolte, 1982): (a) step rate test, (b) bottomhole pressure plot-
ted versus injection rate to infer the values of pc and the fracture extension pressure, (c) combined step rate and flowback
tests (a = rate too low, b = correct rate for pc at curvature reversal and c = rate too high) and (d) shut-in decline test dis-
played on a square-root plot (∆t = shut-in time and tp = injection time into the fracture).

Reservoir Stimulation A9-1


Step rate test
9200
The step rate test is conducted solely to determine pc. pc = 8910 psi

Bottomhole pressure (psi)


In low-permeability reservoirs, the test generally is 8900
performed with completion fluids (e.g., treated
Extension pressure = 8950 psi
water). The use of polymer fluids may be required in
8600
high-permeability reservoirs (Smith, 1985) to control
potential fluid loss and ensure fracture creation at
lower injection rates. 8300
The step rate test can be conducted as the pumping
phase of either the flowback or shut-in decline test. 8000
The time duration of the individual injection steps 0 3 6 9
Injection rate (bbl/min)
should be equal (Appendix Fig. 1a) and can be rela-
tively small (i.e., the time required for the pumps to
Appendix Figure 2. Pressure versus rate analysis for the
change and maintain a constant rate and the pressure step rate test in Fig. 9-8.
to be recorded, typically 1 or 2 min). The injection
rate increments over the successive steps should also
be approximately the same. When the step rate test is should not be used in preference to also obtaining a
conducted as part of a flowback or decline test, the well-defined straight line for matrix injection, which
last step is maintained for a longer duration (i.e., 5 is the only means to ensure that fracture extension
to 10 min) to ensure the creation of a sufficient-size occurred during the test.
fracture. Injection rates typically vary from 1 to The y-axis intercept interpretation was used to ana-
10 bbl/min in moderate-permeability formations, and lyze the step rate test illustrated in Fig. 9-10b. The
the rates in low-permeability formations are about pressure versus rate plot is entirely characterized by
one-half of these values. Ideally, the injection rates the fracture extension line (Appendix Fig. 3), which
for three of the steps should fall below the extension has a y-axis intercept of 4375 psi for the estimated
pressure to define matrix injection prior to fracturing, value of pc.
and a similar set of values should be obtained above An indication of a valid step rate test is that the
the extension pressure. extrapolated pressure for the zero rate of the line rep-
The fracture closure and extension pressure are resenting matrix injection should equal approximately
inferred from a crossplot of the bottomhole pressure the bottomhole pressure preceding the test. The zero-
at the end of each injection step and the injection rate rate pressure is the reservoir pressure, if significant
(Appendix Fig. 1b). The plot is characterized by two amounts of fluids were not previously injected. This
straight lines, with matrix injection represented by the quality control check can also be used while conduct-
line with the steeper slope and fracture extension ing the test when no apparent slope change is
characterized by the shallower sloping line. The inter- observed for the pressure versus injection rate plot.
section of the two lines provides an estimate of the
fracture extension pressure, which is the upper bound 4800

for pc. Typically, the extension pressure is 50 to


Bottomhole pressure (psi)

4700
200 psi greater than pc, as shown on the pressure
versus rate plot in Appendix Fig. 2 for the step rate 4600
test in Fig. 9-8b. This larger value represents effects
from fluid friction pressure within the fracture and 4500
resistance to fracture extension (i.e., toughness).
Laboratory tests (Rutqvist and Stephansson, 1996) 4400
pc = 4375 psi
also indicate that the y-axis intercept of the shallower
4300
sloped line that represents fracture extension on the 0 5 10 15 20 25
crossplot provides a first-order approximation for pc, Injection rate (bbl/min)
even when the steeper line that represents matrix
injection is absent. This interpretation, however, Appendix Figure 3. Pressure versus rate analysis for the
step rate test in Fig. 9-10.

A9-2 Chapter 9 Appendix: Background for Hydraulic Fracturing Pressure Analysis Techniques
Shut-in decline test ful fraction of the in-situ leakoff rate). With the
assumption that a fracture has been created, the pres-
The shut-in decline test can be used with either a step sure response during flowback has two distinctly dif-
rate or calibration test. The decline data are displayed ferent profiles while the fracture is closing and after
on a square-root plot (Appendix Fig. 1d) or a G-plot the fracture closes. Comprehensive simulations (Plahn
(Fig. 9-29) that assumes square-root exposure time et al., 1997) indicate that the fracture pc is identified
for the fluid-loss behavior. The closure pressure is by the intersection of the two straight lines that define
inferred where the slope changes on either plot. The these two periods. The increasing rate of pressure
derivative should be used to magnify the change of decline for the postclosure period results from fluid
slope and enhance its identification. flow through the pinched fracture width (i.e., induced
Either of the specialized plots, however, may be fluid choking) in the near-wellbore region induced by
completely devoid of a significant slope change or fluid flowback. The characteristic “lazy-S” signature
may exhibit multiple slope changes. In general, up exhibited by the pressure during the flowback period
to six events could be associated with a slope change: is in contrast to the multiple inflections commonly
• height recession from the bounding layers observed with the shut-in decline test. Therefore, the
• transition between fracture extension and recession flowback test provides a more objective indication of
closure relative to the decline test.
• fracture closure
Maintaining a constant flowback rate as the pres-
• postclosure consolidation of the polymer filter cake sure decreases is critical for a flowback test. This
and fracture face irregularities objective requires a field-rugged, debris-resistant flow
• reservoir linear flow regulator that both measures and controls the flow-
• reservoir radial flow. back rate. The flow regulator should be preset for the
desired rate at the pressure expected following the
Consequently, the shut-in test commonly fails to end of injection, and it should be isolated by a closed
provide an objective indication of pc and should not valve during pumping. Presetting the flow regulator is
be used as the primary procedure for determining it. best achieved by opening it during the last period of
Experience indicates that the square-root plot may the step rate test to establish the desired flowback rate
provide a better indication of closure for fluids that prior to the actual test. The effect of the additional
do not have effective fluid-loss control from wall- fluid loss can be compensated for by increasing the
building behavior, whereas the G-plot may provide a injection rate. Fluid injection is terminated once the
better indication for fluids with wall-building behav- desired constant flowback rate has been attained, and
ior. The analysis of decline data typically uses both this rate is then maintained throughout the flowback
plots to determine the value of pc. period.
Other specialized plots have been used, although Experience shows that an adjustable choke often
less frequently, to identify pc. These include the log- plugs because of pipe dope and other debris loosened
log shut-in plot (Elbel et al., 1984) and multidimen- into the wellbore during the injection period. A gate
sional derivative analysis. In contrast to the square- valve is preferable for controlling the flowback rate.
root and G-function plots, the interpretation In addition, a pressure sensor and fixed choke at the
philosophy of these plots is based on identifying end of the flowback line can be substituted for a flow
reservoir flow regime changes to obtain bounding meter to reliably measure the flow rate, particularly
values of pc. when the rate is low (i.e., 3 bbl/min or less). Tabu-
lated values for the pressure drop versus the flow rate
through standard choke sizes are used to select the
Flowback test choke size that will provide the pressure change nec-
The preferred method for determining pc is a combi- essary for a reliable pressure measurement at the
nation of the step rate test (with an extended last step) anticipated flowback rate.
and flowback test (Appendix Fig. 1c). The essential The flowback test is frequently repeated for verifi-
element of the flowback test involves a flowback cation and selection of a more optimum flowback
period at a constant rate that is between 1⁄6 and 1⁄4 of rate. The first flowback period should be of sufficient
the last injection rate (i.e., at a rate that is a meaning- duration to ensure that fracture closure in the primary

