You are on page 1of 15

University of Wollongong

Research Online
Faculty of Engineering and Information Sciences -
Faculty of Engineering and Information Sciences
Papers

2012

Vertical drain consolidation with non-Darcian flow


and void ratio dependent compressibility and
permeability
R Walker
University of Wollongong, rohanw@uow.edu.au

B Indraratna
University of Wollongong, indra@uow.edu.au

C Rujikiatkamjorn
University of Wollongong, cholacha@uow.edu.au

Publication Details
Walker, R., Indraratna, B. & Rujikiatkamjorn, C. (2012). Vertical drain consolidation with non-Darcian flow and void ratio dependent
compressibility and permeability. Geotechnique: international journal of soil mechanics, 62 (11), 985-997.

Research Online is the open access institutional repository for the


University of Wollongong. For further information contact the UOW
Library: research-pubs@uow.edu.au
Vertical drain consolidation with non-Darcian flow and void ratio
dependent compressibility and permeability
Abstract
Vertical drains increase the rate of consolidation in soft soils by facilitating faster dissipation of excess pore
water pressure through short, horizontal drainage paths. This paper presents an analytical solution for non-
linear radial consolidation under equal-strain conditions incorporating smear but ignoring well resistance.
Three aspects of non-linearity are considered: (a) non-Darcian flow, (b) a log-linear void-ratio-stress
relationship; and (b) a log-linear void-ratio-permeability relationship. The analytical solution to non-linear
radial consolidation can explicitly capture the behaviour of both overconsolidated and normally consolidated
soils. For non-linear material properties, consolidation may be faster or slower when compared with the cases
with constant material properties. The difference depends on the compressibility/ permeability ratios (Cc/Ck
and Cr/Ck), the preconsolidation pressure and the stress increase. If Cc/Ck < 1 or Cr/Ck < 1 then the
coefficient of consolidation increases as excess pore pressures dissipate, and the corresponding rate of
consolidation is greater.

Keywords
dependent, compressibility, permeability, drain, flow, vertical, darcian, non, consolidation, void, ratio

Publication Details
Walker, R., Indraratna, B. & Rujikiatkamjorn, C. (2012). Vertical drain consolidation with non-Darcian flow
and void ratio dependent compressibility and permeability. Geotechnique: international journal of soil
mechanics, 62 (11), 985-997.

This journal article is available at Research Online: http://ro.uow.edu.au/eispapers/2


Walker, R. et al. (2012). Géotechnique 62, No. 11, 985–997 [http://dx.doi.org/10.1680/geot.10.P.084]

Vertical drain consolidation with non-Darcian flow and void-ratio-


dependent compressibility and permeability
R . WA L K E R  , B. I N D R A R AT NA  a n d C . RU J I K I AT K A M J O R N 

Vertical drains increase the rate of consolidation in soft Les drains verticaux augmentent la vitesse de consolida-
soils by facilitating faster dissipation of excess pore tion dans les sols meubles, en facilitant une dissipation
water pressure through short, horizontal drainage paths. plus rapide de la pression d’eau interstitielle à travers
This paper presents an analytical solution for non-linear des canaux de drainage horizontaux courts. Cette com-
radial consolidation under equal-strain conditions incor- munication présente une solution analytique pour la con-
porating smear but ignoring well resistance. Three as- solidation radiale non linéaire dans des conditions de
pects of non-linearity are considered: (a) non-Darcian contrainte égales, comprenant l’effet smear, mais ne te-
flow, (b) a log-linear void-ratio–stress relationship; and nant pas compte de la résistance de puits. On examine
(b) a log-linear void-ratio–permeability relationship. The trois aspects de non linéarité suivants: (i) écoulement non
analytical solution to non-linear radial consolidation can Darcien, (ii) rapports ratio de vide/contrainte log-liné-
explicitly capture the behaviour of both overconsolidated aire ; et (iii) rapport ratio de vide/perméabilité log-liné-
and normally consolidated soils. For non-linear material aire. La solution analytique pour la consolidation radiale
properties, consolidation may be faster or slower when non linéaire est en mesure de capturer de façon explicite
compared with the cases with constant material proper- le comportement de sols surconsolidés et à consolidation
ties. The difference depends on the compressibility/ normale. Pour les propriétés de matières non linéaires, la
permeability ratios (Cc /Ck and Cr /Ck ), the preconsolida- consolidation peut être plus rapide ou plus lente par
tion pressure and the stress increase. If Cc /Ck < 1 or rapport aux cas à propriétés matérielles constantes. La
Cr /Ck < 1 then the coefficient of consolidation increases différence est tributaire des ratios de compressibilité/
as excess pore pressures dissipate, and the corresponding perméabilité (Cc /Ck et Cr /Ck ), de la pression de pré-
rate of consolidation is greater. consolidation, et de l’augmentation des contraintes. Si
Cc /Ck < 1 on Cr /Ck < 1, le coefficient de consolidation
augmente, les pressions interstitielles excessives se dissi-
KEYWORDS: case history; consolidation; pore pressures; settle- pant et les ratios de consolidation correspondants étant
ment supérieurs.

INTRODUCTION ratio is related to effective stress and permeability by the


Vertical drains increase the rate of consolidation in soft soils following relationships.
by facilitating faster dissipation of excess pore water pressure  
9
through short, horizontal drainage paths. Usually, the simplest e ¼ e0  Cc log (2)
analytical methods used to analyse vertical drain consolidation  09
!
problems simulate a single soil cylinder with constant soil ~
k
properties over a given stress range (Barron, 1948; Hansbo, e ¼ e0 þ Ck log (3)
1981). When material behaviour is non-linear, the simple ~
k0
models are often inadequate. This paper presents analytical
solutions for non-linear radial consolidation under equal- where e is the void ratio; 9 is the effective stress; Cc is the
strain conditions incorporating smear but ignoring well resis- compressibility index; Ck is the permeability index; and e0 ,
tance. For most commonly installed drain lengths in Australia  09 and ~k 0 are the initial values of void ratio, effective stress
and Southeast Asia (less than 18 m), the well resistance is not and permeability respectively before the application of pre-
significant compared with the smear effects (Indraratna & loading.
Redana, 2000). Three aspects of non-linearity are considered: A review of the current literature reveals various attempts
(a) non-Darcian flow, (b) log-linear void-ratio–stress relation- to model the corresponding problem with Darcian flow (e.g.
ship, and (c) log-linear void-ratio–permeability relationship. Basak & Madhav, 1978; Lekha et al., 1998; Indraratna et
In non-Darcian flow, the velocity of flow, v, is related to the al., 2005). Lekha et al. (1998) proposed the average excess
hydraulic gradient, i, by the power law pore pressure, u, for radial consolidation considering log-
linear void-ratio–stress and log-linear void-ratio–permeabil-
v ¼ ~k i n (1) ity relationships as

˜ Th 1  
u ¼  09 
where ~k is the coefficient of permeability under non-Darcian  09  2  Cc =Ck
conditions, and n is the non-Darcian flow exponent. Void " 2Cc =Ck #) (4)
˜ 2Cc =Ck
3 þ 
Manuscript received 12 August 2010; revised manuscript accepted 21
 09
 
March 2012. Published online ahead of print 17 August 2012. 1 ˜
Discussion on this paper closes on 1 April 2013, for further details see ¼1 1þ (5)
p. ii. 2  09
 Centre for Geomechanics and Railway Engineering, Faculty of
Engineering, University of Wollongong, Australia. where ˜ is the change in total stress, Th is the time factor

