You are on page 1of 24

SPE-170698-MS

Rate Transient and Decline Curve Analyses for Continuously


(Dual-Porosity) and Discretely Naturally Fractured Reservoirs
Fikri Kuchuk, SPE; Denis Biryukov and Tony Fitzpatrick, Schlumberger

Copyright 2014, Society of Petroleum Engineers

This paper was prepared for presentation at the SPE Annual Technical Conference and Exhibition held in Amsterdam, The Netherlands, 27–29 October 2014.

This paper was selected for presentation by an SPE program committee following review of information contained in an abstract submitted by the author(s). Contents
of the paper have not been reviewed by the Society of Petroleum Engineers and are subject to correction by the author(s). The material does not necessarily reflect
any position of the Society of Petroleum Engineers, its officers, or members. Electronic reproduction, distribution, or storage of any part of this paper without the written
consent of the Society of Petroleum Engineers is prohibited. Permission to reproduce in print is restricted to an abstract of not more than 300 words; illustrations may
not be copied. The abstract must contain conspicuous acknowledgment of SPE copyright.

Abstract
Forecasting production rates and reserves is important for reservoir management. Analysis of long-term
production data has traditionally been performed using the empirical Arps (1945) decline curves to predict
future production. Fetkovich (1980) combined Arps decline curves with constant-wellbore pressure
solutions to offer a new method for decline-curve analysis. Furthermore, both rate-transient and decline-
curve analyses for naturally fractured reservoirs were performed using Arps (1945) decline curves or
Warren and Root (1963) dual-porosity-type models. These approaches have yielded unsatisfactory results
for naturally fractured reservoirs. Recent studies show that the Warren and Root (1963) dual-porosity-type
models, which do not contain fractures, cannot characterize highly diverse transient behaviors of
continuously and discretely fractured reservoirs because they are inappropriate and incomplete for most
naturally fractured reservoirs. Furthermore, no investigation has been performed on rate-transient behav-
ior of discretely fractured reservoirs.
We investigate rate-transient behavior of continuously (dual-porosity) and discretely naturally frac-
tured reservoirs using semianalytical solutions. These fractured reservoirs can contain periodically or
arbitrarily distributed finite-and/or infinite-conductivity fractures, of different lengths and orientations.
Our results demonstrate that, in terms of rate transients, neither continuously nor discretely fractured
reservoirs behave like the Warren and Root (1963) dual-porosity type model. There are many factors that
dominate the rate-transient behavior of naturally fractured reservoirs, such as fracture conductivities, dip
angles, lengths, and distributions, as well as whether or not fractures intersect the wellbore. Rate transients
associated with these factors are shown for a few continuously and discretely fractured reservoirs with
different well and fracture configurations.
The inverse of the dimensionless pressure (pD) is not a good approximation for the dimensionless rate
(qD), but for most cases, their derivatives behave similarly, and exhibit same or
similar flow regimes. For some cases, they behave very differently. The similarities or variations are true
for any type of reservoirs.
For any reservoir, Arps’ decline curves yield unreasonably high production rates and cumulative
productions, except the exponential decline curve. The exponential decline curve should not be used
without taking into account the change of the wellbore pressure as a function of time. Arps’ decline curves
2 SPE-170698-MS

analysis should not be used for both discretely and continuously fractured reservoirs. An integrated
interpretation methodology is outlined for rate transient analysis in fractured reservoirs.
A few examples for the rate transient behavior of the continuously and discretely fractured fractured
reservoirs are presented. They exhibit different flow regimes depending on fracture distribution, intensity,
and conductivity. We also compare the rate transient derivatives with pressure transient derivatives of
these examples. For most cases, they exhibit similar behaviors. These derivatives and flow regimes
presented are valuable diagnostic tools for rate transient analysis.

Introduction
It has been exactly 101 years since the publication of the first article on the estimation of hydrocarbon-
in-place and reserves for oil reservoirs based on the production history of wells (Huntley, 1913), and 99
years since the publication of the second article (Washburne, 1915). Reserve estimation and production
forecasting continue to be a challenging problem. There is still no standard method agreed upon by the
industry, although they are essential for determining the economic viability of a field, as well as for
booking reserves required by regulatory bodies. For over a century, various methods based on historical
production data have been developed for estimating reserves and predicting performance. Four main
methods are currently used: 1) Material balance (Bertuzzi et al., 1987; Craft and Hawkins, 1959; Muskat,
1945), 2) Decline-type curve analysis (Arps, 1945; Cutter, 1924; Fetkovich, 1980), 3) Rate-transient
analysis (Blasingame et al., 1991; Fetkovich, 1980; Fetkovich et al., 1987; Fraim and Wattenbarger, 1987;
Kuchuk et al., 2005; Palacio and Blasingame, 1993), and 4) Numerical simulation (Carter, 1985).
Raghavan (1993) has given details of the first three methods. Of course, many more papers have been
published in the literature on this subject.
A number of papers have been published on the rate transient behavior of naturally fractured (also
called two-porosity, double-porosity, dual-porosity, or fissured) reservoirs using Barenblatt et al. (1960)
and Warren and Root (1963) type models (Camacho-Velazquez et al., 2008; Da Prat et al., 1981; Ozkan
et al., 1987; Raghavan and Ohaeri, 1981; Sageev et al., 1985). A few papers are published on the rate
transient behavior of triple-porosity (three-porosity) naturally fractured reservoirs (Al-Ahmadi and
Wattenbarger, 2011; Ezulike and Dehghanpour, 2013; Tivayanonda et al., 2012). To our knowledge,
however, there is no study on the rate transient behavior of discretely fractured systems.
The overall objective of this paper is to investigate the rate-transient behavior of continuously
(fractures create a network and communicate hydraulically with each other globally) and discretely
naturally fractured reservoirs using semianalytical solutions. These fractured reservoirs can contain
periodically or arbitrarily distributed finite- and infinite-conductivity fractures, or both, with different
density, lengths, and orientations.

Reservoir Engineering Considerations


Aside from fractured basement reservoirs, naturally fractured reservoirs can be described as sedimentary
formations fractured by tectonic, thermal, or chemical processes. Although Nelson (1985) categorized
naturally fractured reservoirs into four different types, with regard to reservoir simulations and perfor-
mance, and pressure and rate transient test interpretation, Kuchuk and Biryukov (2014) divide fractured
reservoirs into four categories as follows:
1. Fractures create a network, communicate hydraulically with each other globally, and provide the
overall conductivity (permeability) of the reservoir. The matrix provides the overall storage
capacity (porosity), but it should be permeable enough to provide conductivity for flow from the
matrix into the fractures. If matrix permeabilities are high, or similar to fracture permeabilities, or
ultralow, pressure and rate transient behaviors of these types of fractured reservoirs will be similar
to the behavior of homogeneous reservoirs. We call them continuously fractured reservoirs. They
SPE-170698-MS 3

