Sie sind auf Seite 1von 13

See

discussions, stats, and author profiles for this publication at: http://www.researchgate.net/publication/257785775

A simple geometric model of sedimentary rock


to connect transfer and acoustic properties
ARTICLE in ARABIAN JOURNAL OF GEOSCIENCES MARCH 2013
Impact Factor: 1.22 DOI: 10.1007/s12517-013-0863-z

3 AUTHORS, INCLUDING:
Gabor Korvin

Abdulazeez Abdulraheem

King Fahd University of Petroleum and Min

King Fahd University of Petroleum and Min

47 PUBLICATIONS 196 CITATIONS

29 PUBLICATIONS 53 CITATIONS

SEE PROFILE

SEE PROFILE

Available from: Gabor Korvin


Retrieved on: 23 August 2015

Arab J Geosci (2014) 7:11271138


DOI 10.1007/s12517-013-0863-z

ORIGINAL PAPER

A simple geometric model of sedimentary rock to connect


transfer and acoustic properties
Gabor Korvin & Klavdia Oleschko & Abdulazeez Abdulraheem

Received: 13 April 2012 / Accepted: 25 January 2013 / Published online: 12 February 2013
# Saudi Society for Geosciences 2013

Abstract A simple rock model is presented which reproduces the measured hydraulic and electric transport properties of sedimentary rocks and connects these properties with
each other, as well as with the acoustic propagation velocities and elastic moduli. The model has four geometric
parameters (average coordination number Z of the pores,
average pore radius r, average distance between nearest
pores d, and average throat radius ) which can be directly
determined from the measured porosity , hydraulic permeability k, and cementation exponent m of the rock via simple
analytic expressions. Inversion examples are presented for
published sandstone data, and for cores taken from Saudi
Arabian, Upper Jurassic and Permian carbonate reservoirs.
For sandstone, the inversion works perfectly; for carbonates,
the derived rock model shows order-of-magnitude agreement with the structure seen in thin sections. Inverting the
equations, we express the transfer properties , k, and m as
functions of r, d, , and Z. Formulae are derived for the bulk
density Db, formation factor F, and P-wave velocity in terms
of the proposed geometrical parameters.

G. Korvin (*)
Earth Sciences Department and Reservoir Characterization
Research Group, King Fahd University of Petroleum and Minerals,
Dhahran, Saudi Arabia
e-mail: gabor@kfupm.edu.sa
K. Oleschko
Centro de Geociencias, Universidad Nacional Autnoma de
Mxico (UNAM), Km. 15.5 Carretera Quertaro-San Luis Potos,
C. P. 76230, Juriquilla, Quertaro, Mexico
e-mail: olechko@servidor.unam.mx
A. Abdulraheem
Petroleum Engineering Department, King Fahd University,
Dhahran, Saudi Arabia
e-mail: aazeez@kfupm.edu.sa

Keywords Digital rock physics . Petrophysics . Pore-space


model . Rock physics . Pore structure inversion

Introduction
The most exciting recent development in petrophysics has
been the emergence of digital rock physics (Andr et al
2012; Arns et al 2004; Boylan et al, 2002; Dong 2007;
Dong et al. 2008; Dvorkin et al. 2009, 2011; Kalam et al.
2011; Kayser and Ziauddin 2006; Keehm 2003; Knackstedt
et al. 2007, 2009a, b, c; Peng et al. 2012; Rassenfoss 2011;
Saenger et al. Sok et al. 2009; Sorbie and Skauge 2011;
Touati et al. 2009; Zhang and Knackstedt 1995, etc.). In
digital rock physics (DPR), one reconstructs the pore space
of a small (less than 1 cm3) piece of rock by computerized
X-ray tomography, digitizes the pore space, and then numerically simulates various physical processes to obtain
such macroscopic rock properties as permeability, electric
conductivity and elastic moduli (direct quote from Andr et
al. 2012). What is not always emphasized in DPR literature
(see, however, Dvorkin et al. 2011) is that the tiny fragment
of rock selected for analysis must be a statistically representative realization of all possible random rock structures
that could result from the same geologic processes of deposition, compaction, and diagenesis.
As the detailed pore geometries of any two cuttings from
the same rock body are obviously different, there exist
infinitely many equivalent pore structures which would
lead, through computer simulation or laboratory measurement, to the same set of macroproperties. Among all different microstructures which would result in the same
petrophysical macroproperties, there must be a simple effective pore-space model corresponding to that set of macroproperties. The search for this simple effective rock model
had been the key idea of this study: we asked ourselves how

1128

Arab J Geosci (2014) 7:11271138

to find a simple pore-space model, with as few degrees of


freedom as possible, that would correspond to a set of
computed or measured macroscopic rock properties. This
idea is close to Biswals model (Biswal et al. 1999, 2009;
Widjajakusuma et al. 1999) that embedded the microtomographically derived pore space into three stochastic pore
models: two geostatistically constructed and one based on
simulated sedimentary processes.
The geometric rock model what we propose in order to fit
macroscopic properties is a simplified version of Doyens
model (1987). It consists of identical spherical pores
connected with cylindrical tubes of the same radius (as in
Akbar 1993; Akbar et al. 1994). The model has four geometric parameters (average coordination number Z of the
pores, average pore radius r, average distance between
nearest pores d, and average throat radius ).
This rock physics to pore structure inversion is discussed
in Rock physics to rock texture inversion section. In the
Inversion of hydraulic and electric transfer properties to
pore geometry section, we review the direct inversion of
hydraulic and electric transport properties into pore geometry, using the pore model of Doyen (1987) as example. In
the Inversion of hydraulic and electric transfer properties to
pore geometry section, we introduce the equivalent pore
model and express the four geometric parameters from the
measured porosity , hydraulic permeability k, and cementation exponent m of the rock. The relevant Eqs. (2123) or
(2426) are the most important results of this work.
Numerical results, based on these equations, will be given
in the Rock property to pore geometry inversion examples
section (Tables 1, 2, and 3 and Figs. 1, 2, and 3) for six
published (Jorgensen 1988) sandstone data and for our
measurements on 26 carbonate samples from the Upper
Jurassic Arab-D and the Permian Khuff formations (see
Halawani 2000; Meyer et al. 2000; Stenger et al. 2003, for
the geology).
In the Use of the geometric model to connect pore
geometry with transfer- and elastic properties section,
Eqs. (2426) will be inverted, and we express the transfer
properties , k, and m as functions of r, d, , and Z. In the
Computation of the transfer properties and some other rock
Table 1 Comparison of published (Jorgensen 1988)
and computed (by Eq. 25) pore
throat radii

properties from the equivalent rock model section, the bulk


density Db, formation factor F, and P-wave velocity are
expressed in terms of the proposed geometrical parameters,
for the case of complete saturation. The case of incomplete
saturation (The case of incomplete saturation section) is
somewhat special because it needs a further independent
parameter, the saturation exponent n. We do not attempt,
in this paper, to relate n to the pore geometry (in this respect,
see Knackstedt et al. 2007).
Finally, the Discussion and concluding remarks part
gives a critical summary of the model and its applications
and lists the several, still open problems.

