Вы находитесь на странице: 1из 31

Thermo-hydro-

Thermo-hydro-mechanical mechanical
analysis of partially saturated analysis

porous materials
113
Dariusz Gawin
Received March 1995
Technical University of Lodz, Lodz, Poland, and Revised October 1995
Bernhard A. Schrefler
Universita’ di Padova, Padova, Italy
Introduction
In recent years we have observed increasing interest in thermo-hydro-
mechanics of partially saturated porous materials. Several mathematical
models of transport processes have been proposed, e.g.[1-5]. The existing
numerical models, however, do not take properly into consideration the whole
complexity of physical phenomena concurrent with heat, water and gas flow in
deforming porous media. Typical simplifying assumptions are: rigid solid
matrix, e.g.[6-15]; static gas phase (constant gas pressure equal to the
atmospheric pressure), e.g.[1,2,15]; and omission of phase changes (evaporation-
condensation), e.g.[2,5]. Moreover, during computer simulation of drying
processes the boundary conditions (BC) of the first kind, i.e. fixed values of
capillary (water) pressure and temperature, or BC of the second kind, i.e. fixed
vapour flux and heat flux, were usually applied, e.g.[1,5,11]. From the physical
point of view much more correct are the BC of the third kind, i.e. fixed values of
vapour density (capillary pressure) and temperature, and fixed fluxes of vapour
and of heat, e.g.[14-16].
The present paper follows a companion paper by Gawin et al.[17], in which a
fully coupled numerical model to simulate slow transient phenomena involving
flow of heat, water and gas in deforming porous media was first proposed. Here
we emphasize the importance of physically correct modelling and of the
appropriate choice of the boundary conditions. The model is based on strong
physical background and is aimed at handling situations which span from fully
saturated to almost dry conditions. It consists of balance equations of mass,
linear momentum and energy, as well as of the appropriate constitutive
equations.
The macroscopic balance equations applied in the model were obtained
in[1,4,10,18-20] from balance equations at microscopic level by use of spatial
averaging operators[9,10,18-25].

Engineering Computations,
This work has been carried out within the framework of HCM Project ALERT Geomaterials. The Vol. 13 No. 7, 1996, pp. 113-143.
first author’s research has been supported by Phare TEMPUS Project. © MCB University Press, 0264-4401
EC The gas phase is modelled as an ideal gas composed of dry air and water
13,7 vapour, which are considered as two miscible species. Phase changes of water
(evaporation-condensation) and heat transfer through conduction and
convection, as well as latent heat transfer are considered. This model is an
extension of the heat and mass transfer formulation of Baggio et al.[14], based
on Whitaker’s approach[9,10], to the case of deformable porous media. The
114 model makes use of a modified effective stress concept together with the
capillary pressure relationship.
The macroscopic balance equations, after introduction of the constitutive
relationships, are discretized in space by means of the finite element method
and in time by finite differences. The chosen macroscopic primary variables –
gas pressure, capillary pressure, temperature and displacements – correspond
to real measurable quantities directly linked to laboratory practice. This is an
important aspect when selecting the appropriate constitutive equations.
In this paper the governing equations are summarized and the discretized
equations in matrix form are shown. They are solved by means of a Newton-
Raphson type procedure. An error analysis for the discretization in time is
presented. Three examples are solved, which are used for further validation of
the code based on the outlined approach. These examples demonstrate the
importance of proper physical modelling of heat and mass transfer in porous
media. Particular attention is paid to the effects of two-phase flow, as compared
to simpler one-phase flow model (constant gas pressure equal to atmospheric
pressure), to the constraint effects of phase change, which influences strongly
the capillary and gas pressure solution profiles, as well as to the influence of
appropriate choice of the boundary conditions on the solution obtained.

Constitutive equations
The partially saturated porous medium is treated as multiphase system where
the voids of the skeleton are filled partly with liquid water and partly with gas
assumed to behave as an ideal mixture of dry air and water vapour. The state of
the medium is described by gas pressure pg, capillary pressure pc, temperature
T and displacement vector of the solid matrix u.
The saturation of liquid water S is an experimentally determined function of
capillary pressure pc and temperature T:
(1)
while its pressure pl can be expressed at equilibrium state[26], as:

(2)
The equation of state of perfect gases and Dalton’s law, applied to dry air (ga),
vapour (gw) and moist air (g) give:

(3)
Owing to the curvature of the meniscus separating the liquid (water) phase Thermo-hydro-
from the gas phase inside the pores of the medium (considered as a capillary mechanical
porous body), the equilibrium vapour pressure can be obtained from the Kelvin analysis
relationship which gives the relative humidity (RH) of the moist air inside the
pores:

115
(4)

where the water vapour saturation pressure p gws, which depends only on
temperature T, can be obtained from the Clausius-Clapeyron equation or from
empirical correlations, see for example[27].
The constitutive law of the solid phase is introduced through the concept of
modified effective stress σ":

(5)
where σ is the total stress tensor, I the unit tensor, α is the Biot’s constant and p
an average pressure of the mixture of fluids surrounding the grains, which in
the case of small dependence of the Helmholz free energies on void fraction
reduces[26] to the commonly employed relation:

(6)
The last term in (6) has been added because we use here absolute pressure –
pure atmospheric pressure does not cause any deformation of the medium.
The Biot’s constant is introduced to account for the volumetric deformability
of the particles and was shown[2,28,29], to be:

(7)
where KT and KS mean the bulk modules of the porous medium and the solid
phase, respectively.
For a discussion about the validity of the effective stress concept the
interested reader is referred to[30].
The constitutive relationship for the solid skeleton has the form:

(8)
where D is a tangent matrix, dεT = I βs3 dT, is the strain increment caused by
thermo-elastic expansion, βs means the cubic thermal expansion coefficient of
the solid and d ε o represents the autogeneous strain increments and the
irreversible part of the thermal strains[1].
EC Governing equations
13,7 In the following, only slow phenomena and small displacements are considered.
Thermal equilibrium between solid matrix, gas and liquid is assumed, so the
temperature is the same for the three constituents.
The formulation of heat and mass transfer in porous media is obtained
starting from the appropriate local equations expressing the laws of continuum
116 physics (see, for example[31]), specifically the continuity equation for each
species considered, the Navier-Stokes equation for quasi-steady creeping
flow (i.e. with the time-dependent and convective terms neglected) and the
energy equation (enthalpy balance) with viscous dissipation and reversible
work neglected. In general such equations cannot be solved because of the
complex geometry of the porous media, but using the spatial averaging
technique[9,18-20], we can obtain equations averaged on a representative
elementary volume (REV) of the porous medium. The multiphase Darcy
equation, applied to the liquid phase, gives:

(9a)

(9b)

(9c)

while for the gas phase yields:

