Вы находитесь на странице: 1из 11

Determination of Fracture Conductivity in

Tight Formations With Non-Darcy Flow


Behavior
Feng Zhang, SPE, and Daoyong Yang, SPE, University of Regina

Summary fracture and unsteady-state Darcy flow in the matrix. This facili-
In this paper, a mathematical model has been developed and suc- tates displaying changes in the flux distribution along the fracture,
cessfully applied to accurately determine the fracture conductivity whereas non-Darcy flow in a fracture leads to the apparent fracture
in tight formations with non-Darcy flow behavior. A new non- conductivity being significantly less than its true value. Subse-
Darcy flow number is first defined to account for the effect of quently, Umnuayponwiwat et al. (2000) used a similar model to
characteristic length in a hydraulic fracture. A semianalytical investigate the flux distribution along the fracture and interaction
method is then applied to solve the newly formulated mathemati- between the fracture and matrix and found that an increase in the
cal model by discretizing the fracture into small segments, assum- fracture flow would result in an increase in the non-Darcy flow
ing that there exists unsteady flow between the adjacent segments. effects. Because of its difference from Darcy’s condition, buildup
The newly developed model has been validated by simplifying it tests could not be analyzed as superposition of drawdown tests for
to the traditional Forchheimer (i.e., non-Darcy) model and by per- the nonlinearity of non-Darcy problem (Guppy et al. 1982b).
forming numerical simulation with a reservoir simulator as well. By examining non-Darcy effects in the proppant packs with
The pressure response and its corresponding derivative type laboratory tests, Martins et al. (1990) discovered that pressure-loss
curves have been reproduced to examine non-Darcy flow behavior behavior occurs at a high flow rate. Vincent et al. (1999) pointed
under different fracture conductivities. Both relative minimum out that increasing fracture width or placing high-permeability
permeability and characteristic length are found to impose a nega- proppant could increase conductivity. Settari et al. (2000) used a
tive effect on the fracture conductivity. Compared with relative finite-difference simulator for evaluating performance of hydraulic
minimum permeability, characteristic length is a strong function fracturing of high-permeability gas wells, thereby reducing the
dominating the non-Darcy flow behavior in the fractures. It is associated non-Darcy skin effects. It was found that, in a moder-
obvious that the fracture conductivity can be accurately deter- ate- to high-permeability reservoir, the non-Darcy effect is too
mined when non-Darcy flow behavior in the fracture network is large to be neglected. Gill et al. (2003) examined the effect of non-
taken into account. Darcy flow in a low-permeability gas reservoir and subsequently
developed correlations to limit non-Darcy flow within acceptable
ranges. After examining the effect of non-Darcy flow in nanopore
Introduction shale reservoirs, Swami et al. (2012) concluded that the non-Darcy
Primary recovery factor remains low (approximately 5 to 10% of effect is not only significant, but also dependent on pore radius.
original oil in place) in tight oil formations even after long hori- Smith et al. (2004) found that non-Darcy flow should not be simply
zontal wells have been drilled and massively fractured (Manrique considered as the reduction of the true conductivity, whereas Gup-
et al. 2010). This is ascribed to the fact that fluid flow in such tight py’s correlations (Guppy et al. 1982a) would overestimate the
formations can be completely dependent on the fracture network, non-Darcy flow effects. As for non-Darcy effects at low flow rates,
whereas the matrix plays only a source role (Manrique et al. Miskimins et al. (2005) concluded that non-Darcy effects would
2007). Numerous efforts have been made to describe the non- reduce the flow capacity by 5 to 30% in tight gas reservoirs with
Darcy effect on the basis of the traditional Forchheimer equation, relatively low permeability (0.01 md). By incorporating the
the limits of which have been identified (Andrade et al. 1997; Hill unsteady-state solutions in both the fracture and the matrix, Zeng
et al. 2001; Stanley and Andrade 2001). By replacing the inertial and Zhao (2010) pointed out that this solution can be especially
factor in the Forchheimer model with two new parameters—mini- applied to tight formations. The Barree-Conway model has been
mum permeability plateau and characteristic length—the Barree- used to describe the pressure-transient behavior of non-Darcy flow
Conway model is claimed to be capable of describing non-Darcy in hydraulically fractured wells with the finite-difference method
flow behavior from low to high flow rates (Barree and Conway (Al-Otaibi and Wu 2011). Because limited field data have been
2004; Lai et al. 2012). So far, few attempts have been made to made available in the literature, it is essential that the versatile
quantify the fracture conductivity in such hydraulically fractured Barree-Conway model be modified and used to accurately describe
formations because of both its complexity and the existence of non-Darcy flow behavior in hydraulically fractured wells.
non-Darcy flow behavior. Therefore, it is of fundamental and In this paper, a new mathematical model has been developed
practical importance to accurately determine the fracture conduc- and successfully applied to determine fracture conductivity
tivity in tight formations with non-Darcy flow behavior. with non-Darcy flow behavior in a hydraulic fracture. After the
The non-Darcy flow in the fracture has been investigated to unsteady-state solution in the fracture is applied to accurately
examine its effect on the transient pressure behavior. After analyz- model the flow in the tight formation, a semianalytical method is
ing the transient behavior of various conductivities with a finite- developed to obtain the solutions for the flow rates in the fracture
difference simulator, Holditch and Morse (1976) concluded that and matrix. As such, the pressure and its derivative type curves can
non-Darcy flow in the fracture reduced the apparent conductivity. be generated, respectively. The non-Darcy effects on fracture con-
Guppy et al. (1982a) proposed a comprehensive semianalytical ductivity by relative minimum permeability and characteristic
model of non-Darcy flow in wells with finite-conductivity frac- length are discussed in tight formations. In addition, a field case is
tures, assuming there exists steady-state non-Darcy flow in the shown to evaluate the actual conductivity under the non-Darcy
effect in tight formations.
Copyright V
C 2014 Society of Petroleum Engineers

This paper (SPE 162548) was accepted for presentation at the SPE Canadian Theoretical Model
Unconventional Resources Conference, Calgary, 30 October–1 November 2012, and
revised for publication. Original manuscript received for review 12 August 2012. Paper peer Mathematical Formulation. In this study, Fig. 1 shows a verti-
approved 18 June 2013. cal fracture that is fully penetrating along the well in a closed

34 February 2014 SPE Journal


φf , ctf , kf with boundary conditions
z
km h @p
j ¼ qf ;  xf  x  xf . . . . . . . . . . . . . . ð7aÞ
l @y y¼0

φm , ctm , km pðx ! 1; y ! 1; tÞ ¼ pi : . . . . . . . . . . . . . . . . . . . . ð7bÞ


h xf The initial condition for both fracture and formation is

pðx; y; t ¼ 0Þ ¼ pi : . . . . . . . . . . . . . . . . . . . . . . . . . ð7cÞ

For the convenience of analysis, the following definitions are


x made to convert the mathematical models into dimensionless
y forms:

Fig. 1—Schematic of a vertical hydraulic fracture in a closed cy- km hðpi  pÞ km hðpi  pf Þ


lindrical reservoir. pD ¼ ; pfD ¼ ; . . . . . . . . . . . . ð8Þ
Qsc Bg l Qsc Bg l
km t
cylindrical reservoir. There are two assumptions. First, the reser- tD ¼ ; . . . . . . . . . . . . . . . . . . . . . . . . . . . ð9aÞ
/m lctm x2f
voir is isotropic, homogeneous, and bounded by upper and lower
impermeable layers. The fluid flow in the formation is considered x y
xD ¼ ; yD ¼ ; . . . . . . . . . . . . . . . . . . . . . . . . . . ð9bÞ
as Darcy flow, whereas the flow in the fracture can be described xf xf
with the Barree-Conway model (Barree and Conway 2004). Sec-
ond, the matrix porosity /m , compressibility ctm , permeability km , q Qsc Bg
FND ¼ ; . . . . . . . . . . . . . . . . . . . . . . . . . . ð9cÞ
and thickness h are set to be constant. The half-length and width ls wf h
of the vertical fracture is denoted by 2xf and wf , respectively. The qf xf qN
fracture porosity /f , compressibility ctf , and permeability kf are qfD ¼ ; qND ¼ ; . . . . . . . . . . . . . . . . . . . ð9dÞ
Qsc Bg Qsc Bg
also assumed to be constant.
According to the Barree-Conway model (Barree and Conway kf w f
CfD ¼ ; . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð9eÞ
2004), the flow in the fracture is expressed by km xf
@p lv kf /m ctm
¼ 2 3 ; . . . . . . . . . . . . . . . . . . . ð1Þ Cg ¼ : . . . . . . . . . . . . . . . . . . . . . . . . . . ð9fÞ
@x /f ctf km
6 ð1  kmr Þ 7
kf 6
4kmr þ  7 Eqs. 3 through 5b can then be rewritten as
qv 5
1þ  
ls @ @pfD qfD 1 @pfD
d þ ¼ ; . . . . . . . . . . . . . . . . ð10Þ
where l is fluid viscosity, v is superficial velocity, q is fluid den- @xD @xD CfD Cg @tD
sity, kmr is minimum permeability relative to Darcy permeability, 1  kmr
and s is the characteristic length. d ¼ kmr þ ; . . . . . . . . . . . . . . . . . . . . . ð11aÞ
1 þ FND qND
According to the Darcy’s flow model, the flow in the matrix  
can be expressed as @pfD  1 1
d ¼ ; . . . . . . . . . . . . . . . . . . . . . ð11bÞ
@xD xD ¼0 2 CfD
@p lv 
¼ : . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð2Þ @pfD 
@x km ¼ 0: . . . . . . . . . . . . . . . . . . . . . . . . . . . ð11cÞ
@x D xD ¼1
The fracture is considered as a plane source with nonuniform
flux distribution from the matrix to the fracture; i.e., qf ðx; tÞ in Similarly, Eqs. 6 through 7b can be rewritten as
m3 =ðs  mÞ. On the basis of Eq. 1, the expression for fluid flow in  2 
the fracture can be written as @ pD @ 2 pD @pD
þ ¼ ; . . . . . . . . . . . . . . . . . . . . . ð12Þ
  @x2D @y2D @tD
kf @ @pf qf @pf
l @x
d
@x
þ
wf h
¼ /f ctf
@t
; . . . . . . . . . . . . . . . ð3Þ @pD 
 ¼ qfD ;  1  xD  1; . . . . . . . . . . . . . . ð13aÞ
@yD y¼0
where
pD ðxD ! 1; yD ! 1; tD Þ ¼ 0: . . . . . . . . . . . . . . . ð13bÞ
1  kmr
d ¼ kmr þ q qN . . . . . . . . . . . . . . . . . . . . . . . . ð4Þ
1þ The initial condition can be expressed with the dimensionless
ls wf h term

and where qN is the flux inside the fracture. The boundary condi- pD ðxD ¼ 0; yD ¼ 0; tD Þ ¼ 0: . . . . . . . . . . . . . . . . . . . ð13cÞ
tions can be written as
 
kf hwf @pf  Qsc Bg Semianalytical Solution. The semianalytical method applied to
d ¼ ; . . . . . . . . . . . . . . . . . . ð5aÞ
l @x x¼0 2 solve the hydraulic-fracture system has been widely discussed in
 the literature (Guppy et al. 1982a; Umnuayponwiwat et al. 2000;
@pf 
¼ 0: . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð5bÞ Zeng and Zhao 2010). In general, the whole system is divided
@x x¼xf into the matrix system and the fracture system, each of which is
solved separately. The fracture system is first discretized into sev-
The pressure drop in the formation can be expressed by eral segments in each of which the flux qfD can be approximated
  to its average value within the segment. With such an approxima-
km @ 2 p @ 2 p @p
þ ¼ /m ctm ; . . . . . . . . . . . . . . . . . . ð6Þ tion, equations for the matrix and the fracture systems can be
l @x2 @y2 @t solved in the Laplace domain in real time, respectively. Their

February 2014 SPE Journal 35


solutions are then coupled along the fracture interface, leading to p fDi ðxD ; ui Þ ¼ q ND i1  p s ðxD ; xDi1 ; ui Þ  q ND i
a linear system. Solving the linear system, the flux in each seg-
ment can be obtained. Such obtained flux can be substituted into Cg 
 p s ðxD ; xDi ; ui Þ  q ; . . . . . . . . ð19Þ
the first segment to determine the bottomhole pressure (BHP) in CfD ui fD i
the Laplace domain in pseudotime. Finally, the Stehfest inverse
algorithm (Stehfest 1970) can be used to convert the BHP in the where
Laplace domain in real time. Compared with the traditional meth-  
q ND  q ND  q fD i
ods and numerical simulation, the newly developed technique on q ND i1 ¼ ; q ND i ¼ ; q fD i ¼ :
d  xD i1 d  xD i di
the basis of the Laplace domain not only easily incorporates the
skin factor, flow rate, wellbore storage, and conductivity in a uni-                    ð20Þ
fied and consistent framework, but also efficiently and accurately
evaluates the petrophysical parameters together with reservoir For no-flow-boundary segments, 0  x0D  DxD, with a source at
performance for a complex reservoir. x0D0 , van Kruysdijk (1988) provided the solution:
( pffiffiffiffiffiffiffiffiffiffiffiffi
Matrix System. The solution to the matrix system can be 2cosh½ðx0D  x0D0 Þ ui =Cg 
0 0 1
directly obtained by applying the point-source solution in the Lap- p s ðxD ; xD0 ; ui Þ ¼ pffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffi
lace domain to a closed cylindrical reservoir (Ozkan and Ragha- 2 ui =Cg expð2DxD ui =Cg Þ
van 1991), yielding
q ffiffiffiffiffiffiffiffiffiffiffi ffi 2cosh½ðx0 þ x0 Þpffiffiffiffiffiffiffiffiffiffiffi ffi
ui =Cg 
ð1   qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi þ exp jx0D  x0D0 j ui =Cg þ D
p D0
ffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffi expð2DxD ui =Cg Þ
pD ¼ q fD ðx; uÞ K0 u ðxD  xÞ2 þ y2D )
1
qffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffi  qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
K1 ðreD uÞ pffiffiffi þ exp jx0D þ x0D0 j ui =Cg :                ð21Þ
þ pffiffiffi I0 u ðxD  xÞ2 dx;
I1 ðreD uÞ
                   ð14Þ When the source is at x0D0 ¼ xDi–1 and x0D0 ¼ xDi, the final solu-
tion is obtained with the superposition principle:
where K0 is a zero-modified Bessel function of the second type; it
p fDi ðxD ; ui Þ ¼ bi ðxD Þq ND i1 þ ci ðxD Þq ND i þ di q fD i ; . . . ð22Þ
has a singularity at xD ¼ x. Van Kruysdijk (1988) proposed a
method for integration with the singularity. The modified Bessel where
function K0 (x) can be divided into two parts, lnðxÞ and "
8 rffiffiffiffiffiffi#9
K0 ðxÞþlnðxÞ. The singularity in the first part can be integrated >
> ui > >
>
>2cosh ðx D  x D i1 Þ >
>
analytically, whereas the second part, which is free of singular-
1 < C g =
ities, can be integrated simply by discretization. For the second bi ðxD Þ ¼ rffiffiffiffiffiffi rffiffiffiffiffiffi
term in Eq. 14, the ratio of the Bessel functions approaches zero ui > > ui >
>
CfD > 2ðxD i xD i1 Þ >
at a large value of u, and thus in this case the second term will not Cg > :
e C g
1
>
;
contribute to theffi solution. Another Bessel-function integration,
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiu
pffiffiffi 2ðxD xD i1 Þ
i

