Sei sulla pagina 1di 10

717

NOTE / NOTE

Nonlinear consolidation of thin layers subjected to


time-dependent loading
Enrico Conte and Antonello Troncone

Abstract: The paper deals with one-dimensional consolidation of saturated clays with variable compressibility and perme-
ability. A formulation is developed to analyse the consolidation of thin clay layers subjected to time-dependent loading.
Moreover, a simple solution procedure is presented, which makes use of some analytical expressions derived in this study
in conjunction with the Fourier series. Comparisons with other analytical and numerical solutions are shown, and some
aspects of the nonlinear consolidation caused by time-dependent loading are highlighted.
Key words: one-dimensional consolidation, nonlinear theory, time-dependent loading, excess pore-water pressure,
settlement rate.
Résumé : L’article traite de la consolidation unidimensionnelle des argiles saturées avec compressibilité et perméabilité
variables. Une formulation est développée pour analyser la consolidation des couches minces d’argile soumises à un
changement variable avec le temps. En outre, une solution simple est présentée qui utilise des expressions analytiques,
obtenues dans cette étude, et les séries de Fourier. Des comparaisons avec les résultats d’autres solutions théoriques
sont montrées, et quelques caractéristiques de la consolidation non linéaire, qui résulte de l’application d’une charge
variable avec le temps, sont mises en évidence.
Mots-clés : consolidation unidimensionnelle, théorie non linéaire, chargement variable avec le temps, surpression
interstitielle, vitesse de tassement.

Conte and Troncone 725

Introduction Townsend and McVay (1990), Sills (1995), De Boer et al.


(1996), Lancellotta and Preziosi (1997), and Yao et al.
During the consolidation processes in saturated clays, the (2002). Lastly, Bartholomeeusen et al. (2002) discussed the
changes in permeability and compressibility can significantly identification of the soil parameters necessary to analyse real
affect the generation and dissipation of excess pore-water situations properly.
pressure and the rate of settlement. In Terzaghi’s one- From a practical point of view, the nonlinear theory pro-
dimensional theory (Terzaghi 1925), which is the most posed by Davis and Raymond (1965) is the most attractive
widely used in engineering practice, these variations are ig- for current applications because it leads to an analytical solu-
nored in the interests of a governing differential equation, tion similar to that of Terzaghi’s linear theory. Specifically,
which can be readily solved. Since the development of Davis and Raymond developed a consolidation model based
Terzaghi’s theory, many attempts have been made to present on the commonly used e – log σ′ (void ratio – log effective
more general one-dimensional models in which the varia- stress) relationship and the assumption that the coefficient of
tions in compressibility and permeability are incorporated consolidation remains constant during consolidation while
(Barden and Berry 1965; Davis and Raymond 1965; Janbu the compressibility and permeability are both allowed to de-
1965; Mikasa 1965; Viggiani 1973; Burghignoli 1979; Zhuang crease with increasing effective stress. In others words, the
et al. 2005; and others). In this context, Gibson et al. (1967), authors assumed that during consolidation the decrease in
Poskitt (1969), Mesri and Rokhsar (1974), and Schiffman et permeability is compensated for by the decrease in com-
al. (1984) developed comprehensive formulations that ac- pressibility. Further assumptions of the model are that the ef-
count for both material nonlinearity and finite strains. More fect of the soil weight is negligible and the applied load is
recently, noteworthy contributions have been presented by constant with time. In reality, however, during the construc-

Received 31 July 2006. Accepted 22 January 2007. Published on the NRC Research Press Web site at cgj.nrc.ca on 24 July 2007.
E. Conte1 and A. Troncone. Dipartimento di Difesa del Suolo, Università della Calabria, Ponte P. Bucci, Cubo 41b, 87036 Rende
(Cosenza), Italy.
1
Correspnding author (e-mail: conte@dds.unical.it).

