Badar 2020
Badar 2020
Badar 2020
https://doi.org/10.1007/s11242-020-01483-0
Abstract
A successful prediction of the response of poroelastic material to external forces depends
critically on the use of appropriate constitutive relation for the material. The most com-
monly used stress–strain relation that captures the behavior of poroelastic materials satu-
rated with a liquid is that proposed by Biot (J Appl Phys 12:155–164, 1941). It is a linear
theory akin to the generalized Hooke’s law containing material constants such as the effec-
tive bulk and shear modulus of the porous material. However, the effective elastic coeffi-
cients are not known a priori and need to be determined either via separate experiments or
by fitting predictions with measurements. The main cause for this drawback is that Biot’s
theory does not account for the microstructural details of the system. This limitation in
Biot’s model can be overcome by utilizing the constitutive relation proposed by Russel
et al. (Langmuir 5(24):1721–1730, 2008) for the case of colloidal packings. We show that
in the linear limit, the constitutive relation proposed by Russel and coworkers is equivalent
to that of Biot. The elastic coefficients obtained from such a linearization are related to the
micro-structural details of the packing such as the particle modulus, the packing concentra-
tion and the nature of packing, thereby enabling a more effective utilization of Biot’s model
for problems in the linear limit. The derivation ignores surface forces between the particles,
which makes the results also applicable to particles whose sizes are beyond the colloidal
range. We compare the predictions of Biot’s model to those of the linearized model of Rus-
sel and coworker’s for two different one-dimensional model problems: fluid outflow driven
by an applied mechanical load, also termed as the consolidation problem, and wave propa-
gation in a saturated colloidal packing.
* Mahesh S. Tirumkudulu
mahesh@che.iitb.ac.in
Atiya Badar
atiyabadar789@gmail.com
1
Department of Chemical Engineering, Indian Institute of Technology Bombay,
Powai, Mumbai 400076, India
13
Vol.:(0123456789)
A. Badar, M. S. Tirumkudulu
1 Introduction
The formation of saturated colloidal packing during drying is common to many applica-
tions such as in paints, inks, ceramics industry, cosmetics, tablet coating and mud/clay
structures. In many cases, the final dried product is required to be homogeneous and free
of defects. However, drying is often accompanied by processes that generate stresses in the
film leading to their eventual failure. The origin of stresses and their magnitude is closely
linked to the mechanical properties of the colloidal packing and therefore a thorough
understanding of their mechanical properties is needed to improve performance.
The mechanical properties of such saturated packing of particles have typically been
described by the linear poroelasticity model of Biot (1941, 1952, 1955, 1956a, b, 1962a, b)
and Biot and Willis (1957), which is an extension of the generalized Hooke’s law for linear
elastic materials. The total stress in the packing is a sum of the stress in the elastic particle
network and the pore pressure exerted by the interstitial liquid, while the water content
in the packing is related to both the stress and the pore pressure. The elastic coefficients
present in the model are unknowns and their relation to the microstructure of the packing
is not known either. These coefficients are determined separately from experiments such as
the compressibility tests (Biot and Willis 1957; Bemer 2001; Alexander 2015). The linear
theory can be extended further to large deformation (nonlinear) by adopting expressions
from the second-order elasticity for the deformation of the particle network. This results
in additional parameters, Biot (1973), Bowen (1982), Bemer (2001), Coussy (2010) which
again have to be determined from experiments.
In contrast to the aforementioned approach, the constitutive relation of Russel et al.
(2008) (henceforth referred to as Russel’s model) proposed for a colloidal packing satu-
rated with liquid is derived by considering the Hertzian contact model of a pair of contact-
ing particles in the packing (Walton 1987). The model was an improvement over an earlier
approach, Routh and Russel (1999) which considered an approximate deformation model
for a pair of contacting neighbors. Consequently, the constitutive relation incorporates the
micro-structural details of the system such as particle modulus, particle volume fraction
and packing characteristics. Such a first-principles approach connects the macroscopic
behavior of the packing to the properties of its constituents enabling a better understand-
ing of the phenomenon under study. The constitutive relation has been used in recent years
to describe the stress evolution in drying films of colloidal dispersions (Routh and Russel
1999; Tirumkudulu and Russel 2004, 2005; Russel et al. 2008), cracking in drying colloi-
dal films (Singh and Tirumkudulu 2007; Sarkar and Tirumkudulu 2011; Routh 2013) and
buckling of drying drops of colloidal dispersions (Tirumkudulu 2018). While these studies
have employed the linearized version of Russel’s model to solve specific problems, there is
a need to derive the general theory for linear, three-dimensional deformation of colloidal
packings with the goal of relating the micro-structural information of the packing to the
elastic constants of the Biot’s model.
In the present work, we consider a packed bed of particles saturated with liquid and
under stress. First, the nonlinear constitutive relation of Russel et al. (2008) is solved
for the case of one-dimensional soil consolidation problem and the results are compared
numerically, to those obtained using the linear theory of Biot (1941). The numerical value
of the effective elastic coefficients of Biot’s model is obtained by matching the predictions
with those of the nonlinear model. Next, we linearize the nonlinear constitutive relation
in the vicinity of a prestressed state to obtain a linear equation that relates the perturbed
stress to the perturbed strain. The latter is compared with the linear constitutive relation of
13
Mechanics of Saturated Colloidal Packings: A Comparison of…
Biot to show that the two relations are identical. More importantly, matching yields ana-
lytical expressions for the effective elastic coefficients in Biot’s model in terms of micro-
structural details of the particle packing. These results are confirmed by comparing the
elastic constants determined from the consolidation problem with that obtained from the
linearization. The importance of these results can be gauged by comparing them to the
existing models for isotropic linear elastic packings, such as the celebrated Hashin–Shtrik-
man bounds (Hashin and Shtrikman 1963) and the Voigt and Reuss bounds (Reuss 1929).
While these models account for the elastic properties of the individual phases, they nei-
ther incorporate information related to the geometry of the phases nor do they capture the
intrinsically nonlinear deformation of spheres subject to external and surface forces, which
is explicitly accounted for in Russel’s model. Further, when the constituents are very dif-
ferent, such as hard particles surrounded by liquid as in the present case, the bounds are
very separated due to which their applicability and predictive value become very limited
(Mavko et al. 2009). Thus, the novelty of the present work lies in the expressions of the
effective elastic coefficients that are related explicitly to the particle modulus, volume frac-
tion and the nature of packing for a saturated packing of colloidal spheres. The present
study focuses on the properties of saturated colloidal packings and will not consider the
mechanical properties of partially saturated or dry colloidal packings (Lesaine et al. 2018).
In the final section of the paper, we present the differential mass and momentum balance
equations for saturated colloidal packings that are obtained from the linearized equation
of Russel’s and Biot’s stress–strain relations. These equations are useful in describing the
dynamics of poroelastic materials such as crack dynamics and wave propagation. As an
application of the derived equations, we solve the dynamic equations for the problem of
one-dimensional wave propagation through a saturated colloidal packing and compare the
results from the two models.
Biot’s theory of consolidation has been applied extensively to study mechanical behavior
of porous elastic material containing a fluid (Biot 1941, 1952, 1955, 1956a, b, 1962a, b;
Biot and Willis 1957). It is a linear model for stress–strain relation in the porous elastic
material containing a fluid. The material may be either isotropic or anisotropic and the
fluid either compressible or incompressible. It assumes reversibility of stress–strain rela-
tion and being a linear theory, it is applicable for small strains. Here, we will be presenting
equations derived for the isotropic case.
The total stress 𝝈 for a porous elastic material is written as the sum of the contribution
of the particle stress, 𝜙𝝈 p, which is the force supported by the particle per unit total area
and the contribution from the fluid stress, (1 − 𝜙)𝝈 f , which is the force supported by the
fluid per unit total area,
𝝈 = 𝜙𝝈 p + (1 − 𝜙)𝝈 f , (1)
where 𝜙, is the solid volume fraction, 𝝈 and 𝝈 , are the stress in the particle network and
p f
the fluid, respectively. The above relations suggest that the fluid and the particle colloidal
packing must jointly support the local mechanical load to provide mechanical equilibrium.
