Sei sulla pagina 1di 8

Journal of Applied Geophysics 75 (2011) 671678

Contents lists available at SciVerse ScienceDirect

Journal of Applied Geophysics


journal homepage: www.elsevier.com/locate/jappgeo

Elastic moduli of dry rocks containing spheroidal pores based on differential effective
medium theory
Hongbing Li , Jiajia Zhang
Petrochina Company Limited, Research Institute of Petroleum Exploration and Development, Beijing, China

a r t i c l e

i n f o

Article history:
Received 5 July 2011
Accepted 15 September 2011
Available online 24 September 2011
Keywords:
Spheroidal pores
Dry rocks
Differential effective medium theory
Elastic moduli
Sandstone
Mathematical formulation

a b s t r a c t
Differential effective medium (DEM) theory is applied to determine the elastic properties of dry rock with
spheroidal pores. These pores are assumed to be randomly oriented. The ordinary differential equations for
bulk and shear moduli are coupled and it is more difcult to obtain accurate analytical formulae about the
moduli of dry porous rock. In this paper, we derive analytical solutions of the bulk and shear moduli for
dry rock from the differential equations by applying an analytical approximation for dry-rock modulus ratio, in
order to decouple these equations. Then, the validity of this analytical approximation is tested by integrating
the full DEM equation numerically. The analytical formulae give good estimates of the numerical results over
the whole porosity range. These analytical formulae can be further simplied under the assumption of
small porosity. The simplied formulae for spherical pores (i.e., the pore aspect ratio is equal to 1) are the
same as Mackenzie's equations. The analytical formulae are relatively easy to analyze the relationship between
the elastic moduli and porosity or pore shapes, and can be used to invert some rock parameters such as porosity
or pore aspect ratio. The predictions of the analytical formula for the sandstone experimental data show that the
analytical formulae can accurately predict the variations of elastic moduli with porosity for dry sandstones. The
effective elastic moduli of these sandstones can be reasonably well characterized by spheroidal pores, whose
pore aspect ratio was determined by minimizing the error between theoretical predictions and experimental
measurements. For sandstones the pore aspect ratio increases as porosity increases if the porosity is less than
0.15, whereas the pore aspect ratio remains relatively stable (about 0.14) if the porosity is more than 0.15.
2011 Elsevier B.V. All rights reserved.

1. Introduction
The bulk and shear moduli of dry-rock framework are two essential parameters for uid substitution and shear-wave prediction in
rock physics study. Between the elastic moduli of dry porous rock
and porosity there does not exist a unifying formula, which depends
on the provided effective medium theory or empirical relations.
The empirical relations, such as the critical porosity model and
Krief's model (Krief et al., 1990; Nur et al., 1998), assume simply that
the elastic moduli relate to porosity, rather than take into account the
effect of pore geometry. Both the critical porosity model and Krief's
model implicitly predict a constant modulus ratio or Vp/Vs of dry-rock
frame, irrespective of porosity. However, it has been proven that
this assumption is not fully reasonable (Gregory, 1976; Li and
Zhang, 2010; Pickett, 1963). The modulus ratio of dry porous rock
depends not only on porosity but also on the pore geometry (Li and
Zhang, 2010).

Corresponding author.
E-mail addresses: hbingli@petrochina.com.cn (H. Li), zhangjia19861121@yahoo.cn
(J. Zhang).
0926-9851/$ see front matter 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.jappgeo.2011.09.012

The effective medium theory might be used to estimate the elastic


properties from the mineral composition and microstructure of porous rock. The KusterToksz theory (Kuster and Toksz, 1974) and
differential effective medium (DEM) theory (Norris, 1985) are two
popular methods of effective medium theory. The DEM theory has
been developed to the determination of rock electric properties and
effective elastic properties of porous rocks that are dry or saturated
by uid (Berryman et al., 2002; Markov et al., 2005; Norris, 1985;
Sheng, 1991; Zimmerman, 1985). The DEM has the property that it
can never violate rigorous bounds comparing with the KusterToksz
theory (Berryman, 1980; Berryman and Berge, 1996; Berryman et al.,
2002; Kuster and Toksz, 1974). However, since the ordinary differential equations for bulk and shear moduli are coupled, it is difcult to
integrate them to yield accurate analytical formulae for the bulk and
shear moduli. Therefore, the behavior of the bulk and shear moduli
can only be accurately simulated through the numerical solution of
differential equations.
Some of the notable work on dry porous solids has been made in
order to obtain some analytical formulae that are relatively easy to analyze. Zimmerman (1985, 1991) derived implicit analytical solutions for
the differential scheme in the limiting cases of innitely thin cracks and
spherical pores, respectively. A solution for the case of needle-shaped

672

H. Li, J. Zhang / Journal of Applied Geophysics 75 (2011) 671678

pores has recently been given by David and Zimmerman (2011).