Rock physics to rock texture inversion


Inversion of hydraulic and electric transfer properties
to pore geometry
The prediction of the interior structure of sedimentary rocks
from measured bulk physical properties is a nonlinear inverse problem where the number of unknowns is much
larger than the number of measurements. Such problems
allow an infinite number of solutions, out of which solutions
with small variance can be obtained by the Tikhonov regularization method (Lamm 1997); while those solutions
which are most likely to occur in nature are given by the
maximum entropy method (MEM, see Lifshitz and
Pitaevskii (1980) where the technique is explained through
the derivation of Boltzmans barometric equation or
Korvins MEM treatment (1984) of shale compaction).
The first application of MEM to rock physical inversion
was due to Doyen (1987) who determined the statistical
crack geometry in igneous rocks from measured hydraulic
conductivity and/or dc electric conductivity values at a
series of confining pressures. His rock model consisted of
a random distribution of spherical pores connected with
nearby pores by tubes (throats). At a given reference
pressure P=P0, the pore radii r are distributed according to
some probability density function (pdf) n(r); the throat
length l is distributed according to a pdf n(l); the throats

Lithology

k (mD)

Hydraulic radius published


in Jorgensen 1988 r (m)

Geometric model
parameter (m)

Berea sandstone
Pyrex 1
Pyrex 2
Nichola Buff sandstone
Eocene sandstone
Pennsylvania sandstone 1
Pennsylvania sandstone 2

0.22
0.37
0.29
0.20
0.22
0.16
0.21

1.624
1.490
1.474
1.569
1.694
1.635
1.644

890
8.1
3,900
230
340
180
390

7
0.38
12
4.2
4.4
4.1
5.5

8.27
0.53
14.2
4.3
5.26
4.47
5.65

Arab J Geosci (2014) 7:11271138

1129

Table 2 Input parameters for the rock physics to pore structure inversion, carbonate samples
k Permeability (mD)

Porosity

m Cementation exponent

n Saturation exponent

F Formation factor

Dol Wkst/Pkst
Dol Wkst
Dol Wkst
Dol Wkst
Dol Mdst
Dol Mdst
Sucrosic Dol
Sucrosic Dol

0.547
1.41
1.3
1.28
0.071
0.108
56.120
125.0

0.1731
0.2258
0.2137
0.2227
0.1037
0.1203
0.2240
0.2289

1.764
1.755
1.742
1.870
1.900
1.919
1.927
1.994

NA
NA
NA
NA
NA
NA
1.236
1.158

22.06
13.62
14.71
16.58
74.05
58.21
17.86
18.93

Sucrosic Dol
Sucrosic Dol
Sucrosic Dol
Sucrosic Dol
Anhyd Dol
Sucrosic Dol
Sucrosic Dol
Sucrosic Dol
Oomoldic Grst
Oomoldic Grst
Oomoldic Grst
Sucrosic Dol
Sucrosic Dol
Sucrosic Dol
Sucrosic Dol
Oomoldic Grst
Oomoldic Grst
Dol Lst

3.23
122
94.8
0.046
0.003
247.0
2.73
117.0
0.007
0.021
0.002

0.1563
0.3187
0.1990
0.0751
0.0315
0.2623
0.2059
0.2104
0.1257
0.1126
0.0686
0.2796
0.2457
0.2157
0.2965
0.1720
0.1600
0.1820

1.775
1.895
1.913
1.670
1.400
1.836
2.034
1.964
2.117
2.092
1.993
1.972
1.982
1.940
1.947
2.324
2.420
2.009

1.419
1.687
1.464
0.862
NA
1.257
1.251
1.224
NA
NA
NA
1.566
1.452
1.426
1.245
2.195
2.004
2.098

26.95
8.73
21.96
75.49
126.42
11.67
24.89
21.35
80.59
96.32
177.50
12.34
16.15
19.61
10.67
59.77
84.33
30.66

Well

Sample

Lithology

A
A
A
A
A
A
A
A

10-A
15-A
19-A
28-A
41-A
45-A
113-A
114-A

A
A
A
A
A
A
A
A
B
B
B
B
B
B
B
B
B
B

122-A
149-A
247-A
253-A
257-A
268-A
272-A
273-A
42-B
59-B
71-B
274-B
278-B
282-B
283-B
287-B
294-B
355-B

111.0
33.7
58.2
0.103
0.05
93.0

have elliptic cross section with random semiaxes b and c.


The pdf of the throat shape at the reference pressure is n(,
c) where is the aspect ratio =b/c. Using elasticity theory
for the deformation of voids under pressure P (Bernabe et al.
1982; Kuster and Toksz 1974; Zimmerman 1991), the
pressure-dependence of the pdfs r(P), c(P), l(P), and (P)
can be determined. At every pressure step, the pdf n(, c)
satisfies the self-consistency equation (Kirkpatrick 1973).
X

na; c

gh* P gh P; a; c; l


0;
gh P; a; c; l Z2 1 gh* P

1a

na; c

gh* P gh P; a; c; l


0;
gh P; a; c; l Z2 1 gh* P

1b

a;c

X
a;c

for all P, where Z is average coordination number (number of


throats incident to a pore), ge* P; gh* P are the measured bulk
electric and hydraulic conductances at pressure P, and gh =gh
(P,,c,l) and ge =ge (P,,c,l) are theoretical conductances of an
originally (, c)-shaped throat of length l subjected to pressure

P. Doyen (1987) solved the system (1a, b) for the pdfs of pore
size n(r), of throat size n(l), and of throat ellipticity n(, c)
with the maximum entropy method.
In our study, we simplified Doyens model by assuming
an equivalent pore space with as few parameters as possible.
We had been encouraged by the success of the ColeCole
equivalent circuit model of the induced potential effect in
rocks (Cole and Cole 1941; Pelton et al. 1978; Telford et al.
1990) or the three-parameter Thomeer model of mercury
pressure injection (Thomeer 1960, 1983; Clerke 2003),
which are classic examples that equivalent rock models with
only three degrees of freedom can realistically describe the
behavior of a complex geologic medium.
Our geometric rock model consists of spherical pores of
the same size, connected with cylindrical tubes (as in Akbar
1993; Akbar et al. 1994). In the next section, we shall define
the model and derive its parameters from transfer properties
measured at a single pressure step rather than using a series
of pressures as in Doyen (1987). Section Rock property to
pore geometry inversion examples will show numerical
examples for this new kind of rock to texture inversion.