Reservoir Stimulation A9-3


zone is recorded. Pressure during this test, however, For moderate to high fluid-loss conditions, the cali-
should be controlled to prevent the production of bration test generally exhibits a closing period that is
reservoir fluids into the wellbore. The second flow- less than the injection time, and prior estimate of pc is
back test may be preceded by either a step rate test or not required for a shut-in period that approximates
a constant injection rate and should end after the indi- the injection period. Conducting the step rate/flow-
cation of a curvature reversal that clearly confirms the back test after the fracture calibration test reaps the
pc estimate from the first test. following advantages from the prior fluid injection:
The pressure decline for the second test is limited
• Breakdown of the complete gross pay interval is
to obtain an optimum rebound pressure. The rebound
ensured.
pressure is the nearly constant pressure that occurs
following a short period of increasing pressure after • Residual fluid-loss control from unbroken polymer
shut-in of the flowback test (see Fig. 3-26b). This sta- or additives in the fracturing fluid enables fractur-
bilized pressure is generally near to or smaller than pc ing using moderate injection rates with the com-
and provides its lower bound estimate. The rebound pletion fluid.
pressure is also an effective tool for estimating pc • Because the poroelastic effects have stabilized, the
when near-wellbore flow restrictions, as discussed in step rate/flowback test better replicates the stress
Section 9-4.9, are large. During the flowback period, conditions that occurred during the preceding cali-
these near-wellbore effects cause prediction of a bration treatment.
lower estimate of pc because of the additional flow
restriction in the near-wellbore region. The wellbore
rebound pressure, however, is unaffected by the
restriction and should provide an accurate lower
Guidelines for the field application
bound estimate of pc (Plahn et al., 1997). of after-closure analysis
In conclusion, the combination of the upper bound The reservoir response during linear flow and either
estimate of pc from the intersection of the matrix and radial or transitional flow is required to conduct a
fracture extension lines on a step rate test, the lower comprehensive analysis of the after-closure period.
bound of pc determined from the rebound pressure Several factors, however, can prevent obtaining infor-
and the estimate of pc from the y-axis intercept of the mation for one or more of the flow regimes. For
fracture extension line as well as the intersection of example, radial flow may be not be attained within a
the two lines during flowback provides multiple, reasonable time period for the high injection rates and
independent values that establish a firm basis for leakoff-resistant fracture fluids used during conven-
defining pc. tional fracturing operations. In addition, the after-
closure pressure data may be compromised because
of several undesirable communication effects
Test sequence between the wellbore and formation. Therefore, the
Because the mini-falloff test (see Section 9-6.2) that calibration testing sequence must be engineered to
characterizes the production potential should be con- provide the requisite pressure data (Talley et al.,
ducted in an initially undisturbed reservoir, it should 1999). The following guidelines increase the likeli-
always be the first injection within the calibration hood of obtaining a comprehensive, objective after-
testing sequence (Fig. 9-45). The subsequent applica- closure analysis.
tion sequence of the step rate/flowback and calibra-
• Unless a bottomhole shut-off valve is employed,
tion/shut-in tests generally depends on the formation
the reservoir pressure should ideally be equal to
fluid-loss characteristics.
or greater than the hydrostatic pressure of the well-
For low fluid-loss conditions, the step rate/flow-
bore fluid. Vacuum-induced fluid injection violates
back test to establish pc should precede the fracture
the no-flow assumption of the analysis for the
calibration test where the closure time could be long
after-closure period.
(e.g., up to 3 times the injection time). This prior esti-
mate of pc can be used to ensure that the shut-in pres- • The wellbore must be free of gas to ensure that
sure during the subsequent calibration test is moni- correct values of the hydrostatic pressure and
tored during the complete time of fracture closure. injected volume are used in the analysis and to

A9-4 Chapter 9 Appendix: Background for Hydraulic Fracturing Pressure Analysis Techniques
minimize wellbore expansion effects during pres- • It is unlikely that both linear and radial flow will
sure falloff. For gas reservoirs, this can be achieved occur during a decline period (see Section 9-6.2).
by conducting the injection test before a production The testing sequence illustrated in Fig. 9-45 is sug-
period. Alternatively, the gas should be circulated gested to increase the likelihood of obtaining infor-
from the wellbore or bullheaded into the formation. mation pertinent to both the linear and radial flow
An extended shut-in period can be required after periods. The mini-falloff test should be applied to
bullheading to allow the pressure transient to dissi- determine the radial flow parameters. The subse-
pate before liquid injection resumes. A relatively quent calibration test is more likely to attain linear
small volume of gas injection ahead of the fluid is flow. The radial flow response for the calibration
acceptable—e.g., gas is circulated from the inter- test can be anticipated from the reservoir informa-
mediate casing but remains in the shorter liner. tion derived from the mini-falloff test (see “Frac-
• Circulating or bullheading (potentially with a long ture length” in Section 9-6.7).
shut-in period) may be similarly required to spot • To attain radial flow within a reasonable time
the fracturing fluid at the perforations for the frac- frame, the mini-falloff test should adhere to the fol-
ture calibration test. The residual reservoir response lowing injection rate criterion, presented in con-
from injection of a significant volume of low- ventional oilfield units:
efficiency wellbore fluid causes the linear flow
qi ( bbl / min ) k ( md ) CL
analysis to indicate an unrealistically high spurt
h f (ft )
≤ 4 × 10 −6
µ(cp) CR
( pc − pi (psi)).
loss (Talley et al., 1999).
• Like preclosure analysis, after-closure analysis is (1)
an inverse problem that is inherently nonunique
If the fluid loss is controlled by the reservoir, as
(see Section 9-7.2). The objectivity of after-closure
desired for the test, the ratio of the fluid-loss coeffi-
analysis can be improved by obtaining an a priori
cients CL and CR becomes unity and a higher injec-
estimate of the reservoir pressure, particularly if
tion rate is possible. The equation provides an
the after-closure period is abbreviated (e.g., before
equality for a dimensionless time of 1.0 (i.e., the
wellbore vacuum) and either the closure time
beginning of radial flow during the injection period
(Nolte et al., 1997) or spurt loss is inferred. The
on Fig. 9-38). The guideline requires using esti-
reservoir pressure estimate can be obtained
mates of the reservoir parameters and fluid-loss
– as the stabilized bottomhole pressure measured characteristics to design the mini-falloff test. In
prior to fluid injection into the reservoir general, Appendix Eq. 1 provides an operationally
– as the stabilized surface pressure measured prior reasonable rate for radial flow with a short moni-
to fluid injection into an overpressured reservoir toring period in a reservoir with a mobility greater
– from the surface pressure and hydrostatic col- than about 5 md/cp. For reservoirs with lower val-
umn estimated through an accurate measurement ues of mobility, transitional flow resulting from
of the fluid injected to completely fill the well- injection rates greater than guideline can be used
bore for an underpressured reservoir to determine the reservoir parameters with a type-
curve-based analysis.
– from an accurate reservoir pressure gradient
established for the field. • Volume has a minimal effect on dimensionless time
and hence the time for development of radial flow
• In deep or hot reservoirs, bottomhole gauges are
because of the quasistationary value of dimension-
necessary because wellbore fluid expansion from
less time for a constant injection rate. However, a
the decreasing pressure and increasing temperature
minimum volume must be pumped to ensure an
during shut-in decrease the hydrostatic pressure.
accurate measure of the volume injected through
Excessive expansion of the fluid may eventually
the perforations because the transmissibility is pro-
violate the no-flow condition to the degree that the
portional to the injected volume (Eq. 9-93).
longer term data are corrupted, particularly for
residual gas in the wellbore. Mitigation of these • If polymer fluids are used (see Section 9-6.3) the
effects, like for a wellbore vacuum, requires a pressure data obtained after fracture closure can be
downhole shut-in device. corrupted by continued consolidation (i.e., squeez-

