Soils Modelling Using The Boltzmann Equation in The Lagrangian Description of The Continuum

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

Soils modelling using the Boltzmann equation in the

Lagrangian description of the continuum


R. Tamagnini1 and B. Switala1
1
BOKU University of Natural Resources of Vienna, Department of Civil
Engineering and Natural Hazards

Abstract
The paper presents an alternative approach to the classic critical state framework for unsaturated and saturated
soils. The aim of the formulation is to discuss the advantages of the proposed model in some engineering
applications such as the shale gas extraction and the analysis of landslides. The model formulation is based on the
theory of non-equilibrium thermodynamics. The basic formulation is extended in order to account for phase
transitions. The introduction of the term which describes the phase transition represents the jump in the entropy
value of the reference volume due its convective motion. The model is integrated using the Laplace transform and
this allows referring the variable fields to both the initial and current configuration without loss of mathematical
correctness. This happens because phase transitions occur independently to the time scale and it means that
whatever velocity field is applied to the volume, the resulting produced entropy is invariant. The current eulerian
volume preserves the original volume because the phase change occurs for a constant value of the fluid masses.
This implies that the classic concept of undrained conditions does not hold anymore. This is also the real condition
in site because, for example, during the gas reservoir exploitation the dimension of the element boundary cannot
be controlled as in laboratory and then the formulated models for the mechanical description of shear bands
forming at constant sample volume do not represent the reality. Generally this condition is the real condition, in
Physics, in pressure controlled tests. The paper presents some numerical shear tests to demonstrate that the
constitutive equations are not affected by mesh dependency.

Introduction
The thermodynamics of multiphase materials is of great relevance for geotechnical and geomechanical
engineering. Soils mechanics that is the basic theory for the study of geotechnical problems has been developed
including the fluids pressure in the definition of the effective stress [1]. The critical state framework constitutive
equations, that have been derived for the analysis of soils and rocks are effective in problems for normally or
slightly consolidated conditions [2] but they are not suitable for the study of softening problems. The analysis of
softening in porous materials requires more sophisticated technique such as, for example, the second gradient
models [3]. This happens, because softening occurs in a shear band that needs the definition of an internal length
scale for the mesh size independency.

In the present paper, the problem of softening and gas dissociation are approached with a different theory. The
proposed thermodynamic framework includes in its formulation the heat produced during phase transition
phenomena. In unsaturated soils this heat is produced by the evaporation or condensation of the water vapor. In
shale gas the heat is produced by the dissociation of the methane gas molecule.

The reference work for the modeling of unsaturated soils is the BBM, the Basic Barcelona Model [4]. The first
thermodynamics framework for unsaturated soils has been developed by Hassanizadeh and Gray [5]. This model
includes in the eulerian description of the mixture the interfaces between the water and gas phase. Another eulerian
approach for the analysis of unsaturated soils has been provided by Lewis and Schrefler [6] with the integration
scheme for FEM analyses.

These work present interesting insights for the description of multiphase materials, which however have also some
shortcomings. The present paper shows that the proper description of unsaturated soils collapse upon wetting or
the shear band formation in softening materials require necessarily a Lagrangian description, because the
phenomena of phase transition occurs at a constant value of the total amount of the saturating fluids contained in
the reference volume. Then the only possible mathematical description of the variation of porosity due to phase
change is the inclusion in the mathematical model of a sink term for the continuity equation of the gas phase. This
term can be formulated using the Dirac delta function. The resulting deformation is the volume loss due to the
outflow of the gas-water interface particles that represent the a-dimensional interfaces adsorbed on the soils grains
and at the contact between the water and gas.

The model is based on the formulation of a non-equilibrium thermodynamics model, developed by Prigogine [7],
de Groot and Mazur [8] and Kundepudi and Prigogine [9]. The extension of the model is represented by the
extension of the theorem for the mechanical equilibrium, particularly, the application of an undefined velocity to
each component of the mixture results in the presence of a new term in the dissipation equation that is a phase
transition of the species contained in the mixture. This term represents the jump in the entropy needed to balance
the entropy flux due to the flow of energy and matter. This means that the model can describe an open system as
a closed system in which the enthalpy of transformation represents the energy needed to move the current volume
from a point to another with an undefined velocity. The only way to achieve this result is the use of the Laplace
transform.

Thermodynamic model
The development of the constitutive framework is based on the work of Prigogine [7]. Non-equilibrium
thermodynamics is reported in the books of Groot and Mazur [8] and Kondepudi and Prigogine [9]. In the
following, the intensive thermodynamic variables are assumed to be function of the space coordinate vector x and
time t. The intensive variables are: the temperature T(x,t), the fluid pressure pα (x,t) of the α component of the
mixture, the total applied stress tensor σij(x,t) and the chemical potential µk(x,t). The extensive variables are
defined with their specific values and they are: the entropy s(x,t), the internal energy u(x,t) and the specific mass
ρk(x,t) of the component k. Non-equilibrium thermodynamics assumes that the system under study could be in an
overall non-equilibrium condition in which the variables may vary from a point to another but the local equilibrium
is satisfied for each elementary volume of the continuum. This implies that the second law of thermodynamics can
be written as:

dS = dSi + dSe (1)

in which dSe is the change of the entropy of the system due to the exchange of energy and matter and dSi is the
internal production of entropy. The restriction on the sign of the entropy inequality in the case of an open system
applies only to the first term of Eq. (1):

dSi ≥ 0 (2)