(10a)

(10b)

where K is intrinsic permeability tensor, Krg and Krl are relative permeabilities
of the gaseous and liquid phase, vl and v g are the velocities of liquid and
gaseous phase relative to the solid phase and the vector b indicates the specific
body force term (normally corresponding to the acceleration of gravity).
Usually, in the gas phase, the body force term ρgb is negligible. The bubbling
pressure pb is the minimum value of pc on a drainage capillary pressure curve,
at which a continuous gas phase exists in the void space[25]. The critical
pressure p cr , corresponding to equilibrium water saturation value, is the
maximum value of capillary pressure, for which flow of water is possible,
because water exists in the funicular saturation state[25], forming continuous
paths in the voids of porous medium.
For the binary gas mixture (dry air and water vapour), the Fick’s law gives
the following relative average velocities of the diffusing species:
(11) Thermo-hydro-
mechanical
analysis
because in this case we have:
117
(12)

The governing equations cast in conservation form can be written[9], as


follows:

Dry air conservation equation

(13)

Water species (liquid-vapour) conservation equation

(14)

Energy conservation (enthalpy balance)

(15)

Linear momentum balance equation for the whole mixture, neglecting inertial
effects, can be written in terms of total stresses[3,5,32], as:
(16)
where ρ is the averaged density of the mixture:
(17)
φ being the porosity, ρs the density of the solid, ρl the density of the liquid water
and ρg the density of the gas.
EC It is further necessary to define the initial and boundary conditions. The
13,7 initial conditions specify the full fields of gas pressure, capillary pressure,
temperature and displacement:
(18)

118 The boundary conditions can be, see e.g.[22], of the first kind or Dirichlet BC on
Γ i1, of the second kind or Neumann BC on Γ 2i , and of the third kind or Cauchy
BC on Γ i3, where the boundary Γ = Γ i1 ∪ Γ 2i , ∪ Γ i3.
BC of the first kind, when there are prescribed values on the boundary for
gas pressure, capillary pressure, temperature and displacements are as follows:

(19)
BC of the second kind, when we have the space averaged fluxes on the boundary
prescribed, can be written in the form:

(20)

where n is the unit normal vector, pointing toward the surrounding gas, qga,
qgw, ql and qT are respectively the imposed dry air flux, the imposed vapour
flux, the imposed liquid flux and the imposed heat flux, and t is the imposed
traction.
BC of the third kind, when both the value and its normal derivative
(corresponding flux) are prescribed on the boundary, usually occur at the
interface between the porous media and the surrounding fluid. In the case of
heat transfer they correspond to the known Newton’s law of cooling, e.g.[33].
This kind of BC for the water species conservation equation (14) and for the
energy conservation equation (15), can be expressed as follows:

(21)
where ρ gw∞ and T ∞ are, respectively, the mass concentration of the water Thermo-hydro-
vapour and the temperature in the undisturbed gas phase distant from the mechanical
interface, while αc and βc are convective heat and mass transfer coefficients. analysis
It is physically admissible, that BC of the second and of the third kind for the
energy equation can be applied at the same portion of boundary, like, for
example, during action of thermal radiation on the external surface of porous
medium. But even in this case it is possible to express them formally as BC of 119
the third kind, using appropriate equivalent value of air temperature[15].

Discretization in space
Discretization in space of the governing equations is carried out by means of the
finite element method. The notations of Zienkiewicz and Taylor[34,35], are used
in the following, together with vector notation. The unknown variables are
expressed in terms of their nodal values as:

(22)
The integral or weak form of the heat and mass transfer equations obtained by
means of the Galerkin procedure and with the usual choice of shape functions
N, can be expressed in matrix form as follows:

(23)
where all the matrices are listed in the Appendix.
Using the principle of virtual work and taking into account the boundary
condition (20) and (21), the linear momentum balance equation (16) can be
written in a weak form as[1]:
(24)

Application of (2), (6), the effective stress relationship (5), as well as the
definition of the strain matrix relating strain and displacement:
(25)
allows us to express (24) in matrix form:
(26)
EC where fu and coupling matrices Kug , Kuc , Kut are listed in the Appendix.
13,7 The effective stresses are obtained from integration of (8), starting from the
known initial values of the problem.
The above discretization in space of the conservation equations (and of the
other ones required to complete the model) with the standard Galerkin method
(weighted residuals)[1,34,35], results in a non-symmetric, non-linear and
120 coupled system of ordinary differential equations of the form:
(27)
where:
(28)

and the non-linear (matrix) coefficients C(x), K(x) and f(x) are obtained by
assembling the submatrices indicated in (23) and (26).
We assume consistency and convergence of the finite element discretization
in space[34].

Discretization in the time domain and error analysis


The time-discretization is accomplished through a fully implicit finite difference
scheme (backward difference)[34,35]:
(29)
where:

n being the time step number and ∆ t – the time step length.
We focus our attention on the global error:
(30)
where ~ x n+1 is the exact solution of equation (29) and L is the number of
performed iterations for time step n + 1. To simplify calculations, we assume
that there are no round-off errors generated by the algorithm, except those
introduced by the initial values at time t0.
If we replace xn+1 and xn in equation (29) by the corresponding exact values,
we obtain:
(31)

where r n+1 is the local truncation error of equation (29). The calculated
L
numerical solution x n+1 satisfies equation (29), so we can subtract equation (29)
written for x n+1 and xnL respectively from equation (31). This results in:
L
Thermo-hydro-
(32) mechanical
Neither symmetry nor positive definiteness of the matrices are assumed here. analysis
By a recursive application for index n, we have from equation (32) that:

121

(33)

The total error e Ln+1 is influenced by the local truncation error rk, being k ≤
n + 1 and by the initial round-off error e00 made at the starting step, because
then for n = 0 and L = 0 :
(34)

with ~
x 0 = x0, the exact solution of equation (29). Thus:

(35)

is the sufficient stability condition for the applied solution procedure, where
||.|| is the spectral norm. The stability condition and the consistency property:
(36)

together with the iteration convergence, indicated below, are sufficient for the
above procedure to be globally convergent[36].
The interested reader will find more information about consistency analysis
and estimation of the time discretization error in[36].
Considering the non-linearity of the system of equation (29) the solution is
obtained with a Newton-Raphson type procedure:

