Extending Predictive Capabilities To Network Models

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Extending Predictive Capabilities to

Network Models
P.-E. Øren, SPE, Stig Bakke, and O.J. Arntzen, Statoil A/S

Summary However, the difficulty in adequately describing the pore network


We reconstruct three-dimensional (3D) sandstone models that give of reservoir rocks has prevented network models from being used
a realistic description of the complex pore space observed in actual as a predictive tool, thus greatly limiting their application in the oil
sandstones. The reconstructed pore space is transformed into a pore industry.
network that is used as input to a two-phase network model. The In the present work, geostatistical information obtained from
model simulates primary drainage and water injection on the basis image analyses of 2D thin-sections are used to generate a reliable
of a physical scenario for wettability changes at the pore level. We reconstruction of the complex rock/pore system in three dimen-
derive general relationships among pore structure, wettability, and sions. The network representation of the pore space is constructed
capillary pressure for the different pore level displacement mech- from topological and geometrical analyses of the fully character-
anisms that may occur in the network model. ized reconstructed sample. The pore network is used as input to a
We present predicted transport properties for three different two-phase network model that simulates drainage and water injec-
reconstructed sandstones of increasing complexity: Fontainebleau, tion based on a physical model for wettability alteration on the pore
a water-wet Bentheimer, and a mixed-wet reservoir rock. Predicted level. Predicted transport properties for different reconstructed
transport properties are in good agreement with available experi- sandstones are compared with experimental data.
mental data. For the reservoir rock, both the experiments and the
simulated results show that continuous oil films allow low oil Experimental
saturations to be reached during forced water injection. However,
the oil relative permeability is very low. The experimental measurements reported here were performed on
outcrop and on reservoir sandstones. The outcrop rock consisted of
three homogenous core samples that were cut from a single block
Introduction
of Bentheimer quarried sandstone. This strongly water-wet sand-
The microstructure of a porous medium and the physical charac- stone is well-sorted and composed mainly of quartz (70 to 80%),
teristics of the solid and the fluids that occupy the pore space feldspar (20 to 25%), and authigenic clays (2 to 3%). The main clay
determine several macroscopic properties of the medium. These component is pore-filling kaolinite. The reservoir rock consisted of
properties include transport properties of interest, such as perme- two preserved mixed-wet core plugs. This North Sea Lower Brent
ability, formation factor, relative permeability, and capillary pres- reservoir rock is a micaceous (10 to 12%) feltspatic (15 to 19%)
sure. In principle, it should be possible to determine these properties sandstone with abundant pore-filling kaolinite. Table 1 gives
by appropriately averaging the equations describing the physical petrophysical properties for the different plug samples.
processes occurring on the pore scale. The prediction of macro- The Bentheimer plugs were tested at room temperature with
scopic transport properties from the associated pore-scale param- synthetic brine (3% NaCl) and a paraffinic oil. Measurements on
eters is a long-standing issue that has been the subject of much the reservoir core plugs were performed at 70°C with simulated
investigation. One commonly applied tool in this investigation is formation brine and a dead crude oil. Table 2 summarizes fluid
the network model. properties at the specific test conditions.
The premise of the network model is that the void space of a
porous medium can be represented by a network of interconnected Capillary Pressure. Primary drainage capillary pressure functions
pores in which larger pores (pore bodies) are connected by smaller for the Bentheimer core plugs were measured by the porous plate
pores (pore throats). Since the pioneering work of Fatt,1-3 network method. A ceramic porous plate with a gas/water threshold pressure
models have been used extensively to study different displacement of 15 bar was installed on the lower endface of the vertically
processes in simple or idealized porous media.4-20 Seldom, how- oriented sample. Water production was measured at nine different
ever, do such models claim to be representative of reservoir rocks.21 drainage pressures in the range 0 to 5 bar. Table 3 summarizes the
The extension of network modeling techniques to real porous measured capillary pressure data, which were similar for all three
media is hampered by the difficulty of adequately describing the samples.
complex nature of the pore space. Advanced techniques, such as Waterflood and secondary drainage capillary pressure functions
microtomographic imaging21-25 and serial sectioning,26-27 provide for the reservoir core plugs were determined from centrifuge
a detailed description of the pore space at micrometer resolution. displacements. Fluid saturations were calculated from material
In practice, however, information about the microstructure of balance and Karl Fischer analysis at the end of the experiments.
reservoir rocks is limited to two-dimensional (2D) thin-section Table 4 tabulates the capillary pressure data, which were similar for
images and to pore throat entry sizes determined from mercury the two samples. Spontaneous imbibition of water (oil) was mea-
injection. These data are insufficient to construct a 3D pore network sured before the waterflood (secondary drainage) and used to
directly that replicates the microstructure of the medium. As a calculate Amott wettability indices. Both core plugs spontaneously
result, simplifying assumptions about the topology and geometry of imbibed both water and oil and the Amott wettability indices (0.08
the pore structure must be invoked. and 0.1) indicate that the samples are mixed-wet.
Despite these simplifications, network models have proved to be
powerful tools for extrapolating limited measured data and for
developing valuable insight into complex multiphase flow phe- Relative Permeability. Primary drainage and waterflood relative
nomena, such as capillary pressure and relative permeability hys- permeability curves for the Bentheimer plugs were determined by
teresis,12, 28 the effect of wettability,29-32 and three-phase flow.33-38 the steady-state method. The measurements were performed on
vertically oriented samples. The total flow rate (200 cm3/h) was
constant during the experiments. Steady-state relative permeability
Copyright 1998 Society of Petroleum Engineers measurements were obtained by keeping the oil and water flow
This paper (SPE 52052) was revised for publication from paper SPE 38880, first rates constant until the conditions in the core stabilized. The
presented at the 1997 SPE Annual Technical Conference and Exhibition held in San pressure drop across the core was monitored continuously with a
Antonio, Texas, 5–8 October. Original manuscript received for review 21 October
1997. Revised manuscript received 22 June 1998. Revised manuscript approved 2 differential pressure transducer. Relative permeability to each
July 1998. phase was calculated by Darcy’s law.

324 SPE Journal, December 1998


TABLE 1—PETROPHYSICAL PROPERTIES FOR TABLE 4—CAPILLARY PRESSURE DATA FOR THE
BENTHEIMER AND RESERVOIR ROCK SAMPLES RESERVOIR ROCK SAMPLES

r kw Forced Water Injection Secondary Drainage


Sample* Porosity (g/cm3) (md) FRF
Pc Sw Sw Pc Sw Sw
B1 0.232 2.640 2840 11.6 (kPa) (R1) (R2) (kPa) (R1) (R2)
B2 0.241 2.650 2820 12.1
0.0 0.403 0.441 0.0 0.774 0.771
B3 0.237 2.647 2930 12.0
21.8 0.688 0.711 0.9 0.610 0.608
R1 0.297 2.67 358
24.1 0.763 0.771 2.1 0.343 0.349
R2 0.293 2.67 326
27.3 0.802 0.817 4.7 0.232 0.241
213.8 0.820 0.840 11.4 0.208 0.202
B, Bentheimer; R, Reservoir rock.
232.9 0.876 0.890 33.5 0.176 0.193
271.1 0.901 0.936 75.3 0.172 0.189
2139.5 0.918 0.936 145.2 0.167 0.188
TABLE 2—TEST CONDITIONS AND FLUID PROPERTIES
2278.9 0.950 0.950 284.8 0.162 0.188
Property Bentheimer Reservoir Rock

Temperature, °C 20 70 diameter to the quartz cement thickness was used to determine