Berryman et al. (2002) obtained approximate analytical results for the
elastic moduli of dry and fully saturated cracked rock based on DEM
theory by assuming that changes in Poisson's ratio occurring in those
terms proportional to the aspect ratio are negligible to rst-order.
Their results show that Poisson's ratio of dry cracked rock depends on
porosity and the crack aspect ratio. Li and Zhang, in press derived explicit analytical solutions of the bulk and shear moduli for dry rock
with the three specic pore shapes: spherical pores, needle-shaped
pores and penny-shaped cracks in order to decouple the differential
equations by applying an analytical approximation for dry-rock modulus ratio. Their analytical formulae give good estimates of the numerical
results over the whole porosity range for the cases of the three given
pore shapes (Li and Zhang, in press).
Penny-shaped cracks have been used extensively to model cracked
materials. Compared with penny-shaped cracks, spheroidal pores are
more suitably used to model porous materials. However, the expressions of the polarization factors for spheroidal pores are more complicated than that of penny-shaped cracks (Berryman, 1980; Mavko
et al., 1998). It is more difcult to obtain analytical expressions of the
bulk and shear moduli to porosity and pore geometry. Some efforts
have been made to achieve certain results. Keys and Xu (2002) obtained
an approximation for dry-rock bulk and shear moduli by applying the
DEM method to the Kuster and Toksz equations (Kuster and Toksz,
1974) for spheroidal pores and assuming a constant dry-rock Poisson's
ratio in coupled equations but the Poisson's ratio or modulus ratio
derived from their resulting dry-rock approximations is not constant
as they at rst assumed. The approximate solutions for crack-like
pores that are not innitely thin, and needle-like pores that are
not innitely long were obtained by David and Zimmerman
(2011). Their results are highly accurate for 0 b b 0.3, and 3 b b .
The purpose of this paper is to obtain analytical formulae for the
bulk and shear moduli of dry rocks with spheroidal pores based on
DEM theory and use them to characterize the dry-rock moduli dependency on porosity. The paper is organized as follows. First we introduce
the DEM theory and Li and Zhang's (Li and Zhang, 2010) approximate
formula for the modulus ratio of dry porous rock. Next we derive analytical results for the bulk and shear moduli of dry rock from the differential equation by applying Li and Zhang's analytical approximation for
dry-rock modulus ratio to solve the coupled differential equations for
bulk and shear moduli. And then we obtain the rst-order approximations from our analytical formulae under the assumption of small porosity. And then we use the theoretical model that Li and Zhang
(2010) designed for checking these approximations by comparison
with numerical computations for the derivative equation of the
DEM. The agreement between the analytical approximations and
the full DEM for dry rock with different pore aspect ratios is discussed.
Finally the analytical approximations are applied to predict the variations of the bulk and shear moduli with respect to porosity for dry porous glass foam, which were measured by Walsh et al. (1965), and
sandstone data, which were measured by Han (1986), respectively, in
the laboratory.
2. Differential effective medium theory
Differential effective medium theory (Berryman, 1992; Berryman
et al., 2002; Mavko et al., 1998; Norris, 1985) takes the point of view
that a composite material may be constructed by making innitesimal changes in an already existing composite. When the inclusions
are sufciently sparse so that they do not form a single connected
network throughout the composite, it is appropriate to use the DEM
to model its effective elastic behavior. Assume that the host material
has moduli Km and Gm and the inclusion material has moduli Ki and Gi.
Then, the effective bulk and shear moduli of the composite are parameterized by K*(y) and G*(y) where the volume fraction of the inclusion phase is y, which is also called porosity. The coupled systems

of ordinary differential equations for these moduli are (Berryman,


1992)
1y

 i
dK  y 

Ki K y P
dy


 i
dG y 

Gi G y Q
dy

and
1y

where the scalar factors P *i and Q *i are the so-called polarization factors for bulk and shear modulus (Eshelby, 1957; Wu, 1966), and the
subscript i denotes the inclusion phase. These equations are typically
integrated starting from porosity y = 0 with values K*(0) = Km and
G*(0) = Gm, which are assumed here to be the mineral moduli values
for the single homogeneous solid constituent. Integration then proceeds from y = 0 to the desired highest value y = ( is the resulting
porosity in the composite medium).
The factors P *i and Q *i depend, in general, on the bulk and shear
moduli of the inclusion, the effective medium marked by stars and
on the shapes of the inclusions. The polarization factors usually
have been computed from Eshelby's well-known results and Wu's results, which can be found in many places including Berryman (1980)
and Mavko et al. (1998).
3. Elastic moduli approximations for dry rock
For dry rock, Ki = 0 and Gi = 0. Eqs. (1) and (2) become