(37)

where l is the iteration index, and at the end of each iteration the primary
variables are updated as follows:
(38)
EC For the convergence of Newton-Raphson procedure (37) within one time step to
13,7 a unique solution (iteration convergence) the interested reader is referred to any
textbook on numerical methods, e.g.[37].
During computations a problem arises when the medium is fully saturated
because in this case the gas pressure pg is undefined and the capillary pressure
pc has no physical meaning and, on the other hand, the state of the liquid water
122 in the medium is described by two variables (since in such situation there are
only two DOF on top of solid displacements: liquid pressure pl and temperature
T ).
The problem is treated with a formal modification of the relationship
between saturation S and capillary pressure pc: when the saturation S becomes
equal to one, the sign of the capillary pressure is formally set negative and the
value equal to the pressure in the liquid above gas pressure (see equation 2).
For fully saturated state we use equation (9a) for description of liquid flow,
while gas pressure is set to equal to the atmospheric pressure (the dry air
conservation equation is dropped).
In real porous media there is no gas flow for capillary pressures below
bubbling pressure[21,22,25], because gas phase is discontinuous and forms
bubbles. Thus we perform “switching”, element by element, from fully
saturated to partially saturated state equations (or vice versa) when capillary
pressure reaches bubbling pressure value ( pc = pb and S < 1).
In practice, capillary and gas pressure oscillations can arise when this switch
is performed. As already pointed out in[17], these oscillations are possibly due
to the sudden switch of element behaviour (there is a change in governing
equations) in a part of the domain, which in turn causes different convergence
to solution in fully saturated and partially saturated zone of the domain and
produces oscillations in the Newton iteration procedure[38]. Another reason is
occurrence in the matrices C and K (equation 27) of very small diagonal terms
related to gas pressure (because of very low value of gas relative permeability),
which create numerical problems. An efficient way of avoiding the mentioned
oscillations is assumption of a lower limit for the value of relative permeability
of gas phase (e.g. 0.0001, as was assumed in the drainage test example of the
present work), which seems to be physically admissible.
Our approach to fully saturated state differs from that used in[30], where gas
phase is considered as always present in porous medium. In[30] the capillary
pressure-saturation relationship is extended in an asymptotical manner in the
area of formally negative capillary pressures, which are in fact the pressures in
the liquid above the gas pressure. Thus in that approach, gas flow (even if very
small, because of low value of the relative permeability of gas phase) takes place
even for fully saturated state, which is not correct from physical point of view.
Based on the presented discretization, the HMTRA-DEF research computer
code has been developed[17], for the solution of the non-linear and non-
symmetrical system of equations governing heat and mass transfer in a
deforming porous medium. This code is here further validated and some of its
potentialities are shown.
Numerical examples Thermo-hydro-
It is very difficult to choose some appropriate tests to validate the proposed mechanical
model because of lack of any analytical solutions for this type of coupled analysis
problem, where deformations of the solid skeleton are studied together with the
saturated-unsaturated flow of mass and heat transfer. There are also very few
documented laboratory experiments. One of these is the experiment conducted
by Liakopoulos[40], on the isothermal drainage of water from a vertical column 123
of sand. This test was also used by Narasimhan and Witherspoon[41], Schrefler
and Simoni[42], Zienkiewicz et al.[2], as well as by Schrefler and Zhan[30], to
check their numerical models. The same example was solved in[17], but only for
one-phase flow. Here attention is paid to the effects of two-phase flow, as
compared to the previous solution[17]. Then two other examples relating
respectively to drying of a concrete wall subjected to compression[43,44], and
thermoelastic consolidation of partially saturated clay[5,11], are presented.
In all these examples isoparametric Lagrangian elements are used, the same
for the pressures, temperature and displacement fields. Furthermore, linear
elastic material behaviour is assumed.
Two other examples, relating to subsidence due to pumping from a phreatic
aquifer[46], as well as non-isothermal consolidation of fully and partially
saturated elastic porous material[1], solved by means of the proposed numerical
model, were presented in[17].

Drainage test
In the experiment of Liakopoulos[40], a column of Perspex, 1 metre high, was
packed with Del Monte sand and instrumented to measure the moisture tension
at several points along the column. Before the start of the experiment (t < 0)
water was continuously added from the top and was allowed to drain freely at
the bottom through a filter, until uniform flow conditions were established. At t
= 0 the water supply was ceased and the tensiometer readings were recorded.
The porosity, φ = 29.75 per cent, and hydraulical properties of Del Monte sand
were measured by Liakopoulos[40], by an independent set of experiments.
During computations the approximated equations of the Liakopoulos’
saturation-capillary pressure and relative permeability of water-capillary
pressure relationships of the following form:

(39)
were applied.
For numerical purposes the column was simulated by 10 and 20 four-, eight-
and nine-node isoparametric finite elements of equal size and different meshes
in the time domain (∆τ = 10s, 1s or 0.5s) were used giving practically the same
results. At the beginning, besides the uniform flow conditions (i.e. unit vertical
gradient of the potential and pc = 0 on the top surface), mechanical equilibrium
state was assumed. The boundary conditions were the following: for the lateral
EC surface, qT = 0, uh = 0, where uh means horizontal displacement of soil; for the
13,7 top surface, pg = patm, where patm means atmospheric pressure, T = 293.15 K;
for the bottom surface, pg = patm, pc = 0 for t >150s, while water pressure pl was
assumed to change linearly from the initial value to zero for t < 150s, T = 293.15
K, uh = uυ = 0, where uυ means vertical displacement of soil.
Liakopoulos did not measure the mechanical parameters of the soil, so the
124 Young modulus of soil was assumed as E = 1.3 MPa, Poisson’s ratio as ν = 0.4
and Biot’s constant as α = 1, similarly to[17,30,42].
The calculations were performed for one-phase flow (gas pressure was
assumed equal to the atmospheric pressure in partially saturated zone), as well
as for two-phase flow. For the last case the switching between saturated and
unsaturated solution was performed at p c = 2000 Pa (S = 0.998), which
corresponds to bubbling pressure of the analysed sand[40], as well as at pc = 0
Pa, in order to analyse the effect of the “switching” value on the obtained
solution.
The relative permeability of gas phase was assumed according to the
relationship of Brooks and Corey[39]:

(40)
with the additional lower limit Krg ≥ 0.0001.
The resulting profiles of water pressure for two-phase flow with switching at
pc = 2,000 Pa (solid lines with different symbols for different time stations) and
for one-phase flow (dashdot line) are compared with the experimental results of
Liakopoulos[40], (fine lines) in Figure 1, showing better agreement for one-
phase flow. This results from the fact that in[42] Young modulus value has been
numerically “fitted” just for that case. The profiles of vertical displacements,
water saturation and capillary pressure for two-phase flow with switching at pc
= 2,000 Pa (solid lines) and at pc = 0 Pa (dash-dot lines) are compared with the
results for one-phase flow (fine lines) in Figure 2.
There are some differences between one- and two-phase flow solutions for
vertical displacements (Figure 2), although their final values are similar. A
difference between the one- and two-phase flow solution is more dramatic for
saturation of water and capillary pressure (Figures 3 and 4). In the bottom zone,
where no gas flow occurs, the differences are small, but higher up there is a
qualitative change of the soil behaviour caused by the presence of gas under
pressure (see also Figure 5). The solution for two-phase flow with switching at
pc = 2,000 Pa is similar to the one-phase flow solution close to the bottom and
tends to the two-phase flow solution with switching at pc = 0 Pa in the upper
part of the sand column, while in the middle part it forms characteristic
constant-value zones corresponding to switching values of capillary pressure or
saturation. Generally, in the upper zone, where gas flow occurs, the gradients of
saturation and capillary pressure are higher for the two-phase solution, which
Water pressure (Pa) Thermo-hydro-
0
mechanical
–1,000 analysis

