Sie sind auf Seite 1von 24

SPE 166296

Pressure Transient Tests and Flow Regimes in Fractured Reservoirs


Fikri Kuchuk, SPE, and Denis Biryukov, Schlumberger
Copyright 2013, Society of Petroleum Engineers
This paper was prepared for presentation at the SPE Annual Technical Conference and Exhibition held in New Orleans, Louisiana, USA, 30 September2 October 2013.
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. Fractures are common features in many well-known reservoirs. Naturally fractured reservoirs consist of
fractures in igneous, metamorphic, and sedimentary rocks (matrix). Faults in many naturally fractured carbonate
reservoirs often have high-permeability zones and are connected to numerous fractures that have varying conductivi-
ties. Furthermore, in many naturally fractured reservoirs, faults and fractures can be discrete (rather than connected
network dual-porosity systems).
In this paper we investigate the pressure transient behavior of continuously and discretely naturally fractured
reservoirs using semi-analytical solutions. These fractured reservoirs can contain periodically or arbitrarily distributed
finite- and/or infinite-conductivity fractures with different lengths and orientations. In some of our earlier papers,
we showed that in terms of pressure transient, neither continuously nor discretely fractured reservoirs behave like
the Warren and Root (1963) dual-porosity model. Unlike the single-derivative shape of the Warren and Root model,
derivatives of pressure transient tests in fractured reservoirs exhibit more than 10 flow regimes. There are seven
important factors that dominate the pressure transient test-flow regime behavior of fractured and faulted reservoirs;
i.e., fractures and/or faults that: 1) intersect the wellbore parallel to its axis, with a dipping angle of 90◦ (vertical
fractures), including hydraulic fractures; 2) intersect the wellbore with dipping angles from 0◦ to less than 90◦ ; 3)
are in the vicinity of the wellbore; 4) have extremely high- or low-fracture and fault conductivities; 5) have various
sizes and distributions; 6) have high- and low-matrix block permeabilities; and 7) have damaged fractures (skin zone)
due to drilling and completion operations and fluids. All flow regimes associated with these factors are shown for a
number of continuously and discretely fractured reservoirs with different well and fracture configurations. For a few
cases, these flow regimes are compared with those from the field data. History matchings of the pressure transient
data generated by using our discretely and continuously fractured reservoir model with the Warren and Root (1963)
type dual-porosity models are also presented.

Introduction
Both fractures and faults are common features of many well-known reservoirs because many geological formations
are fractured to some extent as a result of stress triggered by Earth’s crust weight, high fluid pressure, tectonic forces,
and/or thermal loading. In this paper, fractures denote both fractures and faults, unless differentiation is necessary.
Fractures occur at a variety of scales from microscopic to continental and are important in hydrogeology, the earth
sciences, and petroleum engineering. The Committee on Fracture Characterization and Fluid Flow (1996) stated
that fractures create traps, act as conduits to oil and gas migration, can present barriers or baffles to fluid flow, and
have a significant influence on the behavior and performance of reservoirs. The Committee on Fracture Characteri-
zation and Fluid Flow (1996) presented a comprehensive overview that covers most geological, hydrogeological, and
fluid flow aspects of fractures. Contrary to our usual assumptions, reservoirs are heterogeneous at many different
scales, particularly in regard to fractures.
Fractures occur at different scales with different physical and flow properties in the same formation. They there-
fore have a significant impact on the flow properties of reservoir rock and reservoir management in applications from
drilling to EOR. Although they could vary spatially and temporally (fractures are very stress sensitive), fundamen-
tally, fractured formations have at least three scales of permeability: fault, fracture, and matrix. Micro fractures
and vugs (small cavities in rock) usually increase the scale of permeability. In reality, many fractured reservoirs
contain layers with different properties. Although important particularly in respect to faults, layered reservoirs are
not included in this study.
In this paper, we investigate pressure transient behavior of both continuously and discretely fractured reservoirs
and identify the flow regimes they exhibit. In some cases we compare these flow regimes with those from the field
data. We also present two examples of history matching of the pressure transient data generated by using the
Biryukov and Kuchuk (2012) model for discretely and continuously fractured reservoirs with the Warren and Root
(1963) dual-porosity model. When we refer to the Warren and Root (1963) dual-porosity model, we are including
all its variations added since the 1960s.
SPE 166296 2

Geological Considerations
Nelson (1985) classified naturally fractured reservoirs as follows:
• Type I: Only the fractures contribute to reservoir storage capacity (porosity) and conductivity (permeability).
• Type II: The fractures provide overall conductivity (permeability) of the reservoir, and the matrix elements
provide overall storage capacity (porosity), but the matrix is permeable enough (low permeability) to provide
conductivity for flow from the matrix into fractures.
• Type III: Both fractures and matrix provide conductivity (permeability) of the reservoir, but overall storage
capacity (porosity) is primarily in the matrix.
• Type IV: The fractures are nonconductive (sealed, filled with minerals, or have zero fault-band permeability).
Both conductivity (permeability) and storage capacity are provided by the matrix. These types of nonconduc-
tive fault and/or fractured systems create compartmentalized reservoirs.
From the reservoir point of view with regard to reservoir simulations and performance, and PTT (pressure tran-
sient test) interpretation, the Nelson (1985) classification disregards a large number of fractured reservoirs and does
not provide any advantages in terms of information about flow in faults, fractures, and the matrix, and particularly,
about fault and fracture conductivities and their interdependence with each other. Kuchuk and Biryukov (2012)
divided fractured reservoirs into four categories as follows:
1. Fractures create a network, communicate hydraulically with each other globally, and provide overall conduc-
tivity (permeability) of the reservoir. The matrix provides overall storage capacity (porosity), but it should be
permeable enough to provide conductivity for the flow from the matrix into the fractures. If matrix perme-
abilities are high or similar to fracture permeabilities, or ultra low, the pressure transient behavior of fractured
reservoirs will be similar to the behavior of homogeneous reservoirs. We call this continuously fractured reser-
voirs. It 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 often have non-orthogonal continuous fracture sets
(Pollard and A., 1988), as well as secondary fractures that are non-orthogonal to each other and to primary
fracture sets.

2. Fractures do not form a continuous conductive network; only a limited number of fractures communicate
hydraulically with each other. 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. Stratified reservoirs that
consist of both fractured and non-fractured 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 non-orthogonal discontinuous fracture sets (Pollard and A., 1988), while only a few have orthogonal
discontinuous fracture sets.
3. Fractures are non-conductive (sealed due to mineralization or faults due to shale smearing). Both conductivity
(permeability) and storage capacity are provided by the matrix. These are called compartmentalized reservoirs.
4. Only fractures are conductive; the matrix has no permeability and/or porosity. Reservoirs with these fractures
are called unconventional fractured basement reservoirs.
The first three types can occur in the same reservoir in different locations, depending on the past tectonic
history of the formation and other mechanical and chemical processes. Moreover, faults in fractured reservoirs will
further complicate the above classification. We therefore infrequently observe only one of the Nelson (1985) fractured
reservoir types; we normally observe a combination of the four types depending on the location of the reservoir (crest,
flank, etc.), the layer properties, and the geologic makeup of the reservoir. In a given fractured reservoir, high and
low conductive and closed fractures with many different sizes, dip angles, and azimuths are distributed throughout
the system, as shown in Fig. 1 (Souche and Rotsch, 2007). The same figure also shows two sub-seismic faults among
a large number of open and closed fractures. It is very important in any fractured reservoir modeling to include both
open and closed fractures and faults.
Although the extent to which rocks exposed at the surface providing analogues of subsurface reservoirs is limited,
outcrop studies provide a basic knowledge of natural fracture systems that can help geologists and engineers make
the most effective use of the limited data on the fracture system in a subsurface reservoir (Odling et al., 1999). Many
fractures on the outcrops at the surface look either more vertical or exactly horizontal. On the other hand, wellbore
fracture images show that dip angles can change from 0◦ to 90◦ , but more frequently they are in the range of 45◦ to
SPE 166296 3

80◦ . To understand the PTTs’ derivative signatures, the structural setting of the formation from wellbore to aquifers
is of prime importance. For instance, fractures intersecting the wellbore dominate PTT behavior throughout the test.
Often, naturally fractured reservoirs are also hydraulically or acid fractured, or simply acidized. Natural fractures
have a large impact on the hydraulic fracturing process and hamper the ability to place proppant in the hydraulic
fracture, as well as caus a large amount of fluid loss. In the case of acidizing treatments, the injected acid might only
penetrate into the conductive fractures without cleaning the damaged or plugged fractures.
Another important fluid flow aspect
Orientation of fractured reservoirs is the presence of
Perpendicular to bedding Conjugated fracture corridors: clusters of parallel
or near-parallel (globally in the same
direction) fractures. Fracture corridor
sizes may vary from less than one meter
to 10 m wide, 100 m high, and 1000 m
Main long with very high permeability, e.g.,
fault above 10 Darcy. Each fracture corridor
can contain hundreds to tens of thou-
sands of fractures (Singh et al., 2008).
Fracture corridors affect all aspects of
drilling, logging, production, and reser-
1400 m 2100 m voir engineering. They significantly af-
fect the PTT behavior of wells, partic-
ularly if they are within a few hundred
meters of the wellbore.
Cemented (mineralized) In general, long dominant conduc-
tive and non-conductive fractures in any
given system should be treated indi-
Open
vidually in any solution, whether it is
single- or multi-phase flow. On the other
hand, low conductivity and/or short frac-
tures can be treated as an enhancement
Sub-seismic fault zone of matrix properties. For this case, ei-
ther a single-porosity model with aver-
Figure 1: Fracture characteristics in a horizontal well, from Souche and Rotsch (2007). age properties, or the Warren and Root
(1963) dual-porosity type model can be
used locally while maintaining global heterogeneity, and particularly anisotropy, because even small fractures may
have a dominant direction.

Reservoir Engineering Considerations


Kuchuk and Biryukov (2012) presented an extensive literature review and the historical development of Warren and
Root (1963) dual-porosity type models with their numerous extensions. Kuchuk and Biryukov (2012) demonstrated
with their analytical solution for continuously naturally fractured reservoirs, and also confirmed with many field
examples, that both Warren and Root (1963) and Kazemi (1969) dual-porosity type models, shown in Fig. 2 (b) and
(c), have a very narrow range of applications for fractured reservoirs and do not contain any fractures. These models
transform a fractured reservoir into an “equivalent” homogenous medium without fractures, as shown in Fig. 3 (a)
and (b). To distinguish between the Warren and Root (1963) dual-porosity sugar-cube (matrix element) geological
model with three orthogonal fracture sets [Fig. 2 (b)], and the Warren and Root (1963) “equivalent” homogenous-
medium model with matrix source terms [Fig. 3 (b)], we call the former the Warren and Root (1963) dual-porosity
geological model and the latter, the Warren and Root (1963) dual-porosity model. Other matrix element geometrical
shapes, such as spheres, cylinders, and parallelepipeds, are also used.
In the Warren and Root (1963) model, the matrix element is actually a control volume V (Representative
Elementary Volume) that contains a large number of physical matrix blocks and fracture segments of the fractured
system (Gringarten, 1984; Raghavan, 1993). A portion of the control volume V, which envelopes the Warren and
Root (1963) matrix elements, creates an “equivalent” homogenous-medium that represents fracture segments, as
shown in Fig. 3 (b). This is a sort of upscaling for the fractured reservoirs. The properties of the control volume V ,
such as porosity and permeability, have nothing to do with the actual fractures, their distribution, or their properties.
Pressure transient diffusion in actual fractures does not behave like it does in the Barenblatt et al. (1960), Warren
and Root (1963), and Kazemi (1969) models, although exceptions may occur. These models sometimes create a
fractured reservoir look-alike behavior; consequently, estimated reservoir parameters do not represent the actual
matrix and fracture properties. Furthermore, the control volume contains a large number of physical matrix blocks.
SPE 166296 4

Therefore, it does not make any physical sense to place a damaged skin region on the control volume matrix blocks.
It is possible to have calcite deposit, called mineralization, on the actual matrix surfaces and in the fractures. In
general, fracture mineralization mostly affects small-aperture fractures, reduces the fracture conductivity, and often
produces non-conductivite closed fractures (see Fig. 1), while many nearby fractures can be conductive. The field
core and log data indicate that fracture mineralization is not uniformly distributed (Fig. 1), and that matrix surface
calcite deposition is not observed frequently. Due to fracture mineralization, continuously fractured reservoirs may
become discrete systems. If there is a reduced permeability region due to calcite deposit (diagenesis) on the whole
surface of each physical matrix block, then it will reduce the overall permeability of the Warren and Root (1963)
matrix element control volume V . In other words, because the matrix element control volume V contains a large
number of physical matrix blocks, the core matrix permeability will differ from the reduced upscale permeability
of the volume V . The calcite deposit regions are distributed non-uniformly in the surface and internal sections of
the matrix element control volume V . Therefore, the calcite deposition will prolong the matrix element transient
period in the Warren and Root (1963) model due to the permeability heterogeneity created by the non-uniformly
distributed calcite deposit regions, rather than acting as a damaged skin on the surface of the matrix element control
volume V .