can also be called dual-porosity-permeability fractured reservoirs. In a sense, both fracture and
matrix have distinct porosities and permeabilities. Often, joints create a set of orthogonal
continuous fracture networks (Pollard and A., 1988). Naturally fractured reservoirs may have
nonorthogonal continuous fracture sets (Pollard and A., 1988), as well as secondary fractures that
are nonorthogonal to each other and to primary fracture sets. As stated above, secondary or micro
fractures may create triple-porosity fractured reservoir behaviors. Dual-permeability systems do
not make sense for naturally fractured reservoirs, except multilayered reservoirs in which some
layers contain fractures and some do not.
2. Fractures do not form a continuous conductive network; only a limited number of fractures
communicate hydraulically with each other locally. Fractures and matrix elements provide
conductivity, but the overall storage capacity (porosity) is in the matrix. These are usually called
discretely fractured reservoirs. It should be pointed out that open and partially sealing faults also
create discrete conductive systems, particularly with secondary or micro fractures. Stratified
reservoirs that consist of both fractured and nonfractured layers stacked vertically should also be
called discretely fractured reservoirs. In reservoirs with a continuous network of faults and/or
fractures, if some are mineralized and/or closed, this process also creates discretely fractured
reservoirs. Many discretely naturally fractured reservoirs have nonorthogonal discontinuous frac-
ture sets (Pollard and A., 1988), whereas only a few have orthogonal discontinuous fracture sets.
3. Fractures are nonconductive (sealed due to mineralization or faults caused by shale smearing).
Both conductivity (permeability) and storage capacity are provided by the matrix. These are called
compartmentalized reservoirs. However, there can be many conductive faults and fractures in a
given compartment. Before showing boundary dominated behavior, these systems can also show
fractured reservoir behavior.
4. Only fractures are conductive; the matrix has no permeability and/or porosity. Reservoirs with
these fractures are called unconventional fractured basement reservoirs.
The above definitions are applied to both conventional and unconventional naturally fractured reser-
voirs, such as shale and tight gas and oil formations. Although there are plenty of naturally fractured
sandstone reservoirs, in general, fractures are abundant in carbonate formations. Nelson (1985) gave an
extensive list of fractured reservoirs, such as Spraberry trend, West Texas; Miocene Monterey Formation,
California; Asmari Formation, Iran; Ain Zalah and Kirkuk fields, Iraq; Dukhan field, Qatar; Cantarell
Mexico; Fahud and Yibal in Oman. Many of these reservoirs contain continuously and discretely fractured
regions depending on the location, whether at the crest or flank.
All naturally fractured reservoirs, particularly carbonate formations, contain secondary (minor) and
micro fractures and vugs (small cavities in rock), where, as stated above, they may exhibit triple-porosity
behavior because their conductivities are significantly lower than those of major fractures. Also, they are
easily filled with sediments, or mineralized, because of their small apertures (see Fig. 1). However, vugs,
unless they are highly connected, cannot create triple-porosity behavior; see Figs. 8 and 9 in Kuchuk and
Biryukov (2014). On the other hand, Abdassah and Ershaghi (1986) and Liu (1981) assumed that fractured
reservoirs may have two different matrix block types with two ␻ and ␭ values for each, where the REV
(Representative Elementary Volume) is uniformly distributed throughout the reservoir. It is highly
unlikely that this level of partitioning at the macroscopic level in the REV can be done at the scale of the
macroscopic Darcy’s law. In other words, pressure diffusion is not sensitive to this level of partitioning
of porous media, including small vugs, unconnected micro fractures, intergranular macroporosity and
intragranular microporosity, etc. For instance, if vugs are directly connected to the fracture system, then
they are a part of the fracture system. If they are isolated, then they are a part of the matrix. Therefore,
having multiple matrix block sizes with different properties in a single REV is not useful for either
4 SPE-170698-MS

Figure 1—Schematic of a naturally fractured triple-porosity formation, where macrofractures, minor (secondary) fractures, and microfractures
provide the conductivity of the system.

Figure 2—Directionally but arbitrarily distributed fractures with dif-


ferent conductivities in a discretely fractured reservoir, where the well is Figure 3—Derivatives for a number of realizations of the discretely
located in the matrix at the center of the coordinate system x ⴝ 0 and y fractured reservoir model shown in Fig. 2.
ⴝ 0, after Morton et al. (2013).

pressure or rate transient tests in fractured reservoirs or in any reservoir for that matter. Of course, in
fractured reservoirs, the matrix block size and its permeability may vary significantly spatially.
The model shown in Fig. 2 was presented by Morton et al. (2013) for a discretely fractured reservoir.
The fractured reservoir solution of Biryukov and Kuchuk (2012) is used to generate the wellbore pressure
data with the input model parameters that are given in Table 2 of the Morton et al. (2013) paper. In
addition to the model parameters, fracture density, length, and spacing have truncated lognormal
distributions and fracture conductivity has a lognormal distribution in the model. The model consists of
about 300 fractures, as shown in Fig. 2, and all fractures are predominately in the East to West direction,
extending about 1640 ft (500 m) from the well. Figure 3 presents derivatives for a number of realizations
SPE-170698-MS 5

Figure 4 —Schematic of p/z versus cumulative production from a natu- Figure 5—Schematic of p/z versus cumulative production from a natu-
rally fractured reservoir. rally fractured reservoir with minor and micro fractures.

of the model. As can be seen from this figure, the derivatives for different cases exhibit many different
look-alike (not identical) behaviors: dual-porosity pseudo-steadystate and transient matrix flow, radially
composite, triple-porosity, etc. Of course, the Morton et al. (2013) model has nothing to do with the
Warren and Root (1963) type models. Therefore, the triple-porosity behavior observed from the well test
in naturally fractured reservoirs is most likely due to the variability of fracture properties: their density,
distributions, and conductivities. It is unlikely that they are due to the microscopic level local variability
of matrix properties.
Conventional and unconventional naturally fractured reservoirs exhibit different depletion patterns
(production regimes that are similar to flow regimes) obtained from material-balance equations that
provide the relationship between the average reservoir pressure (p) divided by the compressibility factor
(z) at a given time, and the corresponding cumulative fluid production G [see (Ahmed and McKinney,
2011)]. Thus, a plot of p/z versus G exhibits a straight line for homogenous reservoirs, which yields the
original gas in place (OGIP). Although the p/z versus G plot is primarily used for gas reservoirs, it is also
used for oil reservoirs. This method has been modified for different systems [see (King, 1993; Moghadam
et al., 2011; Rahman et al., 2006)]. But this method still has many uncertainties at early-times and for
heterogeneous and different fluid systems.
In some naturally fractured reservoirs, first a fracture-network-dominated pseudosteady-state flow
regime will be observed if the matrix is very tight and the fractured system has a finite volume. The
fractured system volume can be finite due to a limited extent of the fracture zone, no-flow boundaries,
sealing faults and fractures, and/or no-flow boundaries due to other producing wells. In this case, the fluid
in fractures is depleted quickly, and is not replenished readily by matrix elements. This normally takes
place before the formation linear flow (although there are exceptions). As the pressure declines more and
more in the fracture system, the matrix contributions become significant, and the flow regime deviates
from the fracture-network pseudosteady-state flow. Figure 4 presents a schematic representation of p/z
versus G for a naturally fractured reservoir exhibiting different depletion patterns: 1) fracture-network
dominated and 2) matrix dominated. We may also observe a similar behavior for shale gas reservoirs when
gas adsorption takes place in the pores of the matrix. As is well known, there are other reservoir support
mechanisms, such as water influx, that could also cause a deviation from the initial straight line. In
addition to the macrofracture-network, if the formation contains an extensive volume of connected minor
(secondary) fractures, and micro fractures, as shown in Fig. 1, we may observe another deviation before
any matrix-dominated volumetric behavior. Figure 5 presents a schematic representation of p/z versus G
for a naturally fractured reservoir exhibiting three different depletion patterns: 1) fracture-network
dominated, 2) minor and micro fracture-network dominated, and 3) matrix dominated. When we add
measurement noise (errors) of average reservoir pressure (p) and the corresponding cumulative fluid
6 SPE-170698-MS

production G (see the points with noise in Fig. 5),


the convexity of p/z versus G increases. If we add
gas adsorption in the case of gas shale reservoirs
after the beginning of matrix depletion in the p/z
versus G curve (Fig. 5), then the curve flattens
further for a long time.
Figure 6 presents a plot of p/z versus G for a
naturally fractured shale gas reservoir. The plot
exhibits three depletion periods and becomes almost
totally flat. This behavior of actual field data is very Figure 6 —Values of p/z versus cumulative production G from Well No.
similar to the conceptual model shown in Figure 5. 6630 in a shale gas reservoir, Lincoln County, WV, from Kucuk et al.
It looks like p/z versus G exhibits four different (1978).
depletion patterns: 1) fracture-network dominated,
2) minor and micro fracture-network dominated, 3) matrix dominated, and 4) matrix with gas adsorption
dominated.
Based on the above analysis, it would be imprecise to use Arps (1945); Fetkovich (1980) decline type
curve analyses, and Agarwal et al. (1999); Blasingame et al. (1991); Carter (1985); Fetkovich (1980);
Fetkovich et al. (1987); Fraim and Wattenbarger (1987); Palacio and Blasingame (1993) decline type
curve and rate-transient analyses for naturally fractured reservoirs. Furthermore, in naturally fractured
reservoirs due to the unsymmetrical distribution of fractures and their directionality, it is very difficult to
use well-known circular or square bounded systems for both constant-pressure and no-flow boundary
conditions. The drainage area for a single well should be constructed based on a carefully built geological
model. Hydraulically fractured horizontal wells will further complicate the drainage area determination.
For instance, the stimulated reservoir volume (SRV) in shale reservoirs with multistage hydraulically
fractured horizontal wells does not make sense if the primary mechanism of production is described by
pressure diffusion in porous media that includes adsorption and Fickian molecular diffusion. Furthermore,
the stimulated reservoir volume (SRV) in shale reservoirs will be misleading if a shale gas formation
contains natural fractures.
In summary, as shown in Figs. 4, 5, and 6, naturally fractured reservoirs, including shale gas with or
without natural fractures, exhibit two to four types of decline periods. Therefore, a single-decline type
Arps curve should not be used, except in the boundary dominated final stage of the decline period.
In the following section, we will present a mathematical model for the rate-transient analysis in
naturally fractured reservoirs. Our model is also applicable to shale gas reservoirs until the beginning of
gas adsorption. For this paper, we will not include gas adsorption and/or Fickian molecular diffusion in
our model. On the other hand, our model is sufficiently flexible to include many geological complexities,
such as faults, fractures, constant or no-flow boundaries, etc.
Mathematical Model for Rate Transient Analysis in Fractured Reservoirs
As discussed above, the Barenblatt et al. (1960) and Warren and Root (1963) type models have been used
(Camacho-Velazquez et al., 2008; Da Prat et al., 1981; Ozkan et al., 1987; Raghavan and Ohaeri, 1981;
Sageev et al., 1985) for rate transient and decline curve analyses. In earlier papers (Biryukov and Kuchuk,
2012; Kuchuk and Biryukov, 2013, 2014), we showed that the Barenblatt et al. (1960) and Warren and
Root (1963) type models are inappropriate and incomplete for naturally fractured reservoirs. This is
because to obtain a reliable average for a REV, the scale of averaging should include a sufficiently large
number of blocks. It is therefore necessary to consider the fact that any infinitesimally small volume
includes not only a larger number of pores, as is assumed in the classical theory of seepage, but also a large
number of blocks. In other words, Warren and Root (1963) assumed that the infinitesimal control volume
contains a large proportion of each of the two constitutive media (the fracture network and matrix).
SPE-170698-MS 7