–2,000 5

125
–3,000

–4,000
10
–5,000 Figure 1.
Comparison of the
–6,000 20 numerical solutions for
Key 20 water pressure (two-
–7,000 5 minutes 30 phase flow with
10 minutes 30 switching at pc = 2,000
20 minutes Pa indicated by solid
–8,000 30 minutes 60 lines and one-phase flow
60 minutes solution indicated by
–9,000 120 minutes dash-dot lines) with
experimental results of
–10,000 Liakopoulos[41] (fine
0 0.2 0.4 0.6 0.8 1 lines)
Height (m)

is qualitatively in accordance with the solution obtained by Schrefler and


Zhan[30].
The profiles of gas pressure for two-phase flow with switching at pc = 2,000
Pa (solid lines) and at pc = 0 Pa (dash-dot lines) are compared in Figure 5. The
visible differences are caused by different assumptions about gas flow (no gas
flow for the zone where S > 0.998 in the first case). There is a small difference of
the under pressure amplitudes, nevertheless a qualitative similarity of the gas
pressure profiles in the zone, where its flow occurs, is obvious.
This example shows that modelling of the transition from fully saturated to
partially saturated condition (and vice versa) is very sensitive to the procedure
adopted.

Drying of a concrete wall


This problem deals with drying of a concrete wall subjected to compression by
a uniform surface pressure, which was solved previously by Lassabatere and
Coussy[43,44]. Because of lack of sufficient data about their solution of the
problem only a qualitative comparison of obtained results will be performed.
A vertical concrete wall of thickness d = 20 cm, initially at constant
saturation of water corresponding to relative humidity (RH) = 90 per cent, is
subjected both to a rapid variation of external relative humidity (i.e. capillary
EC 0
Vertical displacement (m)
13,7
–0.0005

–0.001
126
–0.0015

–0.002

–0.0025 10

10
Figure 2. –0.003 Key 20
Comparison of the two- 5 minutes
phase flow solutions 10 minutes
with switching at pc = –0.0035 20 minutes
2,000 Pa (solid lines) 60 minutes
and at pc = 0 Pa (dash- 120 minutes
–0.004
dot lines) with one- 120
phase flow solution (fine
lines) made for vertical –0.0045
displacements 0 0.2 0.4 0.6 0.8 1
Height (m)

pressure on the wall surface) reaching a final value RH = 50 per cent and to
uniform mechanical pressure t on its faces.
Because of geometrical symmetry of the analysed problem, the solution was
obtained only for one half of the domain (space co-ordinate x = 0 on symmetry
surface).
The initial conditions were: T = 293.15 K, pg = 101,325 Pa, pc = 13.97 MPa
and deformations consistent with mechanical equilibrium.
The boundary conditions were the following:
• for the symmetry surface: qT = 0, ql = 0, qg = 0, uh = 0;
• for the external surface: T = 293.15 K, pg = 101,325 Pa, pc = 93.78 MPa,
t = 14 MPa.
Because of lack of any material data of the concrete analysed, the following
parameters taken from another paper about concrete drying of the same
authors[45], were assumed: porosity φ = 35 per cent, Young’s modulus E = 14
GPa, Poisson’s ratio ν = 0.25 and relationship between capillary pressure and
saturation of the form[45]:

(41)
Saturation (–) Thermo-hydro-
1
mechanical
0.99
analysis
10

0.98
127
20
0.97
10
0.96

0.95

0.94 Key 60
Figure 3.
5 minutes Comparison of the two-
10 minutes phase flow solutions
0.93 20 minutes with switching at pc =
60 minutes 2,000 Pa (solid lines)
120 minutes and at pc = 0 Pa (dash-
0.92 120
dot lines) with one-
phase flow solution (fine
0.91 lines) made for
0 0.2 0.4 0.6 0.8 1 saturation of water
Height (m)

Lassabatere and Coussy[45], use a diffusion equation for description of moisture


–10
transfer, so they give only the value of the global diffusion coefficient D = 10
m2/s. In our solution we used an absolute permeability value K = 3.97 10–21 m2
lying in the range of possible values for this kind of concrete.
Besides we assumed: cubic thermal expansion coefficient βs = 0.9 10–6 K–1,
thermal conductivity λ = 2.5 W/mK, Sirr = 0.09, Biot’s coefficient α = 0.5 and
relative permeabilities of water and gas according to[47]:
(42)
where effective saturation:
Se = ( S – Sirr ) /(1 – Sirr ). (43)
The problem was also solved for Biot’s coefficient α = 1 giving similar results,
but with slightly smaller amplitudes of the variations in the profiles of relative
humidity, capillary pressure and gas pressure, so the only solution for the case
of Biot’s coefficient α = 0.5 is presented here.
For numerical purposes the wall was simulated by ten eight-node
isoparametric elements of different width (4 × 0.25cm, 3 × 1.0cm, 3 × 2cm).
Temporal discretization was performed with an initial step of 0.01 s during the
EC Capillary pressure (Pa)
10,000
13,7
9,000
Key
120 minutes
8,000 60 minutes
20 minutes
128 7,000 10 minutes
5 minutes
6,000 20

5,000
10
4,000
Figure 4.
Comparison of the two- 3,000
5
phase flow solutions
with switching at pc =
2,000 Pa (solid lines) 2,000
and at pc = 0 Pa (dash-
dot lines) with one- 1,000
phase flow solution (fine
lines) made for capillary 0
pressure 0 0.2 0.4 0.6 0.8 1
Height (m)

first 1,000 steps, then of 1 s during the next 90 steps multiplied by 10 after
repeating 90 steps until 107 s. A 3 × 3 Gaussian integration scheme was applied.
The resulting profiles of relative humidity, capillary pressure and gas
pressure are presented in Figures 6-8. In comparison with the results of
Lassabatere and Coussy[43,44], (only relative humidity profiles), a similar
maximum increase of relative humidity (about 1 per cent) close to the wall
surface is observed in Figure 6, although time evolution of obtained profiles as
well as the range of the characteristic hump (much smaller for our solution) over
initial profile are different. Similar phenomena but in the opposite direction (pc
decrease) are visible in Figure 7 also for the capillary pressure profiles.
In Figure 8 a significant increase of gas pressure (about 70 kPa) is observed
first in the zone close to the wall surface, then in the whole wall domain. Time
evolution of gas pressure profiles is qualitatively similar to that of relative
humidity obtained by Lassabatere and Coussy[43,44].
The reason for the observed differences between the actual results and those
of Lassabatere and Coussy[43,44], is phase change, which through the energy
balance equation imposes physical “constraints” on vapour pressure variations.
In order to evaporate a certain amount of water inside the pores we must supply
latent heat of evaporation, which is relatively high in comparison to other terms
of the energy balance equation. This important term was not taken into account
Gas pressure (Pa) Thermo-hydro-
102,000
mechanical
5
analysis
101,000