1130

Arab J Geosci (2014) 7:11271138


Pore size

Table 3 Equivalent rock model computed from Table 2

A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
B
B

10-A
15-A
19-A
28-A
41-A
45-A
113-A
114-A
122-A
149-A
247-A
253-A
257-A
268-A
272-A
273-A
42-B
59-B

4.62
4.65
4.70
4.30
4.22
4.18
4.16
4.01
4.58
4.23
4.19
4.99
7
4.39
3.93
4.07
3.79
3.83

1.9
1.99
2.04
2.34
2
2.04
16.85
27
5.51
12.81
26.14
1.48
0.49
23.48
5.13
28.79
0.73
1.46

0.25
0.34
0.34
0.34
0.12
0.14
2.28
3.41
0.63
2.76
3.13
0.1
0.04
4.30
0.54
3.42
0.03
0.06

6.08
6
6.2
6.92
7.22
7
49.22
77.5
18.05
35.06
78.71
6.05
2.75
67.47
15.06
84.85
2.44
5.04

B
B
B
B
B
B
B
B

71-B
274-B
278-B
282-B
283-B
287-B
294-B
355-B

4.01

0.82

0.02

3.32

4.04
4.13
4.11
3.51
3.41
3.98

21.98
14.23
10.80
2.31
2.22
35.66

3.09
1.80
2.01
0.12
0.09
3.31

62.33
41.84
29.65
6.92
6.76
108.33

r, , and d are given in microns

Pore size
Calculated r (microns)

Well Sample Z Coordination r Mean


Mean throat d Distance
number
pore
radius ()
between
radius ()
pores ()

1000
800
600
400
200
0
0

200

400

600

800

1000

Estimated r (microns)

Fig. 2 Crossplot of the microscopically observed average pore radii


and the model pore radii r (computed by Eq. 24)

The equivalent rock model


Assume that the porosity , permeability k, and cementation
exponent m (figuring in Archie's law, 1942; modern treatment: Perez-Rozales 1982; Worthington 1993; Glover 2009)
are measured for some sedimentary rock at a reference
pressure P. We will construct an equivalent rock model
which has the same measured , k, and m values. The model
consists of a connected system of pores and throats distributed within a homogeneous rock matrix. Pores are spheres
of radius r, and they are distributed in the rock matrix in
such a way that the Euclidian distance between two nearest
pores is d. A pore is connected on the average by Z neighboring pores with tortuous cylindrical pipes (throats) of
radius . The parameters of the equivalent rock model (r,

d
1000

Calculated d (microns)

d
800

600

400

200

Fig. 1 Microscopic image from the sucrosic dolomite sample 268-A


from the Permian Khuff formation. The model parameters (Z=4.39; r=
23.48 , = 4.30 , d=67.47 ) determined from m, k, and using
Eqs. (2) and (2426) show order-of-magnitude agreement with the
geometry of the rock

200

400

600

800

1000

Estimated d (microns)

Fig. 3 Crossplot of the microscopically observed average throat


lengths and the model throat lengths d (computed by Eq. 26)

Arab J Geosci (2014) 7:11271138

1131

d, Z, and ) can be uniquely determined as functions of , k,


and m if we make the following seven assumptions:
(1) The equivalent rock model is only applicable for macroscopically homogeneous and isotropic parts of the
reservoir. In case of heterogeneity or anisotropy, the
model parameters should be varied and then smoothly
fitted together. Optimally, the model is applied to plugsized or cutting-sized rock fragments. In such a small
homogeneous and isotropic rock volume, the theoretical porosity and hydraulic permeability of the equivalent rock model are assumed to be the same as the
measured and k values;
(2) The resistivity for 100 % saturation satisfies Archies
law bulk am w (resistivities are denoted by ,
densities by D to avoid confusion);
(3) The hydraulic permeability of porous rocks obeys the
KozenyCarman equation (Kozeny 1927; Carman
1937; Walsh and Brace 1984) k 1b  3  S 21  t12 ;
spec

(4) In the KozenyCarman equation, C is the hydraulic


tortuosity. We assume that the rock is not very tight, so
that the electric and hydraulic tortuosity values can be
taken as equal. For low-permeability rocks, the hydraulic
permeability can be an order of magnitude higher than
the electric one (Zhang and Knackstedt 1995; Duda et al.
2011; Worthington 2011), or, in case of clayey rocks, it is
divergent at the percolation threshold (Korvin 1992a, b).
(5) The average coordination number and the cementation
exponent m are related as Z m2m1 (Yonezawa and
Cohen 1983);
(6) The hydraulic tortuosity C and formation factor F are
related as C=F (Perez-Rozales 1982; Glover 2009).
(7) During hydraulic flow, the total porosity consists of a
stagnant and a flowing part, =s +f where, according to Perez-Rozales (1982), f =m.
We note that for the determination of (r, d, Z, and ), the
measured resistivity value and Archies saturation exponent n
are not needed.
To relate the transfer properties , k, and m as functions
of (r, d, , and Z), we first use the cementation exponent m to
express the average coordination number Z and the hydraulic tortuosity . According to the self-similar effective medium approximation for granular media (Yonezawa and
Cohen 1983), the average coordination number and the
cementation exponent are related as
Z

2m
;
m 1

Using Archies equation,


t F 

a
a
m
m

where if a is not known, one usually takes its default values:


i.e., a=0.62 for unconsolidated sand; a=0.81 for consolidated sand; a=1 for carbonate formations (Schlumberger
1991). According to Glover (2009), a=1 is the only proper
choice because otherwise lim bulk a w , which is
!1

counterintuitive.
The hydraulic permeability of porous rocks is, by the
KozenyCarman equation (Walsh and Brace 1984):
k

1 3 1
1
  2  2
b
t
S spec

where b is a geometric factor equal to 2 for cylindrical


throats and 3 for cracks and Sspec is the specific surface. If
we measure distances in millimeter, surface areas in square
millimeter, and permeability in millidarcy, Eq. (5) becomes
(Korvin 1992a, b)
k md

1 3
1
1
9
 
2  2  10 :
1
b
t
S spec mm

Taking b=2 and using Eq. (3), the specific surface Sspec is
r
1:5  104:5

4:5 1
10  
S spec 0:5
7
0:5
F
2k
2 t  k
Suppose that there are altogether N uniformly distributed
spherical pores of radius r within a 1 mm1 mm1-mmsized cube V of the rock. Each pore occupies a separate
domain of average volume N1 , that is, the mean distance
between two nearby pores is
1
:
d p
3
N