Source matrix elements


Matrix

Matrix
Matrix
Matrix

Matrix

Fractures
Vugs
Fractures Fractures

Vugs Homogeneous reservoir


Actual reservoir Sugar cube representation Slab representation Wellbore

(a) (b) Warren & Root model (c) Kazemi model (a) Actual reservoir (b) 1D “equivalent” model

Figure 2: Idealization of a naturally fractured heterogeneous porous Figure 3: (a) Actual and (b) the “equivalent” Warren and Root
medium: (a) actual reservoir, (b) the Warren and Root (1963) (1963) and Kazemi (1969) homogenous dual-porosity reservoir
model, and (c) the Kazemi (1969) model. model.

Furthermore, the inner boundary conditions used in the Barenblatt et al. (1960), Warren and Root (1963), and
Kazemi (1969) models are unsuitable for fractured reservoirs. The field data examined by Kuchuk and Biryukov
(2012) do not, in general, exhibit the behavior of these models. Unlike the single derivative shape of these models, the
derivatives of field examples exhibit many different flow regimes depending on fracture length, distribution, intensity,
conductivity, etc. Of course, there might be a few naturally fractured reservoirs, for which these models may work
at middle or late times, where wellbore-intersecting fractures and near-wellbore region with fractures are lumped in
the mathematical model as a near-wellbore negative skin—which itself is lumped in with the wellbore radius. That
means the wellbore and near-wellbore lumped-parameter model will not exhibit linear, bilinear, etc. flow regimes at
early times.
In an earlier paper (Biryukov and Kuchuk, 2012), we presented transient pressure solutions for conductive and
non-conductive faults and fractures. The details of the models and derivations are given in the paper. Here we
investigate pressure transient behavior of naturally fractured reservoirs with conductive fractures.
In general, the near-wellbore region dominates drawdown and buildup (injection and falloff) pressure transient
behavior of a well. This is true for all types of reservoirs. The near-wellbore region begins with the sandface (the
formation that is intersected by the wellbore) and may extend to a few hundred feet laterally. It should be noted that
the distance given here is relative. The actual distance depends of the diffusivity of the system k/(φµct ), although
pressure diffusion in the fluid-filled porous medium due to a pressure or a rate pulse in a well is sensed instantaneously
everywhere in the medium. Here we are referring to observable pressure change at a distance. After a pulse at the
wellbore, pressure starts to diffuse and we observe pressure change at a given space and time above the pressure
gauge resolution and natural background noise.
In fractured reservoirs, the depth of the near-wellbore region may change considerably depending on whether the
well intersects fractures or not, and/or whether the fracture network is continuous or discrete. Figure 4 shows a
possible fracture and formation interface. In this zone, fractures may intersect the wellbore, as shown in Fig. 4(c).
Most formation wellbore-image logs show that fractures intersect the wellbore as swarms or clusters with certain dips
and azimuths. Dip angles are usually higher than 45◦ but very rarely 90◦ (vertical fractures). If fractures do not
intersect the wellbore, as shown in Fig. 4(b), then the formation matrix dominates the flow behavior of the system
until the presence of fractures is sensed by the wellbore.
SPE 166296 5

As shown in Fig. 4, there are more than a few interfaces between the wellbore and fractures, between fractures, and
between fractures and matrix elements in the near-wellbore region that can laterally extend to the whole reservoir.
A combination of these interfaces creates a few possible models for the near-wellbore region and/or the reservoir. Of
course, in reality there are many more possibilities. These are possible models:
1. One or more fractures intersect the wellbore in a continuous fracture network (a). This is the dual-porosity
geological model for which Barenblatt et al. (1960) and Warren and Root (1963) presented solutions, although
their solutions do not contain any explicitly modeled fractures.
2. None of the fractures intersects the wellbore in a continuous fracture network (b). The well is in the matrix
and communicates with fractures through the matrix.
3. The well intersects fractures in a discretely fractured reservoir, where discrete fracture sets may communicate
with each other through the matrix (c). In this system some fractures may intersect each other locally.
4. None of the fractures intersects the wellbore in a discretely fractured reservoir, where discrete fracture sets may
communicate with each other through the matrix (d).

Well Well
Porous Porous
medium medium
(matrix) (matrix)

Fractures
Fractures

(a) (b)
Well Well
Porous Porous
medium medium
(matrix) (matrix)

Fractures Fractures

(c) (d)

Figure 4: Schematic of near-wellbore region in a naturally frac-


tured formation, modified from Bear (1993): (a) fractures in-
tersect the wellbore in a continuous fracture network, (b) frac-
tures do not intersect the wellbore in a continuous fracture
network, (c) fractures intersect the wellbore in a discretely frac-
tured reservoir, (d) fractures do not intersect the wellbore in a Figure 5: Formation resistivity images along the wellbore over one of
discretely fractured reservoir. the tested zones, and the location of packer and probes.

The near-wellbore region can be probed by many different measurements through the wellbore. Cores and
openhole logs, particularly from near-wellbore resistivity images, give information about fracture dip angles, azimuths,
apertures, and distribution. If resistivity images indicate open fractures, it is most likely that they are open and
conductive fractures. If resistivity images indicate closed mineralized fractures, it is still possible for them to be
conductive fractures because after mineralization, tectonic events and continuous stress buildup may easily unconnect
the fracture surface from the mineralized (filing) material. Even one tenth of a millimeter would be enough to create
a very conductive fracture. Wireline formation testers (WFT) (Zeybek et al., 2002) can also be used to characterize
both matrix and fracture permeabilities of this near-wellbore region.
Zeybek et al. (2002) presented a field example, in which a total of 12 interval pressure transient tests (IPTT)
were conducted over the fracture zones in a layered 91-ft carbonate formation with a Wireline formation tester
(WFT)—dual-packer and two observation probes—for the estimation of horizontal and vertical permeabilities and
the delineation of fracture conductivities. The resistivity images are shown in Fig. 5 over the tested zone and the
location of the WFT. The resistivity image to the left is static, whereas the image to the right is dynamically
enhanced. This is a small section of the image log for the entire formation. As shown in the figure, the well is
intersected by fracture sets (swarms) with dipping angles between 30◦ and 60◦ . The resistivity images indicate that
all fractures in this zone are mineralized. However, 12 IPTTs indicated that all fractures in all tested zones except
SPE 166296 6

one were conductive. In other words, the effective permeability of a set of fractures in a given zone is about one or
two orders of magnitude higher than the formation matrix permeability.
As we discussed above, when a vertical well is drilled in a fractured formation, there are only two possibilities in
a continuously or discretely fractured reservoir: 1) the well does not intersect with any fracture, i.e., the well is in
the matrix: (b) and (d) of Figure 4 or 2) the well intersects with one or multiple fractures and a single or multiple
matrix blocks, because well diameters are much larger than the fracture apertures. It is therefore no possibility
for the well to only intersect a fracture(s): (a) and (c) of Figure 4. The matrix block may or may not contribute
flow directly into the wellbore, depending on the matrix permeability and the conductivity of the fracture(s). If the
contrast between matrix and fracture permeabilities is low, then flow occurs from the matrix into the wellbore. As
shown in Figure 4, these are the geological possibilities for well and fractured reservoir interactions:
There are seven important geometrical and geological situations that dominate the PTT flow behavior of fractured
reservoirs:
1. Wellbore-intersecting fractures parallel to the wellbore axis with a dip angle of 90◦ (vertical fractures), including
hydraulic fractures.
2. Wellbore-intersecting fractures with dip angles from 0◦ to less than 90◦ .
3. Non-wellbore-intersecting fractures in the few-hundred-foot vicinity of the wellbore.
4. Fracture conductivities, particularly very high or very low conductivities: kf ≈ ∞ or kf ≈ 0.
5. Fracture densities, lengths, and orientations.
6. Very low matrix block permeabilities or km ≈ 0.
7. Damaged fractures (skin zone) due to drilling and completion operations and fluids and partially contributing
perforations.
Of course, there are many other formation and fracture characteristics that affect the pressure transient behavior of
fractured reservoirs.
Finite-conductivity analytical solutions given in the literature have ignored the presence of the well when it is
intersecting a fracture(s) in hydraulically fractured wells, naturally fractured reservoirs, or both. As the conductivity
becomes low–if, for instance, it is damaged in the vicinity of the wellbore–we observe a radial flow before any
other flow regimes, and then a combination of linear and radial. If fractures intersect the wellbore with dip angles
that are less than 90◦ , we will observe a radial flow before other flow regimes, followed by a combination of radial
and formation linear flow regimes, except in the infinite conductivity case. If the wellbore-intersecting fractures
intersect other nearby fractures, then the formation radial and fracture linear flow regimes may progressively appear,
particularly if fractures have dominant directions. After the fracture linear flow, another formation linear flow may
take place in a second set of near-wellbore fractures. When these flow regimes begin and end, their interactions with
each other depend on the seven factors listed above. The PTT examples given often exhibit radial (m = 0), linear
(m = 1/2), bilinear (m = 1/4), trilinear (m = 1/3), m = 1/5, spherical (m = -1/2) and other flow regimes. For
wellbore-intersecting infinite-conductivity fractures without skin and storage, the first flow regime will always be the
formation linear flow, regardless of their orientations and dip angles.
In some reservoirs, derivatives exhibit a fractured pseudosteady-state flow regime 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, which
means it happens 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 pseudosteady-state flow.
All these factors affect pressure transient behavior of discretely and continuously fractured reservoirs. Pressure
data have the sensitivity to these factors to determine them with certainty provided that we have: 1) a reasonably
constrained geological model, 2) accurate and high resolution pressure measurements, 3) accurate and reasonable
magnitude flow rate measurements, 4) a minimized wellbore storage-dominated period with downhole shut-in tools,
and 5) sufficiently long drawdown and buildup tests. The accuracy and resolution of pressure gauges have continuously
improved since the 1930s and are still improving; the resolution of electronic quartz gauges is now about 0.002 psi
(Kuchuk et al., 2010). The superior resolution has improved the quality of derivatives, so we do not have to severely
smooth PTT data to minimize the distortion of geological signatures. Downhole shut-in and accurate surface systems
have significantly improved the quality of pressure and flow rate measurements. Downhole shut-in tools normally
reduce the wellbore-storage dominated period to minutes from hours, and minimize significantly the effects of fluid
phase segregation in the production string. Many flow regimes of fractured reservoirs have not been observed in the
earlier field data due to long wellbore storage dominated flow periods, as well as heavy smoothing of the pressure
data to obtain smooth derivatives.
SPE 166296 7

Pressure Transient Flow Regimes of Continuously Fractured Reservoirs


The majority of vertical wells are perforated to communicate with natural fractures and/or to create hydraulic
fractures. Most natural fractures do not penetrate the entire pay zone. Therefore, 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 these flow regimes and often distort them for a considerable period of time.
Furthermore, drilling and completion operations and fluids, and 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 connected fractures. As a result, a severe skin-damaged fracture
zone (choked fracture) is formed. The behaviors of damaged infinite-conductivity and/or low conductivity fractures
are initially dominated by wellbore storage effects at early times, particularly without downhole shut-in. In addition,
damaged fractures with dip angles that are less than 90◦ amplify the wellbore storage effects.