Can. Geotech. J. 44: 717–725 (2007) doi:10.1139/T07-015 © 2007 NRC Canada


718 Can. Geotech. J. Vol. 44, 2007

tion of civil engineering works the loads are generally applied where λ is a constant parameter defining the slope of the
gradually in time and in many cases the loading process de- compression curve in the aforementioned bilogarithmic plot.
velops over a long time, so that a significant part of the con- Moreover, in many situations it is reasonable to assume
solidation occurs in this time. This problem was investigated that the changes in cv occurring during the consolidation
recently by Conte and Troncone (2006) who presented a so- process may be neglected (Davis and Raymond 1965;
lution procedure for the analysis of the linear one- Poskitt 1969; Lancellotta 1995). This is due to the fact that
dimensional consolidation in saturated soils subjected to the decrease in kw tends to compensate for the decrease in
general time-dependent loading. mv. As a consequence of this assumption and owing to
In the present paper, a formulation is developed to analyse eq. [3], the following relationships can be written:
the nonlinear consolidation of thin clay layers when the vari-
0.434 C γw c v
ations in the external load with time are considered. Some [5] kw =
analytical expressions are first derived and then incorporated σ′
into the procedure proposed by Conte and Troncone (2006)
to readily achieve the solution to nonlinear one-dimensional and
consolidation. The validity of this solution is proven by the ∂k w 1 ∂ σ′
comparison with analytical and numerical methods. [6] = −0.434C c v γ w 2
∂z σ′ ∂ z
Governing equation Substituting eqs. [5] and [6] into the last term of eq. [1]
yields
The equation governing the one-dimensional consolida-
tion in saturated soils can be written in terms of excess pore- ∂ 2u 1  d q ∂ u  1 ∂ σ′ ∂ u
water pressure as (Lancellotta 1995) [7] +  −  − =0
∂z  d t ∂ t  σ′ ∂ z ∂ z
2 cv
∂ 2u 1  d q ∂u  1 ∂ k w ∂u
[1] +  −  + =0 Moreover, the effective vertical stress can be expressed as
∂z  dt ∂t  k w ∂z ∂z
2 cv
[8] σ′ = σ′o + q − u
where z denotes the spatial co-ordinate; t is time; u is the ex-
cess pore-water pressure depending on both z and t; q is the where σ′o is the initial effective vertical stress, which for a
external load applied at the ground surface, which is a func- thin soil layer may be considered to be constant with depth.
tion of time; kw is the coefficient of permeability; and cv is Therefore, from eq. [8] it follows that
the coefficient of consolidation given by ∂ σ′ ∂u
[9] =− q is only a function of t, is irrelevant to z.
[2] cv =
kw ∂z ∂z
γw m v
and eq. [7] takes the form
in which mv is the coefficient of volume change of the soil
skeleton, and γ w is the unit weight of water. Equation [1] is  1 ∂ 2u  1  2  ∂ u  2  1  d q ∂ u 
[10] −c v  +      =  − 
based on the following assumptions: water and solid parti-  σ′ ∂ z  σ′   ∂ z   σ′  dt ∂ t 
2
cles are incompressible, water flow is described by Darcy’s
law, soil behaviour is governed by the effective stresses, and
which reduces to the equation presented by Davis and
strains are small. Moreover, creep, thermal, and inertial ef-
Raymond (1965) when q is assumed to be constant with
fects are ignored.
time. Equation [10] can be expressed in a simpler form by
Using the results of one-dimensional consolidation tests
using a convenient function w defined as
the coefficient of volume change can be expressed as
0.434 C  σ′   σ′ + q − u 
[3] mv = [11] w = ln   = ln  o 
σ′  σ′o   σ′o 
where σ′ is the effective vertical stress depending on z and t; which is different from that considered by Davis and Raymond
and C is equal to the recompression ratio, RR, or to the (1965).
compression ratio, CR, depending on whether the soil is In fact, differentiating eq. [11] with respect to z
overconsolidated or normally consolidated, respectively. If
∂w 1 ∂u
the compression curve from the laboratory tests is fairly [12] =−
straight or may be approximated by a straight line over the ∂z σ′ ∂ z
relevant range of effective stress, C can be considered as a
2
∂2w
2
constant. With regard to this aspect, it is of interest to note 1 ∂ 2u  1   ∂u 
that when the experimental data are plotted in terms of [13] =− −   
∂z 2 σ′ ∂ z 2  σ′   ∂z 
log v − log σ′ (log specific volume of the soil – log effective
stress), as suggested by Butterfield (1979), the coefficient
and with respect to t
of volume change is given by (Lancellotta 1995)
λ ∂w 1  d q ∂u 
mv = [14] =  − 
∂ t σ′  d t ∂ t 
[4]
σ′
© 2007 NRC Canada
Conte and Troncone 719

and substituting eqs. [13] and [14] into eq. [10] lead to the Fig. 1. The soil layer considered to achieve the solution to eq. [15].
same differential equation as that in Terzaghi’s linear theory,
provided that the excess pore pressure is replaced by w, i.e.,
∂ 2w ∂w
[15] cv =
∂z 2 ∂t