Brine density, g/cm3 1.02 0.99 whether the section cuts close to the grain center. If this ratio is
Brine viscosity, cp 1.06 0.43 below a predefined cutoff value that depends on the degree of
Oil density, g/cm3 0.76 0.83 quartz cementation, the grain diameter is assumed to be nonrep-
Oil viscosity, cp 1.40 3.11
resentative, and the grain is removed from the grain-size distribu-
tion curve.
Oil-water IFT, mN/m 35.0 7.1 The porosity of the sample is defined as the fraction pore area
contained in the BSE image. It is determined by first thresholding
the epoxy-filled pore space from the rock matrix and then mea-
TABLE 3—PRIMARY DRAINAGE CAPILLARY PRESSURE suring the corresponding area. The total amount of clay is deter-
DATA FOR THE BENTHEIMER SAMPLES mined similarly. The BSE signals from clays fall between the epoxy
and quartz/feldspar signals. By thresholding the clay, one may thus
Water Saturation measure the total amount of clay in the image. The clay texture is
Pc determined visually.
(kPa) Sample B1 Sample B2 Sample B3
Sandstone Reconstruction
0 1.0 1.0 1.0 The results from the thin-section analyses are used to reconstruct
2.5 1.0 0.972 0.956 3D sandstone models. The reconstruction algorithm has been de-
5.0 0.692 0.648 0.423 scribed in detail previously.40 The essence of our approach is to
10.0 0.135 0.137 0.129 build sandstone models that are analogs of actual sandstones by
20.0 0.090 0.098 0.093
stochastically modeling the results of the main sandstone-forming
processes—sedimentation, compaction, and diagenesis. In the
40.0 0.073 0.080 0.079 present work, the modeling of the sedimentation and compaction
80.0 0.063 0.068 0.068 processes is identical to that described previously.40 The diagenesis
150.0 0.056 0.062 0.063 modeling has been extended to include nonuniform quartz cement
300.0 0.054 0.057 0.058 overgrowth and microporous pore-lining and pore-filling clays or
500.0 0.051 0.052 0.055
cements.
In our initial attempt40 at modeling quartz cement overgrowth,
the radii of all the sand grains were increased uniformly. As shown
Oil relative permeabilities for the reservoir rock during forced by Schwartz and Kimminau,41 more realistic algorithms can be
water injection (i.e., negative capillary pressure) were determined defined by letting the rate of cement growth depend on the direction
from a single speed (2,500 rev/min 5 270 kPa) centrifuge dis- of growth. For each direction r (measured from the grain center),
placement applying Hagoort’s39 analytical method to compute we define the length l(r) as the distance between the surface of the
relative permeability. Fluid production was monitored with a Cen- original spherical grain and the surface of its Voronoi polyhe-
tas automated centrifuge system. Production data were recorded dron.40, 42 The rate of increase of the distance L(r) between the grain
every 2 seconds at the beginning of the displacement and then at center and the pore-grain surface is controlled by a cement growth
increasing intervals up to every 5 minutes. The experiment ran for exponent a according to the formula41
14 hours.
L~r! 5 Ro 1 min~ala , l !, . . . . . . . . . . . . . . . . . . . . . . . . . . . (1)
Thin-Section Analysis. Thin-sections were prepared from the end where R o 5 the radius of the original spherical grain and a simply
pieces of the core plugs used in the experiments. Backscattered controls the amount of cement growth (i.e., the porosity). Fig. 1
scanning electron microscope (BSE) images of the thin-sections illustrates the effect of the growth exponent a. Positive values of
were binarized and analyzed with a Kontron KS400 image ana- a favor growth of cement in the directions of large l(r) (i.e., the pore
lyzer. An erosion-dilation algorithm was used to partition the sand bodies). There is a clear tendency to form sheet-like pores as the
grain matrix into discrete grains. We measured the diameter of each porosity decreases. This ensures that the pore space remains inter-
grain and constructed a grain-size distribution curve. Cathodolu- connected down to very low porosities. Negative values of a favor
minescence images were used to distinguish quartz cement from the growth toward the directions with the smallest l(r) (i.e., the pore
original detrital grains. throats). Pore throats thus tend to be closed off as the porosity
The measured grain diameter is correct only if the thin-section decreases. This increases the tortuosity of the medium, and the
cuts through the center of the grain. The ratio of the measured grain porosity (i.e., threshold percolation porosity) at which the connec-

SPE Journal, December 1998 325


Fig. 1—Cross sections of computer generated models illustrat- Fig. 2—Cross sections of computer generated models illustrat-
ing the effect of increasing volumes of quartz cement over- ing the effect of increasing volumes of pore-lining (left) and
growth for negative (left) and positive (right) values of the ce- pore-filling (right) clays.
ment growth exponent.

tivity of the pore space vanishes, is higher than that for a . 0. If


a 5 0, quartz cement deposits as concentric overgrowth on the
grains.
Many varieties of clays or cements may precipitate in the pore
space during the diagenetic stages. Diagenetic clays clog and
subdivide pore space and generate microporosity. Pores within the
microporosity are typically one or two orders of magnitude less
than the intergranular pores. We assume that the permeability of
these pores is negligible compared with that of the intergranular
pore space.
The morphology of the common diagenetic clays can be divided
into three broad categories: pore-lining, pore-filling, and pore-
bridging. Pore-lining clays typically form thin crystals that grow
radially outward from the surface of the detrital grains. Typical
examples are chlorite, illite, and smectite. Pore-lining clays are
modeled by randomly precipitating clay particles (voxels) on the
surfaces of the detrital grains or the quartz cement. Fig. 2 shows that
the presence of pore-lining clays reduces the size of the intergranu-
lar pores. However, their effect on altering the effective tortuosity
of the medium is small.
Pore-filling clays, such as pseudohexagonal booklets of kaolin-
ite, are modeled with a clay clustering routine that causes clay
particles (voxels) to precipitate preferentially in randomly selected
pore bodies. Fig. 2 shows that pore-filling clays tend to form tight
clusters of particles that clog or block off pore throats even at Fig. 3—A small cube of a reconstructed pore space. The linear
relatively high porosities. The presence of pore-filling clays thus scale of the model is 100 voxels (0.75 mm) on each side.
significantly increases the tortuosity of the medium.
The morphology of the clay and the total amount of clay to be
precipitated in our models are determined from petrographical
analyses of thin-sections. Pore-bridging clays are currently not allow two or more fluids to flow simultaneously through the same
implemented. pore. Pores that are angular in cross section are a much more
realistic model of the pore structure than the commonly applied
Pore Network. Fig. 3 shows an example of a reconstructed pore cylindrical shape. In the present work, the shape of every pore body
space. Although it is possible to perform flow simulations directly and throat is described in terms of a dimensionless shape factor G,
on the chaotic pore space, either by numerically solving the Navier which is defined as46
Stokes equation43, 44 or by applying a Lattice Boltzmann simula-
tion,45 it can only be done at considerable computational expense. A#
G5 , . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (2)
It is, therefore, convenient to construct a pore network that repli- s2
cates the essential features of the pore space that are relevant to fluid
flow. The transformation of the reconstructed pore space into a pore where A# 5 the average cross-sectional area of the pore body or
network and the topological and geometrical characterization of the throat and s 5 the corresponding perimeter length. The area and
network have been described in detail previously.40 perimeter length are determined by standard image analysis tech-
Pore Shape. A key characteristic of real porous media is the niques. The shape factor replaces the irregular shape of a pore body
angular corners of pores. Angular corners retain wetting fluid and or throat by an equivalent irregular triangular shape. It ranges from

326 SPE Journal, December 1998


zero for a slit-shaped pore to 0.048 for an equilateral triangular
pore.
It is useful to relate G to the actual shape of the pore body or
throat. From elementary geometry, G may be expressed as

SO D
21
3
1 1
G5 , . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (3)
4 tan bi
i51

where b 5 the corner half angle (b1 % b2 % b3). Obviously, a single


value of G corresponds to a range of shapes. The combinations of
b1 and b2 that give a constant G is46

G5
sin~2b1 !
2 S
21
sin~2b1 !
sin~2b2 ! D 22
. . . . . . . . . . . . . . . . . . . . . . . (4)

For a given G, the minimum and maximum allowed values of b2


are found from Eq. 4 by setting b2 5 b1 and b2 5 p/4 2 b1/2,
respectively. In the present work, corner angles are assigned to pore
bodies and throats by first randomly selecting b2 (b2,min % b2 %
b2,max). Then, the remaining angles may be determined from Eq. 3.