The second term can be positive or negative. For a closed system in which dSe is equal to zero the Clausius´
expression is recovered. The local equilibrium assumption requires that the internal production of entropy can be
decomposed into the sum of the local production of entropy of the elementary volumes:

dSi = ∑ dSik ≥ 0 with dSik ≥ 0 (3)


k

This is a stronger assumption with respect to the classic second law as postulated for a closed system. The local
entropy production is defined as:

dSi dsi (x, t )


= Φ(x, t )dV with Φ (x, t ) =
dt V∫
≥0 (4)
dt
The change in the mass of the component k can be expressed by dρk=dρke+dρki where the first term is the change
of the number of moles due to transport and the second term is the internal production due to chemical reactions.
The mass balance equation of the component k can be expressed by the equation:

∂ρ k ∂ρ ke ∂ρ ki
= + = −∇ (v k ρ k ) + Θ[ρ k ] (5)
∂t ∂t ∂t
in which vk is the velocity of the component k and Θ is the amount of the component k produced inside the element
volume by chemical reactions. The eulerian velocity of the centre of mass of the mixture (generally associated
with the motion of the solid skeleton) can be defined as:

v=
∑ρv k k k
(6)
∑ρ k k

The diffusion flow Jk of component k is defined as:

J k = ρ k (v k − v ) (7)

Introducing Eq. (7) in Eq. (5) the following expression can be obtained:

∂ρ k ∂ρ ke ∂ρ ki
= + = −∇ J k − ∇ (ρ k v ) + ∑ j υ jk v j (8)
∂t ∂t ∂t
in which the last term is the internal production of the component k and it can be expressed by the product of the
stoichiometric coefficient υjk and the velocity of reaction:

1 dξ j
vj = (9)
V dt
in which ξ is the extent of the jth reaction.

Eq. (8) can be formulated in terms of the material time derivative d/dt with respect to the motion of the centre of
mass of the mixture:

dρ k
= −∇J k − ρ k ∇ ⋅ v + ∑ j υ jk v j (10)
dt
The equation of motion of the system can be written:

dv
∑ k
ρk
dt
= −∇σ ij + ∑k ρ k Fk (11)

and the force Fk due to an external field ψ k:


Fk = ∇ψ k (12)

have been introduced. The balance of the kinetic energy can be obtained multiplying both terms of Eq. (11) by
the velocity of the centre of mass of the mixture:

1 
d  v2 
ρ 
2 
= −∇(σ ⋅ v ) + σ : ∇v + ∑k ρ k Fk ⋅ v (14)
dt
Considering the relation [8]:
da ∂ρa
ρ = + ∇ (aρv ) (15)
dt ∂t
in which a is a general quantity, Eq. (14) can be rewritten as:

1 
∂ ρv 2 
2  = −∇ 1 ρv 2 v + σ ⋅ v  + σ : ∇v +
∂t

2


∑k ρk Fk v (16)

The rate of the kinetic energy and potential energy can be defined as [8]:

1 
∂ρ  v 2 +ψ 
2  = −∇  ρ  1 v 2 + ψ  v + σ ⋅ v + 
∂t
   ∑ ψ k J k  + σ : ∇v − ∑k J k Fk (17)
 2  k

In an open system the conservation of total energy can be expressed by the equation:

∂e
+ ∇J e = 0 (18)
∂t
in which e is the total specific energy and Je is the energy flux. The total energy e is the sum of the kinetic energy,
the internal energy (the energy that is not associated with the bulk motion) and the potential energy:

1 1 1
e=
2
∑ k
ρ k v 2k + ρu + ρψ = ρv 2 + ∑k ρ k J 2k + ρu + ρψ (19)
2 2
In the following, the kinetic energy of diffusion associated with the vector Jk will be disregarded. The energy flux
can be written as:

J e = J q + ∑k ρ k ek v k + ∑k ψ k J k + σ ⋅ v (20)

In which Jq is the heat flow. Introducing Eq. (20) in (18) and by the use of Eq. (19), the expression for the rate of
the specific internal energy is obtained (using the soils mechanics convention for the sign of the strain and stress):

∂ ( ρu )
= −∇ (J q + ρuv ) + σ : ∇v + ∑ k Fk J k (21)
∂t
Eq. (21) can be reformulated using Eq. (15) :

du
ρ = −∇J q + σ : ∇v + ∑ k Fk J k (22)
dt

The total stress tensor σij can be split into an isotropic component pδij and a deviator component Π:

σ ij = pδ ij + Π ij (23)

And the relation (22) can be rewritten:

du
ρ = −∇ J q + pdv + Π : ∇v + ∑ k J k Fk (24)
dt

in which v is the specific volume. The specific entropy production in the reference volume can be expressed as:
∂ρs
= −∇J s ,tot +Φ (25)
∂t
in which Js,tot is the total entropy flux due to exchange of matter and heat flow. By the use of Eq. (15), Eq. (25)
results:

ρ
ds
dt
( )
= −∇ J s ,tot − ρsv + Φ (26)

In which the entropy due to convection, ρsv has been subtracted from the total flux. Eq. (26) reads:

ds
ρ = −∇ J s + Φ (27)
dt

The expression of the time derivative of the specific entropy production can be obtained using the Gibbs equation
in its Eulerian form with respect to the motion of the centre of mass:

Tds = du − vpdv e − ∑k vµ k dρ k (28)

In which v is the specific volume and pd ve is the elastic reversible part of the volumetric strain work.
Considering the relation ρ = v and introducing Eq. (10) and (22), the following expression can be obtained:
-1

ds 1 1 1 1 1 1
ρ = − ∇J q + σ : dε p − ∑ k Fk J k + ∑ k µ k ∇J k + ∑ k µ k ρ k ∇ ⋅ v − ∑ j A j J j (29)
dt T T T T T T
In Eq. (29), the term Aj is the affinity, the thermodynamic force that drives the chemical reactions and it is defined
as:

A j = ∑ k µ kυ kj (30)

And εp is the plastic small strain tensor. The following relation between the Gibbs energy of the mixture and the
chemical potential holds:

g = g f + g g = ∑k µ k ρ k (31)

The second and the fifth term of Eq. (29) can be cast to define the effective stress of the mixture; the indices f and
g indicate the fluid and gas phases.