This equation has to be solved accounting for the boundary


and initial conditions. Referring to the soil layer schematized
in Fig. 1, the boundary conditions are:
at z = H, where H is the layer thickness, ∂ u ∂ z = 0 (the layer
base is impervious); owing to eq. [12] we have
∂w
[16] =0
∂z

at z = 0 0 , u = 0 (the upper surface is fully permeable); owing the settlement, s, at any time is provided by the equation
to eq. [11] we can write H H

 q  [22] s = ∫ ε d z = 0.434 C ∫ w dz
[17] w = ln 1 +  0 0
 σ′o 
in which q is a function of time. It should be observed that Solution
for the case in which the lower surface of the layer is perme- The solution to eq. [15] with the boundary and initial
able like the upper one (double drainage condition), eq. [17] conditions stated in eqs. [16]–[18] can be achieved using
has to be considered both at z = 0 and z = H. This implies Duhamel’s theorem, which is expressed by the following
that the unknown function w is symmetrical with respect to equation (Carslaw and Jaeger 1959):
the middle plane of the layer and can be obtained from
eq. [15] by imposing eq. [17] at z = 0 and eq. [16] at z = H 2. t
∂w
In other words, as expected, the solution to the single drain- [23] w =∫ F dτ
0
∂t
age condition can be used for the double drainage condition
by interpreting H as the drainage distance. where F = ln(1 + q/σ′o ) is the time function imposed at z = 0
Lastly, the initial condition is u = q at t = 0; therefore, (eq. [17]), and w is the solution to eqs. [15], [16], and [18]
owing to eq. [11] it results when F = 1. As shown in Appendix A, this latter solution
[18] w = 0 becomes
4 ∞ 1
As can be noted, taking advantage of eq. [11] the differen-
tial equation governing the nonlinear consolidation (eq. [10])
[24] w = 1− ∑
π n=1 ( 2 n −1)
sin( α n z)exp( −α n2 c v t)
is reduced to the simplest diffusion equation (eq. [15]) in
which the unknown variable w assumes an initial value where α n = ( 2 n −1) π ( 2H ). In eq. [23], F is evaluated at time
equal to zero (eq. [18]) and boundary values that vary with τ and w at time (t − τ).
time according to a prescribed function (eq. [17]). These If we assume that F varies harmonically with time accord-
boundary and initial conditions are different from those ing to the equation
considered to achieve the solution to Terzaghi’s theory.
[25] F = A cos( ω t) + B sin( ω t)
Therefore, this latter solution cannot be used for the case
under consideration. A solution to eq. [15], which also sat- in which A and B are the amplitudes of F, and ω is the circu-
isfies eqs. [16]–[18], is presented in the next section. Using lar frequency, by substituting τ for t in eq. [25] and (t − τ) for
this latter solution and taking into account eq. [11], the t in eq. [24] and performing the partial derivative of w with
excess pore-water pressure and effective vertical stress at respect to t, eq. [23] leads to
any time and depth can be calculated, respectively, by the ∞
Nϑ Nz
equations [26a] w k = 2∑ X n sin  
n =1 1 + ϑ N  H 
2 4
[19] u = σ′o (1 − e w ) + q
this latter is hence the analytical solution to eqs. [15]–[18]
and when F is expressed by eq. [25]. In eq. [26a],
[20] σ′ = σ′o e w [26b] N = ( 2n −1) π 2
In addition, since the vertical strain, ε, developing within [26c] X n = {( B − AϑN 2 )[exp( −N 2 T v ) − cos( ω t)]
the soil is commonly expressed by the relation
+ ( A + BϑN 2 ) sin( ω t)}
σ′
[21] ε = 0.434 C ln
σ′o [26d] ϑ = c v ( ωH 2)
© 2007 NRC Canada
720 Can. Geotech. J. Vol. 44, 2007