For the fluid within the porous solid, the flow generated by the shear component of the
stress is negligible relative to that generated by the pressure gradient at the continuum
scale. This can be understood from the following simple argument: the fluid flow in the
13
A. Badar, M. S. Tirumkudulu
𝜅p |𝛁P|
porous solid is related to the pressure gradient via the Darcy’s law, V ∼ 𝜂 Lc
with a shear
rate of 𝛾̇ ∼ √
V
𝜅p
in the pores. Here, 𝜅p ∼ a2 is the permeability of the packing, a is the parti-
cle size, 𝜂 is the viscosity of the liquid, and Lc is the length scale corresponding to the mac-
roscopic dimensions of the packing. The shear stress in the fluid phase becomes, 𝜏d ∼ 𝜂 𝛾̇
resulting in |∇P| ∼ La << 1 for very small pore size (assuming pore size of the same order
𝜏d
c
as particle size). Thus, 𝝈 f ≈ −P𝜹 to a good approximation, where P is the pressure in the
pores.
Based on the aforementioned assumptions, Biot proposed the following stress–strain
equations, which can be written in terms of the measurable elastic coefficients (Biot and
Willis 1957),
𝝈 = 𝜆tr(𝝐 p )I + 2𝜇𝝐 p − 𝛼PI (2)
P = M𝜁 − 𝛼Mtr(𝝐 p ), (3)
where Eq. (2) shows that the total effective stress in the bulk material is a combination of
solid part of deformation represented by Hooke’s law and the isotropic pressure for fluid
phase. In addition, Eq. (3) relates pressure in the fluid phase with the mechanical deforma-
tion in the solid part (dilatation term, tr(𝝐p p )) and changes in the fluid volume per unit refer-
( 𝜕up 𝜕u )
ence volume (𝜁 ). Here, 𝜖ij = 21 𝜕xi + 𝜕xj is the component of solid strain, tr(.) represents
p
j i
the trace, 𝜁 = 𝛁 ⋅ (1 − 𝜙)(uf − up ), up and uf are the displacement fields of solid and fluid,
respectively, and 𝛼 measures the ratio of fluid volume squeezed out to the volume change
of solid when compressed. Most importantly, 𝜆 and 𝜇 are elastic coefficients and 1/M meas-
ures the amount of solvent that can be forced into the material under pressure while keep-
ing the volume of the material constant. It is important to note that the elastic coefficients,
𝜆 and 𝜇, are drained properties, indicating that they are mechanical properties of the parti-
cle network alone and must be measured under quasistatic conditions where the fluid is
allowed to drain (leave) or enter freely (MacMinn et al. 2016). The other constants, 𝛼 and
M, give a measure of the compressibility of the material. For saturated and incompressible
porous elastic material containing an incompressible liquid, 𝛼 = 1 and M = ∞ for which
Eq. (3) becomes 𝜁 = tr(𝝐 p ).
Biot and Willis (1957) also derived the general relation between stress and strain in a
prestressed material wherein, it is assumed that the material is isotropic in both prestressed
and incremental/perturbed state. The linearization about the prestressed state relates the
incremental stress to strain and is applicable for small deformations in the vicinity of pre-
stressed state,
(5)
�
△𝜎 f =Q tr(𝝐 p ) + R(𝛁 ⋅ uf )
Here, 𝛥𝝈 and 𝛥𝜎 f are the incremental stresses of the bulk material and the fluid phase,
respectively, in the deformed state. The incremental stress–strain relation based on isotropy
alone shows that there are five elastic coefficients, namely, A, Nb , Q, Q and R.
′
In contrast, Russel’s model (2008) relates the macroscopic stress 𝝈 in the system to
the macroscopic strain 𝝐 (same as 𝝐 p in the Biot’s model) through a nonlinear constitutive
relation, which accounts for the micro-structural details of the porous media. The model
assumes the material to be composed of a compact network of colloidal particles wherein
13
Mechanics of Saturated Colloidal Packings: A Comparison of…
the void spaces are filled with solvent. Particles present in the network are spheres of radius
a, and are viscoelastic in nature. The model focuses on a pair of particles that experi-
ence force ± F exerted by the neighbors, which acts on the particles and is aligned along
the center-to-center vector n of the particles, are shown in Fig. 1. The model ignores any
tangential force on the pair assuming perfect slip at contact point. In addition to the con-
tact force, the particle pair also experiences an interfacial attractive force of order O(𝛾pl ),
which pulls the two particles towards each other at the contact point. Here, 𝛾pl is the surface
energy of the particle–liquid interface since the packing is assumed to be saturated with
liquid. Assuming the strain amplitude to be small |𝝐| << 1, the relation between force and
deformation at the particle level is obtained as
t ( � )
3 2(1 − 𝜈p ) ∫−∞
16 G t −t d 3 �
F + 2.83𝜋a𝛾pl n = − a2 × exp �
(−n.𝝐.n) 2 ndt (6)
𝜏r dt
where, the particle is assumed to be a viscoelastic fluid with a single relaxation time,
2𝜂E(1−𝜈 )
𝜏r = HG p and, G and 𝜈p are, respectively, the shear modulus and the Poisson’s ratio of
the particle. The nonlinearity arises from assuming a Hertzian deformation at particle con-
tact leading to a power-law behavior with an exponent of 3/2.
The next step in the model development is to volume average over a representative vol-
ume containing particles and the surrounding fluid. Since the pressure is spatially homoge-
neous at this length scale, the volume integral over the particle is transformed to a surface
integral via the divergence theorem. This leads to the total surface dipole S that accounts
for the contact force ( F ) exerted by the contacting particles, which for a single contacting
particle becomes,
S = anF
The bulk stress tensor in the saturated colloidal film is then obtained by ensemble averag-
ing for all possible orientation of surface dipoles induced in a particle from all contacting
neighbors,
(4𝜋)2 a3 ∫
3𝜙N
𝝈 = −P𝜹 + Sd2 n
t ( � ) (7)
𝜋 ∫−∞ ∫ dt
1̄ t −t d 1 �
= −P𝜹 + G exp �
(|n.𝝐.n| 2 n.𝝐.n)nnd2 ndt
𝜏
Here, d2 n is the elemental surface area over the sphere’s surface (see “Appendix B”), P is
the pressure in the solvent phase of the packing and G = 2𝜋(1−𝜈 𝜙NG
)
is the effective elastic
p
modulus containing all the micro-structural details of the saturated colloidal packing with
N being the average number of contacting neighbors. In the above derivation, the repre-
sentative volume element (RVE) is the pair of contacting particles surrounded by the fluid
13
A. Badar, M. S. Tirumkudulu
over which integration is performed for all possible orientations of the contacting particles
to obtain the macroscopic stress and strain. Since all contact orientations are assumed in
the homogeneous material, this ensures that RVE is larger than the size of single particle
(and its specific contact distribution) but smaller than the variation at the macroscopic
scale.
The above expression (7) assumes hard particles of O(GPa) with sizes greater than O
(10nm) so that the contribution of the surface energy to the deformation of the particles
becomes negligible compared to the contact forces (𝛾pl ∕Ga ≪ 1). Consequently, the consti-
tutive relation is also applicable to particles whose size is beyond the colloidal range. Fur-
ther, the deformation is assumed to be affine, namely, the deformation at the particle level
is the same as that at the macroscopic level. While this assumption is correct for crystal
packings, errors can be expected for random packing.
To compare Russel’s model with that of Biot, it is necessary that we impose the same
restrictions as those used to derive the Biot’s model, namely that the particle phase is elas-
tic and the material is isotropic. Imposing these restrictions reduces Eq. (7) to
𝜋 ∫
G
𝝈 = −P𝜹 + (|n.𝝐.n|1∕2 n.𝝐.n)nnd2 n. (8)
The remaining part of the paper shall use the above expression for deriving expressions
from Russel’s model. In the next subsection, we solve the well known problem of soil con-
solidation using the two models to obtain the numerical values of elastic constants of the
Biot’s model from the results obtained using Russel’s model.
2.1 Soil Consolidation
Let us consider the prototypical case of a column of soil supporting a load 𝜎33 = −po along
the x3 direction, while the lateral expansion of soil is restricted by impermeable side walls
(see Fig. 2). Further, the liquid is not allowed to escape through the bottom surface. The
top slab, which applies the load, is assumed to be porous so that the liquid can escape
freely through it. In such a situation, the pressure of water at the top of the column is zero
( P = 0, at x3 = 0) while the gradient of water pressure at the bottom of the column is zero
13
Mechanics of Saturated Colloidal Packings: A Comparison of…
( 𝜕x
𝜕P
= 0 at x3 = h). Here, h is the initial height of the column with x3 being positive in the
3
downward direction.