1y

dK  y
i
P
K  ydy

and

1y

dG y
i
Q :
G ydy

In general the DEM Eqs. (3) and (4) are also coupled, as both equations depend on both the bulk and shear moduli of the composite
(Berryman, 1980; Mavko et al., 1998). Therefore, it is difcult to obtain the analytical formula of the bulk and shear moduli for dry rock
from integrating the DEM equations. Li and Zhang (2010) obtained
the analytical formula of modulus ratio for dry rock by combining
Eqs. (3) and (4) and assuming an almost linear relationship between
the difference P *i Q *i and the effective modulus ratio K*(y)/G*(y).
Their formula is given by
Kdry Km

Gdry Gm

1
bKm bKm
a
1

1
aGm aGm

where Kdry and Gdry are the bulk and shear moduli for dry rock respectively, constants a and b are intercept and gradient respectively,
which meet P*i Q*i = a + bK*(y)/G*(y). Constant b is simply set to
the rst derivative of P *i Q*i and then constant a can be calculated
from P*i Q*i = a + bK*(y)/G*(y) where K*/G* = Km/Gm (Li and Zhang,
in press). The expressions of P *i Q*i and its rst derivative for porous
rocks with spheroidal pores are given in Appendix A.
In order to decouple the equations for bulk and shear moduli,
substituting Eq. (5) into (3), and then, integrating it from y = 0 to
gives directly the analytical results for Kdry with spheroidal pores, and

H. Li, J. Zhang / Journal of Applied Geophysics 75 (2011) 671678

then, dividing the Kdry by Eq. (5) gives the corresponding analytical results for Gdry, we have (see Appendix B)

S0 S1 S2 S3

Gdry



S1 S3 S2 S3
S3 Km =Gm
1
Km

S0
bK
bK
a b
1 m m 1
aGm aGm
S0 S1 S2 S3

Kdry


S1 S3 S2 S3
a
S3 Km =Gm
1
Gm

S0 1
bK
bK
a b
1 m m 1
aGm aGm

673

23g3
g
4
2g
, S1
,
, S , S 
23g3 2 3 3 3 232 2g
2 232 2g

where S0 

Fig. 1. The dry-rock bulk and shear moduli (a) for spherical pores ( = 1). The dry-rock bulk and shear moduli (b) for needle-shaped pores ( ). The dry-rock bulk and shear
moduli (c) for spheroidal pores with = 0.1. The dry-rock bulk and shear moduli (d) for spheroidal pores with = 0.01. The dry-rock bulk and shear moduli (e) for spheroidal
pores with = 0.001. A full DEM calculation is shown as black points. The analytical formula in the text is displayed as a solid line. Agreement between the full DEM calculation
and the analytical formula is excellent.

674

H. Li, J. Zhang / Journal of Applied Geophysics 75 (2011) 671678

h
pi
8

1
>
cos 12
b 1

>
2
< 
2 3=2

1
h p
i

,
g

32,

1
>
12
>
3=2 2 1cosh N 1
: 2
1

they are functions of the pore aspect ratio .

similar to the dry-rock approximations of Keys and Xu (2002). The


approximations they developed are
p

15

16

Kdry Km 1
and

3.1. First-order approximation of analytical formulae for dry porous


rocks

Gdry Gm 1

For sufciently small , by using rst-order Taylor series expansion we simplify Eq. (5) as



Kdry Km
K

1 a b m
Gdry Gm
Gm

K
where a b m is actually the value of P *i Q *i at K*(y)/G*(y) =
Gm
Km/Gm.
Therefore, Eqs. (6) and (7) are simplied as (see Appendix B)

S0 S1 S2 S3

Kdry Km


S1 S3 S2 S3
S3 Km =Gm

Km S0

Gm


S1 S3 S2 S3
a
S3 Km =Gm
1



:
Gm
K S
K
1 m 0 ab m
Gm
Gm

S0 S1 S2 S3

Gdry

10

When = 1, i.e. the pore space consists of dry, randomly distributed spherical pores, we have = 2/3, g =2/5, S0 = 3/4, S1 = 8/9,
Km 3 Km 6Km =Gm 12
, then Eqs. (9) and