985
986 WALKER, INDRARATNA AND RUJIKIATKAMJORN
for horizontal consolidation, and  is a smear zone param- where mv is the volume compressibility, and ªw is the unit
eter for Darcian flow. weight of water. When considering material non-linearity,
Equation (4) is a linear function of the time factor Th , the coefficient of consolidation becomes dependent on the
whereby at large values of Th unrealistic negative values of effective stress (under equal strain, the effective stress does
excess pore water pressure are predicted. The excess pore not vary radially). The effective stress can be written as
water pressure should decay to zero: thus equation (4) is  9 ¼  09 þ ˜  W ˜ (8)
unsuitable for estimating u as primary consolidation nears
conclusion. Also, equation (4) is undefined if Cc /Ck ¼ 2. where ˜ is the instantaneous change in total stress, and W
Basak & Madhav (1978) presented a more useful solution is a normalised pore pressure given by
with equation (2), but using a linear relationship between u
permeability and effective stress. Indraratna et al. (2005) W¼ (9)
expressed the excess pore pressure under instantaneous load- ˜
ing as The compressibility of soils previously subjected to higher
  effective stresses (overconsolidated) may increase markedly
8Th0 when the preconsolidation pressure,  p9 , is exceeded. With
u ¼ exp Pav (6a)
 reference to Fig. 1, the compressibility relationships in the
recompression zone ( 9 <  p9 ) and the compression zone
where ( 9 .  p9 ) are described, respectively, by
 
"  1Cc =Ck # 9
1 ˜ e ¼ e0  Cr log (10a)
Pav ¼ 1þ 1þ (6b)  09
2  09
and
Equation (6) is very similar to the linear solution given by    
Hansbo (1981); the main difference is in the parameter Pav :  p9 9
e ¼ e0  Cr log  Cc log (10b)
Pav represents the average between the beginning and end  09  p9
values of the consolidation coefficient. This averaging of
consolidation coefficient over the applied stress increment is The consolidation coefficient involves permeability and vol-
implied in Hansbo’s (1981) solution. Thus equation (6) ume compressibility. Volume compressibility is defined by
simply provides the best choice of consolidation coefficient the relationship
for use with Hansbo’s (1981) equations. Although Indraratna
et al. (2005) recommend using the log-linear void-ratio– 1 @e
mv ¼ (11)
stress relationship (equation (2)) for settlement calculations, 1 þ e0 @ 9
the solution expressed by equation (6) does not reflect the Differentiating equation (10) to find @e/@9 and then substi-
non-linear processes involved with pore pressure dissipation. tuting into equation (11) yields
The proposed model presented herein removes these sim-
plifying assumptions. Hansbo’s (2001) equal-strain solution 0:434Cr
mv ¼ ;  9 <  p9 (12a)
for radial drainage with non-Darcian flow is extended  9ð1 þ e0 Þ
to include the non-linear material properties expressed in
equations (2) and (3). A series solution to the resulting non- and
linear partial differential equation is found, explicitly captur-
ing the variation of permeability and compressibility in the 0:434Cc
mv ¼ ;  9 .  p9 (12b)
consolidation of normally and overconsolidated soil. These  9ð1 þ e0 Þ
solutions can also be used for benchmarking numerical
analysis. The relative change in volume compressibility with effective
stress is thus expressed as
CONSTITUTIVE RELATIONSHIP mv0  9
¼ ;  9 <  p9 (13a)
The horizontal coefficient of consolidation under non- mv  09
Darcian flow, ~ch , can be written as and
~k h  
mv0 Cr  9
~ch ¼ (7) ¼ ;  9 .  p9 (13b)
mv ªw mv Cc  09

e e

e0
⫺Cr
1
1 1
Cc Ck

ef

~ ~ ~
log(σ⬘0) log(σ⬘p) log(σ⬘0 ⫹ Δσ) log(σ⬘) log(kf) log(k0) log(k)

Fig. 1. Void-ratio-dependent compressibility and permeability


VERTICAL DRAIN CONSOLIDATION 987
Substituting equation (10) into equation (3) gives the relative
change in permeability with effective stress as
~k  Cr =Ck
9
¼ ;  9 <  p9 (14a)
~k 0  09

and

~k  ð Cc Cr Þ=Ck  Cc =Ck


 p9 9 rw
¼ ;  9 .  p9 (14b)
~k 0  09  09

Combining equations (7), (8), (13) and (14) results in the rs


following expressions for the stress dependence of ~ch in the
~ ~
recompression and compression zones kh ks

 1Cr =Ck
~ch ˜ W ˜
¼ 1þ  ;  9 <  p9 (15a) re
~ch0  09  09
 ð Cc Cr Þ=Ck
~ch Cr  p9
¼
~ch0 Cc  09
 1Cc =Ck (15b)
˜ W ˜
3 1þ  ;  9 .  p9
 09  09

The value of consolidation coefficient at the start of analysis


is denoted ~ch0 :

DEVELOPMENT OF GOVERNING EQUATION FOR


NON-DARCIAN FLOW
The consolidation coefficient is a function of the coeffi-
cient of volume compressibility, mv, and of the soil per-
meability, which vary according to the effective stress. The
stress dependence of the consolidation coefficient is treated
implicitly and incorporated explicitly in the following.
Vertical drains, installed in a square or triangular pattern,
are usually modelled analytically by considering an equiva-
lent axisymmetric system. Pore water flows from a soil
cylinder to a single central vertical drain with simplified de
boundary conditions. Moreover, this mathematical analysis
considers only the lateral flow, because, for long vertical
Fig. 2. Axisymmetric unit cell
drains the contribution from the vertical consolidation coeffi-
cient (cv ) is insignificant. Therefore the question of aniso-
tropy does not affect the mathematical derivations directly. pore water flow in the smear and undisturbed zones are
For simplicity, the authors have assumed that the compressi- respectively given by (Hansbo, 2001)
bility parameter mv is the same for both smear and the  
1 @us n
undisturbed zone. Fig. 2 shows a unit cell with external v¼~ ks ; n>1 (16a)
radius re , drain radius rw and smear zone radius rs : Depend- ªw @ r
ing on the soil stiffness, the size and shape of the mandrel,
and the installation method, the extent of the smear zone (rs ) and
can usually be estimated as rs ¼ (2 to 3)rm , where rm is the  
1 @u n
equivalent radius of the mandrel (Hansbo, 1981). Ghande- v¼~
kh ; n>1 (16b)
harioon et al. (2010) showed that the driving of the mandrel ªw @ r
does follow the theory of lateral cavity expansion in a non-
linear manner. Weber et al. (2010) showed that a very where ~k h is the undisturbed horizontal permeability for non-
clearly structured smear zone was observed from environ- Darcian flow; ~ k s is the horizontal permeability in the smear
mental scanning electron microscope photographs. The soil zone; u is the excess pore water pressure in the undisturbed
permeability and compressibility parameters decrease non- zone; and us is the excess pore water pressure in the smear
uniformly with the radius within and just outside the smear zone.
zone, as vividly discussed elsewhere by Indraratna & Redana The rate of fluid flow through the internal face of the
(1997, 1998a) and Walker (2006), based on measurements hollow cylindrical slice with internal radius r is
from large-scale consolidation tests. However, in order to
simplify the mathematical formulation of already complex 2rv (17)
equations, reduced lateral soil permeability was assumed as The rate of volume change in the hollow cylindrical slice
an equivalent average (constant) across the smear zone with internal radius r and outer radius re is
annulus. The lateral coefficient of consolidation (ch ) is there-
fore changed proportionately too. This assumption may lead   @
 r2e  r 2 (18)
to some overestimation of the settlement. The velocity of @t
988 WALKER, INDRARATNA AND RUJIKIATKAMJORN
X
1  2 j
where @/@t is the one-dimensional strain rate. For instant- f1=ng j y
aneous loading the strain rate can be expressed as gð yÞ ¼ ny 11= n (24a)
j¼0
j!½ð2 j þ 1Þn  1 N
@ @u
¼ mv (19)
@t @t (The derivation of g(y) is given in the Appendix.)
where u is the average excess pore pressure, and mv is the Equation (24a) can be formulated using a recurrence
volume compressibility of the soil. For continuity, the relationship as
volume changes in equations (17) and (18) can be equated; X
1
rearranging the resulting expression using equations (16) and g ð yÞ ¼ a j ð yÞ (24b)
(19), the pore pressure gradient in the smear and undisturbed j¼0
zones is represented by ny 11= n
" a0 ¼
 #1=n n1
@us r2e @u ~kh y2  2 (24c)
 ð rw ªw Þ11= n
 1 2 y 1=n y ð jn  n  1Þð2 jn  n  1Þ
@y 2~ch @ t ~
ks N a j ¼ a j1
N jnð2 jn þ n  1Þ
(20a)
The average excess pore pressure satisfies the algebraic
and
expression
 1= n  1= n   ð rs ð re
@u r2e @u y2  r2e  r2w u ¼ 2 rus ð rÞdr þ 2 ruð rÞdr
 ð rw ªw Þ11= n
 1 2 y 1=n (25a)
@y 2~ch @ t N rw rs
(20b)
or, in the transformed coordinate system
where N ¼ re /rw , and a change of variable has been made "ð ðN #
such that y ¼ r/rw : (The derivation of equation (20) is given 2 s

