You are on page 1of 7

Constitutive modeling

Copyright 2005 Taylor & Francis Group plc, London, UK


Up-scaling analysis with rigorous error estimates for poromechanics in
random polycrystals of porous laminates

James G. Berryman
University of California, Lawrence Livermore National Laboratory, USA

ABSTRACT: A detailed analytical model of random polycrystals of porous laminates has been developed.
This approach permits detailed calculations of poromechanics constants as well as transport coefficients. The
resulting earth reservoir model allows studies of both geomechanics and fluid permeability to proceed semi-
analytically. Rigorous bounds of the Hashin-Shtrikman type provide estimates of overall bulk and shear moduli,
and thereby also provide rigorous error estimates for geomechanical constants obtained from up-scaling based
on a selfconsistent effective medium method. The influence of hidden or unknown microstructure on the final
results can then be evaluated quantitatively. Descriptions of the use of the model and some examples of typical
results on the poromechanics of such a heterogeneous reservoir are presented.

1 INTRODUCTION functions of the parameters, have been expressed only


algorithmically not as analytical formulas. But, in
An explosion of new results on rigorous bounding some cases it has been possible to carry the analysis
methods (Milton 2002; Torquato 2002) has been seen further than was done in the published literature. When
over the last fifty years. Effective medium theory, this is true, a self-consistent effective medium formula
although very useful in many circumstances, neverthe- is then straightforward to obtain from the resulting
less has not seen a similar explosion of new results. So expressions. Then, as expected, resulting predictions
a question that naturally arises is whether we can con- are consistent with the bounds.
struct new effective medium formulas directly from In Section 2, results from a double-porosity geome-
these known bounds? Skeptics will immediately ask: chanics analysis are presented. These results are
Why do we need to do this at all if the bounds are avail- general and do not depend explicitly on the spa-
able? But the answer to this question is most apparent tial arrangement of the two porous constituents.
in poromechanics, where the bounds are often too far Microstructure enters these formulas only through the
apart to be of much practical use. overall drained bulk modulus K . Then, in Section 3,
Hill (1952) was actually the first to try construct- a preferred microstructure that of a locally lay-
ing estimates from bounds. He showed that the Voigt ered medium is imposed. This microstructure has
(1928) and Reuss (1929) averages/estimates in elastic- the advantage that it forms hexagonal (or transversely
ity were in fact upper and lower bounds, respectively. isotropic) crystals locally. Then, if we assume these
Then he proceeded to suggest that estimates of rea- crystals are jumbled together randomly to form an
sonable accuracy were given by the arithmetic or overall isotropic medium, we have the random poly-
geometric means obtained by averaging these two crystal of porous laminates reservoir model. Hashin-
bounds together. Thus, the Voigt-Reuss-Hill estimates Shtrikman bounds are known for such polycrystals
were born. Better bounds than the Voigt and Reuss composed of grains having hexagonal symmetry. So
bounds are now known and no doubt some attempts bounds are easily found. From the form of the bounds,
to update Hills approach have been made. However, we also obtain estimates of both overall bulk modulus
to make a direct connection to traditional approaches and shear modulus, thus completing the semianalytical
of effective medium theory, we apply a more tech- poromechanics model. Examples are computed in
nical approach here in order to obtain estimates of Section 4, and results summarized in the final section.
upscaled constants using the known analytical struc-
ture of the bounds, especially for Hashin and Shtrik-
man (1962) bounds. When this mathematical structure 2 DOUBLE-POROSITY GEOMECHANICS
is not known, then it proves very worthwhile to
expend the effort required to find this structure. Indeed The main results can be derived using uniform expan-
many of these bounds, which are complex nonlinear sion, or self-similar, methods analogous to ideas