I0 ½ u ðxD  xÞ2 , is free of singularity and thus can be evaluated þe Cg


; . . . . . . . . . . . . . . . . . . ð23aÞ
by numerical integration. 8 " rffiffiffiffiffiffi#9
Fracture System. Because dðxD ; tD Þ is a function of time and >
> ui > >
>
>2cosh ðxD i  xD Þ C > >
space, the mathematical model for the fracture system is a strong 1 < g =
nonlinear equation. To solve such a nonlinear equation, the frac- ci ðxD Þ ¼ rffiffiffiffiffiffi rffiffiffiffiffiffi
ui > > ui >
>
ture system is first discretized into several segments, in each of CfD > 2ðxD i xD i1 Þ >
which the mathematical model can be solved analytically. Zeng Cg > :
e C g
1
>
;
and Zhao (2010) provided a detailed derivation of the solution in qffiffiffiffi
u i
the fracture system. 2ðxD i xD Þ
Cg
An average di ðtD Þ is used to simplify the dðxD ; tD Þ distribution þe ; . . . . . . . . . . . . . . . . . . .ð23bÞ
in each segment. A pseudotime is defined on the basis of the aver- Cg
age di ðtD Þ; that is, di ¼  : . . . . . . . . . . . . . . . . . . . . . . . . . . . . ð23cÞ
CfD ui
ð tD
sDi ¼ di ðfÞdf: . . . . . . . . . . . . . . . . . . . . . . . . . . ð15Þ Continuity Conditions. There are two continuity conditions at
0 the interface between any two adjacent segments in the fracture
and boundary between the matrix and the fracture: pressure conti-
The mathematical model in segment i (xDi–1  xD  xDi) is nuity and flow-rate continuity. Because the solution of the fracture
written as system is derived from the pseudotime, pressure or flux of any ad-
jacent grids cannot be directly equalized at the interface in this
@ 2 pfD qfDi ðtD Þ 1 @pfD study. Therefore, a new continuity condition on the basis of the
 ¼ ; . . . . . . . . . . . . . . . . . . . ð16Þ
@x2D di CfD Cg @sDi Stehfest algorithm for the inverse Laplace transformation is pro-
posed. As such, the solution can be obtained in the Laplace do-
with boundary conditions at xDi–1 and xDi, main, incorporating variable flow rates, skin effect, wellbore
    storage, and variable conductivities more easily.
@pfD  1
qND  @pfD  1
qND  Because the pressure in the real-time domain should be same
¼ ; ¼  :
@xD xDi1 CfD d xDi1 @xD xDi CfD d xDi along the boundary between the fracture and matrix, we have
                   ð17Þ pD ðtD i Þf ¼ pD ðtD i Þm : . . . . . . . . . . . . . . . . . . . . . . . . ð24aÞ
The flux, qfD(xD,tD), can be approximated to its average value
The solution for the fracture system is derived from the pseu-
in each fracture segment i (xDi–1  xD  xDi):
dotime Laplace transformation, whereas the solution for the ma-
trix system is derived from the real-time Laplace transformation.
qfD ðxD ; tD Þ ¼ qfDi ðtD Þ: . . . . . . . . . . . . . . . . . . . . . . . ð18Þ
Eq. 24a can be rewritten as
The solution to Eqs. 10 through 11c can be obtained with a
pD ðs i Þf ¼ pD ðtD i Þm : . . . . . . . . . . . . . . . . . . . . . . . ð24bÞ
source function that can be written as

36 February 2014 SPE Journal


ðp
1 2 3 4 ... ... N-3 N-2 N-1 1 2 3 4 ... ... N-2 N-1 N p0
mðpÞ ¼ dp0 : . . . . . . . . . . . . . . . . . . . . . ð27Þ
0 uðp0 ÞZðp0 Þ
1 ∗ ∗ ∗∗
2 ∗ ∗ ∗ ∗∗ Because all computations are derived from the Laplace trans-
3 ∗ ∗ ∗ ∗ ∗ form, wellbore storage can be easily added with the well-known
4 ∗ ∗ ∗ ∗ ∗ identity
... .... ... p wD
... ... ... p wD ¼ : . . . . . . . . . . . . . . . . . . . . . . . ð28Þ
1 þ u2 CD p wD
N-3 ∗ ∗ ∗ ∗ ∗
N-2 ∗ ∗ ∗ ∗ ∗
Model Validation
N-1 ∗ ∗ ∗ ∗
∗ ∗ ∗
Forchheimer Equation. The newly developed model can be
1 ∗ ∗ ∗ ∗
∗ ∗ ∗
simplified to the Forchheimer equation because it is the most
2 ∗ ∗ ∗ ∗ ∗
∗ ∗ ∗ ∗ ∗ ∗
widely used flow model for describing non-Darcy flow behavior
3 ∗ ∗ ∗ ∗ ∗
∗ ∗ ∗ ∗ ∗ ∗
(Al-Otaibi and Wu 2011). The common factor between the Bar-
4 ∗ ∗ ∗ ∗ ∗
∗ ∗ ∗ ∗ ∗ ∗ ree-Conway model and the Forchheimer equation is the apparent
... ... ... permeability of non-Darcy flow, which is defined by Eqs. 29a and
... ... ... b, respectively,
N-2 ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ 2 3
N-1 ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗
6 ð1  kmr Þ 7
∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ ∗ kapp ¼ k6 7; . . . . . . . . . . . . . . . . . . ð29aÞ
N 4kmr þ  qv 5

Fig. 2—The coefficient matrix for solving the matrix/hydraulic- sl
fracture system. k
kapp ¼   ; . . . . . . . . . . . . . . . . . . . . . . . ð29bÞ
bkqv