Flow in the Network Fig. 4 —Fd(ur, G) vs. ur when water is present in at least one of the
We wish to simulate two-phase flow in our pore networks. The corners. The results for each G are the average of 1,000 real-
displacing fluid is injected through an external reservoir that is izations.
connected by pore throats to every pore body on the inlet side of
the network. The displaced fluid escapes through the outlet face on
the opposite side. Periodical boundary conditions are imposed For a pore body or throat i, the critical capillary pressure P *c is
along the sides parallel to the main direction of flow. given by
We assume that capillary forces dominate at the pore scale. This
P*c, i 5 Pi 1 Gi xi , . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (6)
is a reasonable assumption for low capillary number (1026 or less)
processes47 that are typical of most reservoir displacements. In the where P and G 5 input parameters and x varies uniformly between
following, we present a precise description of the different pore- zero and one. The first term on the right represents the disjoining
scale displacement mechanisms that may occur in the pore network pressure, whereas the second term represents the film curvature. If
during primary drainage and water injection. P *c is exceeded in an oil filled pore body or throat, the contact angle
must change.54 We assume that the new value of u r , 908 2 b 1 .
Primary Drainage. Initially, the network is fully saturated with This ensures that water is always present in at least one of the
water and is strongly water-wet. Then, oil enters the network corners.
representing migration into the reservoir. At every stage of the Primary drainage and the film rupturing process continue until a
process, oil invades the available pore body or throat with the maximum capillary pressure P c, max is reached. If P c, max exceeds P *c
lowest threshold capillary pressure. This forms the basis for the in a pore body or throat, the water film collapses and the wettability
invasion percolation algorithm48, 49 used previously to model drain- changes. In this case, the corners contain bulk water and remain
age processes. water-wet. Away from the corners, adsorption of hydrophobic
Threshold or entry capillary pressures are calculated with the components make the pore wall oil-wet. We refer to these pore
Mayer and Stowe50 and Princen51-53 (MS-P) method. The details of bodies and throats as being mixed-wet. Pore bodies and throats
the calculations are given in Appendix A. The threshold capillary where P c, max , P *c remain water-wet.
pressure is governed by both the pore shape and the receding
contact angle u r . Water Injection. The capillary pressure drops during water in-
jection, and water first invades the available water-wet pore bodies
g~1 1 2 ÎpG!cos ur and throats. Pore-scale displacement mechanisms for strongly wet-
Pec 5 Fd ~ur , G!, . . . . . . . . . . . . . . . . . . (5)
r ting systems have been described by Lenormand et al.55 There are
three types of displacements: piston type, pore body filling, and
where r 5 the inscribed radius of the pore body or throat. In general, snap-off. We summarize the behavior and the threshold capillary
the function F d is dependent on the particular corner angles and is pressures for each of these displacement in mixed-wet pores.
not universal for a specific G. In practice, however, F d varies little Piston Type. This refers to the displacement of oil from a pore
for a given G. This is illustrated in Fig. 4, which shows the full throat by an invading interface initially located in an adjoining
variation in F d when water is present in at least one of the corners. water filled pore body. If there is no contact angle hysteresis, the
For strongly water-wet systems (i.e., u r 5 08), F d 5 1. threshold capillary pressure is the same as for drainage (Eq. 5). In
a more realistic scenario, the advancing contact angle u a is different
Wettability Alteration. When oil initially invades a water filled from u r . This is always true for mixed-wet systems in which u a may
pore body or throat, a stable water film protects the pore surface be much larger than u r .
from wettability change by adsorption. At a critical capillary In this case, there is a range of capillary pressure in which the
pressure, the film collapses to form a molecular thin film. This invading interface remains pinned. As the capillary pressure drops,
allows surface active components in the oil to adsorb on the pore the interface remains fixed in place, and the contact angle adjusts
surface. The capillary pressure at which the water film ruptures to a new value u h . The hinging angle u h can acquire any value
depends on the curvature of the pore wall and on the shape of the between u r and u a . Water enters the throat after the capillary
disjoining pressure isotherm. Kovscek et al.54 present a detailed pressure is lowered sufficiently so that u a is reached. The calcu-
analysis of this scenario for star-shaped pores. Blunt32 recently lation of the threshold capillary pressure for this case is given in
extended the analysis to square pores and presents a parametric Appendix A.
model for the critical capillary pressure at which the water film Solutions for the normalized threshold capillary pressure (P Nc 5
collapses. A similar approach is adopted here. rP ec / g ) show that spontaneous imbibition (P Nc . 0) by piston type

SPE Journal, December 1998 327


displacement always occurs for u a less than 90° and may occur for
u a much greater than 90° (Fig. 5). The maximum advancing angle
at which spontaneous imbibition occurs is given by

ua, max 5 cos21 F 24G O


cos~ur 1 bi !
, G . . . . . . . . . . . . (7)
PNd
c 2 cos u r 1 12G sin ur

where P Ndc 5 rP c, max/g. The maximum advancing angle for spon-


taneous imbibition, which determines the residual water saturation
in oil-invaded pores, depends on P c, max, G, and u r . Fig. 6 illustrates
this and shows that u a, max and, therefore, spontaneous imbibition
decreases when the residual water saturation is reduced.
If the advancing angle is greater than u a, max, the water invasion
is forced. Forced water invasion is analogous to the oil drainage
process. If u a . 908 1 b 1 , the entry capillary pressure may be
determined from Eq. 5. If u a , 908 1 b 1 , the threshold capillary
pressure is given by
2g cos ua
Pec 5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (8)
r
Pore Body Filling. The threshold capillary pressure for pore
body filling is limited by the largest radius of curvature required to Fig. 6 —Effect of initial water saturation on the maximum contact
invade the pore body. This radius depends on the size of the pore angle at which spontaneous imbibition by piston type displace-
body and on the number of connecting throats filled with oil. For ment occurs. The receding angle is zero.
a pore body with coordination number z, there are z 2 1 such
mechanisms, referred to as I 1 to I z21 . If only one of the connecting
throats contains oil (i.e., I 1 ), the filling of the pore body is similar hysteresis, the AM’s advance smoothly along the pore walls as the
to that of a piston type invasion and the threshold capillary pressure capillary pressure drops. At a critical point, the AM’s fuse together
is the same as for the corresponding piston type displacement. and the center of the pore space spontaneously fills with water. For
The threshold capillary pressures for the I 2 to I z21 mechanisms a strongly water-wet system, this occurs at a threshold capillary
are more complex, especially for the chaotic networks used in the pressure P ec 5 g /r.
present work. Blunt32 presents a parametric model for this, and a If there is contact angle hysteresis, the AM’s remain pinned at
similar approach is used here. If u a , u a, max, the mean radius of the position established at P c, max until the hinging angle in the
curvature for filling by an I n mechanism is computed as sharpest corner equals u a . Further decrease in the capillary pressure

S D
causes the AM to advance toward the center of the pore space.
Eventually, this AM meets another AM, causing snap-off. If u a ,
O
n
1
Rn 5 r 1 bi ri xi , . . . . . . . . . . . . . . . . . . . . . . . . (9) 908 2 b 1 , the curvatures of the AM’s are positive and snap-off
cos ua p occurs at a positive capillary pressure. The capillary pressure at
i51
which this occurs can be calculated from elementary geometry
where r p 5 the pore body radius, b i 5 input parameters, r i 5 the (Appendix A). In contrast to piston type invasion, spontaneous
radii of the oil filled throats, and x i 5 random numbers between imbibition by snap-off only occurs for u a , 908 (Fig. 7).
zero and one. The threshold capillary pressure is P c, n 5 2 g /R n . If If u a . 908 2 b 1 , the curvatures of the AM’s are negative and
u a . u a, max, the water invasion is forced, and the threshold capillary the invasion is forced. Once the hinging angle in the sharpest corner
pressure is the same as for the I 1 mechanism. has increased to u a , the AM advances toward the center of the pore
space. This causes the absolute value of the negative curvature to
Snap-Off. Snap-off refers to the invasion of an oil filled pore space decrease. The AM is unstable and the pore body or throat imme-
by arc menisci (AM’s), which always exist in the corners of oil
filled pore bodies and throats. In the absence of contact angle

Fig. 7—Normalized threshold pressure for snap-off. The reced-


Fig. 5—Normalized threshold pressure for piston displacement. c 5 2. The results for each G are the
ing angle is zero and PNd
c 5 2.
The receding angle is zero and PNd average of 1,000 realizations.