Because the definition of two continua is required– one for the porous blocks and one for the fracture
network–it follows that the application of this model assumes the existence of a Representative Elemen-
tary Volume (REV) that is common to both subsystems (Bear, 1993). If kf ⬎⬎ km, where kf and km are the
fracture and matrix permeabilities, respectively, a common REV for both subsystems cannot be defined
on a reasonable scale. Therefore, the Barenblatt et al. (1960) and Warren and Root (1963) dual-porosity
models can only be valid if kf and km have a similar order of magnitude. Consequently, they only are
applicable to reservoirs with very small matrix sizes and low fracture conductivities without definite
preferential fracturing directions.
Kuchuk and Biryukov (2014) examined many field well tests published in the literature. With a few
exceptions, none of them shows the behavior of the Warren and Root (1963) dual-porosity type models.
These examples exhibit very diverse pressure behaviors of discretely and continuously fractured reser-
voirs. Unlike the single derivative shape of the Warren and Root (1963) model, the derivatives of these
examples exhibit many different flow regimes [see Figs. 8 and 9 in Kuchuk and Biryukov (2013)],
depending on fracture distribution, and on their intensity and conductivity.
The linearity of the diffusivity equation allows the application of Duhamel’s superposition theorem for
both constant-pressure or constant-rate inner boundary condition, which can be expressed (van Everdin-
gen and Hurst, 1949) as
(1)

where pwfD is the dimensionless wellbore pressure for the constant-rate inner boundary condition, and qwfD
is the dimensionless wellbore flow rate for the constant-pressure inner boundary condition. Thus, if a
model is not applicable for computing pwfD, it will also not be applicable for computing qwfD. Conse-
quently, as in the pressure transient case, Barenblatt et al. (1960) and Warren and Root (1963) type models
are inappropriate and incomplete for rate transient analyses in naturally fractured reservoirs. It should be
understood that the Laplace transform variable s in dimensionless equations should correspond to tD.
Biryukov and Kuchuk (2012) presented transient pressure solutions for conductive and non-conductive
faults and fractures. The details of the solutions and derivations are given in the paper. The mesh-free
analytical-element technique that they presented treats the domain of interest as a whole, rather than
discretizing it into small grid blocks, as with finite-difference or finite-element methods. These mesh-free
solutions allow us to capture properties and particularities of the transient pressure behavior otherwise
inaccessible to mesh-based methods, as well as to obtain highly accurate solutions. They have used the
correct inner boundary conditions, which includes wellbore-intersecting fractures, direct wellbore con-
tribution if fracture conductivities are low, and when the well is only in the matrix without intersecting
any fractures, to capture early time behavior accurately for both continuously and discretely fractured
reservoirs. Their fractured reservoir model includes wellbore storage and skin effects in bounded and
unbounded systems. These fractured reservoirs may contain periodically or arbitrarily distributed finite-
and/or infinite-conductivity fractures with different lengths, intensities, and orientations.
From Duhamel’s superposition theorem, the rate-pressure convolution can be written [Eq. 3.88 in
Kuchuk et al. (2010)] as
(2)

where qm is the measured wellbore flow rate, qwf is the wellbore rate impulse response including the
pressure drop due to the damage skin and wellbore storage effects for the constant-pressure wellbore
boundary condition, ⌬pm ⫽ po – pm, pm is the measured pressure, and rw is the wellbore radius. On the
other hand, the measured pressure in the wellbore in terms of the measured flow rate can be written as
8 SPE-170698-MS

(3)

where gwf is the wellbore pressure impulse response including damage skin and wellbore storage effects,
and can be expressed [Eq. 3.52 in Kuchuk et al. (2010)] as
(4)

where L⫺1 denotes the inverse Laplace transform operator and s is the Laplace transform variable. The
Laplace transform of Eq. 3 can be written as
(5)

which gives
(6)

where gwf given by Eq. 4 and rw is dropped for convenience. Substituting Eq. 4 in Eq. 6 yields
(7)

where ⌬pw is the wellbore pressure response including the skin effect and the integral of gw. In the time
domain, Eq. 7 in the convolution form can be written as
(8)

where
(9)

The terms qm and ⌬pm are flow rate and pressure measurements in the wellbore, respectively. As can
be seen from Eqs. 2 and 8, the measured flow rate and pressure in the wellbore can be expressed in a few
different forms in terms of the pressure impulse response due to the constant-rate wellbore boundary
condition, or the rate impulse response due to the constant-pressure boundary condition.
The dimensionless wellbore flow rate from Eq. 8 can be written as
(10)

where
(11)

In this paper, we use Eq. 10 to obtain the wellbore flow rate with wellbore skin and storage for naturally
fractured reservoirs.
One of the main deficiencies of decline curve and rate transient analyses has been the assumption of
a constant pressure (⌬pm in Eq. 8) at the sandface. When ⌬pm varies as a function of time even for
homogenous reservoirs, production data can match different Arps (1945) decline curves, which has
nothing to do with the constant-pressure Arps (1945) decline curves. For tight and shale gas reservoirs,
it is almost impossible to sustain the wellbore pressure at a constant value for a long time, and the system
SPE-170698-MS 9

often does not reach a pseudosteady-state condition. Although Agarwal et al. (1999); Blasingame et al.
(1991); Fetkovich et al. (1987); Fraim and Wattenbarger (1987); Palacio and Blasingame (1993) have
considered a variable ⌬pm, these type curve rate-transient analysis techniques are limited applications
because they can only entertain simple geometrical reservoir models. These type curve techniques are
insufficient for estimating more than a few parameters, and do not incorporate statistics of estimated
parameters that include confidence intervals and correlation coefficients, etc. Some techniques do not use
primary measurements, and use superficial variables such as tir ⫽ Qm(t)/qm(t). Using these superficial
variables does not bring any substantial value to rate-transient analysis because and the
transformation also causes statistical distortion of primary measurements. If we assume qm(t) ⫽ ␣t, then
Qm(t) ⫽ ␣t2/2, which gives tir ⫽ t/2, but with a numerical integration error. If we assume qm(t) ⫽ qi
exp(⫺␣t), then Qm(t) ⫽ (qi/␣)[1 ⫺ exp(⫺␣t)] ⫽ (qi/␣) ⫺ qm(t)/␣, which also can be written as qm(t) ⫽
1 ⫺ (␣/qi)Qm(t). Thus, the cumulative production, Qm(t), will be a linear function of production rate, qm(t),
or qm(t) will be a linear function of Qm(t) for the linear pressure diffusion in any bounded system when
the system reaches the pseudo-steadystate condition. This is due to a property of the exponential function,
which is 兰 exp(x) dx ⫽ exp(x). This fact has been used to justify or re-derive the Arps (1945) decline
curves, [see Eq. 12.182 in Page 517 of Raghavan (1993)].
It should be stated that Palacio and Blasingame (1993) and Agarwal et al. (1999) used the Gladfelter
et al. (1955) reciprocal productivity index ]normalized pressure ⌬pm(t)/qm(t)] idea to obtain normalized
rate qm(t)/⌬pm(t) (normalized rate integral is also used), which is roughly equal to the constant-rate
solution qwf given in Eq. 2 for very simple flow geometries, such as radial and linear ]see Kuchuk (1990)].
Kuchuk (1990) referred to the Gladfelter et al. (1955) reciprocal productivity index technique as the
Gladfelter deconvolution. Consequently, using normalized rate to obtain the constant-rate solution is
nothing more than the Gladfelter rate-pressure deconvolution. Kuchuk (1990) summarized the validity of
the Gladfelter deconvolution as
1. The Gladfelter deconvolution is a good approximation of qwf or ⌬pwf if the flow rate varies linearly
with time for simple flow geometries.
2. It does not work if the flow rate varies arbitrarily.
As stated above, because of the linearity of the diffusivity equation (see Eq. 1), if a model does not
work for the constant-rate case, it will also not work for the constant-pressure case.
If we assume an exponential decline for a reservoir, then the dimensionless flow rate during the
pseudo-steadystate flow period [see Eq. 5.50 in Page 115 of Raghavan (1993)] can be written as
(12)

