Вы находитесь на странице: 1из 15

Material instability with stress localization

J.D. Goddard
Department of Mechanical and Aerospace Engineering
University of California, San Diego, La Jolla, CA 92093-0411, USA
Abstract
This paper is concerned with a possible form of material instability arising from
a non-unique dependence of stress on deformation and leading to heterogeneous
states of stress in otherwise homogeneous deformations. Following a discussion of
certain theories and experiments that suggest this type of stress behavior, an anal-
ysis is given of some possible quasi-static bifurcations of steady-state stress elds
for extensional motions of isotropic elastic solids and viscoelastic uids. These in-
volve extended lamellar and axisymmetric phase structures having common exten-
sion. For certain other restricted forms of nonlinear elasticity, the classical Eshelby
(1957) theory is employed to describe dilute ellipsoidal phases having a deformation
dierent from the globally imposed mean.
Key words: material instability, stress localization, nonlinear rheology, buckling
instability, granular media, polymeric drag reduction, Eshelby theory
PACS: 01.30.-y
1 Introduction
The concept of material (constitutive) instability is now rmly established in
continuum mechanics, particularly in solid mechanics, where it is manifest ex-
perimentally in various modes of localized deformation (strain localization)
and failure, such as necks, shear bands and damage zones[14]. Such instability,
the analog of phase transition in thermostatics, is mathematically associated
with loss of ellipticity in the underlying eld equations and often traceable
to non-convexity [5] in the form of strain softening or other non-monotone
stress-deformation behavior. This is illustrated paradigmatically Fig. 1, where
Email address: jgoddard@ucsd.edu (J.D. Goddard).
Article appearing in Journal of Non-Newtonian Fluid Mechanics102(2002) 251-261

(a) (b) (c)


Fig. 1. Non-monotone stress. (a) Ductile failure. (b) Gradual strain softening (or
shear thinning) with coexistent strains. (c) Catastrophic stress jumps.
Figs. 1(b)-(c) allow for coexistence of dierent states or phases, with the
possibility of hysteretic transitions. Fig. 1(c), which is similar to the behavior
associated with disproportionation and plastic ow in dry liquid foams [6],
is qualitatively dierent from Figs.1 (a)-(b), in that it allows for the possibility
of stress localization, the subject of the present paper.
Although axiomatic in continuum mechanics that the local stress in a material
body is determined uniquely by the past history of deformation, reecting the
so-called Principle of Determinism[7], there are prominent examples to the
contrary in structural mechanics, such as the (Euler) buckling of elastic rods
[8]. Here, as in more complex structures, one has bifurcations between states
with quite dierent axial load-displacement behavior. Therefore, to the extent
that real materials may possess mechanically unstable microstructures, for
example foamed and ber-reinforced solids [914] and liquid crystalline and
micellar systems [1517,19,18], such instability should also be manifest in their
continuum-level behavior.

(a) (b) (c)