l
By use of the Stehfest algorithm for the inverse Laplace trans-
where kapp is the apparent permeability and b is the inertial flow
formation, we have
parameter.
    When the apparent permeability of the two models is set to be
ln2 XN
ln2 ln2 XN
ln2
Vj p j ¼ Vj p j : . . . . . . . . ð25Þ the same, the inertial flow parameter b is solved with
si j¼1 si f tD j¼1 tD m
lð1  kmr Þ
b¼ ; . . . . . . . . . . . . . . . . . . . . . . . . . ð30Þ
Accordingly, the pressure-continuity condition for the fracture kðsl þ kmr qvÞ
and matrix system can be obtained as where fluid superficial velocity v can be approximated by use of a
    constant production rate and cross-sectional area of flow in a well.
1 ln2 1 ln2
p j ¼ p j : . . . . . . . . . . . . . . . . . ð26aÞ Eq. 30 can be used to correlate the Forchheimer equation and the
si si f tD tD m Barree-Conway model. According to the parameters listed in
Table 1 (Vincent et al. 1999; Settari et al. 2000; Gill et al. 2003),
Similarly, the pressure- and flux-continuity conditions for the the tight gas reservoir is assumed to have no boundary effects (i.e.,
node between two adjacent segments are, respectively, expressed a sufficiently large drainage radius is assumed and used). Fig. 3
as shows an excellent agreement in the transient stage between the
    Forchheimer equation simplified from the newly developed model
1 ln2 1 ln2
p j ¼ p j ; . . . . . . . . . . . . . . . . ð26bÞ and the original Barree-Conway model. There is a very minor mis-
si si f siþ1 siþ1 f
match when entering the bilinear period because the flow rate will
    change significantly at the initial stage, which will then cause a
1 ln2 1 ln2
q j ¼ q j : . . . . . . . . . . . . . . . . ð26cÞ large difference. When reaching the pseudoradial-flow period, the
si si f siþ1 siþ1 f
flow rate will be stabilized and the difference will disappear.
A linear equation is then obtained by applying the previously
discussed continuity conditions. The unknown variables are com- Numerical Simulation. A numerical-simulation model is also
posed of the flux from the matrix to the fracture at each segment used for validation. The numerical model is built with a reservoir
along the fracture, q fDi (i ¼ 1, 2, …, N), and the flow rate inside simulator (IMEX 2010, Computer Modelling Group Ltd.). The
the fracture at each node, q NDi (i ¼ 0, 1, 2, …, N). The flow rates 3D grid system is 20201 with cell dimensions of 202020
at Nodes 0 and N are given as boundary conditions. Therefore, the ft. The single cell at the fracture location was locally refined with
number of unknown variables is 2N–1. Coupling the solutions at the keyword “hydraulically fractured wells.” The number of
the node between two adjacent segments inside the fracture yields refined blocks in each direction is 771. The non-Darcy effect
N–1 equations, and coupling the solutions at the interface between was modeled with the keyword “non-Darcy option.”
the matrix and fracture leads to N equations. Thus, there are 2N–1 The Geertsma (1974) correlation was used to calculate inertial
variables with 2N–1 equations, and the linear system can be flow parameter b. The corresponding kmr and s were obtained with
solved. Fig. 2 presents the format of the coefficient matrix. Eq. 30. The model was built as a tight gas reservoir with porosity
The flux from the matrix to the fracture and the flow rate inside of 15% and permeability of 0.1 md. Fig. 4a shows a top view of
the fracture at each node are obtained by solving this linear equa- the 3D grid system, whereas input parameters are listed in Table 2.
tion system. Substituting the flux at Segment 1 and the flow rate As can be seen in Fig. 4b, a good agreement exists between the
at Node 1 into the original equation for Segment 1 yields the BHP results from this study and those from the Computer Modelling
in the Laplace domain in pseudotime. Subsequently the BHP in Group simulation. It is not surprising that a close match was not
the real-time domain can be obtained with the inverse Laplace obtained at the first 0.1 hour. This is because use of transient flow
algorithm, as proposed by Stehfest (1970). conditions yields a large pressure drop at initial production.
The solution can also be used for a real-gas flow, provided that
pseudopressure transformation has been made (Lee and Holditch Sensitivity Analysis
1982; Spivey 1984). The solution in terms of pseudopressure is Unlike with traditional models based on the Forchheimer equa-
defined by tion, the non-Darcy effect is determined by use of two

February 2014 SPE Journal 37


TABLE 1—BASIC PARAMETERS USED FOR SIMULATING PRESSURE-TRANSIENT RESPONSE
IN A TIGHT GAS FORMATION (VINCENT ET AL. 1999; SETTARI ET AL. 2000; GILL ET AL. 2003)

Parameter Value Unit

Reservoir and Fluid Data Reservoir pressure, pi 3,000 psi


Reservoir permeability, km 0.1 md
Reservoir porosity, / 0.1 fraction

Reservoir temperature, T 190 F
Pay-zone thickness, h 60 ft
Total compressibility, ct 1.010–5 psi–1
Initial viscosity, l 0.018 cp
Density, q 0.0625 lbm/ft3
Production rate, Qsc 10 MMscf/D
Hydraulic-Fracture Data Fracture half-length, xf 800 ft
Fracture width, wf 0.02 ft
Fracture permeability, kf 180,000 md
Characteristic length, s 2.89104 ft–1
Relative minimum permeability, kmr 0.01 fraction
Equivalent b (Forchheimer) 1.96105 ft–1

dimensionless parameters (i.e., kmr and FND) with the Barree-Con- that FND imposes a smaller effect on kmr as it is decreased. When
way model. Accordingly, effects of these two parameters on pres- FND is set to be a small value, the non-Darcy effect will not affect
sure response are discussed here because effects of wellbore the pressure response significantly, even at a very small kmr value.
storage (CD), fracture conductivity (CfD), and fracture diffusivity
(Cg) have been extensively analyzed and discussed (Cinco-Ley. Non-Darcy Number (FND). Figs. 7 and 8 present drawdown
et al. 1978; Guppy et al. 1982a; Lee and Holditch 1982). type curves for FND ¼ 1, 10, and 100 at kmr ¼ 0.01 and 0.10,
respectively. As shown in Fig. 7, the pressure drop and its
Relative Minimum Permeability (kmr). In the Barree-Conway
model, kmr is first introduced, which is caused by streamlining or
flow heterogeneity. The drawdown type curves for kmr ¼ 0.01,
0.10, and 0.50 at FND ¼ 100 are depicted in Fig. 5. The typical
one-fourth slope at early times of both pressure and pressure-de-
rivative curves is exhibited to show a bilinear-flow region for
non-Darcy effect. Then, a linear flow with a slope of 0.5 follows
at the end of the bilinear-flow period. Finally, a pseudoradial flow
is attained as the flux in the fracture has been stabilized. As the Well-1
value of kmr decreases, the pressure drop and its derivative curves
increase because of strong non-Darcy effect. It is found that the
non-Darcy effect becomes significant when kmr becomes less than
0.50. As kmr approaches 1.00, however, the non-Darcy effect
tends to disappear and the pressure response is found to be very
similar to the Darcy case.
Fig. 6 presents the type curves for kmr ¼ 0.01, 0.10, and 0.50 at
FND ¼5. When FND is set to be a smaller value, the pressure and its
derivative curves also show a trend similar to that of the case at a (a)
higher FND value. However, the non-Darcy effect becomes less
3
significant compared with that at a high FND value. This means 10

CMG
1 This work
10 2
10
This study
Forchheimer equation
dp (psi)

1
10
pwD, dpwD/dlntD

0
10

0
10
–1
10

–1
10 –5 –4 –3 –2 –1 0 1 2 3 4
10 10 10 10 10 10 10 10 10 10
t (hr)
–2
10 –5 –4 –3 –2 –1 0 1 2 3 (b)
10 10 10 10 10 10 10 10 10
tD Fig. 4—(a) Top view of the 3D grid system and (b) model valida-
tion between this study and the Computer Modelling Group
Fig. 3—Model validation for the Forchheimer equation. simulation results.

38 February 2014 SPE Journal


TABLE 2—BASIC PARAMETERS USED FOR SIMULATING PRESSURE-TRANSIENT RESPONSE

Parameter Value Unit

Reservoir and Fluid Data Reservoir permeability, km 0.10 md


Reservoir porosity, / 0.15 fraction
Pay-zone thickness, h 20 ft
Total compressibility, ct 9.210–5 psi–1
Viscosity, l 0.026 cp
Density, q 0.0612 lbm/ft3
Production rate, Qsc 0.1 Mscf/D
Initial reservoir pressure, pi 5,000 psi
Hydraulic-Fracture Data Fracture half-length, xf 100 ft
Fracture width, wf 0.02 ft
Fracture permeability, kf 5,000 md
Equivalent b (Forchheimer) 2.332107 1/ft
Relative minimum permeability, kmr 0.1 fraction
Characteristic length, s 2.86104 1/ft

1 1
10 10

Linear
Flow
0 0
10 10
pwD, dpwD/dlntD

pwD, dpwD/dlntD
Bilinear
Flow
Pseudo-
–1 Radial Flow –1
10 10

FND = 5, kmr = 0.01


FND = 100, kmr = 0.01
–2 –2 FND = 5, kmr = 0.10
10 FND = 100, kmr = 0.10 10
FND = 5, kmr = 0.50
FND = 100, kmr = 0.50
kmr = 1.00 (Darcy)
kmr = 1.00 (Darcy)