S2 = 4/3, S3 = 8/9, a b
Gm 4 Gm 9Km =Gm 18
(10) become
Kdry Km

Gdry Gm

1
3K
1 m
4Gm
1
:
6Km =Gm 12

1
9Km =Gm 8

11

12

where p and q are exponents dened by Eqs. (15) and (16) in Keys
and Xu (2002). Therefore, the dry-rock approximations of Keys and
Xu (2002) are special cases of Eqs. (6) and (7).
4. Theoretical and experimental examples
We now consider some applications of the analytical formulae and
their rst-order approximations for the bulk and shear moduli of dry
rock.
First we give a comparison between the analytical formulae or
their approximations and numerical results of the differential equations. We take quartz as the host medium, having Km = 37 GPa and
Gm = 44 GPa, and the solid's Poisson's ratio is equal to 0.0742. We
use the fourth-order RungeKutta scheme to get numerical solutions
of the differential Eqs. (3) and (4). The step size used in RungeKutta
scheme relates to the pore geometry. The step size for = 1 (i.e.
spherical pore) is set to y = 0.1, the step size for (i.e.
needle-shaped pore) is also set to y = 0.1, and the step sizes are
set to y = 0.1 for = 0.1, y = 0.01 for = 0.01 and y = 0.001 for
= 0.001 respectively. We show comparisons of the analytical and
numerical results: (1) spherical pores in Fig. 1a, (2) needle-shaped
pores in Fig. 1b, (3) spheroidal pores with = 0.1 in Fig. 1c, (4) spheroidal pores with = 0.01 in Fig. 1d, and (5) spheroidal pores with
= 0.001 in Fig. 1e. In Fig. 1, solid points represent the numerical results. The solid line represents the analytical results calculated by
Eqs. (6) and (7) whose constant b is simply set to the rst derivative
of P *i Q *i and constant a can then be calculated from P *i Q *i = a +
bK*(y)/G*(y) where K*/G* = Km/Gm = 0. 8409 (see Appendix A). The
results show that the analytical methods are in sufciently good
agreement with the numerical results in all cases.
Fig. 2 shows the relation between porosity and bulk modulus determined from compressibility measurements for the glass foam samples of Walsh et al. (1965). The porosity of the glass samples ranges
from 70% to near zero and the pores are nearly spherical. We apply
Eqs. (6) and (11) to predict the bulk modulus of dry rock. The bulk

Eqs. (11) and (12) are the same as Mackenzie's equations which
were derived for a rock containing a low concentration of spherical
pores (Mackenzie, 1950; Walsh, 1965; Walsh et al., 1965). This
shows that our analytical formulae, i.e., Eqs. (6) and (7), are valid
and more general than Mackenzie's for spherical pores.
When , i.e. the pore space consists of dry, randomly distributed needle-shaped pores, we have = 1, g =1, S0 = 1, S1 = 1,
Km
13 Km
12

, then
S2 = 4/3, S3 = 4/3,
ab

15 Gm 15Km =Gm 5
Gm
Eqs. (9) and (10) become
Kdry Km

Gdry Gm

1
K
1 m
Gm
1

:
3
12

1
15 15Km =Gm 5

13

14

Furthermore, if we let the denominators of Eqs. (9) and (10) be


equal to 1 (e.g. when b 0), we obtain the approximations which

Fig. 2. Bulk moduli of porous glass foam computed from compressibilities measured by
Walsh et al. (1965). Measured values (black rectangles) are compared to theoretical
predictions of DEM theory. The solid curve is obtained from the accurate formula for
spherical pores while the dashed line is computed from the approximate formula.
The dotted curve is the theoretical predictions of SC theory for spherical pores.