in the Appendix.) u¼ 2 yu ð yÞd y þ yuð yÞd y (25b)


N 1 1 s s
By using the binomial expansion, the terms involving y
on the right-hand side of equation (20) can be represented Substituting equation (23) into equation (25) and performing
by a series as the appropriate integrations gives the excess pore pressure
 1=n 1 f1=ng  2 j
(and governing equation) as
y2 X y
1 2 y 1= n ¼ y 1=n
j
(21)  1= n  1= n
N j! N r2e @u ~ch0
j¼0 u ¼ ð rw ªw Þ11=n
  (26)
2~ch0 @ t ~ch
where {x}m , sometimes called the Pochhammer symbol or
rising factorial, is defined by with
( )
f xg m ¼ xð x þ 1Þð x þ 2Þ . . . ð x þ m  1Þ, f xg 0 ¼ 1 1 2g ð N Þ  k1=n [2gð1Þ þ gð1Þð N 2  1Þ]
¼ 2
(22) N 1 þðk1=n  1Þ[ g ð sÞð N 2  s 2 Þ þ 2g ð sÞ]
(27a)
The series expansion performed in equation (21) and others
below are best obtained using computer algebra systems where k ¼ k~h =~k s and g( y) is the product of y and equation
(CAS). For example, entering the text ‘(1-(x^2)/(N^2))^(1/ (24a), integrated with respect to y to give
n) x^(-1/n)’ into the Wolfram Alpha LLC. (2011) website X1 f1=ng j
 2 j
y
yields, after simplification, equation (21). Substituting equa- gð yÞ ¼ n2 y 31= n
tion (21) into equation (20) and integrating (with the bound- j¼0
j!½ð2 j þ 1Þn  1 ½ð2 j þ 3 Þn  1  N
ary conditions us ¼ 0 at y ¼ 1, and u ¼ us at y ¼ s, where (27b)
s ¼ rs /rw ) yields the following pore pressure expressions
Equation (27b) can be formulated using a recurrence rela-
 1=n  1=n
r2e @u ~ch0 tionship as
us ð yÞ ¼ ð rw ªw Þ 11= n

2~ch0 @ t ~ch X
1
g ð yÞ ¼ a j ð yÞ (27c)
!1=n (23a)
~k h j¼0
3 [ g( y)  g(1)]; 1 < y , s n2 y 31= n
~k s
a0 ¼
ð n  1Þð3n  1Þ
 2 (27d)
and y ð jn  n  1Þð2 jn  n  1Þ
 1=n  1=n a j ¼ a j1
N jnð2 jn þ 3n  1Þ
r2 @u ~ch0
uð yÞ ¼ ð rw ªw Þ11= n
 e
2~ch0 @ t ~ch The pore pressure at any point in the soil can now be related
8 !1=n 9 to the average excess pore water pressure by substituting
< ~ = equation (26) into equation (23). The resulting expressions
kh
3 : g ð yÞ  g ð sÞ þ [ g(s)  g(1)];; 1 < y , N for pore water pressure in the smear and undisturbed zones
~
ks
are
(23b) !1= n
u ~
kh h i
where g(y) is the integral of equation (21) with respect to y, us ð yÞ ¼ gð yÞ  g ð1Þ (28a)
given by  ~k s
VERTICAL DRAIN CONSOLIDATION 989
8 !1= n 9
< ~ = W 1 n  Ł1 n
u kh   ~
a0 ¼
uð yÞ ¼ g ð yÞ  g ð s Þ þ gð sÞ  g ð1Þ ; 1 n
: ~
ks  1 1þ j n
ð j  nÞð j  Æ=Ck Þ  90 W  Ł1þ j n
(28b) ~
aj ¼ ~
a j1 1þ
jð1 þ j  nÞ ˜ W j n  Ł j n
Usually the expressions for excess pore water pressure would
involve explicit functions of time (Hansbo, 1981, 2001); (33c)
however, as shown below, the solution of equation (26) gives
consolidation time as a non-invertible function of u: The For the special case of Darcian flow (i.e. derive equation
corresponding time is determined when a suitable value for (32) with n ¼ 1) the recurrence relationships become
u is specified.  
W
~
a0 ¼ ln
Ł
ANALYTICAL SOLUTION OF THE GOVERNING   1
Æ  09
EQUATION ~
a1 ¼ 1  1þ ð W  ŁÞ
Equation (26) can be rewritten as Ck ˜
   1
@W ~ch0 ð j  1Þð j  Æ=Ck Þ  09 W j  Łj
@ T~ ¼ n (29) ~
aj ¼ ~
a j1 1 þ
W ~ch j2 ˜ W j1  Ł j1
(33d)
where T~ is a modified time factor defined by
Equation (32) can now be written in the concise notation as
 