–3 –3
10 –5 –4 –3 –2 –1 0 1 2 3
10
–5 –4 –3 –2 –1 0 1 2 3
10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10
tD tD

Fig. 5—Type curves for non-Darcy effect of relative minimum Fig. 6—Type curves for non-Darcy effect of relative minimum
permeability kmr 5 0.01, 0.10, 0.50, and 1.00 at FND 5 100, permeability kmr 5 0.01, 0.10, 0.50, and 1.00 at FND 5 5,
respectively. respectively.

derivative curves increase as the non-Darcy number FND increased to 1.00, the Barree-Conway model can be converted
increases. This is attributed to the fact that the non-Darcy effect is into Darcy’s model so that the flow will approach the Darcy flow
increased as FND is increased. Obviously, as relative minimum condition. As can be seen in Figs. 5 and 7, the non-Darcy effect
permeability kmr is decreased, the non-Darcy effect becomes will become more evident at a smaller relative minimum perme-
much stronger. Meanwhile, the non-Darcy number FND becomes ability (kmr < 0.5) and a larger non-Darcy number (FND > 1).
more sensitive than that at a larger kmr. This is because, as kmr is Comparing Fig. 6 with Fig. 8, it can be found that both relative

1 1
10 10

0 0
10 10
pwD, dpwD/dlntD
pwD, dpwD/dlntD

–1 –1
10 10

kmr = 0.01, FND = 100 kmr = 0.10, FND = 100


–2 kmr = 0.01, FND = 10 –2 kmr = 0.10, FND = 10
10 10
kmr = 0.01, FND = 1 kmr = 0.10, FND = 1
kmr = 1.00 (Darcy) kmr = 1.00 (Darcy)

–3 –3
10 –5 –4 –3 –2 –1 0 1 2 3
10 –5 –4 –3 –2 –1 0 1 2 3
10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10
tD tD

Fig. 7—Type curves for non-Darcy effect with non-Darcy num- Fig. 8—Type curves for non-Darcy effect with non-Darcy num-
ber FND 5 1, 10, and 100 at kmr 5 0.01, respectively. ber FND 51, 10, and 100 at kmr 5 0.10, respectively.

February 2014 SPE Journal 39


TABLE 3—BASIC PARAMETERS USED FOR SIMULATING PRESSURE-TRANSIENT
RESPONSE IN A TIGHT OIL FORMATION

Parameter Value Unit

Reservoir and Fluid Data Reservoir permeability, km 5.00 md


Reservoir porosity, / 0.18 fraction
Pay-zone thickness, h 20 ft
Total compressibility, ct 1810–6 psi–1
Viscosity, l 1.8 cp
Density, q 60 lbm/ft3
Production rate, Qsc 195 STB/D
Hydraulic-Fracture Data Fracture half-length, xf 200 ft
Fracture width, wf 0.02 ft
Fracture permeability, kf 10,000; 2,000 md
Fracture conductivity, CfD 100; 20 dimensionless

TABLE 4—BASIC PARAMETERS USED FOR SIMULATING PRESSURE-TRANSIENT


RESPONSE IN A TIGHT GAS FORMATION

Parameter Value Unit

Reservoir and Fluid Data Reservoir permeability, km 0.01 md


Reservoir porosity, / 0.15 fraction

Reservoir temperature, T 710 R
Pay-zone thickness, h 20 ft
Total compressibility, ct 1.6310–4 psi–1
Viscosity, l 0.018 cp
Density, q 0.0511 lbm/ft3
Production rate, Qsc 1 MMscf/D
Hydraulic-Fracture Data Fracture half-length, xf 200 ft
Fracture width, wf 0.02 ft
Fracture permeability, kf 10,000; 2,000 md
Fracture conductivity, CfD 100; 20 dimensionless

minimum permeability kmr and non-Darcy number FND can affect Darcy flow on flow performance in a tight oil and gas formation,
the pressure response. However, non-Darcy number FND, which the basic data set (Cinco-Ley et al. 1978; Guppy et al. 1982a;
usually results in a large pressure drop at a large kmr, is more sen- Umnuayponwiwat et al. 2000) used in this study is tabulated in
sitive than the relative minimum permeability kmr. Similarly, Fig. Tables 3 and 4, respectively. It is assumed that fracture diffusiv-
8 shows that pressure drop and its derivative curves increase as ity (Cg) and relative minimum permeability (kmr ) are to be 105
the non-Darcy number is increased. and 0.01, respectively.

Effect of Non-Darcy Flow in Tight Formations Tight Oil Formations. Although the effect of non-Darcy flow on
Because of the low production rate in a tight formation, the non- the transient pressure responses of hydraulically fractured wells
Darcy effect on transient pressure behavior has not yet been con- has been examined, few attempts have been made in tight oil for-
sidered. To respectively analyze and discuss the effect of non- mations at a low production rate. In the tight-oil-formation exam-
ple, the reservoir and fluid data are sourced from Cinco-Ley et al.
(1978). As for a tight oil formation with km ¼ 5.00 md, it is
4
10 assumed that there is a relatively high-conductivity fracture with
CfD ¼ 100.0. As shown in Eq. 14, FND is a rate-dependent factor.
Therefore, oil-production rate is assumed to be constant (195
Derivative, Δpwf, dΔpwf /ln(t)
Pressure Change and Its

10
3 STB/D) so that FND is a function only of s. Barree and Conway
(2004) suggested that s is related to the particle size or size distri-
bution of the porous media. In this study, the value of s is set
2
according to the examples provided by Al-Otaibi and Wu (2011).
10 As can be seen in Fig. 9, when s is increased from 100 to 10,000
ft–1, the value of FND is set to be 22.001, 2.200, and 0.220,
τ = 100, FND = 22.001
τ = 1000, FND = 2.200
whereas responses for two Darcy flow scenarios with CfD ¼ 47.9
10
1
τ = 10000, FND = 0.220 and CfD ¼ 100.0 are included.
kmr = 1.00 (Darcy) Comparing the response between non-Darcy flow and Darcy
kmr = 1.00, CfD = 47.9 flow (CfD ¼ 100.0), non-Darcy flow is found to induce an addi-
0 tional pressure drop. As s is decreased, FND is increased, leading
10 –3 –2 –1 –0 1 2 3 4
10 10 10 10 10 10 10 10 to a strong non-Darcy flow effect. As observed by Guppy et al.
t, Hours (1982a) and Umnuayponwiwat et al. (2000), the additional pres-
sure drop because of the non-Darcy effect can also be expressed
Fig. 9—Type curves for non-Darcy effect in a tight oil formation by a reduction in the apparent conductivity of the fracture. In Fig.
at CfD 5 100.0. 7, the Darcy scenario of CfD ¼ 47.9 can be matched with the non-

40 February 2014 SPE Journal


2.0 0.5
tD = 1.0E–5
tD = 1.0E–5
tD = 1.0E–4
tD = 1.0E–4
tD = 1.0E–2 0.4
tD = 1.0E–2
1.5 tD = 1.0E0
tD = 1.0E0
tD = 1.0E3
tD = 1.0E3
0.3

qND
qfD

1.0

0.2

0.5
0.1

0.0 0.0
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
xD xD

Fig. 10—Flux distribution from the matrix to the fracture in a Fig. 11—Flux distribution inside the fracture in a tight oil forma-
tight oil formation at CfD 5 100.0 and FND 5 22.001. tion at CfD 5100.0 and FND 5 22.001.