H. Li, J. Zhang / Journal of Applied Geophysics 75 (2011) 671678

modulus and mineral modulus ratio of the glass samples used in the
analysis were taken to 46.3 GPa and Km/Gm = 1.5185 from Walsh et
al. (1965), respectively. The solid's Poisson's ratio is equal to 0.23.
The pore aspect ratio is = 1. Constants a and b used in Eq. (6) are
a =1.1684, b = 0.8778, respectively. As shown in Fig. 2, the solid
line represents the prediction of Eq. (6) and the dashed line represents the prediction of Eq. (11). We also use the self-consistent (SC)
theory (Berryman, 1980; Mavko et al., 1998) for estimating elastic

675

moduli of a composite. In SC modeling the glass were represented


as spherical grains and the pore space also as spheres. Theoretical
prediction of the SC theory is displayed as the dotted line. We see
that all these estimates do well at porosities below about 20%, but
for intermediate porosities below about 50%, the spherical approximation overestimates the modulus while the accurate formula for
spherical pore provides do somewhat better at estimating the measured values. There is only one sample at porosity up to 70% close

Fig. 3. Result analysis of the experimental sandstone data measured by Han (1986) and theoretical estimations of the formulae for spheroidal pores. (a) Bulk modulus versus porosity for experimental data. (b) Shear modulus versus porosity for experimental data. (c) Bulk and shear moduli versus porosity for 10 samples of clean sandstone and the theoretical predictions. (d) The estimated pore aspect ratio versus porosity. (e) Comparison of bulk modulus between the experimental result and the theoretical calculation from the
estimated pore aspect ratio. (f) Comparison of shear modulus between the experimental result and the theoretical calculation from the estimated pore aspect ratio.

676

H. Li, J. Zhang / Journal of Applied Geophysics 75 (2011) 671678

to the spherical approximation but away from the accurate formula.


Similar result for the same data set was also obtained from the differential scheme presented by Zimmerman (1991).
Figs. 1 and 2 show that Eqs. (6) and (7) are accurate at Poisson's ratios 0.0742 and 0.23, not like many approximate effective medium
models work well for Poisson's ratios near 0.2 or 0.25, but become inaccurate for large or small values as pointed out by David and Zimmerman
(2011).
We use 69 dry sandstone data of Han (1986) to analyze the predictions of the formulae for spheroidal pores. The data were obtained by
measuring wave velocities in dry-rock samples at conning pressure
40 MPa. The porosity of these samples ranges from 5 to 30%, and the
clay content range in these samples is between 0 and 50%. Fig. 3ab
shows the relation between porosity and moduli of dry frame using
Han's dry sandstone data (Han, 1986). The measured results indicate
that both the bulk and shear moduli of dry sandstone decrease as porosity or clay content increases. The bulk and shear moduli of quartz used
in the analysis were taken to 40 GPa and 43.7 GPa, respectively. The
bulk and shear moduli of clay were taken to 20.9 GPa and 6.85 GPa, respectively. These were provided by Han (1986). The bulk and shear
moduli of composite grains, including clay, are calculated using Voigt
ReussHill's averaging method. The Poisson's ratios of composite solids
range from 0.12 to 0.24. First we apply Eqs. (6) and (7) to predict the
moduli of 10 samples of clean sandstone with porosity. The pore aspect
ratio is set to = 0.145. The solid's Poisson's ratio for clean sandstone is
equal to 0.1. As shown in Fig. 3c, the dry bulk modulus prediction is displayed as a solid curve, the dry shear modulus prediction is displayed as
a dashed curve. The theoretical predictions can reect the trend that the
dry-rock moduli vary with porosity. But not all the theoretical predictions of samples are very good agreement with the measured data. Because the pore structure of all the samples is not all the same, it is
inappropriate to predict the elastic modulus with the porosity by applying a single pore aspect ratio for all samples. Each sample should have
its own pore aspect ratio.
The estimation of the pore structure of a rock from its measured
velocities was presented early by Cheng and Toksz (1979), based
on the KusterToksz model (Kuster and Toksz, 1974), their computational process is more complicated. Here for Han's shaly sandstone
we assume that the rock whose matrix is made up of sand and clay
contains pores with a single aspect ratio. Therefore, we can easily estimate the pore aspect ratio of each sandstone sample by minimizing
the expression between the experimental moduli and theoretical predictions from Eqs. (6) and (7). Fig. 3d shows the estimated results of
the pore aspect ratio. Note that the pore aspect ratio increases as porosity increases if the porosity is less than 0.15, whereas the pore aspect ratio remains relatively stable (about 0.14) if the porosity is
more than 0.15. The higher the clay content, the lower the pore aspect
ratio, but if the porosity is less than 0.15, there is no clear correlation
between the pore aspect ratio and clay content. The assumption on
porous rock with single aspect ratio is different from Xu and White's
model (Xu and White, 1995) where two aspect ratios are used, one
for the sand fraction and another for the clay fraction, but these two
aspect ratios have to be xed.
After the pore aspect ratio is given, we can apply Eqs. (6) and (7)
to predict the elastic moduli of dry rock. Fig. 3e and f is a comparison
between the predicted and measured data. Observe that for the majority of samples the accurate formulae provide good estimates.
There is only for three samples where the predictions of the shear
modulus are deviated from the measured data. This fully shows that
the effective elastic moduli of these sandstones can be reasonably
well characterized by spheroidal pores.
5. Conclusions
By applying an analytical approximation for dry-rock modulus
ratio, analytical solutions for the bulk and shear moduli of dry rock