On the average, each pore is connected with Z neighboring pores by means of tortuous throats with circular cross
section of radius . Denote the total length of these throats
by L. We write up two equations to determine N, L, r, and .
The first equation expresses that the total pore space in rock
volume V is built up of N spherical pores and of a total
length L of cylindrical throats:
N

4 r3 p
L  d 2 p:
3

while (see Perez-Rozales 1982; Glover 2009) tortuosity is


given by

The second equation states that the total pore surface within
the microscopic cube (i.e., the specific surface) is the sum of
the spherical pore surface areas minus the holes on this surface
cut out by the throats, plus the areas of the throat walls:

t F  :

S spec N  4r2 p

NZ d 2 p 2pd  L

10

1132

Arab J Geosci (2014) 7:11271138

As <<r, we neglect the term which is quadratic in :


S spec N  4r2 p 2pd  L:

11

To eliminate L from Eqs. (9) and (11), note that


LN

Z
 d  t:
2

12

The meaning of the factor 1/2 is that every throat belongs


to two pores, the C factor expresses the tortuosity of the pipe
connecting two pores a distance d apart. By Eqs. (2), (3),
and (8),
LN

N 2m1 t
Z
m
2
 d  t mp
 F
N3 
3
2
m
1
2 N

13

Using Eq. (13) in Eqs. (9) and (10):


N

4r3 p
m
2
d2p N 3 
F
3
m 1

S spec N  4r2 p 2d p N 3 

 F:

14

15

4 r3 p
m
2
d2 p  N 3 
  F s f :
3
m 1
16

According to Perez-Rozales (1982),


f m

17a

m :

17b

Matching the two terms in Eq. (14) with Eqs. (17a) and
(17b), respectively, and solving for r and ,
r
m 1
3 3
r
 1
4p
N3

r
m 1 1
1

 m 1  1 :
m
pF
N3

1
1
N3

, substitute Eqs. (18) and (19) to Eq. (15) in

place of r and , and use Eq. (7) to obtain


p
1
2k  F
4:5 2p 
p

1 10

N3
" 
 2 r#
3 m 3
1
m

 m1  F :

 2
4p
p m 1
20
Finally, from Eqs. (18) and (19),
r
p
m 2p  2k  F
4:5 3 3
p

r 10

4p

" 
 2 r#
3 m 3
1
m

 2

 m1  F
4p
p m 1

21

s
p
m 1 m 1 2p  2k  F
p

d 10 4:5

m
pF
m 
" 
 2 r#
3 m 3
1
m

 m1 F ;

 2
4p
p m 1
22

The physical meaning of Eq. (14) is that the total


porosity is the sum of a hydraulically inactive stagnant
part (consisting of the pores) and of a part where the
fluid is actually moving (the flowing porosity constituted by the throats):
N

To find

18

19

and from Eqs. (8) and (20),


p
4:5 2p  2k  F
p
d 10


" 
 2 r#
3 m 3
1
m

 m1 F :

 2
4p
p m 1

23

Equations (21) to (23), combined with Eq. (2) completely


solve the problem of expressing (r, d, Z, and ) as functions
of , k, and m.
If we accept Glovers argument (2009, his Eq. 7) that one
must have a=1 in Archies second equation, then F 1m ,
and Eqs. (2123) simplify to
r
p
3 3
m 2p  2k
4
 m p
r 10 
4p
 10
" 
r#
2

3 m 3
1
m


24

 2
4p
p m 1
s
p
m 1 2m 1 2p  2k
4
d 10

 m p
m
p
 10
" 
r#
2

3 m 3
1
m



 2
4p
p m 1

25

Arab J Geosci (2014) 7:11271138

p
2p  2k
p

m  10
" 
 2 r#
3 m 3
1
m

 :

 2
4p
p m 1

d 10

1133

of the cases, it could not be determined from the micrographs by visual inspection.

26

Rock property to pore geometry inversion examples


Numerical results, based on Eqs. (5) and (2426) are given
in Tables 1, 2, and 3 for seven published sandstone samples
(taken from Jorgensen 1988) and for our laboratory measurements and microscopy of 26 carbonate samples from the
Upper Jurassic Arab-D and the Permian Khuff formations
(Halawani 2000; Meyer et al. 2000; Stenger et al. 2003).
Note that if k is in millidarcy in these equations, r, , and d
will be obtained in millimeter. For convenience, however, in
Tables 1 and 3, these values are shown in microns. For the
seven sandstone samples (Table 1) only one geometric
property, the hydraulic radius had been published, and this
is indeed in near-perfect agreement with the throat radius
computed by Eq. 25. For carbonate rocks, we cannot expect
such a close agreement because most carbonates have double or triple porosity (Warren and Root 1963; Abdassah and
Ershaghi 1986; Lucia 1998; Ahr 2008; Pal 2012; Pulido et
al. 2007, 2011; etc.), as carbonate pore space consists of
inter-matrix pores (for which our model is applicable), vugs,
and fractures. Also, for triple-porosity rocks, the value of
the transfer factors (also called shape factors) between
different kinds of pores (first introduced by Warren and
Root 1963; more recent developments: Lim and Aziz
1995; Hassanzadeh and Pooladi-Darvish 2006) are still
poorly understood. The geometric model should be modified for triple-porosity rocks; this will be discussedas a
future research problemin Discussion and concluding
remarks section. To check whether for carbonate rocks
the computed model parameters are not physically meaningless, that is, they have the same order of magnitude as the
microscopically determined pore radius r, average distance
between nearest pores d, and average throat radius , we
analyzed the micrographs of 26 carbonate samples. Figure 1
is a typical example; it is the microscopic image of a thin
section from sample 268-A (a sucrosic dolomite from the
Permian Khuff formation). Visual inspection shows that the
computed values (Z=4.39; r= 23.48 , = 4.30 , d=
67.47 ) are reasonably close (in an order-of-magnitude
sense) to the geometry of the rock. Further results are given
as in Tables 2 and 3 and as crossplots in Figs. 2 and 3. In
spite of the large scatter, there is an order-of-magnitude
agreement between computed and observed results, and
we are optimistic that the model can be extended and finetuned for carbonates. We note that Tables 2 and 3 and the
crossplots do not contain the throat radius because, in most