The strain in the system is 𝝐 = 𝜖33 e3 e3 since the side walls are smooth and the compres-
sion of soil is only along the x3 direction. The stress tensor from Russel’s model (8) can
be obtained by evaluating the surface integral over all orientation of surface dipoles (see
“Appendix B”),
̄ (1
G
) 1
𝜎ij = 𝛿ij + 𝛿i3 𝛿3j |𝜖33 | 2 𝜖33 − P𝛿ij (9)
2 3
Because of the top load (𝜎33 = −po), the strain can be related to the pressure,
1 3
|𝜖33 | 2 𝜖33 =
̄
(P − po ) (10)
2G
The two possible solutions for 𝜖33 are,
( )2
3 3 2
(11)
𝜖33 = (P − po ) 3
2Ḡ
or
( )2
3 3 2
(12)
𝜖33 =− (−P + po ) 3 ,
2Ḡ
where, V3 and V3 , are the velocities of the particle and liquid phase in the x3 direction,
p f
respectively, and 𝜙 is the particle volume fraction. The transport of the fluid phase through
the porous media is described by Darcy’s law,
f p
𝜅p 𝜕P
(1 − 𝜙)(V3 − V3 ) = − . (14)
𝜂 𝜕x3
Recognizing that 𝜕V3 ∕𝜕x3 ≡ 𝜕𝜀33 ∕𝜕t , (14) is substituted in (13) to relate the time variation
p
where the permeability is assumed to a constant. On substituting 𝜖33 from (12) into the
above equation (since P − po < 0), the non-dimensional consolidation equation from Rus-
sel’s model becomes,
𝜕 P̄ 2̄
̄ 3 𝜕 P,
1
𝜕 ̄t
= KR (−P) (16)
𝜕 x̄ 32
13
A. Badar, M. S. Tirumkudulu
( 1 )
po3 𝜏𝜅p
where the diffusion coefficient of consolidation in the problem is KR = 3
with
aR ≡
2∕3
2 h2 aR 𝜂
𝜕 P̄ 𝜕 2 P̄
𝜕t ̄
= KB 2 (17)
𝜕 x̄ 3
( )
𝜏𝜅p
where the corresponding diffusion coefficient of consolidation is KB = ,
aB ≡
h2 aB 𝜂
1−2𝜈
=
2𝜇(1−𝜈)
is Biot’s elastic coefficient and 𝜈 is the Poisson’s ratio of the packing.
1
𝜆+2𝜇
The partial differential equations for both models are subject to the following bound-
ary conditions (Fig. 3),
P̄ = −1, x̄ 3 = 0
𝜕 P̄
= 0, x̄ 3 = 1 (18)
𝜕 x̄ 3
P̄ = 0, ̄t = 0
The last condition comes from the fact that the strain should be zero initially.
The equations are solved numerically. For discretization, we used the forward differ-
ence scheme for first-order time derivative, while the central difference scheme was
used for the second-order derivative in space. The nonlinear coefficient containing the
pressure term in Russel’s model was calculated by taking the average value of the pres-
sure values from the forward and backward node at the previous time step. This ensured
smooth convergence at each time step. The stability of finite difference schemes for dif-
fusion equations was achieved by ensuring KB,R 𝛥x 𝛥t
2 <
1
2
. It should be noted that there
3
exists an analytical solution to the consolidation problem using Biot’s model in terms of
a Fourier series. Given that the accuracy of the solution depends on the number of terms
considered, we solve numerically the Biot’s model for comparison.
The pressure field for the two models is plotted in Fig 3 at three different time steps.
The pressure increases from top to bottom of the column and reaches a steady-state after
a long time ( ̄t ≳ 1). Water is transported from the interior where the pressure is high to
the surface with lower pressure. This continues till the load is fully supported by the
13
Mechanics of Saturated Colloidal Packings: A Comparison of…
compacted particle network at the top and fluid pressure at the bottom. At this point, the
top plate stops moving and liquid is no longer squeezed out of the porous network.
In plotting Fig 3, the value of KB and KR has been adjusted so that the pressure profiles
from the two models are close to each other for all times. We found a good match when
KB ≈ 2R , which relates the elastic properties of the Biot’s model in terms of the elastic
K
where the terms on the left apply to the macroscopic properties of the packing (subscript
B) while that on the right pertain to the particles (subscript R). These results are general
for the one-dimensional consolidation problem and effective properties can be determined
once the external pressure is known. Since the above methodology does not yield the
expression for an effective Poisson’s ratio, one may assume the same value as the particles.
A proper linearization, which is detailed in the next section, eliminates this problem.
The importance of Russel’s model can be appreciated when we consider a specific
example of soil consolidation for a column of soil composed of silica particles, which is
saturated with water and undergoes the aforementioned consolidation process. The Pois-
son’s ratio and the shear modulus of the silica particles are 𝜈p = 0.19 and G = 32.3GPa ,
respectively. The viscosity of water is 𝜂 = 10−3 Pa s while the particle volume fraction is
taken to be that for random close packing, 𝜙 = 0.64 along with N = 6. The value of 𝜏 is set
as 1s although the value could also be obtained by setting KB = 1 and then determining the
value of 𝜏 . In that case, the chosen value of 𝜏 = 1s corresponds to a bed thickness of O(m)
for particles of size O(𝜇m). The permeability of the packing is given by the well known
3 a2
Carman–Kozeny equation (McCabe et al. 2005), 𝜅p = (1−𝜙) 45𝜙2
, which for micron sized par-
ticles is about 2.53 × 10 m . Generally, permeability changes with pore structure since
−15 2
the particles are expected to deform, but for small deformations, this dependence can be
neglected. Finally, we assume a normal stress of po = 1 MPa.
Using the aforementioned numerical values, we obtain KR = 2.42 × 10−3. Since the
two pressure fields match when KB ≈ KR ∕2, Biot’s effective shear modulus for the packing
𝜈 = 𝜈p is obtained as 𝜇 = 0.18GPa , which is two orders lower than that for the particles.
Clearly, the effective shear modulus for the system is lower than the shear modulus of the
particles, and the above exercise enables determination of the effective elastic constants for
the Biot’s model from the results of the nonlinear model of Russel and coworkers.
Figure 4 shows the corresponding strain in the soil at each time step determined from
Biot’s and Russel’s model. The strain in the soil increases from the top to the bottom of
the column and varies with time until it reaches steady value at long times. Here, the strain
13
A. Badar, M. S. Tirumkudulu
for the Biot’s model has been evaluated from the effective shear modulus obtained earlier
(Fig. 3). Since Russel’s theory is nonlinear in strain, the strain values for the two cases are
different for the same pressure fields although the magnitudes are similar and much less
than 1.
While the previous section compared the numerical results from the Russel’s model to that
from Biot’s model to determine the effective elastic constants for the latter model, a more
formal approach is to linearize the Russel’s model about a prestressed condition to deter-
mine the expressions for the effective elastic constants (Biot and Willis 1957). In this sec-
tion, we will restrict ourselves to the case where the prestress is isotropic and derive the
linear constitutive relations in terms of the incremental stress and strain.
Before proceeding with the detailed derivation, it is useful to study the graphical repre-
sentation of the procedure detailed in this section. Figure 5 presents the stress versus strain
plot for a material with a nonlinear constitutive relation, 𝜎 ∼ G𝜖 3∕2. If we were to linearize
this equation at some finite strain, 𝜖o, then the relation between the incremental stress and
incremental strain about the base solution, is given by the straight line that is tangent to the
curve at the point (𝜖o , 𝜎o) with the slope being the modulus of the linearized constitutive
relation. Of course, the effective modulus itself is now dependent on the base state, as it
should be. But for small strains, the linearized equation may well approximate the behav-
ior of the material. This aspect of linearization should be kept in mind while applying the
results of the derivation to the Biot’s model.
The total strain in the system can be written as a sum of initial isotropic strain and the
incremental/perturbed strain due to deformation in the material beyond the prestressed
state,
(19)
′
𝝐 = −𝜖o (e1 e1 + e2 e2 + e3 e3 ) + 𝝐 , 𝜖o > 0
where, 𝜖o is the initial isotropic compressional strain and 𝝐 is the perturbed strain due to
′
the incremental deformation. All quantities in subsequent discussion with (.)o and (.) rep-
�
resent variables in the prestress and the incremental state, respectively. The above strain
expression can be understood with the help of Fig. 6. The first term on the RHS of (19)
13
Mechanics of Saturated Colloidal Packings: A Comparison of…
(a) (b)
(c)
Fig. 6 a Initial unstressed state. b Pre-stressed state. c Incremental deformation about the prestressed state
represents the deformation in the system going from the unstressed state (Fig. 6a) to pre-
stressed state (Fig. 6b) while the second term denotes the deformation from the prestressed
state (Fig. 6b) to the incremental deformation state. Substitution of above expression for
strain in the Russel’s model (8) and retaining only the linear terms in incremental strain
gives,
3 ′
|n.𝝐.n|3∕2 ≈ − 𝜖o3∕2 + 𝜖o1∕2 (n.𝝐 .n) where, n = n1 e1 + n2 e2 + n3 e3 .