Reservoir Stimulation A9-5


ing) of the fracture faces and filter cake. The con- treatment efficiencies are derived subsequently in this
solidation period lasts for a time approximating the Appendix. The exponent generally decreases through-
combined injection and closing times. The calibra- out the injection time, but the change is relatively
tion test shut-in pressure should therefore be moni- small and can be ignored. For the commonly used
tored for a time interval that is at least 2 to 3 times crosslinked polymer fluids, the exponent is typically 0.6.
the total closure time tc. Similarly, the shut-in Three mechanisms govern the fluid filtration process
period for a mini-falloff test should be at least 4 to during fracturing, as discussed in Section 6-4. The
5 times the total closure time. These guidelines for square root of exposure time relation, introduced by
the shut-in time increase the likelihood that ade- Carter (1957), is applied to predict the cumulative
quate, representative pressure data are obtained for effect of these mechanisms on the rate of fluid loss.
a valid after-closure analysis. A more general expression used in this Appendix
includes fluid-loss behavior dominated by a non-
Newtonian filtrate. Consequently, in contrast to Eq. 5-1,
the flux uL, defined as the rate of fluid loss per unit
Mathematical relations for fluid loss leakoff area, is expressed as
Derived in this section of the Appendix to Chapter 9
2 CL
are the relations used for pressure evaluation prior to uL = , (5)
fracture closure. (t − τ)1−θ
where t is the time since the beginning of pumping,
CL is the fluid-loss coefficient, and θ is the fluid-loss
Fluid-loss relations exponent. The coefficient 2 in Appendix Eq. 5
Two fundamental relations are required to derive the accounts for leakoff over the two walls of the fracture.
expressions for fluid loss. The first relation describes The specific fluid-loss volume vL, defined as the
the nature of fracture growth with respect to time. fluid-loss volume VL per unit leakoff area, is obtained
The second relation is based on the expression for by the time integration of Appendix Eq. 5:
fluid loss introduced by Carter (1957). t

(t − τ(a)) .
2 CL
vL = ∫ uL dt =
θ
The evolution of the fracture area is assumed to (6)
follow a power law relation with time in which the 0
θ
area monotonically increases with time. The proper- Parlar et al. (1995) showed that θ is related to the
ties of the injected fluid and the pump rate are power law exponent nf of the filtrate that invades the
assumed to be relatively constant. The power law reservoir during the leakoff process:
expression relates any intermediate fracture area a
created at a time τ to the total fracture area A at the nf
θ= . (7)
current time t: 1 + nf
α
a  τ As discussed in Chapter 8, a filter cake is deposited
= (2)
A t by a wall-building fluid along the fracture in low-per-
meability reservoirs and within the formation in high-
1/ α
τ  a permeability reservoirs. A Newtonian filtrate (i.e.,
= , (3)
t  A water) is created during the process. Under these con-
ditions, nf = 1 and Appendix Eq. 5 reduces to Eq. 5-1,
where α is referred to as the area exponent. The which is Carter’s square root of exposure time rela-
exponent α is also the log-log slope of A versus t, as tion for the fluid-loss rate (i.e., θ = 1⁄2). Non-wall-
shown by differentiating Appendix Eq. 3 with respect building fracturing fluids invade high-permeability
to time: reservoirs. The resulting filtrate fluid is typically non-
t dA Newtonian, with nf < 1 and θ < 1⁄2. In addition, the
α= . (4) extensional viscosity behavior of viscoelastic filtrates
A dt
above a threshold filtration rate can exhibit relatively
The value of α depends on the fluid efficiency. large values of nf (Chauveteau et al., 1986). The
Bounding values for α corresponding to low and high value of θ in this case is greater than 1⁄2 and can

A9-6 Chapter 9 Appendix: Background for Hydraulic Fracturing Pressure Analysis Techniques
approach unity. A value of θ that is different from ∆tD, which is defined as the ratio of the shut-in time
1
⁄2 has also been proposed to model the effect of ∆t to the pumping time tp:
natural fissures (Soliman et al., 1990).
t − t p ∆t
∆t D = = = tαD − 1. (15)
tp tp
Fluid-loss volume with the Carter-based
leakoff model The time functions f(tαD,α,θ) and g(tαD,α,θ) can be
expressed in terms of ∆tD:
The rate of fluid loss associated with Carter-based

1

f ( ∆t D , α, θ) = ∫
leakoff behavior (i.e., the contribution of CL over an
∆t D ≥ 0 (16)
(1 + ∆t − ξ1/ α )
1−θ
elemental leakoff area da) can be obtained by substi- 0 D
tuting Appendix Eq. 3 into Appendix Eq. 5:
1

g( ∆t D , α, θ) =
1
(1 + ∆t D − ξ1/ α ) dξ ∆t D ≥ 0 . (17)
θ

2r C da
q L ( da, t ) = 1p−θ L 1−θ , (8)
t p (tαD − ξ1/ α ) θ0

where rp is the ratio of the fracture surface area avail- Valkó and Economides (1993b) showed that the
able for fluid loss to the gross fracture area and tp is functions f(tαD,α,θ) and g(tαD,α,θ) are part of the
the injection or pumping time. The dimensionless hypergeometric family of functions or their subset
parameters tαD and ξ are defined as of incomplete beta functions (Meyer and Hagel,
1989). Either of these function families is relatively
t complicated, but simple analytical expressions can
tαD = (9)
tp be obtained for a limited set of values, as discussed
subsequently.
a
ξ= , (10)
Af
where Af is the fracture surface area at the end of
Cumulative fluid-loss volume
pumping. The total fluid-loss volume at the end of pumping
The total rate of fluid loss is obtained by the inte- VLp comprises the cumulative contributions of its
gration of Appendix Eq. 8 over the fracture area: CL fluid-loss component VLp,C and spurt VL,S:

VLp = VLp,C + VL ,S .
( )
2 rp CL A f (18)
qL Af , t = f (tαD , α, θ), (11)
t 1p−θ
An expression for VLp,C can be obtained by substi-
where the function f(tD,α,θ) is defined as tuting ∆tD = 0 into Appendix Eq. 13:
1
dξ VLp,C = VL ,C ( ∆t D = 0) = 2 rp CL t pθ A f g0 (α, θ), (19)
f (tαD , α, θ) = ∫ tαD ≥ 1. (12)
0 (t αD −ξ )
1/ α 1−θ
where the function g0(α,θ) represents the value of the
g-function in Appendix Eq. 17 when ∆tD = 0:
An expression for the CL component of the fluid-
1

(1 − ξ1/ α ) dξ .
loss volume VL,C is similarly obtained by substituting
g0 (α, θ) = g( ∆t D = 0, α, θ) =
1 θ

Appendix Eqs. 9 and 10 into Appendix Eq. 6 and θ0∫ (20)


integrating over the area:
Spurt occurs only for wall-building fluids. It ac-
VL ,C = 2 rp C t A f g(tαD , α, θ),
θ
L p
(13) counts for the fluid-loss volume prior to the creation
of a filter cake and is applicable only during the frac-
where the function g(tαD,α,θ) is defined as ture propagation period. Following the spurt period,
1 wall-building fluids exhibit a Newtonian filtrate during
g(tαD , α, θ) =
1
( tαD − ξ1/ α ) dξ tαD ≥ 1.
θ

θ0∫ (14) the fluid-loss process (i.e., nf = 1 or θ = 1⁄2). The total


volume of fluid lost to spurt follows from definition
The functions f(tαD,α,θ) and g(tαD,α,θ) are usually of the spurt coefficient Sp:
presented in terms of the dimensionless shut-in time

Reservoir Stimulation A9-7


VL ,S = 2 rp Sp A f . (21) θ = 1⁄2 in Appendix Eq. 41, whereas Appendix Eq. 40
provides the upper bound. An upper bound of unity is
The cumulative fluid-loss volume during injection assumed because it provides simple, closed-form
is obtained by substituting Appendix Eqs. 19 and 21 expressions. The bounding values of α for a wall-
into Appendix Eq. 18: building fluid are therefore as in Eq. 9-43:
1
VLp = 2 rp κCL t p A f g0  α, θ = 
1 1 < α < 1.
θ= (22)
 2 2 2
Analytical equations for f(∆tD,α,θ), g(∆tD,α,θ) and
1
VLp = 2 rp C t A f g0 (α, θ)
θ
L p θ≠ . (23) g0(α, θ) are obtained by substituting the bounding
2 values of α into Appendix Eqs. 16, 17 and 20,
Appendix Eq. 22 is expressed in terms of the spurt respectively, and integrating the resulting expressions
factor κ, which provides for the increase in fluid loss (Nolte, 1979; Smith, 1985):
over the no-spurt condition:
sin −1 1 + ∆t −1/ 2
( D)
1
 α=
f ( ∆t D ) θ= 1
Sp
κ = 1+ = 2
( )
. (24)
 1 2 (1 + ∆t D ) − ∆t 1D/ 2 α =1
1/ 2
g0 α, θ = CL t p
2

 
2
The rate of increase in the fracture area decreases (28)
significantly at the end of pumping. Spurt-dependent
 1 + ∆t sin −1 1 + ∆t −1/ 2 + ∆t 1/ 2
( D) ( D)
1
fluid loss therefore also reduces relatively quickly fol- α=
g( ∆t D ) θ= 1
D
lowing the cessation of fluid injection and is assumed = 4 2
to terminate at the time t = tp. The total fluid-loss vol-
ume during a shut-in period is represented entirely by
2
3
(
 (1 + ∆t D ) − ∆t D3/ 2
3/ 2
) α =1

its CL component VLs,C (∆t): (29)

VLs ( ∆t ) = VLs ,C ( ∆t ). (25) π 2 α=


1
g0 θ= 1 = 2 (30)
An expression for VLs(∆t) results from subtracting 2 4 3 α = 1.
Appendix Eq. 19 from Appendix Eq. 13:
All these functions are well behaved. Their values at
VLs ( ∆t ) = 2 rp CL t A f any other value of α can be obtained through simple
interpolation between bounding values.
 1 
× g ∆t D , α, θ =  − g0  α, θ =   θ =
1 1
  2   
2  2
Non-Newtonian filtrate control
(26)
A non-Newtonian filtrate occurs with fracturing fluids

(
VLs ( ∆t ) = 2 rp CL t θ A f g( ∆t D , α, θ) − g0 (α, θ)) θ ≠ .
1 that exhibit a power-law-based rheology and do not
2 develop an effective filter cake. In this case, nf ≠ 1
and if the filtrate controls fluid loss, θ deviates from
(27)
its commonly assumed value of 1⁄2. These conditions
limit the general analytical expressions to g0 only.
Explicit integration of Appendix Eq. 20 gives
Newtonian filtrate control
Fracturing fluids produce a Newtonian filtrate follow- Γ (1 + θ)Γ (1 + α )
g0 (α, θ) = , (31)
ing the deposition of a filter cake. For this case, nf = 1, θΓ (1 + α + θ)
and Appendix Eq. 7 indicates that θ = 1⁄2. Analytical
expressions for fluid loss can be derived for bounding where Γ(x) is the gamma function.
values of α. The lower bound value corresponds to Analytical expressions for the fluid-loss rate and
negligible efficiency and is obtained by substituting volume functions (i.e., Appendix Eqs. 16 and 17,