Fig. 2. (a) Stress blowup, and stress jumps with coexistent states having (b) the
same, and (c) dierent states of deformation
Fig. 2 provides a schematic illustration of the kind instability envisaged in
the present work. In particular, Fig. 2 (a) illustrates stress blowup in steady
elongational ow, of a type found theoretically for suspensions of elastic spheres
in viscous liquids [20], which resembles that found for elementary linear bead-
2
spring models of polymers [21], or for the bursting of liquid droplets [22,23].
By contrast, Fig. 2(b) illustrates the steady elongational behavior of the coil-
stretch model of de Gennes [2428], as well as the shear behavior predicted
in [29] for gas-solid dispersions in simple shear.
Finally, Fig. 2(c) is intended to suggest the possibility of distinct phases, with
dierences in both stress and strain, a possibility discussed further below. The
schematic diagrams in Figs. 1-2 cannot of course represent the full complexity
of tensorial stress-strain relations, for which the basic issue is one of convexity
in a higher dimensional space [5,30].
As another example, we note that the discontinuous shear thickening in col-
loidal dispersions, observed experimentally by Homan [31] and attributed to
microscopic instability, may be subject to similar phenomenological interpre-
tation. However, the author is unaware of any hysteresis observed in these
measurements. Finally, we should mention buckling instabilities in granular
assemblies, as a plausible explanation of stress inhomogeneity in the form
of force chains [3234], which may reect a bifurcation into strong and
weak phases [35], but perhaps of the buckling type depicted in Fig. 1(c).
The purpose of the present work is to provide a reasonably elementary, yet
general rheological theory of bifurcation and spatial heterogeneity of steady-
state stress in isotropic elastic solids and viscoelastic uids. In the section im-
mediately following, we focus attention on steady extensional motion, which
appears more conducive to coexistent states having common strain but dif-
ferent stress than, say, simple-shear, where the opposite is true [19]. Then,
we consider briey the possible eect of more complex deformations for a re-
stricted class of elastic materials. We note that all the examples illustrated in
Fig. 2 involve critical points with innite derivatives of steady-state stress.
While this suggests a strong instability of perturbed stress elds, a complete
stability analysis would require more complete rheological models than those
considered here.
The aim here is not to work out detailed solutions for specic rheological mod-
els but rather to indicate the possibility of such solutions for a fairly broad
class of constitutive equations. As with countless other viscoelastic phenom-
ena, nonlinear elasticity provides a useful point of departure in the investiga-
tion of nonlinear eects.
2 Quasi-static and piecewise-homogeneous extension
At this point, we restrict the analysis to isotropic materials, recalling that the
constitutive equation for anisotropic elastic solid can be written in the implicit
3
form
H
ij
(T
jk
, B
lm
) = 0, i, j, . . . , m = 1, 2, 3, or H(T, B) = 0 (1)
where T = T
T
is Cauchy stress, B = B
T
the (right Cauchy-Green or Finger)
strain, and H = H
T
is a symmetric isotropic tensor-valued map (technically
speaking, 1
6
1
6
1
6
), such that:
H(QTQ
T
, QBQ
T
) = QH(T, B)Q
T
(2)
for arbitrary real orthogonal tensors Q
T
= Q
1
. We assume that H has a
nite number of piecewise dierentiable branches on which T can expressed
as an explicit function of B and vice versa.
We recall that B can be replaced by any isotropic function of B, e.g.
E =
1
2
(B1), (3)
which reduces to the standard innitesimal strain measure in the small-strain
limit. On the other hand, if E in (3) is replaced in (1) by the stretching (strain
rate)
D =
1
2
v + v
T
, (4)
where v is the velocity gradient, a representation of the form (1) also applies
to the steady pure stretching of simple uids, where by denition D can be
taken as independent of time at a given material point, in a frame where the
vorticity
W =
1
2
v v
T
(5)
vanishes identically.
It is worthwhile to regard (1) as the steady-state limit of a viscoelastic or
hypoelastic model of the form:

T = H(T, v), (6)


where the dot denotes the material derivative. A dynamic model such as (6)
would of course be required for a stability analysis and, hence, a comprehensive
4
bifurcation analysis of the various steady-state solutions of (1) to be identied
below.
We focus attention mainly on two special deformations:
(1) ) homogeneous extension or pure stretch, where the displacement or ve-
locity elds are given on a cartesian system xyz, respectively, by
u =
x
x, v =
y
y, w =
z
z (7)
where the principal strains or strain rates are constant, and
(2) ) axisymmetric extension, where these quantities are given on a cylindrical
polar system rz, respectively, by
u = u(r), v = 0, w =
z
z (8)
with
z
constant.
In both the above cases we can replace B or D and T in (1), respectively, by
diagonal forms, in the rst case by
E = diag(
x
,
y
,
z
), and T = diag(
x
,
y
,
z
), (9)
and in the second by
E = diag(
du
dr
,
u
r
,
z
), and T = diag(
r
,