with spheroidal pores from the differential equation have been derived. The analytical formulae are functions of the bulk and shear
moduli of the mineral grain, porosity and pore geometry. The pore geometry is characterized by a linear relation whose intercept a and
gradient b might be approximately calculated from the values of the
difference between the polarization factors P and Q and its rst derivative where the effective modulus ratio is equal to the mineral grain
modulus ratio.
The analytical formulae can be further simplied to obtain rstorder approximate formulae under the assumption of small porosity.
When the aspect ratio is equal to 1, i.e., the material contains the inclusion with ideal spherical pores, the simple rst-order formulae are
the same as Mackenzie's equations for low-porosity media. This testies that our approximate formulae are valid and more general
than Mackenzie's for spherical pores.
The numerical examples show that the analytical formulae are in
sufciently good agreement with the numerical results. The experimental examples show that the analytical formulae can accurately
predict the variations of elastic moduli with porosity for dry sandstones. The analytical formula of bulk and shear moduli for dry rock
can be used to t the experimental data using an apparent aspect
ratio, which was determined by minimizing the error between theoretical predictions and experimental measurements. The effective
elastic moduli of these sandstones can be reasonably well characterized by spheroidal pores. For sandstone there is a positive correlation
between the pore aspect ratio and porosity.
Acknowledgments
This work was supported by the Important National Science and
Technology Specic Projects2011ZX05001 of China. We thank the editors and reviewers for many helpful comments and suggestions that
improve this paper. We also thank the Research Institute of Petroleum Exploration and Development, Petrochina Company Limited
for allowing us to publish this work.
Appendix A. P *i Q *i and its rst derivative for dry rock with
spheroidal pores
For dry spheroidal void the polarization factors are given by
(Berryman, 1980; Mavko et al., 1998)
P

1
T ;
3 iijj

A1

1
1
Tijij Tiijj ;
5
3

A2

where the tensor Tijkl relates the uniform strain eld at innity to
the strain eld within an spheroidal inclusion (Wu, 1966). Berryman (1980) gave the formulations for calculating P and Q as
Tiijj

3F1
;
F2

A3

Tijij

F1 2
1 F F F6 F7 F8 F9
4 5
;
F2 F3 F4
F2 F4

A4

where
F1 1 A


3
3
5
4
g R g
;
2
2
2
3



3
R
F2 1 A 1 g 3g 5 B34R
2
2
h

i
A
2
A 3B34R g R g 2
2

A5

A6

H. Li, J. Zhang / Journal of Applied Geophysics 75 (2011) 671678



2
3
1 2
A4
R2
F3 1
g R15
2
2

A7

F 9 Ag4BR ;

R R ;
A
F4 1 3 gRg;
4

A8


4
F5 A R g g B34R;
3

A9

F6 1 A1 gRg  B13  4R;

A10

A
F7 2 9 3gR5 3g  B34R;
4

A11



g

F8 A 12R R1 5R3 B134R;


2
2

A12

F9 Ag R1R B34R;

A13

G
A i 1;
G

A14

3G
R
;
3K 4G

A16

A17

A18

3
5

g 4=3 R ;
2
2

A19

A20
2

A29

F 1 F2 F1 F 2
F
F
2 32 42
F22
F3 F4



F 4 F5 F4 F 5 F 6 F7 F6 F 7 F 8 F9 F8 F 9 F2 F4 F4 F5 F6 F7 F8 F9 F 2 F4 F2 F 4

:
F22 F42

Therefore, from Eqs. (A1), (A2), (A29) and (A30) we give PQ


and its rst derivative with respect to K/G as
6
1
T T ;
15 iijj 5 ijij
6
1
T T :
15 iijj 5 ijij