A9-8 Chapter 9 Appendix: Background for Hydraulic Fracturing Pressure Analysis Techniques
respectively) can be obtained for only select values The evolution of the fracture area for high fluid
of their arguments that are applicable to specific field efficiencies is obtained from this expression as
applications. The upper bound of α = 1 applies also qi t
for θ ≠ 1⁄2, and the corresponding functions are Af = η → 1, (39)
w

f ( ∆t D , α = 1, θ) =
1
θ
[
(1 + ∆t D )θ − ∆t Dθ ] (32) where 〈w〉 is the fracture width averaged over the

fracture area. The bounding expressions for fracture


penetration in Eqs. 9-32 through 9-34 are derived by
g( ∆t D , α = 1, θ) =
1
θ(1 + θ)
( [ 1+θ
]
1 + ∆t D ) − ∆t 1D+θ . (33) combining Appendix Eqs. 37 and 39 with the fracture
width relations in Eq. 9-27, for which the fracture
Simple analytical expressions for these functions area Af is proportional to the half-length L for the
also result for the value of θ approaching unity, PKN and KGD models and to R2 for the radial model.
which is applicable for time-independent fluid-loss For the common occurrence of Newtonian fracturing
conditions. This fluid-loss behavior can be exhibited fluid filtrate (i.e., θ = 1⁄2):
by viscoelastic fluids with a rapidly increasing exten- L ∝ t 1/ 2 η→ 0
sional viscosity above a threshold filtration rate: PKN
L ∝ t ( 2 n + 2 ) ( 2 n +3 ) η→1
f ( ∆t D , α, θ = 1) = 1 (34)
L ∝ t 1/ 2 η→ 0
KGD
g( ∆t D , α, θ = 1) =
1
+ ∆t D . (35) L ∝ t ( n+1) ( n+2 ) η→1
1+ α
The integral forms for both functions are well R ∝ t 1/ 4 η→ 0
behaved and therefore can be evaluated by numerical Radial
R ∝ t ( 2 n+2 ) ( 3 n+6 ) η → 1.
integration for other values of their arguments.
The time dependency of the net pressure pnet for the
bounding values of η is obtained by combining its
Bounding values for fluid efficiencies dependency on the fracture penetration for the three
The fracture penetration and net pressure can be rep- fracture geometry models in Eq. 9-25 with the three
resented for the bounding values of high and low preceding equations to produce Eqs. 9-35 through 9-37:
fluid efficiencies. When the efficiency η → 0, all the
pnet ∝ t 1/ 4 ( n+1) η→ 0
injected volume can be assumed to have leaked off. It PKN
follows from Eq. 9-5 that pnet ∝ t 1 ( 2 n+3) η→1

Vi ≈ VLp . (36) pnet ∝ t − n /2 ( n+1) η→ 0


KGD
From the relation Vi = qit and combining Appendix pnet ∝ t − n ( n+2 ) η→1
Eq. 36 with the generalized expression for fluid loss
during pumping in Appendix Eq. 23: pnet ∝ t −3n /8( n+1) η→ 0
Radial
pnet ∝ t − n ( n+2 ) η → 1.
qi t
Af = η→ 0 (37)
2 rp CL t θ g0
∝ t 1−θ . Bounding values for the area exponent
For the alternative situation of extremely high fluid The bounding values for the area exponent α corre-
efficiencies (η → 1), the fluid-loss volume can be spond to negligible (η → 1) and dominant (η → 0)
ignored. In this case, Eq. 9-5 reduces to fluid loss to the formation. The upper bound α1 in
Eq. 9-43 assumes not only negligible fluid loss but
Vi ≈ Vfp . (38) also a constant fracture width. Because the width
generally increases during injection, the upper bound

Reservoir Stimulation A9-9


for the area growth is less than unity and independent (
VLp ( ∆tso ) = 2 rp A f CL t p g( ∆t Dso ) − g0 , ) (46)
of the loss exponent θ. It is derived from the condi-
tion of η → 1 in Eqs. 9-32 through 9-34: where the dimensionless time ∆tDso is defined as the
ratio of the time after screenout ∆tso to that at the
(2 n + 2) (2 n + 3) PKN onset of screenout tso.

α1 = (n + 1) (n + 2) KGD η → 1. (40) The change in the area-averaged fracture width fol-
( 4n + 4) (3n + 6) Radial lowing a screenout is obtained by dividing the terms

in Appendix Eq. 42 with those in Appendix Eq. 44
The lower bound α0 follows from Appendix Eq. 37: and substituting the fluid-loss expressions from
Appendix Eqs. 45 and 46:
α0 = 1 − θ η → 0. (41)
∆w 1   g( ∆t Dso ) − g0  
For the typical value of θ = 1⁄2, Appendix Eq. 41 = ∆t Dso − [1 − ηso ]  . (47)
shows that α0 = 1⁄2, which is the value used in Eq. 9-43. wso ηso   κ so g0  

An expression for the treatment efficiency at any


Postscreenout pressure relations time after a screenout η(∆tDso) can be obtained by
combining the definition of η from Eq. 9-4 with the
The material-balance relations following a peripheral width multiplier expression in Appendix Eq. 47:
screenout (i.e., constant fracture area following a
screenout) can be written in a manner similar to
η( ∆t Dso ) = 1 −
(1 − η )
so
 g( ∆t Dso ) − g0 
1+
(1 + ∆t Dso )  .
Eq. 9-5 (Nolte, 1990). At any additional time follow- (48)
κ so g0 
ing a screenout ∆tso, the change in fracture volume is
equal to the additional injected slurry volume minus Appendix Eqs. 47 and 48 are equivalent to
the volume of fluid lost to the formation through Eqs. 6L-6 and 6L-7.
leakoff: Based on the proportionality between net pressure
∆ w A fso = qi ∆tso − VLp ( ∆tso ) ,
and width, differentiating Appendix Eq. 47 obtains
(42)
the following expression:

1 dpnet ( ∆tso )
where the subscript so identifies the corresponding
1 d ∆w
parameter value at screenout. Similarly, an expression =
for the fracture volume at screenout Vfso can be wso dt pnet ,so dt
written as
=
1  f ( ∆t Dso ) 
Vfso = ηsoViso , (43) tso ηso 1 − [1 − ηso ] , (49)
 κ so g0 
where ηso is the fluid efficiency at screenout and Viso
is the volume injected prior to the screenout. where pnet,so is the net pressure at the screenout and β
For an assumed constant injection rate Viso = qitso, is assumed to be unity (Fig. 9-20). From the defini-
Eq. 9-6 and Appendix Eq. 43 can be combined: tion of efficiency, it can be readily shown that

(w A) f = ηso qi tso = ηsoVLp (tso ) (1 − ηso ) , (44) tso ηso


=
tη( ∆tso )
. (50)
pnet ,so pnet ( ∆tso )
so

where VLp(tso) represents the cumulative volume of


fluid lost prior to screenout and can be expressed in a Combining Appendix Eqs. 49 and 50 produces the
manner similar to Appendix Eq. 22 for a wall-building following relation for the log-log slope of the net
fluid: pressure after a screenout (Nolte, 1990):
VLp (tso ) = 2 rp κ so CL tso A f g0 (α ) , (45) t dpnet ,so
=
pnet ,so dt
where κso is the spurt factor evaluated at tso.
Because the same set of conditions as that follow- (
1 − [1 − ηso ] f ( ∆t Dso ) (κ so g0 ))
( ( ) (κ g ))
.
1 − [1 − ηso ] [1 + ∆t Dso ] 1 + g( ∆t Dso ) − g0
ing the end of injection is assumed to exist following
so 0
a screenout, the expression for VLp(∆tso) is similar to
that in Appendix Eq. 26: (51)