Use of the geometric model to connect pore geometry


with transfer and elastic properties
Computation of the transfer properties and some other rock
properties from the equivalent rock model
In what follows, the physical units will always be noted by
square brackets, []. The geometric model parameters r (in
millimeter), d (in millimeter), (in millimeter), and Z (positive real number) can be used in a straightforward way to
compute m (cementation exponent), (porosity, 01), k
(permeability), and D (density). We also need a which is the
lithology-dependent dimensionless constant in Archies law
bulk am w . Accepting the empirical laws of Archie
(1942) and Wyllie et al. (1956, 1958), the formation factor
F, rock resistivity , density D, and P-wave velocity VP can
be estimated as follows.
The cementation factor is obtained from Eq. 2:
m

Z
Z

27

From Eq. (8), the number of spherical pores in a small


volume V of 1 mm3 is
N

1
;
d3

28

that is, from Eq. (19),


"

m
m

 2 # 2m1 1
d
 pa 
:
1
d

29

This immediately yields the dimensionless formation factor from Archies law:
0

"

 2 #
a @
m
d
 pa 
F m

m 1
d

1
2m 1

1
A;

30a

and the tortuosity, by Eq. (3), (based on Perez-Rozales 1982;


Glover 2009):
0
t : F

a
m

@ a

"

m
m

 2 #
d
 pa 
1
d

m 1
2m 1

1
A:
30b

In case of fully saturated rocks, Eq. (29) immediately


gives the bulk density, bulk resistivity, and (by

1134

Arab J Geosci (2014) 7:11271138

Wyllies empirical equation) the approximate P-wave


velocity as

(S). By Archies laws (35a, b), we would have then S am


S am S n w , from where

Dbulk Dfluid 1

S S

bulk

Dmatrix

a
m  fluid

V P;Wyllie

V fluid

31

32

1
:

V1matrix

33

In Eqs. (3133), all densities are in kilogram per cubic


meter, resistivities in ohm meter, and velocities in meter per
second. As well known (Mavko et al. 1998), Wyllies Eq. (33)
is a heuristic, approximate equation for <0.2 porosities. A
theoretically exact expression for P-wave velocity, based on
Biots theory (1956) and also valid in case of partial saturation, will be derived in The case of incomplete saturation
section (Eq. 44).
The hydraulic permeability k is first expressed in millidarcy units:


2m1
h
 d 2 i 2m
8
8
1
2m1
1
m
k millidarcy 510



510
a2
a2  m 1  pa  d
S 2spec


h
 d 2 i 2m1 m1 2
2
d
m
 4pd 3r 2pma

pa

m 1 d2 m 1
d

34a
For the further calculations, we convert the permeability
to SI units (in square meter, see Turcotte and Schubert 1982,
p 382):

k m2 9:8697  10

16

k millidarcy:

34b

The case of incomplete saturation


By the laws of Archie (1942), the resistivity of a rock,
whose pore space is only partially saturated with a conducting fluid of resistivity w, is
S

a
S
m

35a

where n is the saturation exponent. The saturation S is


normalized to lie between 0 and 1. In particular,
S1

m w

35b

How could partial saturation be incorporated to the


equivalent rock model? There are two possible approaches:
The first is to assume that the same rock model describes
electric conduction in case of partial saturation as well, but
instead of w, the fluid has a saturation-dependent resistivity

36

(special cases are (1)=w, (0)=). The second possibility


would be to assume that, in case of partial saturation, it is the
coordination number Z which depends on S, that is, only a
fraction of throats contains conducting brine, the rest contains
nonconducting hydrocarbon. By Eq. (2), this would force the
cementation exponent m to also depend on S and to satisfy
S

a
S
m

mS

w ;

37

which would imply the functional equation


mS m S n :

38

This model, however, is unphysical because for S0, it


implies mS ! 0 ; i:e: ; mS ! 1 ; because 0 < < 1,
which contradicts the observed fact (Mavko et al. 1998) that m
is usually close to 2 and rarely exceeds 3.
Thus, only Eq. (36) is feasible, and for a partially saturated rock, the resistivity is given by Eqs. (30a) and (36):
"
 2 # 2mm 1
m
d
S : a
pa
S n w :
39
m 1
d
Consider now the acoustics of the fully or partially,
saturated rock model. According to Biots theory (1956),
the saturated bulk modulus differs from the dry bulk modulus in a term K (, S):
K S K dry K ; S :
We estimate first the effective bulk modulus of the fluid,
through the Reuss (1929) average (Mavko et al. 1998):


S
1 S 1
K porefiller :

;
40
K fluid
K air
then we compute
K ; S

h
K solid 1

K solid

K dry
K dry
K solid

2

solid
KKporefiller

i;

41

and finally get


K S K dry K ; S :

42

All compressibilities in Eqs. (4345) are in Pascal=kilogram per meter per square second, the values of the moduli
Kdry, Ksolid, Kfluid, Kair must be specified. The density, in case
of partial saturation, is
D; S S Dfluid 1

S Dair 1

Dmatrix
43

Arab J Geosci (2014) 7:11271138

1135

The shear modulus in micropascal kilogram per meter


per square second of the rock is independent of saturation or
of the fluid properties. Recalling that

q
p
K 4

3
44
VP
D
D ; V Shear
and using quantities already calculated in Eqs. (42) and (43),
r
V P :

K S 4
3
D;S ;

V shear :

D;S :

45

Discussion and concluding remarks


The proposed equivalent geometric model of sedimentary
rocks belongs to the family of effective medium models
(modern examples: Kachanov 1994; Sayers and Kachanov
1995; Schubnel and Guguen 2003; Fortin et al. 2005; etc.),
its parameters (Z, r, , and d) can be easily derived from a
few measured rock properties (k, , and m). The converse is
also true: from the values (Z, r, , and d), one can calculate
the bulk rock properties (k, , and m). If the specific matrix
and fluid properties are also known, the elastic constants and
the P- and S-velocities can be calculated both for the fully
saturated case and for partial saturation. The dc resistivity
can be computed in case of complete saturation by Archies
law, but for partial saturation, we need a further rock property, the saturation exponent n. We note that, while in digital
rock physics, n can be estimated by simulation (Knackstedt
et al. 2007), in our model, we could not derive it in terms of
(Z, r, , and d) or (k, , and m) using physical arguments.
Also, while the parameters (Z, r, , and d) are sufficient to
find first the coordination number m (by Eq. 2) and then to
calculate single-phase permeability (using Eq. 34a), we
have not succeeded to describe relative permeabilities for
two-phase flow. This failure is possibly explained by the
computer simulation study of Sok et al. (2002), who
reported that, for a proper description of two-phase flow in
a pore network, one needs to specify network topology at
higher order than the coordination number, by using coordination number sequences (Grosse-Kunstleve et al. 1996).
The geometrical model implies that mechanical properties
(density, P- and S-velocities, elastic moduli) depend on the
pore coordination number Z, because all of them depend on
porosity , by Eq. (29) depends on m, and by Eq. (2), m is
related to Z. This is in accord with granular medium theories
(e.g., Brandt 1955; Walton 1987) where porosity, grain-tograin coordination number, and elastic properties are linked,
though the relation between grain-to-grain and pore-to-pore
coordination numbers still needs a separate careful study.
Through the entire study, we have heavily relied on the
KozenyCarman equation (Walsh and Brace 1984) k 1b

