You are on page 1of 24

Transport in Porous Media 1, (1986), 155-178. 9 1986 by D. Reidel Publishing Company.

155

A Coupled Finite Element Model for the Consolidation of Nonisothermal Elastoplastic Porous Media
R. W. L E W I S

Institute of Numerical Methods in Engineering, University College of Swansea, Swansea SA2 8PP, Wales, U.K.
C. E. M A J O R A N A and B. A. S C H R E F L E R

Istituto di Costruzioni, University of Padua, Italy.


(Received: 21 August 1985; revised: 18 December 1985)
Abstract. A coupled finite element model for the analysis of the deformation of elastoplastic porous media due to fluid and heat flow is presented. A displacement-pressure temperature formulation is used for this purpose. This formulation results in an unsymmetric coefficient matrix, even in the case of associated plasticity. A partitioned solution procedure is applied to restore the symmetry of the coefficient matrix. The partitioning procedure is an algebraic one which is carried out after integration in the time domain. For this integration, a two-point recurrence scheme is used. The finite element model is applied to the investigation of nonisothermal consolidation in various situations.

Key words. Three-field interaction problem; temperature field, flow field, displacement field; geothermal systems, porous media, elastoplastic analysis, partitioned solution procedure, fully coupled finite element models, nonisothermal consolidation.

1. Introduction
A numerical model is presented in this paper which fully takes into account the interaction between the stress, flow, and temperature fields in saturated elastoplastic porous media. This model allows for the investigation of land subsidence in connection with geothermal energy production for a given geothermal system as well as for the estimation of the quantity of recoverable energy and the optimum extraction rate of mass and energy from such a system. Analyses of this type can also be applied to the design of thermal and/or hydraulic fracturing stimulation of oil reservoirs and for more accurate interpretation of well tests when thermal effects are taken into account [1]. Another important area of application regards the study of the effects of radioactive waste disposal in clay layers on the ocean bed [2]. The decay of radioactive material produces heat which causes a temperature rise and an expansion in both the pore

156

R.W. LEWIS,C. E. MAJORANA,AND B. A. SCHREFLER

water and the soil skeleton. The difference of the volume increase of pore water and voids results in the consolidation of the clay beds. If the effective stress reduction is too great, a progressive failure may occur which has to be avoided, since the clay is the main long-term barrier against the diffusion of radionuclides [2]. Hence, a tool for predictive analyses is important in this case. Since failure is involved, such a tool requires the capability of elastoplastic analysis. The effects of heat and fluid in soil have been investigated mainly in connection with geothermal reservoir analysis. The study of geothermal reservoirs has recently received much attention but the deformation characteristics have been considered only in a small number of papers, especially when coupled with heat and mass transfer equations. The governing equations in earlier studies on this subject are reviewed and thoroughly discussed by Corapcioglu and Karahanoglu [3]. Some of the more recent studies, which are relevant to the model presented in this paper are now briefly discussed. The effect of geothermal production, as well as reinjection, on the deformation of geothermal systems has been dealt with by Lippmann et al. [4], where the numerical model for the mass and energy equations are combined with the numerical solution of Terzaghi's consolidation equation. The reservoir temperature and pore pressure are solved by interlacing equations for mass and energy and the pore pressure changes are then used to determine volumetric and vertical deformation. Brownell et al. [5] discussed the full interaction of a porous solid matrix and fluid in geothermal systems, including momentum and energy transfer and the dependence of porosity and permeability upon fluid and solid stresses. The ground-surface subsidence of Wairakei has been examined in terms of a calculated two-phase fluid flow and the local geology by Pritchett et al. [6]. A similar study for a geopressured reservoir is presented by Garg et al. [7], considering dissolved methane in the fluid. Aktan and Farouq Ali [8] study the thermal stress induced by hot water injection using thermoelastic stress-strain relationships. Similarly, Ertekin [9] has presented a two-dimensional two-phase fluid flow, a three-dimensional heat flow, and a two-dimensional displacement model for a hot water flooded oil reservoir. He used the method of finite differences to solve the fluid and energy flow equations and then applied a finite element model to determine the displacements. Bear and Corapcioglu [10] developed a mathematical model to simulate the areal distribution of fluid pressure, temperature, land subsidence, and horizontal displacements, due to hot water injection into a thermoelastic confined and leaky aquifer. Their mathematical model is derived by averaging a three-dimensional model over the vertical thickness of the aquifer under the assumption that the aquifer is thin in relation to the horizontal distances. Borsetto et al. [1] presented a numerical model for the heat and mass transfer