σˆ ij = ( p + g )δ ij + Π ij (32)

The Eq. (29) can be rewritten as:

ds 1 1 1 1 1
ρ = − ∇J q + σˆ : dε p − ∑ k Fk J k + ∑ k µ k ∇J k − ∑ j A j J j (33)
dt T T T T T

Note that the rate of the term in Eq. (31) for isothermal condition can be written as:

dg = ∑k ρ k dµ k + ∑k µ k dρ k = −v g dp g − vw dpw + ∑k µ k dρ k (34)
in which the Gibbs-Duhem equation has been used. Introducing the porosity φ, the degree of saturation Sr and the
gas saturation degree Sg as follows:

Vv Vw Vg
φ= Sr = Sg = = 1 − Sr (35)
V Vv Vv

the rate of Eq. (32) can be formulated as:

)
dσ ij = dσ ij − φdp g δ ij + φS r ( dp g − dp f )δ ij + ∑k µ k dρ k δ ij (36)

The first three terms of Eq. (36) are similar to the expression proposed by Bishop [10]. His equation is defined as:

dσ ij′ = dσ ij − dpgδij + χ (dpg − dpf )δij (37)

Eq. (37) also agrees with the formulation proposed by Coussy [11] in which the energetic term U can be identified
by the last term in Eq. (36). This term vanishes in stationary condition during a phase transition (contrary to the
model of Coussy). A simplified version of Eq. (37) has been proposed by Lewis and Schrefler [6] in which χ is
replaced by Sr. In the sake of simplicity this last equation will be adopted in the discussion of the numerical
algorithm. The following expressions can be used to reformulate Eq. (33):

µk µ J  µ 
∇J k = ∇ k k  − ∇ k J k (38)
T  T  T 

1 J  J
∇J q = ∇ q  − ∇T q2 (39)
T T  T

Rearranging Eq. (33):

ds  J q − µk J k  1 1 1 1 µ 
ρ = −∇  − 2 J q ∇T + σˆ : dε p + ∑k Fk J k − ∑k T∇ k J k +
dt  T  T T T T T  (40)
1
− ∑ Aj J j
T j

or in the equivalent form:

ds  J − µk J k  1 1 1  µ   1
ρ = −∇ q  − 2 J q ∇T + σˆ : dε p − ∑ k J k  T∇ k  − Fk  − ∑ j A j J j (41)
dt  T  T T T  T   T

Eq. (41) can be rearranged decomposing the contribution of the isotropic component of the stress and the deviatoric
component, for the sign convention the second component of the strain work has a different sign with respect to
the isotropic work (soils mechanics convention for the isotropic component of stress and strains, isotropic
compression is positive).

ds  J q − µk J k  1 1 1 1  µ  
ρ = −∇  − 2 J q ∇T + pdv − Π : ∇v − ∑k J k  T∇ k  − Fk  +
dt  T  T T T T  T  
1
− ∑ Aj J j
T j
(42)
The second component of the strain work can be rearranged in the following way:

1  Πv  1 1
− Π∇v = −∇  + v∇Π − 2 vΠ∇T (43)
T  T  T T

Introducing Eq. (43) in Eq. (42):

 J q − µk J k + Πv  1
 − 2 (J q + vΠ )∇T + pˆ dv − v k ρ k (∇µ k − Fk ) −
ds 1 1
ρ = −∇
dt  T  T T T
1   µk   1
+ ∑ J  T∇ 
T k   T
k  − Fk  − ∑ j A j J j

(44)
 T

in which the equilibrium and the Gibbs-Duhem equation have been used. From the comparison of Eq. (41) and
Eq. (27) the flow Js can be identified by the first term of Eq. (41):

 J − µ k J k + Πv 
J s = −∇ q  (42)
 T 

The local entropy inequality of Eq. (4) can be finally written as:

= Φ = − 2 (J q + Πv )∇T + pˆ dv − ∑k ρ k v k (∇µ k − Fk ) − ∑ j A j J j ≥ 0 (43)


dsi 1 1 1 1
dt T T T T

Eq. (43) has the following structure:

Φ = ∑α Fα Jα ≥ 0 (45)

in which Fα is the vector of the thermodynamic driving forces (the gradient of the velocity field, the gradient of
the chemical potential and the chemical affinity) and Jα is the vector of the fluxes (the effective stress tensor, the
diffusion flow and the velocity of reaction). In non-equilibrium thermodynamics, when the system is close to the
equilibrium, the fluxes of Eq. (45) can be expressed as a linear combination of the forces:

J k = ∑α Lkα Fα (46)

in which Lkα are called the phenomenological coefficients. Onsager [12] has shown that the reciprocal relations
hold:

Lkα = Lαk (47)

It is interesting to investigate when the constitutive relations in Eq. (46) remain valid imposing an arbitrary velocity
field vα to each component of the mixture. An important theorem of non-equilibrium thermodynamics to deal with
this problem has been developed by Prigogine. This theorem applies in the hypothesis of a negligible value of the
strain tensor. In the following this hypothesis is removed to analyse the applicability of the theorem in slope
stability analyses in which the strains of the continuum are accounted for. This theorem states that: if the local
mechanical equilibrium in Eq. (11) is satisfied (the inertial term is disregarded) the entropy production Φ in Eq.
(44) is invariant under the transformation [9]:

ρ k v k → J ′k = ρ k v k + ρ k v a (48)

in which va is an arbritary “drift” velocity that is applied to all components k of the mixture. The theorem can be
demonstrated using the Gibbs-Duhem equation. The Gibbs energy is equal to, at the local level:
dg = − sdT + ρδ ij ε ij dσ ij + ∑ k µ k dρ k = ∑ k µ k dρ k + ∑ k ρ k dµ k (50)

Considering constant isothermal conditions and that the term ρδijεij is equal to one the following relation can be
obtained:

∇σ ij = ∑k ρ k ∇µ k (51)

Considering only the term relating to the flow of component k and the chemical reaction and introducing the
hypothesis in Eq. (48) and Eq. (51) in Eq. (44) it can be obtained (in isothermal conditions) and in the hypothesis
of negligible value of the gradient of the velocity va:

Φ = ∑ k (∇ µ k − Fk )v k ρ k + ∑ k (∇ σ ij − ρ k Fk )v a − ∑ j A j J j ≥ 0 (52)

If Eq. (11) holds (with negligible inertia) the third term multiplying va vanishes and this demonstrates the theorem
(neglecting the gradient of va). Now consider the condition of non-negligible gradient of the velocity field. Eq.
(10) reads:

dρ k
= −∇J ′k − ρ k ∇ ⋅ v − ∇ ( ρ k v a ) + ∑ j υ jk v j (53)
dt

Applying the velocity va, Eq. (52) reads :

Φ=−
1
(J q + Πv )∇T + 1 pˆ dv − µ k ∂ρ k − 1 ∑k v k ρ k (∇µ k − Fk ) − 1 ∑ j A j J j ≥ 0
T 2
T ∂t T T
The third term in Eq. (36) vanishes if a phase transition occurs inside the REV. This condition is verified when:

∂ρ w ∂ρ g
µ g = µ w and =− (55)
∂t ∂t

Note the variation of the kinetic energy inside the REV due to va is null and then the expression of e is the same
of Eq. (19), the term belongs to the internal potential energy contrary to what proposed by Lin and Wu [13] even
in the case of dry soils such as the desert sands, because the sheared electrochemical forces of the air-vapour-solid
skeleton interfaces are present in any case. This term is not accounted for in any soils mechanics model and it is
un-correctly attributed to the second order work. The specific mass of the two components can be expressed with
a single equation including the interface between the two phases [14]:

ρα = ρ w + I (ρ g − ρ w ) (56)

in which I is the Dirac delta function that has an integral equal to zero when x belongs to the water phase and it is
equal to one when x is in the gas phase. Then Eq. (39) can be rewritten as follows:

∂ρ w ∂ρ g ∂I
=− − (ρ g − ρ w ) (57)
∂t ∂t ∂t

Using Eq. (57) the following kinematic constraint is obtained:

∂(φS w + φ − φS w ) ∂φ ∂I
= = (ρ w − ρ g ) = (ρ w − ρ g ) ∫ v* ⋅ n dΓ (58)
∂t ∂t ∂t Γ

in which v* is the velocity of the moving interface and n is the normal to the interface Γ. Note that the interface
velocity is independent with respect to va [15]. Eq. (42) states that a collapse of the porous network equal to the
phase transition of gas in water must occur to have an invariant expression of the entropy inequality during a so
called ( in soils´ mechanics ) “bifurcated” solution. Then the entropy preserves the mathematical structure of Eq.
(41):

Φ = ∑α Fα Jα ≥ 0 (59)

Recapitulating, in the hypothesis of:

a) Local mechanical equilibrium (Eq. (11)),


b) Negligible inertia forces,
c) A variation of porosity equal to the internal production of liquid due to phase transformation from gas to
water (Eq. (58)).

For each value of a drift velocity va the local production of entropy is invariant with respect to the transformation
J k → J ′k = J k + ρ k v a and the constitutive equations can be expressed with the Eq. (45) exactly as in the
equilibrium state in which va is not applied. Overall non-equilibrium condition can be modelled with the same
constitutive equations valid for the local equilibrium, in the hypothesis that a collapse strain due to a phase
transition of vapour into liquid is accounted for.

In order to investigate the consequences of the new formulation, the Prigogine´s theorem can be used. The theorem
assumes that: “in the linear regime, the total entropy production in a system subjected to flow of energy and matter
dSi/dt= ʃσ·dV reaches a minimum value at a non-equilibrium stationary state”. This means that if the entropy is
expressed by the equation:

Φ = ∫ ( F1 J 1 + F2 J 2 ) dV (60)

with a linear combination of the forces and fluxes. If a force is maintained at a fixed value, for example F2, this
implies that the flux J1 must be equal to zero. The Curie principle implies that in isotropic media, such as the one
considered, fluxes and forces couple only if they have the same order of symmetry, this means that scalar fluxes
couple with scalar forces, vectorial with vectorial and so on. This means that at t0 the equilibrium represented by
the condition (51) implies that:

J µ 
J q = −Πv and dSe = ∇ k k 
 T  (61)
The term related with the phase transition is related to the convective term of the entropy that is disregarded in the
formula (26). For the theorem, if the condition of phase transition in Eq. (55) is satisfied, the isotropic component
of the effective stress must be constant, at the same time. If a phase transition takes place the following condition
must be satisfied:

∇ v = dv = 0
(62)

This implies that:

∂ρ g
(µ g − µw ) =0 (63)
∂t

When a phase transition takes place the following condition holds:

∑ dρ
k
k =0
(64)
and then results in:

d ( ρs ) ds dρ ds ∂ρ k ∂µ
= ρ +s =ρ and ∇ ( vρ s ) = ∇ ( vµ k ρ k ) = µ k + ρk k (65)
dt dt dt dt ∂t ∂t
Demonstrating that the new formulation of the entropy production includes the convective term of the entropy,
this is the production of the internal entropy due to the motion of the reference point, particularly the point
describing the coexistence of the “interface” between the gas and liquid phase. In soils´ mechanics this part of the
inequality is also known as the Water Retention Curve (WRC). This is also the amount of the energy that is stored
in the interfaces and that flows out from the reference volume. This energy is also the enthalpy of formation that
is released during the dissociation of gas. It will be shown that this energy can be described by the Boltzmann
equation.

Integration of the rate equation


The constitutive model adopted for this work is the extension of the modified Cam clay model to the partial
saturation conditions [16]. The modification of the original scheme consists of the possibility of modifying the
plastic surface position in the elastic predictor stage. This condition is used to obtain the collapse strains that are
needed to satisfy the constraint in Eq. (58) and to use the Bishop´s formulation consistently with the
thermodynamically based expression in Eq. (37).

The elastic predictor stress (also called trial state, indicated with the superscript tr) can be defined by the equation:

dσ trn+1 = D e dε n+1 (61)

in which De is the fourth order elastic tensor and dε is the increment of the total strain tensor computed during the
global iterative linearization of the finite element procedure. The stress tensor is computed from the equation:

σ trn +1 = σ n + dσ trn+1 (62)

in which σij is the stress value at the end of time n. The preconsolidation pressure pc varies during the predictor
stage according to the following expression:

pctr = pcn exp(− b∆Sr ) (64)

where b is a constitutive parameter. The increment of Sr is computed during the global linearization through the
value of the capillary pressure.

The plasticity check can be obtained by incorporating Eq. (63) and (64) into the equation of the yield surface of
the Cam clay model and computing its value at time n+1:

2
 q tr 
f tr
=  (
 + p tr p tr − pctr ) (65)
M 

If the above expression is less than zero, the trial stress is inside the elastic domain, the total strains are elastic and
the trial stress is the actual stress. Otherwise a plastic corrector stage is required .

The plastic corrector updates the stress starting from the following equation:
 ∂f 
( )
dσ n +1 = D e dε n +1 − dε np+1 = D e  dε kl n+1 − dλn +1
∂σ
 (66)
 n +1 

where dεp is the increment of the plastic strains and dλ is the increment of the plastic multiplier. Note that the
model assumes that a flow rule is associated with the plastic potential coinciding with the yield surface.
Comparison with Eq. (61) leads to:

∂f
dσ ij n+1 = dσ trij n+1 − Dijkl
e
dλn+1 (67)
∂σ
n +1

The update of stress and internal variable can be obtained by solving the following system of the four scalar
equations:

 pn +1 = pntr+1 − Kdλn +1 (2 pn +1 − pc n )

q = q tr − 3G 2qn +1
 n +1 n +1
M2

p  ∂f  (68)
= pctr expθ n dλn +1  = pctr exp(θ n dλn +1 (2 pn +1 − pc n +1 ))
 c n + 1
∂p 
  n +1 

 q 
2

 f n +1 =  n +1  + pn +1 ( pn +1 − pc n +1 ) = 0
  M 

The above non-linear system of four equations contains four unknowns pn+1, qn+1, pcn+1 and dλn+1. K is the elastic
bulk modulus and G is the elastic shear modulus. This equation system can be solved using the Newton-Raphson
iterative method. Rewriting Eq. (68) in compact form:

 pn +1 − pntr+1 + Kdλn +1 (2 pn +1 − pc n ) = 0

q − q tr + 3G 2qn +1 = 0
 n +1 n +1
M2
p ⇒ A (x ) = 0 (69)
 c n +1 − pc exp(θ n dλn +1 (2 pn +1 − pc n +1 )) = 0
tr

 2
 q n +1 
 f n +1 =   + pn +1 ( pn +1 − pc n +1 ) = 0
  M 

in which xT is the vector of the unknowns. The Newton-Raphson technique solves the system by linearizing Eq.
(69) as:

−1
∂A(x n )  ∂A 
A(x n +1 ) = A(x n ) + dx n +1 = 0 ⇒ dx n +1 = −   A(x n ) (70)
∂x  ∂x 

The solution is obtained iteratively when the values of the updated variables satisfy the inequality:

A(x n + dx n +1 ) ≤ Res (71)


in which Res are prescribed residuals. If the hardening law is modified to account for the effect of the degree of
saturation, the consistency (that is the fourth equation of the system (69)) can be obtained with the expression:

∂f ∂f ∂f ∂f ∂f ∂f
df = dσ ij + dpc = dσ ij + θpc dλ − bpc dSr = 0 (72)
∂σ ij ∂pc ∂σ ij ∂pc ∂p ∂pc

The plastic multiplier is then obtained:

1  ∂f ∂f  ∂f ∂f
dλ =  dσ ij − bpc dSr  with H = − θpc (73)
H  ∂σ ij ∂pc  ∂pc ∂p

If Eq. (73) is introduced in the third equation of system (69), this can be written as (for simplicity only an isotropic
stress path is considered):

 
 
1  ∂f ∂f  ∂f
pc n +1 = pc n +1 expθ  ∆p − bpc ∆S r   =
tr

 ∂f ∂f  ∂p ∂pc  ∂p 
 − ∂p θp´c ∂p 
 c 
(74)
 