and where w is provided by eq. [24]. Once the function w is ob-


tained, the excess pore pressure and the effective vertical
[26e] T v = c v t H 2
stress at any time and depth can be calculated using eqs. [19]
It should be observed that eq. [26a] can be used to calcu- and [20], respectively.
late wk when 0 < z ≤ H, whereas different mathematical op- Likewise, the expression for the settlement, s, at any time is
erations are required to achieve a solution that is valid also M
for z = 0. However, this leads to complicated expressions, [33] s = so + ∑ sk
which are very cumbersome to use, especially for practical k =1
applications. Therefore, we have preferred to maintain the
where s o is given by
solution in the simpler form given by eq. [26a] for 0 < z ≤
H, and to employ directly eq. [17] to determine the values of Ao
wk at z = 0. [34] s o = 0.434CH
Lastly, owing to eq. [22] the expression for the settlement 2
at any time, sk , is  8 ∞
1 

ϑ
× 1 − 2 ∑ ( 2n −1) 2 exp [ − ( 2n −1) 2 π 2 T v / 4 ] 
[27] s k = 0.868 CH ∑ Xn  π n =1 
n=1 1 + ϑ
2 4
N
The derived analytical expressions can be incorporated Applications
into the procedure proposed by Conte and Troncone (2006) In this section, a series of comparisons with theoretical
to achieve a solution to eq. [15] when a more general time- results are presented to assess the accuracy of the proposed
dependent function than eq. [25] is considered for F. Spe- solution. Moreover, some aspects of the nonlinear consoli-
cifically, the procedure first requires that the prescribed dation caused by time-dependent loading are highlighted.
function F is expanded in a finite number M of harmonic
As the first example, the results published by Davis and
components using a Fourier series
Raymond (1965) are considered. These results concern a thin
Ao M layer of saturated soil subjected to a uniform load q that is in-
[28] F= + ∑ [ Ak cos( ωk t) + Bk sin( ωk t)] stantaneously applied at the layer surface and thereafter is
2 k =1
held constant with time. The base of the layer is impervious,
where the series amplitudes Ak and Bk associated with the whereas the upper surface is permeable. The excess pore pres-
frequency ωk are provided, respectively, by sures were calculated by Davis and Raymond (1965) at z = H
T
for different values of the time factor and initial effective ver-
2 tical stress. These results are compared in Table 1 with those
[29] Ak =
T ∫ F cos( ωk t) dt obtained using the present solution when q σ′o = 0.5, 1, 3, 7,
0
and 15. In this table the excess pore pressures are expressed
and as a percentage of the load intensity. In addition, the values of
T
the degree of consolidation settlement Us are shown. This lat-
2 ter variable is defined as
[30] Bk =
T ∫ F sin(ωk t) dt
0 st
[35] Us =
and T is the period of F, which is related to ωk by the expres- sc
sion ωk = 2k π T (with k = 1, 2,…, M). It is of interest to
note that, in the calculations, T must be assumed to be where st is the settlement at the time of analysis, and s c is the
greater than the final time of analysis. Moreover, Ao can be settlement at the end of the consolidation process. Although
obtained from eq. [29] when ωk = 0, i.e., for the case under consideration, the solution to eq. [15] could
be directly obtained from the equation
T
2
[31] Ao =
T ∫ F dt [36] w = w ln(1 + q σ′o )
0
the values shown in Table 1 were calculated using the com-
It should be noted that if the time-dependent loading con- plete procedure described in the previous section, with a
sists of a series of step loads, as for example occurs in the view to assessing the accuracy of the proposed solution by
case of fills, simple closed-form expressions can be obtained comparing the results with those obtained by the analytical
for Ak, Bk, and Ao using eqs. [29]–[31]. expression derived by Davis and Raymond (1965), which, as
Then, for each of the components of F considered in said, can only be used when the load is constant with time.
eq. [28], the corresponding function w k is calculated using As can be seen, the present solution and that developed by
eqs. [26a]–[26e] in which A and B are replaced by Ak and Davis and Raymond provide essentially the same results
Bk , respectively, and ω is replaced by ωk . Lastly, the sought- both in terms of u q and Us. Moreover, the results prove that
after solution is achieved by superimposing all the harmonic the pore-pressure dissipation depends on the ratio, q σ′o .
components of wk, i.e., The greater this ratio, the slower the pore-pressure dissipation.
Ao M On the other hand, the degree of settlement is independent of
[32] w= w + ∑wk q σ′o and is equal to that predicted by Terzaghi’s linear theory,
2 k =1 as previously emphasized by Davis and Raymond (1965).
© 2007 NRC Canada
Conte and Troncone 721