 3  S 21  t12 (Eq. 5). This is an empirical law which has


spec

been improved lately using percolation theory (Guguen


and Dienes 1989; Korvin 1992a; Benson et al. 2006a, b).
Even though empirical, the KozenyCarman Law is extremely useful and versatile. Its Sspec factor makes it possible
to incorporate pore size distribution in the model, while the
hydraulic tortuosity can describe the permeability behavior
at the percolation threshold (Korvin 1992a, b). Throughout
the paper, we have mixed empirical laws (Archies,
Wyllies, and that of KozenyCarman) with theoretical
ones (such as Kirkpatricks Eq. 1a, 1b or Biots theory),
and even the theoretical equations have empirical elements
(as the Reuss average, Eq. 40, hidden in Biots theory) or ad
hoc assumptions (such as the self-consistency assumption in
Kirkpatrick 1973; the maximum entropy assumption in
Doyen 1987; or the effective media approximation in
Yonezawa and Cohen 1983). Such a mixture of theoretical
and empirical equations naturally results in empirical equations of limited validity, and our main equations, Eqs. (21)
to (23), or their special cases (24) to (26) for a=1) should be
considered empirical, as the majority of equations of petrophysics and well log analysis are.
As seen in section Rock property to pore geometry
inversion examples, the geometric model works well for
sandstones; for carbonates, the resulting pore parameters are
not physically impossible, but they show only order-ofmagnitude agreement and very poor correlation with the
microscopic rock structure. We see no difficulty to similarly
model fracture porosities or vug porosities which are more
appropriate for carbonates, but the integration of the three
models into a single triple-porosity modelwhich will be
the next step of this researchis a great challenge.
Throughout the paper, we assumed the statistical homogeneity and isotropy of the rock volume, which might be true
for a small cutting analyzed in DPR, less true for a plugsized sample, and is an obviously invalid assumption on
reservoir scale. Issues of upscaling the model to reservoir
scale, making it heterogeneous and anisotropic, are among
the further tasks to be solved.
Acknowledgments The authors gratefully acknowledge the excellent conditions for research provided by their home institutions, the
King Fahd University of Petroleum and Minerals, Saudi Arabia, and
Universidad Nacional Autnoma de Mxico. Thanks are due to
Dr. Nabil Akbar who, many years ago, suggested the topic and
acquainted us with his 1993 Stanford Ph.D. thesis. Parts of the work
have been previously presented at two workshops: the Symposium
Arreglos de Fractura (UNAM, Mexico City, 25 February, 2002) and
at the Theoretical Physics Day (KFUPM, Dhahran, Saudi Arabia, 13
May, 2007). The work of Dr. G.K. and Dr. A.A. has been partly
supported by the King Abdulaziz City for Science and Technology
through project no. 08-OIL82-4, National Science Technology Innovation Plan. Dr. G.K. is grateful for the financial support from the
project no. 168638 SENERCONACYT-Hidrocarburos Yacimiento Petrolero como un Reactor Fractal (project leader Professor Klavdia

1136
Oleshko) which enabled him to visit the Research Group of Professor
Oleshko in Juriquilla, Quertaro, Mexico, and to work with her on
some critical aspects of this research.

Appendix: Notations
a

Dimensionless constant in Archies law


bulk am w
=b/c
Aspect ratio of the throats cross
sections in the model of Doyen (1987)
b
Geometrical constant in Eq. (5). For
circular tubes b=2
b, c
Semiaxes of the elliptical throats cross
section in the model of Doyen (1987)
d
Model parameter, average geometric
distance between two nearest pores
(millimeter)
Dfluid, Dmatrix
Specific densities (in gram per cubic
etc.
centimeter)

Model parameter, mean throat radius


(millimeter)
F
Formation factor in Archies law (1942)

Total porosity, fraction (01)


s, f
Stagnant, resp. flowing (hydraulically
effective) parts of the porosity (see
Eq. 16, and Perez-Rozales 1982; Glover
2009)
ge , g h
Electric, resp. hydraulic conductances
in the Kirkpatrick equation (Eq. 1)
ge* ; gh*
Bulk electric, resp. hydraulic
conductances in the Kirkpatrick
equation (Eq. 1)
k
Permeability (in square meter)
Kdry, Ksolid, Kfluid, Specific bulk moduli (in Pascal)
Kair etc.,
l
Throat length in the model of Doyen
(1987)
l
Average distance followed by the fluid

path (in millimeter), dl t
L
Total length of the throats in a unit
volume of rock (in millimeter)
m
Cementation exponent in Archies law
m(S)
Saturation-dependent cementation
exponent, Eqs. (37) and (38)

Shear modulus in Pascal


n
Saturation exponent (in Archies law)
n(r)
Pore radius probability distribution
function in the model of Doyen (1987)
n(, c)
Throat-shape probability distribution
function in the model of Doyen (1987)
N
Number of pores in a unit volume
(1 mm1 mm1 mm) of rock

Arab J Geosci (2014) 7:11271138

P, P0
r
100%, bulk, s,
w
(S)
S
Sspec [1/mm]
C
V
Vfluid, Vmatrix etc.
Z

Pressure in the model of Doyen (1987)


Model parameter, mean pore radius
(in millimeter)
Specific resistivities in ohm meter
Saturation-dependent resistivity in Eqs.
(36) and (37)
Saturation, between [0,1]
Specific surface (i.e., total surface area
of a unit volume of rock)
Hydraulic tortuosity
Rock volume for which the geometric
model holds
Specific P-wave velocities in meter per
second
Coordination number (average number
of throats emerging from one pore)