where
(13)

where CA is the shape factor, , and A is the drainage area of the reservoir, and ␥ ⫽
0.577216. . . is the Euler constant. The Laplace transform of qwfD is given as [see Eq. 5.48 in Page 115
of Raghavan (1993)]
(14)

and the Laplace transform of Eq. 2 in dimensionless variables can be written as


(15)

Substituting Eq. 14 in Eq. 15 yields


10 SPE-170698-MS

(16)

For large tD as s ¡ 0, Eq. 16 becomes


(17)

Since , Eq. 17 in the time domain becomes


(18)

Equation 18 confirms the linear relationship between pmD and QmD. A plot of pmD as a function of QmD
yields the initial pressure, cumulative production, and the drainage area. It should be noted that because
is large, i.e. 2␲/( ) is small, thus it will take a long time to have s ⬍⬍ 2␲/( ) in Eq. 16.
As suggested by Fetkovich (1980), in terms of dimensionless rate and time, Arps’ decline equation can
be written as
(19)

where qD ⫽ q(t)/qi and tdD ⫽ Dit. Now let us examine the ultimate predicted dimensionless cumulative
production (reserve) from Arps’ decline curves by integrating by Eq. 19 with respect to tdD as
(20)

As expected for the exponential decline (b ⫽ 0), the ultimate predicted dimensionless cumulative
production (QD) becomes 1. For harmonic decline (b ⫽ 1), as b ¡ 1 then QD ¡ ⬁. This also can be shown
for b ⫽ 1 as
(21)

Of course, an infinite cumulative production is an impossiblity in any bounded system. For instance,
for solution-gas drive reservoirs, b ⫽ 2/3 (Raghavan, 1993), then QD ⫽ 3, and QD ⫽ 10 for b ⫽ 0.9. For
the same volume, these are impossible cumulative production values. Nonlinearities in rock (compaction,
etc), two-phase or three-phase flow, fluid properties, and formation heterogeneity may distort the behavior
of exponential decline. As a result, the production behavior may not follow the exponential Arps’ decline
curve for a while, which is not empirical but is a property of pressure diffusion in a bounded system.
However, none of these effects will make the bounded volume based cumulative production (QD ⫽ 1)
doubled (b ⫽ 0.5), tripled (b ⫽ 2/3), or decupled (b ⫽ 0.9). We have ignored the abandonment pressure
in Eq. 20, but that will not change the unphysical behavior (yields very high production rate and
cumulative production) of Arps’ decline curves. Therefore, the validity of Arps’ decline curves should be
seriously questioned for all types of reservoirs.
Unneland et al. (1998) used Eq. 2 with measured ⌬pm(t) and qm(t), and the Arps (1945) exponential and
hyperbolic decline curves for two examples, respectively, to obtain the parameters of the decline curves.
Although Unneland et al. (1998) used wellbore flow rate and pressure measurements directly, but using
the hyperbolic decline curves will not yield a reliable production rate and cumulative production, except
the exponential decline as mentioned above.
Kuchuk et al. (2005) used both rate-pressure deconvolution and nonlinear regression techniques to
identify the system and to estimate its parameters with the confidence intervals from measured wellbore
flow rate and pressure data, without using unnecessary transformations of the direct measurements. It
should be stated that for gas reservoirs, the pseudo-pressure and sometimes pseudo-time transformations
SPE-170698-MS 11

Figure 7—Measured and computed wellbore pressures and flow rates Figure 8 —The Arps (1945) decline type curve analysis for Field Exam-
for Field Example 2, from Kuchuk et al. (2005). ple 2 given by Kuchuk et al. (2005).

must be used if we want to take advantage of


analytical solutions for nonlinear regression. It is
also worth noting that because surface flow rate
measurements may be noisy and not very accurate,
and are acquired infrequently, rate-pressure decon-
volution and taking the derivative of rate data may
not always be possible. On the other hand, although
permanent downhole monitoring systems provide
accurate pressure data, measuring downhole rate is
not very common for most wells. Thus Eq. 2 with
measured ⌬pm(t) and qm(t) and the best available
geological model of the reservoir should be used
with nonlinear regression to obtain reservoir param-
eters. For this purpose, the maximum likelihood
estimation method should be used instead of the Figure 9 —Nonlinear regression analysis of the deconvolved unit-
pressure rate with an Arps’ decline curve for Field Example 2, from
weighted least square, commonly used in reservoir Kuchuk et al. (2005).
engineering applications, because it is very difficult
to assign weights to both surface and wellbore flow rate measurements. Also, obtaining the confidence
intervals is indispensable for any rate transient interpretation, including the pseudo-steadystate flow period
because of the high uncertainty in flow rate measurements.
To demonstrate the Kuchuk et al. (2005) rate-transient analysis technique, we present Field Example
2 given by Kuchuk et al. (2005). In this example, both downhole pressure and flow rate were acquired
simultaneously by a permanent downhole system, as shown in Fig. 7. The data quality of both pressure
and rate is very good. Figure 8 presents a semilog plot of the Arps (1945) decline curve analysis of the
downhole flow rate data shown in Fig. 7. Arps’ decline analysis yields: the exponent b ⫽ 9.2, Di ⫽
0.0000295 hr⫺1, and qi ⫽ 24570 B/D. Although the data matches Arps’ decline curve perfectly, the value
of b is unacceptably high because the downhole pressure variation is not taken into account. Kuchuk et
al. (2005) applied the rate-pressure deconvolution algorithm, which is very similar to what is given by von
Schroeter et al. (2004) for this downhole pressure and flow rate data. Using nonlinear regression analysis,
the deconvolved (constant-pressure) rate data is matched to the exponential decline curve, as shown in Fig.
9. The match is fairly good, except that there is a slight mismatch towards the end of the data, and yields:
b ⫽ 0.155, Di ⫽ 0.00237 hr⫺1, and qi ⫽ 1355 B/D/psi. The other details, confidence intervals, etc., are
12 SPE-170698-MS

Figure 10 —Comparison of the dimensionless flow rates (qD) obtained from Kuchuk and Brigham (1979) denoted as Kuchuk 1979, Kuchuk (1990)
denoted as Kuchuk 1990, and this study.

given in Kuchuk et al. (2005). This example dem-


onstrates the importance of including pressure vari-
ations in the exponential decline curve, which has
an analytical foundation.
Field Example 2 is an exceptional case because
normally flow rate (production) and pressure are
infrequently measured at the surface, and the data
are noisy and not very accurate. It is therefore very
difficult to deconvolve surface-measured flow rate
and pressure data for a long term decline and rate-
transient analysis. We have presented this example
simply to demonstrate the drawbacks of the Arps
Figure 11—Warren and Root (1963) uniformly distributed fracture and
(1945) decline curve analysis, with the assumption matrix block geological model, after Kuchuk and Biryukov (2014).
that the average reservoir pressure and the bottom-
hole flowing pressure are constant during the entire
production period. As mentioned above, it is almost
impossible to sustain the wellbore pressure at a
constant value for a long time, and tight reservoirs
often do not reach a pseudosteady-state condition.
Except for pressure transient tests with long produc-
tion periods, there is no technique that indicates
when the system reaches a pseudosteady-state con-
dition.
Because of these drawbacks with the Arps (1945)
decline curve analysis, measured qm(t) with ⌬pm(t)
and the best available geological model of the res-
ervoir should be used for nonlinear regression to
obtain reservoir parameters. The results of the non- Figure 12—Dimensionless rate of a fracture-intersecting well in the
linear regression Field Example 2 is given in Ku- Warren and Root (1963) geological model (shown in Fig. 11) for various
fracture conductivities.
SPE-170698-MS 13

Figure 13—Pressure derivatives of a fracture-intersecting well in the Figure 14 —Rate derivatives of a fracture-intersecting well in the War-
Warren and Root (1963) geological model (shown in Fig. 11), after ren and Root (1963) geological model shown in Fig. 11).
Kuchuk and Biryukov (2014).