,
z
), (10)
representing principal values
j
and
j
, j = 1, 2, 3. Then, (1) reduces to three
generally nonlinear equations of the form
h
i
(
j
,
k
) = 0, i, j, k = 1, 2, 3 (11)
with 1,2,3 representing x, y, z or r, , z, respectively.
In the special case of incompressible materials, the three relations (11) reduce
to two relations involving only two each of the deviatoric quantities, say,
j
and
j
which satisfy
3

i=1

i
= 0, and
3

i=1

i
= 0 (12)
For the simple deformations (9) and (10), we now investigate the multiplicity
of solutions of (11) and the quasi-static equation of equilibrium:
T = 0, (13)
5
in spatially unbounded regions.
Lamellar bifurcation
It is clear that (13) admits piecewise homogeneous solutions of the type (9)
provided the tractions and displacements or velocities can be matched at the
boundaries of the various homogeneous regions or phases. One class of such
solutions is obtained with parallel lamellar or smectic phases, lying perpen-
dicular to the x direction, say, as illustrated in Fig. 2 and distinguished here
by primed and unprimed quantities. For example, with all the quantities

y
=

y
,
z
=

z
, and
x
=

x
(14)
specied, (11) provides three equations for
x
,
y
,
z
and three identical
equations for

x
,

y
,

z
.
Of course, if the solution to the above set of simultaneous equations is unique,
then the phases are identical and the corresponding globally homogeneous
state is unique. Otherwise, one has the possibility of a space-lling stack of
distinct lamellae, each of arbitrary thickness and each representing a distinct
solution to the equations in question. Here, as with thermostatic phase tran-
sitions, higher-gradient or non-local eects are required to set length scales
and regularize eld equations, and such eects may also inuence material
instability [42,4,43].
In the case of incompressible materials, we have
x
+
y
+
z
= 0, with only
two equations for the deviatoric stresses
y
,
z
and their primed counterparts.
z
x
y
'
Fig. 3. Lamellar bifurcation
6
Axisymmetric bifurcation
One class of axisymmetric solutions is represented by (8) and (10) with all
relevant variables being independent of and z. Then, (13) reduces to a single
dierential equation:
r
d
r
dr
+
r

= 0 (15)
For the sake of simplicity, we assume that the relations (11) can be solved
explicitly for the s, with
du
dr
= E
1
(
r
,

,
z
) (16)
u
r
= E
2
(
r
,

,
z
) (17)

z
= E
3
(
r
,

,
z
) (18)
Obviously, (15)-(18) represent a set of two ODEs subject to two algebraic
constraints. We note that certain special integrals are given by Green and
Adkins[36] for the case of purely elastic (hyperelastic) solids endowed with
strain-energy function, with considerable simplication arising from the as-
sumption of incompressibility. For given, constant
z
the essential problem is
to integrate these equations, subject to continuity of
r
(r) and u(r) and to
certain regularity conditions at r = 0. To this end, we further assume that
(18) possesses a solution, possibly non-unique, for
z
in terms of
r
,

and
z
,
so that
z
can be eliminated from (16)-(18), after which
z
can be treated as
a constant parameter. Then, letting
:=
u
r
, and := log r, (19)
(15)-(18) can be replaced by the autonomous set
d
r
d
=


r
(20)
d
d
= (
r
,

) (21)
= (
r
,

) (22)
where
7
(
r
,

) :=E
2

r
,

,
z
(
r
,

) (23)
(
r
,

) :=E
1

r
,

,
z
(
r
,

) E
2

r
,

,
z
(
r
,

) (24)
and
z
is an implicit parameter.
Once again, we have simultaneous ODEs (20)-(21) with algebraic constraint
(21), and once again we could employ one of the possibly multiple solutions of
the latter to eliminate

and, hence, to reduce (20)-(21) to simultaneous ODEs


for
r
and . Without making this explicit, we see that the non-uniqueness
associated with leads to the possibility of solution branches exhibiting any
number of radial discontinuities in azimuthal (hoop) stress

, depicted in
Fig. 4, and shear strain , while maintaining continuity of radial stress
r
and
displacement u.
r
'

Fig. 4. Axisymmetric bifurcation
In the case of an incompressible material, the continuity equation:
du
dr
+
u
r
+
z
= 0,
together with regularity (absence of a mass source) at r = 0, imply that
=
z
/2. Then (20)-(22) reduce to
d
d
=
r