10
100,000 129

99,000

98,000

Key 120
97,000 5 minutes
10 minutes Figure 5.
60
20 minutes Comparison of the gas
60 minutes 20 mins pressure profiles (the
96,000 120 minutes two-phase flow solution)
for switching at pc =
2,000 Pa (solid lines)
95,000 and at pc = 0 Pa
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 (dashdot lines)
Height (m)

by Lassabatere and Coussy[43,44]. Hence in the present solution, instead of


significant changes of capillary pressure (relative humidity), much more
sensible variations of gas pressure are observed. This confirms that phase
change is of importance, being a physical constraint for capillary pressure
change, and should be taken into consideration to obtain physically correct
results.

Thermoelastic consolidation
This problem deals with thermoelastic consolidation of partially saturated clay
for which previous solutions are known[5,11]. In[11] the case of constant
permeability and thermal properties was studied in detail, while in[5] the
problem was solved for the relationships between relative permeabilities of
water and air, saturation of water and the capillary pressure proposed by
Brooks and Corey[39]. Dakshanamurthy and Frelund[11], have analysed four
different cases, i.e. isothermal/non-isothermal consolidation/swelling, whereas
here only the non-isothermal consolidation problem is studied.
The layer, of 10cm thickness, of partially saturated subgrade soil (S = 0.445),
initially in a state of equilibrium, is subjected to a sudden environmental
change. The surface temperature jump of 15K and capillary pressure jump of
140kPa cause simultaneous heat and mass transfer which physically
EC 0.95
Relative humidity (–)
13,7
0.9
Key Time (s)
1.00E+02
0.85
130 5.00E+02
0.8 1.00E+03
2.00E+03
0.75
3.00E+03
5.00E+03
0.7
1.00E+04
0.65 1.00E+05
1.00E+06
0.6
1.00E+07
Figure 6.
The actual solution in 0.55
the zone close to the
surface of the concrete 0.5
wall – relative humidity 0.09 0.0925 0.095 0.0975 0.1
Distance X(m)

corresponds to the drying process caused by warm dry-air flux. Porosity,


density, absolute permeability, as well as thermal conductivity and capacity of
the soil were assumed the same as in[11]. Because of lack of any mechanical
data for the soil analysed, a Young modulus E = 60MPa, Poisson ratio ν =
0.2857 and cubic thermal expansion coefficient β s = 0.9 × 10–6, lying in the
range of possible values for this kind of clay[48], were assumed. The
relationships of Brooks and Corey[39] for relative permeabilities of water and
gas, as well as for saturation-capillary pressure curve were assumed as follows:

(44)

where S e is defined by equation (43), bubbling pressure p b = 133.813kPa,


irreversible saturation Sirr = 0.3216, λ = 2.308, which corresponds to the given
saturation-capillary pressure values for consolidation case in[11].
The initial conditions were: T = 283.15K, pg = 102kPa, pc = 280kPa and
deformations consistent with mechanical equilibrium state.
The boundary conditions were the following: for the lateral surface, qT = 0,
ql = 0, qg = 0, uh = 0; for the top surface, T = 298.15K, pg = 102kPa, pc = 420 kPa;
for the bottom surface, qT = 0, ql = 0, qg = 0, uν = 0.
Capillary pressure (Pa) Thermo-hydro-
1.00E+08
mechanical
9.00E+07 analysis
Key Time (s)
1.00E+02
8.00E+07
0.00E+00 131
7.00E+07 1.00E+03
2.00E+03
6.00E+07 3.00E+03
5.00E+03
5.00E+07
1.00E+04
4.00E+07 1.00E+05
1.00E+06
3.00E+07 1.00E+07
Figure 7.
2.00E+07 The actual solution in
the zone close to the
surface of the concrete
1.00E+07 wall – capillary
0.09 0.0925 0.095 0.0975 0.1 pressure
Distance X(m)

For numerical purposes the column of soil was simulated by 10 eight-node


isoparametric elements of different height (2 × 0.5cm, 7 × 1.0cm, 1 × 2cm).
Temporal discretization was performed with an initial step of 0.001 h during
first 100 steps and multiplied by 10 after repeating 90 steps until 1,000 hours. A
3 × 3 Gaussian integration scheme was applied.
The resulting profiles of temperature, saturation and water pressure are
compared in Figures 9, 11 and 13 with results of Schrefler at al.[8], (indicated by
dashed fine line). The slight differences between temperature profiles (Figure 9),
could be caused by different temporal discretization (much finer for our
solution) and latent heat transfer. The reason for the visible differences between
saturation profiles (Figure 11) as well as capillary pressure profiles ( Figure 13),
is the different shape of curves for relative permeabilities-capillary pressure and
saturation-capillary pressure relationships used in[5] (different parameters pb,
Sirr and λ for relations of Brooks and Corey[39]). The gas pressure profiles,
presented in Figure 15, show that gas pressure is practically constant in this
case. The resulting profiles of vertical displacement are presented in Figure 17.
The boundary conditions on the top surface of soil assumed in[11] are very
artificial from the physical point of view[15,16]. Therefore the same example
was solved for more correct, natural boundary conditions as follows: convective
heat transfer coefficient α c = 23.2 W m –2 K –1, ambient temperature T∞ =
EC Gas pressure (Pa)
1.90E+05
13,7
1.80E+05 Key
Key
1.00E+02 Time (s)
1.00E+02
1.70E+05 5.00E+02
5.00E+02
132 1.00E+03
1.60E+05 1.00E+03
2.00E+03
2.00E+03
1.50E+05 3.00E+03
3.00E+03
5.00E+03
5.00E+03
1.40E+05
1.00E+04
1.00E+04
1.30E+05 1.00E+05 1.00E+05
1.00E+06 1.00E+06
1.20E+05 1.00E+07 1.00E+07
1.10E+05
Figure 8.
The resulting profiles of
gas pressure in the 1.00E+05
concrete wall 0 0.02 0.04 0.06 0.08 0.1
Distance X(m)