328 SPE Journal, December 1998


diately fills with water. The threshold capillary pressure for this The boundary condition is that u 5 0 along the walls of the pore
process depends on the curvature of the AM when it begins to body or throat. Eq. 16 was solved by use of a standard finite
move. It is given by difference method. Hydraulic conductances were determined by
integrating the computed velocity field over the pore body or throat
cos~ua 1 b1 ! cross section.
Pec 5 Pc, max ua % 1808 2 b1 . . . . . . . . . . . . . . (10)
cos~ur 1 b1 ! Fig. 8 shows computed conductances for different shaped tri-
angular pores. The functional dependency between g and the shape
21 factor G is almost linear and is closely approximated by
and Pec 5 Pc, max ua ^ 1808 2 b1 . . . . . . . . . . (11)
cos~ur 1 b1 !
3A2 G 3r2 A
Formation of Oil Films. Piston like advance in a mixed-wet g5 5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (18)
5m 20m
pore body or throat leaves the center of the pore space filled with
water. If u a . 908 1 b 1 . u a, max, a film of oil may be left
sandwiched between the water in the corner(s) and the water in the If oil occupies the center of a pore body or throat with water present
center. These films may significantly increase the connectivity of as AM’s in the corners, the oil conductance is found from Eq. 18,
the oil phase and allow oil mobility down to very low saturations. but with A replaced by the cross-sectional area occupied by oil.
Similar to Blunt,32 we assume that the oil film in a corner is stable When water is present as AM in a corner i with a contact angle
until the two oil/water interfaces on either side of the film meet. The less than 908 2 b i , the water conductance is
critical capillary pressure at which an oil film in a corner i collapses
is given by r2w Aw, i
gw, i 5 , . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (19)
Cw, i m
Pc
5
sin bi
S1 2 d2
2Pc, max cos~ur 1 bi ! d cos bi 1 Î1 2 d2 sin2 bi
, . . . . . (12) D where C w 5 a dimensionless flow resistance factor that accounts for
the reduced water conductivity close to the pore walls. Numerical
where d 5 2 1 cos u a /sin b i . solutions of the corner flow problem56 show that C w depends on the
corner geometry, the contact angle, and the boundary condition at
Fluid Volumes and Conductances the oil/water interface. The total conductance is the sum of all the
Pore bodies and throats are assumed to have a triangular cross corner conductances.
section with area A 5 r 2 /4G. When water is present as AM’s in If the contact angle is greater than 908 2 b i , the curvature of the
the corners of a pore body or throat, the area occupied by the water AM is negative and we do not have an exact expression for the
is given by conductance. In this case, we use an approximation based on Eq.
19 that may be written as

OF S DG
3
cos u cos~u 1 bi ! p u 1 bi
Aw 5 r2w 2 12
2
, . . . . . . (13) Aw,
sin bi
i
2 90 gw, i 5 , . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (20)
i51 Cw, i ki m
where r w 5 g /P c 5 radius of curvature of the AM’s and u can be
where C w is evaluated at u 5 0° and k is given by
u r , u h , or u a , depending on the circumstances. When oil films are

S D
present in the corners of a pore body or throat, the area occupied cos bi p bi
by the oil is ki 5 2 12 . . . . . . . . . . . . . . . . . . . . . . . . . . (21)
sin bi 2 90

OF S DG
3
cos ua cos~ua 2 bi ! p ua 2 bi
Ao 5 r2w 2 21 2 Aw , If oil films are present in the corners of a pore body or throat, the
sin bi 2 90 oil conductance is found from Eq. 20, but with A w replaced by A o .
i51
This is an approximation because the geometry of the oil film is
. . . . . . . . . . . . . . . . . . . . . . . . . . (14) different from that of a water AM.32
where A w is given by Eq. 13 and the sum includes only the corners
where oil films exist. The volume of water or oil is assumed to be
the fraction of the cross section occupied by the phase multiplied
by the total pore body or throat volume. The volume of trapped fluid
is computed by use of the capillary pressure at which the fluid was
first trapped. The overall saturation of each phase is found by
adding the volume of each phase in every pore body and throat and
dividing by the total pore volume of the network.

Conductances. The hydraulic conductance of completely filled


pore bodies and throats is computed by assuming a Poiseuille law
relation between the flow rate q and the pressure gradient ¹p,
q 5 2g¹p. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (15)
At low Reynolds numbers, incompressible flow in a straight hor-
izontal pore body or throat may be described in dimensionless form
as
­2 u ­2 u
1 5 21, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (16)
­Y2 ­Z2
where the independent space variables y 5 Y =A and z 5 Z =A.
The dimensionless velocity u is defined as
Fig. 8 —Computed dimensionless conductances (gm/A2) as a
vx m function of the shape factor for different triangular shaped
u5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (17)
A~2dp/dx! pores.

SPE Journal, December 1998 329


Macroscopic Properties. For laminar flow, the flow rate of fluid
i between two connecting nodes I and J is given by
gi, IJ
qi, IJ 5 ~ p 2 pi, J !, . . . . . . . . . . . . . . . . . . . . . . . . . . . (22)
LIJ i, I
where L IJ 5 the spacing between the pore body centers. The
effective conductance g i, IJ is assumed to be the harmonic mean of
the conductances of the throat and the two pore bodies themselves,
i.e.,
LIJ
5
Lt 1 LI
1 1 S
LJ
gi, IJ gi, t 2 gi, I gi, J
, D . . . . . . . . . . . . . . . . . . . . . . . . (23)

where the subscript t 5 the pore throat. We impose mass conser-


vation at each pore body, which means that

O J
qi, IJ 5 0, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (24)

where J runs over all the pore throats that are connected to Pore
Body I. Eqs. 22 through 24 give rise to a set of linear equations for
the pore body pressures. Fig. 9 —Standard deviation in local porosity for a reconstructed
The permeability of the network is computed by imposing a model and a micro-CT image of a Fontainebleau sandstone.
constant pressure gradient across the network and letting the system
relax by use of a conjugate gradient method to determine the pore
body pressures. From the pressure distribution, we calculate the
total flow rate and, thus, the permeability by use of Darcy’s law.
Relative permeabilities are computed similarly. At various stages
of the displacement, we compute the pressure drop in each phase
separately. The saturation and the total flow rate in each phase are
also calculated. Then, relative permeabilities may be determined
from Darcy’s law.
Because of the analogy between Poseuille’s law and Ohm’s law,
the flow of electrical current in the network is also described by
Eqs. 22 through 24, but with pressure replaced by voltage and
hydraulic conductance by electrical conductance. The electrical
conductance g e is assumed to depend only upon the geometry of the
pore body or throat (i.e., pore walls are insulating) and is given by
ge 5 sw A, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (25)
where s w 5 the electrical conductivity of the fluid that fills the pore
network. The electrical conductivity, s, of the pore network is
computed by applying a constant voltage drop across the network
and then applying Eq. 22 to find the flow of current between each
pore body. From this, one may calculate the total current and, thus,
the formation factor, F F 5 s w / s .