Standard Flow Regimes


Before presenting flow regimes for fractured reservoirs, in this short section we will describe the standard (common)
flow regimes. In general, we may observe one or more of them for a given well in a given reservoir as a function of time.
Flow regime identification that leads to model identification is a crucial step in pressure transient test interpretation.
Skin and wellbore storage affect most of the early time behavior of well tests and therefore distort the flow regime
signature (behavior) until their effects diminish. Downhole flow rate measurements for drawdown tests and downhole
shut-in for buildup tests minimize skin and wellbore storage effects. It is customary in pressure transient testing
to divide flow regime as follows: 1) Early time period, 2) Middle time period, and 3) Late time period. Figure 6
presents flow regimes for a fully penetrated circular well in a 1D radial reservoir with constant pressure and no-flow
outer boundaries. As shown in Fig. 6, the early-time period derivatives look like an unsymmetrical bell-shape curve
with four distinct periods: 1) a unit slope (m = 1) period, 2) a maximum arch interval, 3) a negative slope (m = −1)
period, and 4) a transition period before the infinite-acting radial flow regime(Kuchuk et al., 2010). The flow periods
leading up to the radial flow regime are due to skin and wellbore storage effects, while the sandface flow rate increases
slowly to reach a stabilized-constant flow rate. The skin and wellbore-storage dominate the early time period; during
the middle time period the derivative exhibits an infinite-acting radial flow regime. If there is no inhomogeneity or
heterogeneity in the reservoir, the radial flow regime continues until the effects of the constant pressure boundary
or the no-flow outer boundary are felt by the wellbore, as shown in Fig. 6. The derivative for the constant-pressure
boundary or for the no-flow outer boundary deviates from the infinite-acting radial behavior during the late time
period, where the steady-state flow regime or pseudosteady-state regime is observed.

10 CD=0, S=0 & pseudosteady-state spherical blocks


CD=1000, S=10 & pseudosteady-state spherical blocks
8 CD=0, S=0 & transient spherical blocks
7 CD=1000, S=10 & transient spherical block
infinite-acting 6 CD=0, S=0 & homogenous model
100 m=1 constant-pressure 5
noflow bounded
4
m = -1
6
5 pseudo-steady state, m = 1 3
unit slope
4
2
Derivatives, psi

3
dpD/dlntD

first radial third radial


2
infinite acting radial 1 second radial
8
10 7
6
5
6 steady-state 4
5
4 3
3
2
m = 1/2
2
early time middle late time
time
0.1
1 2 3 4 5 6 7 8
10 10 10 10 10 10 10
0.01 0.1 1 10 100 tD
Time, hr
Figure 7: Derivatives and flow regimes for a dual-porosity model
Figure 6: Derivatives and flow regimes for a reservoir with constant- (Warren and Root, 1963) with and without wellbore storage and
pressure or no-flow boundary conditions, and wellbore storage and skin effects for various matrix block flow conditions in an infinite
skin effects, after Kuchuk et al. (2010). reservoir, after Kuchuk and Biryukov (2012).

In general, there can be more than a few flow regimes in each period. For instance, a system may exhibit linear,
spherical, and radial flow regimes in the early time period. Therefore, the above three flow periods shown in Fig. 6
are only for a well in a circular reservoir with either a no-flow boundary or a constant pressure outer boundary. As
we show in this paper, many more flow regimes can be observed in fractured reservoirs. Fractured reservoir flow
regimes can appear along with and in between the fundamental flow regimes shown in Fig. 6.
SPE 166296 8

In this paper, we often refer to the Warren and Root (1963) dual-porosity type model behavior and its derivatives.
For completeness, Fig. 7 presents dimensionless pressure derivatives for the pseudosteady-state and transient flow
conditions in spherical matrix blocks, where λ = 10−7 and ω = 0.06, with and without wellbore storage and skin.
For this case, the wellbore radius is used for the definition of the dimensionless time (unlike in Eq. 1). The same
figure also presents the dimensionless derivative for a homogenous system without wellbore storage and skin effects
for comparison. As can be seen in this derivative plot, we have named the flow regimes as: first radial, second
(transition) radial, and third radial.
All synthetic example pressure transient data for fractured reservoirs are generated with the general fractured
reservoir model presented by Biryukov and Kuchuk (2012). In this paper the following dimensionless variables are
used and defined as (in the oilfield units)

km h 0.0002637km t 5.6146C kf b x y
pD = [po − pwf (t)], tD = 2
, CD = 2
, FD = , xD = , yD = , (1)
141.2qµ µ(φct )m lw 2π(φct )m hlw km lw lw lw

where kf , b, and lw denote the fracture permeability, aperture, and the half length or a reference length, respectively;
km , φm , and (ct )m denote the matrix permeability, porosity, and total compressibility, respectively; C denotes
wellbore storage coefficient; h denotes the reservoir thickness; po and pwf denote the initial reservoir and flowing
wellbore pressures, respectively; q denotes a reference flow rate; µ denotes fluid viscosity; and t denotes time.

Wellbore-Intersecting Fracture
When a vertical well intersects a fracture that is parallel to the wellbore axis (90◦ ), we may observe a fracture
linear (a), bilinear (b), formation linear (c), and trilinear (d), as shown in Fig. 8, assuming the contrast between
the matrix and fracture permeabilities is high. Depending on the fracture conductivity and length, we may observe
only one or a few of these flow regimes in chronological order. These flow regimes are the same as those observed
from transient behavior of hydraulically fractured wells; see Cinco-Ley and Samaniego-V (1981b). If the fracture
has infinite conductivity, only a formation linear flow regime (c) will be observed. If more than one vertical fracture
intersects the wellbore, which is highly unlikely in vertical wells, the flow regimes will be distorted, except in the
infinite conductivity case, where the formation linear flow will be observed. Due to very low fracture conductivity
and/or high matrix permeability, flow entries may take place from matrix blocks directly into the wellbore.
When a vertical well intersects a finite-conductivity fracture with an angle that is less than 90◦ , we may observe
a fracture radial (e) and formation linear (f), as shown in Fig. 8, assuming the contrast between the matrix and
fracture permeabilities is high. This fracture radial flow will yield the fracture permeability and aperture product.
As in the vertical fracture case, the infinite-conductivity fractures will exhibit a formation linear flow, as also shown
in Fig. 8 (f).
After these flow regimes, one or more other flow regimes can be observed for a given fractured formation. We
may also observe an elliptical flow regime, which occurs before the formation matrix radial flow, as shown in Fig. 9
(a). However, for most fractured formations, the formation matrix radial flow will not occur unless the well is in the
matrix to begin with, or the extension of the fractured zone is short compared to the entire tested zone. A few more
flow regimes can be observed for both vertical and deviated fractures if the fractures are partially penetrating (they
do not cross the entire formation from the top and bottom boundaries). Therefore we may observe spherical, as shown
in Fig. 9 (a), hemispherical, and ellipsoidal flow regimes. The occurrences and duration of the flow regimes shown in
Figs. 8 and 9 for both vertical and deviated fractures depend on both the matrix and fracture permeabilities, and
the length of the fractures. All these flow regimes occur in a certain chronological order.
It is difficult to understand the pressure transient behavior of fractured reservoirs without understanding the
behavior of a single-wellbore-intersecting fracture. Because of the complexity of fractured reservoir modeling, we
cannot create a single fractured reservoir model to show all possible flow regimes. In this paper we often look
at the behavior of systems locally. In other words, we look at the individual behavior of certain features of the
model. In reality, the effects of fractured reservoir features may interfere with or mask each other. The complexity
of the behavior is increased when we add the effects of skin and wellbore storage. All pressure transient data are
affected, sometimes significantly by skin and wellbore storage. Therefore it is crucial to understand the behavior
of a single wellbore-intersecting fracture with skin and storage effects (Cinco-Ley and Samaniego-V, 1977, 1981a;
Evans, 1971; Raghavan, 1976) to identify flow regimes that take place around the wellbore and fractures. Many flow
regimes for a single vertical fracture have been discussed in the literature (Cinco-Ley and Samaniego-V, 1981b; Lee
and Brockenbrough, 1986; Soliman, 2009), but we are looking at them from a different perspective, particularly in
connection to fractured reservoirs.
Figure 10 presents derivatives for a single vertical wellbore-intersecting fully-penetrated infinite-conductivity
fracture with various dimensionless wellbore storage values. The skin S value is 1 for all cases, and is zero for CD = 0
(without wellbore storage). For all the storage cases, all derivatives clearly exhibit a positive unit-slope (m = 1)
wellbore-storage dominated period. The derivative without the wellbore storage effect (CD = 0) exhibits a 1/2-slope
formation linear flow regime. As can be observed from Fig. 10, even if it is distorted, the derivative for CD = 0.001
SPE 166296 9

Well
Well

Fracture

Fracture
Fracture intersecting wellbore
axis in parallel

Fracture intersecting wellbore


with an angle
(a) Fracture linear flow

(a) Elliptical (transition) flow (b) Spherical flow (c) Directional fracture linear flow
(e) Radial flow
(b) Bilinear flow

(f ) Formation linear flow


(c) Formation linear flow

(d) Directional fracture-formation (e) Directional fracture formation (f) Pseudo radial flow
(d) Trilinear flow
bilinear flow linear flow

Figure 8: Schematics of flow regimes for a vertical well intersect a Figure 9: Schematics of common flow regimes for a vertical well
partially penetrating fracture parallel to, and with an angle to the intersect a partially penetrating fracture parallel to, and with an
wellbore axis. angle to the wellbore axis.

still reveals the fractured nature of the well, because after a minimum (dip) we can observe a short 1/2-slope period.
For the CD = 0.01 case, it is difficult to say that the well intersects an infinite-conductivity fracture. The derivative
behavior for the CD = 0.1 case is very strange: After the unit-slope period, it immediately becomes flat (its value
approaches 0.5—a radial flow regime) almost without a transition period after a maximum arch interval due to the
wellbore storage (see Fig. 6).
In this paper, if a synthetic derivative example behavior is similar to some field data derivative behavior, we
present them as well. The objective of showing these field examples is not to claim that that is the model for a given
test because we do not have enough geoscience data to make such a claim. It could easily be a look-alike behavior.
Our objective is to show that there is an alternative model to the Warren and Root (1963) dual-porosity type models.

1
10000 10000
Recorded pressure data
Unity slope due to afterflow
dpD/dlntD

1000 1000
Δp and derivative, psi

Pressure derivative
Δp and derivative, psi

Half slope for linear flow


CD = 0 100 100
0.1 CD = 0.001
m=1 CD = 0.01 m = 1/2
CD = 0.1 10 10
m=1
m = 1/2 m=1 Recorded pressure data
1 1 Unity slope due to afterflow
Pressure derivative
0.01 Half slope for linear flow
-5 -4 -3 -2 -1 0 1 2 0.1 0.1
10 10 10 10 10 10 10 10
0.01 0.1 1 10 100 1000 0.01 0.1 1 10 100 1000
tD Δt, hr Δt, hr

Figure 10: Derivative of a single wellbore-intersecting fully- Figure 11: Derivative plots of the buildup tests before and af-
penetrated infinite-conductivity fracture with skin and storage in ter fracturing in the Spraberry reservoir, from Guo and Schechter
an infinite reservoir. (1998).