298.15K, pg = 102kPa, convective mass transfer coefficient β c = 0.0209kg


m –3, relative humidity of surrounding air RH = 50 per cent. Besides, the
temperature dependence of capillary pressure-saturation relationship, as well

Temperature (K)
299
297
295
293
291
Figure 9. 289
The resulting
temperature profiles 287
during thermoelastic
285
consolidation of clay
solved for BC of the first 283
kind (solid line), 0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1
compared with the Height (m)
solution of Schrefler
et al. – case 1[5] (dashed Key
line) 1 hour 10 hours 100 hours 300 hours 1,000 hours
5 hours 50 hours 200 hours 500 hours
Temperature (K) Thermo-hydro-
299
mechanical
297 analysis
295
293
133
291
289
287
285 Figure 10.
283 The resulting
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1 temperature profiles
Height (m) during thermoelastic
consolidation of clay
Key solved for BC of the
1 hour 10 hours 100 hours 300 hours 1000 hours third kind
5 hours 50 hours 200 hours 500 hours

as dependence of thermal conductivity and capacity on saturation of water


were taken into account.
The same space discretization as before was used, while temporal
discretization was performed with an initial step of 0.01h during the first 1,000
steps, then 0.1h during the next 900 steps and 0.25h until 1,000 hours.
The resulting profiles of temperature, saturation, capillary pressure, gas
pressure and vertical displacement are presented in Figures 10, 12, 14, 16 and 18

Saturation (–)
0.45
0.44
0.43
0.42
0.41
0.40 Figure 11.
The resulting saturation
0.39 profiles during
thermoelastic
0.38
consolidation of clay
0.37 solved for BC of the first
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1 kind (solid line),
Height (m) compared with the
solution of Schrefler
Key et al. – case 1[5] (dashed
1 hour 10 hours 100 hours 300 hours 1000 hours line)
5 hours 50 hours 200 hours 500 hours
EC respectively, showing significant differences in comparison to the previous
13,7 solution (Figures 9, 11, 13, 15 and 17). The temperature gradients are much
smaller for this case, Figure 10, while the final value of surface temperature is
reached after about 1,000 hours (not immediately, as it was assumed before).
Saturation in the zone close to the upper surface increases slightly at the
beginning of the process analysed because of temperature increase (Figure 12).
134 Then it decreases, reaching its final value at the surface after about 50 hours
(not immediately, as it was assumed in the previous solution). The saturation
gradients in the zone close to the surface are significantly higher in comparison
to the previous solution (Figure 11). The capillary pressure profiles behave
accordingly (Figure 14), causing considerable displacements first of all in the
surface zone, (Figure 15), though its values are incomparable with the previous
solution because of different final value of capillary pressure at the surface. The
changes of gas pressure are bigger than for the previous solution, although they
still do not exceed 85Pa (see Figure 16).
It can be seen that the process analysed is strongly influenced by boundary
conditions. The behaviour of the soil for those more realistic and more correct
ones from the physical point of view is completely different.

Conclusions
A fully coupled model for simulating heat and mass transfer in deformable
porous materials regarding also phase change phenomena (evaporation,
condensation and latent heat transfer) has been presented. The model is based

Saturation (–)
0.48

0.46

0.44

0.42

0.40

0.38

0.36

0.34
Figure 12.
The resulting saturation 0.32
profiles during 0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1
thermoelastic Height (m)
consolidation of clay
solved for BC of the Key
third kind 1 hour 10 hours 100 hours 300 hours 1,000 hours
5 hours 50 hours 200 hours 500 hours
420
Capillary pressure (kPa) Thermo-hydro-
mechanical
400 analysis
380

360 135
340
Figure 13.
320 The resulting capillary
pressure profiles during
300 thermoelastic
consolidation of clay
solved for BC of the
280
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1 first kind (solid line),
Height (m) compared with the
solution of Schrefler
Key et al. – case 1[5] (dashed
1 hour 10 hours 100 hours 300 hours 1,000 hours line)
5 hours 50 hours 200 hours 500 hours

on a strong physical background that allows us clearly to identify the


constitutive equations and the other coefficients needed to characterize the
analysed medium. The model results in a set of non-linear and coupled partial
differential equations. They are discretized in space using a finite element
method and in time using finite differences and solved for gas pressure,

Capillary pressure (kPa)


100,000

10,000

1,000

1h-10h
Figure 14.
100 The resulting capillary
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1 pressure profiles during
Height (m) thermoelastic
consolidation of clay
Key solved for BC of the
50 hours 200 hours 500 hours third kind
100 hours 300 hours 1,000 hours
EC capillary pressure, temperature and displacements as primary variables.
13,7 Because of the non-linearity of the equations a Newton-Raphson approach is
used for solution. The validity of the approach has been demonstrated by the
good agreement between simulation results and experimental data of
Liakopoulos[40]. Further examples show both the robustness of the model in
dealing with various problems and the appreciable influence of proper physical
136 modelling (regarding phase changes, realistic boundary conditions) on the
evolution of the phenomena.

Gas pressure (Pa) Key


102,000.1 1 hour
5 hours
10 hours
102,000.0
50 hours
100 hours
Figure 15. 200 hours
The resulting gas 101,999.9
pressure profiles during 300 hours
thermoelastic 500 hours
consolidation of clay
solved for BC of the first 101,999.8 1,000 hours
kind 0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1
Height (m)

Gas pressure (Pa)


102,100 Key
1 hour
102,080
5 hours
102,060
10 hours
102,040 50 hours
102,020 100 hours
Figure 16. 102,000 200 hours
The resulting gas
pressure profiles during 101,980 300 hours
thermoelastic 500 hours
101,960
consolidation of clay
solved for BC of the 101,940 1,000 hours
third kind 0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1
Height (m)
0.00E+00
Vertical displacement (m)
Key Thermo-hydro-
1 hour mechanical
–1.00E–05 5 hours analysis
–2.00E–05 10 hours
50 hours
–3.00E–05
100 hours 137
–4.00E–05
200 hours Figure 17.
–5.00E–05 300 hours The resulting vertical
displacement profiles
–6.00E–05 500 hours during thermoelastic
consolidation of clay
1,000 hours
–7.00E–050 solved for BC of the first
0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1 kind
Height (m)

Vertical displacement (m)


0.00E+00 Key
1 hour
–5.00E–04 5 hours
10 hours
–1.00E–03
50 hours
–1.50E–03 100 hours
200 hours
–2.00E–03 Figure 18.
300 hours The resulting vertical
500 hours displacement profiles
–2.50E–03 during thermoelastic
1,000 hours consolidation of clay
–3.00E–03 solved for BC of the
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1 third kind
Height (m)