Results
Fontainebleau Sandstone. The sandstone reconstruction algo-
rithm was validated by comparing a reconstructed sample of Fig. 10 —Horizontal (top) and vertical (bottom) two-point corre-
Fontainebleau sandstone with microtomographic images of the lation functions for the Fontainebleau sandstone. Micro-CT im-
actual sample. Fig. 9 compares the standard deviation in local age (left), reconstructed model (right) (1 unit 5 7.5 mm).
porosity for the two samples. Local porosity distributions were
measured by means of a moving box technique with a logarithmic
increase in box size. This ensures that both local and global Thin-section analyses reveal that it is well sorted with an average
variations in porosity are captured. The model closely mimics the grain size of around 200 mm. The porosity and permeability,
actual sample, except at large box sizes, where the variation in however, vary widely. The variation in these properties is predom-
porosity is too small. This suggests that the reconstructed model inantly caused by variation in the volume of quartz cement over-
does not capture all the global heterogeneities that are present in the growth. On the basis of the reconstructed model, we generated
real sample. models with different porosities by varying the amount of quartz
The horizontal and vertical two-point correlation functions were cement overgrowth. The value of the cement growth exponent a
computed as a function of varying box size and distance, thus was 0.6 for all the models.
covering both local and global correlations. Fig. 10 shows these Fig. 11 compares the predicted permeability vs. porosity trend
correlation functions, which may be represented as 2D surfaces. with the experimental data of Bourbie and Zinszner.57 Although the
The correlation functions for the two samples are very similar. measured permeabilities span nearly five orders of magnitude, the
Statistical analysis shows that there is no significant differences predicted permeability vs. porosity trend is in good agreement with
within the level of noise we experience (1,000 replications for each the experimental one. Fig. 11 shows that there is a change in the
box size/distance combination). permeability vs. porosity trend at approximately 10 to 13% poros-
Bourbie and Zinszner57 performed extensive laboratory mea- ity. Our simulation results show that this corresponds to the porosity
surements of the hydraulic properties of Fontainebleau sandstone. at which quartz cement begins to close off pore throats. This
The Fontainebleau sandstone is an aeolan fine grained quartzite. increases the tortuosity of the medium, and the permeability de-

330 SPE Journal, December 1998


imental data, whereas the predicted permeabilities are approxi-
mately 30% higher than the measured ones. This suggests that we
overestimate the hydraulic conductance of pore bodies and throats.
One possible explanation for this may be that surface roughness is
not accounted for in our expression (Eq. 18) for the hydraulic
conductance.
Primary drainage and waterflood displacements were simulated
on each of the realizations. In all the simulations, the receding
contact angle was zero, whereas the advancing contact angle for
oil-invaded pore bodies and throats was randomly distributed
between 30 and 50°. This is in agreement with contact angle
measurements on quartz.60 The corner resistance factor C w was
calculated assuming a no-flow boundary condition at the oil/water
interface.56
Fig. 13 compares simulated and measured drainage capillary
pressures. The calculated capillary pressures are the average for the
six realizations. The predicted capillary pressure curve is in good
agreement with the measured data, and we correctly predict the
threshold displacement pressure and the low connate water satu-
ration. For the high capillary pressures established at the end of the
primary drainage simulations, oil has invaded almost all the inter-
Fig. 11—Comparison between predicted and measured abso- granular pore space. The connate water saturation is mainly made
lute permeability for Fontainebleau sandstone. up of water present in the corners of oil-invaded pore bodies and
throats and water present in the microporosity, which is assumed
to be impermeable.
clines more rapidly. Fig. 12 compares predicted and measured58, 59 Fig. 14 compares simulated primary drainage and waterflood
formation factors. The predicted formation factors are also in good relative permeability curves with experimental data. Although the
agreement with the measured ones, except for some deviations at predicted results are the average of only six realizations, the
high porosities where we tend to underestimate the formation comparison shows that the network model reproduces the experi-
factor. These results demonstrate that our sandstone models and mental characteristics of the measured data and correctly predicts
their network representations adequately predict single-phase trans- the waterflood residual oil saturation. At low water saturations, the
port properties for this texturally and diagentically simple quartz predicted water permeability is larger than the measured one. This
sandstone. suggests that we overestimate the water conductance in the corners
of oil-invaded pore bodies and throats. The expression for the
Bentheimer Sandstone. Thin section images of the Bentheimer corner water conductance (Eq. 19) assumes a perfectly sharp corner
core plugs described earlier were analyzed and used to reconstruct with smooth pore walls. In reality, the pore walls are irregular and
two realizations of each sample. Table 5 summarizes the main input rough and their curvature is not zero. Both surface roughness and
parameters for the reconstruction algorithm. Table 6 compares pore wall curvature can reduce the water conductance.
simulated petrophysical properties for the reconstructed samples
with measured values. Reservoir Rock. The main clay component in the reservoir rock
The predicted porosities are 2 to 3 porosity units (p.u.) less than samples is pore-filling kaolinite. In addition, the samples contain
the plug porosities that were measured with a helium porosimeter. lesser amounts of clay pseudomorphs after potassium-feldspar,
The discrepancy between measured and predicted porosity is most carbonate cement, pyrite, and trace amounts of chlorite. For sim-
likely caused by differences in the amount of microporosity, which plicity, we model all this as pore-filling clay or cement. Table 5
is difficult to determine accurately from thin-section analyses. gives input parameters for the reconstruction algorithm. Two re-
Predicted formation factors are in good agreement with the exper- alizations of each core plug were reconstructed. Table 6 gives the
predicted porosities and permeabilities.
The predicted porosities are about 5 p.u. less than the helium
porosities. This may be because we underestimate the amount of
microporosity or because the thin sections are not fully represen-
tative of the core plugs (i.e., heterogeneity). The predicted perme-
abilities are in fair agreement with the measured values.
As mentioned previously, the reservoir rock samples display a
mixed wettability. Unfortunately, it is not possible to know a priori
what fraction of the pore space that is mixed-wet nor the contact
angles that should be assigned to the water-wet and mixed-wet
regions of the pore space. Instead, we investigate how the water-
flood residual oil saturation, S or , depends on the fraction f of
oil-invaded pore bodies and throats that become mixed-wet.
Fig. 15 shows how S or depends on the fraction of mixed-wet pore
bodies and throats. In all the simulations, the advancing contact
angle for mixed-wet pore bodies and throats is 180°, whereas u a for
the water-wet regions is 40°. The receding angle is 20°. For small
f, the mixed-wet regions of the pore space do not form a continuous
cluster. So, oil is trapped in the mixed-wet regions of the pore space
and S or increases almost linearly with f. At a critical f (around 0.62),
mixed-wet pore bodies and throats first form a continuous flow path
to the outlet. Forced water injection can then displace oil, and S or
Fig. 12—Comparison between predicted and measured forma- drops sharply. These results are in qualitative agreement with those
tion factors for Fontainebleau sandstone. reported by Blunt.32

SPE Journal, December 1998 331


TABLE 5—INPUT PARAMETERS FOR THE SANDSTONE RECONSTRUCTION ALGORITHM

Parameter B1 B2 B3 R1 R2

Grain radius, mm 90 to 303 90 to 303 90 to 303 21 to 74 21 to 74


Clay content, wt% 2.5 2.5 2.5 20.0 19.0
Clay texture* m m m p-f p-f
Compaction factor 0.05 0.05 0.05 0.0 0.0
Cement exponent 20.2 20.2 20.2 1.0 1.0
Intergranular porosity 0.194 0.204 0.201 0.210 0.218
Size, mm3 2.55 2.55 2.55 1.02 1.02
Resolution, mm 10.0 10.0 10.0 4.0 4.0

* m 5 mixture; p-f 5 pore-filling.

TABLE 6—COMPARISON BETWEEN PREDICTED AND MEASURED PETROPHYSICAL


PROPERTIES FOR BENTHEIMER AND RESERVOIR ROCK SAMPLES

Permeability
Porosity (md) Formation Factor

Sample* Expected Predicted Expected Predicted Expected Predicted

B1 0.232 0.210 2,840 3,650 11.6 9.4


B2 0.241 0.219 2,820 3,906 12.1 11.2
B3 0.237 0.216 2,930 3,764 12.0 11.1
R1 0.297 0.243 358 429 18.6
R2 0.293 0.245 326 472 17.7

* B 5 Bentheimer; R 5 Reservoir rock.

included in the model. To capture the presence of these small pores,


the resolution of the reconstructed model must be increased. This,
however, can only be done at considerable computational expense.
The agreement between the predicted and measured oil relative
permeabilities is encouraging. In spite of the necessarily approxi-
mate nature of the distribution of water-wet and mixed-wet pores
as well as the uncertain estimates for both the contact angles and
the oil film conductance, the predicted and measured oil relative
permeability curves display the same generic behavior. Both the
experiments and the simulations show that continuous oil films
allow low oil saturations to be reached during forced water injec-
tion. However, the oil relative permeability is very low.
The fair agreement between the predicted and measured transport
properties for the three different reconstructed sandstones is en-
couraging. Although we need to examine a much larger group of
petrographically and petrophysically heterogeneous rock samples,
these preliminary results cautiously suggest that network modeling
techniques may be used to a priori predict single and multiphase
transport properties for real rocks.