2
Substituting the above expression in (8) and integrating (see “Appendix C”) gives the
stress–strain relation in the prestressed state,
( )
4 o 3∕2
(20)
o o
𝝈 = −P 𝜹 − G 𝜖o e1 e1 + e2 e2 + e3 e3
3
and the corresponding relation between the incremental stress and the incremental strain,
( )
� � 6E � 1 � 1 � 4 �
𝜎11 = −P + o 𝜖11 + 𝜖22 + 𝜖33 − G 𝜖o3∕2
5 3 3 3
( )
� � 6Eo 1 � � 1 � 4 �
𝜎22 = −P + 𝜖11 + 𝜖22 + 𝜖33 − G 𝜖o3∕2
5 3 3 3
( )
� � 6E 1 � 1 � � 4 �
𝜎33 = −P + o 𝜖11 + 𝜖22 + 𝜖33 − G 𝜖o3∕2
5 3 3 3 (21)
� 4 �
𝜎12 = Eo 𝜖12
5
� 4 �
𝜎23 = Eo 𝜖23
5
� 4 �
𝜎31 = Eo 𝜖31
5
o 1∕2 o � �
𝜙o NG
Here, Eo = G 𝜖o , G = 2𝜋(1−𝜈p )
, and G = 𝜙 NG
2𝜋(1−𝜈p )
. Further, 𝜙o is the volume fraction in
the prestressed state while 𝜙 is the change in the volume fraction on reaching the incre-
′
mental state.
To obtain a self consistent set of equations, the pressure in the system needs to be related
to the mechanical deformation of particles and the variation in the fluid content. We start
with the particle volume balance wherein we follow a unit volume of particle packing as it
13
A. Badar, M. S. Tirumkudulu
deforms. Since the particles in the unit volume before and after the incremental deforma-
tion remain the same, the volume fraction of the particles in the two states can be related,
( )
� � �
𝜙 (1 − 𝜖o + 𝜖33 )(1 − 𝜖o + 𝜖11 )(1 − 𝜖o + 𝜖22 ) − 𝜙o (1 − 𝜖o )3 = 0. (22)
Rearranging
′
the above expression and retaining only the linear terms in incremental varia-
bles ( 𝜙𝜙o << 1) gives,
( o ) ( � )
� 𝜙 𝜙
𝜖kk = (1 − 𝜖o ) − 1 ≈ (1 − 𝜖o ) − o . (23)
𝜙 𝜙
Substituting (23) in (24) relates the change in the liquid content to the trace of the incre-
mental strain,
�
𝜁 = (1 − 𝜖o )2 𝜖kk . (25)
′ o
Similarly, G can be written in terms of G from (23) as
�
� o 𝜖kk
G = −G (26)
1 − 𝜖o
On substituting the above relation in stress–strain equation (21) leads to a full set of lin-
earized constitutive equation,
(27)
′ ′ ′ � ′
𝝈 = 𝜆tr(𝝐 p )I + 2𝜇𝝐 p − 𝛼P I and 𝜁 = 𝛽tr(𝝐 p )
On comparing the coefficients with those of Biot’s model (2–3), we have M = ∞, which
confirms the assumption of incompressibility in Russel’s model. However, the linearization
procedure introduces an extra coefficient, 𝛽 = (1 − 𝜖o )2, due to the finite stress in the pre-
stressed state. The latter asymptotes to 1 in the limit of negligible strain in the prestressed
state making it identical to Biot’s model for the fully saturated packing.
One of the most important consequences of the linearization process is that the elastic
constants of the linearized constitutive relation can be expressed in terms ( of the
) micro-
structural properties of the packing, 𝜆 = 5 Eo 6(1−𝜖 ) , 2𝜇 = 5 Eo and 𝜈 = 4 2𝜖 +3 . These
1 3+7𝜖o
4 3+7𝜖o 4
o o
13
Mechanics of Saturated Colloidal Packings: A Comparison of…
Thus, it is now possible to utilize these in the Biot’s model where the elastic constants con-
tain information on the nature of packing.
The presence of the base state strain term in these expressions requires prior knowl-
edge of the magnitude of the strain. The value of the strain may be estimated from the
prestressed state. For example, for the one-dimensional consolidation problem dis-
cussed in the previous section, the stress is related to the pressure and the strain in the
prestressed state. Since the pressure at the top of the column is zero, the base strain at
the top surface can be obtained from the knowledge of the normal stress exerted on the
surface of column (20),
( )2∕3
po
𝜖o ∼ .
̄o
G
The strain at the bottom is zero. For the problem solved in the previous section for silica
particles, the strain is estimated to be 𝜖o ∼ 10−3. The corresponding values of the effective
constants for the packing are listed in Table 2, whose numerical values are in the range of
those obtained in the previous section.
In comparison to the above calculated values, the upper (+) and lower (−)
Hashin–Shtrikman bounds for the shear modulus values may be obtained from,
± f2
𝜇HS = 𝜇1 +
(𝜇2 − 𝜇1 )−1 +
2f1 (K1 +2𝜇1 ) (28)
5𝜇1 (K1 + 34 𝜇1 )
and yields values of 0 and 15 GPa, respectively. The shear and bulk modulus of silica are
32 GPa and 36.8 GPa, respectively. For water, bulk modulus is taken as 2.1GPa. Here,
K1 = is the bulk modulus, 𝜇1,2 is the shear modulus of individual phases and f1,2 is the vol-
ume fraction of individual phases. The upper bound is computed when the stiffest material
is termed 1 and the lower bound is computed when the softest material is termed 1. Simi-
larly, the Voigt (subscript V) and Reuss (subscript R) bounds may be obtained from,
13
A. Badar, M. S. Tirumkudulu
N
∑
𝜇V = fi 𝜇i
i=1
N
(29)
1 ∑ fi
=
𝜇R 𝜇
i=1 i
which give 0 and 20 GPa, respectively. Clearly, the values obtained from the linearization
lie between both bonds but closer to the lower bond.
The previous section focused on linearizing the Russel’s constitutive relation and obtain-
ing the equivalent elastic constants for the Biot model. However, there are many problems
where the dynamics of the poroelastic media becomes important. For example, Biot’s
model has been used extensively to study consolidation of soil (and discussed in the previ-
ous sections), filtration of solvent from polymers (Doi 2009), fractures in permeable rocks
(Rice 1968; Rudnicki 1985), hydraulic fractures in oil industry (Detournay 2004, 2016),
cracking in drying colloidal/polymer films (Goehring et al. 2015) and buckling of colloidal
drops (Doi 2009). In all the aforementioned cases, the mass and the momentum balances
for the two phases are solved using the Biot’s model as the constitutive equation. In this
section, we shall derive the linearized governing equations for the Russel’s model for the
generalized three-dimensional deformation, and compare the derived equations with those
of Biot’s model.
The mass balance for the particle and liquid phases are given by (see “Appendix D” for
derivation),
𝜕𝜌p 𝜙
= −𝛁 ⋅ (𝜌p 𝜙V p ) (30)
𝜕t
and,
𝜕𝜌f (1 − 𝜙) ( )
= −𝛁 ⋅ 𝜌f (1 − 𝜙)V f . (31)
𝜕t
p∕f
Here, V p∕f = 𝜕u𝜕t is the velocity of the particle(p)/fluid(f) phase, up∕f is the displacement
of particle/fluid phase and 𝜌p∕f is the density of the particle/fluid phase. The above equa-
tions are valid even when either or both phases are compressible. The total mass balance
for both particle and fluid phase together is obtained by adding the two equations,
𝜕 ( )
(𝜌p 𝜙 + 𝜌f (1 − 𝜙)) = −𝛁 ⋅ 𝜌p 𝜙V p + 𝜌f (1 − 𝜙)V f . (32)
𝜕t
To derive the momentum balance for the system, it is important to consider the inertial
effects of solid and fluid phases as these effects are prominent in fracture problems pertain-
ing to fracture in rocks and in colloidal/polymer films. The governing equations for dynam-
ics of wave propagation in saturated colloidal systems are presented here. The convective
acceleration term is neglected since the material is stationary in the prestressed state so that
only terms linear in incremental variables are considered. Further, body forces are typically
much smaller than those from stresses.