References
1. Lewis, R.W. and Schrefler, B.A., The Finite Element Method in the Static and Dynamic
Deformation and Consolidation of Porous Media, John Wiley, New York, NY (to appear).
2. Zienkiewicz, O.C., Xie, Y.M., Schrefler, B.A., Ledesma, A. and Bicanic N., “Static and
dynamic behaviour of soils: a rational approach to quantitative solutions, II, semi-
saturated problems”, Proceedings of the Royal Society of London., Vol. A 429, 1990,
pp. 311-21.
3. Coussy, O., Mécanique des Milieux Poreux, Editions Technip, Paris, 1991.
4. de Boer, R., Ehlers, W., Kowalski, S. and Plischka J., Porous Media, a Survey of Different
Approaches, Fachbereich Bauwesen der Universitaet Gesamthochschule – Essen, Heft 54,
1991.
EC 5. Schrefler, B.A., Zhan Xiaoyong and Simoni, L., “A coupled model for water flow, airflow
and heat flow in deformable porous media”, International Journal for Numerical Methods
13,7 in Heat Fluid Flow, Vol. 5, 1995, pp. 531-47.
6. Philip, J.R. and de Vries, D.A., “Moisture movements in porous material under temperature
gradients”, Transactions of American Geophysics Union, Vol. 38 No. 2, 1957, pp. 222-32.
7. de Vries, D.A., “Simultaneous transfer of heat and moisture in porous media”, Transactions
of American Geophysics Union, Vol. 39 No. 5, 1958, pp. 909-16.
138 8. Luikov, A.V., “Systems of differential equations of heat and mass transfer in capillary-
porous bodies (review)”, International Journal of Heat Mass Transfer, Vol. 18, 1975,
pp. 1-14 .
9. Whitaker, S., “Simultaneous heat, mass and momentum transfer in porous media: a theory
of drying”, Advances in Heat Transfer, Vol. 13, 1977.
10. Whitaker, S., “Heat and mass transfer in granular porous media”, Advances in Drying,
Vol. 1, 1980.
11. Dakshanamurthy, V. and Frelund, D.G., “A mathematical model for predicting moisture
flow in an unsaturated soil under hydraulic and temperature gradients”, Water Resources
Research, Vol. 17, 1981, pp. 714-22.
12. Thomas, H.R., “Modelling two-dimensional heat and moisture transfer in unsaturated
soils, including gravity effects”, International Journal for Numerical and Analytical
Methods in Geomechanics, Vol. 9, 1985, pp. 573-88.
13. Thomas, H.R. and King, S.D., “Coupled temperature/capillary potential variation in
unsaturated soils”, Journal of Engineering Mechanics, Vol. 117 No. 11, 1991, pp. 2475-90.
14. Baggio, P., Bonacina, C. and Strada, M., “Trasporto di calore e di massa nel calcestruzzo
celulare” (in Italian), La Termotecnica, Vol. 45 No. 12, 1993, pp. 53-60.
15. Gawin, D. and Klemm, P., “A model of coupled heat and moisture transfer with phase
changes in porous building materials”, Archives of Civil Engineering, Vol. 40 No. 1, 1994,
pp. 89-104.
16. Krischer, O. and Kroell, K., Trocknungstechnik, Band1, 3 Aufl. Die wissenschaftlichen
Grundlagen der Trocknungstechnik, Springer Verlag, Berlin-Goettingen-Heidelberg, 1978.
17. Gawin, D., Baggio, P. and Schrefler, B.A., “Coupled heat, water and gas flow in deformable
porous media”, International Journal for Numerical Methods in Fluids, Vol. 20, 1995,
pp. 969-87.
18. Hassanizadeh, M. and Gray, W.G., “General conservation equations for multiphase
systems: 1. Averaging technique”, Advances in Water Research, Vol. 2, 1979, pp. 131-44.
19. Hassanizadeh, M. and Gray, W.G., “General conservation equations for multiphase
systems: 2. Mass, momenta, energy and entropy equations”, Advances in Water Research,
Vol. 2, 1979, pp. 191-203.
20. Hassanizadeh, M. and Gray, W.G., “General conservation equations for multiphase
systems: 3. Constitutive theory for porous media flow”, Advances in Water Research,
Vol. 3, 1980, pp. 25-40.
21. Bear, J., Hydraulics of Groundwater, McGraw Hill, New York, NY, 1979.
22. Bear, J., Dynamics of Fluids in Porous Media, Dover, New York, NY, 1988.
23. Bachmat, Y. and Bear, J., “Macroscopic modelling of transport phenomena in porous media.
1: The continuum approach”, Journal of Transport in Porous Media, Vol. 1, 1986, pp. 213-40.
24. Bachmat, Y. and Bear J., “Macroscopic modelling of transport phenomena in porous media.
2: Applications to mass, momentum and energy transfer”, Journal of Transport in Porous
Media, Vol. 1, 1986, pp. 241-69.
25. Bear, J. and Bachmat, Y., Introduction to Modeling of Transport Phenomena in Porous
Media, Kluwer, Dordrecht, 1990.
26. Gray, W.G. and Hassanizadeh, S.M., “Unsaturated flow theory including interfacial Thermo-hydro-
phenomena”, Water Resources Research, Vol. 27 No. 8, 1991, pp. 1855-63.
27. ASHRAE Handbook, Fundamentals Volume, ASHRAE, Atlanta, 1993.
mechanical
28. Biot, M.A. and Willis, P.G., “The elastic coefficients of the theory of consolidation”, Journal analysis
of Applied Mechanics, Vol. 24, 1957, pp. 594-601.
29. Skempton, A.W., “Effective stress in soils, concrete and rocks”, Pore Pressure and Suction
in Soil, Gutterworth, Stoneham, MA, 1961, pp. 4-16. 139
30. Schrefler, B.A. and Zhan Xiaoyong, “A fully coupled model for water flow and airflow in
deformable porous media”, Water Resources Research, Vol. 29, 1993, pp. 155-67.
31. Baggio, P. and Bonacina, C., “Introduction to modeling heat and mass transfer in porous
building materials”, Quaderno di Istituto di Fisica Tecnica dell’ Università di Padova,
Vol. 146 (in print).
32. Schrefler, B.A., Simoni, L., Li, X. and Zienkiewicz O.C., “Mechanics of partially saturated
porous media”, in Desai, C.S. and Gioda, G. (Eds), Numerical Methods and Constitutive
Modelling in Geomechanics, CISM Courses and Lectures, No. 311, Springer-Verlag, New
York and Berlin, 1990, pp. 169-209.
33. Moran, M.J. and Shapiro, H.N., Fundamentals of Engineering Thermodynamics, 2nd ed.,
John Wiley & Sons, Chichester, 1993.
34. Zienkiewicz, O.C. and Taylor, R.L., The Finite Element Method, Vol. 1, 4th ed., McGraw Hill,
London, 1989.
35. Zienkiewicz, O.C. and Taylor, R.L., The Finite Element Method, Vol. 2, 4th ed., McGraw Hill,
London,1991.
36. Turska, E. and Schrefler, B.A., “On convergence conditions of partitioned solution
procedures for consolidation problems”, Computational Methods in Applied Mechanics
and Engineering, Vol. 106, 1993, pp. 51-63.
37. Marchuk, G.I., Methods of Numerical Analysis, Springer Verlag, New York, 1975.
38. Forsyth, P.A. and Simpson, R.B., “A two-phase, two-component model for natural
convection in a porous medium”, International Journal of Numerical Methods in Fluids,
Vol. 12, 1991, pp. 655-82.
39. Brooks, R.N. and Corey, A.T., “Properties of porous media affecting fluid flow”, Journal of
Irrigation Drain, Division of the American Society of Civil Engineering, Vol. 92 No. IR2,
1966, pp. 61-8.
40. Liakopoulos, A.C., “Transient flow through unsaturated porous media”, PhD thesis,
University of California, Berkeley, CA, 1965.
41. Narasimhan, T.N. and Witherspoon, P.A., “Numerical model for saturated-unsaturated
flow in deformable porous media. 3. Applications”, Water Resource Research, Vol. 14, 1978,
pp. 1017-34.
42. Schrefler, B.A. and Simoni L., “A unified approach to the analysis of saturated-unsaturated
elastoplastic porous media”, in Svoboda, G. (Ed.), Numerical Methods in Geomechanics,
A.A. Balkema, Rotterdam, 1988, pp. 205-12.
43. Lassabatere, T. and Coussy, O., “Retrait de dessiccation et fluage induit dans les milieux
poreux”, Actes du XI congrès francais de mécanique, Lille-Villeneuve d’Ascq, Vol. 3, 1993,
pp. 273-6.
44. Lassabatere, T., “Thermodynamics of unsaturated porous media: a micro-macro
approach”, Proceedings of First Forum of Young European Researchers, Liège, 1993,
pp. 97-101.
45. Lassabatere, T. and Coussy, O., “Retrait de dessiccation: un modèle continu poromécanique,
à paraitre” in Les Actes des Jounées des Sciences de l’ingénieur, Presqu’ile de Giens, 1994.
EC 46. Safai, N.M. and Pinder, G.F., “Vertical and horizontal land deformation in a desaturating
porous medium”, Advances in Water Resources, Vol. 2, 1979, pp. 19-25.
13,7
47. Nasrallah, S.B. and Perre, P., “Detailed study of a model of heat and mass transfer during
convective drying of porous media”, International Journal of Heat Mass Transfer,
Vol. 31 No. 5, 1988, pp. 957-67.
48. Lambe, T.W. and Whitman, R.V., Soil Mechanics, John Wiley & Sons, Chichester, 1969.
140
Appendix
The governing equations (23) and (27) in the discretized form are shown here in detail, using the
notation of Lewis and Schrefler[1]:
Thermo-hydro-
mechanical
analysis