The Spraberry reservoir is one of the most well-documented fractured oil reservoirs. Guo and Schechter (1998)
presented the derivative plots of two buildup tests before and after hydraulic fracturing in the Spraberry reservoir,
as shown in Fig. 11 [Figs. 1 and 4 of the Guo and Schechter (1998) paper]. As can be seen from Fig. 11, these
derivatives behave similarly to the derivatives shown in Fig. 10 for the wellbore storage values CD = 0.001 and 0.01.
SPE 166296 10

Both derivatives given in Fig. 11 exhibit a positive unit-slope (m = 1) wellbore-storage dominated period, lasting
about 10 and 3 hr. The wellbore-storage dominated period for the post-fracture buildup test is much shorter (3 hr)
than the one for the pre-fracture buildup test. This indicates that the fracturing reduced the skin significantly.
Both derivatives also exhibit a 1/2-slope period after a minimum (dip). Therefore, it is most likely that some of the
fractures intersect the wellbore, perhaps the less conductive ones. If the well is intersecting primary fractures, then
it must have been severely damaged or plugged.
In naturally fractured reservoirs, fracture heights are variable and often less than the formation thickness. We
call these fractures partially-penetrating fractures. The pressure transient behavior of partially-penetrating fractures
(Igbokoyi and Tiab, 2008; Raghavan et al., 1978; Rodriguez et al., 1984) is usually different from that of fully-
penetrating fractures. As in the fully-penetrating case, we look at the behavior of partially-penetrating fractures in
an infinite homogenous reservoir in order to isolate their behavior from other effects. Figure 12 presents derivatives for
a single wellbore-intersecting partially-penetrated infinite-conductivity fracture with various dimensionless wellbore
storage and skin values. The penetration ratio b = hf /h is 0.1, where hf and h are the heights of the fracture and
formation, respectively. As can be observed from Fig. 12, the derivative without the wellbore storage and skin effects
(CD = 0, S = 0) exhibits a 1/2-slope formation linear flow regime. After that, the derivative goes down slightly
due to the fracture partial-penetration effect and then becomes flat before reaching the final radial flow regime. For
all the wellbore storage and skin cases, all derivatives clearly exhibit a positive unit-slope (m = 1) wellbore-storage
dominated period, and they all go through a maximum value before dipping down. The derivative behavior for
the CD = 0.1, S = 1 case looks like a typical derivative of a fully-penetrated circular well in an infinite reservoir
with the wellbore storage and skin effects (see Fig. 6). Figure 13 compares the derivatives of fully-penetrating and
partially-penetrating infinite-conductivity fractures. Their behaviors are different, as can be seen in Fig. 13, although
both exhibit a 1/2-slope formation linear flow regime at early times. However, the partial-penetration effect distorts
the 1/2-slope period at early times due to the vertical flow into the fracture from the top and bottom regions.

1
1
m = 1/2
dpD/dlntD
dpD/dlntD

0.1
0.1 CD = 0, S=0
CD = 0.001, S = 1 fully-penetrating
m=1
CD = 0.01, S = 1 m = 1/2 partially-penetrating
m=1 CD = 0.1, S = 1

0.01 0.01
-5 -4 -3 -2 -1 0 1 2 -5 -4 -3 -2 -1 0 1 2
10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10
tD tD

Figure 12: Derivatives of a single wellbore-intersecting partially- Figure 13: Comparison of derivatives of fully-penetrating and
penetrated infinite–conductivity fracture with skin and storage in partially-penetrating infinite-conductivity fractures in an infinite
an infinite reservoir. reservoir.

Figure 14 presents derivatives for wellbore-intersecting partially-penetrated finite-conductivity fractures with


various dimensionless fracture conductivity, wellbore storage, and skin values. The penetration ratio b is also 0.1
for this example. The derivatives become more and more similar to the derivative of a fully-penetrated circular well
with the wellbore storage and skin with a unit slope as the fracture conductivity decreases. In the derivatives for
high fracture conductivities, we observe the effect of the fracture, but the derivatives exhibit a 1/2-slope formation
linear flow regime only for the CD = 0, S = 0 case. As we can see from these two examples, most derivatives are
dominated by wellbore storage and skin effects for fully- and partially-penetrated infinite- and finite-conductivity
fracture cases. Therefore, downhole shut-in is essential for identifying both of these fully- and partially-penetrated
fractures. This is particularly important for low conductivity fractures.
Now let us look at the PTT behavior of a continuously fractured system, where fractures are in a network and
communicate with each other throughout the system, as shown in Fig. 15. This is a 2D representation of the Warren
and Root (1963) dual-porosity geological model, where the matrix blocks consist of 65.6 ft by 65.6 ft by h (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. 15, 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
SPE 166296 11

(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.
As shown in Fig. 16, the derivatives for this continu-
1 ous fractured reservoir exhibit quite different flow regimes
m = 0 (0.5) at early times depending on the fracture conductivities. No-
tice that for the high conductivity (FD = 1, 000 and 10, 000)
Derivative, dpD/dlntD

fractures, the reservoir goes to the pseudosteady-state flow


0.1 regime (m = 1) very quickly; after that, the matrix blocks
FD = 0.1, CD = 0, S = 0
FD = 1, CD = 0, S = 0 start contributing, which develops as a formation linear flow.
m = 1/2 FD = 100, CD = 0, S = 0
FD = 0.1, CD = 0.001, S = 1 In these high conductivity fractures, the fluid in the fractures
m=1 FD = 1, CD = 0.001, S = 1
FD = 100, CD = 0.001, S = 1 is quickly depleted, and is not replenished readily by the ma-
0.01
trix elements. As the pressure declines more and more in the
10
-6
10
-5
10
-4
10
-3
10
-2
10
-1
10
0
10
1
10
2 fracture system, the matrix contributions become significant,
Dimensionless time, tD and the flow regime deviates from the fracture pseudosteady-
state flow regime, which takes place before the formation
Figure 14: Derivatives of a single wellbore-intersecting partially- linear flow (although there are exceptions). If the fracture
penetrated fracture with different conductivities, and skin and system is connected throughout the reservoir, the fracture
storage in an infinite reservoir.
network may drain most of the fluid in the fracture system
before a significant matrix contribution, if its permeability is very low. In reality we may see the effects of outer
boundaries, such as other producing and/or injection wells, sealing faults, acquirer, gas cap, etc., before a significant
matrix contribution. Although it is not shown in the figure, a very short-duration fracture-linear flow also takes place
in these high conductivity fractures before the pseudosteady-state flow regime. For lower conductivity fractures, we
do not observe a pseudosteady-state flow regime; after several transition periods, we observe the formation linear flow
for FD = 1 and 10. The derivatives for FD = 0.001, 0.01, and 0.1 do not exhibit a 1/2-slope formation linear flow
regime. Fractures with low conductivities (FD = 0.001 and 0.01) almost do not affect the behavior of the system.
All derivatives exhibit stabilized flat (m = 0 slope) periods at early times, except for FD = 1, 000 and 10, 000. As
can be seen from Fig. 16, the stabilized flat period is very short for FD = 100. The derivatives FD = 0.1, 1, and 10
exhibit two stabilized flat periods each.

1
m = 0 (0.5) m = 0 (0.5)

m = 1/2
Derivative, dpD/dlntD

0.1
Matrix

Well
FD = 0.001
Fracture FD = 0.01
0.01 FD = 0.1
FD = 1
m=1
2lw
FD = 10
FD = 100
FD = 1000
2lw Matrix
0.001 FD = 10000

0.0001
0.0001 0.001 0.01 0.1 1 10 100
Dimensionless time, tD

Figure 16: Derivatives of a fracture-intersecting well in the War-


ren and Root (1963) geological model (shown in Fig. 15), where
Figure 15: Warren and Root (1963) uniformly distributed fracture fractures and matrix blocks are uniformly distributed with different
and matrix block geological model. fracture conductivities.

Figure 17 compares the behavior of high and low fracture densities in the Warren and Root (1963) geological
model (shown in Fig. 15) for a fracture-intersecting well. As can be observed in this figure, the apparent conductivity
of the system increases at early and middle times with increasing fracture density. In other words, the wellbore
pressure drop increases at early and middle times with decreasing fracture density. The fracture density effect is
more pronounced for medium fracture conductivities: FD = 1, 10, and 100.
The derivatives for FD = 100 shown in Fig. 17 are almost the same as the derivative shown in Fig. 18 (Kuchuk
and Biryukov, 2012). The buildup data given in this figure is presented by Ayestaran et al. (1989) from a Tertiary
limestone 350-ft thick reservoir that has a low matrix porosity but is extensively naturally fractured.
SPE 166296 12

LD = low density
1
HD = high density m = 0 (0.5)
100 m = 1/2
m = 1/2
Derivative, dpD/dlntD

0.1

FD = 0.1 HD
FD = 1 HD 10

(tD/CD)p'D
FD = 10 HD
0.01 FD = 100 HD
FD = 10000 HD
m=1 FD = 0.1 LD m=0
FD = 1 LD
FD = 10 LD
0.001 FD = 100 LD 1
FD = 10000 LD m=1

0.0001
0.1
0.0001 0.001 0.01 0.1 1 10 100
1 2 3 4 5
Dimensionless time, tD 10 10 10 10 10
tD/CD
Figure 17: Comparison of derivatives for high and low fracture
densities in the Warren and Root (1963) geological model (shown Figure 18: Pressure derivative of a buildup test from a fracture
in Fig. 15) for a fracture-intersecting well. reservoir, modified from Ayestaran et al. (1989).

Figure 19 presents derivatives for finite-conductivity frac-


10
tures in the Warren and Root (1963) geological model (shown
in Fig. 15), where the well intersects a fracture with CD =
0.001, S = 5. The wellbore storage and skin effects change
the behavior of derivatives significantly when we compare
Derivative, dpD/dlntD

1 Fig. 16 with Fig. 19. First, the fracture pseudosteady-


m = 0 (0.5) state flow regime for FD = 10, 100, 1, 000, and 10, 000
almost disappears. Second, the stabilized flat periods for
m = 1/2
m=1 low conductivities also disappear. Instead, we have a well-
FD = 0.001
0.1 FD = 0.01 bore storage and skin dominated flow regime (m = 1). The
FD = 0.1
FD = 1 derivatives for FD = 0.001 and 0.01 exhibit an infinite-
FD = 10
m=1 FD = 100 acting radial flow regime after the wellbore storage and skin
FD = 1000
FD = 10000 effects. The derivative for FD = 0.1 exhibits a radially-
0.01 composite look-alike behavior. Finally, the derivatives for
0.0001 0.001 0.01 0.1 1 10 100 FD = 1, 10, 100, 1, 000, and 10, 000 appear to have
Dimensionless time, tD the dual-porosity Warren and Root (1963) model behav-
ior. However, the minimum values of these derivatives are
Figure 19: Derivatives of a fracture-intersecting well with CD = much smaller that those from the Warren and Root (1963)
0.001, S = 5 in the Warren and Root (1963) geological model model; see Fig. 7. This provides an important piece of
(shown in Fig. 15), where fractures and matrix blocks are uniformly identification evidence.
distributed with different fracture conductivities.
Well in the Matrix
Now let us assume that the well does not intersect any of the fractures in the above 2D Warren and Root (1963)
dual-porosity geological model, and is in a rectangular parallelepiped (a square base) matrix block at the center of
the fracture network, as shown in Fig. 15. All other properties of the model are the same as the previous model.
The derivatives for this case shown in Fig. 20 exhibit different flow regimes at middle times. First, they exhibit a
formation radial flow because the well is in a matrix block, and then they all go downward rapidly at early times.
The start of this downward trend depends on the size of the matrix blocks. Fractures with low conductivities
(FD = 0.001 and 0.01) hardly affect the behavior of the system, as shown in Fig. 20. For the high fracture
conductivities of FD = 100, 1, 000, and 10, 000, the derivatives exhibit a pseudosteady-state flow regime (m = 1);
after that, matrix blocks start contributing, which develops as a formation linear flow—a 1/2-slope period, as shown
in Fig. 16. The derivatives exhibit stabilized flat (m = 0 slope) periods for FD = 1 and 10, and are correlated with
the composite fracture-matrix system permeabilities. The derivatives for FD = 1 and 10 exhibit a dual-porosity
Warren and Root (1963) model look-alike behavior. This can be clearly seen from the second radial derivative curve
in Fig. 7 for the spherical matrix blocks with the transient flow condition. After all derivatives go through minimum
values, they finally reach the formation of radial flow because the area of the fracture network is finite.
Figure 21 presents derivatives for finite-conductivity fractures in the Warren and Root (1963) geological model
SPE 166296 13

10
1
m = 0 (0.5) m = 0 (0.5)

Derivative, dpD/dlntD
m = 1/2
Derivative, dpD/dlntD

1
0.1 m = 0 (0.5)
m=1
m = 1/2

FD = 0.001 FD = 0.001
FD = 0.01 0.1 FD = 0.01
FD = 0.1
0.01 FD = 0.1
FD = 1
FD = 1
m=1 FD = 10 FD = 10
FD = 100 FD = 100
FD = 1000 FD = 1000
FD = 10000 FD = 10000

0.01
0.001 0.0001 0.001 0.01 0.1 1 10 100
0.0001 0.001 0.01 0.1 1 10 100 Dimensionless time, tD
Dimensionless time, tD

Figure 21: Derivatives of a well in the matrix (not fracture inter-


Figure 20: Derivatives of a well in the matrix (not fracture inter- secting) with CD = 0.001, S = 5 of the Warren and Root (1963)
secting) of the Warren and Root (1963) geological model (shown geological model (shown in Fig. 15), where fractures and matrix
in Fig. 15), where fractures and matrix blocks are uniformly dis- blocks are uniformly distributed with different fracture conductivi-
tributed with different fracture conductivities. ties.