Table 1. Comparison of Davis and Raymond’s (1965) analytical expression and the present solution: percentage values of excess pore-
water pressure u q and the degree of settlement Us, for different values of the time factor, Tv, and the ratio of loading intensity to the
initial effective vertical stress, q σo′ .
u q (%) u q (%) u q (%) u q (%) u q (%) Us (%)
for q σo′ = 0.5 for q σo′ = 1.0 for q σo′ = 3.0 for q σo′ = 7.0 for q σo′ = 15.0 for all q σo′
Davis and This Davis and This Davis and This Davis and This Davis and This Davis and This
Tv Raymond study Raymond study Raymond study Raymond study Raymond study Raymond study
0.02 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0 100.0 16.0 15.7
0.04 99.9 99.9 99.9 99.9 100.0 100.0 100.0 100.0 100.0 100.0 22.6 22.4
0.08 98.0 98.0 93.3 98.3 98.8 98.9 99.2 99.3 99.5 99.5 31.9 31.8
0.1 95.9 95.9 96.4 96.5 97.6 97.6 98.4 98.4 99.0 99.0 35.7 35.6
0.2 80.7 80.7 82.9 83.0 87.6 87.7 91.4 91.4 94.1 94.2 50.4 50.4
0.3 65.4 65.5 68.7 68.7 75.8 75.9 81.9 82.0 86.8 86.9 61.3 61.3
0.4 52.5 52.6 56.1 56.1 64.3 64.3 71.7 71.7 78.1 78.1 69.8 69.8
0.5 41.9 41.9 45.3 45.4 53.6 53.6 61.4 61.5 68.5 68.6 76.4 76.4
0.6 33.3 33.3 36.4 36.4 44.1 44.1 51.7 51.8 58.9 58.9 81.6 81.6
0.7 26.3 26.3 29.0 29.1 35.9 36.0 42.9 43.0 49.7 49.8 85.6 85.6
0.8 20.8 20.8 23.1 23.1 29.0 29.0 35.2 35.2 41.3 41.4 88.7 88.8
0.9 16.4 16.4 18.3 18.3 23.3 23.3 28.5 28.6 34.0 34.0 91.2 91.3
1.0 12.9 12.9 14.4 14.4 18.5 18.6 23.0 23.0 27.6 27.6 93.1 93.2
2.0 1.11 1.12 1.27 1.27 1.68 1.69 2.16 2.16 2.67 2.68 99.4 99.5
3.0 0.09 0.10 0.11 0.11 0.14 0.15 0.18 0.19 0.23 0.24 100.0 100.0
4.0 0.01 0.01 0.01 0.01 0.01 0.01 0.02 0.02 0.02 0.02 100.0 100.0

Fig. 2. Comparison of linear and nonlinear consolidation: loading Fig. 3. Comparison of linear and nonlinear consolidation: excess
process. pore-water pressure u at z = H, normalized with respect to the
greatest intensity of loading qu, versus time factor Tv.

Except for the loading process, which in this case consists


of three step loads (Fig. 2), the results in Figs. 3 and 4 refer
to the same case study. These figures show a comparison be-
tween the linear and nonlinear theories in terms of excess
pore pressure at z = H (Fig. 3) and degree of settlement
(Fig. 4) against the time factor. The excess pore pressures
are normalized with respect to the greatest intensity of load-
ing, q u (Fig. 2). Moreover, the results of the nonlinear con- calculated using the nonlinear theory depend on q u σ′o ,
solidation theory are presented for two values of q/ σ′o that unlike in the previous example where a constant load was
are 3 and 15. The calculations for the linear consolidation considered. These values differ significantly from those ob-
were performed using the solution developed by Conte and tained when the linear theory is used, although the same co-
Troncone (2006). As can be noted, the linear consolidation efficient of consolidation was assumed in the calculations.
gives an underestimation for both the settlement rate and Another case examined in this study concerns an ideal
excess pore-pressure dissipation with respect to the nonlin- embankment constructed on a soil deposit consisting of a
ear consolidation. Moreover, Fig. 4 shows that in the present clay layer sandwiched between an upper layer of sand and a
case wherein the load is time-dependent, the values of Us lower layer of gravel (Fig. 5). Owing to the large lateral ex-
© 2007 NRC Canada
722 Can. Geotech. J. Vol. 44, 2007