 
 1  ∂f ∂f  ∂f 
( )
= pc n exp − b∆S r exp θ
∂f

∂f  ∂p
∆p − 
bpc ∆S r 
 ∂pc  ∂p 
 − θ p´c 
 ∂pc ∂p 
tr
where the trial value of the preconsolidation pressure pc n +1 has been introduced. Eq. (74) shows that (it can be
demonstrated that b is equal to θ using the Maxwell symmetries):

 ∆p  ∆p
pc n +1 = pc n exp − b∆Sr + + b∆Sr  ( )
⇒ ln pcn+1 − ln( pc n ) = ∆ ln( pc ) = (75)
 p  p

Eq. (75) shows as the effect of Sr on the hardening law vanishes during the integration. This happens, because the
rate of the volumetric plastic strains and the degree of saturation are not independent variables (see Eq. (73)) and
thus the consistent tangent operator can be simply expressed by:

∂σ n+1
dσ n+1 = dε n+1 = Dep
n+1dε n+1 (76)
∂ε n+1

In the case of the model of Lewis and Schrefler this consideration does not hold. The presented constitutive
framework differs from the model of Jommi and Di Prisco [17] and Buscarnera and Nova [18].

The numerical approach can be investigated using the Laplace transformation [19]. The integration of the
hardening parameter can be formulated as:

∆S r T
dpc − b  − b ∆TS r t 
dt * = pc n +1e − b∆S r − pc n − b∆S r L{pc (t )} = pc n +1 − pc n
t
∫ dt e  − b∆S r ∫ pc e
− b∆S r t *
T
dt =  pc e
T  0 T

(77)
in which T is the time interval, t* is the ratio t/T and the operator L indicates the Laplace transformation. Eq. (77)
can be rearranged as:
pcn+1  e−b∆S r − 1
L{pc (t )} = (78)
b  ∆Sr 

Tamagnini [19] showed that the evolution of pc can be described with the inverse transformation that is represented
by the sum of a step function u(t) and a delayed step function u(t-b).

pcn+1
L−1{pc (S r )} = [u (t ) − u (t − b )] (79)
b
The graphical representation is shown in Fig. 2. Tamagnini [19] also showed that using the Maxwell´s symmetries
it is possible to demonstrate that the parameter b is equal to the parameter θ in Eq. (73). It is then interesting that
for each value of the parameter b along the x axis the resulting integral of the inverse transformation provides the
value of pcn+1. The limit of the transformation when the value of b and then the value of the hardening modulus H
approaches zero, which corresponds to the bifurcation condition is:

 1 − e −b∆S r 
lim L( pc (t ) ) = lim pc n +1   = pc n +1 ×1 (80)
b →0 b →0
 b∆S r 

and the integral results in:

∂pc ∂I ( x − xs )

V
∂t
dV = ∫ pc n+1
V
∂t
dV = ∫ pcn+1Vn n ⋅ dΓ
Γ
(81)

in which Ι is the Dirac delta function that is equal to zero everywhere except when x is equal to xs ,pcn+1Vn is the
momentum of the interface Γ between the liquid and gas and n is the unit normal to the interface surface. The
result in Eq. (80) can be obtained by applying the de l´Hôpital theorem. This result shows that the adopted
integration scheme provides an integral value that is equivalent to the integral value obtained during an integration
that accounts for a hardening modulus H that is equal to zero (implying the mechanical instability), but with a
value of H that is different from zero. This numerical feature overcomes the mesh dependency as shown in the
following sections. The evolution of the integral (79) can be presented like in Figure 1. The integration of Eq. (79)
can be also obtained from the definition of the Gibbs energy.

∂ρ k p ∂ρ
∫ ∇(J µ ρ ) = ∫ µ
k k k
∂t
k = ∫ cn+1 (u(0) − u(t − b) ) k =
b ∂t
(82)
v (λ − k ) ∂v
= ∫ pcn+1 (u(0) − u (t − b) ) s
v ∂t
The integration by parts of Eq. (82) leads to:

v (λ − k ) ∂v v
∫ p (u (0) − u(t − b) ) v ∂t = (λ − k ) v
cn +1
s s
pcn+1 ln v −
(83)
v v v v
+ ∫ (λ − k ) s pcn +1 ln vδ (v − v(t )) = (λ − k ) s pcn +1 ln f − (λ − k ) s pcn+1 ln(∆Φ)
v v vi v

Where the last term represents the space filled by the gas forming during the dissociation of the methane molecule.
Φ is the Lagrangian porosity. The Eq. (83) is the Boltzmann equation for the entropy. The first term is equal to
zero if a transition occurs, because it is not the eulerian sample volume that is measured, but its mass. The water
that flows out from the sample, which is used to compute the deformations, includes the gas which phase change
is not accounted for in volume computation. The change in entropy is computed by the second term. Two
considerations have to be underlined. First the diagonal term of the equilibrium represents the entropy without a
dependency on the temperature. The second consideration is that it is averaged on the solid skeleton fraction
because this is the only part of the mixture which mass fraction can vary if a transition of phase occurs between
the fluids in the Eulerian space. This means that the gas is the part of the continuum that cannot move during the
exchange of mass in the classic FEM formulation (the adsorbed water). This part of the volume can be recovered
during shearing or volume collapse. Note that in classic soils mechanics models the term related to the Boltzmann
equation is not present and this means that in the presented formulation the solid skeleton phase is removed from
the equilibrium during the integration, because the second term represents its entropy. This agrees with the Curie
principle. Figure 1 shows that for each values of b (that is the point on the time axis) the areas of the
rectangles are the same and equal to pcn+1. The value of b is:

vg v0 ∆t
b= = (82)
vs (λ − k ) vs (λ − k )