Darcy scenario of CfD ¼ 100.0 and FND ¼ 2.200 all the time. The bore from the formation through the fracture enter at the fracture
traditional analysis method can result in more than 50% error in the toe.
estimation of conductivity. This is the reason that the type curves For comparison, a relatively low-conductivity fracture is con-
that can be matched with the conventional Darcy model would sidered with CfD ¼ 20.0 in the same formation by changing the
yield an underestimated fracture conductivity if the non-Darcy fracture permeability to a smaller value. As depicted in Fig. 12,
effect exists (Guppy et al. 1982a, 1982b; Umnuayponwiwat et al. FND is still the same as that in a high-conductivity fracture. The
2000). Therefore, the fracture half-length together with other frac- transient pressure and its derivative curves show a trend similar to
ture parameters cannot be accurately determined if the apparent those in a high-conductivity fracture, except that a larger pressure
fracture conductivity is evaluated with the conventional method. drop can be obtained at the early stages. This is ascribed to the fact
To better understand the non-Darcy effect, Fig. 10 presents the that low fracture conductivity can also cause an additional pressure
flux distribution from the matrix to the fracture as a function of drop, which is similar to that of the non-Darcy effect. Fig. 13
time at FND ¼ 22.001. In this case with a strong non-Darcy effect, presents flux distribution along the fracture with a low conductiv-
it is found that the fluid enters from the heel of the fracture at the ity. As shown in Fig. 13, the flux at early time is larger than that in
very beginning, resulting in a large pressure drop until the pseu- the high-conductivity fracture at the same location, directly leading
doradial flow is established. As time proceeds, the flux decreases to a larger pressure drop. In the late stage, the flux distribution is
at the heel of the fracture but increases at its toe. This is ascribed to similar to that of a high-conductivity scenario, indicating that the
the fact that a strong non-Darcy effect causes resistance in the frac- fluid entering the fracture enters less at its heel and more at its toe.
ture, forcing the fluids to largely enter from the heel at the begin- In practice, bilinear-flow analysis proposed by Cinco-Ley and
ning of well production. When pseudoradial flow is established, Samaniego-V. (1981) is widely used to estimate fracture conduc-
the flux distribution stabilizes and remains unchanged. Because tivity. It is worthwhile, however, to examine how non-Darcy
the resistance force becomes smaller, most of the fluids enter the behavior will affect bilinear-flow behavior (Umnuayponwiwat
fracture at its toe, causing a smaller pressure drop. It should be et al. 2000). As shown in Fig. 14, the pressure change is increased
noted that the additional pressure drop caused by the non-Darcy following a straight line in the one-fourth of time at a low conduc-
effect is rate dependent and that the flux distribution along the frac- tivity. This is because bilinear flow exists at the beginning of pro-
ture is a function of time until the pseudoradial flow begins. duction. At the same one-fourth of time, the pressure change is
To perform further illustration, flux distribution inside the reduced as s increases. This means that non-Darcy flow leads to
fracture is plotted in Fig. 11. At the beginning of well production, an increase in the pseudopressure drop. Similarly, Fig. 15 shows
the flux inside the fracture generally decreases at each node, the same trend for a scenario with a higher conductivity. Obvi-
though it is decreased sharply in the half of the fracture closer to ously, there exists non-Darcy flow behavior in a tight oil forma-
the wellbore. This is because most of the fluids entering the well- tion even at a low production rate.
3.0
4
10 tD = 1.0E–5
tD = 1.0E–4
2.5
Derivative, Δpwf, dΔpwf / ln(t)

tD = 1.0E–2
Pressure Change and Its

3 tD = 1.0E0
10 tD = 1.0E3
2.0
qfD

2 1.5
10

τ = 100, FND = 22.001 1.0


1
τ = 1000, FND = 2.200
10 τ = 10000, FND = 0.220
0.5
kmr = 1.00 (Darcy)

10 –3
0 0.0
–2 –1 –0 1 2 3 4 0.0 0.2 0.4 0.6 0.8 1.0
10 10 10 10 10 10 10 10
t, Hours xD

Fig. 12—Type curves for non-Darcy effect in a tight oil forma- Fig. 13—Flux distribution from the matrix to the fracture in a
tion at CfD 5 20.0. tight oil formation at CfD 5 20.0 and FND 5 22.001.

February 2014 SPE Journal 41


100
140
τ = 100, FND = 22.001
τ = 100, FND = 22.001
120 τ = 1000, FND = 2.200

Pressure Change, Δpw


80
Pressure Change, Δpw
τ = 1000, FND = 2.200
τ = 10000, FND = 0.2200
τ = 10000, FND = 0.2200
100
60
80

60 40

40
20
20

0 0
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
Δt1/4, Hours Δt1/4, Hours

Fig. 14—Bilinear-flow analysis with non-Darcy effect at Fig. 15—Bilinear-flow analysis with non-Darcy effect at
CfD 5 20.0. CfD 5 100.0.

11 2.5
10
tD = 1.0E–5
Derivative, Δm(pwf), dΔm(pwf) /ln(t)
Pseudo-Pressure Change and Its

tD = 1.0E–4
2.0 tD = 1.0E–2
10 tD = 1.0E0
10
tD = 1.0E3

1.5
qfD
9
10
1.0
τ = 1000, FND = 121.328
8 τ = 5000, FND = 24.266
10 τ = 10000, FND = 12.133
kmr = 1.00 (Darcy)
0.5
kmr = 1.00, CfD = 9.65

7
10 –2 –1 –0 1 2 3 4 5 6 7 0.0
10 10 10 10 10 10 10 10 10 10 0.0 0.2 0.4 0.6 0.8 1.0
t, Hours xD

Fig. 16—Type curves for non-Darcy effect in a tight gas forma-


tion at CfD 5 100.0. Fig. 17—Flux distribution from the matrix to the fracture in a
tight gas formation at CfD 5 100.0 and FND 5 121.328.

Tight Gas Formations. The non-Darcy effect has been widely


discussed and analyzed in gas formations (Martins et al. 1990;
Vincent et al. 1999; Smith et al. 2004). The gas-production rate is seen, there is a large additional pseudopressure drop at the early
set to be constant (1 MMscf). In the tight gas formation, the value stages. Comparing the curve of FND ¼121.328 with the Darcy
of FND is set to be 121.328, 24.266, and 12.133 as s is increased condition, it is found that the pressure drop is larger than that of
from 1,000 to 10,000 ft–1, whereas responses for two Darcy flow the tight oil formation. This is because the FND can be set to be a
scenarios with CfD ¼ 9.6 and CfD ¼ 100.0 are included. larger value for a higher gas-production rate. It is also found that
Fig. 16 presents the type curves for the non-Darcy effect in a the Darcy scenario of CfD ¼ 9.6 can be matched with the the non-
tight gas formation with high fracture conductivity. As can be Darcy scenario of CfD ¼ 100.0 and FND ¼ 24.266. Comparing
with the case in tight oil formation, the traditional method will
cause a larger error in the estimation of conductivity because non-
11
10 Darcy effect is usually larger in the tight gas formation. Fig. 17
Derivative, Δm(pwf), dΔm(pwf) / ln(t)
Pseudo-Pressure Change and Its

shows the flux distribution from the matrix to the fracture at a


high fracture conductivity. As shown in Fig. 17, the trend is very
10
10 similar to that in the tight oil formation; however, there is more
fluid entering into fracture from the matrix, finally resulting in a
larger pseudopressure drop.
9
Fig. 18 shows the type curves for the non-Darcy effect in a
10 tight gas formation with a low fracture conductivity. Compared
with the high-conductivity-fracture scenario, the bilinear flow is
τ = 1000, FND = 121.328
τ = 5000, FND = 24.266
more significant than that in a high-conductivity fracture mainly
8
10 τ = 10000, FND = 12.133
because of low conductivity other than the non-Darcy effect.
kmr = 1.00 (Darcy) Fig. 19 presents the flux distribution from the matrix to the frac-
ture at low fracture conductivity. As can be seen, the flux is larger
7 than that in a high-conductivity fracture at the early stages. Even
10 –2 –1 –0 1 2 3 4 5 6 7
10 10 10 10 10 10 10 10 10 10 at a late stage, there still exists more fluid entering from the heel
t, Hours than occurs in a high-conductivity fracture, which cannot be
observed in the tight oil formation. This is because both the high
Fig. 18—Type curves for non-Darcy effect in a tight gas forma- non-Darcy number and low fracture conductivity can induce more
tion at CfD 5 20.0. resistance strength on the flux decrease at the heel.