79
Copyright 2005 Taylor & Francis Group plc, London, UK
used in thermoelasticity by Cribb (1968) and in drained bulk modulus K (1) of material 1, the corre-
single-porosity poroelasticity by Berryman and Milton sponding Biot and Willis (1957) parameter (1) , and
(1991). Cribbs method provided a simpler derivation the Skempton (1954) coefficient B(1) .The volume frac-
of earlier results on thermoelastic expansion coeffi- tion (1) appears here to correct for the difference
cients. Our results also provide a simpler derivation between a global fluid content and the corresponding
of results obtained by Berryman and Pride (2002) local variable for material 1. The main special charac-
for the double-porosity coefficients. Related meth- teristic of a Gassmann (1951) porous material is that it
ods in micromechanics are called the theory of is composed of only one type of solid constituent, so it
uniform fields by some authors (Dvorak and Ben- is microhomogeneous in its solid component, and in
veniste 1997). addition the porosity is randomly, but fairly uniformly,
First assume two distinct phases at the macro- distributed so there is a well-defined constant porosity
scopic level: a porous matrix phase with the effective (1) associated with material 1, etc.
(1)
properties K (1) , G (1) , Km , (1) (which are drained bulk To proceed, we ask this question: Is it possible
(1) (2) (1)
and shear moduli, grain/mineral bulk modulus, and to find combinations of pc = pc = pc , pf , and
porosity of phase 1 with analogous definitions for (2)
pf so that the expansion or contraction of the system
phase 2), occupying volume fraction V (1) /V = (1) of
the total volume and a macroscopic crack or joint is spatially uniform or self-similar? Or, can we find
phase occupying the remaining fraction of the volume uniform confining pressure pc , and pore-fluid pres-
(1) (2)
V (2) /V = (2) = 1 (1) . sures pf and pf , so all these scalar conditions can
There are three distinct pressures: confining pres- be met simultaneously? If so, then results for system
(1) constants can be obtained purely algebraically with-
sure pc , pore-fluid pressure pf , and joint-fluid
(2) (1) (2) out ever having to solve equilibrium equations. We
pressure pf . Treating pc , pf , and pf as the (1) (2)
initially set pc = pc = pc , as this condition of uni-
independent variables in our double porosity
form confining pressure is clearly necessary for this
theory, we define the dependent variables e
(1) (1) (2) self-similar thought experiment to be a valid solution
V /V , (1) = (V Vf )/V , and (2) = (v of stress equilibrium equations.
(2)
Vf )/V , which are respectively the total volume So, the first condition to be considered is the
dilatation, the increment of fluid content in the matrix equality of the strains of the two constituents:
phase, and the increment of fluid content in the
joints. The fluid in the matrix is the same as that in
the cracks or joints, but the two fluid regions may be
in different states of average stress and, therefore, need
to be distinguished by their respective superscripts.
Linear relations among strain, fluid content, and
pressure take the symmetric form If this condition is satisfied, then the two constituents
are expanding or contracting at the same rate and it is
clear that self-similarity prevails. If we imagine that
(1)
pc and pf are fixed, then we need an appropriate
(2)
value of pf , so that (3) is satisfied. This requires that
following Berryman and Wang (1995) amd Lewallen
and Wang (1998). It is easy to check that 11 = 1/K ,
where K is the overall drained bulk modulus of the
system. We now find the remaining five constants for
a binary composite system.
The components of the system are themselves
porous materials 1 and 2, but each is assumed to be showing that, for undrained conditions, pf can
(2)
what we call a Gassmann material satisfying
almost always be chosen so the uniform expansion
takes place.
(2)
Using (4), we now eliminate pf from the remain-
ing equality so

for material 1 and a similar expression for material


2. The new constants appearing on the right are the

80
Copyright 2005 Taylor & Francis Group plc, London, UK
(2) (1)
where pf (pc , pf ) is given by (4). Making the sub- Performing the corresponding calculation for (2)
(1) produces formulas for a32 and a33 . Since (11) is already
stitution and then noting that pc and pfwere chosen
symmetric in component indices, the formula for a32
independently and arbitrarily, we find the resulting provides nothing new. The formula for a33 easily seen
coefficients must each vanish. The equations we obtain to be identical in form to a22 , but indices 1 and 2 are
are interchanged.
All five of the nontrivial coefficients of double
porosity are now determined.
These results show how the constituent properties
and K, , B up-scale at the macrolevel for a two-constituent
composite. We find

With a11 known, (6) can be solved directly, giving

and
Similarly, with a13 known, substituting into (7) gives

Note that all important formulas [(8), (9), (11)(14)]


depend on the overall bulk modulus K of the system.
So, three of the six coefficients are known. This quantity must be determined independently either
To evaluate the remaining coefficients, we consider by experiment or by another analytical method.
what happens to fluid increments during self-similar
expansion. We treat only material 1, but the equa-
tions for material 2 are completely analogous. From 3 UP-SCALING RANDOM POLYCRYSTALS
the preceding equations, OF POROUS LAMINATES