CONSOLIDATION OF NONISOTHERMAL ELASTOPLASTIC POROUS MEDIA

157

and stress field in transient coupled conditions using an elastic stress-strain relationship. The governing equations are partitioned prior to the discretization in the time domain and, for their solution, require a predictor for the displacements, pressure, and temperature. A different partitioning procedure for the governing equations used in this paper was presented by Schrefler [11]. In connection with the analysis of nuclear waste disposal in geological clay formation, Borsetto et al. [12] discuss the constitutive relationship for clay under the combined action of heating, elastoplastic deformation, and ground-water flow. The dependence of the coefficient of permeability of the temperature is also included. The resulting governing equations can be solved in principle to obtain displacements, pressure, temperature, and porosity. Since the solution is expected to require a considerable numerical effort, Borsetto et al. propose simplifications such as the uncoupling of the heat flow equation by reducing the number of independent unknowns to be solved simultaneously. Thermomechanically-induced creep in clays in situations where elasticity of the medium only slightly affects the flow field was studied by Dawson and Chavez [13] using a coupled creeping-flow model for saturated porous media. Booker and Savvidou [2] studied the consolidation around a heat source in thermoelastic soil. The temperature field is uncoupled from the determination of displacements and pressure by neglecting the mechanical contributions to energy balance and the convective terms. An analytical solution is given for a spherical heat source and a point heat source buried deep in clay. The point source solution is then integrated over the volume of a cylindrical canister. Thermoelastic consolidation is also investigated by Aboustit et al. [14] using a general variational principle. The fluid is assumed as incompressible and convection is neglected. The coupling terms between pressure and temperature do not appear in their formulation. Only the couplings between displacement and temperature and displacement and pressure are taken into account. This results in a symmetric coefficient matrix. Numerical results of a case of one-dimensional thermoelastic consolidation and a field application for the Centralia coal gasification site are presented. In view of the aforementioned studies, a finite element model is formulated in the next section for the simulation of heat and fluid flow in deforming porous media, using elastoplastic stress-strain relationships. In the formulation, the assumption of thermal equilibrium between the solid and fluid phases is taken into account and cross-transport phenomena, such as the Soret or Dufour effects are neglected, as well as thermal dispersion [15, 16]. Thermal dispersion may be introduced into the model if required by a particular problem by means of a modified conductivity matrix. Part of this research was presented in a paper by Lewis and Kaharanoglu [17] to an international conference on numerical methods in thermal problems in Venice, Italy.

158

R . w . LEWIS, C. E. MAJORANA, AND B. A. SCHREFLER

2. Governing

Equations

Nonisothermal porous media are modelled by utilizing the equilibrium equation together with the governing equations of heat and fluid flow through such a porous medium. The isothermal equilibrium and fluid flow equations are well known. For their complete derivation, the interested reader is referred to [18-20] for the case of saturated one-phase flow, two-phase flow, and the general case of multiphase-flow, respectively. The modifications due to the introduction of heat are shown below for the case of a single liquid saturating the porous medium. These modifications are, in principle, also applicable to multiphase flow, where the same argument applies to the mass continuity equation of each fluid phase.

2.1. NONISOTHERMAL SOLID PHASE BEHAVIOUR

As in the isothermal case, the total stress vector o. (with tension positive) is decomposed into a hydrostatic component mp and an effective stress o.'
cr = o.' - m p (1)

in which m = [1, 1, 1, 0, 0, 0] r and p denotes the pore pressure (compression positive). The constitutive equation relating the effective stresses to the strains of the skeleton may be written for a general nonlinear material as do'' = I1T (de - dec - dev - deT - deo) where de represents the total strain of the skeleton, d~c = c d t is the creep strain,
dep =

(2)

(2a)

-m(dp/3 K~)

(2b)

represents the overall volumetric strains caused by uniform compression of the particles by the pore fluid, with Ks being the bulk modulus of the solid phase, deT = m -~ d T (2c)

is the reversible part of the strain increment due to temperature increment d T;/3s is the thermal expansion coefficient of the solid phase and deo represents all other strain increments not directly associated with stress changes (swelling, chemical, the irreversible part of the thermal strains, etc.), i.e., the autogeneous strains. The reversible part of the thermal strains is separated from the irreversible part because, actually, only a few indications exist about the temperature-dependence of yield surfaces and hardening parameters of soils. This temperature-depen-

CONSOLIDATION OF NONISOTHERMALELASTOPLASTIC POROUS MEDIA

159

dence is, hence, neglected in what follows. The tangent matrix DT and the creep function c are dependent on the level of effective stress tr' and also, if strain effects are considered, on the total strain of the skeleton. The temperature dependence of DT and c can easily be introduced in the model when warranted by sufficient information. The incremental equilibrium equation in the total stress state is written as

Ia~5~Tdtrdf~-fn6uTdbdf~-Ir

di dF= 0

(3)

by invoking the principle of virtual work [21]. Here u is the displacement vector and db and di stand for any changes of external forces due to body and boundary force loadings, specified in the domain ~ and at its boundary F, respectively. Incorporating the effective stress relationship (1) and the constitutive relationship (2) into Equation (3), the following equation is obtained

ot ~ Tn Op fn6~TDr~dl~--In6~Tm~td~--Iao~ l/Tm~ a --fa6~TDTcdf~--fa6eTDTmlS"OTdl)--fa~%TDTOe~


where

1 ~f~ --

Ot

Ot

(4)

^ (~ 0t = I~ (~uT 0b 0t d~ -t- Ir ~I[IT 0t dl~* (4a)

2.2. FLOW EQUATION FOR A SINGLE PHASE IN A DEFORMING NONISOTHERMAL POROUS MEDIUM The equation governing the flow of a single fluid phase through a deforming isothermal porous medium is defined by

VrKV(p+ ywZ)+m /T _ m T D T ] ~
+ [ 1 - th_~ th Ks Kw where K Kw ~b yw z Qe is is is is is is the the the the the the

Ok
3K~ / 0t

mTDTc 3K~ (5)

1 mTDTm]OP+OOe=o, (3Ks) 2 J Ot Ot

permeability matrix, bulk modulus, porosity, specific weight of the fluid, elevation above some datum, volumetric outflow of the fluid per unit volume of the solid.

Three terms which contribute to the rate of fluid accumulation [19] in Equation (5), are now modified for the nonisothermal case.