= const. where :=
1
2
(
r
+

) (25)
giving the mean in-plane mean stress in terms of the in-plane shear stress

. The latter is piecewise constant by virtue of the deviatoric forms


of (16)-(18), whose multiple solutions allow once more for contiguous regions
with dierent (constant) values of
z
and
r

.
Returning to the compressible material, we note that one possible solution of
(20)-(22) is given by the transversely-isotropic, homogeneous state


r
const., and


r
const. (26)
8
Such a solution serves to represent a far-eld state for r and another
homogeneous state for r R having dierent values for the constants in (26).
The transition between these states is then governed by (20)-(22), subject to
the relevant continuity conditions at r = R.
The above type of solution, or its incompressible counterpart, represents an
axisymmetric lamentary (or annular) structure embedded in a large body of
material in a dierent state in the far eld. Although this structure resembles
the birefringent strands or pipes of [37,38], the latter appear to arise from
a strongly inhomogeneous ow eld. In any event, since the axial stress in
a lament
z
can be much higher than that in the surrounding material, as
suggested by Fig. 2(b), a dilute array of non-interacting laments could make
a major contribution to the axial stress. If the two dierent stress states are
assumed to represent the macromolecular coil-stretch transition envisaged in
[24], the resulting model is relevant to phenomena such as the suppression
of turbulence by high molecular-weight polymers [39] or, for that matter, by
other orientable slender bodies [40,41]. However, this type of problem will
generally involve a dierence in axial extension
z
between lament and far-
eld and, hence, axial shears,
rz
,
rz
, of a type found e. g. in the analysis of
linear-elastic brous composites [44]. This requires a modication of the above
theory, and one possibility is considered next.
3 Dierential Extension
In the preceding examples, at least one principal extension is identical in the
two phases, and the question arises as to possibility of phases having distinct
extensions, as depicted schematically in Fig. 2(c). For example, consider the
axisymmetric extension treated in the preceding section, with imposed global
mean <
z
>
<
z
>=
z
+ (1 )

z
, (27)
where
z
and

z
denote volume averages over two distinct phases occupying
respective volume fractions and 1 . With a similar expression for mean
stress
z
), the mean state is then represented by a point lying on the line
segment connecting points on the upper and lower branches in Fig. 2(c) and
related by a well-known lever rule.
For the general nonlinear form (1) an example has not been found in the
present work of a two-phase structure with completely distinct principal ex-
tensions. However, some progress can be made for the situation, illustrated
schematically in Fig. 5, where the relation (1) involves at least one linear
9

(a) (b) (c)

L
L
L
L*
N
N
Fig. 5. Bifurcations with linear (L) and nonlinear (N) branches.
branch, on which T and E are linearly related, as we now show.
Dilute ellipsoids in a linear matrix
The classical theory of Eshelby [45] for the static stress eld around an el-
lipsoidal inclusion in an innite linear elastic matrix has been applied to a
wide range of problems involving various heterogenieties in elastic bodies [46],
including thermostatic equilibrium between dierent phases of the same mate-
rial. In a similar way, it can be applied to the two-phase structure postulated
in the present work, provided the external or matrix phase is represented by a
linear branch of (1), which we now assume to have the standard linear-elastic
form:
T
ij
= C
ijkl
E
kl
, or T = (E (28)
where ( = [C
ijkl
] represents the elastic constants. Here as below, we employ
upper case script to denote fourth-rank tensors regarded as linear transfor-
mations of second-rank tensors, with essentially all being symmetric. Further-
more, we relax our assumption of isotropy, supposing (1) to be replaced by a
relation appropriate to anistropic materials, which generally involves a depen-
dence on the nite rotation [7].
Following the analysis of Eshelby [45,46], consider an innite linear elastic
region subject to an (innitesimal) homogeneous strain E

at innity and a
displacement u = E
0
x at points x on the boundary
x Ax = A
ij
x
i
x
j
= 1, with A = A
T
= const., (29)
of a solitary, couple-free ellipsoidal inclusion undergoing a hypothetical homo-
geneous strain E
0
. It follows from Eshelbys results that the matrix exerts a
10
traction T
0
n on the surface of the inclusion, where n = Ax/[Ax[ is the unit
outer normal, and T
0
is a constant stress, corresponding to a hypothetical
uniform stress eld inside the inclusion, given by
T
0
= (o
1
E