13
Mechanics of Saturated Colloidal Packings: A Comparison of…
The particle phase momentum balance (see “Appendix E”) is given by,
( ) 𝜂 𝜕
𝛁 ⋅ 𝜙𝝈 p − (1 − 𝜙)2 (V p − V f ) = (𝜌p 𝜙V p ), (33)
𝜅 𝜕t
and the corresponding balance equation for the fluid phase is given by (see “Appendix E”),
( ) 𝜂 𝜕( )
𝛁 ⋅ (1 − 𝜙)𝝈 f + (1 − 𝜙)2 (V p − V f ) = 𝜌 (1 − 𝜙)V f . (34)
𝜅p 𝜕t f
The first term on the LHS in the particle and the fluid momentum equations corresponds
to the diffusive transport of momentum while the second term is the drag force acting on
the particle phase due to fluid flow. The latter can be obtained by considering net pressure
of fluid P acting on the faces of particle phase. The Darcy’s law for fluid phase is used to
replace the pressure gradient with fluid relative velocity,
𝜅p
(1 − 𝜙)(V f − V p ) = − 𝛁P (35)
𝜂
Finally, the term on RHS of (34) relates to the inertia of the particle/fluid phase. The addi-
tion of particle and fluid momentum balances gives the total momentum balance in the
system,
𝜕( )
𝛁⋅𝝈 = 𝜌p 𝜙V p + 𝜌f (1 − 𝜙)V f . (36)
𝜕t
We are now in a position to derive the mass and momentum balance equations for incre-
mental deformation about the prestressed state, which has constant density (incompress-
ible) and zero initial velocities for both phases. The incompressible nature of the individual
phases does not prohibit compressibility of the bulk packing since variation in the macro-
scopic mass density is enabled by changes in pore volume. Only when the two phases have
identical density and are incompressible will the bulk packing also be incompressible. The
particle phase mass balance in terms of the incremental volume fraction is given by,
�
𝜕𝜙
(37)
′
= −𝜙o 𝛁 ⋅ V p
𝜕t
while that for the fluid phase is,
�
𝜕𝜙
(38)
�
= (1 − 𝜙o )𝛁 ⋅ V f .
𝜕t
The total mass balance becomes,
�
𝜕𝜙 ( �)
(39)
�
(𝜌p − 𝜌f ) = −𝛁 ⋅ 𝜌p 𝜙o V p + 𝜌f (1 − 𝜙o )V f
𝜕t
Similarly, the particle phase momentum balance equation is derived for the incremental
state as,
�
� ( � o) 𝜂 � � 𝜕V p
𝜙o 𝛁 ⋅ 𝝈 p + 𝛁 ⋅ 𝜙 𝝈 p − (1 − 𝜙o )2 (V p − V f ) = 𝜌p 𝜙o , (40)
𝜅p 𝜕t
13
A. Badar, M. S. Tirumkudulu
Finally, the total momentum balance in terms of the incremental variables is given by,
( )
′ 𝜕 o p� f�
(42)
o
𝛁⋅𝝈 = 𝜌 𝜙 V + 𝜌f (1 − 𝜙 )V
𝜕t p
For the fluid momentum balance (41), we substitute 𝝈 f = −P 𝜹 and 𝝈 f = −Po 𝜹 to obtain
� � o
the balance equation in terms of the pressure, volume fraction and displacement,
�
𝜕 2 uf 1 𝜕 �
(44)
� � �
−(1 − 𝜙o )𝛁P + Po 𝛁𝜙 = 𝜌f (1 − 𝜙o ) + (1 − 𝜙o )2 (uf − up )
𝜕t2 𝜅 𝜕t
𝜅 p� f�
where, 𝜅 = 𝜂p . The above field Eqs. (43) and (44) have eight unknowns as ui=1..3, ui=1..3, P
′
and 𝜙 . The above six equations along with equations for particle balance (37) and total
′
mass balance (39) are the full set of governing equations that describe the dynamics of sat-
urated poroelastic material in the incremental state.
One class of problems where the aforementioned theory is applicable relates to frac-
ture in fully saturated porous media, which is of importance in the field of geophysics
and, oil and gas extraction. Examples include initiation of earthquake from fault zones,
recovery of groundwater in the fractured rocks and the recent technique for extracting
natural gas from porous rocks using hydraulic fracturing or fracking (Alexander 2015;
Rudnicki 1985; Rice and Michael 1976; Detournay et al. 1989; Madsen 1978; Cleary
1978; Detournary and Alexander 1991; Verruijt 1969). The latter involves injecting a
fracturing fluid into a well-bore to initiate and propagate a fracture in the direction per-
pendicular to the in situ minimum compressive stress (Detournary and Alexander 1991;
Alexander 2015). A corresponding problem in colloids is the propagation of cracks in
drying films of colloidal dispersions (Dufresne et al. 2006; Sengupta and Tirumkudulu
2016). In all these problems, the mechanical and pressure driven fractures occur inside
the porous media and the fracture propagation speed is often less than the body wave
speed. Consequently, the inertial forces can be neglected (Atkinson and Craster 1991),
due to which the above dynamic field equation for saturated colloidal system reduces to,
p� p� �
𝜇∇2 ui + (𝜆 + 𝜇)𝛁𝜖kk − 𝛼𝛁P = 0 (45)
p�
𝜅p
𝜕𝜖kk
(46)
�
= M ̄ ∇2 𝜖 p
𝜕t 𝜂 c kk
( )
(𝜆+2𝜇)(1−𝜖o ) 𝜙o 𝜅
where, M
̄c =
𝛼
+ Po 1−𝜙 o
. As before, 𝛼 = 1, and 𝜂p M
̄ c is the effective diffusion
coefficient of consolidation.
13
Mechanics of Saturated Colloidal Packings: A Comparison of…
Biot (1962b) derived the equivalent mass and momentum balance equation that describe
wave propagation in porous elastic media saturated with fluid,
𝜕2
𝜇∇2 up + (𝜇 + 𝜆c )𝛁e − 𝛼M𝛁𝜁 = (𝜌up + 𝜌f w) (47)
𝜕t2
𝜕2 𝜂 𝜕w
𝛁(𝛼M e − M𝜁 ) = (𝜌 up + mw) + (48)
𝜕t2 f 𝜅p 𝜕t
𝜕2
𝜇∇2 up + (𝜇 + 𝜆)𝛁e − 𝛼𝛁P = (𝜌 𝜙up + 𝜌f (1 − 𝜙)uf ). (49)
𝜕t2 p
𝜌
Similarly, we substitute the expressions for w, 𝜁 and m = 1−𝜙f
in the fluid momentum bal-
ance (48) and then multiply both sides of the equation with (1 − 𝜙) to obtain,
𝜕 2 uf 1 𝜕
−(1 − 𝜙)𝛁P = 𝜌f (1 − 𝜙) + (1 − 𝜙)2 (uf − up ) (50)
𝜕t2 𝜅 𝜕t
Equations (49) and (50) are the equations for the dynamics of poroelastic material derived
using Biot’s model and may be compared with the corresponding Eqs. (43, 44) obtained
by linearizing Russel’s model. The two sets of equations are identical except for the Po 𝛁𝜙′
term in (44), which arises due to a finite pressure in the prestressed state.
For the case of negligible inertia, the governing equations of Biot’s model reduce to,
𝜕e 𝜅p
= M c ∇2 e (52)
𝜕t 𝜂
𝜅
where e = 𝛁 ⋅ up is dilatation in the particle phase, Mc = 𝜆+2𝜇
𝛼
, here 𝛼 = 1, and 𝜂p Mc is the
diffusion coefficient of consolidation. Again, the above equation set is equivalent to that
obtained from linearizing Russel’s model (45, 46). Note that the term containing the pre-
̄ c.
stressed pressure is absorbed in the definition of M
In this subsection, we solve the basic problem of wave propagation in saturated colloidal
packing using the above derived linearized incremental state momentum and mass bal-
ances and compared them with those of Biot’s. The goal here again is to show the equiva-
lence of the two linear model, especially when inertia of the medium is also considered.
Consider a layer of saturated colloidal packing of thickness L, which is supported at the
bottom by a rigid and impermeable wall, and is bound by a zero-stress permeable surface at
the top (see Fig. 7). The bottom wall is excited by a sinusoidal vertical displacement and the
13
A. Badar, M. S. Tirumkudulu
interest is to understand how the pressure and strains propagate in time and space. One such
example at large length scales is that of earthquakes (Alexander 2015).