160

R.W. LEWIS, C. E. MAJORANA, AND B. A. SCHREFLER

The fluid density pw and the solid density ps are a function of pressure and temperature. The temperature-dependence of the fluid density has to be taken into account in the rate of changes of fluid density, expressed for a single fluid phase (water) and for the isothermal case by, [20] 4)

10pw 1 apwOp r - 4) pw Ot pw Op Ot Kw Ot"

(6)

Equation (6) here becomes

10pw ChOw Ot

4) 10pw I O _ _ p _ + p 4) 10pw OT E -~p I T Ot E-O-f- . - ~

(7)

where, by definition

OPwI _ Ow and OPwI =-pwflw. Op r Kw O-T p

(8)

Here, flw is the thermal expansion coefficient of the fluid. The dependence of the solid density upon temperature can be taken into account in the following way. The rate of change of grain volume due to the pressure changes, given by

1 - 4) Op K~ Ot'

(9)

expresses the dependence of the solid density upon pressure changes. Taking into account the definition of the rock matrix compressibility lIKe, which is equal to the unit change in rock matrix density per unit change in pore pressure, Equation (9) can be written as

1 - 4) Op Ks Ot

(1 - 4 ) ) - -

1 002 ps Ot"

(10)

Since Ps = P~ (P, T), Equation (10) becomes

(1-4))oT\OpITOt
where, by definition,

OT pO-t

(11)

Ops I 1 ON T = P s E

and

Opt = OZ p -Ps~s.

(12)

The rate of change of grain volume due to pressure and temperature changes may, hence, be written as

1 - 4) Op (1 - 4))/3s O_T_T. Ks Ot Ot

(13)

CONSOLIDATION OF NONISOTHERMAL ELASTOPLASTIC POROUS MEDIA

161

If the temperature-dependence of the solid density is neglected, the second term of Equation (13) disappears. One further factor contributing to the rate of fluid accumulation needs a modification. The change of grain size due to effective stress changes, taking into account the constitutive relationship (2) becomes 1 mrDT(Oe_f m Op -3Ks \Ot 3K~Ot cm tSsOT~ ~--~/.

(14)

The sum of the factors (13) and (14) represent the contribution of the volumetric compression of the solid phase to the fluid accumulation. Finally, the fluid density also appears in the first term of the flow equation, together with the dynamic viscosity. In fact, the permeability matrix K may be expressed as

K = kpw_____gg,
IX

(15)

where k is the absolute permeability matrix and IX is the dynamic viscosity. Information about p~ = pw (p, T) and IX = Ix(p, T ) = Ix(T) can be found in [10]. Both fluid density and viscosity are assumed not to change with space. Therefore, continuity Equation (5), taking into account Equations (7), (9), (11) and (14), may be written as

+ q

(
__mTDT m

mrDTe

3K~

Ks

Kw

(3Ks) 2

Op+
0t (16)

. 1 r~ r , OQ~ - ~b/3w-(1-~b)/3s +--m3Ks t, Tm--3]-- Ot =0.

Approximations may be introduced into Equation (16). By neglecting for instance creep, the volume changes of the solid phase due to changes in stress and the volume changes in the pore water pressure, Equation (16) reduces to
_V T k V(p+pwgz)

{;

} + m oOE--~)[~wOT OZ'-l-O(~e=o t -OT-(1-~b)/3" Ot Ot

(17)

This form of the continuity equation was used by Booker and Savvidou [2].

2.3. ENERGY TRANSPORT T H R O U G H A DEFORMING POROUS MEDIUM

Again, as in the previous section, the equation is developed for the case of a single fluid phase saturating the porous medium. Both the solid and the fluid are

162

R.W. LEWIS, C. E. MAJORANA, AND B. A. SCHREFLER

assumed to be at the same temperature at coincident points. Furthermore, it is assumed that no chemical reactions occur that produce heat, or change relative amounts of the two constituents. The energy balance equation states that the rate of inflow into a control volume will be balanced by the increase in the internal energy

pwCw(V r T ) V " - V T H -F OOh _ 0 [[(1 - dO)OsC~+ ChpwCw]T} Ot Ot


where H is the conduction heat flux vector, osC~ is the heat capacity of the solid phase, OwC~ is the heat capacity of the fluid phase, V~ is an apparent velocity of the fluid, T is the increase in temperature over an initial equilibrium state, Oh is the outflow of heat per unit volume of solid.

(18)

The conductive flow of heat in the porous medium is governed by Fourier's law
H = -kV T, (19)

where k is the thermal conductivity matrix of the soil. By combining Equation (18) with Equation (19) the partial differential equation governing the transport of energy through a porous medium is obtained as [22] - V r k V T + O{[(1 - ch)psCs + (ppwC~]T}- pwC~(V

Tz)va-OOhot =0
(20)

Equation (20) is rewritten, considering pressure and temperature dependent fluid and solid densities and keeping the other parameters constant.
-

vTkVT+(1-r162 OT OQh = 0. + [(1 -- d~)psG + 4~p.Cw] _~ - pwC.(V m r ) v " -at at


(21)

Substituting, from Equations (7), (8), (11) and (12), for the appropriate terms in Equation (21) produces (l-q0 ~ k~-7-p~/3~ at/ (~ Ot
. . . .

+ [(1 - c~)p~C~+ $p~C~] O__T_T_ V TkV T - p~C~(V r T ) V Ot

0Qh --0. Ot

(22) The convective heat transport term in Equation (22) is a pressure and temperature combined term (via the fluid velocity Va).