A9-10 Chapter 9 Appendix: Background for Hydraulic Fracturing Pressure Analysis Techniques
The bounding values of the log-log slope are for describing the pressure in the reservoir are the
obtained by substituting the appropriate values for η pressure drop caused by the near-face leakoff effects
and the corresponding relations for f(∆tDso), g(∆tDso) (i.e., filter cake and filtrate) ∆pnf and the pressure dif-
and g0 that can be obtained from Appendix Eqs. 28, ference in the reservoir ∆pR:
29 and 30, respectively.
∆pT = p f − pi = ∆pnf + ∆pR t < tc (52)

Comparison of fixed-length ∆pT = ∆pR ; ∆pnf = 0 t > tc . (53)


and propagating fractures† In general, each of the pressure differences depends
Fracture propagation and closure lead to characteristic on time and position as the fracture propagates and
pressure and fluid-loss distributions along the inter- closes.
face between the fracture and the reservoir. The time As shown in Eq. 9-76, ∆pR is represented by the
history of these distributions establishes the boundary added contributions from the two sources of fluid
conditions for the reservoir response before and after loss: the pressure difference associated with the CL
fracture closure. These conditions are generally dif- component of fluid loss ∆pRC and the component
ferent from those experienced by fixed-length frac- associated with spurt loss ∆pRS. Nolte et al. (1997)
tures; however, there are also similar conditions that provided the linear flow expression for ∆pR that
are shared by these two types of fractures. The simi- includes the time dependence for ∆pRS in Eq. 9-84,
lar conditions enable applying established fixed- which is applicable both before and after closure.
length relations to a propagating fracture. The back- The remainder of this section of the Appendix
ground for identifying these differences and assumes that near-face effects and spurt loss are neg-
similarities is reviewed in this section of the ligible and focuses on the CL component of the pres-
Appendix to Chapter 9 along with the fixed-length sure difference:
relations that are applicable to the propagating case. ∆pRC = ∆pT ; ∆pRS = ∆pnf = 0. (54)
Specifically, this section of the Appendix provides
Abousleiman et al. (1994) expressed the general
• conditions that enable adapting fixed-length relations
relation for ∆pRC in terms of an integral equation
to a propagating fracture
using a Green’s function approach and requiring
• definition of a generalized reservoir leakoff coeffi- specification of the fluid-loss history. Another
cient that is applicable to fracture propagation approach for describing ∆pRC is to identify several
within all flow regimes simplifying features for a propagating fracture. These
• framework for type-curve analysis of the after- are the conditions of approximately constant pressure
closure period. and dimensionless time during fracture propagation.
Equations 9-23 and 9-24 characterize the pressure
Throughout this section, the propagating fracture
profile within the fracture and its magnitude at the
is assumed to have a rectangular shape (i.e., Af = 2hf L)
wellbore in terms of net pressure. These equations
and a square root of time dependency of the fluid loss
indicate that the pressure is only weakly dependent
(i.e., θ = 1⁄2 in Appendix Eq. 5).
on fracture length and the position in the fracture,
with the potential exception of the region near its tip.
Pressure characterization for It can, therefore, be concluded that the overall spatial
a propagating fracture and temporal pressure variations within a fracture are
relatively small in comparison with the total pressure
The total pressure difference ∆pT between the fluid drop between the fracturing fluid and the undisturbed
pressure in the propagating fracture pf and the initial reservoir pressure. An approximately constant and
reservoir pressure pi can be divided into three compo- uniform pressure is a primary assumption for deriv-
nents, as proposed by Howard and Fast (1957) and ing the individual leakoff coefficients (see Section 6-4)
illustrated in Fig. 5-17. The components of interest and the primary reason for their successful applica-
tion as invariant parameters. This pressure condition

This section by K. G. Nolte, Schlumberger Dowell.
is also reflected in the approximately constant injec-

Reservoir Stimulation A9-11


tion pressure during propagation at a constant injec- whereas the fixed-length case is generally associated
tion rate. The simultaneous condition of constant with a conductive fracture.
pressure and constant rate provides one of the pri-
mary distinctions between a propagating and a fixed-
length fracture. The fixed-length fracture requires an Fluid-loss characterization for
increasing pressure difference to maintain a constant a propagating fracture
production rate or a decreasing rate for a constant- Figure 9-35 illustrates the spatial distributions of the
pressure condition. From the reservoir perspective, fluid-loss rate during propagation and the specific
this difference occurs because a propagating fracture loss volume after propagation. The discussion of this
has an approximately constant or stationary value of figure in Section 9-6 observes that the specific vol-
dimensionless time with its associated condition of ume distribution is similar to that expected from a
time invariance for pressure and flow rate. uniform-flux fixed-length fracture. This similarity
provides the basis for the apparent fracture length
relation given by Eq. 9-78. A second type of fixed-
Propagation with a stationary length fracture assumes uniform pressure, or equiva-
dimensionless time lently infinite conductivity. Both of these fixed-length
The dimensionless time T (Eq. 9-75) can be com- fractures were discussed and characterized by
bined with the power law relation for fracture area Gringarten et al. (1974).
versus time (Appendix Eq. 2) and an efficiency-based The two types of fixed-length fractures can be
approximation for the area exponent α (Eq. 9-44) to compared to the after-closure behavior of the impulse
obtain injection shown in Fig. 9-38 for a propagating frac-
0.5 η ture. Sidebar 9G summarizes the impulse description
T  tp 
=  , (55) of after-closure radial flow behavior by the time
Tp  t  derivative of dimensionless pressure. The impulse
where Tp is the dimensionless time at the end of description can also be applied to linear and transi-
pumping and the approximation used is α ≈ 0.5 + tional flow. For a propagating fracture, this applica-
0.25η, which covers the values given by Appendix tion requires the apparent length relation to transform
Eqs. 40 and 41 for typical fracturing conditions. the dimensionless time (Eq. 9-79) for defining the
Appendix Eq. 55 shows that T is stationary and equal dimensionless pressure of a fixed-length fracture. The
to its value at the end of pumping for a fracture with derivative of dimensionless pressure is the same for
vanishing efficiency (i.e., η → 0). The equation also the two fixed-length fractures in linear and radial flow
shows that T retains only a weak dependence on time (Eq. 9-95) and accurately represents the impulse
for moderate values of efficiency. For example, for behavior in Fig. 9-38 for these flow regimes.
η = 0.5, T decreases only 19% during the second half However, the fixed-length fractures have different
of a treatment. Furthermore, Fig. 9-38 shows that a transitional flow behaviors (e.g., Gringarten et al.,
10-fold change in the dimensionless time is required 1974) and represent the transitional behavior in
before any meaningful change occurs in the reservoir Fig. 9-38 with differing degrees of accuracy. The
flow regime. The reservoir flow regime, therefore, is uniform-flux fracture has a deviation of less than
even more weakly dependent on dimensional time 5% from the propagating case for the complete time
than the case for dimensionless time. It can be con- range shown in the figure (i.e., that applicable to the
cluded that typical conditions for fracture propagation impulse representation). The transitional behavior for
result in essentially a stationary dimensionless time the uniform-pressure case deviates by almost 25%
and reservoir flow regime. from that for the propagating fracture. The maximum
The approximately stationary value of T provides deviation in both cases occurs slightly before the
another primary difference between a propagating knee time. These quantitative results confirm the
fracture and a fixed-length fracture for which dimen- cited qualitative inferences from Fig. 9-35, that the
sionless time increases with increasing dimensional after-closure behavior of a propagating fracture can
time. An additional difference for a calibration treat- be represented by a uniform-flux fixed-length fracture.
ment is that it does not retain fracture conductivity,