(shown in Fig. 15), where the well is in the matrix with CD = 0.001, S = 5. The first formation radial flow disappears
when this figure is compared with Fig. 20 because of the wellbore storage and skin effects. As can be seen from Fig. 21,
after the unit slope period, all derivatives go downward rapidly at early times. The derivatives for low conductivities
(FD = 0.001 and 0.01) exhibit an infinite-acting radial flow regime after the wellbore storage and skin effects. The
derivatives for FD = 0.1, 1, 10, 100, 1, 000, and 10, 000 appear to have the dual-porosity Warren and Root (1963)
model behavior; see Fig. 7. These derivatives go through minimum values and finally reach the formation of radial
flow. As shown in Fig. 21, the derivatives for the high fracture conductivities of FD = 100, 1, 000, and 10, 000 exhibit
a 1/2-slope period formation linear flow regime at middle to late times. Due to the wellbore storage and skin effects,
the pseudosteady-state flow regime (m = 1) due to the high fracture conductivities almost disappears; compare
Fig. 21 with Fig. 20. Notice also that the values of the minimum of the derivative have also changed.
The derivatives for FD = 0.001, 0.01, and 0.01 shown in Fig. 21 look like a few of the derivatives presented by
Corre (1990) for buildup tests from the fractured Palanca oil field in offshore Angola, as shown in Fig. 22 (Kuchuk
and Biryukov, 2012).

1 HD = high density LD = low density


m = 0 (0.5) m = 0 (0.5)

m = 1/2
Derivative, dpD/dlntD

10 0.1
PAL 3 (Test 2) PAL 103
PAL 208
Derivative, psi

FD = 0.1 HD
FD = 1 HD
1 0.01 FD = 10 HD
FD = 10000 HD
m=1 FD = 0.1 LD
FD = 1 LD
FD = 10 LD
FD = 1000 LD
PAL 3 (Test 3)
0.001
0.1 0.0001 0.001 0.01 0.1 1 10 100
0.1 1 10 100 1000 10000
tD/CD
Dimensionless time, tD

Figure 23: Comparison of derivatives for high and low fracture


Figure 22: Derivatives for seven buildup tests from vertical wells in densities in the Warren and Root (1963) geological model (shown
the fractured Palanca oil field given by Corre (1990). in Fig. 15) for a well in the matrix.

Figure 23 compares the behavior of high and low fracture densities in the Warren and Root (1963) geological
model (shown in Fig. 15) for a well in the matrix. As can be observed in this figure, the apparent conductivity
of the system increases at early and middle times with increasing fracture density. In other words, the wellbore
pressure drop increases at early and middle times with decreasing fracture density. The fracture density effect is
more pronounced for medium fracture conductivities: FD = 1, 10, and 100.
SPE 166296 14

Pressure Transient Flow Regimes of Discretely Fractured Reservoirs


Very few papers have been published on the pressure transient behavior of discretely fractured systems. Most of
these studies have focused on the pressure behavior of a single sealing or leaky fault, or a single fracture. Kuchuk
and Biryukov (2012) presented a literature review and some flow regimes of discretely fractured reservoirs. In
general, Discrete Fracture Network (DFN) models are also used frequently as continuously fractured models for
characterization of fractured reservoirs. In both continuous and DFN models, the fractures are explicitly modeled
in the solutions of Biryukov and Kuchuk (2012). Here we present a number of examples for discretely fractured
reservoirs for the investigation of their pressure transient behavior.

Wellbore-Intersecting Fracture
This model was first presented by Kuchuk and Habashy (1997) for an infinite discretely fractured reservoir that has
40 parallel equally-spaced finite-conductivity fractures with an infinite extent in the y direction, where the distance
between two adjacent fractures is 20 ft. While keeping the Kuchuk and Habashy (1997) model fundamentally the
same, we have changed fracture lengths from infinity to 1000 ft from the x axis, as shown in Fig. 24, For this case,
a fully-penetrating vertical well intersects the 20th fracture. The fractures in the model have the same conductivity
with a finite small width for a given derivative case.
Figure 25 presents derivatives for various dimensionless finite fracture conductivities without wellbore storage
and skin. As can be seen from this figure, the derivatives for the fracture conductivities of FD = 100 and 106
exhibit a 1/2-slope formation linear flow regime. The derivatives for FD = 0.1, 1, and 10 exhibit a 1/4-slope
bilinear flow regime at early times. Thus, the fracture linear flow regime must have occurred at very early times.
The transition from the bilinear flow regime of FD = 1 to the formation linear flow is very different from a single
wellbore-intersecting finite-conductivity fracture. The derivatives for FD = 0.001, 0.1, and 1 exhibit a radially
composite look-alike behavior. The derivative of FD = 0.01 reaches a radially composite look-alike behavior at very
early times, but the difference between the two radial flow regimes is very small. Of course, it would be a mistake
to use just any radially composite model, including the composite dual-porosity Warren and Root (1963) model,
to interpret this behavior. Cinco-Ley (1996) and Kikani and Walkup (1991) have discussed dual-porosity radially
composite models in connection with the Warren and Root (1963) model.

m=0
Well
+∞
-∞
Derivative, dpD/dlntD

m = 1/3

Fractures
2lw = 1000’ m = 1/4

0.1
20 Matrix
FD = 0.01
m = 1/4 FD = 0.1
FD = 1
FD = 10 m = 1/3
FD = 100
FD = 1000
6
FD = 10
m = 1/2
m = 1/4

-∞ +∞
0.01
-5 -4 -3 -2 -1 0 1 2
10 10 10 10 10 10 10 10
Dimensionless time, tD

Figure 25: Derivatives of a discretely fractured reservoir shown in


Fig. 24 for various finite conductivities without skin and storage
Figure 24: A schematic of a periodically and discretely fractured effects in an infinite reservoir, where one of the fractures intersects
reservoir. with the wellbore.

Figure 26 presents derivatives for various dimensionless fracture conductivities with CD = 0.001 and S = 0.
As shown in the figure, the derivatives for FD = 100 and 106 exhibit a 1/2-slope formation linear flow regime.
As expected, the wellbore storage does not have any noticeable effect on the derivative behavior, although it will if
FD  ∞ and tD < 10−5 . The derivative for FD = 10 exhibits a 1/2-slope period in the interval of 10−5 < tD < 10−4 ,
but this has nothing to do with the fracture or formation linear flow; it is simply a transition region from the unit
slope period to the formation linear flow regime. The derivatives for FD = 1, 10, 100, and 106 exhibit a 1/3-slope
trilinear flow regime at middle times. The derivatives for FD = 0.1 and 1 exhibit a radially composite look-alike
behavior. Finally, the derivative of FD = 0.01 looks like the behavior of an infinite-acting radial flow starting at
early times.
Figure 27 presents derivatives for various dimensionless fracture conductivities with CD = 0.001 and S = 5. This
is the same as the previous case, except that we have a skin value of 5. Adding skin has changed the derivative
behavior significantly, even for very large values FD , the largest of which is 106 . The 1/3-slope trilinear flow regime
at middle times of the derivatives for FD = 1, 10, 100, and 106 is not affected by the wellbore storage and skin.
The derivative for FD = 0.1 exhibit a radially composite look-alike behavior. As in the previous case, the derivative
of FD = 0.01 looks like the behavior of an infinite-acting radial flow starting at early times.
SPE 166296 15

2 10

1 6 FD = 0.01
5
8
4
FD = 0.1
6 m=0 FD = 1
Derivative, dpD/dlntD

Derivative, dpD/dlntD
3
4 FD = 10
2 FD = 100
2
m = 1/3 FD = 10
6

1
0.1 m=1 FD = 0.01
8 6 m=0
6 FD = 0.1 5
4
4 m = 1/2 FD = 1
FD = 10 3

m = 1/2 FD = 100 2 m = 1/3


2
6
FD = 10
0.01 0.1
-5 -4 -3 -2 -1 0 1 2 3 0.0001 0.001 0.01 0.1 1 10 100 1000
10 10 10 10 10 10 10 10 10
tD tD

Figure 26: Derivatives of a discretely fractured reservoir shown in Figure 27: Derivatives of a discretely fractured reservoir shown in
Fig. 24 for various finite conductivities with CD = 0.001 and S = 0 Fig. 24 for various finite conductivities with CD = 0.001 and S = 5
in an infinite reservoir, where one of the fractures intersects the in an infinite reservoir, where one of the fractures intersects with
wellbore. the wellbore.

Figure 28 presents derivatives for various dimensionless fracture conductivities with CD = 0.01 and S = 0.
For FD = 100 and 106 , a 1/2-slope formation linear flow regime is clearly apparent from the derivative plot given
in Fig. 28. As can be observed from this figure, the wellbore storage does not have any noticeable effect on the
derivative behavior of the high conductivity fractures. When we compare this figure with Fig. 26, the wellbore
storage of CD = 0.01 dominates much more the derivative behavior of low conductivity fractures than the wellbore
storage of CD = 0.001.

1
8
6 m=0 10
Derivative, dpD/dlntD

4
Δp and derivative, psi

2 m = 1/3

0.1 m=1 FD = 0.01


8
FD = 0.1
6
m = 1/2 FD = 1
4 FD = 10 1.0
FD = 100
2 6
FD = 10 m = 1/3
0.01 m=1
0.0001 0.001 0.01 0.1 1 10 100 1000
tD
0.1
0.01 0.1 1 10 100
Figure 28: Derivatives of a discretely fractured reservoir shown in Δt, hr
Fig. 24 for various finite conductivities with CD = 0.01 and S = 0
in an infinite reservoir, where one of the fractures intersects with Figure 29: Pressure derivative of a buildup test from a carbonate
the wellbore. reservoir, 3.3in from Djatmiko and Hansamuit (2010).