Figure 15—Distributions of fractures in a discretely fractured reservoir, Figure 16 —Dimensionless rates of the discretely fractured reservoir
where the well is located at (0,0), from Kuchuk and Biryukov (2012). (Fig. 15), where the well does not intersect any fracture, for various
fracture conductivities.

chuk et al. (2005). The geological model for Field Example 2: the well was located 500 ft from two sealing
faults intersecting at a 90° angle with good aquifer support. For naturally fractured reservoirs, however,
geological model building is a highly involved task.
The integrated pressure transient interpretation methodology for naturally fractured reservoirs given by
Morton et al. (2012) can also be applied to rate transient analysis and production forecasting. Of course,
there will be some modifications, particularly if high-precision downhole pressure data and its diagnostic
derivatives are not available. However, most of the workflow for the geological model building will be
the same. As it is well known, naturally fractured reservoirs comprise a discrete or connected network of
fractures and a matrix (host formation) with different porosities and permeabilities. For naturally fractured
reservoirs, the following parameters are important for model building:
1. Fracture connectivities.
2. Fracture density.
3. Fracture orientations.
4. Fracture lengths.
5. Fracture spacings.
6. Faults and fractures.
7. Fault types and properties.
In addition, the formation matrix properties should be included in the reservoir model. The main steps
of the rate transient analysis:
1. Model identification—a model is found that behaves in the same way as the actual system
14 SPE-170698-MS

Figure 17—Dimensionless pressure derivatives of the discretely frac- Figure 18 —Dimensionless rate derivatives for the discretely fractured
tured reservoir (Fig. 15), where the well does not intersect any fracture, reservoir (Fig. 15), where the well does not intersect any fracture, for
for various fracture conductivities. various fracture conductivities.

(considered as the parameterization of the physical system).


2. Model parameter estimation—the model parameters are adjusted until the model output matches
satisfactorily the measured data.
3. Model verification—the consistency of the interpretation model is verified by measuring the
mismatch between the data and the model, or by comparison with other data.
As stated above, it is difficult to perform the model identification step with surface flow rate and
pressure measurements. This should be done with pressure transient well tests if they are available. It
should be emphasized that the above methodology can be applied to any type of reservoir, but
characterizing naturally fractured reservoirs is the most challenging.

Examples for Rate Transient Behavior of Naturally Fractured Reservoirs


The majority of vertical wells are perforated to communicate with natural fractures as well as to create
hydraulic fractures. Most natural fractures do not penetrate the entire pay zone. Wellbore storage, skin due
to scaling and plugging of perforations, damage on the surface of fractures, limited entry due to
perforations, and partially penetrated fractures complicate and often distort rate behavior of fractured
reservoirs for a considerable period of time. Furthermore, drilling and completion operations and fluids,
as well as mud plugging, may damage wellbore-intersecting natural fractures. In some cases, the fracture
damage zone may extend beyond the wellbore-intersecting fractures because of very heavy drilling mud
losses into all of the connected fractures. As a result, a severely damaged fracture zone (choked fracture)
may form. The behaviors of damaged infinite-conductivity and/or low-conductivity fractures are initially
dominated by wellbore storage effects.
In addition, damaged fractures with dip angles that are less than 90° amplify the wellbore storage
effects. If the wellbore-intersecting fractures are highly conductive without a substantial amount of skin,
then wellbore storage effects for the rate transient analysis are negligible for naturally fractured reservoirs
with a reasonable good matrix permeability. However, for tight reservoirs, where the matrix permeability
could vary from nanodarcy to microdarcy, wellbore storage effects should not be ignored.
All these fracture-damaging factors significantly affect the productivity of naturally fractured reser-
voirs with or without hydraulic fractures. Our naturally fractured reservoir solutions can be easily
extended to investigate the effect of these fracture-damaging factors on rate transient and well produc-
tivity. The effects of these factors on pressure transient behavior of the naturally fractured reservoirs with
and without hydraulic fractures were investigated by Kuchuk and Biryukov (2013). In this paper,
however, we will not investigate these effects on the rate transient behavior of naturally fractured
reservoirs due to the page limit restriction. It will be covered in the future. In this paper, we will
investigate the rate transient behavior of a few different types of naturally fractured reservoirs.
SPE-170698-MS 15

Figure 19 —Dimensionless rates for a bounded version of the fractured Figure 20 —Dimensionless rate derivatives for a bounded version of the
reservoir shown in Fig. 15 with an exponential decline curve matching. fractured reservoir shown in Fig. 15.

All synthetic rate transient examples for fractured reservoirs are generated with the fractured reservoir
model presented by Biryukov and Kuchuk (2012). In this paper, the following dimensionless variables are
used and defined (in the oilfield units). The dimensionless wellbore pressure and flow rate for the
constant-pressure inner boundary condition are given as
(22)

pwf is the constant pressure at the inner wellbore boundary. The dimensionless wellbore pressure and
other dimensionless variables for constant-rate inner boundary condition are given as
(23)

where kf, b, and lw denote the fracture permeability, aperture, and the half length or reference length,
respectively; km, ␾m, and (ct)m denote the matrix permeability, porosity, and total compressibility,
respectively; C denotes the wellbore storage coefficient; h denotes the reservoir thickness; po and pwf
denote the initial reservoir and flowing wellbore pressures, respectively; q denotes the reference flow rate;
␮ denotes fluid viscosity; and t denotes time.
In the next sections, we will present the dimensionless flow rate (qmD ⫽ qD) and the value of its
derivative for the constant-pressure inner boundary condition as a function of tD.
For most cases, the rate derivative , corresponds to the pressure derivative . As stated
above, taking the derivative of surface or wellbore flow rate measurements is difficult; nevertheless, it still
makes sense to present them along with the dimensionless flow rate data.
Wellbore-Intersecting Single Fracture
It is important to understand the behavior of a single-wellbore-intersecting fracture (hydraulic or natural)
to interpret the pressure and rate transient behavior of fractured reservoirs. As will be shown later, when
a well intersects a fracture(s) it dominates the behavior of the system for a considerable time period,
particularly if the length of the fracture is long and the formation is tight, las in shale gas and oil reservoirs.
In this paper we often look at the individual behavior of certain features of the model. In reality, the effects
of fractured reservoir features may interfere with each other. The complexity of the behavior increases
when we add the effects of skin and wellbore storage. The pressure transient behavior of a single
wellbore-intersecting fracture has been well studied, and many flow regimes have been discussed in the
literature (Cinco-Ley and Samaniego-V, 1977, 1981a,b; Evans, 1971; Kuchuk and Biryukov, 2013, 2014;
Lee and Brockenbrough, 1986; Raghavan, 1976; Soliman, 2009). The rate transient behavior of a single
wellbore-intersecting fracture has also been studied (Guppy et al., 1981, 1988; Nashawi and Malallah,
2007; Prats et al., 1962). Only Nashawi and Malallah (2007) have investigated the rate transient derivative
behavior of a single fracture using published solutions for flow regimes.
16 SPE-170698-MS

Figure 21—Dimensionless pressure derivatives for the discretely frac- Figure 22—Dimensionless rate derivatives for the discretely fractured
tured reservoir (Fig. 15) with a FD ⴝ 1000 hydraulic fracture in the NW reservoir (Fig. 15) with a FD ⴝ 1000 hydraulic fracture in the NW to SE
to SE direction. direction.

Figure 23—Dimensionless rate for the discretely fractured reservoir Figure 24 —Dimensionless rates for the discretely fractured reservoir,
(Fig. 15) with a FD ⴝ 1000 hydraulic fracture in the NW to SE direction. shown in Fig. 2.