141

Nomenclature
B strain matrix relating strain and displacement
b specific body force[m s–2 ]
Cp effective specific heat of the porous medium[J kg–1 K–1]
Cpg specific heat of the gas mixture[J kg–1 K–1]
EC Cps specific heat of the solid matrix[J kg–1 K–1]
13,7 Cpl specific heat of the liquid phase (water)[J kg–1 K–1]
D tangent matrix[Pa]
Deff effective diffusivity of the gas mixture[m2 s–1]
E Young modulus[Pa]
I unit tensor
K intrinsic permeability tensor[m2]
142 KS bulk modulus of the solid phase[Pa]
KT bulk modulus of the porous medium[Pa]
Krg relative permeability of the gas phase[–]
Krl relative permeability of the liquid phase[–]
k hydraulic conductivity[m s–1]
M molar mass of the gas mixture (moist air)[kg kmol–1]
Ma molar mass of the dry air[kg kmol–1]
Mw molar mass of the water vapour[kg kmol–1]
n unit normal vector
p average pressure of the mixture[Pa]
patm atmospheric pressure[Pa]
pb bubbling pressure[Pa]
pc capillary pressure[Pa]
pcr critical value of capillary pressure[Pa]
pg pressure of the gas phase[Pa]
pgw water vapour partial pressure[Pa]
pgws water vapour saturation pressure[Pa]
pl liquid water pressure[Pa]
qga dry air flux imposed on the boundary
qgw vapour flux imposed on the boundary
ql liquid flux imposed on the boundary
qT heat flux imposed on the boundary
R gas constant[8314.41 J kmol–1 K–1 ]
S liquid phase volumic saturation[liquid Vol./pore Vol. ]
T temperature[K]
T∞ temperature in the undisturbed gas phase distant from the interface[K]
t time[s]
t traction imposed on the boundary[Pa]
u displacement vector of the solid matrix[m]
uh horizontal displacement[m]
uν vertical displacement[m]
vg velocity of the gaseous phase[m s–1]
vl velocity of the liquid phase[m s–1]
vdga relative average diffusion velocity of dry air species[m s–1]
vdgw relative average diffusion velocity of water vapour species[m s–1]

Greek symbols
α Biot’s constant
αc convective heat transfer coefficient[W m–2 K–1]
βc convective mass transfer coefficient[m s–1]
βs cubic thermal expansion coefficient of the solid[K–1]
∆hvap specific enthalpy of evaporation[J kg–1]
∆t time step[s]
ε elastic strain
εT thermoelastic strain
εν volumetric strain
Γ1i part of the boundary Γ, where BC of the first kind are applied for the i-variable (the Thermo-hydro-
imposed value of i-variable) mechanical
Γ2i part of the boundary Γ, where BC of the second kind are applied for i-variable (the imposed
value of flux corresponding to the i-variable) analysis
Γ3i part of the boundary Γ, where BC of the third kind are applied for the i-variable (the
imposed value of the i-variable and of flux corresponding to it)
λeff effective thermal conductivity[W m–1 K–1]
µg gas phase dynamic viscosity[Pa s] 143
µl liquid phase dynamic viscosity[Pa s]
ν Poisson’s ratio[–]
ρ effective density of the porous medium[kg m–3]
ρg gas phase density[kg m–3]
ρ ga mass concentration of dry air in the gas phase[kg m–3]
ρ gw mass concentration of water vapour in the gas phase[kg m–3]
ρ gw∞ mass concentration of water vapour in the undisturbed gas phase distant from the
interface[kg m–3]
ρl liquid phase density[kg m–3]
σ total stress tensor[Pa]
σ" effective stress tensor[Pa]
φ porosity[pore Vol./total Vol.]

Вам также может понравиться