The derivative for FD = 1 shown in Fig. 28 looks like the derivative shown in Fig. 29 (Kuchuk and Biryukov,
2012). The buildup data given in this figure is presented by Djatmiko and Hansamuit (2010) from a carbonate
reservoir with vugs and karst.
After the early time flow regimes due to wellbore-intersecting fractures in discretely fractured reservoirs, the
formation matrix pseudo-radial (called simply radial) flow, as shown in Fig. 9 (f), may occur if the other fractures
are sufficiently far from the wellbore-intersecting fractures. However, most of the time we observe the effect of the
matrix because the derivative curve goes up without attaining the final slope of the formation matrix radial flow,
and then goes downward because of the effects of the near-wellbore fractures, which are non-intersecting. Fractures
are quite directional in most discretely fractured reservoirs. Therefore, we may observe the flow regimes shown in
Fig. 9: directional fracture linear (c), directional fracture-formation bilinear (d), and directional fracture formation
linear (e) flow regimes. Another formation matrix pseudo radial flow, shown in Fig. 9 (f), may occur if fractures
cease to exist farther away from the fractured zone.
Now we present another discretely fractured reservoir model with arbitrarily distributed discrete fractures with
various conductivities FD = 10, 000, 1, 000, 100, and 10, as shown in Fig. 30. In this model, many fractures,
particularly low conductivity fractures, intersect each other. All fractures are predominately in the east-west and
north-east directions. Here and there, the system fractures create local fracture networks, while the whole fracture
network is discontinuous. This is a very realistic model in the sense that in most naturally fractured reservoirs,
fractures are arbitrarily distributed with definite directions (usually not more than two), and different lengths and
SPE 166296 16

conductivities. In reality, fractures are created by geotectonic forces, and/or thermal processes. Therefore, the
arbitrary distribution does not imply a totally random distribution. When secondary and tertiary fracturing, and/or
possible pressure solution or pressure dissolution take place, the geology of fractured formations becomes complicated.
In this model, low conductivity secondary fractures are not orthogonal to the primary fractures, as shown in Fig. 30,
where a vertical well intersects one of the FD = 10, 000 fractures at the center of the system. Figure 31 presents

FD = 10 5 N 1
FD = 100 4

Derivative, dpD/dlntD
FD = 1000 3 m = 1/3
m=1
FD = 10
4 2 0.1
y coordinate

1
Well CD = 0, S = 0
0 CD = 0.0001, S = 0
0.01 m = 1/2 CD = 0.001, S = 0
-1
CD = 0.01, S = 0
-2 CD = 0.0001, S = 5

-3
0.001
-4 -5 -4 -3 -2 -1 0 1 2 3
10 10 10 10 10 10 10 10 10
-5 Dimensionless time, tD
-5 -4 -3 -2 -1 0 1 2 3 4 5
x coordinate E
Figure 31: Derivatives of a wellbore-intersecting fracture in an ar-
Figure 30: Wellbore-intersecting fracture in an arbitrarily dis- bitrarily distributed discrete fracture network model with different
tributed discrete fracture network model with different fracture fracture conductivities with skin and storage in an infinite reservoir
conductivities. (see Figure 30).

derivatives for the model described in the previous paragraph with various dimensionless wellbore storage and skin
values. Because of the high conductivity of the well-intersecting fracture, we observe a 1/2-slope formation linear
flow regime for all storage values, except for the CD = 0.0001 and S = 5 case. For this case, the 1/2-slope formation
linear flow was basically dominated by the wellbore storage and skin. All derivatives also exhibit a 1/3-slope trilinear
flow regime as a combination of directional flow in the fractures, and formation linear flow from the matrix. Some
field examples presented by Kuchuk and Biryukov (2012) also show the late-time 1/3-slope trilinear flow regime.
Finally, all derivatives reach the formation of radial flow because of a finite radius of the fractured region of the
reservoir.

Well in the Matrix


As shown in Fig. 4(d), fractures do not intersect wells in many discretely fractured reservoirs. Therefore, for this case
the first flow regime will be an infinite-acting radial flow regime without wellbore storage and skin effect. When the
well is in the formation matrix, the effects of wellbore storage and skin effect will be significant, particularly if the
matrix permeability is low. If fractures are far enough from the wellbore, we may still observe the first infinite-acting
radial regime with wellbore storage and skin effect, particularly with the downhole shut-in system. For these types
of discretely fractured reservoirs, pressure transient tests are a valuable tool to be able to observe the fracture nature
of the reservoir because wellbore image data will not show any fractures. Furthermore, these reservoirs are good
candidates for hydraulic fracturing to connect the wellbore to the fracture network. Pressure transient tests will also
provide information for hydraulic fracture design, particularly to determine the minimum length of the fracture to
intersect with natural fractures.
First let us look at the pressure transient behavior of a well in a reservoir with a single vertical infinite conductivity
fracture, where the well does not intersect the fracture. It is very unlikely for any reservoir to have a single fracture or
fault. However, in any discretely fractured reservoirs, where the well does not intersect any fractures, first we see the
effect of the closest fracture(s) to the wellbore without wellbore storage and skin effect. Therefore, it will be useful to
look at the transient behavior of a near-wellbore single fracture. There are more than a few papers published on the
subject: see Abbaszadeh et al. (2000); Cinco-Ley et al. (1976); Kuchuk and Habashy (1997); Matthai et al. (1998);
Tiab and Kumar (1980); Yaxley (1978). Here we show another look-alike behavior of the dual-porosity Warren and
Root (1963) model.
Figure 32 presents derivatives for a single wellbore-nonintersecting fully-penetrated infinite-conductivity fracture
with various distances from the wellbore to the fracture: d = 33, 131, and 328 ft, where the fracture half length is
328 ft from the x axis. As can be observed from Fig. 32, the derivatives without the wellbore storage and skin effects
(CD = 0, S = 0) show little deviation from the formation infinite-acting radial regime. With the exception of d =
33 ft, the deviations are too small to be observed in real field data. In fact, there is no observable deviation for d =
328-ft. However, if the fracture length becomes much larger than the distance, the effect will be observable. In other
words, if d ≈ 2lw , the effect of the single fracture will be very small. if d  2lw , the effect of the single fracture will
always be observable. For d = 33-ft, a moderate-magnitude wellbore storage (CD = 0.001) distorts the derivative so
SPE 166296 17

much that its behavior becomes a look-alike behavior of the dual-porosity Warren and Root (1963) model.
Now let us look at the pressure transient behavior of a
4 discretely and periodically fractured reservoir. This is a simple
d = 0 ft, wellbore intersecting
d = 33 ft
d = 131 ft
2 d = 328 ft
d = 33 ft & CD= 0.001
2D reservoir, where we uniformly distribute a 32.8-ft (10-m)
Derivative, dpD/dlntD

1
8
m = 1 fracture in every 65.6 ft (20 m) in the x and x directions. The
6
4
fractures are lined up in in the east-west direction, and the
2
distance between two adjacent fractures is 65.6 ft, as shown
0.1
in Fig. 33 (Biryukov and Kuchuk, 2012). In this system, the
8
6 well is located in the matrix formation in the center and about
4 m = 1/2
46 ft away farm a fracture, i.e. it does not intersect fractures As
2
shown in Fig. 33, the matrix initially dominates the behavior
0.01
0.001 0.01 0.1 1 10
of the system, after which the derivatives for different fracture
Dimensionless time, tD conductivities go downward gradually, become flat, and then
go up slightly. Although the effect of periodic small fractures
Figure 32: Derivatives of an infinite conductivity vertical frac- is observable but it is small. The early time radial flow regime
ture at various distances to the wellbore. will yield the matrix formation parameters. These derivatives,
as shown in Fig. 34, behave like the derivatives of the radially composite reservoir. The late time radial flow regime
will yield some sort of weighted average of the formation and fracture permeabilities. This is not a simple averaging.
For instance, we do not observe any difference between the derivative values of FD = 10, 100, and, 106 . its behavior
becomes a look-alike behavior of the dual-porosity Warren and Root (1963) model.

0.54
6
FD = 10
FD = 100
FD = 1
20 m 0.52 FD = 10
Fracture
10 m
Derivative, dpD/dlntD
FD = 0.1
FD = 0.01

Matrix 0.5

Well
0.48

0.46

0.44
2 3 4 5 6 7 8 9 2 3 4 5 6 7 8 9
0.001 0.01 0.1
tD

Figure 34: Pressure derivatives for a periodically fractured reser-


Figure 33: A schematic of a periodically fractured reservoir with voir with short directional fractures with different conductivities
short directional fractures. (Fig. 33).

This model was first presented by Kuchuk and Habashy (1997) for an infinite discretely fractured reservoir. Its
details were given previously and in Fig. 24. A fully penetrating vertical well is located in the formation between
the 20th and 21st fractures, as shown in Fig. 24, and its distances to these fractures are 5 ft and 15 ft, respectively.
All fracture lengths are the same and the fracture half length is lw = 500 ft from the x axis, as shown in Fig. 24.
Figure 35 presents derivatives for this discretely fractured reservoir, where the well is in the matrix with various
dimensionless fracture conductivities without wellbore storage and skin. As can be seen from Fig. 35, the derivative
of FD = 0.01 is the same as the matrix derivative, i.e., FD = 0.01 does not affect the derivative behavior of the
system. The derivatives for FD = 0.1 and 1 exhibit a radially composite look-alike behavior. The derivatives for
FD = 10, 100, and 1000 exhibit a 1/2-slope linear flow regime, and then a 1/3-slope trilinear flow regime at middle
times. All derivatives reach the formation radial flow regime at late times.
Figure 36 presents derivatives for the same discretely fractured reservoir, as shown in Fig. 24 with various
dimensionless fracture conductivities with CD = 0.0001 and S = 5, where the well is in the matrix. When we
compare the derivatives shown in this figure with the derivative shown in Fig. 35, the early time behavior is dominated
by skin and storage effects, shown as a unit (m=1) slope period. The early radial flow regime disappears due to
formation matrix, as can be observed in Fig. 36. The derivative of FD = 0.01 exhibits a typical homogenous reservoir
behavior with skin and storage effects. As shown in Fig. 35 (without skin and storage effects), the derivatives for
FD = 0.1 and 1 exhibit a radially composite look-alike behavior, and the derivatives for FD = 10, 100, and 1000
exhibit a 1/2-slope linear flow regime, and then a 1/4-slope bilinear flow regime at middle times. When we compare
the derivatives shown in Fig. 36 with those in Fig. 27—in the latter case the wellbore intersects one of the fractures
but the model is the same (Fig. 24)—we see that they exhibit different flow regimes at middle times. It should be
noticed that the fractured nature of the system is still apparent. When we increase CD from 0.0001 to 0.001 with
the same S = 5, as shown in Fig. 37, the 1/2-slope linear flow regime disappears, while the 1/3-slope trilinear flow
SPE 166296 18

10
FD = 0.01
1 FD = 0.1
FD = 1
FD = 10
m=0 m=0 FD = 100
FD = 1000

Derivative, dpD/dlntD
1
Derivative, dpD/dlntD

m = 1/3 m=0

m=1 m = 1/3
0.1
0.1 m = 1/2
FD = 0.01 m = 1/3
m = 1/2 FD = 0.1
FD = 1 m = 1/3
FD = 10
FD = 100
FD = 1000

0.01
-5 -4 -3 -2 -1 0 1 2
0.01 10 10 10 10 10 10 10 10
10
-5 -4
10
-3
10
-2
10
-1
10 10
0
10
1
10
2 Dimensionless time, tD
Dimensionless time, tD

Figure 36: Derivatives of a discretely fractured reservoir shown in


Figure 35: Derivatives of a discretely fractured reservoir shown in Fig. 24, for various finite conductivities with the skin and storage
Fig. 24, for various finite conductivities without skin and storage effects (CD = 0.0001 and S = 5) in an infinite reservoir, where the
effects in an infinite reservoir, where the well is in the matrix. well is in the matrix.

regime is still apparent.

10 10
FD = 0.01
CD = 0, S = 0
FD = 0.1
CD = 0.0001 & S = 0
FD = 1
CD = 0.0001 & S = 5
FD = 10
CD = 0.0001 & S = 10
FD = 100
CD = 0.001 & S = 0
FD = 1000
CD = 0.001 & S = 5
m=0
Derivative, dpD/dlntD

Derivative, dpD/dlntD

CD = 0.001 & S = 10
1

m=0
0.1 m = 1/3
m = 1/3
m=1 m = 1/3 m = 1/2

0.1 0.01
0.0001 0.001 0.01 0.1 1 10 100 10
-5 -4
10
-3
10 10
-2 -1
10 10
0 1
10
2
10
Dimensionless time, tD Dimensionless time, tD

Figure 37: Derivatives of a discretely fractured reservoir shown in Figure 38: Derivatives of an arbitrarily distributed discrete fracture
Fig. 24 for various finite conductivities with the skin and storage network model with different fracture conductivities, where the well
effects (CD = 0.001 and S = 5) in an infinite reservoir, where the is in the matrix with skin and storage in an infinite reservoir (see
well is in the matrix. Figure 30).