42 February 2014 SPE Journal


3.0 1000
tD = 1.0E–5
4.5

Derivative, Δpwf, dΔpwf /ln(t)


tD = 1.0E–4

Pressure Change and Its


3.0 tD = 1.0E–2
tD = 1.0E0
3.5 tD = 1.0E3
3.0
100
qfD

2.5

2.0

1.5
Field Data
Non-Darcy Flow
1.0 Darcy Flow
0.5
10
0.0 0.1 1 10 100
0.0 0.2 0.4 0.6 0.8 1.0 t, Hours
xD
Fig. 20—Type-curve match of the field data on an oil well in a
Fig. 19—Flux distribution from the matrix to the fracture in a tight formation.
tight gas formation at CfD 5 20.0 and FND 5 121.328.

c ¼ coefficient defined in Eq. 26b, dimensionless


C ¼ wellbore-storage coefficient, bbl/psi
Case Study Cg ¼ fracture diffusivity, fraction
The case study is a drawdown test on an oil well that was first Cf ¼ fracture conductivity, md-ft
reported by Cinco-Ley et al. (1978). Fig. 20 shows the type-curve ct ¼ total compressibility, psi–1
match between the field data on the oil well in a tight formation d ¼ coefficient defined in Eq. 23c, dimensionless
and the solution of the semianalytical model of this study. Two D1 ¼ rate-dependent non-Darcy number, (STB/D)–1 or (Mscf/
cases of Darcy and non-Darcy flow are considered to examine the D)–1
non-Darcy effect on oil wells. As shown in Fig. 20, there exists an FND ¼ non-Darcy number, dimensionless
excellent agreement between the field measurements and solu- h ¼ pay-zone thickness, ft
tions of the model proposed in this study for both Darcy and non- m ¼ pseudopressure, psi2/cp
Darcy flow cases. As for the Darcy case, the dimensionless frac- I0 ðxÞ ¼ modified Bessel function of first kind, zero order
ture conductivity CfD is calculated to be 6.267, which is very close I1 ðxÞ ¼ modified Bessel function of first kind, first order
to the original result of 2p. One of the matches for the non-Darcy k ¼ permeability, md
case corresponds to CfD ¼ 6.274, FND ¼ 0.47, and kmr ¼ 0.12. This K0 ðxÞ ¼ modified Bessel function of second kind, zero order
result is consistent with the previous analysis that the non-Darcy p ¼ pressure, psi
effect in a tight oil formation is usually small. It should be noted Qsc ¼ flow rate, STB/D or Mscf/D
that it is possible to obtain multiple combinations of Darcy and r ¼ radius, ft
non-Darcy parameters by adjusting some parameters. Therefore, t ¼ time, hours
it is recommended that at least two flow tests be required to accu- u ¼ Laplace-transform variable
rately determine the fracture conductivity, relative minimum per- v ¼ velocity, ft/sec
meability, and non-Darcy number. wf ¼ fracture width, ft
xf ¼ fracture half-length, ft
Conclusions b ¼ non-Darcy flow coefficient, ft–1
In this study, a mathematical model incorporating the Barree- d ¼ parameter defined in Eq. 4, dimensionless
Conway model has been developed to analyze non-Darcy flow l ¼ fluid viscosity, cp or pseudotime
behavior in a hydraulic fracture. The newly developed technique q ¼ fluid density, lbm/ft3
is validated with the traditional Forchheimer model and further s ¼ characteristic length, ft–1
with numerical simulation. The pressure response is determined / ¼ porosity, fraction
by three parameters: fracture conductivity (CfD), non-Darcy num-
ber (FND), and relative minimum permeability (kmr). The FND is Subscripts
more sensitive than the relative minimum permeability (kmr), app ¼ apparent
resulting in a larger pressure drop even at a larger kmr. In a tight D ¼ dimensionless
gas formation, it is found that significant non-Darcy effect can e ¼ external boundary
exist even at a low production rate, whereas in a tight oil forma- f ¼ fracture
tion, it is not easy to observe such a non-Darcy effect because of i ¼ initial condition, or index, ith node
the low production rate. However, when s is decreased, the non- j ¼ number of summation
Darcy effect becomes significant even in a tight oil formation m ¼ matrix
with a low production rate. When the non-Darcy effect exists, the mr ¼ minimum relative
traditional analysis method can result in more than 50% error in N ¼ node
the estimation of conductivity. Finally, a field case is analyzed on sf ¼ sandface
the basis of the newly developed semianalytical model to show w ¼ wellbore
that the results are consistent with theoretical studies. It is sug- wf ¼ wellbore flowing
gested that at least two flow tests be performed to accurately eval-
uate fracture conductivity, relative minimum permeability, and
non-Darcy number. Acknowledgments
The authors acknowledge a Discovery Grant from the Natural Sci-
Nomenclature ences and Engineering Research Council of Canada and an inno-
b ¼ coefficient defined in Eq. 23a, dimensionless vation fund from the Petroleum Technology Research Centre to
Bg ¼ formation volume factor, bbl/STB or bbl/Mscf D. Yang.