Th0 ˜ n1 T~ ¼ F ½ W , Cr , 1
T~ ¼ 8 n (30a) (34)
 rw ªw
~ch0 t Equation (34) is valid during recompression: that is, when
Th0 ¼ (30b)
4r2e W > Wp , where Wp corresponds to the preconsolidation
pressure. Wp is calculated from
Using a Taylor power series, expansion of equation (15a)  
about the point W ¼ 0 gives (in the recompression zone)  09  p9
Wp ¼ 1  1 (35)
 ð1Cr =Ck Þ ˜  09
~ch0 ˜
¼ 1 þ
W n~ch  09 The time factor required to reach the preconsolidation
  (31) pressure, T~p , is determined by substituting Wp into equation
X f1  Cr =Ck g j
1
 09 j n
3 1þ ðW Þ (34): hence
j! ˜
j¼0
T~p ¼ F ½ W p , Cr , 1 (36)
Substituting equation (31) into equation (29), and integrating
with the initial condition W ¼ 1 at t ¼ 0, results in the Now the power series representation of equation (15b) is
following time factor-normalised pore pressure relationship substituted into equation (29) and solved with the initial
in the recompression zone condition W ¼ Wp at T~ ¼ T~p : The resulting expression for
 ð1Cr =Ck Þ consolidation in the compression phase is
˜  ð Cc Cr Þ=Ck
T~ ¼  1 þ Cc  p9
 09 T~ ¼ F ½ W , Cc , W p  þ F ½ W p , Cr , 1
2 3 Cr  09
  j
X1 f1  C =C g
r k j  9 (37)
34 ðW  1Þ5
0
1þ j nþ1

j¼0
j!ð j  n þ 1Þ ˜
Equation (37) can be used for normally consolidated soils
(32) by putting Cr ¼ Cc and Wp ¼ 1. The mv0 value in T~ should
always be calculated using the recompression index, Cr :
To avoid writing such large expressions as the right-hand Substituting a given normalised excess pore pressure into
side of equation (32), a shorthand notation is used, whereby equations (34) and (37), the resulting value of T~ can be
a function F, depending on parameters Æ, Ł and W, is found. The time required to reach the specified degree of
described by consolidation is then found from equation (30). Equation
 ð1Æ=Ck Þ (32) is undefined for integer values of n; however, values
˜ very close to integer values give the appropriate consolida-
F ½ W , Æ, Ł ¼  1 þ tion times (e.g. use n ¼ 1.0001, not n ¼ 1).
 09
2 3 When Cr /Ck ¼ Cc /Ck ¼ 1 (i.e. ~ch does not change during
  j
X 1 f1  Æ=C g
k j  9 consolidation), equation (34) is numerically equivalent to
34 1þ
0 ðW j nþ1
Ł j nþ1 Þ5 Hansbo’s (2001) non-Darcian radial consolidation equation.
j¼0
j!ð j  n þ 1Þ ˜ For the case of Darcian flow n ¼ 1, the corresponding value
of T~Darcy is given by
(33a)
8Th0
Equation (33a) can be formulated using a recurrence rela- T~Darcy ¼ (38)

tionship as
 ð1Æ=Ck Þ X1 The  parameter can then be any of those derived for
˜
F ½ W , Æ, Ł ¼  1 þ ~
aj (33b) Darcian flow conditions. The following expressions for 
 09 l¼0 were derived for
990 WALKER, INDRARATNA AND RUJIKIATKAMJORN
(a) a smear zone of constant reduced permeability (Hansbo, Table 1. Parameters used to produce consolidation curves shown
1981) in Fig. 3
 
N ˜ = 09
 ¼ ln þ ðkÞln ð sÞ  0:75 (39a) Cc /Ck chf /ch0
s
0.500 5.758 2.600
(b) similar to (a), but plane-strain conditions (Indraratna & 0.500 3.840 2.200
Redana, 1997) 0.500 2.240 1.800
"   # 0.500 0.960 1.400
3 3
2 1 2ð N  sÞ 2ð N  sÞ 1.000 1.000 1.000
¼k 1 2  þ
3 N 3ð N  1Þ N 2 3ð N  1Þ N 2 1.115 6.000 0.800
1.263 6.000 0.600
(39b) 1.471 6.000 0.400
(c) a smear zone with parabolic variation of permeability in
the smear zone (Walker & Indraratna, 2006)
  2  
N 3 kð s  1Þ s consolidation would occur. Analysis including the non-linear
 ¼ ln  þ 2 ln pffiffiffi
s 4 s  2ks þ k k soil properties is particularly relevant for soils where rel-
atively high values of ˜ = 09 will lead to large changes in
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffi pffiffiffiffiffiffiffiffiffiffiffi! (39c)
sð s  1Þ kðk  1Þ kþ k1 ch , and thus potentially large deviations in degree of con-
 ln pffiffiffi pffiffiffiffiffiffiffiffiffiffiffi solidation compared with calculations based on constant soil
2ð s 2  2ks þ kÞ k k1
properties. For chf /ch0 values close to unity, each 10% in-
where k is the ratio of undisturbed permeability to the crease in chf /ch0 decreases the time to reach 90% consolida-
permeability at the drain/soil interface. Equation (39c) tion by approximately 5% (compared with the ideal case of
was simplified by ignoring insignificant terms, and can be chf /ch0 ¼ 1). Each 10% decrease in chf /ch0 increases the time
used when n . 10; otherwise the full expression to reach 90% consolidation by approximately 10%.
presented by Walker & Indraratna (2006) is more
appropriate, usually for smaller drain spacing. When
n ¼ 1 and Cr /Ck ¼ Cc /Ck ¼ 1, equation (34) is numeri- SETTLEMENT
cally equivalent to Hansbo’s (1981) radial consolidation Primary consolidation settlements, r, result from changes
equation. in effective stress. Once excess pore water pressures are
found (i.e. based on W ¼ u=˜ ), then the settlements can
be calculated using the following expressions. If the initial
and final effective stresses ( 09 and  09 þ ˜ ) fall in the
EFFECT OF Cc /Ck ON NORMALLY CONSOLIDATED recompression zone then settlements are calculated with
SOIL

For normally consolidated soils, if Cc /Ck , 1, then ch HCr ˜
r¼ log 1 þ ð1W Þ (40a)
increases over time, and the consolidation rate is faster. If 1 þ e0  09
Cc /Ck . 1 then ch decreases over time, and the consolidation
rate is slower. The degree to which consolidation is slowed
where H is the depth of soil. If the initial and final effective
or hastened is controlled by the normalised factor ˜ = 09 :
stresses fall in the compression zone then settlements are
Taken together, Cc /Ck and ˜ = 09 give an indication of the
given by
total change in ~ch over the consolidation period, in accor-
dance with equation (15). In practice, Cc /Ck ranges between