One of the classic papers on elliptical flow in porous media by Prats et al. (1962) presented the first
solutions for transient behavior of an infinite-conductivity vertically fractured well produced at a constant
pressure and constant rate in a closed elliptical reservoir. The second paper on an infinite-conductivity
vertically fractured well produced at a constant pressure in a bounded reservoir was published by Locke
and Sawyer (1975). Since then more papers were published on the subject.
The dimensionless pressure for an infinite-conductivity vertically fractured well is given by Gringarten
et al. (1974)
(24)

and the Laplace transforms of Eq. 24 can be written (Kuchuk, 1990) as


(25)

where S(z) is expressed as


(26)

K0 and K1 are the modified Bessel functions, L0 and L1 are the modified Struve functions [see
(Abramowitz and Stegun, 1972)], , , ␯1 ⫽ 0.1299514, and ␯2 ⫽ 0.8700486. It
should be noted that ␯1 and ␯2 are slightly modified (Kuchuk, 1990) compared to the values given by
Gringarten et al. (1974).
SPE-170698-MS 17

Figure 10 presents the comparison of our dimen-


sionless flow rate with that from Table 2 for K ⫽ ⬁
in Kuchuk and Brigham (1979) and Kuchuk (1990)
for an infinite-conductivity vertically fractured well
in a homogeneous infinite system. Our solution
compares very well with the Kuchuk and Brigham
(1979) solution, for which qD was not obtained from
the dimensionless pressure solution, pD, but they
solved the pressure diffusion directly for a constant-
pressure inner boundary condition. On the other
hand, qD for Kuchuk (1990) is obtained from pD Figure 25—Dimensionless rate derivatives for the discretely fractured
given by Eq. 25 with Eq. 1. For this study, qD is reservoir, shown in Fig. 2.
obtained from pD from Eq. 1 using the solutions
given by Biryukov and Kuchuk (2012). All three qD values compare with each other exceptionally well.
This agreement also verifies that qD obtained from pD via Eq. 1 is very accurate.
Often qD is approximated by the inverse of pD. This is not true universally. For instance, for the linear
flow, using Eq. 1 and yields
(27)

The rate derivative from Eq. 27 can be written as


(28)

and the pressure derivative as


(29)

As in the pressure and rate case, the pressure and rate derivatives are also different, but both will
exhibit a 1/2-slope (linear flow regime) on a log-log plot.
For the bilinear flow, using Eq. 1 and [see Eq. 11.19
in page 297 of Soliman (2009) for pD] yields
(30)

where FD is the dimensionless fracture conductivity and ⌫ is the gamma function. The rate and pressure
derivatives can be written as
(31)

As can be seen from Eq. 31 the pressure and rate and their derivatives are different, but both will
exhibit a 1/4-slope (bilinear flow regime) on a log-log plot.

Rate Transient Behavior of Continuously Fractured Reservoirs


Now let us look at the rate-transient behavior of a continuously fractured system, where fractures are in
a network and communicate with each other throughout the system, as shown in Fig. 11. This example
is given by Kuchuk and Biryukov (2014) and a 2D representation of the Warren and Root (1963)
18 SPE-170698-MS

dual-porosity geological model, where the matrix blocks consist of 65.6 ft by 65.6 ft by h (the reservoir
thickness) (20 m by 20 m by h) rectangular parallelepipeds with a square base, and fractures between them
with an aperture b, a length 65.6 ft (2lw), and a height h. As shown in Fig. 11, the square matrix blocks
(in 2D) with a height h (red bricks in the figure) are uniformly distributed in the formation, and the
fractures are orthogonal and uniform with the aperture (white space between bricks). In this model the
fracture network is finite in the reservoir, and the well intersects one of the fractures at the center of the
fracture network. Fig. 12 presents dimensionless flow rate as a function of dimensionless time for FD ⫽
0.001, 0.01, 0.1, 1, 10,100 and 1,000. As can be observed in this plot, the flow rate for FD ⫽ 1,000 starts
declining much faster after tD of 10 because the fluid in the fractures is quickly depleted before being
replenished by the matrix elements.
Figures 13 and 14 present the pressure and rate derivatives for the model shown in Fig. 11. As can be
seen from these figures, the pressure derivatives, which have well defined flow regimes, such as
pseudosteady-state (m ⫽ 1), fracture-radial (m ⫽ 0), and formation linear (m ⫽ 1/2) flow regimes, do not
correspond to the rate derivatives. Only the rate derivatives for FD ⫽ 10 and 100 exhibit similar behavior
(pseudosteady-state and formation linear) to the corresponding pressure derivatives. As can be seen from
Fig. 14, some of the rate derivatives intersect each other. In other words, the pressure and rate derivatives
exhibit very dissimilar behaviors. We will explain this behavior in Appendix B with another example,.
Very low conductivity (FD ⫽ 0.001, and 0.01) fractures do not affect both pressure and rate derivatives,
as can be seen in Figs. 13 and 14.

Rate Transient Behavior of of Discretely Fractured Reservoirs


We simulated the rate transient behavior of a reservoir configuration containing 40 fractures (Fig. 15),
which are vertical and disjointed, predominately in the North West to South East (NW to SE) direction
in the homogeneous porous medium (matrix). Figure 16 presents the dimensionless rate for this reservoir.
The fracture conductivity affects the rate behavior, but not very significantly. Figures 17 and 18 present
the pressure and rate derivatives for the model shown in Fig. 15. As can be seen from these figures, the
pressure derivatives correspond very well to the rate derivatives. Neither derivative exhibits a recogniz-
able flow regime, but both show a three-zone radially composite reservoir look-alike behavior.
The dimensionless flow rates are also obtained for a bounded version of the fractured reservoir shown
in Fig. 15, and then are we matched the exponential decline curve, as shown in Fig. 19. An exponential
decline curve matches these rates very well but deviates slightly towards the end. The rate derivatives, as
shown in Fig. 20, show the deviation from the boundary-dominated pseudo-steadystate flow regime (m ⫽
1). For continuously fractured reservoirs, if the fractures have the same conductivity, the matrix blocks are
the same size, and both are uniformly distributed throughout the bounded system. First the connected
fracture network and then the matrix blocks will reach a pseudo-steadystate flow condition. On the other
hand, for continuously fractured reservoirs, if the fractures have different conductivities and the matrix
blocks have different sizes and/or different permeabilities, there will not be a single pseudo-steadystate
condition for either the fractured system or the matrix system (see Fig. 5). For instance, some large matrix
blocks are still in a transient condition, while smaller matrix blocks could be in a pseudo-steadystate
condition. This is also true for discretely fractured reservoirs with nonuniform distributions of fracture and
matrix densities and properties (sizes, permeabilities, conductivities, etc.). This is what we observe in Fig.
20.
Now let us assume that the well in the fractured reservoir shown in Fig. 15 is hydraulically fractured
in the NW to SE direction, where FD ⫽ 1000 and lwD ⫽ 5 (the dimensionless hydraulic fracture length).
For this model, Figs. 21 and 22 present the pressure and rate derivatives. As can be seen from these
figures, the pressure derivatives correspond very well to the rate derivatives with recognizable flow
regimes.
SPE-170698-MS 19

Figure 23 shows the dimensionless rates for this hydraulically fractured reservoir. The same figure also
shows the dimensionless rate for FD ⫽ 1000 without the hydraulic fracture. Although the fracture
conductivity somewhat affects the rate behavior, the hydraulic fracture very significantly increases the
flow rate.
We simulated the rate transient behavior of a fractured reservoir, shown previously in Fig. 2, and its
pressure derivative in Fig. 3. Figures 24 and 25 present dimensionless rate and its derivative. As can be
seen from these figures, the fracture conductivity affects the rate and derivative behaviors considerably.
As can also be observed from Figs. 3 and 24, the pressure derivatives correspond somewhat to the rate
derivatives, but there are differences.