3.1 Elasticity of layered materials


Next, to determine the overall bulk and shear moduli
of the reservoir, assume a typical building block of the
random system is a small grain of laminate material
whose elastic response for a transversely isotropic
(2) (1)
Again substituting for pf (pc , pf ) from (4) and (hexagonal) system can be described locally by:
noting that the resulting equation contains arbitrary
(1)
values of pc and pf , the coefficients of these
terms must vanish separately. Resulting equations are
a12 + a23 (1 K (2)/K (1) )/(2) = (1) (1) /K (1) , and
a22 + a23 ((1) K (2)/(2) K (1) ) = (1) (1) /B(1) K (1) . Solv-
ing these, we obtain

where ij are the usual stress components for


i, j = 1 3 in Cartesian coordinates, with 3 (or z)
and being the axis of symmetry (the lamination direc-
tion for such a layered material). Displacement ui is
then related to strain component eij by eij = (ui /xj +
uj /xi )/2. This definition introduces some conve-
nient factors of two into the 44, 55, 66 components
of the stiffness matrix shown in (15).
For definiteness we also assume that the stiffness
matrix in (15) arises from the lamination of N isotropic

81
Copyright 2005 Taylor & Francis Group plc, London, UK

constituents having bulk and shear moduli Kn , n , The quantity Gcff is the energy per unit volume in a
in the N > 1 layers present in each building block. It grain when a pure uniaxial shear strain of unit magni-
is important that the thicknesses dn always be in the tude [i.e., (e11 , e22 , e33 ) = (1, 1, 2)/ 6], whose main
same proportionin each of these laminated blocks, compressive strain is applied to the grain along its axis
so that fn = dn / n dn . But the order in which lay- of symmetry (Berryman 2004a,b).
ers were added to the blocks is not important, as The Reuss (1929) average KR for bulk modulus can
Backuss formulas (Backus 1962) for the constants also be written in terms of stiffness coefficients as
show. For the overall behavior for the quasistatic (long
wavelength) behavior of the system we are studying,
Backuss results [also see Berryman (1998), Milton
(2002), Berryman (2004a)] state that
The Reuss average for shear is

r
that defines Geff i.e., the energy per unit volume in a
grain when a pure uniaxial shear stress of unit magni-
This bracket notation can be correctly viewed as a line tude [i.e., (11 , 22 , 33 = (1, 1, 2)/ 6], whose main
integral along the symmetry axis x3 . compressive pressure is applied to a grain along its
The bulk modulus for each laminated grain is that axis of symmetry.
given by the compressional Reuss average KR of the For each grain having hexagonal symmetry,
corresponding compliance matrix sij [the inverse of two product formulas hold (Berryman 2004b):
the usual stiffness matrix cij , whose nonzero compo- v
3KR Geff = 3KV Geff r
= + /2 = c33 (c11 c66 ) c13
2
.
nents are shown in (15)]. The result is e = e11 + e22 + The symbols stand for the quasi-compressional
e33 = /Keff , where 1/Keff = 1/KR = 2s11 + 2s12 + and quasiuniaxial- shear eigenvalues for the crystalline
4s13 + s33 . grains. Thus, it follows that
Even though Keff = KR is the same for every grain,
since the grains themselves are not isotropic, the over-
all bulk modulus K of the random polycrystal does not
necessarily have the same value as KR for the individ- is a general formula, true for hexagonal symmetry.
ual grains (Hill 1952). Hashin- Shtrikman bounds on
K for random polycrystals whose grains have hexag- 3.2.2 Hashin-Shtrikman bounds
onal symmetry (Peselnick and Meister 1965; Watt It has been shown elsewhere (Berryman 2004a,b) that
and Peselnick 1980) show in fact that the KR value the Peselnick-Meister-Watt bounds for bulk modulus
lies outside the bounds in many situations (Berryman of a random polycrystal composed of hexagonal (or
2004a). transversely isotropic) grains are given by