HCc ˜
0.5 and 2.0 (Berry & Wilkinson, 1969; Mesri & Rokhsar, r¼ log 1 þ ð1  W Þ (40b)
1974), with Ck taken from the empirical relation Ck ¼ 0.5e0 : 1 þ e0  09
Figure 3 gives consolidation curves based on the ratio
between the final and initial consolidation coefficients (chf If the initial effective stress is less than the preconsolidation
and ch0 respectively) for Darcian flow. Although the same pressure,  p9 , and the final effective stress is greater than  p9
change in ch can be obtained with different Cc /Ck and then settlements in the recompression zone are the same as
˜ = 09 values, the resulting difference in degree of consoli- in equation (40a); settlements in the compression zone are
dation is minimal. The actual values of Cc /Ck and ˜ = 09 then expressed as
used in producing Fig. 3 are given in Table 1. Using differ- 

ent parameter values from those in Table 1 will result in H ˜
r¼ ð Cr  Cc ÞlogðOCRÞ þ Cc log 1 þ ð1  W Þ
slightly faster or slower consolidation. Faster consolidation 1 þ e0  90
would occur if ch /ch0 was more than unity; otherwise, slower
(40c)
0
Degree of consolidation: %

where OCR is the overconsolidation ratio, defined by


20
 9p
40 OCR ¼ (41)
 09
60
Chf /Cho ⫽ 2·6 The total primary settlement, r1 , can be calculated by
2·2
80 1·8
1·4 1·0
putting W ¼ 1 in the above equations.
0·8 0·6 0·4 0·2
100
0·1 1 10
Tho /μ
DEGREE OF CONSOLIDATION BASED ON
SETTLEMENT AND PORE PRESSURE
Fig. 3. Consolidation curves for normally consolidated soil under The degree of consolidation, as determined by pore
Darcian flow conditions pressure dissipation, is simply given by
VERTICAL DRAIN CONSOLIDATION 991
u 100
Uh ¼ 1  ¼1W (42)
˜
80
The degree of consolidation, based on settlement, is written
as
Δσ/σ⬘0 ⫽ 8
r 60 4
U hs ¼ (43)

Uhs: %
2
r1
1

_
The relationship between Uhs and W depends on the stress 40
history of the soil.
In equation (40), when the stress range is either comple-
tely in the recompression zone or in the compression zone, 20
the term Uhs is then related to W by
 
log 1 þ ð˜ = 09 Þð1  W Þ 0
U hs ¼ (44a) 0 20 40 _ 60 80 100
log½1 þ ð˜ = 09 Þ Uh: %
(a)
For other cases, if  9 <  p9 then 100
 
log 1 þ ð˜ = 09 Þð1  W Þ
U hs ¼
ð1  Cc =Cr Þ logðOCRÞ þ Cc =Cr log½1 þ ð˜ = 09 Þ 80
(44b)

If  9 .  p9 then 60 Δσ/σ⬘0 ⫽ 8
4
Uhs: %

ð1  C c =Cr Þ logðOCRÞ  2
_

þCc =Cr log 1 þ ð˜ = 09 Þð1  W Þ 40


U hs ¼ (44c) 1
ð1  Cc =Cr Þ logðOCRÞ
þCc =Cr log½1 þ ð˜ = 09 Þ
20
The degrees of consolidation based on pore pressure and
settlement are different (Leroueil et al., 1990; Indraratna et
0
al., 2005). For equation (44a) Uh lags behind Uhs depending 0 20 40 _ 60 80 100
on ˜ = 09 , as shown in Fig. 4. When determining the Uh: %
degree of consolidation for normally consolidated soils by (b)
settlement data, it is important to note that the effective Fig. 5. Comparison between degree of consolidation based on
stress in the soil will be less than expected, particularly settlement and pore pressure for overconsolidated soil
during the middle stages of consolidation, if Uh is assumed (OCR 1.5): (a) Cc /Cr 2; (b) Cc /Cr 10
to be equal to Uhs : A large variety of behaviour can occur
for overconsolidated soil, as shown in Fig. 5, depending on
the value of Cc /Cr , OCR and ˜ = 09 : Dissipation of excess overconsolidated soil, emphasises the need for accurate
pore pressure is fast during the recompression stage, because determination of the soil stress history, and the care needed
Cr is low relative to Cc , while the resulting settlements are when specifying construction milestones based on degree of
relatively small compared with the ultimate settlement. This consolidation.
means that Uh will be greater than Uhs during recompression,
but Uhs may or may not become greater than Uh during the
compression stage. The many parameters involved, and the APPROXIMATION FOR ARBITRARY LOADING
subsequent large variety of behaviour relating Uhs to Uh for The consolidation behaviour expressed by equation (37) is
valid only for instantaneous loading. Unfortunately, including
100 arbitrary time-dependent loading such as Conte & Troncone
Cc /Cr ⫽ 1
OCR ⫽ 1 (2009), or even a simple ramp load (Walker & Indraratna,
2009), will make equation (29) non-separable and, to the
80
authors’ knowledge, analytically unsolvable. However, arbi-
Δσ/σ⬘0 ⫽ 8 trary loading can be simulated by subdividing a continuous
4 loading function into a finite number of instantaneous step
60 2
1
loads (see Fig. 6). The only restriction is that average excess
Uhs: %

pore pressure cannot become negative (u . 0). For example,


_

40 slightly decreasing loads associated with submergence of fill


can be modelled, but the swelling associated with preload
removal (caused by dissipation of negative pore pressure)
20 cannot be modelled with the consolidation equations pre-
sented.
For the mth loading stage, as shown in Fig. 6, at T~ m the
0 load increases from ˜ m1 to ˜ m : The loading stage ends
at T~ mþ1 : The excess pore water pressure at the end of the
0 20 40 _ 60 80 100
Uh: %
last increment, um , is known (e.g. in the first loading step
Fig. 4. Comparison between degree of consolidation based on um ¼ 0). The pore pressure after the load application, uþm , is
settlement and pore pressure for normally consolidated soil given by
992 WALKER, INDRARATNA AND RUJIKIATKAMJORN
Δσ
Actual load (d) Step 4: calculate T~ in small increments of W, with
equation (47), until T~ . T~ mþ1 :
(e) Step 5: calculate umþ1 : Either use the last W value
obtained in step 4, or interpolate between the last two
Δσm values of W found in step 4.
Δσm⫺1 ( f ) Step 6: repeat steps 3 to 5 for each loading stage.
Piecewise (g) Step 7: convert the time factor values obtained in above
load
steps to time values using equation (30).

~ In the steps just described, ˜ changes, and in the mth


~ ~
Tm Tm⫹1 T
loading stage is equal to ˜ m : For cases where soil proper-
(a) ties vary with depth, the soil profile can be divided into
u sublayers and the above process performed for each sublayer.
The above steps have been implemented in a VBA/spread-
Actual load sheet program that can be downloaded from http://www.uow.
edu.au/eng/research/geotechnical/software/index.html.

u⫹
m
ILLUSTRATIVE EXAMPLE
u⫺
m A pore pressure and settlement analysis has been per-
formed on a soil/drain system with the following properties:
rw ¼ rs ¼ 0.07 m; re ¼ 1.4 m; Cc ¼ 0.7; Cr ¼ 0.175; Ck ¼
~
Tm
~
Tm⫹1 0.875;  09 ¼ 10 kPa;  p9 ¼ 10, 20, 30 kPa; n ¼ 1.001, 1.3;
(b) ªw ¼ 10 kN/m3 ; e0 ¼ 1; H ¼ 1 m. No smear zone is mod-
elled. The resulting average pore pressure and settlement
Fig. 6. Schematic diagram of piecewise loading in terms of plots are shown in Fig. 7. There are a few salient points to
loading rate combined with resulting excess pore water pressure note from this figure. For the two overconsolidated soils
( p9 ¼ 20, 30 kPa), the change from recompression to com-
uþm ¼ um þ ˜ m  ˜ m1 (45) pression can be observed during the first ramp loading stage.
At the preconsolidation pressure, the slope of the excess
The normalised pore pressure at the beginning of the load pore pressure plot sharply increases, reflecting the slower
increment, W þm , is then described by rate of consolidation during compression. Just prior to the
preconsolidation pressure being reached, the dissipation of
uþm
W þm ¼ (46) pore pressure in the highly overconsolidated soil
˜ m ( p9 ¼ 30 kPa) actually exceeds that generated by the load
application; the pore pressure reduces despite load still being
The initial conditions for solution of equation (29) have now applied. This occurs because Cr /Ck , 1, resulting in an
been found: W ¼ W þm at T~ ¼ T~ m : There are three cases to increasingly rapid consolidation reaching a maximum just
consider in the solution of equation (29). If the preconsoli- prior to the preconsolidation pressure. The difference be-
dation pressure has been exceeded in previous loading steps, tween the Darcian flow of Fig. 7(a) and the non-Darcian
then the normalised pore pressure in the mth loading step is flow of Fig. 7(b) is small for the highly overconsolidated
governed by soils. For the normally consolidated soil ( p9 ¼ 10 kPa), the
 ð Cc Cr Þ=Ck difference is illustrated during the first ramp loading stage.
 