CONSOLIDATION OF NONISOTHERMALELASTOPLASTIC POROUSMEDIA

163

Again, in many cases a simplification of Equation (22) can be justified. In particular by neglecting the temperature and pressure dependence of the fluid and solid densities and by neglecting the convective term, the energy transport equation becomes [(1 - 4~)p~C~+ 4pwCw] a T _ V r k V T 0 Q ~ = O.
ot ot

(23)

This uncoupled form of the energy transport equation was used by Booker and Savvidou [2]. In the case of the analysis of a heat source buried in clay, this uncoupling seems justified everywhere in the analysed volume except very close to the heat source where large hydraulic gradients could be generated [11]. The extension of the energy transport equation to multiphase flow can be carried out by again assuming that the temperature is the same for the fluids and the solid. The individual equations are combined into a single equation for the porous medium as a whole [2].

2.4. BOUNDARYCONDITIONS Together with the above equations, the model satisfies the following boundary conditions (a) Equilibrium of boundary stresses and external loads, already incorporated in the equilibrium equation (4). (b) Prescribed displacements,
u = ub (24)

(c) Continuity of the fluid flow across the boundary, -nTKV (P+ z)-~t~ = 0 (25)

where n is the unit normal and Oqo[Ot is the outflow rate per unit area of the boundary surface. (d) Prescribed pore pressures, p = pb (26)

(e) The continuity of convective and conductive heat flow across the boundary
nTpwCwV"T - nTkV T - Oq--2~= 0 Ot

(27)

where Oqn[Ot is the heat outflow rate per unit area of the boundary surface. (f) Prescribed temperature
T = T b.

(28)

164
3. Finite Element

R.W. LEWIS, C. E. MAJORANA, AND B. A. SCHREFLER


Formulation

Application of the standard finite element discretization to the governing equations (4), (16), (22) and boundary equations (24)-(28) and approximating the variation of displacement u, pore pressure p and temperature T by the nodal values and shape functions as [21] u = N~, p = Np, T = N2" (29)

will form the following system of coupled equations K aft + L ~__~t + T U d"[' Of dt -d-t-: c + ~ ' T d~ d P + T p dT = O F + G ' L -~- + Up + S ~-~ dt 0 t di dT OTG TL -d~ + TRI" + TS d--t : 0---t--

(30)
(31)

(32)

with the matrices K . . . . . T G given in Appendix A, and where B is given by = B~. Equations (30)-(32) are more general than the governing equations presented in [17]. The above system of coupled equations is now written in a more concise form for further developments.
- d~ -

(33)

where g = S TL TU TP TS

(34a)

(2 =

[i~
H
O

(34b) TR

o~
and

%]

(34c)

tit

(34d)

CONSOLIDATION OF NONISOTHERMAL ELASTOPLASTIC POROUS MEDIA

165

As can be easily recognized, the coefficient matrix B is no longer symmetric, even if DT is symmetric. The problem is no longer linear, even in the case of linear constitutive relationship and constant material parameters, due to the presence of the temperatures in T L and TS. A linearization is possible by means of the simplifications indicated above. Integration of Equation (32) in the time domain by means of a two point recurrence scheme produces the 'monolithic' system of equation

= [B -

Atk(1

--

a)~S]k,~itk +

Fk,,Atk

(35)

where
O/ - -

t-tk Ark'

Atk = tk+1 --

tk .

3.1. PARTITIONED SOLUTION PROCEDURE

Symmetry can be restored in equation (35) by means of a partitioning procedure shown by Schrefler [11]. This procedure results in the following system of equation,

= [ B - Atk(1 -- a)(2]~'~ -t13i--"E--Xptk+Atk+ ~Atk where the indices k, a have been omitted,

(36)

BZ =

S
0 TS

(37a)

TU BE = 0 TL and xp-'k+a'~is a suitable predictor of the vector 2. Such a predictor may be expressed as [1] Xp-'~+a'k= (1 - 3,)~'k + 7i'~+A'k (38) (37b)

where 0 ~< 3' ~< 1. Usually 3" = a. If y = 0, the predictor is a one-term extrapolator, given by the last solution. Unfortunately in this case, the last solution predictor produces uncoupling between the temperature field and the displacement and pressure fields. For 3, ~ 0 an iterative solution technique is needed [23].

166

R. W. LEWIS, C. E. MAJORANA, AND B. A. SCHREFLER

To avoid this, the following techniques may be used: (a) The predictor (38) is substituted by Xp-'~+A'~= (1 + 7)i '~ -- 3'i '~-a'~ (39)

with the correction for the time step length changes of the type 3' --- 3,(Atg/Atk_l). (b) Equation (36) is solved by taking i'~ = ~t~-a,~, ~'~+atk = i'~ (40a) (40b)

For the solution of equation (36) a staggered scheme is the most appropriate one, where first 1"t~+a~ is predicted to solve a modified consolidation analyser

+ oF+0t

Ztk_ JTU/T,~+~,~[TpI p

(41)

[0t
Then the temperature field is solved to obtain "l't~+a'~ [TS + aAtkTR]T ~+~'~

= [TL T S - (1 - atAtkTR]

+TGAt k - TL~;~+~'..

(42)

The next time integration step proceeds with Equation (41). A stability analysis of the partitioned system (36) was carried out [11]. For the linear case and a predictor given by Equation (38) with 3' = a, the stability of the solution requires that [2B~+ 2BE+ A t ( 2 a - 1)1~] is positive definite and (; nonnegative definite. Since (2 is nonnegative definite, B~ is positive definite and B~ is nonnegative definite, stability can easily be ensured by setting c~>i 89 This is the same requirement as for the monolithic system of Equation (35). A last term extrapolator may be used if a three-level scheme is adopted. For instance, Lees' algorithm together with the same partitioning as above results in

- ~ A t k C i '~ +[B-~AtkC],~i '~-a'~ BEi'~+a'~ +2AtkF,~.