Figure 38 presents derivatives for the model described in Fig. 30. This is another discretely fractured reservoir
model with arbitrarily distributed discrete fractures with various conductivities FD = 10, 100, 1, 000, and 10, 000.
All fractures are predominately in the east-west and north-east directions. In this model, many fractures, particularly
low conductivity fractures, intersect each other. In this case, the well is in the matrix, and we look at its pressure
behavior with various wellbore storage and skin effects. Because the well is in the matrix, we first observe the
formation matrix radial flow, as shown in Fig. 38, for the CD = 0 and S = 0 case. After this flow regime, we observe
a brief 1/2-slope formation linear flow regime, and then a 1/3-slope trilinear flow regime. Due to wellbore storage
and skin effects, all other derivatives exhibit first a unit slope period, and then a wellbore storage and skin transition
period. After the wellbore storage and skin dominated flow periods, the derivatives exhibit a brief 1/2-slope formation
linear flow regime and then a 1/3-slope trilinear flow regime. Finally, all derivatives reach the formation radial flow
because of the finite radius of the fractured region of the reservoir.
The model shown in Fig. 39 was presented by Morton et al. (2013) for a discretely fractured reservoir. The
fractured reservoir model presented by 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 concentration, 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. 39, and all fractures are predominately in the east-west direction, extending about 1640 ft (500 m) from the well.
The model is one of the realizations of the Morton et al. (2013) discretely fractured reservoir (selected derivatives
were presented by Morton et al. (2013) for different realizations). Figure 40 presents derivatives for a number of
realizations of the model. As can be seen from this figure, the derivatives for different cases exhibit many different
SPE 166296 19

look-alike 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 the Warren and Root (1963) type models. Morton
et al. (2013) showed that matching pressures and derivatives from the model shown in Fig. 39 to the Warren and
Root (1963) models yields unphysical fractured reservoir parameters.

600 N

400
1

200

Derivative, dpD/dlntD
m=0 m=0
y axis, m

0 well

-200

-400

-600 0.1
-600 -400 -200 0 200 400 600
0.0001 0.001 0.01 0.1 1 10 100 1000
x axis, m E
Dimensionless time, tD
Figure 39: Directionally but arbitrarily distributed fractures differ-
ent conductivities in a discretely fractured reservoir, where the well
is located in the matrix at the center of the coordinate system Figure 40: Derivatives for a number of realizations of the discretely
x = 0 and y = 0, after Morton et al. (2013). fractured reservoir model shown in Fig. 39.

History Matching with the Warren and Root (1963) Dual-Porosity Type Models
As shown above, much of the behavior of many continuously and discretely fractured reservoirs creates the look-
alike behaviors of the Warren and Root (1963) type models. This is particularly true for fracture-intersecting wells
with wellbore storage and skin effects, and for wells completed in the matrix. In this section we history match two
synthetic well test data [generated with Biryukov and Kuchuk (2012) model] with the Warren and Root (1963) and
Kazemi (1969) “equivalent” dual-porosity models. The objective of presenting the history (type-curve) matching
of these two pressure transient test examples is to show how incorrect it would be to rely on history matching for
obtaining fractured reservoir parameters from the “equivalent” homogenous dual-porosity models.

Synthetic Example 1
This geological model was presented by Kuchuk and Habashy (1997) for an infinite discretely fractured reservoir; its
details were given previously and in Fig. 24. A fully penetrating vertical well is located in the formation between
the 20th and 21st fractures, as shown in Fig. 24, and its distances to these fractures are 5 ft and 15 ft, respectively.
All fracture lengths are the same, and the fracture half length is lw = 500 ft from the x axis. All fractures in the
model have the same conductivity: FD = 1000. We assume that the damage skin is 5 and the dimensionless wellbore
storage is CD = 0.0001. The input model parameters given in Table 1 are used with the Biryukov and Kuchuk (2012)
model to generate the wellbore pressure data.

Table 1: Formation and fluid properties for Synthetic Example 1.

φm fraction 0.2 µ cp 1.0 (ct )m psi−1 1.5×10−5 q B/D 500


km mD 1 kf mD 7.62 × 107 h ft 100 rw ft 0.354
S 5 CD 0.0001 FD 1000 lw ft 500
b mm 2 (0.0066 ft)
where b denotes the fracture aperture.

Figure 41 shows the history matches of pressure change and derivative of the example data (denoted as measured)
with the five different Warren and Root (1963) type dual-porosity pressure transient solution with different matrix
shapes and flow conditions. As can be observed from Fig. 41, both the pressure change and its derivative (denoted
as measured) are matched very well with the five different dual-porosity models. The estimated values of model
SPE 166296 20

parameters for each model are shown in Table 2. As can be seen from Fig. 41, all the models match the measured
(synthetic) data very well, but none of the estimated parameters given in Table 2 relates to any of the parameters of
the discretely fractured reservoir described above with its parameters given in Table 1. First, the actual model is an
infinite reservoir, but all five dual-porosity models are bounded by either two parallel or two perpendicular no-flow
boundaries. As can be seen in Table 2, the estimated fracture system permeability varies from 5 to 7.5 mD, while
the input fracture permeability kf is 7.62 × 107 mD (Table 1). The input skin S is 5, and the estimated skin values
vary from 37 to 65. The ω values vary from 0.01 to 0.32, and the λ values vary from 1.00 × 10−6 to 1.00 × 10−5 .
The characteristic length of a matrix block given
10
4 Δp pressure from the definition of the interporosity flow coefficient,
λ, can be written (Eq. 1-10 of the Warren and Root
Δp measured (1963) paper) as
der. measured
Δp & Derivative, psi

Δp model 1
der. model 1 s
3 Δp model 2
10 der. model 2
2
km rw
Δp model 3
der. model 3
` = 4.78 , n = 1, 2, 3, (2)
Δp model 4 λkf
der. model 4
Δp model 5
der. model 5
2 where kf and km are the fracture and matrix perme-
10
abilities, respectively, and rw is the wellbore radius.
Using the rw , km , and kf values given in Table 1, and λ
derivative = 1.00×10−6 (one of the λ values given in Table 2), the
1
10 estimated characteristic length ` of the matrix block
0.01 0.1 1 10 100 is 0.19 ft. This value of the characteristic length does
Time, hr not make any sense. The matrix block characteristic
length formula is given in Eq. 1-13 of the Warren and
Figure 41: Comparison of the pressure changes and derivatives (denoted Root (1963) paper as
as model) from the history match of various dual-porosity Warren and
Root (1963) type models to the pressure change and derivative data 3abc
generated with the (Biryukov and Kuchuk, 2012) model (denoted as `= , n = 3, (3)
measured). ab + bc + ac
where a, b, and c are the dimensions of the matrix
blocks. For any geometrical shape with a, b, c > 1 ft,
the lower limit for ` is 1: this would be a 1 ft by 1 ft by 1 ft cube. Basically, ` is the length of the sides of any cubic
matrix block. For any rectangular parallelepiped matrix blocks, `  1 if a, b, c > 1 ft.

Table 2: Nonlinear regression analysis results for different dual-porosity models of Synthetic Example 1.

Parameters Model 1 Model 2 Model 3 Model 4 Model 5


k, mD 5 5 7.5 7.5 7.5
C, B/psi 0.00145 0.00145 0.00145 0.00145 0.00145
S 40 37 65 65 65
matrix PSS† PSS PSS slab TR§ sphere TR
ω 0.17 0.23 0.08 0.32 0.01
λ 1.35E-05 1.30E-05 1.20E-05 1.00E-06 1.00E-05
Lwn , ft NA 270 ∞ NA ∞
Lws , ft 280 270 160 155 155
Lwe , ft 280 ∞ 270 270 270
Lww , ft -∞ -∞ -∞ -∞ -∞
where † pseudo-steadystate matrix flow, § transient matrix flow,
Lwn , Lws , Lwe , and Lww denote the distance from the wellbore to North, South, East,
and West boundaries, respectively.

Synthetic Example 2
The previous example is for an infinite discretely fractured reservoir (see Fig. 24). Although its derivative exhibits
a Warren and Root (1963) dual-porosity type-model look-alike behavior, the history matching of its behavior with
Warren and Root (1963) dual-porosity type models yields incorrect estimates of the reservoir parameters. The next
example is a continuously fractured system, as shown in Fig. 15, where fractures are in a network and communicate
with each other throughout the system. It is a Warren and Root (1963) dual-porosity geological model, where
65.617 ft by 65.617 ft by h (20 m by 20 m by h) rectangular parallelepiped matrix blocks (red bricks in the figure)
are uniformly distributed in the formation. The fractures (white space between the bricks in Fig. 15) are orthogonal
and uniform, with a length 2lw = 65.617 ft. The well intersects one of the fractures at the center of the fracture
SPE 166296 21

network, as shown in Fig. 15. All fractures in the model have the same conductivity: FD = 100. It is assumed that
the damage skin is 0 and the dimensionless wellbore storage is CD = 0.001. The fracture network is finite in the
reservoir with a drainage radius re of 1640 ft. Again, the Biryukov and Kuchuk (2012) fractured reservoir model is
used to generate the wellbore pressure data with the input model parameters that are given in Table 3.
Table 3: Formation and fluid properties for Synthetic Example 2.

φm fraction 0.2 (ct )m psi−1 1.5×10−5 µ cp 1.0 q B/D 500


km mD 1 kf mD 107 h ft 100 rw ft 0.354
S 0 CD 0.001 FD 100 re ft 1640
lw ft 32.80 b mm 2 (0.0066 ft)
where re denotes drainage radius.

Figure 42 shows the history matches of the pressure change


108 and derivative of the second example data (denoted as mea-
6 Δp pressure sured) with the six different dual-porosity Warren and Root
4
(1963) models with different matrix shapes and flow con-
Δp & Derivative, psi

2
ditions. As can be observed from Fig. 42, both the pres-
18 Δp measured sure change and its derivative (denoted as measured) are
6 der. measured
4 derivative
Δp model 1
der. model 1
matched very well with the six different dual-porosity mod-
2
Δp model 2
der. model 2
els (denoted as model). The estimated values of the model
Δp model 3
der. model 3 parameters for each model are as shown in Table 4. As can
0.18 Δp model 4
6
der. model 4
Δp model 5
be seen from Fig. 42, all the models match the measured
4 der. model 5
Δp model 6 data very well, but none of the estimated parameters given
der. model 6
2 in Table 4 relates to the parameters of the Warren and Root
0.01 (1963) dual-porosity continuously-fractured geological model
0.001 0.01 0.1 1 10 100 described above, with the parameters given in Table 3. The
Time, hr actual model is a no-flow bounded circular reservoir, but all
six dual-porosity models are bounded rectangular reservoirs.
Figure 42: Comparison of the pressure changes and derivatives The estimated distances from the wellbore to the no-flow
from the history match of various dual-porosity Warren and Root boundaries vary from 200 to 2600 ft, while the radius of the
(1963) models to the pressure change and derivative data gen- actual model is r = 1640 ft. The input fracture permeability
e
erated with the (Biryukov and Kuchuk, 2012) model.
kf is 107 mD, and the estimated kf values vary from 1000 to
2500 mD. The input skin S is 0 and the estimated skin values vary from -2.7 to 3. The ω values vary from 0.2 to
0.45, and the λ values vary from 7.00×10−7 to 10−6 . Again using the rw , km , and kf values given in Table 3, and
λ = 1.00×10−6 (one of the λ values given in Table 4), the estimated characteristic length ` of the matrix block is
0.54 ft from Eq. 2. This value does not make any sense. For rectangular parallelepiped matrix blocks of this model
with the dimensions of 65.617 ft by 65.617 ft by 100 ft, the estimated characteristic length ` from Eq. 3 would be 74 ft.
In this example, we use Warren and Root (1963) dual-porosity geological model in the Biryukov and Kuchuk (2012)
fractured reservoir solution to generate the well test data. Again, we obtain a Warren and Root (1963) dual-porosity
type-model look-alike derivative behavior, and its history matching with Warren and Root (1963) dual-porosity type
models yields incorrect estimates of the reservoir parameters with an unphysical matrix-block-size estimate.
Table 4: Nonlinear regression analysis results for different dual-porosity models of Synthetic Example 2.