þ C c  p9 For non-Darcian flow the pore pressure curve is slightly
~ ~
T ¼ T m þ F W , Cc , W m (47a)
Cr  09 flatter, because, owing to the power law flow relationship,
the higher pore pressures result in faster flow. The settlement
If the preconsolidation pressure has not been exceeded in plots illustrate the importance of minimising the disturbance
previous load steps, then the expressions for normalised pore caused by vertical drain installation, which can to some
pressure in the recompression and compression zones are extent destroy any existing structure in the soil. Hence the
  preconsolidation pressure may be lowered, which, as shown
T~ ¼ T~ m þ F W , Cr , W þm (47b) in Fig. 7, causes greater settlements for the same pressure
increase. This example shows that, with the equations pre-
and sented above, almost any primary consolidation problem
involving radial drainage can be modelled (provided the
 ð Cc Cr Þ=Ck
Cc  p9 effective stress increases with time). When more than one
T~ ¼ T~ m þ F ½ W , Cc , W p  þ T~ soil layer is present, the analysis can simply be repeated
Cr  09 (47c) with different soil properties (Walker & Indraratna, 2009).
 
¼ T~ m þ F W p , Cr , W þm
LABORATORY VERIFICATION
The consolidation behaviour can now be described at any
The large-scale consolidometer at the University of Wol-
point. The following list of steps describes the process of
longong is a steel cylinder 450 mm in diameter and 950 mm
constructing a graph of pore pressure against time. The
high, with the ability to monitor the pore pressure and
process is best automated in a computer program.
settlement response of soils under load. Since 1995, the
(a) Step 1: approximate the arbitrary loading by a finite consolidometer has been used by University of Wollongong
number of instantaneous step loads. researchers to study various aspects of vertical drain con-
(b) Step 2: convert the loading times to time factor values, T~ solidation (Indraratna & Redana, 1995, 1998a, 1998b; In-
using equation (30). draratna et al., 2002, 2004). Fig. 8 shows a schematic
(c) Step 3: calculate W þm using equations (45) and (46). diagram of the large-scale consolidometer. The cylindrical
VERTICAL DRAIN CONSOLIDATION 993
100 Pulley
Excess pore pressure: kPa

Line connecting to mandrel


80
Load Hoist
60
LVDT
40 σ⬘p: kPa To vacuum pump
10 20 30 Pressure gauge
20 To data logger Loading piston

0 Top plate
To air jack compressor
0·1 Pressure
chamber
Movable plate
Settlement: m

0·2
30 Porous material
20 (Two half cylinders)
0·3
Steel cell
10
Soil 2 mm Teflon sheet
0·4
0 1 2 3 4 5 Vertical drain
Th0 ⫽ ch0t /4r 2e
(a) Piezometer points

100
Excess pore pressure: kPa

80
Load
60 Fig. 8. Schematic diagram of large-scale consolidation apparatus
σ⬘p: kPa (Indraratna & Redana, 1998a, with permission from ASCE)
40
10 20 30
20
transformer) transducer is placed on top of the piston to
0 monitor surface settlement. Pore water pressures are moni-
tored by strain gauge type pore pressure transducers installed
through small holes in the steel cell at various positions in
0·1
the soil. Transducers are easily located on the cell periphery,
or can be placed within the clay by using small-diameter
Settlement: m