Conclusions
Fig. 13—Comparison between predicted and measured primary
drainage capillary pressure for the water-wet Bentheimer sand- 1. Petrographical data obtained from image analysis of 2D thin-
stone. sections are used to reconstruct model sandstones, which give a
realistic description of the complex microstructure exhibited in real
sandstones.
The experimentally determined S or for the reservoir rock is 5%. 2. Image analysis techniques are used to transform the recon-
From Fig. 15, this corresponds to f 5 0.85. Fig. 16 shows predicted structed pore space into a pore network that can be used as input
waterflood capillary pressure and relative permeability curves for to a network simulator.
this case. The simulated capillary pressure curve is in fair agree- 3. Predicted permeabilities and formation factors of a recon-
ment with the experimental data except at low oil saturations. At structed Fontainebleau sandstone model correspond well with pub-
low oil saturations, the simulated capillary pressure curve displays lished data over a wide range of porosity.
a sharper break than does the experimental data. This suggests that 4. Predicted primary drainage and waterflood relative perme-
the reconstructed sample has a narrower pore size distribution than ability curves for a water-wet Bentheimer sandstone are in good
the actual sample and that some of the smaller pores are not agreement with experimental data.

332 SPE Journal, December 1998


Fig. 14 —Comparison between predicted and measured drainage (left) and waterflood (right) relative permeabilities for a water-wet
Bentheimer sandstone.

P c 5 capillary pressure, N/m2


P *c 5 threshold capillary pressure for water film collapse, Eq. 6,
N/m2
P c, max 5 maximum capillary pressure reached during primary
drainage, N/m2
q 5 volumetric flow rate, m3/sec
r 5 radius, m
R n 5 radius of curvature for pore body filling, m
s 5 perimeter length, m
S or 5 residual oil saturation
u 5 dimensionless velocity
v 5 velocity, m/sec
a 5 cement growth exponent
b 5 corner half angle, degrees
g 5 interfacial tension, N/m
G 5 input parameter, Eq. 6
u 5 contact angle, degrees
m 5 viscosity, kg/m sec
P 5 parameter controlling water film collapse
r 5 density, kg/m3
s 5 electrical conductivity
Fig. 15—Waterflood residual oil saturation for the reconstructed
f 5 porosity
reservoir rock as a function of the fraction of pore bodies and
throats contacted by oil, which become mixed-wet.
Subscripts
a 5 advancing
5. Simulated waterflood capillary pressures and oil relative per- d 5 drainage
meabilities for a mixed-wet reservoir rock are in fair agreement e 5 electrical
with experimental measurements. eff 5 effective
h 5 hinging
Nomenclature min 5 minimum
a 5 parameter controlling the amount of cement growth
max 5 maximum
A 5 cross-sectional area, m2 o 5 oil
b 5 input parameter, Eq. 9 p 5 pore body
C 5 corner resistance factor pd 5 primary drainage
f 5 fraction of mixed-wet pore bodies and throats pt 5 piston type
F F 5 formation factor s 5 solid
g 5 conductance r 5 receding
G 5 pore body or throat shape factor t 5 pore throat
k 5 permeability, m2 w 5 water
k r 5 relative permeability
l 5 distance between surface of grain and its Voronoi poly-
hedron, m Superscripts
L 5 length, m e 5 entry
L b 5 distance between pinned contact line and corner, m d 5 drainage
p 5 pressure, N/m2 N 5 normalized

SPE Journal, December 1998 333


Fig. 16 —Comparison between simulated and measured waterflood capillary pressures and relative permeabilities for a mixed-wet
reservoir rock.

Acknowledgments 16. Bryant, S., King, P.R., and Mellor, D.W.: “Network Model Evaluation
The authors acknowledge Den Norske Stats Oljeselskap A/S (Sta- of Permeability and Spatial Correlation in a Real Random Sphere
toil) for granting permission to publish this paper. We are indebted Packing,” Transport in Porous Media (1993) 11, 53.
to Svein H. Midtlyng for performing the experimental work. We 17. Billiotte, J., De Moegen, H., and Øren, P.E.: “Experimental Micro-
also graciously thank David Stern, Exxon Research Production Co., Modeling and Numerical Simulation of Gas/Water Injection/With-
for providing us with the microtomography data. drawal Cycles as Applied to Underground Gas Storage Reservoir,” SPE
Advanced Technology Series (April 1993) 133.
References 18. McDougall, S.R. and Sorbie, K.S.: “The Combined Effect of Capillary
1. Fatt, I.: “The Network Model of Porous Media. I. Capillary Pressure and Viscous Forces on Waterflood-Displacement Efficiency in Finely
Characteristics,” Trans., AIME (1956) 207, 144. Laminated Porous Media,” paper SPE 26659 presented at the 1993 SPE
2. Fatt, I.: “The Network Model of Porous Media. II. Dynamic Properties Annual Technical Conference and Exhibition, Houston, 3–6 October.
of a Single Size Tube Network,” Trans., AIME (1956) 207, 160. 19. Blunt, M.J. and Scher, H.: “Pore Level Modelling of Wetting,” Physical
3. Fatt, I.: “The Network Model of Porous Media. III. Dynamic Properties Review E (December 1995) 52, 6387.
of Networks With Tube Radius Distribution,” Trans., AIME (1956) 207, 20. Øren, P.E. et al.: “Prediction of Relative Permeability and Capillary
164. Pressure From Pore-Scale Modelling,” Proc., European Conference on
4. Larson, R.G., Scriven, L.E., and Davis, H.T.: “Percolation Theory of the Mathematics of Oil Recovery, Leoben, Austria (3–6 September
Two-Phase Flow in Porous Media,” Chemical Engineering Science 1996).
(1981) 36, 57. 21. Hazlett, R.D.: “Simulation of Capillary-Dominated Displacements in
5. Koplik, J.: “Creeping Flow in Two-Dimensional Networks,” J. Fluid Microtomographic Images of Reservoir Rocks,” Transport in Porous
Mechanics (1982) 119, 219. Media (1995) 20, 21.
6. Koplik, J. and Lasseter, T.J.: “Two-Phase Flow in Random Network 22. Dunsmuir, J.H. et al.: “X-Ray Microtomography: A New Tool for the
Models of Porous Media,” SPEJ (February 1985) 22, 89. Characterization of Porous Media,” paper SPE 22860 presented at the
7. Chan, D.Y.C. et al.: “Simulating Flow in Porous Media,” Physical 1991 SPE Annual Technical Conference and Exhibition, Dallas, 6–9
Review A (1988) 38, 4106. October.
8. Touboul, E., Lenormand, R., and Zarcone, C.: “Immiscible Displace- 23. Coles, M.E. et al.: “Computed Microtomography of Reservoir Core
ments in Porous Media: Testing Network Models by Micromodel Ex- Samples,” Proc., Intl. SCA Meeting, Stavanger, Norway (12–14 Sep-
periments,” paper SPE 16954 presented at the 1987 SPE Annual Tech- tember 1994).
nical Conference and Exhibition, Dallas, 27–30 September.
24. Coles, M.E. et al.: “Developments in Synchroton X-Ray Microtomog-
9. Lenormand, R., Touboul, E., and Zarcone, C.: “Numerical Models and
raphy with Applications to Flow in Porous Media,” paper SPE 36531
Experiments on Immiscible Displacements in Porous Media,” J. Fluid
presented at the 1996 SPE Annual Technical Conference and Exhibi-
Mechanics (1988) 189, 165.
tion, Denver, Colorado, 6–9 October.
10. Lenormand, R.: “Flow Through Porous Media: Limits of Fractal Pat-
25. Baldwin, C.A. et al.: “Determination and Characterization of the Struc-
terns,” Proc., Royal Society, London A (1989) 423, 159.
ture of a Pore Space from 3D Volume Images,” J. Colloid and Interface
11. Blunt, M.J. and King, P.R.: “Macroscopic Parameters From Simulation
Science (1996) 181, 79.
of Pore Scale Flow,” Physical Review A (1990) 42, 4780.
12. Jerauld, G.R. and Salter, S.J.: “The Effect of Pore-Structure on Hys- 26. Lin, C. and Cohen, M.H.: “Quantitative Methods of Microgeometric
teresis in Relative Permeability and Capillary Pressure: Pore-Level Modelling,” J. Applied Physics (1982) 53, 4152.
Modelling,” Transport in Porous Media (1990) 5, 103. 27. Holt, R. et al.: “Petrophysical Laboratory Measurements for Basin and
13. Blunt, M.J. and King, P.R.: “Relative Permeabilities From Two- and Reservoir Evaluation,” Marine and Petroleum Geology (1996) 13, No.
Three-Dimensional Pore-Scale Network Modelling,” Transport in Po- 4, 383.
rous Media (1991) 6, 407. 28. Dixit, A.B., McDougall, S.R., and Sorbie, K.S.: “A Pore-Level Inves-
14. Blunt, M.J., King, M., and Scher, H.: “Simulation and Theory of tigation of Relative Permeability Hysteresis in Water-Wet Systems,”
Two-Phase Flow in Porous Media,” Physical Review A (1992) 46, 7680. SPE Journal (June 1998) 115.
15. Bryant, S. and Blunt, M.J.: “Prediction of Relative Permeability in 29. McDougall, S.R. and Sorbie, K.S.: “The Impact of Wettability on
Simple Porous Media,” Physical Review A (1992) 46, No. 4, 2004. Waterflooding: Pore-Scale Simulation,” SPERE (August 1995) 208.