A9-12 Chapter 9 Appendix: Background for Hydraulic Fracturing Pressure Analysis Techniques
In contrast to this after-closure comparison, the where reservoir linear flow is assumed (Howard and
uniform-pressure fracture provides essentially the Fast, 1957) and the definition is in terms of the total
same flux distribution in radial flow as that for the pressure difference between the fracture and the ini-
fluid loss from an inefficient propagating fracture. tial reservoir pressure. A definition in Eq. 6-91 is in
This comparison is shown on Appendix Fig. 4. The terms of the pressure difference ∆pRC between the fil-
fixed-length distribution from Gringarten et al. (1974) trate/reservoir interface and the initial reservoir pres-
was described as the stabilized flux distribution. The sure. This pressure difference is defined in Appendix
spatial variation of the fluid loss during propagation Eq. 54 and denoted as ∆pc in Eq. 6-91. Combining
was obtained by combining the fracture growth Appendix Eq. 56 and Eq. 6-92 gives the following
power law relation (Appendix Eq. 2) for vanishing ratio between the pressure differences and fluid-loss
efficiency (i.e., area exponent α = 1⁄2) and the square coefficients:
root of time leakoff behavior (Appendix Eq. 8). ∆pRC CL
= . (57)
∆pT Cc
Uniform-pressure fixed-length
fracture (Gringarten et al., 1974) When the near-face effects and spurt are negligible,
10
Inefficient, propagating fracture the two pressure differences are equal (i.e., Appendix
8 Eq. 54) and Appendix Eq. 57 then indicates the
Normalized flux

expected result that the combined and reservoir coeffi-


6 cients are also equal. This result reflects the expected
Stabilized distribution
4
radial flow, TD ≥ 1.0 reservoir behavior based on a linear relation between
the pressure difference and flux. This behavior is also
2 required for the radial flow reservoir coefficient:
∆pRC CL
Linear flow = 0.1 0.01 0.001
0
0.5 0.6 0.7 0.8 0.9 1.0 = . (58)
Dimensionless position, ξ
∆pT CR
The previously cited application of pD,up for defin-
Appendix Figure 4. Comparison of flux distribution for ing CR requires expression of the dimensionless pres-
a uniform-pressure fixed-length fracture and a fluid-loss-
dominated propagating fracture. sure (Eq. 12-2) in terms of the quantities for the CL
component of fluid loss:
Gringarten et al. also noted that the stabilized 2π khL ∆pRC t p ( ),
distribution is independent of its prior history. ( )
pD,up Tp =
µq L ,C
(59)
Furthermore, it is spatially the same as that for a
propagating fracture with vanishing efficiency and where the fluid-loss height hL = rphf and qL,C is the
spurt. It follows that the dimensionless pressure leakoff rate associated with Carter-based fluid loss.
response pD,up for the uniform-pressure fixed-length Substituting for the expression for qL,C with vanishing
fracture applies to a propagating fracture when the efficiency and spurt from Appendix Eq. 11 and the
pressure is evaluated at its stationary value of the dimensionless rate function f = π/2 for the specified
dimensionless time Tp (i.e., pD,up(Tp)). This dimen- conditions (Nolte, 1986a):
sionless pressure value applies for radial flow with
vanishing efficiency and spurt and therefore provides
the basis for subsequently defining the reservoir ( )
pD,up Tp =
k tp
µ CL L
∆pRC t p ( ) η → 0, κ → 1. (60)
leakoff coefficient CR for these conditions.
Before defining this coefficient, the conventional An expression for the radial coefficient CR can then
definition of the reservoir leakoff coefficient must be be obtained by combining Appendix Eq. 59 with the
considered: definitions of dimensionless time (Eq. 9-75), Cc
(Appendix Eq. 56) and the pressure ratio (Appendix
k φ ct Eq. 58):
Cc = ∆pT , (56)
πµ Cc
CR = πTp η → 0, κ → 1.
( )
(61)
pD,up Tp

Reservoir Stimulation A9-13


Limiting values for pD,up(Tp) were given by The second relation in Appendix Eq. 64 indicates
Gringarten et al. (1974): an expanded range of application relative to that

( )
assumed for the derivation of Appendix Eq. 61. The
Linear flow pD,up Tp = πTp Tp < 0.01 expanded range results from numerical simulations
Radial flow pD,up (T )
p
1
2
[
= ln Tp + 2.2 ] Tp > 3. (Abousleiman et al., 1994; Nolte, 1998) that indicate
that Appendix Eq. 64 approximates (i.e., within a 5%
(62) error) the reservoir coefficient for transitional flow.
More generally, the simulations indicate that
They also provided a general relation for pD,up(Tp)
Appendix Eq. 64 is approximately valid (i.e., within
in terms of special functions. This relation can be
a 10% error) for moderate values of efficiency
approximated with an error of less than 2% by
(η < 0.5) and with any reservoir flow regime.
 πTp − 0.58Tp Tp < 0.16 Therefore, CR, as defined by Appendix Eq. 61,