0·2 stainless steel tubes. The LVDT and pore pressure transdu-
30 cers are connected to a PC-based data logger.
20 The soil is subjected to an initial preconsolidation pres-
0·3
10 sure (usually  p9 ¼ 20 kPa), until the settlement rate becomes
negligible (strain rate , 1 mm/day). The load is then re-
0·4
0 1 2 3 4 5 moved, and a single vertical drain is installed using a
Th0 ⫽ ch0t /4r 2e rectangular steel mandrel that acts as a protector for the
(b) drain and a guide during penetration. Sand compaction piles
may be installed with a circular pipe mandrel. Depending on
Fig. 7. Non-linear radial consolidation for non-Darcian flow the purpose of the test, vertical and horizontal samples may
exponent: (a) n 1.001; (b) n 1.3 be cored to investigate the effect of drain installation on soil
properties. Once the drain has been installed, the required
loading sequence is applied to the sample.
cell consists of two stainless steel sections bolted together The consolidation test described below was conducted by
along the two joining flanges. Top and bottom drainage can Rujikiatkamjorn (2006) to verify the radial consolidation
be facilitated by placing a geotextile on the cell base and model of Indraratna et al. (2005) expressed by equation (6).
clay surface. If only radial drainage is required, the per- The model of Indraratna et al. (2005) determines the best
meable geotextile is replaced by an impervious plastic sheet. value of ch to use in Hansbo’s (1981) consolidation equation
The clay is thoroughly mixed with water to ensure full with reference to Cc , Ck and the stress range. The soil
saturation, and the resulting slurry is placed in the consoli- consisted of reconstituted alluvial clay from the Moruya
dation cell in layers approximately 20 cm thick. It is not floodplain. The clay size particles (,2 m) accounted for
necessary to fill the entire consolidometer, since an internal about 40–50% of the soil specimen. For this Moruya clay,
‘riser’ can be used to transfer loads from the loading piston the water content and plastic limits were 45% and 17%
to the shortened sample. The ring friction expected with respectively. The saturated unit weight was 17 kN/m3 : In two
such a large height/diameter ratio (1.5–2) is almost elimi- separate tests, the soil was subjected to an initial preconsoli-
nated by using an ultra-smooth Teflon membrane around the dation pressure,  p9 ¼ 20 kPa and 50 kPa for five days. The
cell boundary (friction coefficient less than 0.03). Surcharge load was removed, a 100 mm 3 4 mm band drain was
loading with a maximum capacity of 1200 kN can be centrally installed, and the preconsolidation loads were re-
applied by an air jack compressor system by way of a rigid applied. Drainage in the vertical direction was prevented.
piston 50 mm thick. Vacuum loading with a maximum Once settlements became negligible, the load was increased
capacity of 100 kPa can be applied through the central hole by 30 kPa and 50 kPa respectively (i.e. ˜ ¼ 30, 50 kPa).
of the rigid piston. An LVDT (linear variable differential Settlements were monitored after this point, with the soil
994 WALKER, INDRARATNA AND RUJIKIATKAMJORN
consolidating in the compression range. The properties of ˜ = 09 ¼1.5 and 1 produces ratios of final to initial consoli-
the soil–drain system are given in Table 2, based on the dation coefficient of 1.39 and 1.28 respectively, for Test 1 and
smear zone with constant reduced permeability. Only for the Test 2. Fig. 3 showed that for such values of chf /ch0 the
purpose of analysis, Darcian flow was considered (i.e. difference between the non-linear equations and the Hansbo
n ¼ 1.001). The degree of consolidation based on settlement (1981) equation are small. This explains the small difference
was calculated and compared with laboratory data, the between the analysis based on the proposed relationship and
proposed model, and Indraratna et al. (2005). The compari- that of Indraratna et al. (2005). Larger and more significant
son of the three approaches is shown in Fig. 9. discrepancies between the two approaches would be expected
Figure 9 shows good agreement between the measured and for combinations of Cc /Ck and ˜ = 09 producing chf /ch0
predicted values of degree of consolidation for both the values much greater or much less than one.
proposed non-linear equations and Indraratna et al. (2005). As
mentioned above, in this case the method of Indraratna et al.
(2005) will give the same results as Hansbo (1981). The APPLICATION TO A CASE STUDY AND THE
proposed equations give a slightly better match in the early IMPLICATIONS OF PARAMETER n
stages of consolidation; the difference is less for Test 2 than At the site of the Second Bangkok International Airport
for Test 1. Both methods are expected to converge to 100% (SBIA), several embankments were constructed over soft clay
consolidation at longer times, as both are based on the same stabilised by vertical drains and vacuum pressure, combined
value of ultimate settlement. As the consolidation coefficient with surcharge loading. The surface settlement and excess
is increasing during consolidation (Cc /Ck , 1), by using an pore water pressure at the middle of two embankments, TV2,
average value of ch as in Indraratna et al. (2005), settlement is incorporating vacuum loading and vertical drains are analysed
expected to be overpredicted in the initial stages of consolida- here. The basic soil parameters were reported by Indraratna et
tion, as in Fig. 9. The combination of Cc /Ck ¼ 0.64 and al. (2005). The initial in situ soil properties are assumed to be
the same for both Darcian and non-Darcian flow analysis. A
summary of the properties used for analysis is given in Table
Table 2. Parameters used in analysis (Indraratna et al., 2005) 3. Soil layers are divided into 1 m thick sublayers (0.5 m for
when required), with stresses and pore pressure calculated at
Parameter Test 1 Test 2 sublayer midpoints. Surcharge load reduces with depth ac-
. . cording to a Boussinesq stress distribution under the centre of
Cc 0 29 0 29
Ck .
0 45 .
0 45
a trapezoidal embankment with 40 m base width and 36 m
Influence radius, re : m 0.45 0.45 crest width. Vacuum pressure load factor reduces from unity
Equivalent drain radius, rw : m 0.066 0.066 at surface to zero at 15 m depth. Linear load graphs in Fig.
Smear zone radius, rs : m 0.2 0.2 10 are discretised into small step loads every 1.744 days.
N ¼ re /rw 6.79 6.79 Geometry/smear properties for all layers are: kh /ks ¼ 2,
s ¼ rs /rw 3.02 3.02 re ¼ 0.5 m, rw ¼ 0.025 m, rs ¼ 0.1 m. The analysis in detail
Initial horizontal permeability, kh0 : 3 1010 m/s 4.4 4 can be found in the VBA/spreadsheet mentioned above. Fig.
kh /ks 1.5 1.5 10 presents the loading history with both measured surface
Initial void ratio, e0 1 .
0 95 settlements at the embankment centreline and pore water
ch0 : 3 103 m2 /day 1.20 2.68
Initial height, H: m .
0 925 .
0 87
pressures 3 m below the embankment centreline.
Preconsolidation pressure,  p9 : kPa 20 50 To study the effect of non-Darcian flow, the predicted
Load, ˜ : kPa 30 50 settlements and pore pressures are calculated based on n
values varying between 1 and 1.5 (vacuum pressure is
treated as an equivalent surcharge). For settlement predic-
tions (Fig. 10(b)), when n ¼1, a good agreement between
0
the predicted and measured results can be obtained. When n
20
increases, the predicted settlements overestimate the meas-
Calculated from laboratory ured data. The predicted and measured pore pressures are
Test 1 settlement data
40 Proposed model
shown in Fig. 10(c). As expected, the pore pressure dissi-
Uhs: %

pates rapidly when n increases. Within 60 days, the agree-


Indraratna et al. (2005)
ment between field observations and predictions is better
_

60
according to Darcian flow. However, beyond 60 days, the
80
Test 2 prediction with non-Darcian flow (n ¼ 1.5) is more realistic.

100
0 20 40 60 MODEL LIMITATIONS AND FUTURE
Time: days
RECOMMENDATIONS
Fig. 9. Comparison between non-linear model and Indraratna et Although the proposed analytical model is capable of
al. (2005) determining degree of consolidation based on both pore

Table 3. Parameters used in analysis for embankment TV2

Layer Unit weight:  p9 : kPa Cr Cc Ck e0 kh0 :


kN/m3 3 1010 m/s

0.0–2.0 m 16 58 0.06 0.37 0.9 1.8 90.3


2.0–8.5 m 15 45 0.08 1.6 0.4 2.8 38.1
8.5–10.5 m 15 70 0.05 1.7 1.2 2.4 18.1
10.5–12.0 m 16 80 0.03 0.95 0.9 1.8 7.68
 Constant within a layer.
VERTICAL DRAIN CONSOLIDATION 995
can still be used effectively, for example under rail tracks,
Surcharge as demonstrated by Indraratna et al. (2010).
40 (c) Lateral soil displacements based on elliptical cavity
Loading: kPa

expansion theory (ECET) have been recently analysed


0 (Ghandeharioon et al., 2010); however, for simplicity, the
Vacuum mathematical formulations related to ECET have not
been incorporated in the current analysis.
⫺40 (d ) The non-Darcian flow equation adopted in this study is
similar to Hansbo (2001) and Sathananthan & Indraratna
(2006). However, the scope of possible flow is limited by
40 80 120
Time: days
the value of n and the type of algebraic expression.
(a) (e) Well resistance is ignored in the equations, as the authors
have shown elsewhere that unless the vertical drains
exceed 20 m or so in length, the well resistance is
insignificant compared with the smear effects (Indraratna
Field & Redana, 2000).
Settlement: m

0·4 n ⫽ 1·001 ( f ) Indraratna & Redana (1997) proved, using large-scale


n ⫽ 1·2
n ⫽ 1·5
consolidation tests, that even within the smear zone, the
lateral permeability kh and hence the lateral coefficient of
0·8 consolidation ch can vary. Subsequently, Walker &
Indraratna (2006) introduced a parabolic fit to this
variation. However, for simplicity of the mathematical
formulations, the authors have used a constant kh within
40 80 120 the smear zone, but it is a reduced value (an equivalent
Time: days
(b)
average) compared with the higher kh in the outer
undisturbed zone. Use of a parabolic fit may provide a
more comprehensive model, but the associated cumber-
some mathematical solutions may not necessarily give a
Pore pressure: kPa

0 much greater accuracy.