334 SPE Journal, December 1998


30. Heiba, A.A., Davis, H.T., and Scriven, L.E.: “Effect of Wettability on 55. Lenormand, R., Zarcone, C., and Sarr, A.: “Mechanism of the Dis-
Two-Phase Relative Permeabilities and Capillary Pressures,” paper SPE placement of One Fluid by Another in a Network of Capillary Ducts,”
12172 presented at the 1983 SPE Annual Technical Conference and J. Fluid Mechanics (1983) 135, 337.
Exhibition, San Francisco, 5–8 October. 56. Ranshoff, T.C. and Radke, C.J.: “Laminar Flow of Wetting Liquid
31. Blunt, M.J.: “Effects of Heterogeneity and Wetting on Relative Per- Along the Corners of a Predominantly Gas-Occupied Noncircular
meability Using Pore Level Modeling,” SPE Journal (March 1997) 70. Pore,” J. Colloid and Interface Science (1988) 121, 392.
32. Blunt, M.J.: “Pore Level Modeling of the Effects of Wettability,” SPE 57. Bourbie, T. and Zinszner, B.: “Hydraulic and Acoustic Properties as a
Journal (December 1997) 449. Function of Porosity in Fontainebleau Sandstone,” J. Geophysical
33. Soll, W.E. and Celia, M.A.: “A Modified Percolation Approach to Research (1985) 90, 11524.
Simulating Three-Fluid Capillary Pressure-Saturation Relationships,” 58. Jacquin, C.G.: “Correlation entre la Permeabilite et les Caracteristiques
Advances in Water Resources (1993) 16, 107. Geometriques du gres de Fontainebleau,” Review Institut Francois du
34. Øren, P.E., Billiotte, J., and Pinczewski, W.V.: “Pore-Scale Network Petrole (1964) 19, 921.
Modelling of Waterflood Residual Oil Recovery by Immiscible Gas 59. Fredrichs, J.T., Greaves, K.H., and Martin, J.W.: “Pore Geometry and
Flooding,” paper SPE 27814 presented at the 1994 SPE/DOE 9th Transport Properties of Fontainebleau Sandstone,” Intl. J. Rock Me-
Symposium on Improved Oil Recovery, Tulsa, Oklahoma, 17–20 April. chanics and Mining Science and Geomechanics (1993) 30, No. 7, 691.
35. Øren, P.E. and Pinczewski, W.V.: “Fluid Distribution and Pore-Scale 60. Hjelmeland, O., Selle, O.M., and Rueslåtten, H.: “The Effects of Rock
Displacement Mechanisms in Drainage Dominated Three-Phase Flow,” Properties on Wettability of Oil Reservoirs—An Investigation of North
Transport in Porous Media (1995) 20, 105. Sea Systems,” Proc., European Symposium on Enhanced Oil Recovery,
Hamburg, Germany (27–29 October 1987).
36. Fenwick, D.H. and Blunt, M.J.: “Pore Level Modelling of Three Phase
61. Morrow, N.R.: “Pore Level Displacement Mechanisms,” paper pre-
Flow in Porous Media,” Proc., European IOR Symposium, Vienna,
sented at the 1990 First Intl. Symposium on Evaluation of Reservoir
Austria (15–17 May 1995).
Wettability and Its Effect on Oil Recovery, Socorro, New Mexico,
37. Pereira, G.G. et al.: “Pore-Scale Network Model for Drainage Domi-
18–21 September.
nated Three-Phase Flow in Porous Media,” Transport in Porous Media
62. Ma, S., Mason, G., and Morrow, N.R.: “Effect of Contact Angle on
(1996) 24, 167.
Drainage and Imbibition in Regular Polygonal Tubes,” Colloids and
38. Mani, V. and Mohanty, K.K.: “Effect of the Spreading Coefficient on Surfaces A (1996) 117, 273.
Three-Phase Flow in Porous Media,” J. Colloid and Interface Science
(1997) 187, 45.
Appendix A—Threshold Pressures
39. Hagoort, J.: “Oil Recovery by Gravity Drainage,” SPE Journal (June
1980) 139. Mason and Morrow46 applied the MS-P method to derive a general
40. Bakke, S. and Øren, P.E.: “3-D Pore-Scale Modeling of Sandstones and expression in terms of G for the drainage threshold capillary
Flow Simulations in the Pore Networks,” SPE Journal (June 1997) 136. pressure in strongly wetted triangular pores. This analysis was later
41. Schwartz, L.M. and Kimminau, S.: “Analysis of Electrical Conduction extended to include the effect of contact angle and contact angle
in the Grain Consolidation Model,” Geophysics (October 1987) 52, hysteresis in equilateral triangular pores.61 In the present work, we
1402. generalize this model to any shape triangular pores. Our compu-
42. Roberts, J.N. and Schwartz, L.M.: “Grain Consolidation and Electrical
tations are similar to those reported by Ma et al.62 for the displace-
Conductivity in Porous Media,” Physical Review (1985) B31, No. 9,
ment curvature of a meniscus in regular polygonal tubes.
5990.
43. Adler, P.M., Jacquin, C.G., and Quiblier, J.A.: “Flow in Simulated Drainage. Capillary forces prevent oil from spontaneously enter-
Porous Media,” Intl. J. Multiphase Flow (1990) 16, 691. ing water filled throats. Oil can only enter an available throat if the
44. Adler, P.M., Jacquin, C.G., and Thovert, J.F.: “The Formation Factor of capillary pressure exceeds the threshold capillary pressure. At the
Reconstructed Porous Media,” Water Resources Research (1992) 28, threshold capillary pressure, oil enters the throat with a fixed
No. 6, 1571. curvature and displaces water from the central part of the throat,
leaving some of the water as AM’s in the corners. In the absence
45. Ferreol, B. and Rothman, D.H.: “Lattice-Boltzmann Simulations of
of gravity, the curvature of these AM’s is the same as the curvature
Flow Through Fontainebleau Sandstone,” Transport in Porous Media
of the invading interface.
(1995) 20, 3.
The MS-P method for calculating threshold capillary pressures
46. Mason, G. and Morrow, N.R.: “Capillary Behaviour of a Perfectly
relies on equating the curvature, 1/r d , of the AM’s to the curvature
Wetting Liquid in Irregular Triangular Tubes,” J. Colloid and Interface
of the invading interface. If the AM’s were displaced a small
Science (1991) 141, 262.
distance dx, the work of the displacement must be balanced by the
47. Hilfer, R. and Øren, P.E.: “Dimensional Analysis of Pore Scale and change in surface free energy,
Field Scale Immiscible Displacement,” Transport in Porous Media
(1996) 22, 53. Pc Aeff dx 5 ~Low gow 1 Los gos 2 Los gws !dx, . . . . . . . . . . . . (A-1)
48. Wilkinson, D. and Willemsen, J.F.: “Invasion Percolation: A New Form
of Percolation Theory,” J. Physics A (1983) 16, 3365. where the subscript s 5 the solid, A eff 5 the effective area occupied
49. Chandler, R. et al.: “Capillary Displacement and Percolation in Porous by oil, L os 5 the length of the solid wall in contact with oil, and L ow
Media,” J. Fluid Mechanics (1982) 119, 249. 5 the perimeter length of the AM’s. From Young’s equation, g os
50. Mayer, R.P. and Stowe, R.A.: “Mercury Porosimetry-Breakthrough
2 g ws 5 g ow cos u r , and Eq. A-1 simplifies to
Pressure for Penetration Between Packed Spheres,” J. Colloid and Pc 1 Low 1 Los cosur Leff
Interface Science (1965) 20, 893. 5 5 5 . . . . . . . . . . . . . . . . . . . (A-2)
gow rd Aeff Aeff
51. Princen, H.M.: “Capillary Phenomena in Assemblies of Parallel Cyl-
inders. I. Capillary Rise Between Two Cylinders,” J. Colloid and The curvature of the invading interface at the condition for invasion
Interface Science (1969) 30, 60. is, thus, given as the ratio of the effective perimeter to the effective
52. Princen, H.M.: “Capillary Phenomena in Assemblies of Parallel Cyl- area. A eff, L os , and L ow are readily determined from elementary
inders. II. Capillary Rise in Systems With More Than Two Cylinders,” geometry and are given by
J. Colloid and Interface Science (1969) 30, 359.