Where vg is the volume of the gas and v0 is the velocity of deformation of the lagrangian reference volume. The
independence of the value of the pv work with respect to time implies that the velocity can be also referred to the
velocity of the initial configuration and then in the equilibrium at t0 the velocity is referred to the initial volume
at time t0. This means that at the boundary of the reference configuration the displacement that are measured are
due to the gas phase motion. This consideration suggests that some techniques such as for example the axis
translation technique can induce misinterpretation of the results during phase transition phenomenabecause for
example in the oedometer at constant value the constrain can produces a variation in the deviator component of
the total stress. This can be observed in Eq. (58) in which the variation of porosity is obtained at the reference
configuration because the variation due to the motion of the point is zero.
Numerical analysis
The model is applied in the simulation of two sets of shear tests. In the first set, the sample is free to change the
original volume, in the second one; the boundary conditions are imposed in order to keep the original volume
constant. The two sets can be discussed to understand the difference between the classic constitutive modelling
that uses the Terzaghi stress and the proposed approach in which the gas enthalpy is modelled as a phase transition
phenomena. The three different meshes of the first set of analyses are reported in Fig.2. Fig. 3 reports the results
of the simulation that is divided in three stages. The first is an isotropic compression at constant suction; the second
is a deviatoric compression at constant suction (s1/s2 is the ratio of the vertical and horizontal total stress) and third
and last step is a wetting that is obtained by reducing the suction at the boundary of the sample.

The results presented in Fig. 3 show that the analyses are not affected by mesh dependency. This result is important
for the analyses of landslides where the instability is produced by the infiltration of water at the slope profile.

It is important to understand that for the condition (61) the component of the deviator stress depends only on the
flux of the latent heat at the boundary that for all the samples is the same. This is a typical definition of the external
work for an open system with phase transition phenomena in which the temperature gradient is not affecting the
results (during the early stage of the dissociation process).

It is interesting to apply to the same samples the constrains of the total volume, in the sense of the total area of the
sample, while the hydraulic boundary conditions can change with the constrain of maintaining the divergence of
the entropy due to the constant motion in order to satisfy the condition in Eq. (65). The variation of the pv work
of the water phase corresponds to the variation of the pv work of the gas phase at the equilibrium. This condition
can be obtained in laboratory, only if the gas phase can flow through the conduct of the liquid even if this is
maintained closed after the device for the pressure measurement. The result of the measure will be the cavitation
of the gas that produces an expansion of the measuring chamber. This produces the misleading in interpretation of
the triaxial tests in undrained conditions, because it is not the Terzaghi effective stress that is changing, but it is
the measure of the enthalpy of the gas, particularly the suction will remain constant contrary to what predicted by
classic models (because it is the radius of the bubble that is changing and not the pressure, see the equation (65)
with the discontinuity in chemical potential, the hysteresis) and the latent heat will increases. This produces the
changes in the stress. This means that the constitutive models for liquefaction that use the Terzaghi stress are not
able to model the exploitation of gas during the hydraulic fracturing, because they can find the equilibrium without
changing the temperature and then the variation of the deviator stress can be only due to the deformation of the
solid skeleton. This is in contradiction with the basic postulate of the effective stress that implies an undeformable
solid skeleton. The proposed approach is simpler and cost effective and not only thermodynamically consistent. It
is important to note that in the proposed model the gradient of the temperature can be arbitrary, because phase
transition occurs at constant temperature with a jump in the entropy value. This means that the involved energy
during fracturing is not used to heat the material but to produce the dissociation of the gas molecule. Fig. 4 shows
the formation of the shear bands due to the imposed boundary condition.

Fig. 5 shows that even in the case of constant volume tests the stress-strain curve is not affected by mesh
dependency. Fig. 4 explains as the variation of porosity in Eq. (83) cannot flow out of the mesh for the constraint
on the sample side displacements. This boundary constrains induce the formation of shear bands inside the mesh,
compressing the solid matrix. The shear bands are produced by the release of entropy during gas dissociation. The
interpretation of this test can give also some insights in other branches of porous mechanics such as the cancer
research with regards to the elasticity of human tissues. In this case the entropy production represents the product
of a chemical reaction accurring when the produced enthalpy cannot flow out from the boundary due to the constant
value of the area of the domain (closed system with no exchange of fluid masses but only heat that in absence of
a temperature gradient cannot flow). This means that the condition could be verified for example in cells that have
lost their capacity to deform during the changing pressure of the internal fluids, this condition corresponds to the
elastic expansion in hysteretic stress path condition (in regards to the water retention curve), because the human
body maintain its temperature constant and the only way to balance the latent heat is to decrease the fluid pressure
producing a swelling.
For geotechnical applications, it is very important to consider the gap between the area of the first kind of test and
the second one. This gap is the change in the lagrangian porosity due to the exploitation of the gas. All geo-
mechanics models for porous rocks and soils consider the eulerian volume constant (even for lagrangian
formulations and second gradient enhancement) This means that the sample can change the shape but not area. In
reality, in the presented model, the eulerian volume remains constant for the constraint on the phase transformation
(the divergence of the mixture velocity is zero), but the total volume changes due to the sink term in the gas
continuity equation. The evolution of the total sample volume during the test of Fig. 3 is reported in Fig. 6.