+ (1 o
1
)E
0
(30)
where 1 is the fourth-rank idemfactor and o
1
is the inverse of the so-called
Eshelby tensor o = [S
ijkl
], which depends on ( and A. Mura [46] lists the S
ijkl
for an isotropic matrix and other special forms of (, and for various special
cases of A (spheroids, elliptical cylinders, etc.).
Since a homogeneous strain E
0
of the inclusion induces the homogeneous stress
T
0
in (30), the pair E
0
, T
0
serves to represent at least one solution to (1) and
(13) in the interior of the ellipsoidal inclusion. Hence, with a linear branch
of (1) representing the innite matrix, a second, possibly nonlinear branch
represents the ellipsoidal inclusion. This idea underlies the suspension model
of Roscoe [20] and has also been invoked recently by Ledbetter and Dunn [47]
in their treatment of the thermostatic phase transition between a nonlinear
elastic inclusion and a linear elastic matrix.
In the case of nite, non-interacting ellipsoids with volume fraction << 1,
it is therefore possible to have a strain E
0
in the inclusion which diers from
the mean strain E

in the matrix, neither of which is restricted to purely


extensional deformations. Then, given an imposed global strain
E) = E
0
+ (1 )E

, (31)
and T
0
, E
0
, E

presumably can be expressed in terms of E), by means of


(31), (30) and the appropriate branch of (1). This is particularly obvious in
the special case of a linear-elastic inclusion, represented schematically by Fig.
5(c), since the substitution of the form T
0
= (