(43)

The respective merits of the different solution procedures outlined in this section will be thoroughly investigated in another paper [25]. The monolithic system and a staggered solution based on a two level scheme are used for the examples in the following section.

CONSOLIDATION OF NONISOTHERMAL ELASTOPLASTIC POROUS MEDIA

167

4. Numerical Examples
The following examples are used to validate the finite element programme based on the approach presented in the previous sections and to show its possibilities. E X A M P L E 1. Therrno-elastic consolidation. The first example is that of a thermo-elastic 1-D consolidation solved previously by Aboustit et al. [14]. A column of linear elastic material is subjected to a unit surface pressure and a constant surface temperature T = 50. The following data were assumed: L = 7, E = 6000, v = 0.4, K = 4 x 10 -6, A ~ - 0.2, (pc), = 40, /3, = 0.9 x 10 -6. The coupling matrix T P of Equation (31) is equal to zero. This matrix is not taken into account in [14]. In the present solution the matrix T U 9 ~'o retained by Aboustit et al. in the energy transport equation, is not included. Its contribution is, in fact, negligible as will be seen in Figure 6 by comparison between the results of [14] and our results. T h e finite-element discretization used is shown in Figure 1. The
P=I

II

2
9

permeable

AT=50
i

2"7

impermeable

zI
y~
L.

~/

Fig. 1. 1-D model problem and finite element mesh.

168

R.W. LEWIS, C. E. MAJORANA, AND B. A. SCHREFLER

p o r e p r e s s u r e is e q u a l to zero at the top surface; e v e r y w h e r e else the surface of the b o d y is sealed a n d insulated. T h e time step v a l u e s used are i n d i c a t e d in T a b l e

I, (scheme I).
Table I. Time stepping (scheme I) Time interval 0.01 0.1 10 100 1000 Number of time steps 10 10 10 10 20

For comparison also a more refined mesh in the time domain was used, (scheme II). This second scheme assumes 10 x 1 steps after the 10 0.1 steps of scheme I. Scheme II resulted in slight differences during only a few time steps after the change from At = 0.1 to 1. It was, hence, not further applied. T h e values of a used in the two-level time stepping scheme are 0.5 and 0.875 (the second one according to [14]). Spatial oscillations of the temperature distribution was observed with a = 0.5 and a 2 x 2 spatial integration scheme, after suddenly applying the prescribed temperature. These oscillations disappeared after a few time steps and were not observed using a 3 x 3 Gaussian integration scheme. The temporal pore pressure oscillations vanished with a = 0.875. T h e errors associated with a sudden change of At died out in a few time steps. T h e temperatures of some nodal points are plotted versus time in Figure 2. T h e temperature increase of the body is very slow. After 1000 days the temperature

T
50_

40

30
20

10
0 1 0 -2

10 -1

10 o

101

10 2

10 3

10 4

Fig. 2. Thermo-elastic consolidation: temperature versus time at different nodes.

CONSOLIDATION OF NONISOTHERMAL ELASTOPLASTIC POROUS MEDIA


Uz

169

0.5.10 ~

0.4,10 -3

0,3.10 ~

0,2.10 <

0.1.10a

0
10 -2

10 -1

10 0

101

10 2

10 3

10 4

Fig. 3. Displacements versus time at different nodes.

in the b o t t o m half of the sample is still below 20 degrees. This explains the settlement histories of Figures 3, 6 and 7 where a reversal of the displacement is observed only after that time. T h e pore pressures versus time at several nodes are shown in figure 4. T h e damping effect of the 3 3 Gaussian integration scheme on the temperature propagation along the column is clearly shown in Figure 5, where a comparison with a 2 2 integration scheme is made at two different time values. After t = 1, the t e m p e r a t u r e distributions are the same; this justifies the use of the 2 2 rule throughout the solution. T h e settlement versus time of the top surface is shown in Figure 6 for the following situations: isothermal consolidation, applied surface temperature, temperature change and surface load applied simultaneously. These values were

pd

t_

10 -2

10 -1

100

10~

102

10 a

10 4

Fig. 4. Pore pressures versus time at different nodes.

170
50

R . W . LEWIS, C. E. MAJORANA, AND B. A. SCHREFLER


50 50 50

/
>
<
t = 10-2 t =10 - 2 t = 10 - 1 t = 10-1

(a)

(b)

(a)

(b)

Fig. 5. Pore pressures at two different time values using (a) 3 3 Gauss points in the spatial integration, (b) 2 2 Gauss points in the spatial integration.

obtained with the assumption of incompressible fluid and solid grains. The effect of the compressibility of the fluid and solid grains with bulk moduli K w = 0.43 101~ and Ks = 0.14 x 101~ are also shown in Figure 6. Finally, for comparison, the results obtained by Aboustit et al. [14] are drawn with dots. These results are given in tabular form in [14]. For most of the values, the comparison is excellent except for two values. The difference is probably due to a printing error in Table I of [14]; in fact, an agreement between the dots and our values is obtained for a time value which is ten times smaller. On the other hand, an

Uz

<0

0.5.10:

0.4.10-:

0.3.10 <

0.2-10 :

0.1-10 ~

0
10:2 10-I 10 0 101 10 2 10 3 10 4 t

Fig. 6. Surface settlements versus time for the 1-D example problem of Figure 1. (a) isothermal consolidation only, (b) applied surface temperature only, (c) nonisothermal consolidation with incompressible fluid and solid grains, (d) nonisothermal consolidation with the actual values of the bulk moduli. Q values obtained from [14].