A31
A32

For dry rocks, Ki = 0 and Gi = 0, so A = 1, B = 0. Eq. (A1) becomes


g
3
3
K K 4
1 g

2 2
2
2
G G 3
:
P

K
2
2
2
g 23 2g
3
3
G

A33

Eq. (A33) indicates that P is the function of K/G. Similarly, we


can also get the expression of Q, it is also the function of K/G. Therefore, P Q is also the function of K/G. In general, P *i Q *i varies slowly
with K*(y)/G*(y), it can be approximated by a linear relationship (Li
and Zhang, 2010), i.e., P *i Q *i = a + bK*(y)/G*(y), constants a and b
can be estimated by a linear regression method. Here, we make the
further approximation. We simply set constant b to the rst derivative of P *i Q *i at K*(y)/G*(y) = Km/Gm, i.e., b = (P *i Q *i) and constant a can be then taken as a = (P *i Q *i) (P *i Q *i)Km/Gm.
Appendix B. Derivation of the bulk and shear moduli for dry
porous rock with spheroidal pores and their
rst-order approximations
Substituting Eq. (A33) into Eq. (3), we have

  

g
3
3
K
K
4

1
g


dK y
2 2
2
2
G G 3
1y 

:

K
K ydy
2
2
2
g 23 2g 
3
3
G


A
1

g R;
2
2

A21

F 4 gR ;
4

F5

T ijij

F22


h

i
A
A

2
2

F 2 3g 54B A 3B 7g6 8R g 2
R;
2
2

F3

3 F 1 F2 F1 F 2

In Eqs. (A14)(A16), K and Ki are the bulk moduli for the rock matrix and inclusion material, respectively; G and Gi denote the shear
moduli for the rock matrix and inclusion material, respectively; is
the aspect ratio of the inclusion.
Taking the K/G derivative of Eqs. (A5)(A13) and (A16), we have
F 1 A

T iijj

PQ

2
32:
12

A28

where the superscript' denotes rst derivative. From Eqs. (A19)


(A28) we obtain the rst derivatives of Tiijj and Tijij with respect to
K/G. i.e.

PQ
A15

A27

A30

1 Ki Gi
B

;
3 K
G

h
pi
8

1
>
cos 12
b 1

>
< 
2 3=2
1
h
i
p

1
>
>
3=2 2 1cosh N 1
: 2
1

677

A22

B1

Rearranging the right side of Eq. (B1) gives the result


4

A g 4B R ;
3

A23

1y

 

dK  y
K y
S S S S
S0 
S1 S2 S3 1 3 2  3

K ydy
G y
S3 K y=G y

B2

F 6 Ag 4B1:R ;

A24

where



A

F 7 5 3g 4B R ;
4

A25

S0

23g3
;
462 4g

B3


g 5

F 8 A 2 4B1 R ;
2 2

A26

S1

g
;
23g3

B4

678

H. Li, J. Zhang / Journal of Applied Geophysics 75 (2011) 671678

S2

4
;
3

B5

Dividing Eq. (B11) by Eq. (15) gives


h
S0 S1 S2 S3

2g
:
S3 
3 232 2g

B6

Expanding the last term around y = 0 on the right hand side in


Eq. (B2) using rst-order Taylor series, we have

1y

dK  y
S0
K  ydy

K  y
S1 S2 S3 S1 S3 S2 S3
G y

3

Km i
i 
P Q K  K
6
7

Gm
6
7
m
6
7
1
G Gm y
6
7

6
7
2
S
1y

K
=G

K
=G

3
m
m
6
7
3
m
m
4
5
2

)
:

B7
Substituting Eq. (5) into (B7), then integrating it from y = 0 to ,
and since 0 y and 0 1, omitting the quadratic term with respect to (i.e., omitting the second term within the square bracket in
Eq. (B7)), we have
K

S0
 dry
b
K
K

a
lnK 
ln a b m b m 1y

Gm
Gm
Km

h
S0 S1 S2 S3

1y

S1 S3 S2 S3
S3 Km =Gm

B8

i 


 :

0

Eq. (B8) can be easily rearranged to give


h
S0 S1 S2 S3

S1 S3 S2 S3

S3 Km =Gm

1
Kdry Km 
S0 :
bK
bK
a b
1 m m 1
aGm aGm

B9

Dividing Eq. (B9) by Eq. (5) gives


h
S0 S1 S2 S3

S1 S3 S2 S3

S3 Km =Gm

1
Gdry Gm 
S0 1 :
bKm bKm
a b
1

1
aGm aGm

B10

For sufciently small , (1) a 1a, Eq. (B9) is simplied as


h
S0 S1 S2 S3

Kdry Km

S1 S3 S2 S3
S3 Km =Gm

Km S0

Gm

i
:

B11

Gdry Gm

S1 S3 S2 S3
S3 Km =Gm


i
1 KGm S0 a b GKm
m

i
a

B12

References
Berryman, J.G., 1980. Long-wavelength propagation in composite elastic media. The
Journal of the Acoustical Society of America 68, 18091831.
Berryman, J.G., 1992. Single-scattering approximations for coefcients in Biot's equations of poroelasticity. The Journal of the Acoustical Society of America 91,
551571.
Berryman, J.G., Berge, P.A., 1996. Critique of explicit schemes for estimating elastic
properties of multiphase composites. Mechanics of Materials 22, 149164.
Berryman, J.G., Pride, S.R., Wang, H.F., 2002. A differential scheme for elastic properties
of rocks with dry or saturated cracks. Geophysical Journal International 151,
597611.
Cheng, C.H., Toksz, M.N., 1979. Inversion of seismic velocities for the pore aspect-ratio
spectrum of a rock. Journal of Geophysical Research 84, 75337543.
David, E.C., Zimmerman, R.W., 2011. Elastic moduli of solids containing spheroidal
pores. International Journal of Engineering Science 49, 544560.
Eshelby, J.D., 1957. The Determination of the Elastic Field of an Ellipsoidal Inclusion,
and Related Problems. Proceedings of the Royal Society of London Series A 241,
376396.
Gregory, A.R., 1976. Fluid saturation effects on dynamic elastic properties of sedimentary rocks. Geophysics 41, 895921.
Han, D.-H., 1986. Effects of porosity and clay content on acoustic properties of sandstones and unconsolidated sediments, PhD thesis, Stanford University.
Keys, R.G., Xu, S., 2002. An approximation for the XuWhite model. Geophysics 67,
14061414.
Krief, M., Garat, J., Stellingwerff, J., Ventre, J., 1990. A petrophysical interpretation using
the velocities of P and S waves (fullwaveform sonic). The Log Analyst 31, 355369.
Kuster, G.T., Toksz, M.N., 1974. Velocity and attenuation of seismic waves in two
media, Part I, Theoretical considerations. Geophysics 39, 587606.
Li, H., Zhang, J., 2010. Modulus ratio of dry rock based on differential effective medium
theory. Geophysics 75 (2), N43N50.
Li, H., Zhang, J., in press. Analytical Approximations of Bulk and Shear Moduli for Dry
Rock Based on Differential Effective Medium Theory, Geophysical Prospecting.
doi:10.1111/j.1365-2478.2011.00980.x.
Mackenzie, J.K., 1950. The Elastic Constants of a Solid Containing Spherical Holes. Proceedings of the Physical Society of London 63B (l), 211211.
Markov, M., Levine, V., Mousatov, A., Kazatchenko, E., 2005. Elastic properties of
double-porosity rocks using the differential effective medium model. Geophysical
Prospecting 53, 733754.
Mavko, G., Mukerji, T., Dvorkin, J., 1998. The Rock Physics Handbook: Tools for Seismic
Analysis in Porous Media. Cambridge University Press, New York, pp. 307309.
Norris, A., 1985. A differential scheme for the effective moduli of composites. Mechanics of Materials 4, 116.
Nur, A., Mavko, G., Dvorkin, J., Galmudi, D., 1998. Critical porosity: a key to relating
physical properties to porosity in rocks. The Leading Edge 17, 357362.
Pickett, G.R., 1963. Acoustic character logs and their applications in formation evaluation. Journal of petroleum technology 15, 650667.
Sheng, P., 1991. Consistent modeling of the electrical and elastic properties of sedimentary rocks. Geophysics 56, 12361243.
Walsh, J.B., 1965. The effect of cracks on the compressibility of rocks. Journal of Geophysical Research 70, 381389.
Walsh, J.B., Brace, W.F., England, A.W., 1965. Effect of porosity on compressibility of
glass. Journal of the American Ceramic Society 48, 605608.
Wu, T.T., 1966. The effect of inclusion shape on the elastic moduli of a two-phase material. International Journal of Solids and Structures 2, 18.
Xu, S., White, R.E., 1995. A new velocity model for claysand mixtures. Geophysical
Prospecting 43, 91118.
Zimmerman, R.W., 1985. The effect of microcracks on the elastic moduli of brittle materials. Journal of Materials Science Letters 4, 14571460.
Zimmerman, R.W., 1991. Elastic moduli of a solid containing spherical inclusions. Mechanics of Materials 12, 1724.

Potrebbero piacerti anche