E
0
into (30) immediately
yields a linear relation between E
0
and E

involving the elasticities (, (

of
matrix and inclusion. The resulting linear-elastic formula can also be derived
by somewhat less direct methods [46].
As a nal issue, we note that a solitary cylindrical or lamellar structure of the
type considered in the preceding sections can be represented, respectively, by
an ellipsoid having one or two innite axes. In this case, it follows from the
formulae given by Mura [46] (pp. 80-88) for linear isotropic elastic matrices
that the Eshelby tensor o becomes singular, with unbounded inverse o
1
. This
suggests that, barring some extreme rheological nonlinearity giving rise to a
virtual slip between phases, such as that envisaged in the nonlinear uid
model of [48], it is generally not possible to sustain a dierential extension
between indenitely extended phases, in directions lying parallel to the phases.
11
While this conjecture is relevant to a number of elds, particularly composite
materials, its resolution would take us well beyond the scope of the present
article.
In closing, it should be emphasized that, in contrast to the preceding sections,
the analysis of the present section is generally restricted to elastic solids, since
the representation (1) is valid at best as an approximation for restricted classes
of viscoelastic uids.
4 Conclusions
The preceding analysis serves to establish the theoretical possibility of hetero-
geneous stress elds arising from rheological models with non-unique stress. If
the admissible stress states have widely disparate magnitudes, this serves to
dene stress localization, which is to be distinguished the more familiar strain
localization usually associated with material instability. However, the steady-
state analysis of the present paper generally does not provide a description of
the evolution and stability of various heterogeneous states.
The possible importance to several phenomena, such as force transmission in
particulate media or the suppression of uid turbulence by polymer additives,
would seem to warrant further theoretical and experimental investigation.
Acknowledgement
Partial support from the U.S. National Aeronautics and Space Administration
(Grants NAG3-1888 and NAG3-2465), and the National Science Foundation
(Grant CTS-9510121) is gratefully acknowledged. Special thanks are due to
Professor Andreas Acrivos, Mentor worthy of Odysseus.
References
[1] S. S. Antman, E. R. Carbone, Shear and necking instabilities in nonlinear
elasticity, J. Elast. 7 (1977) 125151.
[2] R. Abeyaratne, N. Triantafyllidis, The emergence of shear bands in plane strain,
Int. J. Solids Struct. 17 (1981) 11131134.
[3] I. Vardoulakis, Rigid granular plasticity model and bifurcation in the triaxial
test, Acta Mech. (Austria) 49 (1983) 5779.
12
[4] J. D. Goddard, M. Alam, Shear-ow and material instabilities in particulate
suspensions and granular media, Particulate Sci. Tech. 17 (1999) 6996.
[5] J. M. Ball, Convexity conditions and existence theorems in nonlinear elasticity,
Arch. Ration. Mech. Anal. 63 (1977) 337403.
[6] D. A. Reinelt, A. M. Kraynik, Simple shearing ow of dry soap foams with
tetrahedrally close-packed structure, J. Rheology 44 (2000) 453471.
[7] C. Truesdell, W. Noll, The nonlinear Field Theories of Mechanics, Vol. III/3 of
Encyclopedia of Physics, Springer-Verlag, New York, 1965.
[8] A. Love, A Treatise on the Mathematical Theory of Elasticity, 4th Edition,
(Cambridge University Press) Dover, New York, 1944.
[9] M. Kurashige, Instability of a ber-reinforced elastic slab subjected to axial
loads, Trans. ASME, J. Appl. Mech. 46 (1979) 839843.
[10] R. Abeyaratne, N. Triantafyllidis, An investigation of localization in a porous
elastic material using homogenization theory, Trans. ASME, J. Appl. Mech. 51
(1984) 481486.
[11] W. E. Warren, A. M. Kraynik, C. M. Stone, A constitutive model for two-
dimensional nonlinear elastic foams, J. Mech. Phys. Solids 37 (1989) 717733.
[12] G. Geymonat, S. Muller, N. Triantafyllidis, Homogenization of nonlinearly
elastic materials, microscopic bifurcation and macroscopic loss of rank-one
convexity, Arch. Ration. Mech. Anal. 122 (1993) 231290.
[13] R. M. Christensen, Material instability for ber composites, Trans. ASME, J.
Appl. Mech. 61 (1994) 476477.
[14] H. X. Zhu, N. J. Mills, J. F. Knott, Analysis of the high strain compression of
open-cell foams, J. Mech. Phys. Solids 45 (1997) 18751904.
[15] M. Delaye, R. Ribotta, G. Durand, Buckling instability of the layers in a
smectic-a liquid crystal, Phys. Lett. A 44A (1973) 139140.