CONSOLIDATION OF NONISOTHERMAL ELASTOPLASTIC POROUS MEDIA


Uz

171

05-10-"

0.4.10-:3

O" 3"10-~

1~//'

0.2.10

,/(/(b)
" ~ " " "

0.1.10 ~

J
, i

~'s///

0
10 -2

10 1

10~

101

10 2

103

10 4

t
flw = 0

Fig. 7. Surface settlements versus time. (a) thermal expansion coefficient of the fluid (diagram c of Figure 6). (b)/3w = 0.63 10-5.

inspection of the settlement histories for p and T only excludes the two values indicated in [14]. T h e influence of the thermal expansion coefficient of the fluid and, hence, of the first derivative of T in the continuity equation (16) was also investigated, assuming flw = 0 . 6 3 10 -5. Figure 7 shows the effects on the displacements, Figure 8 the effects on the pore pressure. T h e effects of convection, which produce a real coupling between the pressure and the t e m p e r a t u r e fields have also been investigated and found to be negligible in this case. T h e P e t r o v - G a l e r k i n and upwinding techniques were used for this purpose. As already indicated, the effect of the heat generated by elastic deformation was negligible. This was also verified by actually introducing the matrix ~o 9T U T in the third equation of our model.

10-2 10-1 10~ 10~ 102 10 3 10 4 t Fig. 8. Pore pressures versus time at three different nodes obtained with (a) /3w = 0, (b) /3w = 0.63 x 10-5.

172

R.W. LEWIS, C. E. MAJORANA, AND B. A. SCHREFLER

E X A M P L E 2. Thermo-elastoplastic consolidation. T h e same problem as in E x a m p l e 1 was also investigated, assuming an elastoplastic behaviour with a t e m p e r a t u r e - i n d e p e n d e n t yield function. A critical state yield function of the form
F = q2 _ M2sppc(2 _ P/Pc) = 0

was assumed,

where

M~s = slope of the failure line in the q - p space, Pc = m e a s u r e of the current size of the yield surface related to the strain t~ardening p a r a m e t e r hi = S deP by the expression
Pc = Pco e x p ( h l / x ) ,

where

Pco = half preconsolidation pressure,


X-l+eo' eo = initial void ratio, ,~ = slope of virgin consolidation line on e - In p plot, K --- slope of swelling line on e - In p plot.

T h e material properties assumed were: Mcs = 1.05, = 0.14, K = 0.05, eo = 0.90 [26]. T h e initial elastic modulus, Poisson's ratio and the permeability tensor were the same as for E x a m p l e 1. T h e time-step values used in this instance are indicated in T a b l e II.
Table II. Time steps different time intervals Time interval 0.01 0.1 1 2 5 10 20 Number of time steps 10 10 10 10 10 15 35 for

T h e settlement history for different values of the preconsolidation pressure is shown in Figure 9. For Pc0 = 0.40 some yielding a p p e a r e d at the beginning of the transient, then the material behaviour was simply elastic; with pco = 0.030, 0.35, 0.375 the reversal of the settlement was not observed. T h a t means that the plastic flow due to the unit pressure applied is higher than the thermal expansion throughout the transient. T h e pore pressure is drawn versus time in Figure 10 for pco = 0.35. T h e other values of Pco considered before give a similar behaviour. It

CONSOLIDATION OF NONISOTHERMAL ELASTOPLASTIC POROUS MEDIA


U Z

173

o.1 .lo-!

08.10-a

o6.1o-~

o.410!

///
/ / /

therrno-ela.stic
consolidation

Q2.10-~

10 .2

10 -1

10 0

101

10 2

10 a

10 4

Fig. 9. Thermo-elastoplastic consolidation of the model problem of Figure 1: settlements versus time for different values of the preconsolidation pressure.

should be remembered that nonisothermal conditions also require a temperaturedependent yield function. This is actually being investigated. E X A M P L E 3. Thermo-elastic consolidation around a cylindrical heat source. In the previous examples, the flow of pore water and of temperature with time was in the vertical direction only. For this reason nodes at the same depth had identical pore pressure and temperature histories. In this examples, a true axisymmetric problem is investigated. The effects of a cylindrical radiating source, buried in a thermoelastic soil, was examined by Booker et al. in [2], where an analytic solution for a point-heat source was numerically integrated over the surface of a cylindrical canister. The solution was shown for a particular case of a cylindrical heat source in

10 .2

10 "1

10 ~

101

10 2

10 3

10 4

Fig. 10. Thermo-elastoplastic consolidation: pore pressures versus time of three different nodes for Pco = 0.35.

174

R . W . LEWIS, C. E. MAJORANA, AND B. A. SCHREFLER

"l

t,

Fig. 11.

Finite element discretization of the radial cross-section containing a cylindrical heat source.

which a'/a,, = J, v = 0.4, c/r = 2, with a' coefficient of v o l u m e expansion of the soil, a , =/3s (1 - qS) +/3,~b, c coefficient of consolidation, K coefficient of thermal diffusivity. This solution is used as c o m p a r i s o n with our numerical model. T h e finite e l e m e n t m e s h is s h o w n in Figure 11, w h e r e ro is the radius of the heat source. A set of possible field data which satisfy the a b o v e ratios are: E =- 6000, v = 0.4, K = 0.4 10 -5, h = 1.02857, (pc)o = 40, /3s = 0.9 10 -6, /3w = 0.63 10 -5. T h e heat source was simulated by a c o n s t a n t heat input of 1000 for e a c h of the two elements of the source.
I | I

T/T~

1. Z

1.0 0.8

r/r~ ~ i/,~'J~ . fr

O.Z

0.1

10

100

1000 [
analytical solution of [2], G

Fig. 12. Temperatureversus time for a cylindrical heat source: - finite clement solution.