OF S DG
3
53. Princen, H.M.: “Capillary Phenomena in Assemblies of Parallel Cyl- cos ur cos~ur 1 bi ! p ur 1 bi
inders. III. Liquid Columns Between Horizontal Parallel Cylinders,” J. Aeff 5 A 2 r2d 2 12
Colloid and Interface Science (1970) 34, 171.
sin bi 2 90
i51
54. Kovscek, A.R., Wong, H., and Radke, C.J.: “A Pore-Level Scenario for
the Development of Mixed Wettability in Oil Reservoirs,” AIChE J. r2
5 2 r2d S1 , . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (A-3)
(1993) 39, No. 6, 1072. 4G

SPE Journal, December 1998 335


O
3 These equations constitute a nonlinear system of equations that can
r cos~ur 1 bi ! r
Los 5 2 2rd 5 2 2rd S2 , . . . . . . . (A-4) be solved numerically for the radius of curvature r pt and, thus, the
2G sin bi 2G threshold capillary pressure P c 5 g /r pt . The maximum advancing
i51
angle at which spontaneous imbibition occurs is given by Eq. 7 and

O
2prd
3 corresponds to the limit when L eff 5 0. If u a . u a, max, the threshold
and Low 5 ~90 2 ur 2 bi ! 5 rd S3 . . . . . . . . . . . . . (A-5) capillary pressure may be determined from Eqs. 5 or 8.
180
i51

After a bit of algebra, we obtain a quadratic expression for r d that


Snap-Off. If u a , 908 2 b 1 , snap-off occurs when the advancing
has the solution
AM in the sharpest corner meets one of the other AM’s. For small
r cos ur ~21 6 Î1 1 4GD/cos2 ur ! values of u a , all the AM’s advance toward the center of the pore
rd 5 , . . . . . . . . . . . . . (A-6) space at u a and snap-off occurs when the AM’s in the two sharpest
4GD
corners meet. This occurs at a threshold capillary pressure of
where D 5 S 1 2 2S 2 cos u r 1 S 3 . The radius of curvature, r d , is
given by the valid root (smaller radius than the inscribed radius r)
of Eq. A-6 and may be expressed as Pc 5
g
r S
cosua 2
2 sin ua
cot b1 1 cot b2
. D . . . . . . . . . . . . . . . . (A-14)
r r
5 Pc 5 cos ur ~1 1 2 ÎpG! Fd ~ur , G!, . . . . . . . . . . (A-7)
rd gow For larger values of u a , the AM in the largest corner may remain
pinned, whereas the two other AM’s advance at u a . Snap-off then
where the function F d ( u r , G) is given by occurs either when the two advancing AM’s meet or when the AM
1 1 Î1 1 4GD/cos2 ur in the sharpest corner meets the pinned AM. The threshold capillary
Fd ~ur , G! 5 . . . . . . . . . . . . . . . . . (A-8) pressure for the first case is given by Eq. A-14, whereas the
1 1 2 ÎpG threshold capillary pressure in the second case is
In general, F d ( u r , G) is dependent on the particular corner angles
and is not universal for a specific G. However, if AM’s are present
in all the corners, the expression for D becomes Pc 5
r S
g cos ua cot b1 2 sin ua 1 cos uh3 cot b3 2 sin uh3
cot b1 1 cot b3
. D
D5p 12 S ur
60 D
1 3sinur cosur 2
cos2 ur
4G
. . . . . . . . . . . . . (A-9) . . . . . . . . . . . . . . . . . . . . . . . . . . (A-15)

In this case D is dependent only on u r and F d ( u r , G) is universal The event that actually takes place is the one occurring at the
for a particular G. highest capillary pressure. Because the hinging angle depends on
the prevailing capillary pressure, the above equation cannot be
Piston Type. If there is contact angle hysteresis, the threshold solved analytically. If u a . 908 2 b 1 , the curvatures of the AM’s
capillary pressure for piston type invasion during water injection is are negative and the threshold capillary pressure is given by
different from that during drainage. The invading interface enters Eq. 10 or 11.
the throat after the curvature is lowered sufficiently that u a is
reached. The oil/water interfaces of the AM’s in the throat remain
pinned at the position L b established at the end of the primary SI Metric Conversion Factors
drainage, i.e., bar 3 1.0* E105 5 Pa
cos~ur 1 bi ! cp 3 1.0* E203 5 Pazs
Lbi 5 rpd , . . . . . . . . . . . . . . . . . . . . . . . . . . (A-10) °F (°F232)/1.8 5 °C
sin bi
*Conversion factors are exact. SPEJ
where r pd 5 g /P c, max and L bi 5 the distance between the pinned
contact line and the corner i. As the capillary pressure drops, the
hinging angle of the pinned AM’s adjusts to give the same curvature
as for the invading interface. Provided that u a is not too large, the Pål-Eric Øren is at Statoil Research Centre in Trondheim, Nor-
way. e-mail: peoe@statoil.com. His research interests include
invading interface meets these AM’s at zero contact angle. The
physics of multiphase flow in porous media and reservoir engi-
radius of curvature r pt of the AM’s may be calculated by equating neering. He holds a BS degree in chemical engineering from
r pt to A eff /L eff . The effective area A eff is given by Eq. A-3, but with the U. of Michigan and a PhD degree in petroleum engineering
r pt substituted for r d and u r replaced by the hinging angle from the U. of New South Wales. Øren served as a member of

S D
rpd the Annual Meeting Program Committee and has served on
uh, i 5 cos21 cos~ur 1 bi ! 2 bi . . . . . . . . . . . . . . . . (A-11) Annual Meeting Technical Committees. Stig Bakke is a geolo-
rpt gist at Statoil Research Centre in Trondheim. e-mail
stiba@statoil.com. His research interests include microscale res-
The effective perimeter L eff is given by ervoir description, numerical modeling of geological pro-

S D
cesses, and 2D and 3D image analysis. Before joining Statoil, he

O O
3 3
r 2prpt worked at IKU in Trondheim. Bakke holds an MS degree in
Leff 5 cosua 22 Lbi 1 ai , . . . . . . . . . (A-12) geology from Norwegian Technical U. Ole Jacob Arntzen is a
2G 180
i51 i51 research scientist at Statoil Research Centre in Trondheim.
e-mail oja@statoil.com. His research interests include mul-
where the angle a i is given by tiphase flow in porous media, probability, and statistics. Be-

ai 5 sin21 S Lbi sinbi


rpt
. D . . . . . . . . . . . . . . . . . . . . . . . . . . (A-13)
fore joining Statoil, Arntzen worked at the Norwegian De-
fense Research Establishment. He holds an MS degree in
applied mathematics.

336 SPE Journal, December 1998

You might also like