3.2 Bounds for random polycrystals


3.2.1 Voigt and Reuss bounds: hexagonal symmetry
For hexagonal symmetry, the nonzero stiffness v
where Geff v
(Geff ) is the uniaxial shear energy per unit
constants are: c11 , c12 , c13 = c23 , c33 , c44 = c55 , and volume for a unit applied shear strain (stress). The sec-
c66 = (c11 c12 )/2. ond equality follows directly from the product formula
The Voigt (1928) average for bulk modulus of (22). Parameters are defined by
hexagonal systems is well-known to be

Similarly, for the overall shear modulus G , we have


In (24), values of G (shear moduli of isotropic
comparison materials) are given by inequalities

where the new term appearing here is essentially


defined by (18) and given explicitly by and

82
Copyright 2005 Taylor & Francis Group plc, London, UK
The values of K (bulk moduli of isotropic comparison storage material contains no fractures, and therefore is
materials) are then given by algorithmic equalities not sensitive to those mechanical effects on the overall
reservoir. The behavior of a23 also shows little dis-
persion as this value is always quite close to zero.

35
K-
derived by Peselnick and Meister (1965) and Watt and K*
Peselnick (1980). Also see Berryman (2004a). 30
K+

Bulk modulus (GPa)


Bounds on the shear moduli are then given by K-u
25 Ku
K+u
20

15

10
where and are given by
5
0.75 0.8 0.85 0.9 0.95 1
Volume fraction of the storage phase

KV is the Voigt average of the bulk modulus as defined Figure 1. Bulk modulus bounds and self-consistent esti-
previously. mates for the random polycrystal of porous laminates model
of a Navajo sandstone reservoir.

4 EXAMPLE Double porosity coefficients for


navajo sandstone
Navajo sandstone is one possible host rock for which 0.12
the required elastic constants have been measured 0.1 a11
(Coyner 1984). Table 1 displays the values needed in a22
0.08
the double-porosity theory presented here.
The drained bulk moduli of the storage and frac- 0.06
aij (GPa-1)

ture phases are used in the effective medium theory 0.04


of Section 3 to determine the overall drained and 0.02 a23 a33
undrained bulk moduli of the random polycrystal of 0
laminates system. Results for the self-consistent esti- -0.02
mates (Berryman 2004a) and the upper and lower
-0.04 a12
bounds are all displayed in Figure 1. [Note: A cor- a13
rection must be applied to (28) before computing the -0.06
effective constants.] Observed dispersion is small over -0.08
the range of volume fractions considered. Then these 0.75 0.8 0.85 0.9 0.95 1
Volume fraction of the storage phase
drained K values are used in the formulas of Sec. 2
to determine both estimates and bounds on the dou-
Figure 2. Values of double-porosity coefficients aij for a
bleporosity coefficients. These results are displayed in system similar to Navajo sandstone. Values used for the
Figure 2. The results for a11 essentially repeat results input parameters are listed in Table 1. For each coefficient,
shown in Figure 1, but for the inverse of K . three curves are shown, depending on which estimate of the
The coefficients a12 , a22 , and a23 show little dis- overall bulk modulus is used: lower bound (dot-dash line),
persion. This is natural for a12 and a22 because the self-consistent (solid line), or upper bound (dashed line).

Table 1. Input parameters for Navajo sandstone model of double-porosity system.

(1) (2)
Ks Ks K (1) Ks K (2)
(GPa) (GPa) (GPa) (GPa) (GPa) (1) (1) (2) (2)

34.0 34.5 16.5 34.5 1.65 0.15 0.118 0.017 0.354

Note: Poissons ratio and porosity are dimensionless.