We derive an analytical solution for a one-dimensional plane wave in a homogeneous
poroelastic medium. In this case, only one component of displacement (u2) for the two
phases exists (Fig 7) with x2 being the direction of wave propagation. In such a case, the
p� p�
displacement and the pressure field can be expressed as u2 = u2 (x2 , t) and P = P (x2 , t).
� �
where 𝜔 is the frequency, ℜ is the real part of the complex variable and uo (̃x2 ) is deter-
mined from boundary condition of the problem. Substituting (54) in (53) and dividing the
governing equation throughout by ei𝜔̃t gives a simple second-order equation in uo,
d2 uo
− Buo = 0
A (55)
dx2 2
[( ) ]
( )2
where the coefficients are A = Ē d L𝜏 and B = 1 i𝜔𝜏
.
�
𝜌f − 𝜌 + (𝜌
1−𝜖o f
− 𝜌 ) 𝜔2 + 𝜅(1−𝜖o )
The possible solution for the above differential equation is
13
Mechanics of Saturated Colloidal Packings: A Comparison of…
− x2 direction, respectively. The constants C and D will be determined from the boundary
condition of oscillating bottom wall and zero-stress permeable surface at the top of the sat-
urated packing,
p�
ũ 2 (̃x2 , ̃t) = ℜ{ei𝜔̃t } at x̃ 2 = 0, and
(57)
�
𝜎̃ 22 = 0 at x̃ 2 = 1,
�
P̃ = 0 at x̃ 2 = 1.
Solving the problem with the aforementioned boundary conditions gives the displacement
field,
[ ( )]
� Ed L G sinh(n(̃x2 − 1)) i𝜔̃t
𝜎̃ 22 (̃x2 , ̃t) = Uo
n+ 2 +H n e (60)
L 𝜏 n cosh(n)
[ ] ( )( )( )2
�
Po 𝜙o
where, G = ( 1−𝜖
𝜌 i𝜔𝜏
− 𝜌f )𝜔2 − 𝜅(1−𝜖 and H = 1−𝜙o
𝜏
L2
. The above analysis
o o) 1−𝜖o
( )2
using the Biot’s model gives identical results with coefficients A = Ed L𝜏 ,
[ ] [ ]
B = 2𝜌f − (𝜌 + 𝜌 )𝜔 + 𝜅 , G = (𝜌 − 𝜌f )𝜔 − 𝜅 and H = 0.
i𝜔𝜏 i𝜔𝜏
� 2 � 2
For a colloidal packing composed of silica particles and saturated with water, the pre-
dictions from Russel’s linearized model are plotted in Figs. 8 and 9. Here, Uo = 1 cm/s,
L = 10 m, Po = 106 Pa, 𝜏 = 107.4 s, while the properties of the packing are the same as
before (Sect. 2.1). This gives: A = 1.0754 × 1011 and B = −7.1182 × 107 − i 2.1507 × 1013
for the Russel’s linearized model while the corresponding values for the Biot model are
nearly identical: A = 1.0753 × 1011 and B = −7.1182 × 107 − i 2.1505 × 1013. Note that
the extra term Po 𝛁𝜙′ in (44) for the Russel’s model, which arises due to the existence of
a finite pressure in the prestressed state, contributes negligibly to the values of A and B in
this particular example. However, this term could become significant if the pressure in the
prestressed state takes values in the range of drained effective elastic coefficients.
13
A. Badar, M. S. Tirumkudulu
depths. The magnitude of the pressure amplitude at the bottom layer of saturated packing is
large due to the oscillating wall but decays towards to the top.
5 Conclusion
We present the nonlinear constitutive relation for deformation of a saturated colloidal pack-
ing proposed by Russel et al. (2008) and show that for small strains, the equations may be
linearized to obtain a form that is identical to the linear constitutive relation proposed by
Biot (1941). The elastic coefficients obtained from such a linearization are related to the
micro-structural details of the packing such as the particle modulus, the packing concentra-
tion and the nature of packing, thereby enabling a more effective utilization of Biot’s model
for problems in the linear limit.
Table 2 presents the effective elastic constants for the Biot’s model obtained from the
linearization of Russel’s model for fixed prestress strain in three-dimensions. Apart from
the particle properties, the calculations require the knowledge of the strain in the prestress
state to determine the elastic coefficients for the packing. For 𝜖o ≪ 1, 𝜈 ≈ 1∕4 and
1∕2
𝜙o NG𝜖
𝜆 ≈ 𝜇 = 5𝜋(1−𝜈o ) .
p
Two examples are presented where the predictions of Biot’s model are compared to
those of the linearized model of Russel’s model for two one-dimensional model problems:
fluid outflow driven by an applied mechanical load and wave propagation in a saturated
colloidal packing. While the emphasis in this study has been on packings of colloidal par-
ticles, the derivation of the constitutive relation (8) neglects the dispersion and interfacial
forces implying that the expressions for the effective elastic coefficients presented here are
equally applicable to packings of larger particles.
Acknowledgements A.B. thanks IIT Bombay for financial support. The project was funded in part by
Department of Science and Technology, India.
13
Mechanics of Saturated Colloidal Packings: A Comparison of…
Appendix
A: List of Variables
This section lists all variables and their definitions used in the main text.
Variables Definitions
𝛼 Biot–Willis coefficient
𝛥𝝈 Incremental stress of bulk material
𝛥𝝈 f Incremental stress in the fluid phase
𝝐p = 𝝐 Strain in the solid/particle network/macroscopic strain
𝜖o Initial isotropic compressive strain
𝝐
′
Perturbed strain
𝜂 Viscosity of liquid
𝜸̇ Shear rate
𝛾pl Surface energy of the particle–liquid interface
𝜅p Permeability of the packing
𝜹 Second-order identity tensor
𝜆 Lame’s first constant
𝜇 Shear modulus
𝜇HS±
Hashin–Shtrikman’s upper and lower bounds for effective shear modulus
𝜇V Voigt bound for effective shear modulus bounds
𝜇R Reuss bound for effective shear modulus bounds
𝜇i Shear modulus of the ith phase
𝜈p Poisson’s ratio of particle
𝜈 Poisson’s ratio of packing
𝜔 Angular frequency
𝜙 Solid volume fraction/particle phase volume fraction
𝜙o Volume fraction in the prestress state
𝜙
′
Perturbed/incremental volume fraction
𝜌p Mass density of particle phase
𝜌f Mass density of fluid phase
𝜌 Bulk density
𝝈 Total effective stress/macroscopic stress
𝝈p Particle network stress
𝝈f Fluid stress
𝝈o Stress in the prestress/base state
𝝈
′
Perturbed/incremental stress
𝝉d Shear stress in fluid phase
𝝉v Viscoelastic stress in the particle
𝜏 Characteristic time scale in soil consolidation and one-dimensional wave problems
𝜏r Relaxation time in Russel’s model
𝜁 Change in fluid or liquid content
A = 𝜆, obtained by linearization in the vicinity of prestressed state by Biot
a Particle radius
13
A. Badar, M. S. Tirumkudulu
Variables Definitions
aR Inverse of longitudinal wave modulus from Russel’s model