[16] M. Cagnon, M. Gharbia, G. Durand, Column buckling instability in a discotic
liquid crystal, Phys. Rev. Lett. 53 (1984) 938940.
[17] P.D. Olmsted, C.-Y.D. Lu, Coexistence and phase separation in sheared
complex uids, Phys. Rev. E 56 (1997) R5558.
[18] P.D. Olmsted, C.-Y.D. Lu, Phase separation of rigid-rod suspension in shear
ow, Phys. Rev. E 60 (1999) 43974415.
[19] P.D. Olmsted, Two-state shear diagrams for complex uids in shear ow,
Europhys. Lett. 48 (1999) 339345.
[20] R. Roscoe, On the rheology of a suspension of viscoelastic spheres in a viscous
liquid, J. Fluid Mech. 28 (1967) 273293.
[21] R. C. Bird, R.B. Armstrong, O. Hassager, Dynamics of polymeric liquids, Wiley,
New York, 1977.
13
[22] D. Barthes-Biesel, A. Acrivos, Deformation and burst of a liquid droplet freely
suspended in a linear shear eld, J. Fluid Mech. 61 ( 1973) 121.
[23] A. Acrivos, T. S. Lo, Deformation and breakup of a single slender drop in an
extensional ow, J. Fluid Mech. 86 pt.4 (1978) 641672.
[24] P. G. de Gennes, Coil-stretch transition of dilute exible polymers under
ultrahigh velocity gradients, J. Chem. Phys. 60 (1974) 50305042.
[25] J. M. Wiest, L. E. Wedgewood, R. B. Bird, On coil-stretch transitions in dilute
polymer solutions, J. Chem. Phys. 90 (1989) 587594.
[26] M. L. Manseld, The coil-stretch transition of polymers in external elds, J.
Chem. Phys. 88 (1988) 65706580.
[27] E. M. Sevick, D. R. M. Williams, Coil-stretch transitions for grafted polymers
in spatially varying ows, Europhysics Letters 31 (1995) 357362.
[28] Y. Termonia, Coil-stretch transition in deformation ows, J. Polymer Sci., Part
B 38 (2000) 24222428.
[29] T. Heng-Kwong, D. L. Koch, Simple shear ows of dilute gas-solid suspensions,
J. Fluid Mech. 296 (1995) 211245.
[30] R.W. Ogden, nonlinear Elastic Deformations, Ellis Horwood/Halsted/ Wiley,
1984.
[31] R. L. Homan, Discontinuous and dilatant viscosity behavior in concentrated
suspensions. ii. theory and experimental tests, J. Colloid Interface Sci. 46 (1974)
491506.
[32] A. Drescher, G. De-Josselin-de Jong, Photoelastic verication of a mechanical
model for the ow of a granular material, J. Mech. Phys. Solids 20 (1972) 337
351.
[33] J. D. Goddard, Nonlinear elasticity and pressure-dependent wave speeds in
granular media, Proc. Roy. Soc. Lond. A 430 (1990) 105131.
[34] L. Vanel, D. Howell, D. Clark, R. P. Behringer, Memories in sand: Experimental
tests of construction history on stress distributions under sandpiles, Phys. Rev.
E, Stat. Phys. Plasmas Fluids Relat. Interdiscip. Top. 60 (1999) R50405043.
[35] F. Radjai, D. E. Wolf, M. Jean, J.-J. Moreau, Bimodal character of stress
transmission in granular packings, Phys. Rev. Lett. 80 (1998) 6164.
[36] A. Green, J. Adkins, Large Elastic Deformations, 2nd Edition, Oxford
University Press, London, 1970.
[37] O. G. Harlen, J. M. Rallison, M. D. Chilcott, High-deborah-number ows of
dilute polymer solutions, J. Non-Newton. Fluid Mech. 34 (1990) 319349.
[38] O. G. Harlen, E. J. Hinch, J. M. Rallison, Birefringent pipes: the steady ow of
a dilute polymer solution near a stagnation point, J. Non-Newton. Fluid Mech.
44 (1992) 229265.
14
[39] J. L. Lumley, Drag reduction in turbulent ow by polymer additives, J. Polym.
Sci. Macromol. Rev. 7 (1973) 263290.
[40] G. K. Batchelor, Slender-body theory for particles of arbitrary cross-section in
stokes ow, J. Fluid Mech. 44 pt.3 (1970) 419440.
[41] C. B. Weinberger, J. D. Goddard, Extensional ow behavior of polymer
solutions and particle suspensions in a spinning motion, Int. J. Multiph. Flow
1 (1974) 465486.
[42] N.Triantafyllidis, E.C. Aifantis A gradient approach to localization of
deformation. I. Hyperelastic materials, J. Elast., 16 (1986) 225237.
[43] C.-y. D. Lu, P.D. Olmsted, R.C. Ball, Eects of Nonlocal Stress on the
Determination of Shear Banding Flow, Phys. Rev. Lett 84 (2000) 642645.
[44] W. B. Russel, A. Acrivos, On the eective moduli of composite materials:
slender rigid inclusions at dilute concentrations, Z. Angew. Math. Phys. 23
(1972) 434464.
[45] J.D. Eshelby, The determination of the elastic eld of an ellipsoidal inclusion
and related problems, Proc. Roy. Soc. A241 (1957) 376396.
[46] T. Mura, Micromechanics of defects in solids, Second, revised edition, Martinus
Nijho Publishers, 1987.
[47] H. Ledbetter, M.L. Dunn, Equivalence of Eshelby inclusion theory
and Wechsler-Lieberman-Read, Bowles-Mackenzie martensite crystallography
theory, Mat. Sci. Eng. A285(2000) 180-185.
[48] J.D. Goddard, Tensile Behavior of Power-Law Fluids Containing Oriented
Slender Fibers, J. Rheology 22 (1978) 615-622.
15

Вам также может понравиться