( )
pD,up Tp ≅  1 represents the “general reservoir” leakoff coefficient
( [
 2 ln Tp + 0.22 + 2.2 ] ) Tp > 0.16. within the accuracy required for fracture design and
evaluation purposes.
(63)
Appendix Eq. 63 can be used to show that during
Comparison of Appendix Eqs. 62 and 63 shows transitional and radial flow, pD,up(Tp) ≤ √πTp.
that the approximations provided by the latter equa- Appendix Eq. 64 therefore implies that the general
tion are defined by adding a term to each of the limit- coefficient CR is larger than the linear flow coefficient
ing cases. The second approximation in Appendix Cc for fracture propagation under these flow condi-
Eq. 63 can be obtained by applying the apparent time tions. This result has been reported by Hagoort
multiplier (1 + 0.22/Tp) to the dimensionless time. Its (1980) and Valkó and Economides (1997). For exam-
inclusion extends the applicability of the logarithmic- ple, for a dimensionless time Tp = 1, the dimension-
based radial flow relation to a dimensionless time that less pressure pD,up = 1.21 and √πTp = 1.78. For these
is about 1⁄20 of the value normally required for radial values, Appendix Eq. 64 indicates that the general
flow (i.e., Tp = 0.16 in Appendix Eq. 63 versus Tp = coefficient CR is about 1.5 times larger than Cc under
3 in Appendix Eq. 62). The apparent time multiplier these late transitional flow conditions. For a larger
for application with Appendix Eq. 63 has the same dimensionless time Tp = 10, CR becomes larger by
form as that for a similar development introduced in a factor of 2.5.
Eq. 9-83 for the after-closure behavior of a propa- This observation of a larger leakoff coefficient gen-
gating fracture. These two relations are seemingly erally applies to the mini-falloff test because the test
different because the relation for a propagating frac- design should be based on Tp > 1 and the achieve-
ture contains a different constant (i.e., 0.14 instead ment of reservoir-controlled fluid-loss conditions
of 0.22). However, this difference occurs because the (Appendix Eq. 1). Larger values of the reservoir coef-
dimensionless times corresponding to the two cases ficient do not affect the fluid loss for most proppant
differ by the square of the apparent length fraction treatments where near-face effects are designed into
faL from Eq. 9-79. For the assumed conditions of van- the fluid system to control the fluid-loss behavior.
ishing efficiency and spurt, Eq. 9-80 indicates that Combining Appendix Eqs. 58 and 61 provides the
faL = π/4. Applying this value to the dimensionless general relation for the CL component of the pressure
time for the fixed-length case indicates that the con- difference ∆pRC for all flow regimes and in terms of
stants for the two apparent time relations are actually the total pressure difference and combined fluid-loss
equivalent. coefficient.
For reservoir linear flow, Appendix Eq. 62 shows
that pD,up(Tp) ≈ √πTp. It follows from Appendix
Eq. 61 that Type-curve-based analysis
Cc Linear flow The dimensionless pressure pD,up can also be used to
 develop type-curve analyses for general after-closure
CR =  πTp conditions. The normalized pressure difference and
( )
C Transitionaland radial flow.
 pD,up Tp c
 pressure derivative variables and the log-log slope in
(64) Fig. 9-38 illustrate several characteristics of the after-

A9-14 Chapter 9 Appendix: Background for Hydraulic Fracturing Pressure Analysis Techniques
closure pressure response that motivate its analysis lished by the matching procedure, the dimensionless
within a type-curve framework. For example, the pressure pD,up(Tp) can be obtained using Appendix
character, or shape, of the curves depends on the Eq. 63.
dimensionless time Tp. Also, a suitable match pres- This type-curve analysis has the following relation
sure for the analysis can be defined as the ratio of the between the after-closure pressure and the pressure-
dimensional pressure variables and the normalized difference curve Rp:

[ ]
pressure variables shown on the figure.
The development of type curves applicable to ∆pR (t ) = p(t ) − pi = m p Rp Tp , (t − tc ) tc , η, κ . (69)
generalized fluid-loss conditions requires a relation
For fracture propagation in flow regimes other than
among the average value of the fluid-loss rate q—L, CL
well-established linear and radial flow, Rp must be
component of the fluid-loss rate qL,C, injection rate qi,
defined by numerical simulation.
fluid efficiency η and spurt factor κ. This relation can
The matching procedure uses a pair of type curves:
be obtained by combining the rate versions of Eq. 9-6
one for the pressure difference and one for the pres-
and Appendix Eq. 22:
sure derivative. Each quantity can be defined using an
q L = κ q L ,C = (1 − η) qi . (65) appropriate time function. The square of the linear
flow time function FL(t/tc) given by Eq. 9-88 is pre-
For general values of efficiency and spurt, the ferred because it provides a consistent representation
reservoir pressure difference at the end of pumping of the after-closure period for the reservoir response
∆pR(tp) provides a convenient quantity to use as the in any flow regime, as discussed in Section 9-6.7.
match pressure (i.e., the multiplying factor for the Furthermore, the pressure difference and the pressure
type curves). The relation between ∆pR(tp) and ∆pRC derivative are presented in terms of the inverse of FL2
at the end of injection can be obtained from Eqs. 9-76 because this presentation provides the conventional
and 9-84 and Appendix Eq. 58: representation of increasing time from left to right
along the x-axis. Appendix Fig. 5 illustrates these
κ + 1 κ + 1  CL
( )
∆pR t p =  ∆p t =  ( )
 2  RC p  2  CR T p
∆p t . ( ) curves for the case of vanishing efficiency and spurt
(e.g., applicable to a mini-falloff test).
(66) The type-curve analysis begins by matching the
Rearranging Appendix Eqs. 65 and 66 and substi- character (i.e., shape) of the pressure derivative for
tuting them into Appendix Eq. 59 provides a more the data to the character of one of a collection of type
general form of the dimensionless pressure: curves based on different values of Tp. This character
matching defines Tp. The selected curve for Tp is then
 2κ  2 π kh
( ) ( )
vertically translated to match the pressure derivative
pD,up Tp =   mp ∆pR t p → m p ,
 (1 − η)(1 + κ )  µqi
of the data, and the resulting form of the pressure
derivative defines the match pressure mp:
(67)
where the fluid-loss height hL = h and the role of 1.0
∆pR(tp) is introduced as the match pressure mp. The Pressure difference
Pressure derivative
transmissibility can be determined from this dimen- η→0 κ→0

sionless relation:

kh  (1 − η)(1 + κ )  qi pD,up Tp ( )
Rp(t)

=
µ   2 πm . (68) Tp = 0.001
2κ  p

In addition to determination of the match pressure for


Tp = 2.00 Tp = 0.05
the type-curve analysis, the dimensionless time at the 0.1
end of pumping Tp is required to define pD,up for 0 10 100
1/{FL(t/tc)}2
Appendix Eq. 68. The value of Tp can be obtained by
matching the character, or shape, of the data to that
of the type curves. Once a value of Tp has been estab- Appendix Figure 5. Example type curves for negligible
efficiency and spurt.

Reservoir Stimulation A9-15


  (i.e., ∆pR(tp) inferred from the after-closure data).
dp(t )
{F (t t )} { }
dRp
= m p  FL (t tc )  Relative to Appendix Eq. 66, ∆pT (tp) = p(tp) – pi, with
2 2
2 .
{
d FL (t tc ) }  {
d FL (t tc )} 
L c 2

  p(tp) defined by the ISIP, and when ∆pR(tp) ≈ ∆pT (tp)


in the absence of spurt, the reservoir and total leakoff
(70) coefficients are approximately equal. Under these
conditions, essentially the entire pressure difference
The transmissibility is then estimated from Appendix
∆pT is within the reservoir and the reservoir controls
Eq. 68 using the two type-curve parameters Tp and mp.
the fluid-loss behavior. Conversely, when ∆pR(tp) <<
The initial reservoir pressure is extracted by applying
∆pT(tp), a negligible fraction of the total pressure drop
Appendix Eq. 69 over the time range of the match.
occurs in the reservoir, and the reservoir does not
Additional information concerning the nature of
have a meaningful role in controlling fluid leakoff.
the fluid loss can be obtained from the value of mp

A9-16 Chapter 9 Appendix: Background for Hydraulic Fracturing Pressure Analysis Techniques

You might also like