February 2014 SPE Journal 43


References Effects. J. Cdn. Pet. Tech. 39 (5): 56–63. http://dx.doi.org/10.2118/00-
Al-Otaibi, A.M. and Wu, Y.S. 2011. An Alternative Approach to Model- 05-04.
ing Non-Darcy Flow for Pressure Transient Analysis in Porous and Smith, M.B., Bale, A., Britt, L.K., et al. 2004. An Investigation of Non-
Fractured Reservoirs. Paper SPE 149123 presented at the SPE/DGS Darcy Flow Effects on Hydraulic Fractured Oil and Gas Well Perform-
Saudi Arabia Section Technical Symposium and Exhibition, Al-Kho- ance. Paper SPE 90864 presented at the SPE Annual Technical Con-
bar, Saudi Arabia, 15–18 May. http://dx.doi.org/10.2118/149123-MS. ference and Exhibition, Houston, Texas, 26–29 September. http://
Andrade, J.S. Jr., Almeida, M.P., Medes Filho, J., et al. 1997. Fluid Flow dx.doi.org/10.2118/90864-MS.
through Porous Media: The Role of Stagnant Zones. Phys. Rev. Lett. Stanley, H.E. and Andrade J.S. Jr. 2001. Physics of the Cigarette Filter: Fluid
79 (20): 3901–3904. http://dx.doi.org/10.1103/PhysRevLett.79.3901. Flow through Structures with Randomly-Placed Obstacles. Physica A
Barree, R.D. and Conway, M.W. 2004. Beyond Beta Factors: A Complete 295 (1–2): 17–30. http://dx.doi.org/10.1016/S0378-4371(01)00140-6.
Model for Darcy, Forchheimer, and Trans-Forchheimer Flow in Po- Stehfest, H. 1970. Algorithm 368: Numerical Inversion of Laplace Trans-
rous Media. Paper SPE 89325 presented at the SPE Annual Technical forms [D5]. Commun. ACM 13 (1): 47–49. http://dx.doi.org/10.1145/
Conference and Exhibition, Houston, Texas, 26–29 September. http:// 361953.361969.
dx.doi.org/10.2118/89325-MS. Spivey, J.P. 1984. An Investigation of the Use of Pseudotime in Transient
Cinco-Ley, H. and Samaniego-V., F. 1981. Transient Pressure Analysis Test Analysis of Gas Wells. PhD dissertation, Texas A&M University,
for Fractured Wells. J. Pet Tech 33 (9): 1749–1766. http://dx.doi.org/ College Station, Texas (May 1984).
10.2118/7490-PA. Swami, V., Clarkson, C.R. and Settari, A. 2012. Non-Darcy Flow in Shale
Cinco-Ley, H., Samaniego-V., F. and Domı́nguez A, N. 1978. Transient Nanopores: Do We Have a Final Answer? Paper SPE 162665 pre-
Pressure Behavior for a Well with a Finite-Conductivity Vertical Frac- sented at the SPE Canadian Unconventional Resources Conference,
ture. SPE J. 18 (4): 253–264. http://dx.doi.org/10.2118/6014-PA. Calgary, Alberta, Canada, 30 October–1 November. http://dx.doi.org/
Geertsma, J. 1974. Estimating the Coefficient of Inertial Resistance in 10.2118/162665-MS.
Fluid Flow Through Porous Media. SPE J. 14 (5): 445–450. http:// Umnuayponwiwat, S., Ozkan, E. and Pearson, C.M. 2000. Effect of Non-
dx.doi.org/10.2118/4706-PA. Darcy Flow on the Interpretation of Transient Pressure Response of
Gill, J.A., Ozkan, E. and Raghavan, R. 2003. Fractured-Well-Test Design Hydraulically Fractured Wells. Paper SPE 63176 presented at the SPE
and Analysis in the Presence of Non-Darcy Flow. SPE Res Eval & Annual Technical Conference and Exhibition, Dallas, Texas, 1–4 Oc-
Eng 6 (3): 185–196. http://dx.doi.org/10.2118/84846-PA. tober. http://dx.doi.org/10.2118/63176-MS.
Guppy, K.H., Cinco-Ley, H., Ramey, H.J. Jr., et al. 1982a. Non-Darcy Van Kruysdijk, C.P.J.W. 1988. Semianalytical Modeling of Pressure Tran-
Flow in Wells with Finite-Conductivity Vertical Fractures. SPE J. 22 sients in Fractured Reservoirs. Paper SPE 18169 presented at the SPE
(5): 681–698. http://dx.doi.org/10.2118/8281-PA. Annual Technical Conference and Exhibition, Houston, Texas, 2–5
Guppy, K.H., Cinco-Ley, H. and Ramey H.J. Jr. 1982b. Pressure Buildup October. http://dx.doi.org/10.2118/18169-MS.
Analysis of Fractured Wells Producing at High Flow Rates. J. Pet Vincent, M.C., Pearson, C.M. and Kullman, J. 1999. Non-Darcy and Mul-
Tech 34 (11): 2655–2666. http://dx.doi.org/10.2118/10178-PA. tiphase Flow in Propped Fractures: Case Studies Illustrate the Dra-
Hill, R.J., Koch, D.L. and Ladd, A.J.C. 2001. Moderate Reynolds Number matic Effect on Well Productivity. Paper SPE 54630 presented at the
Flows in Ordered and Random Arrays of Spheres. J. Fluid Mech. 448: SPE Western Regional Meeting, Anchorage, Alaska, 26–27 May.
243–278. http://dx.doi.org/10.1017/S0022112001005936. http://dx.doi.org/10.2118/54630-MS.
Holditch, S.A. and Morse, R.A. 1976. The Effects of Non-Darcy Flow on Zeng, F. and Zhao, G. 2010. The Optimal Hydraulic Fracture Geometry
the Behavior if Hydraulically Fractured Gas Wells. J. Pet Tech 28 Under Non-Darcy Flow Effects. J. Pet. Sci. Eng. 72 (1–2): 143–157.
(10): 1169–1179. http://dx.doi.org/10.2118/5586-PA. http://dx.doi.org/10.1016/j.petrol.2010.03.012.
Lai, B., Miskimins, J.L. and Wu, Y.S. 2012. Non-Darcy Porous Media
Feng Zhang is an MASc graduate student in petroleum systems
Flow According to the Barree and Conway Model: Laboratory and Nu-
engineering in the Faculty of Engineering and Applied Sci-
merical Modeling Studies. SPE J. 17 (1): 70–79. http://dx.doi.org/ ence at the University of Regina, Regina, Saskatchewan, Can-
10.2118/122611-PA. ada. He holds a BSc degree in geology engineering from the
Lee, W.J. and Holditch, S.A. 1982. Application of Pseudotime to Buildup China University of Petroleum, Beijing, China. His main research
Test Analysis of Low-permeability Gas Wells with Long-Duration interests include non-Darcy flow in porous media, pressure-
Storage Distortion. J. Pet Tech 34 (12): 2877–2887. http://dx.doi.org/ transient analysis, reservoir modeling, and simulation. He is a
10.2118/9888-PA. member of SPE.
Manrique, M., Muci, V.E. and Gurfinkel, M.E. 2007. EOR Field Experien- Daoyong Yang is a professor in and the chair of the Petroleum
ces in Carbonate Reservoirs in the United States. SPE Res Eval & Eng Systems Engineering Program in the Faculty of Engineering
10 (6): 667–686. http://dx.doi.org/10.2118/100063-PA. and Applied Science at the University of Regina. He holds BSc
Manrique, M., Thomas, C., Ravikiran, R., et al. 2010. EOR: Current Status and PhD degrees in petroleum engineering from the China
and Opportunities. Paper SPE 130113 presented at the SPE Improved University of Petroleum and a PhD degree in petroleum sys-
Oil Recovery Symposium, Tulsa, Oklahoma, 24–28 April. http:// tems engineering from the University of Regina. Previously,
dx.doi.org/10.2118/130113-MS. Yang worked for 3 years as a petroleum engineer in the Petro-
China TuHa Oilfield Company. He also worked for 4 years as a
Martins, J.P., Milton-Tayler, D. and Leung, H.K. 1990. The Effects of
reservoir engineer in the Software Development and Informa-
Non-Darcy Flow in Propped Hydraulic Fractures. Paper SPE 20709 tion Centre of Petroleum Engineering. Yang’s major research
presented at the SPE Annual Technical Conference and Exhibition, areas include reservoir modeling and simulation, formation
New Orleans, Louisiana, 23–26 April. http://dx.doi.org/10.2118/ evaluation, phase behavior, production optimization, CO2
20709-MS. enhanced oil recovery and storage, pressure-transient analy-
Miskimins, J.L., Lopez-Hernandez, H.D. and Barree, R.D. 2005. Non- sis, artificial-lift methods, interfacial interactions in enhanced-
Darcy Flow in Hydraulic Fractures: Does It Really Matter? Paper SPE oil-recovery processes, heavy-oil recovery, and unconven-
96389 presented at the SPE Annual Technical Conference and Exhibi- tional-resources exploitation. He has authored or coauthored
tion, Dallas, Texas, 9–12 October. http://dx.doi.org/10.2118/96389- more than 90 refereed-journal papers and conference
papers. Yang was the recipient of the SPE Canada Region Re-
MS.
gional Formation Evaluation Award in 2011 and the SPE Can-
Ozkan, E. and Raghavan, R. 1991. New Solutions for Well-Test-Analysis ada Region Regional Reservoir Description and Dynamics
Problems: Part 1–Analytical Considerations. SPE Form Eval 6 (3): Award in 2013. He is an active member of SPE, the American
359–368. http://dx.doi.org/10.2118/18615-PA. Chemical Society, and the Association of Professional Engi-
Settari, A., Stark, A.J. and Jones, J.R. 2000. Analysis of Hydraulic Fractur- neers and Geoscientists of Saskatchewan and is a registered
ing of High Permeability Gas Wells to Reduce Non-Darcy Skin professional engineer in Saskatchewan, Canada.

44 February 2014 SPE Journal

Вам также может понравиться