aB Inverse of longitudinal wave modulus from Biot’s model
F Force acting on the particle pair
fi Volume fraction of ith phase
G Particle shear modulus
̄
G Effective elastic modulus
H Initial thickness
h Height of the soil consolidation column
I Identity tensor
KR Diffusion coefficient of consolidation from Russel’s model
KB Diffusion coefficient of consolidation from Biot’s model
K1 Bulk modulus of material termed 1
L Thickness of saturated colloidal system in 1D wave problem
Lc Characteristic length scale for macroscopic dimensions
M Biot modulus
m Effective mass in Biot’s model
N Number of contacting neighbor
Nb = 𝜇, obtained by linearization in the vicinity of prestressed state by Biot
n Unit normal vector on the surface of the particle
P Fluid pressure in the pores/pressure in the solvent phase packing
Po Fluid pressure in the prestress state
P
′
Fluid pressure in the perturbed/incremental state
po Normal load in soil consolidation problem
′
Q, Q , R Elastic constants obtained by linearization in the vicinity of prestressed state by Biot
S Surface dipole
Uo Sinusoidal vertical displacement in 1D wave problem
up Displacement field in the solid/particle network
uf Displacement field in the fluid
V Representative volume in Russel’s model
V Fluid velocity
Vp Velocity of particle phase
Vf Velocity of fluid phase
B: Soil Consolidation
1 1
|n ⋅ 𝝐 ⋅ n| 2 n ⋅ 𝝐 ⋅ n = |𝜖33 | 2 𝜖33 n33 ; where,
n = cos(𝜙)sin(𝜃)e1 + sin(𝜙)sin(𝜃)e2 + cos(𝜃)e3
[ 𝜋∕2 2𝜋
∫ ∫𝜃=0 ∫𝜙=0
1 1
∴ |𝜖33 2
|𝜖33 n3z nnd2 n = |𝜖33
2
|𝜖33 (cos 𝜃)3 nn sin 𝜃d𝜃d𝜙
(61)
𝜋 2𝜋 ]
∫𝜃=𝜋∕2 ∫𝜙=0
− (cos 𝜃)3 nn sin 𝜃d𝜃d𝜙
𝜋 1
= |𝜖33 | 2 𝜖33 (e1 e1 + e2 e2 + 4e3 e3 )
6
13
Mechanics of Saturated Colloidal Packings: A Comparison of…
The radial vector on the surface of the particle is written in Cartesian coordinates,
n = cos(𝜙)sin(𝜃)e1 + sin(𝜙)sin(𝜃)e2 + cos(𝜃)e3
∫ ∫𝜃=0 ∫𝜙=0
I1 = −𝜖o3∕2 nnd2 n = −𝜖o3∕2 nn sin 𝜃d𝜃d𝜙
3∕2
−4𝜖o
= 𝜋(e1 e1 + e2 e2 + e3 e3 )
3
𝜋 2𝜋
∫ ∫𝜃=0 ∫𝜙=0
′ ′
I2 = (𝜖o1∕2 (n.𝝐 .n))nnd2 n = 𝜖o1∕2 (n.𝝐 .n)nn sin 𝜃d𝜃d𝜙
𝜋 2𝜋
∫𝜃=0 ∫𝜙=0
′
= 𝜖o1∕2 𝝐 ∶ nn nn sin 𝜃d𝜃d𝜙
(
′ 4𝜋 4𝜋
= 𝜖o1∕2 𝝐 ∶ e e e e + e e e e
15 2 2 3 3 15 2 3 2 3
4𝜋 4𝜋 4𝜋
+ e e e e + e e e e + e e e e
15 2 3 3 2 15 3 2 2 3 15 3 2 3 2
4𝜋 4𝜋 4𝜋 4𝜋
+ e e e e + e e e e + e e e e + e e e e
15 3 3 2 2 5 3 3 3 3 15 1 1 3 3 15 1 3 1 3
4𝜋 4𝜋 4𝜋 4𝜋 4𝜋
+ e e e e + e e e e + e e e e + e e e e + e e e e
15 1 3 3 1 15 3 1 1 3 15 3 1 3 1 15 3 3 1 1 15 1 2 2 1
4𝜋 4𝜋 4𝜋 4𝜋 4𝜋
+ e e e e + e e e e + e e e e + e e e e + e e e e
15 2 1 1 2 15 2 1 2 1 15 2 2 1 1 5 2 2 2 2 5 1 1 1 1
)
4𝜋 4𝜋
+ e1 e1 e2 e2 + e e e e
15 15 1 2 1 2
[
4𝜋 � 1 � 1 �
= 𝜖o1∕2 (𝜖 + 𝜖 + 𝜖 )e e
5 11 3 22 3 33 1 1
4𝜋 � 1 � 1 � 4𝜋 � 1 � 1 �
+ (𝜖22 + 𝜖11 + 𝜖33 )e2 e2 + (𝜖33 + 𝜖11 + 𝜖22 )e3 e3
5 3 3 5 3 3
4𝜋 � �
+ (𝜖 + 𝜖21 )(e1 e2 + e2 e1 )
15 12
]
4𝜋 � � 4𝜋 � �
+ (𝜖13 + 𝜖31 )(e1 e3 + e3 e1 ) + (𝜖23 + 𝜖32 )(e2 e3 + e3 e2 )
15 15
D: Mass Balance
The mass balance equations for the particle and the fluid phase are given by (see Fig.10 for
shell balance):
13
A. Badar, M. S. Tirumkudulu
1. Particle phase
(Particle mass in) − (Particle mass out)
= (Accumulation of particle mass inside the volume V) ;
( )
p p
𝜌p 𝜙V1 𝛥x2 𝛥x3 |x1 − 𝜌p 𝜙V1 𝛥x2 𝛥x3 |x1 +𝛥x1
( ) (62)
p p
+ 𝜌p 𝜙V2 𝛥x1 𝛥x3 |x2 − 𝜌p 𝜙V2 𝛥x1 𝛥x3 |x2 +𝛥x2
( )
p p
𝜕(𝜌p 𝜙)
+ 𝜌p 𝜙V3 𝛥x2 𝛥x1 |x3 − 𝜌p 𝜙V3 𝛥x2 𝛥x1 |x3 +𝛥x3 = 𝛥x1 𝛥x2 𝛥x3
𝜕t
On dividing by 𝛥x1 𝛥x2 𝛥x3 and taking the limit as 𝛥xi → 0 gives,
( p p p )
𝜕(𝜌p 𝜙) 𝜕(𝜌p 𝜙V1 ) 𝜕(𝜌p 𝜙V2 ) 𝜕(𝜌p 𝜙V3 )
=− + + (63)
𝜕t 𝜕x1 𝜕x2 𝜕x3
13
Mechanics of Saturated Colloidal Packings: A Comparison of…
2. Fluid phase
( )
Fluid mass in − (Fluid mass out) =
( )
Accumulation of fluid mass inside the volume V
( )
f f
𝜌f (1 − 𝜙)V1 𝛥x2 𝛥x3 |x1 − 𝜌f (1 − 𝜙)V1 𝛥x2 𝛥x3 |x1 +𝛥x1
( )
f f
+ 𝜌f (1 − 𝜙)V2 𝛥x1 𝛥x3 |x2 − 𝜌f (1 − 𝜙)V2 𝛥x1 𝛥x3 |x2 +𝛥x2
( )
f f
+ 𝜌f (1 − 𝜙)V3 𝛥x2 𝛥x1 |x3 − 𝜌f (1 − 𝜙)V3 𝛥x2 𝛥x1 |x3 +𝛥x3
𝜕(𝜌f (1 − 𝜙))
= 𝛥x1 𝛥x2 𝛥x3 ;
𝜕t
On dividing by 𝛥x1 𝛥x2 𝛥x3 and taking the limit as 𝛥xi → 0 gives,
𝜕(𝜌f (1 − 𝜙))
=
𝜕t
( f f f ) (64)
𝜕(𝜌f (1 − 𝜙)V1 ) 𝜕(𝜌f (1 − 𝜙)V2 ) 𝜕(𝜌f (1 − 𝜙)V3 )
− + +
𝜕x1 𝜕x2 𝜕x3
E: Momentum Balance
Next, the shell balance for the momentum is considered where both body force and con-
vective acceleration are ignored (Fig. 11),
13
A. Badar, M. S. Tirumkudulu
f p
𝜅p 𝜕(1 − 𝜙)P 1
(1 − 𝜙)(Vi − Vi ) = −
𝜂 𝜕xi 1−𝜙
p p p
𝜕𝜙𝜎13 𝜕𝜙𝜎23 𝜕𝜙𝜎33 𝜂 f p 𝜕 p
+ + + (1 − 𝜙)2 (V3 − V3 ) = (𝜙𝜌p V3 ) (67)
𝜕x1 𝜕x2 𝜕x3 𝜅p 𝜕t
13
Mechanics of Saturated Colloidal Packings: A Comparison of…
f f f
𝜕(1 − 𝜙)𝜎13 𝜕(1 − 𝜙)𝜎23 𝜕(1 − 𝜙)𝜎33 𝜂 f p
+ + − (1 − 𝜙)2 (V3 − V3 )
𝜕x1 𝜕x2 𝜕x3 𝜅p (70)
𝜕 f
= ((1 − 𝜙)𝜌f V3 )
𝜕t
13
A. Badar, M. S. Tirumkudulu
Substituting the above relation in the total mass balance (39) and particle balance (37), and
combining the two gives,
( �)
1 𝜕𝜁 𝜕w
− =𝛁⋅ (74)
(1 − 𝜖o )3 𝜕t 𝜕t
On integrating the above equation with respect to time relates the variation of the fluid con-
tent to the dilatation term for the fluid phase and volume of injected fluid (Q),
1 �
− 3
𝜁 =𝛁⋅w +Q (75)
(1 − 𝜖o )
In our analysis Q = 0 since no external source supplies fluid to the packing. Next, we sub-
stitute 𝜁 in terms of the dilatation term for the particle phase (from (25)) into the above
equation,
1 � �
𝛁 ⋅ up = −𝛁 ⋅ w (76)
1 − 𝜖o
Taking divergence of equations (71) and (72) and replacing 𝛁 ⋅ w with 𝛁 ⋅ up from (76)
′ ′
and rearranging, reduces the number of independent variables from 7 to 4 in the two
equations,
� � � 𝜕 2 up
�
𝜌f 𝜕 2 up�
𝜇∇2 up + (𝜆 + 𝜇)𝛁(𝛁 ⋅ up ) − 𝛁P = 𝜌 − (77)
𝜕t2 1 − 𝜖o 𝜕t2
and,
� ( )( ) � �
1 𝜕up � Po 𝜙o p� 𝜌 𝜕 2 up
= 𝛁P + 𝛁(𝛁 ⋅ u ) + (𝜌 f − ) (78)
𝜅(1 − 𝜖o ) 𝜕t 1 − 𝜖o 1 − 𝜙o 1 − 𝜖o 𝜕t2
13
Mechanics of Saturated Colloidal Packings: A Comparison of…
The above exercise results in an arbitrary constant of integration in both equations, which
is neglected without loss of generality. The above two equations can be further combined
into one by substituting 𝛁P from (78) into (77), resulting in the final form of governing
′
References
Alexander, H.D.C.: Theory and Applications of Transport in Porous Media Poroelasticity, vol. 27. Springer,
Berlin (2015)
Atkinson, C., Craster, R.: Plane strain fracture in poroelastic media. Math. Phys. Sci. 434(1892), 605–633
(1991)
Bemer, E.: Poromechanics : From linear to nonlinear poroelasticity and poroviscoelasticity. Oil Gas Sci.