Fig. 4. Comparison of linear and nonlinear consolidation: degree Fig. 5. Layout of the embankment considered.
of settlement Us versus time factor Tv.

k wo σ′o
[39] cv = = 0.026 m 2 / day
0.434 CR γ w

In addition, when C c C k = 1 the constitutive relationships


incorporated into Plaxis (eqs. [37] and [38]) are equivalent
to those considered in the present solution (eqs. [3] and [5]).
The sand of the upper layer and the embankment material
were modelled as a linear elastic-perfectly – plastic material
(Mohr–Coulomb model) undergoing a drained behaviour.
The Young’s modulus and Poisson’s ratio for these soils are
100 MPa and 0.15, respectively. Moreover, the strength pa-
rameters are ϕ′ = 35° with c′ and ψ null. To evaluate the
initial stress field, the coefficient of earth pressure at rest,
Ko, for both the sand and clay layers was evaluated by the
tent of the embankment compared with the thickness of the expression suggested by Jaky (1944)
clay layer, one-dimensional consolidation can be expected to [40] K o = 1 − sin ϕ′
occur directly under the centre of the embankment. The clay
layer, which is of interest here, is 3 m thick and lies at a Lastly, the lower layer of gravel was assumed to be rigid
depth of 3 m from the ground surface. The water table is lo- and fully permeable. Figure 6 shows the employed finite ele-
cated at the ground surface, and the submerged unit weight of ment mesh, which consists of 15 node elements. The con-
the involved soils is 10 kN/m3. Therefore, the initial effective struction of the embankment was simulated in two steps
vertical stress at the middle height of the clay layer is 45 kPa. corresponding to the loading process schematized in Fig. 7.
The analysis of this case study was first performed using The maximum intensity of loading was qu = 80 kPa.
the finite element code Plaxis (Brinkgreve et al. 2003), in The degree of settlement at the top of the clay layer and the
which the behaviour of the clay was modelled by a Cam- normalized excess pore pressure u q u at the middle height of
Clay type model (soft soil model) with constitutive relation- this layer are plotted against the time factor in Figs. 8 and 9, re-
ships provided by the following nonlinear expressions: spectively. Both of these variables were calculated at the em-
bankment centerline. The results obtained using the proposed
[37] e = e o − 0.434 C c ln( σ′ σ′o ) solution along with the value of cv given by eq. [39] and the
loading process shown in Fig. 7, are also reported in the same
[38] k w = k wo ( σ′o σ′) C c Ck
figures for comparison purposes. As can be seen, the agree-
ment is very good between the present solution and the finite
where e is the void ratio, Cc is the compression index, Ck is element method, in spite of their different levels of complexity.
the permeability index, and e o and k wo are the initial values
of void ratio and coefficient of permeability, respectively. Conclusions
Obviously, the finite element method allows a complete
analysis of the case study considered to be performed; A formulation has been presented for the nonlinear one-
however, it requires a number of soil parameters and com- dimensional consolidation of thin clay layers subjected to
putational costs much greater than the proposed solution, time-dependent loading. It has been shown that when the
which is on the contrary simpler to use and requires only a changes in the coefficient of consolidation are negligible,
knowledge of cv. the differential equation governing nonlinear consolidation
In the Plaxis calculations it was assumed that the clay is can be reduced to a simple diffusion equation (like that of
normally consolidated with compression ratio CR = 0.12, Terzaghi’s linear theory), by using a convenient variable.
shearing resistance angle ϕ′ = 30°, and cohesion intercept However, the boundary and initial conditions, which have
c′ and dilatancy angle ψ both equal to zero. Moreover, to be imposed to achieve the solution, are different from
eo = 0.8, k wo = 3 × 10 − 4 m/day, and C c C k was kept at unity, those considered in Terzaghi’s theory; hence this cannot be
which is a value lying in the range observed by Mesri and used for the problem under consideration. A simple solu-
Rokhsar (1974) for many soils. Owing to this latter assump- tion procedure has been proposed to readily obtain the tran-
tion, the coefficient of consolidation is constant and takes sient response of clay layers to time-dependent loading.
the following value: This procedure makes use of some analytical expressions
© 2007 NRC Canada
Conte and Troncone 723

Fig. 6. The finite element mesh used in the analysis.