Conclusions
In this paper we presented a literature review for rate transient and decline curves analyses generally and
specifically for naturally fractured reservoirs. We presented solutions for the rate transient behavior of a
vertical well in naturally fractured reservoirs containing periodically or arbitrarily distributed finite-
and/or infinite-conductivity fractures, as well as hydraulic fractures. These solutions are applicable to
continuously and discretely fractured reservoirs. As in the pressure transient case, it is shown that the
Warren and Root (1963) dual-porosity type models do not capture the transient rate behavior of most
fractured reservoirs because they do not contain fractures. The Warren and Root (1963) model is
inappropriate and incomplete for the interpretation of rate transient behavior of fractured reservoirs.
It is shown that the inverse of pD is not a good approximation for qD. Although they are different, for
most cases, the rate derivatives tn correspond to the pressure derivatives correspond to the
pressure derivative . In other cases, they behave differently.
It is shown that for any reservoir, Arps’ decline curves yield unreasonably high production rates and
cumulative productions, except the exponential decline case. The exponential decline curve should not be
used without taking into account the change of wellbore pressure as a function of time. Arps’ decline
curves analysis should not be used for both discretely and continuously fractured reservoirs. An integrated
interpretation methodology is outlined for rate transient analysis in fractured reservoirs.
We presented a few examples of the rate transient behavior of discretely and continuously fractured
reservoirs. These examples exhibit many different flow regimes depending on fracture distribution,
intensity, and conductivity. We also compared the rate transient derivatives with the pressure transient
derivatives. For most cases, they exhibit similar behaviors. The derivatives and flow regimes presented are
valuable diagnostic tools for rate transient analysis.

Acknowledgments
The authors are grateful to Schlumberger for permission to publish this paper, and would like to thank
Kirsty Morton for valuable discussions.

Nomenclature
A ⫽ well drainage area
b ⫽ decline curve exponent
C ⫽ shape factor or wellbore storage constant
ct ⫽ compressibility
Di ⫽ initial decline rate coefficient
F ⫽ fracture conductivity
g ⫽ reservoir impulse response
G ⫽ cumulative fluid production
h ⫽ formation thickness
20 SPE-170698-MS

k ⫽ permeability
l ⫽ fracture half-length
p ⫽ pressure
q ⫽ flow rate
Q ⫽ cumulative production
r ⫽ radius or radial coordinate
s ⫽ Laplace transform variable
S ⫽ skin factor
t ⫽ time
x ⫽ coordinate
y ⫽ coordinate
z ⫽ vertical coordinate or compressibility factor
␩ ⫽ diffusivity
␮ ⫽ viscosity
␾ ⫽ porosity
␬ ⫽ defined by Eq. 9

Subscripts
D ⫽ dimensionless
f ⫽fracture
o ⫽ initial or original
m ⫽ matrix or measured
w ⫽wellbore
wf ⫽ wellbore flowing

References
Abdassah, D. and Ershaghi, I. 1986. Triple-porosity systems for representing naturally fractured
reservoirs. SPE Formation Evaluation, 1(2):113–127.
Abramowitz, M. and Stegun, I. A. 1972. Handbook of Mathematical Functions. Dover, New York.
Agarwal, R., Gardner, D., Kleinsteiber, S., and Fussell, D. 1999. Analyzing well production data using
combined-type– curve and decline– curve analysis concepts. SPE Reservoir Engineering, 2(5):478 –486.
Ahmed, T. and McKinney, P. 2011. Advanced Reservoir Engineering. Gulf Publishing Co, Houston,
TX, 1st edition.
Al-Ahmadi, A. H., and Wattenbarger, R. A. 2011. Triple-porosity models: One further step towards
capturing fractured reservoirs heterogeneity. Paper SPE-149054-MS, the SPE/DGS Saudi Arabia Section
Technical Symposium and Exhibition, Al-Khobar, Saudi Arabia.
Arps, J. 1945. Analysis of decline curves. Transactions of the AIME, 160(1):228 –247.
Barenblatt, G. I., Zeltov, Y. P., and Kochina, I. 1960. Basic concepts in the theory of seepage of
homogeneous liquids in fissured rocks. PMM (Journal of Soviet Applied Mathematics and Mechanics),
24(5):1286 –1303.
Bear, J. 1993. Flow and Contaminant Transport in Fractured Rock, chapter Modeling Flow and
Contaminant Transport in Fractured Rocks, pages 1–37. Bear, J. and Tsang, C.F. and de Marsily, G,
(editors). Academic Press, Inc, San Diego, CA., US, 1st edition.
Bertuzzi, A. F., Fetkovich, M. J., Poettmann, F. H., and Thomas, L. K. 1987. Petroleum Engineering
Handbook, chapter Wellbore Hydraulics. Edited by H. B. Bradley. Society of Petroleum Engineers,
Richardson, Texas, 1st edition.
SPE-170698-MS 21

Biryukov, D. and Kuchuk, F. 2012. Transient pressure behavior of reservoirs with discrete conductive
faults and fractures. Transport in Porous Media, 95:239 –268. 10.1007/s11242-012-0041-x.
Blasingame, T., McCray, T., and Lee, W. 1991. Decline curve analysis for variable pressure
drop/variable flowrate systems. Paper SPE 21513-MS, SPE Gas Technology Symposium, 22–24 January,
Houston, Texas.
Camacho-Velazquez, R., Fuentes-Cruz, G., and Vasquez-Cruz, M. 2008. Decline-curve analysis of
fractured reservoirs with fractal geometry. SPE Reservoir Evaluation & Engineering, 11(3):606 –19.
Carter, R. 1985. Type curves for finite radial and linear gas-flow systems: Constant-terminal-pressure
case. SPE Journal, 25(5):719 –728.
Cinco-Ley, H. and Samaniego-V, F. 1977. Effect of wellbore storage and damage on the transient
pressure behavior of vertically fractured wells. Paper SPE 6752, SPE Annual Technical Conference and
Exhibition, Denver, Colorado, 9 –12 October.
Cinco-Ley, H. and Samaniego-V, F. 1981a. Transient pressure analysis: Finite conductivity fracture
case versus damaged fracture case. Paper SPE 10179, SPE Annual Technical Conference and Exhibition,
San Antonio, Texas, 4 –7 October.
Cinco-Ley, H. and Samaniego-V, F. 1981b. Transient pressure analysis for fractured wells. J. of
Petroleum Technology, 33(9):1749 –1766.
Craft, B. C., and Hawkins, M. 1959. Applied Petroleum Reservoir Engineering. Prentice-Hall,
Englewood Cliffs, New Jersey.
Cutter, W. W. 1924. Estimation of undergound oil reserves. Technical Report 228, U. S. Bureau of
Mines Bulletin.
Da Prat, G., Cinco-Ley, H., and Ramey, Jr. H., 1981. Decline curve analysis using type curves for
two-porosity systems. SPE Journal, 21(3):354 –362.
Evans, J. G. 1971. The use of pressure buildup information to analyze non-respondent vertically
fractured oil wells. Paper SPE 3345, SPE Rocky Mountain Regional Meeting, Billings, Montana, 2– 4
June.
Ezulike, D. O., and Dehghanpour, H. 2013. Characterizing tight oil reservoirs using dual- and
triple-porosity models. Paper SPE-167126-MS, SSPE Unconventional Resources Conference Canada, 5–7
November, Calgary, Alberta, Canada.
Fetkovich, M. 1980. Decline curve analysis using type curves. J. of Petroleum Technology, 32(6):
1065–1077.
Fetkovich, M., Vienot, M. E., Bradley, M. D., and Kiesow, U. G. 1987. Decline curve analysis using
type curves: Case histories. SPE Formation Evaluation, 2(4):637–656.
Fraim, M. L., and Wattenbarger, R. A. 1987. Gas reservoir decline-curve analysis using type curves
with real gas pseudopressure and normalized time. SPE Formation Evaluation, pages 671–682.
Gladfelter, R., Tracy, G., and Wilsey, L. 1955. Selecting wells which will respond to production-
stimulation treatment. Drill. and Prod. Prac., API, pages 117–129.
Gringarten, A. C., Ramey, H. J., and Raghavan, R. 1974. Unsteady–state pressure distributions created
by a well with a single infinite— conductivity vertical fracture. SPE J., 14(4).
Guppy, K. H., Cinco, H., and Ramey, H. 1981. Transient flow behavior of a vertically fractured well
producing at constant pressure. Paper SPE-9963-MS, unsolicited.
Guppy, K. H., Kumar, S., and Kagawan, V. D. 1988. Pressure transient analysis for fractured wells
producing at constant pressure. SPE Formation Evaluation, 3(1):169 –178.
Houze, O. P., Kikani, J., and Horne, R. N. 2009. Transient Well Testing, chapter 17-Permanent
Gauges and Production Analysis, pages 133–178. Monograph Series 23, Edited by M. Kamal. Society of
Petroleum Engineers, Richardson, 1st edition.
Huntley, L. 1913. Possible causes of the decline of oil wells. Bureau of Mines, 52:32–.
22 SPE-170698-MS