References
Abdassah D, Ershaghi I (1986) Triple-porosity systems for representing naturally fractured reservoirs, SPE Formation Evaluation,
April 1986, 113127
Ahr WA (2008) Geology of carbonate reservoirs. The identification,
description, and characterization of hydrocarbon reservoirs in
carbonate rocks. Wiley, Hoboken
Akbar NA (1993) Seismic signatures of reservoir transport properties
and pore fluid distribution, Ph. D. Thesis, Stanford University,
Stanford, CA
Akbar N, Mavko G, Nur A, Dvorkin J (1994) Seismic signatures of
reservoir properties and pore fluid distribution. Geophysics
59(8):12221236
Andr H, Combaret N, Dvorkin J, Glatt E, Han J, Kabel M, Keehm Y,
Krzikalla F, Lee M, Madonna C, Marsh M, Mukerji T, Saenger
EH, Sain R, Saxena N, Ricker S, Wiegman A, Zhan X (2012)
Digital rock physics benchmarksPart 1. Imaging and segmentation; Pt. 2. Computing effective properties. Comput Geosci
50:2532. doi:10.1016/j.cageo.2012.09.005
Archie GE (1942) The electric resistivity log as an aid in determination
of some reservoir characteristics. Trans AIME 146:5462
Arns C, Knackstedt MA, Val Pinczewski W, Martys NS (2004) Virtual
permeametry on microtomographic images. J Petrol Sci Eng
45(12):4146
Benson P, Meredith P, Schubnel A (2006) Examining the role of void
space fabric in the permeability development of crustal rock with
pressure. J Geophys Res 111: Art. No. B12203
Benson A, Schubnel A, Vinciguerra S, Trovato C, Meredith P, Young
RP (2006) Modelling the permeability evolution of micro-cracked
rocks from elastic wave velocity inversion at elevated hydrostatic
pressure. J Geophys Res 111: Art. No. B04202
Bernabe Y, Brace WF, Evans B (1982) Permeability, porosity, and pore
geometry of hot-pressed calcite. Mech Mater 1:173183
Biot MA (1956) Theory of propagation of elastic waves in a fluidsaturated porous solid. I. Low frequency range. J Acoust Soc Am
28:168178
Biswal B, Mauwart C, Hilfer R, Bakke S, ren PE (1999) Quantitative
analysis of experimental and synthetic microstructures for sedimentary rock. Physica A 273:452475

Arab J Geosci (2014) 7:11271138


Biswal B, ren P-E, Held RJ, Bakke S, Hilfer R (2009) Modeling of
multiscale porous media. Image Anal Stereol 28(1):2334
Boylan AL, Waltham DA, Bosence DWJ, Badenas B, Aurell M (2002)
Digital rocks linking forward modeling to carbonate facies. Basin
Res 14(3):401415
Brandt H (1955) A study of the speed of sound in solid granular media.
Trans ASME 22:479486
Carman P (1937) Fluid flow through a granular bed. Trans Inst Chem
Eng 15:150167
Clerke EA (2003) Beyond porosity-permeability relationshipsdetermining pore network parameters for the Ghawar Arab-D using the
Thomeer method. Geofrontier (Dhahran, Saudi Arabia) 1(3):12
17
Cole KS, Cole RH (1941) Dispersion and absorption in dielectrics. Part
1. Alternating current fields. J Chem Phys 9:341
Dong H (2007) Micro CT imaging and pore network extraction, Ph. D.
Dissertation, Imperial College, London
Dong H, Fjeldstad S, Alberts L, Roth S, Bakke S, ren P-E (2008)
Pore network modeling on carbonate: a comparative study of
different micro-CT network extraction methods. Soc. Core Anal.
Intl. Symp. UAE, Oct. 292 Nov., 2008. Paper SCA2008-31.
Doyen PE (1987) Crack geometry of igneous rocks: a maximum
entropy inversion of elastic and transport properties. J Geophys
Res 92(B8):81698181
Duda A, Koza Z, Matyka M (2011) Hydraulic tortuosity in arbitrary
porous media flow. Phys Rev E 84:036319
Dvorkin J, Derzhi N, Fang Q, Nur A, Nur B, Grader A, Baldwin C,
Tono H, Diaz E (2009) From micro to reservoir scale: permeability from digital experiments. Lead Edge 28:1446
Dvorkin J, Derzhi N, Diaz E, Fang Q (2011) Relevance of computational rock physics. Geophysics 76(5):E141E153
Fortin J, Schnubnel A, Guguen Y (2005) Elastic wave velocities and
permeability evolution during compaction of Bleuswiller sandstone. Int J Rock Mech Min Sci 42:873889
Glover PW (2009) What is the cementation exponent? A new interpretation. Lead Edge 28:8285
Grosse-Kunstleve RW, Brunner GO, Sloane NJA (1996) Algebraic
description of coordination sequences and exact topological densities for zeolites. Acta Crystallogr A52:879889
Guguen Y, Dienes J (1989) Transport properties of rocks from statistics and percolation. Math Geol 21:113
Halawani MA (2000) Stratigraphic column for the phanerozoic rocks
of Saudi Arabia, Techn. Rept. BRGM-TR-2000-3, 95 pp
Hassanzadeh H, Pooladi-Darvish M (2006) Effects of fracture boundary conditions on matrix-fracture transfer shape factor. Transp
Porous Media 64:5171
Jorgensen DA (1988) Using geophysical logs to estimate porosity,
water resistivity, and intrinsic permeability, USGS Water-Supply
Paper # 2321, 24 pp
Kachanov M (1994) Elastic solids with many cracks and related problems. Adv Appl Mech 30:259345
Kalam Z, Al Dayyani T, Grader A, Sisk C (2011) Digital rock physics
analysis in complex carbonates, World Oil, 232, May 2011
Kayser A, Ziauddin M (2006) A closer look at pore geometry. Oilfield
Review, Spring, 2006, 4-14
Keehm Y (2003) Computational rock physics transport properties in
porous media and applications. Ph. D. Dissertation, Stanford
Kirkpatrick S (1973) Percolation and conduction. Rev Mod Phys
45(4):574588
Knackstedt M, Arns C, Sheppard A et al (2007) Archies exponents in
complex lithologies derived from 3D digital core analysis. Annual
Logging Symp. of SPWLA, SPWLA Paper UU:1-1C
Knackstedt M, Madadi M, Arns Ch, Baechle G, Eberli G, Weger R
(2009a) Carbonate petrophysical parameters derived from 3d
images. Search & Discovery article #40393, posted March 20,
2009