Fig. 7. Comparison of the present solution and the finite element Fig. 9. Comparison of the present solution and the finite element
method: loading process. method: degree of settlement Us versus time factor Tv.

Fig. 8. Comparison of the present solution and the finite element


method: excess pore-water pressure u at z = H 2, normalized with
respect to the greatest intensity of loading qu, versus time factor Tv.

derived in this paper in conjunction with the Fourier series.


Comparisons with numerical and analytical methods have
proven the validity of the proposed solution. Moreover, the
presented results have shown that when the external load-
ing is time-dependent, both the degree of consolidation set-
tlement and excess pore pressure calculated using the
nonlinear theory significantly depend on the ratio of the
loading intensity to the initial effective vertical stress.
Therefore, they differ from those predicted by the linear
theory, even though the same coefficient of consolidation
is assumed in the calculations.

References
Barden, L., and Berry, P. 1965. Consolidation of normally consoli-
dated clay. Journal of the Soil Mechanics and Foundation Divi-
sion, ASCE, 91(SM5): 15–35.
Bartholomeeusen, G., Sills, G.C., Znidar i , D., Van Kesteren,
W., Merckelbach, L.M., Pyke, R., Carrier III, W.D., Lin, H.,
© 2007 NRC Canada
724 Can. Geotech. J. Vol. 44, 2007