King, G. 1993. Material-balance techniques for coal-seam and devonian shale gas reservoirs with
limited water influx. SPE Reservoir Engineering, 8(1):67–72.
Kuchuk, F. 1990. Gladfelter deconvolution. SPE Formation Evaluation, 5(3).
Kuchuk, F. and Biryukov, D. 2012. Transient pressure test interpretation from continuously and
discretely fractured reservoirs. Paper SPE 158096, SPE Annual Technical Conference and Exhibition, San
Antonio, Texas, 8 –10 October.
Kuchuk, F. and Biryukov, D. 2013. Pressure transient tests and flow regimes in fractured reservoirs.
Paper SPE 166296-MS, SPE Annual Technical Conference and Exhibition, New Orleans, Louisiana, 30
September- 2 October 2013.
Kuchuk, F. and Biryukov, D. 2014. Transient pressure test interpretation from continuously and
discretely fractured reservoirs. SPE Reservoir Evaluation & Engineering, 17(01):82–97.
Kuchuk, F., Hollaender, F., Gok, I., and Onur, M. 2005. Decline curves from deconvolution of
pressure and flow-rate measurements for production optimization and prediction. Paper SPE 96002, SPE
Annual Technical Conference and Exhibition, Richardson, Texas, 9 –12 October.
Kuchuk, F., Onur, M., and Hollaender, F. 2010. Pressure Transient Formation and Well Testing:
Convolution, Deconvolution and Nonlinear Estimation. Elsevier, New York.
Kuchuk, F. J., and Brigham, W. E. 1979. Transient flow in elliptical systems. SPE J., 19(6):401–410.
Kucuk, F., Alam, J., and Streib, D. L. 1978. Proceedings, reservoir engineering aspects and resource
assessment methodology of eastern gas shales. Second Eastern Gas Shales Symposium, Morgantown,
West Virginia, Oct. 16 –18.
Lee, S.-T., and Brockenbrough, J. R. 1986. A new approximate analytic solution for finite-conduc-
tivity vertical fractures. SPE Formation Evaluation, 1(1):75–88.
Liu, C. Q. 1981. Exact solution for the compressible flow equations through a medium with
triple-porosity. Applied Mathematics and Mechanics, 2(4):457–462.
Locke, G. H., and Sawyer, W. K. 1975. Constant pressure injection test in a fractured reservoir–
history match using numerical simulation and type curve analysis. Paper SPE 5594-MS, SPE Annual
Technical Conference and Exhibition, 28 September-1 October, Dallas, Texas.
Moghadam, S., Jeje, O., and Mattar, L. 2011. Advanced gas material balance in simplified format. J.
of Canadian Petroleum Technology, 50(1):90 –98.
Morton, K., Booth, R., Chugunov, N., Fitzpatrick, A., and Kuchuk, F. 2013. Global sensitivity analysis
for natural fracture geological modeling parameters from pressure transient tests. Paper SPE 164894, SPE
EUROPEC/EAGE Annual Conference and Exhibition, 10 –13 June, London, United Kingdom.
Morton, K., de Brito Nogueira, P., Booth, R., and Kuchuk, F. 2012. Integrated interpretation for
pressure transient tests in discretely fractured reservoirs. Paper SPE 154531, SPE EUROPEC/EAGE
Annual Conference and Exhibition, 4 –7 June, Copenhagen, Denmark.
Muskat, M. 1945. The production histories of oil producing gas-drive reservoirs. J. of Applied Physics,
pages 147–159.
Nashawi, I. S., and Malallah, A. H. 2007. Well test analysis of finite-conductivity fractured wells
producing at constant bottomhole pressure. Journal of Petroleum Science and Engineering, 57(3– 4):303–
320.
Nelson, R. A. 1985. Geologic Analysis of Naturally Fractured Reservoirs. Gulf Publishing, Houston,
Houston, Texas, 1st edition.
Olarewaju, J. and Lee, W. 1987. An analytical model for composite reservoirs produced at either
constant bottomhole pressure or constant rate. Paper SPE 16763-MS, SPE Annual Technical Conference
and Exhibition, 27–30 September, Dallas, Texas.
Ozkan, E., Ohaeri, U., and Raghavan, R. 1987. Unsteady flow to a well produced at a constant
pressure in a fractured reservoir. SPE Formation Evaluation, 2(2):186 –200.
SPE-170698-MS 23

Palacio, J. and Blasingame, T. 1993. Decline-curve analysis with type curves - analysis of gas well
production data. Paper SPE 25909-MS, SPE Low Permeability Reservoirs Symposium, 26 –28 April,
Denver, Colorado.
Pollard, D. D., and A., A. 1988. Progress in understanding jointing over the past century. Geological
Society of America Bulletin, 100(8):1181–1204.
Prats, M., Hazebroek, P., and Strickler, W. R. 1962. Effect of vertical fractures on reservoir
behavior— compressible fluid case. SPE J., 2(2):87–94.
Raghavan, R. 1976. Some practical considerations in the analysis of pressure data. J. of Petroleum
Technology, 28(10):1256 –1268.
Raghavan, R. 1993. Well Test Analysis. Prentice Hall, Boston.
Raghavan, R. and Ohaeri, C. 1981. Unsteady flow to a well produced at constant pressure in a
fractured reservoir,. Paper SPE 9902, SPE California Regional Meeting, 25–26 March, Bakersfield,
California.
Rahman, N., Kok, L., and Zaoral, K. 2006. A new method for computing pseudo-time for real gas flow
using the material balance equation. J. of Canadian Petroleum Technology, 45(10).
Sageev, A., Da Prat, G., and Ramey, Jr. H., 1985. Decline curve analysis for double-porosity systems.
Paper SPE 13630, SPE California Regional Meeting, 27–29 March, Bakersfield, California.
Soliman, M. 2009. Transient Well Testing, chapter 11–Well-test analysis of hydraulically fractured
wells, pages 281–331. Monograph Series 23, Edited by M. Kamal. Society of Petroleum Engineers,
Richardson, 1st edition.
Tivayanonda, V., Apiwathanasorn, S., Ehlig-Economides, C., and Wattenbarger, R. A. 2012. Alter-
native interpretations of shale gas/oil rate behavior using a triple porosity model. Paper SPE-159703-MS,
SPE Annual Technical Conference and Exhibition, 8 –10 October, San Antonio, Texas, USA.
Unneland, T., Manin, Y., and Kuchuk, F. 1998. Application of permanent pressure and rate measure-
ment systems in well monitoring and reservoir description: Field cases. SPE Formation Evaluation,
1(3):224 –230.
van Everdingen, A. and Hurst, W. 1949. The application of the laplace transformation to flow
problems in reservoirs. Trans., AIME, 186: 305–324.
von Schroeter, T., Hollaender, F., and Gringarten, A. 2004. Deconvolution of well-test data as a
nonlinear total least-squares problem. SPE J., 9 (4):375–390.
Warren, J. E., and Root, P. J. 1963. The behavior of naturally fractured reservoirs. SPE J., 3(3):245–
255.
Washburne, C. 1915. Estimation of oil reserves. AIME Bull, 98: 46 –97.
24 SPE-170698-MS

Appendix A
Dissimilar Behavior of Pressure, Rate and Derivatives

In this appendix, we give an example for which the pressure and rate, and their derivatives behave quite differently. For this
example, we will use a radially composite reservoir because its solution ]see Olarewaju and Lee (1987) for the qD solution]
and derivatives [see Fig. 7.56 in page 161 of Houze et al. (2009) for the pD derivative] are well known. This is a two-zone
radially composite reservoir, where we keep the outer zone permeability k2 constant and vary the inner zone permeability k1.
As can be seen from Figs. 26, 27, 28, and 29, the pressure and rate, and their derivatives behave significantly differently for
the k1/k2 ratios greater than 10. Notice from Fig. 28 that the high k1/k2 ratio pressure derivatives exhibit a pseudosteady-state
flow regime (m ⫽ 1) very quickly because the fluid in the inner zone is quickly depleted and not replenished readily by the
low permeability outer zone. This effect manifests itself very differently in the rate and its derivative behavior. Consequently,
we should not assume that the pressure and rate, or their derivatives behave correspondingly for all reservoir and well
configurations.

Figure 26 —Dimensionless pressures for various k1/k2 ratios in the Figure 27—Dimensionless rates for various k1/k2 ratios in the two-zone
two-zone radially composite reservoir. radially composite reservoir.

Figure 28 —Dimensionless pressure derivatives for various k1/k2 ratios Figure 29 —Dimensionless rate derivatives for various k1/k2 ratios in the
in the two-zone radially composite reservoir. two-zone radially composite reservoir.

You might also like