CONSOLIDATION OF NONISOTHERMAL ELASTOPLASTIC POROUS MEDIA

175

P/PN 0.20
0.16

0.12

0.1
finite element solution.

10

100 [
analytical solution of [2], Q

Fig. 13. Pore pressures versus time for cylindrical heat source: - -

The temperature and pressure histories for three values of r/ro are shown in Figures 12 and 13. In these diagrams the temperature is normalized with respect to its maximum value TN reached at the surface midpoint of the heat source. The pressure is given as a ratio of pN, the maximum pore water pressure at the surface midpoint of the heat source if the soil were impermeable [2], and [ is given by t = r 2 t / K . The agreement between the analytical solution and the finite element solution is remarkable. The modelling by finite elements was found to be sensitive to the ratio between the characteristic length of the domain (finite in this model) and the size of the heat source. For this purpose, three different cases were investigated. Taking the axisymmetric domain (Figure 11) 8 x 8 in section and a radius of the heat source ro = ~ so that [ = t, a solution was obtained for the following values of the length of the cylinder: 2.0, 0.5, 1.25. In the first case the temperature variation was sufficiently close to that predicted with the analytical model; the maximum temperature at r/ro = 1 and z = 0 was T = 35.53. The pore pressure was underestimated until [ = 1, overestimated after [-~ 1: at [ = 100 the pore pressures were about twice the analytical ones. The second case showed an overall underestimation of the temperature

Table III. Time 0.1 1 10 100 1000 10000

Displacements at r/ro = 1, 2, 5; z = 0
r/ro = 1 r/ro = 2 r/ro = 5

0.1961 0.6689 1.0755 1.1352 1.0142 1.0000

0.1516 0.9103 1.7819 1.9160 1.6758 1.6462

0.0580 0.6036 2.7405 3.2868 2.7263 2.6928

176

R. W. LEWIS, C. E. MAJORANA, AND B. A. SCHREFLER

which attained the maximum value T = 21.68 at the same point as above. The pore pressures were very high at i = 1, then they dissipated rapidly showing a substantial coincidence with the analytical curves between i = 10 ,rod i = 100. The third case (h = 1.25) was the most satisfactory, both for the temperatures and the pore pressures. Their variations are those shown in Figures 12 and 13. The displacements of three nodes along the r-axis are indicated in Table III (the values are normalized, with respect to u~ (r/ro = 1)).

5. Conclusions
A numerical model for the analysis of thermoelastoplastic porous media is presented, which fully takes into account the interaction between the displacements, pressure, and temperature fields. Appropriate solution techniques for such interaction problems are shown. The model is applied to plane strain thermoelastic and thermoplastic consolidation and to the investigation of the consolidation around a cylindrical heat source buried in saturated clay. The agreement with existing solutions, where applicable, is remarkable. The model will now be applied to the investigation of the subsidence of the geothermal field of Abano in northeast of Italy and to the investigation of the upheaval of Pozzuoli near Naples, due to a heat source of volcanic origin.

Appendix A
List of the matrices K . . . . , T G of Equations (30)-(32).

K = Ia - BTDTB d~,
L = i m (B'rmN n - BTDT ~--~N ) dgl, T U = I n B r D r m ff~ N d~Q, L r = In ( N T m B - N T D r ~ 3K ms B) dD''

H = f n ( V N ) T _k V N dD~,

Kw

(3Ks)2 m r D T m N dD.,

1 ~ _~ mTDTm] N d12' T P = In N T [ -- */3w -- (1 -- *)~s + 3K

CONSOLIDATION OF NONISOTHERMAL ELASTOPLASTIC POROUS MEDIA

177

9 L-- Io

- '"

T R = fc~ (V N)T~ V N d " + fc~(VN)TpwCwV~N da,

TS = Ia NT[(1 -- 4~)psC~+ ~bp~C~- (1 - 4>)p~C~/3~Nf - ~b/3~C~f3~N'Y]Nd~,


C = - Ia BTDTc d~,

c3f_ c g t I ~ NT O tc3b d ~ F

Ir NT c~tdF -

Ot

c3~~ dfl IaBTDTo~ '

IrNTq~ dF, G=-J,r T N/ 1 ~T m


= -

DTc+Qe d ~ -

Ia(VN) T k v p ~ g z d ~ ,

TG=-IrNTqhdF-I NTQhd~.
Acknowledgement The authors wish to thank the CNR (Italy) for financing this project under Contract Number 83.02501.07.