Technol. 56(6), 531–544 (2001)
Biot, M.A.: General theory of three-dimensional consolidation. J. Appl. Phys. 12(2), 155–164 (1941)
Biot, M.A.: Propagation of elastic waves in a cylindrical bore containing a fluid. J. Appl. Phys. 23(9), 997–
1005 (1952)
Biot, M.A.: Theory of elasticity and consolidtion for a porous anisotropic solid. J. Appl. Phys. 26(2), 182–
185 (1955)
Biot, M.A.: General solution of the equations of elasticity and consolidation for a porous material. J. Appl.
Mech. 78, 91–96 (1956a)
Biot, M.A.: Propagation of elastic waves in a cylindrical bore containing a fluid. J. Acoust. Soc. Am. 28(2),
168–178 (1956b)
Biot, M.A.: Generalized theory of accoustic propagation in porous dissipative media. J. Acoust. Soc. Am.
34(9), 1254–1264 (1962a)
Biot, M.A.: Mechanics of deformation and acoustic propagation in porous media. J. Appl. Phys. 33(4),
1482–1498 (1962b)
Biot, M.A.: Nonlinear and semilinear rheology of porous solids. J. Geophys. Res. 78(23), 4924–4937 (1973)
Biot, M.A., Willis, D.: Propagation of elastic waves in a cylindrical bore containing a fluid. J. Appl. Mech.
26(2), 594–601 (1957)
Bowen, R.A.Y.M.: Compressible porous media models by use of the theory of mixtures. Int. J. Eng. Sci.
20(6), 697–735 (1982)
Cleary, M.P.: Moving singularities in elasto-diffusive solids with applications to fracture propagation. Int. J.
Solids Struct. 14(2), 81–97 (1978)
Coussy, O.: Mechanics and Physics of Porous Solids, Chap 4. Wiley, Hoboken , 61–83 (2010)
Detournay, E.: Propagation regimes of fluid-driven fractures in impermeable rocks. Int. J Geomech.
4(March), 35–45 (2004)
Detournay, E.: Mechanics of hydraulic fractures. Ann. Rev. Fluid 48, 311–339 (2016)
Detournary, E., Alexander, H.D.C.: Plane strain analysis of a stationary hydraulic fracture in a poroelastic
medium. Int. J. Solids Struct. 27(13), 1645–1662 (1991)
Detournay, E., Cheng, A.D., Roegiers, J.C., McLennan, J.: Poroelasticity considerations in in situ stress
determination by hydraulic fracturing. Int. J. Rock Mech. Min. Sci. Geomech. Abstr. 26(6), 507–513
(1989)
Doi, M.: Gel dynamics. J. Phys. Soc. Jpn. 78(5), 1–15 (2009)
Dufresne, E.R., Stark, D.J., Greenblatt, N.A., Cheng, J.X., Hutchinson, J.W., Mahadevan, L., Weitz, D.A.:
Dynamics of fracture in drying suspensions. Langmuir 22(17), 7144–7147 (2006)
Goehring, L., Nakahara, A., Dutta, T., Kitsunezaki, S., Tarafdar, S.: Desiccation Cracks and their Patterns:
Formation and Modelling in Science and Nature. Wiley, Hoboken (2015)
Hashin, Z., Shtrikman, S.: A variational approach to the elastic behavior of multiphase materials. J. Mech.
Phys. Solids 11, 127–140 (1963)
13
A. Badar, M. S. Tirumkudulu
Lesaine, A., Bonamy, D., Gauthier, G., Rountree, C.L., Lazarus, V.: Highly porous layers of silica nano-
spheres sintered by drying: scaling up of the elastic properties of the beads to the macroscopic mechan-
ical properties. Soft Matter 14, 3987–3997 (2018)
MacMinn, C.W., Dufresne, E.R., Wettlaufer, J.S.: Large deformations of a soft porous material. Phys. Rev.
Appl. 5(044020), 1–30 (2016)
Madsen, O.: Wave-induced pore pressures and effective stresses in a porous bed. Geotechnique 28(044020),
377–393 (1978)
Mavko, G., Mukerji, T., Dvorkin, J.: The Rock Physics Handbook vol. 27. Cambridge University Press,
Cambridge (2009)
McCabe, W.L., Smith, J.C., Harriot, P.: Unit Operations of Chemical Engineering, 7th edn. McGraw-Hill,
New York (2005)
Reuss, A.: Berechnung der Fliessgrenzen von Mischkristallen auf Grund der Plastizitatsbedingung fur Eink-
ristalle. Z. Ang. Math. Mech. 9, 49–58 (1929)
Rice, J.R.: A path independent integral and the approximate analysis of strain concentration by nitches and
cracks. J. Appl. Mech. 35(35), 379–386 (1968)
Rice, R.J., Michael, P.C.: Some basic stress difussion solution for fluid saturated elastic porous media with
compressible constituents. Rev. Geophys. Space Phys. 14(2), 227–241 (1976)
Routh, A.F.: Drying of thin colloidal films. Rep. Prog. Phys. 76(4), 046603–046633 (2013)
Routh, A.F., Russel, W.B.: A process model for latex film formation : limiting regimes for individual driving
forces. Langmuir 9(5), 7762–7773 (1999)
Rudnicki, J.W.: Effect of Pore Fluid Diffusion on Deformation and Failure of Rock. Northwestern Univ,
Evanston (1985)
Russel, W.B., Wu, N., Man, W.: Generalized Hertzian model for the deformation and cracking of colloidal
packings saturated with liquid. Langmuir 5(24), 1721–1730 (2008)
Sarkar, A., Tirumkudulu, M.S.: Asymptotic analysis of stresses near a crack tip in a two-dimensional col-
loidal packing saturated with liquid. Phys. Rev. E 051401, 1–10 (2011)
Sengupta, R., Tirumkudulu, M.S.: Dynamics of cracking in drying colloidal sheet. Soft Matter 12, 3149–
3155 (2016)
Singh, K.B., Tirumkudulu, M.S.: Cracking in drying colloidal films. Phys. Rev. Lett. 98(218302), 1–4
(2007)
Tirumkudulu, M.S.: Buckling of a drying colloidal drop. Soft Matter 14, 7455–7461 (2018)
Tirumkudulu, M.S., Russel, W.B.: Role of capillary stresses in film formation. Langmuir 08544(4), 2947–
2961 (2004)
Tirumkudulu, M.S., Russel, W.B.: Cracking in drying latex films. Soft Matter 15(4), 4938–4948 (2005)
Verruijt, A.: Elastic Storage of Aquifers. Flow through Porous Media, pp. 331–376. Academic Press, New
York (1969)
Walton, K.: The effective elastic moduli of a random packing of spheres. J. Mech. Phys. Solids 35(2), 213–
226 (1987)
Publisher’s Note Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional affiliations.
13