Parameters Model 1 Model 2 Model 3 Model 4 Model 5 Model 6


k, mD 1350 1350 1350 1000 2500 1900
C, B/psi 0.14 0.15 0.15 0.13 0.16 0.16
S -1.1 -1.1 -1.1 -2.7 3 1.24
matrix PSS† slab TR§ sphere TR slab TR PSS NA
ω 0.45 0.32 0.2 0.32 0.45 NA
λ 7.00E-07 1.00E-06 1.00E-06 1.00E-06 7.00E-07 NA
Lwn , ft 1100 1100 1100 1300 1600 1400
Lws , ft 1100 1100 1100 1200 200 660
Lwe , ft 1100 1100 1100 1200 1300 1150
Lww , ft 2100 2100 2100 1600 2600 2300
where † pseudo-steadystate matrix flow, § transient matrix flow,
Lwn , Lws , Lwe , and Lww denote the distance from the wellbore to North, South, East,
and West boundaries, respectively.
SPE 166296 22

Discussion of Results
The main objective of presenting derivatives from various types of continuously and discretely fractured reservoirs
with different fracture and wellbore configurations, and with skin and wellbore storage effects, is to make geoscientists
and reservoir engineers more aware of the pressure transient behavior of these systems for reservoir characterization.
This is necessary because otherwise we would keep using the “equivalent” homogenous medium Warren and Root
(1963) type models for the characterization of fractured reservoirs without regard to different fracture and wellbore
configurations, or many geological processes. This practice has been going on since the 1960s.
This study on the pressure transient behavior of fractured reservoirs is hardly the last word on the subject.
There remain many unsolved problems, for instance how to model and upscale thousands, if not tens of thousands,
fractures without losing the characteristics of the pressure transient signature of the system and for multiphase flow.
The Warren and Root (1963) dual-porosity type models without fractures are not the solution. The new solutions can
now handle about a thousand fractures with many different fracture conductivities, densities, lengths, orientations,
etc., and wellbore configurations, such as intersecting or non-intersecting with fractures. This is a significant change
and an improvement over the Warren and Root (1963) dual-porosity type models. We anticipate that the new
solutions will eventually be extended to handle several thousand of fractures.
Most derivatives presented here include wellbore storage and skin effects. These effects hinder well test system
identification and interpretation for fractured reservoirs, and for any reservoir in general. Wellbore storage obstructs
identification and interpretation severely if the wellbore-intersecting fractures are damaged or some of the perforations
that connect the wellbore to the fractures are not active, the fractures have low conductivities, or the well is in the
matrix—a non-fracture intersecting well. To minimize wellbore storage and skin effects on pressure transient behavior,
it is essential to use downhole shut-in tools for better characterization of fractured reservoirs.

Conclusions
In this paper we investigated the pressure transient behavior of continuously and discretely fractured reservoirs
using semi-analytical solutions for a vertical well. These fractured reservoirs can contain periodically or arbitrarily
distributed finite- and/or infinite-conductivity fractures with different lengths and orientations. The wells in these
fractured systems can intersect one or more fractures, and can be located in the matrix blocks with wellbore storage
and skin effects. It is shown that if a well intersects a fracture, then it dominates the pressure transient behavior of
both continuously and discretely fractured reservoirs. If the well is in the matrix, we observe a matrix-block radial
flow regime until the effects of the fractures are felt in the wellbore. After the first radial flow, the derivatives go
downward rapidly at early times until they reach a minimum, depending on conductivities.
It is shown that the low dimensionless conductivities of FD = 0.001 and 0.01 do not really affect the transient
behavior (derivatives) of both continuously and discretely fractured reservoirs. The derivatives of low and medium
conductivity fractures intersecting the wellbore with dip angles less than 90◦ may exhibit a fractured radial flow
first, followed by a formation linear flow. Low and medium conductivity fractures amplify the wellbore storage
and skin effects. The conductivities of FD = 1, 10, 100, 1, 000, 10, 000, and ∞ dominate transient behavior
(derivatives) of both continuously and discretely fractured reservoirs. The derivatives for high conductivities (FD =
1, 000, 10, 000, and ∞) tend to exhibit a 1/2-slope formation linear flow regime. Medium and high conductivity
fractures with short lengths in discretely fractured reservoirs exhibit flow regimes that show linearly and radially
composite reservoir look-alike derivative behaviors. The derivatives for high fracture conductivities, except for
FD = ∞, exhibit a pseudosteady-state flow regime (m = 1) at early times due to rapid depletion of the fracture
network in continuously fractured reservoirs if the matrix permeability is low. This flow regime usually takes place
before the formation linear flow.
It is shown that the derivatives of pressure transient tests from fractured reservoirs exhibit more than 10 flow
regimes, depending on whether the well intersects fractures or is in the matrix; whether the fractures are fully or
partially penetrating; the conductivity, size, distribution, and orientation of the fractures; and finally, the wellbore
storage and skin effects.
If we history (type-curve) match pressure transient tests from continuously and discretely fractured reservoirs to
the Warren and Root (1963) “equivalent” dual-porosity type models, the resulting reservoir model and its estimated
parameters would be incorrect with an unphysical matrix-block-size estimate.

Nomenclature
C = wellbore storage constant, RB/psi [m3 /Pa]
ct = compressibility
F = fracture conductivity
h = formation thickness
k = permeability
l = fracture half-length
p = pressure
q = flow rate
SPE 166296 23

r = radius or radial coordinate


S = Skin factor
t = time
x = coordinate
y = coordinate
z = vertical coordinate
η = diffusivity for pressure
µ = viscosity
φ = porosity
Subscripts
D = dimensionless
f = fracture
o = initial or original
m = matrix
w = wellbore

Acknowledgments
The authors are grateful to Schlumberger for permission to publish this paper, and would like to thank Kirsty Morton
and Amine Ennaifer for their contributions for understanding pressure transient tests from naturally fractured reservoirs,
particularly for Amine who performed the history matching of two examples with the Warren and Root (1963) type models.

References
Abbaszadeh, M., Asakawa, K., Cinco-Ley, H., and Arihara, N. 2000. Interference testing in reservoirs with conductive faults
or fractures. SPE Reservoir Evaluation & Engineering, 3(5):426–434.

Ayestaran, L., Nurmi, R., Shehab, G., and El Sisi, W. 1989. Well test design and final interpretation improved by integrated
well testing and geological efforts. Paper SPE 17945, Middle East Oil Show, Bahrain, 11–14 March.

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.

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.

Cinco-Ley, H. 1996. Well–test analysis for naturally fractured reservoirs. Journal of Petroleum Technology, 48(1):51–54.

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. Journal of Petroleum Technology,
33(9):1749–1766.

Cinco-Ley, H., Samaniego-V, F., and Dominguez, N. 1976. Unsteady state behavior for a well near a natural fracture. Paper
SPE 6019, SPE Annual Technical Conference and Exhibition, New Orleans, 3–6 October.

Committee on Fracture Characterization and Fluid Flow 1996. Rock Fractures and Fluid Flow:Contemporary Understanding
and Applications. The National Academies Press.

Corre, B. 1990. Characterization of fracture networks from well tests using a new analytical solution. Paper SPE 20533, SPE
Annual Technical Conference and Exhibition, New Orleans, Louisiana, 23–26 September.

Djatmiko, W. and Hansamuit, V. 2010. Well test analysis of multiple matrix-to-fracture fluid transfer in fractured-vuggy
reservoir. Paper SPE 130557, International Oil and Gas Conference and Exhibition in China, Beijing, China, 8–10 June.

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.

Gringarten, A. 1984. Interpretation of tests in fissured and multilayered reservoirs with double-porosity behavior: Theory and
practice. Journal of Petroleum Technology, 36(4):549–564.

Guo, B. and Schechter, D. 1998. Use of single-well test data for estimating permeability anisotropy of the naturally fractured
Spraberrytrend area reservoirs. Paper SPE 39807, SPE Permian Basin Oil and Gas Recovery Conference, Midland, Texas,
25–27 March.
SPE 166296 24

Igbokoyi, A. and Tiab, D. 2008. Pressure transient analysis in partially penetrating infinite conductivity hydraulic fractures
in naturally fractured reservoirs. Paper SPE 116733, SPE Annual Technical Conference and Exhibition, Denver, Colorado.

Kazemi, H. 1969. Pressure transient analysis of naturally fractured reservoirs with uniform fracture distribution. SPE Journal,
9(4):451–462.

Kikani, J. and Walkup, G. W. 1991. Analysis of pressure-transient tests for composite naturally fractured reservoirs. SPE
Formation Evaluation, 6(2):176–182.

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 2012.

Kuchuk, F. and Habashy, T. 1997. Pressure behavior of laterally composite reservoirs. SPE Formation Evaluation, 12(1):47–56.

Kuchuk, F., Onur, M., and Hollaender, F. 2010. Pressure Transient Formation and Well Testing: Convolution, Deconvolution
and Nonlinear Estimation. Elsevier, New York.

Lee, S.-T. and Brockenbrough, J. R. 1986. A new approximate analytic solution for finite-conductivity vertical fractures. SPE
Formation Evaluation, 1(1):75–88.

Matthai, S. K., Aydin, A., Pollard, D. D., and Roberts, S. 1998. Simulation of transient well-test signatures for geologically
realistic faults in sandstone reservoirs. SPE Journal, 3(1):62–76.

Morton, K., Booth, R., Chugunov, N., Fitzpatrick, A., and Kuchuk, F. 2013. Global sensitivity analysis for natural frac-
ture geological modeling parameters from pressure transient tests. Paper SPE 164894, SPE EUROPEC/EAGE Annual
Conference and Exhibition, 10–13 June, London, United Kingdom.

Nelson, R. A. 1985. Geologic Analysis of Naturally Fractured Reservoirs. Gulf Publishing, Houston, Texas.

Odling, N. E., Gillespie, P., Bourgine, B., and Castaing, C. 1999. Variations in fracture system geometry and their implications
for fluid flow in fractured hydrocarbon reservoirs. Petroleum Geoscience, 5(4):373–384.

Pollard, D. D. and A., A. 1988. Progress in understanding jointing over the past century. Geological Society of America
Bulletin, 100(8):1181–1204.

Raghavan, R. 1976. Some practical considerations in the analysis of pressure data. Journal of Petroleum Technology,
28(10):1256–1268.

Raghavan, R. 1993. Well Test Analysis. Prentice Hall, Boston.

Raghavan, R., Uraiet, A., and Thomas, G. W. 1978. Vertical fracture height: Effect on transient flow behavior. SPE Journal,
18(4):265–277.

Rodriguez, F., Horne, R., and Cinco-Ley, H. 1984. Partially penetrating fractures: Pressure transient analysis of an infinite
conductivity fracture. Paper SPE 12743, SPE California Regional Meeting, Long Beach, California, 11–13 April.

Singh, S., Abu-Habbiel, H., Khan, B., Akbar, M., Etchecopar, A., and Montaron, B. 2008. Mapping fracture corridors in
naturally fractured reservoirs: an example from Middle East carbonates. Firstbreak, 26(5):109–113.

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, Dallas, 1st edition.

Souche, L. and Rotsch, M. 2007. An end-to-end approach to naturally fractured reservoir modeling: Workflow and implemen-
tation. SEG/EAGE Research Workshop 2007, Perugia, Italy, September 3-6.

Tiab, D. and Kumar, A. 1980. Detection and location of two parallel sealing faults around a well. Journal of Petroleum
Technology, 32(10):1701–1708.

Warren, J. E. and Root, P. J. 1963. The behavior of naturally fractured reservoirs. SPE Journal, 3(3):245–255.

Yaxley, L. M. 1978. Effect of a partially communicating fault on transient pressure behaviour. SPE Formation Evaluation,
2(6):590–598.

Zeybek, M., Kuchuk, F., and Hafez, H. 2002. Fault and fracture characterization using 3d interval pressure transient tests.
Paper SPE 78506, Abu Dhabi International Petroleum Exhibition and Conference, Abu Dhabi, UAE, 13–16 October.

Das könnte Ihnen auch gefallen