1137
Knackstedt M, Latham S, Sheppard A, Vaslet T, Arns C (2009b)
Digital rock physics: 3D imaging of core material and correlations
to acoustic and flow properties. Lead Edge Jan. 2009, 2833
Korvin G (1984) Shale compaction and statistical physics. Geophys J
R Astron Soc 78:3550
Korvin G (1992a) A percolation model for the permeability of
kaolinite-bearing sandstones. Geophys Trans 37(23):177209
Korvin G (1992b) Fractal models in the earth sciences. Elsevier,
Amsterdam
Kozeny J (1927) ber kapillare Leitung der Wasser in Boden,
Sitzungs-Ber. Akad Wiss Wien 136:271306
Kuster GT, Toksz MN (1974) Velocity and attenuation of seismic
waves in two-phase media. Pt. 1, Theoretical formulations.
Geophysics 39(5):587618
Lamm PK (1997) Solution of ill-posed Volterra equations via variable
smoothing Tikhonov regularization. In: Engl EW et al (eds)
Inverse problems in geophysical applications. SIAM, 92-108
Lifshitz EM, Pitaevskii LP (1980) Statistical physics, Part 1.
Pergamon, New York
Lim KT, Aziz K (1995) Matrix-fracture transfer shape factors for dualporosity simulators. Petrol Sci Eng 13:169178
Lucia FJ (1998) Carbonate reservoir characterization. Springer, Berlin
Mavko G, Mukerji T, Dvorkin J (1998) The rock physics handbook.
Tools for seismic analysis in porous media. Cambridge University
Press, Cambridge, UK
Meyer FO, Price RC, Al-Raimi SM (2000) Stratigraphic and petrophysical characteristics of core Arab-D super-k intervals,
Hawiyah area, Ghawar field, Saudi Arabia. GeoArabia
5(3):355384
Pal M (2012) A unified approach to simulation and upscaling of singlephase flow through vuggy carbonates. Int J Numer Methods
Fluids 69:10961123
Pelton WH, Ward SH, Halloff PG, Still WR, Nelson PH (1978)
Mineral discrimination and removal of inductive coupling with
multifrequency induced polarization. Geophysics 43:588609
Peng S, Hu Q, Dultz S, Zhang M (2012) Using X-ray computed
tomography in pore structure characterization for a Berea sandstone: resolution effect. J Hydrol 472473(23):254261
Perez-Rozales C (1982) On the relationship between formation resistivity factor and porosity. SPE J (Aug., 1982): 531-536
Pulido H, Samaniego FV, Cinco-Ley H, Rivera J, Guadalupe G (2007)
Triple porosity modedouble permeability with transient hydraulic diffusivity in naturally fractured reservoirs, Proceedings of
32nd Workshop on Geothermal Reservoir Engineering, Stanford
University, Stanford, California, January 2224, 2007, SGP-TR183
Pulido H, Galicia-Munoz G, Valds-Prez AR, Diaz-Garcia F Improve
Reserves estimation using interporosity skin in naturally fractured
reservoirs. Proceedings of 32nd Workshop on Geothermal
Reservoir Engineering, Stanford University, Stanford,
California, January 31February 2, 2011, SGP-TR-191
Rassenfoss S (2011) Digital rocks out to become a core technology. J
Petrol Technol May, 2011, 3641
Reuss A (1929) Berechnung der Fliessgrenzen von Mischkristallen auf
Grund der Plastizittsbedingung fr Einkristalle. Z Angew Math
Mech 9:4958
Sayers CM, Kachanov M (1995) Microcrack induced elastic wave
anisotropy of brittle rocks. J Geophys Res 100:41494156
Schlumberger (1991) Log interpretation principles/applications.
Schlumberger Educational Services, Houston, Texas
Schubnel A, Guguen Y (2003) Dispersion and anisotropy in cracked
rocks. J Geophys Res 108:2001. doi:10.1029/2002JB001824
Sok RM, Knackstedt MA, Sheppard AP, Pinczewski WV, Lindquist
WB, Venkatarangan A, Paterson L (2002) Direct and stochastic
generation of network models from tomographic images; effect of
topology on residual saturations. Transp Porous Media 46:345372

1138
Sok R, Varslot T, Ghous A, Latnam S, Sheppard AP, Knackstedt MA
(2009) Pore scale characterization of carbonates at multiple
scales: integration of MCT, BSEM and FIBSEM, International
Symp. of Soc. Core Anal
Sorbie KS, Skauge A (2011) Can network modeling predict two-phase
flow functions? International Symp. Soc. Core Anal., Austin, TX,
USA, 1821 Sept., 2011. Paper SCA2011-29
Stenger B, Pham T, Al-Afaleg N, Lawrence P (2003) Tilted original
oil/water contact in the Arab-D reservoir, Ghawar field, Saudi
Arabia. GeoArabia 8(1):939
Telford WM, Geldart LP, Sheriff RE (1990) Applied geophysics.
Cambridge University Press, Cambridge
Thomeer JHM (1960) Introduction of a pore geometrical factor defined by
a capillary pressure curve. Petr Trans AIME 219(TN 2057):354358
Thomeer JHM (1983) Air permeability as a function of three porenetwork parameters. J Petr Technol April, pp. 809-814
Touati M, Suicmez S, Funk J, Cinar Y, Knacksted M (2009) Pore
network modeling of Saudi Aramco rocks: a comparative study.
SPE Saudi Arabia Section Techn. Symp., 911 May, 2009, Al
Khobar, SA
Turcotte DL, Schubert G (1982) Geodynamics, applications of continuum physics to geological problems. Wiley, New York
Walsh JB, Brace WF (1984) The effect of pressure on porosity and the
transport property of rock. J Geophys Res 89B(11):94259431

Arab J Geosci (2014) 7:11271138


Walton K (1987) The effective moduli of a random packing of spheres.
J Mech Phys Solids 35:213226
Warren JE, Root PJ (1963) The behavior of naturally fractured reservoirs. SPE J 426:245255
Widjajakusuma J, Biswal B, Hilfer R (1999) Quantitative prediction of
effective material properties of heterogeneous media. Comput
Mater Sci 16(70)
Worthington PF (1993) The uses and abuses of the Archie equations, 1:
the formation factor-porosity relationship. J Appl Geophys
30(3):215228
Worthington PF (2011) The petrophysics of problematic reservoirs. J
Petrol Technol December 2011, 8897
Wyllie MRJ, Gregory AR, Gardner GHF (1956) Elastic wave velocities in heterogeneous and porous media. Geophysics 21:4170
Wyllie MRJ, Gregory AR, Gardner GHF (1958) An experimental
investigation of factors affecting elastic wave velocities in porous
media. Geophysics 23:459493
Yonezawa F, Cohen MH (1983) Granular effective medium approximation. J Appl Phys 54:28952899
Zhang X, Knackstedt MA (1995) Direct simulation of electrical and
hydraulic tortuosity in porous solids. Geophys Res Lett
22(17):23332336
Zimmerman RW (1991) Compressibility of sandstones. Elsevier,
Amsterdam

Das könnte Ihnen auch gefallen