References
1. Borsetto, M., Carradori, G., and Ribacchi, R., 1981, Coupled seepage, heat transfer and stress analysis with applications to geothermal problems, R. W. Lewis, K. Morgan, and O. C. Zienkiewicz (eds.), Numerical Heat Transfer, Wiley, Chichester. 2. Booker, J. R., and Savvidou, C., 1985, Consolidation around a heat source, Int. J. Num. Anal. Methods Geom. 9, 173-184. 3. Corapcioglu, M. Y. and Karahanoglu, N., 1980, Simulation of geothermal production, in T. Nejat Verizoglu (ed.), Alternative Energy Sources H, Vol. 5, Hemisphere Publ. Co., N.Y., pp. 1895-1918. 4. Lippmann, M. J., Narasimhan, T. N., and Witherspoon, P. A., 1976, Numerical simulation of reservoir compaction in liquid dominated geothermal systems, Proc. 2nd. Int. Symp. Land Subsidence, International Association of Hydrological Sciences, Anaheim, Calif., Pub. No. 121, pp. 179-189. 5. Brownell, D. H., Garg, S. K., and Pritchett, J. W., 1977, Governing equations for geothermal reservoirs, Water Resour. Res. 12, 929-934. 6. Pritchett, J. W., Garg, S. K., Brownell, D. H., Rice, L. F., Rice, M. H., Riney, T. D., and Hendrickson, R. R., 1976, Geohydrological environmental effects of geothermal power production, Phase IIa, Systems, Science and Software, Report SSS-R-77-2998, La Jolla, Calif.

178

R. W. LEWIS, C. E. MAJORANA, AND B. A. SCHREFLER

7. Garg, S. K., Pritchett, J. W., Rice, M. H., and Riney, T. D., 1977, U.S. Gulf Coast geopressure geothermal reservoir simulation, Final Report, Systems, Science and Software, Rep. SSS-R-773147, La Jolla, Calif. 8. Aktan, T. and Farouq Ali, S. M., 1978, Finite element analysis of temperature and thermal stresses induced by hot water injection, Soc. Pet. Eng. J., 457-469. 9. Ertekin, T., 1978, Numerical simulation of the compaction-subsidence phenomena in a reservoir for two-phase non-isothermal flow, PhD Thesis, The Pennsylvania State Univ., Penn. 10. Bear, J. and Corapcioglu, M. Y., 1981, A mathematical model for consolidation in a thermoelastic aquifer due to hot water injection or pumping, Water Resour. Res. 17, 723-736. 11. Schrefler, B. A., 1985, A partitioned solution procedure for geothermal reservoir analysis, C.A.N.M., 1, 53-56. 12. Borsetto, M., Cricchi, D., Hueckel, T., and Peano, A., 1984, On numerical models for the analysis of nuclear waste disposal in geological clay formations, in R. W. Lewis, E. Hinton, P. Bettes, and B. A. Schrefler, (eds.), Numerical Methods for Transient and Coupled Problems, Pineridge Press, Swansea, pp. 603-618. 13. Dawson, P. R. and Chavez, P. F., 1980, Thermomechanically coupled creeping flow of saturated cohesive porous media, Proc. 3rd. Int. Con]:. Finite Elements in Flow Problems, Banff, pp. 268-279. 14. Aboustit, B. L., Advani, S. H., and Lee, J. K., 1985, Variational principles and finite element simulations for thermo-elastic consolidation, Int. J. Num. Anal. Methods Geom. 9, 49-69. 15. Sauty, J. P., Gringarten, A. C., Menjoz, A., and Landel, P. A., 1982, Sensiible energy storage in aquifers 1. Theoretical study, Water Resour. Res. 18, 245-252. 16. Doughty, C., Hellstr6m, G., Tsang, C. F., and Claesson, J., 1982, A dimensionless parameter approach to the thermal behaviour of an aquifer thermal energy storage system, Water Resour. Res. 18, 571-587. 17. Lewis, R. W. and Karahanoglu, N., 1981, Simulation of subsidence in geothermal reservoirs, in R. W. Lewis, K. Morgan, and B. A. Schrefler, (eds.), Numerical Methods in Thermal Problems, Vol. II, Pineridge Press, Swansea, pp. 326-335. 18. Zienkiewicz, O. C., Norris, V. A., Winnicki, L. A., Naylor, A. J., and Lewis, R. W., 1978, A unified approach to the soil mechanics problems of offshore foundations, in O. C. Zienkiewicz, R. W. Lewis, and K. G. Stagg, (eds.), Numerical Methods in Offshore Engineering, Wiley, Chichester, Ch. 12, pp. 361-412. 19. Lewis, R. W. and Schrefler, B. A., 1982, A finite element simulation of the subsidence of a gas reservoir undergoing a waterdrive, in R. H. Gallagher, D. H. Norrie, J. T. Oden, and O. C. Zienkiewicz, Finite Elements in Fluids, Vol. 4, Wiley, Chichester, pp. 179-199. 20. Schrefler, B. A., 1984, The finite element method in soil consolidation (with applications to surface subsidence), PhD Thesis, University College of Swansea. 21. Zienkiewicz, O. C., 1977, The Finite Element Method, McGraw-Hill, London. 22. Combarnous, N. A. and Bories, S. A., 1975, Hydrothermal convection in saturated porous media, V. T. Chow, (ed.), in Advances in Hydroscience, Vol. 10, Academic Press, New York, pp. 231-307. 23. Schrefler, B. A. and Vitaliani, R., 1985, Partitioned solution procedure for educational software on microcomputers: the case of a coupled consolidation analysis, Proc. Int. Con]:. Education, Practice and Promotion of Computational Methods in Engineering Using Small Computers (EPMESC), Macao, 5-9 August. 24. Park, K. C., 1985, Stabilization of partitioned solution procedure for pore fluid soil interaction analysis, I.J.N.M.E. 19, 1669-1673. 25. Schrefler, B. A., Majorana, C. E., and Lewis, R. W., A comparison between different simulators for geothermal reservoirs (to be published). 26. Siriwardane, H. J. and Desai, C. S., 1981, Two numerical schemes for nonlinear consolidation, I.J.N.M.E. 17, 405--426.

You might also like