83
Copyright 2005 Taylor & Francis Group plc, London, UK
The two coefficients that show a significant level of Berryman, J. G. and S. R. Pride (2002). Models for comput-
dispersion are a13 and a33 , where the third stress is the ing geomechanical constants of double-porosity materials
(2) from the constituents properties. J. Geophys. Res. 107,
pore pressure pf of the fracture or joint phase. We 10.1029/2000JB000108.
generally expect that the joint phase is most tightly Berryman, J. G. and H. F. Wang (1995). The elastic coef-
coupled to, and therefore most sensitive to, the fluctu- ficients of double-porosity models for fluid transport in
ations in overall drained bulk modulus K . So all these jointed rock. J. Geophys. Res. 100, 2461124627.
results are qualitatively consistent with our intuition. Biot, M. A. and D. G. Willis (1957). The elastic coefficients of
the theory of consolidation. J . Appl. Mech. 24, 594601.
Coyner, K. B. (1984). Effects of Stress, Pore Pressure, and
5 CONCLUSIONS Pore Fluids on Bulk Strain, Velocity, and Permeabil-
ity in Rocks. Ph. D. thesis, Massachusetts Institute of
Technology, Cambridge, Massachusetts.
The methods presented have been successfully applied Cribb, J. L. (1968). Shrinkage and thermal expansion of a two
to determine geomechanical parameters for one reser- phase material. Nature 220, 576577.
voir model assuming Navajo sandstone is the host Dvorak, G. J. and Y. Benveniste (1997). On microme-
rock. Although the details differ, the basic ideas used chanics of inelastic and piezoelectric composites. In
above for elastic and poroelastic constants can also T. Tatsumi, E. Watanabe, and T. Kambe (Eds.), Theoreti-
be used to obtain bounds and estimates of electrical cal and Applied Mechanics 1996, Amsterdam, pp. 6581.
formation factor and fluid permeability for the same Elsevier Science.
random polycrystal of porous laminates model. Space Gassmann, F. (1951). ber die Elastizitt porser Medien.
Vierteljahrsschrift der Naturforschenden Gesellschaft in
constraints force us to pursue these issues further Zrich 96, 123.
elsewhere. Hashin, Z. and S. Shtrikman (1962). A variational approach
to the theory of the elastic behaviour of polycrystals. J.
ACKNOWLEDGMENTS Mech. Phys. Solids 10, 343352.
Hill, R. (1952). The elastic behaviour of a crystalline aggre-
gate. Proc. Phys. Soc. London A65, 349354.
Work performed under the auspices of the U.S. Depart- Lewallen, K. T. and H. F. Wang (1998). Consolidation of
ment of Energy by the University of California, a double-porosity medium. Int. J. Solids Structures 35,
Lawrence Livermore National Laboratory under Con- 48454867.
tract No. W-7405-ENG-48 and supported specifically Milton, G. W. (2002). The Theory of Composites. Cambridge,
by the Geosciences Research Program of the DOE UK: Cambridge University Press.
Office of Basic Energy Sciences, Division of Chemical Peselnick, L. and R. Meister (1965). Variational method of
Sciences, Geosciences and Biosciences. determining effective moduli of polycrystals: (A) Hexag-
onal symmetry, (B) trigonal symmetry. J. Appl. Phys 36,
28792884.
REFERENCES Reuss, A. (1929). Berechung der Fliegrenze von Mis-
chkristallen auf Grund der Plastizittsbedingung fr
Backus, G. E. (1962). Long-wave elastic anisotropy produced Einkristalle. Z. Angew. Math. Mech. 9, 4958.
by horizontal layering. J. Geophys. Res. 67, 44274440. Skempton, A. W. (1954). The pore-pressure coefficients A
Berryman, J. G. (1998). Transversely isotropic poroelastic- and B. Geotechnique 4, 143147.
ity arising from thin isotropic layers. In K. M. Golden, Torquato, S. (2002). Random Heterogeneous Materials:
G. R. Grimmett, R. D. James, G. W. Milton, and P. N. Sen Microstructure and Macroscopic Properties. New York:
(Eds.), Mathematics of Multiscale Materials, New York, Springer.
pp. 3750. Springer-Verlag. Voigt, W. (1928). Lehrbuch der Kristallphysik. Leipzig:
Berryman, J. G. (2004a). Bounds on elastic constants for Teubner.
random polycrystals of laminates. J . Appl. Phys. 96, Watt, J. P. and L. Peselnick (1980). Clarification of the
42814287. Hashin-Shtrikman bounds on the effective elastic moduli
Berryman, J. G. (2004b). Poroelastic shear modulus depen- of polycrystals with hexagonal, trigonal, and tetragonal
dence on pore-fluid properties arising in a model of thin symmetries. J. Appl. Phys. 51, 15251531.
isotropic layers. Geophys. J. Int. 127, 415425.
Berryman, J. G. and G. W. Milton (1991). Exact results for
generalized Gassmanns equations in composite porous
media with two constituents. Geophysics 56, 19501960.

84
Copyright 2005 Taylor & Francis Group plc, London, UK

You might also like