The figure shows that at the end of the test the total volume is changed of about the 2% even if the eulerian
reference volume has maintained its volume constant. This means that if a gas fracturing analysis is performed
with a classic model, with the second gradient plasticity, the resulting outflow of the gas can be underestimated of
about the 2% That means for a reservoir of trillions of cubic meters, a huge economic loss. Obviously this
consideration should be adapted to the type of rock encountered in the particular problem.

Conclusion
The paper has shown a thermodynamically based analysis of shear softening in multiphase porous materials. The
analysis has been conducted using two different approaches. In the first, the sample is free to deform and in the
second the volume is held fixed. That means, the divergence of the entropy flow does not produce any effect on
the volume of the domain. The analysis has demonstrated that the occurrence of the shear band is due to the
constraint on fixed area of the sample. If the constraint on the boundary nodes is removed the instability is diffuse
and there are no shear bands. This means, that the shear bands phenomenon is due to formation of cavities in the
porous matrix, that are due to the geometric impossibility for the porosity to flow out of the mesh. This is the
constraint, that is generally applied in laboratory in biaxial or triaxial undrained tests. The interesting consideration
is that the model formulation can predict an “undrained condition” with the outflow of the enthalpy of gas
dissociation. This is due to the fact that entropy jump is modelled properly in the hardening law. The flow out of
entropy corresponds to the volume loss due to the gas exploitation. In the present analysis the previous expansion
due to the hysteretic behaviour of the water retention is not modelled, the phase of the molecule inflow with an
osmotic change in pressure. The recovery of this expansion is the gas out-flow (in the analysis is the 2%) The
analysis suggests that classic model for soils mechanics with a second gradient enhancement can produce economic
misevaluation of the capability of gas depletion and potential environmental problems.
12

0 4 8 12 16 20
Time

Figure 1: evolution of the preconsolidation pressure for different values of b

Figure 2: different meshes for the same soil sample


3
Coarse
Intermediate
Fine
2.5
σ1/σ2

1.5

0 0.04 0.08 0.12 0.16 0.2


εs

Figure 3: shear tests simulation for the same soils sample. First stage compression and application of
the deviator stress, second stage wetting, s1/s2 is the ratio between the vertical and the horizontal
component of the total stress, es is the deviator strains.

Figure 4: Different meshes for the shear tests with constant area and changing the hydraulic
boundary conditions.
Figure 5: Results from the shear driven tests with constant flow of entropy by the sides.

Figure 6: Evolution of volume due to wetting


ACKNOWLEDGENTS

The authors wish to thank the European Commission for the financial support within its 7th Frame Work
Program to the following projects: MUMOLADE (Multiscale Modelling of Landslides and Debris Flows),
Contract Agreement No. 289911 within Marie Curie ITN; REVENUES (Reinforced Vegetation Numerical
Evaluation of Slopes), Grant Agreement No.: 324466, within the Industry-Academia Partnerships and Pathways
(IAPP).

REFERENCES

[1] Terzaghi K (1943) Theoretical Soil Mechanics. Willey and Sons Ltd., New York

[2] Roscoe KH, Burland JB (1968) On the Generalized Stress-Strain Behaviour of Wet Clay. In:
Engineering Plasticity, Cambridge, pp 535-609

[3] Chambon (1994)

[4] Alonso E, Gens A, Josa A (1990) Constitutive model for partially saturated soils. Géotechnique
40(3):405-430

[5] Hassanizadeh SM, Gray WG (1990) Mechanics and thermodynamics of multiphase flow in porous
media including interphase boundaries. Advances in Water Resources 13(4):169-186

[6] Lewis RW, Schrefler BA (1998) The finite element method in the static and dynamic deformation
and consolidation of porous media. Wiley & Sons Ltd, Chichester

[7] Prigogine I (1947) Etude thermodynamique des phénomines irreversibles. Desoer, Liege.

[8] de Groot SR, Mazur P (1994) Non-equilibrium thermodynamics. Dover Publications Inc., New
York

[9] Kondepudi D, Prigogine I (1998) Modern Thermodynamics: From Heat Engines to Dissipative
Structures. Wiley and Sons Ltd, Chichester

[10] Bishop AW (1959) The principle of effective stress. Teknisk Ukeblad 39(2):197-213

[11] Coussy O (2004) Poromechanics. Wiley & Sons Ltd., Chichester

[12] Onsager L (1931) Reciprocal Relations in Irreversible Processes. I. Phys. Rev. 37:405-426

[13] Lin J, Wu W (2012) Numerical study of miniature penetrometer in granular material by discrete
element method. Philosophical Magazine 92(28-30): 3474-3482

[14] Juric D, Tryggvason G (1998) Computation of boiling flow. Int. J. of Multiphase flow 24(3):387–
410

[15] Houlsby GT (1997) The Work Input to an Unsaturated Granular Material. Géotechnique
47(1):193-196
[16] Tamagnini R (2004) An extended Cam-clay model for unsaturated soils with hydraulic hysteresis.
Geotechnique 54(3): 223-228

[17] Jommi C, di Prisco C (1994) A simple theoretical approach for modelling the mechanical
behavior of unsaturated granular soils. Proc., Conf. Il ruolo dei fluidi nei problemi di ingegneria
geotecnica. Mondovi, pp 167-188

[18] Buscarnera G, Nova R (2009) An elastoplastic strain hardening model for soil allowing for
hydraulic bonding-debonding effects. Int. J. Numer. Anal. Meth. Geomech. 33(8):1055–1086

[19] Tamagnini R, Mavroulidou M, Gunn MJ (2010) Numerical integration and analysis of


equilibrium in unsaturated multiphase media. Proc., 7th European Conference on Numerical Methods
in Geotechnical Engineering, NUMGE 2010. Trondheim, Norway

You might also like