Penumadu, D., Winterwerf, H., Masala, S., and Chan, D. soft soils. International Journal of Numerical and Analytical
2002. Sidere: numerical prediction of large-strain consolida- Methods in Geomechanics, 26(2): 139–161.
tion. Géotechnique, 52(9): 639–648. Zhuang, Y.C., Xie, K.H., and Li, X.B. 2005. Nonlinear analysis of
Brinkgreve, R.B.J., Broere, W., and Waterman, D. 2003. Plaxis consolidation with variable compressibility and permeability.
2D—Version 8. User’s manual. A.A. Balkema, Rotterdam, Journal of Zhejiang University (Engineering Science), 6A(3):
Netherlands. 181–187.
Burghignoli, A. 1979. Consolidazione monodirezionale e creep
delle argille. Rivista Italiana di Geotecnica, 13(3): 204–213. Appendix A. Solution to eqs. [15], [16], and
Butterfield, R. 1979. A natural compression law for soils (an ad-
vance on e-log p′). Géotechnique, 29(4): 469–480.
[18] when w is kept at unity at the
Carslaw, H.S., and Jaeger, J.C. 1959. Conduction of heat in solids. boundary.
Oxford University Press, Oxford. The differential equation under consideration is
Conte, E., and Troncone, A. 2006. One-dimensional consolidation
under time-dependent loading. Canadian Geotechnical Journal, ∂ 2w ∂w
43(11): 1107–1116. [A1] cv = , at 0 < z < H , t > 0
∂z 2 ∂t
Davis, E.H., and Raymond, G.P. 1965. A non-linear theory of con-
solidation. Géotechnique, 15(2): 161–173. where z and t are the independent variables, w is the unknown
De Boer, R., Schiffman, R.L., and Gibson, R.E. 1996. The origins of function depending on both z and t, and cv is a constant pa-
the theory of consolidation: the Terzaghi–Fillunger dispute. rameter. To achieve the solution to eq. [A1] when w = 1 at the
Géotechnique, 46(2): 175–186. boundary, the following boundary conditions are considered:
Gibson, R.E., England, G.L., and Hussey, M.J.L. 1967. The theory
of one-dimensional consolidation of saturated clays. Géotechnique, [A2] w = 1, at z = 0, t > 0
17(3): 261–273.
∂w
Jaky, J. 1944. The coefficient of earth pressure at rest. Journal of [A3] = 0, at z = H, t > 0
the Society of Hungarian Architects and Engineers, pp. 355– ∂z
358.
Janbu, N. 1965. Consolidation of clay layers based on non-linear together with the initial condition
stress–strain. In Proceedings of the 6th International Conference [A4] w = 0, at t = 0, 0 ≤ z ≤ H
on Soil Mechanics and Foundation Engineering, Montréal, Que.
Vol. 2, pp. 83–87. To reduce eq. [A1] to a separable partial differential equa-
Lancellotta, R. 1995. Geotechnical Engineering. A.A. Balkema, tion, we put
Rotterdam, Netherlands.
Lancellotta, R., and Preziosi, L. 1997. A general nonlinear mathemat-
[A5] w = X +Y
ical model for soil consolidation problems. International Journal of where X is a function of both z and t, and Y is a function
Engineering Sciences, 35(10/11): 1045–1063. only of z. These latter functions satisfy, respectively, the fol-
Mesri, G., and Rokhsar, A. 1974. Theory of consolidation for lowing relations:
clays. Journal of the Geotechnical Engineering Division, ASCE,
100(GT8): 889–904. d 2Y
Mikasa, M. 1965. The consolidation of soft clay—a new consolida-
[A6] =0
dz 2
tion theory and its application. Japanese Society of Civil Engi-
neering, pp. 21–26. [In Japanese.] [A7] Y = 1, at z = 0
Poskitt, T.J. 1969. The consolidation of saturated clay with variable
permeability and compressibility. Géotechnique, 19(2): 234–252. dY
[A8] = 0, at z = H
Schiffman, R.L., Pane, V., and Gibson, R.E. 1984. The theory of dz
one-dimensional consolidation of saturated clays. In Proceedings
of the American Society of Civil Engineers Symposium on Sedi- and
mentation Consolidation Models—Predictions and Validation.
Edited by R.N. Yong and F.C. Townsend. pp. 1–29. ∂2X ∂X
[A9] cv =
Sills, G.C. 1995. Time dependent processes in soil consolidation. ∂z 2 ∂t
In Proceedings of the International Symposium on Compression
and Consolidation of Clayey Soils. Edited by Yoshikuni and [A10] X = 0, at z = 0, t > 0
Kusakabe. pp. 875–890.
Terzaghi, K. 1925. Erdbaumechanik auf Bodenphysikalisher ∂X
[A11] = 0, at z = H, t > 0
Grundlage. Deuticke, Vienna, Austria. ∂z
Townsend, F.C., and McVay, M.C. 1990. SOA: large-strain con-
solidation predictions. Journal of Geotechnical Engineering, [A12] X = −Y , at t = 0, 0 ≤ z ≤ H
ASCE, 116(2): 166–176.
Viggiani, C. 1973. Non-linear one-dimensional consolidation of thick It can be easily shown that function Y satisfying eq. [A6]
clay layers. In Proceedings of the 8th International Conference on with the spatial boundary conditions eqs. [A7] and [A8] is
Soil Mechanics and Foundation Engineering, Moscow, Vol. 4.3, [A13] Y = 1
pp. 119–120.
Yao, D.T.C., Oliveira-Filho, W.L., Cai, X.C., and Znidar i , D. At the same time, the solution to eq. [A9] with the bound-
2002. Numerical solution for consolidation and desiccation of ary and initial conditions eqs. [A10]–[A12] can be derived

© 2007 NRC Canada


Conte and Troncone 725

using a mathematical procedure similar to that developed by 4 ∞ 1


Taylor (1948) to achieve the solution to the equation governing [A15] w = 1 − ∑
π n=1 ( 2n −1)
sin( α n z) exp( − α 2n c v t)
Terzaghi’s one-dimensional consolidation (Terzaghi 1925). The
resulting equation takes the form:
4 ∞ 1 References
[A14] X = − ∑
π n=1 ( 2n −1)
sin( α n z) exp( − α 2n c v t)
Taylor, D.W. 1948. Fundaments of soil mechanics. John Wiley &
Sons, New York.
where α n = ( 2n −1) π ( 2H ). Terzaghi, K. 1925. Erdbaumechanik auf Bodenphysikalisher
Consequently, by substituting eqs. [A13] and [A14] into Grundlage. Deuticke, Vienna, Austria.
eq. [A5] the complete solution to eqs. [A1]–[A4] is given as

© 2007 NRC Canada

Potrebbero piacerti anche