(g) The model assumes full saturation and steady-state flow
(both Darcian and non-Darcian). The unsaturated condi-
⫺20 tions at the time of installation of relatively dry drains
and the occurrence of air gaps (occluded air) at the drain/
soil interfaces (Indraratna et al., 2004) have not been
⫺40 considered in this analysis. The current solutions will be
further enhanced if the effects of soil/drain interfaces for
40 80 120
Time: days partially saturated conditions are incorporated.
(c) (h) In situ (undisturbed) soils usually possess a characteristic
structure, whose behaviour can be significantly different
Fig. 10. Embankment TV2: (a) loading history; (b) surface from that of the same material in a reconstituted state, as
settlements at centreline; and (c) pore pressure at 3 m depth explained below.
below centreline
(i) Soil structure creates a material that is initially quite
stiff at relatively low stress levels.
(ii) When a soil with structure is remoulded/reconsti-
pressure dissipation and strain, it has some inherent limit- tuted, usually there is a reduction in the voids ratio.
ations, either through the assumptions made, or through (iii) During virgin yielding, structured soil is generally
simplifications applied to facilitate the mathematical formu- more compressible than the reconstituted soil (Liu &
lations and solutions in this analytical approach. Carter, 2002).
Some of these limitations are highlighted below. A short preconsolidation period in the laboratory may cause
minimal effect on soil settlements and excess pore pressure
(a) The equations do not capture the strain-rate effects dissipation. The use of reconstituted soil parameters to
directly, as the model is a stress-based approach, and does predict field behaviour may result in the underestimation of
not incorporate creep effects. For a given time– soil settlement and excess pore pressure. The selection of
settlement curve, the rate of dissipation of excess pore soil parameters (reconstituted or in situ soil) should be based
pressure tends to be slower when considering creep that on soil conditions to obtain accurate predictions of soil
constitutes part of the total time-dependent settlement. settlement and excess pore pressure.
However, extension of this work is currently being carried
out at the University of Wollongong adopting a large-
strain radial consolidation theory. CONCLUSION
(b) The proposed model does not include the effects of An analytical solution to non-linear radial consolidation
vertical consolidation through cv , because, for long problems has been presented and verified against a large-
vertical drains (say exceeding 12 m or so) with relatively scale laboratory test. The evolution of pore pressure with
close spacing (1.0–1.5 m), the radial consolidation is time described by equation (37) is valid for both Darcian
predominantly governed by ch : Therefore the question of and non-Darcian flow, and can capture the behaviour of
anisotropy (ch /cv or kh /kv ratio) is not a major concern in overconsolidated and normally consolidated soils. Consolida-
this analysis. However, for short vertical drains with tion may be faster or slower when compared with the cases
larger drain spacing, the contribution from cv will with constant material properties for non-linear material
become increasingly relevant, and this is to be captured properties. The difference depends on the compressibility/
as future work where short prefabricated vertical drains permeability ratios (Cc /Ck and Cr /Ck ), the preconsolidation
996 WALKER, INDRARATNA AND RUJIKIATKAMJORN
pressure ( p9 ) and the stress increase (˜ = 09 ). If Cc /Ck , 1 Dividing equation (53) by equation (54) gives
or Cr /Ck , 1, then the coefficient of consolidation increases f1=ng j y 2 jð1= nÞþ1
as excess pore pressures dissipate, and the corresponding
consolidation is faster. If Cc /Ck . 1 or Cr /Ck . 1, then the
a j ð yÞ j!N 2 j 2 j  ð1=nÞ þ 1
¼ (55)
coefficient of consolidation decreases as excess pore pres- a j1 ð yÞ f1=ng j1 y 2ð j1Þð1= nÞþ1
sures dissipate, and the corresponding consolidation becomes ð j  1Þ!N 2ð j1Þ 2ð j  1Þ  ð1=nÞ þ 1
slower. For the case where Cc /Ck ¼ 1, the solution is iden- a j ð yÞ f1=ng j ð j  1Þ! N 2ð j1Þ
tical to Hansbo (2001). The equations presented give an ¼
a j1 ð yÞ f1=ng j1 j! N2j
analytical solution to non-linear radial consolidation that can (56)
be used to verify purely numerical methods. Many vertical y 2 jð1= nÞþ1 2ð j  1Þ  ð1=nÞ þ 1
drain problems where vertical drainage is negligible and 3 2ð j1Þð1= nÞþ1
y 2 j  ð1=nÞ þ 1
effective stresses always increase can be analysed with the
approximation for arbitrary loading. a j ð yÞ 1=n þ j  1 y 2 2 jn  n  1
¼ (57)
a j1 ð yÞ j N 2 2 jn  1 þ n
a j ð yÞ y 2 ð jn  n  1Þð2 jn  n  1Þ
ACKNOWLEDGEMENTS ¼ 2 (58)
a j1 ð yÞ N jnð2 jn þ n  1Þ
The authors would like to express their gratitude to the
Australian Research Council and Coffey Geotechnics (Syd-
ney and Brisbane) for supporting this area of research over a
long period of time. The first and third authors’ PhD studies NOTATION
at the University of Wollongong (UOW) under the super- Cc compression index
Ck permeability change index
vision of Professor Buddhima Indraratna (second author)
Cr recompression index
were sponsored through UOW’s Australian Postgraduate ~ch horizontal coefficient of consolidation for non-Darcian flow
Award and International Postgraduate Research Scholarship chf final value of horizontal consolidation coefficient
scheme respectively, with further financial support by ~ch0 initial horizontal coefficient of consolidation for non-
Queensland Department of Main Roads during that time. Darcian flow
de diameter of influence area
e void ratio
APPENDIX e0 initial void ratio
Derivations of equations (20a) and (20b) ef final void ratio
F function
By equating equations (17) and (18), the radial flow rate is
g function
assumed to be equal to the rate of volume change of soil mass in the
H height of soil; drainage length
vertical direction. Therefore
i hydraulic gradient; integer
  @ j integer
2rv ¼  r2e  r 2 (48) ~
@t k non-Darcian coefficient of consolidation
~
kh undisturbed non-Darcian horizontal permeability
Substituting equations (16) and (19) gives ~
ks smear zone non-Darcian horizontal permeability
"   #1= n
 r2e  r 2 mv ªw @u m integer
@us
¼ ; n>1 (49a) mv coefficient of volume compressibility
@r 2r~k s @t mv0 initial value of volume compressibility
"   #1= n N ratio of influence radius to drain radius
@u  r2e  r 2 mv ªw @u n non-Darcian flow index
¼ ; n>1 (49b)
@r 2r~k h @t OCR overconsolidation ratio
Pav averaging factor to account for changing coefficient of
Substituting N ¼ re /rw and y ¼ r/rw in equations (49a) and (49b) consolidation
yields equations (20a) and (20b). r radial coordinate
re radius of influence area
rm equivalent radius of mandrel
rs radius of smear zone
Derivation of g(y)
rw radius of drain
From equation (20), g(y) can be expressed as
s ratio of smear zone radius to drain radius
ð X1 f1=ng  2 j Th time factor for horizontal consolidation
y
g ð yÞ ¼ y 1= n
j
(50) Th0 horizontal time factor calculated from initial parameters
j! N
T~
j¼0
modified time factor
1 f1=ng ð
X j T~Darcy Darcian time factor
g ð yÞ ¼ y 2 j1= n (51) T~ m modified time factor at application of mth instantaneous
j¼0 j!N 2 j
load
X
1 f1=ng
j y 2 jð1= nÞþ1 T~p modified time factor at preconsolidation pressure
g ð yÞ ¼ (52) t time
j¼0 j!N 2j 2 j  ð1=nÞ þ 1
U degree of consolidation
Uh average degree of consolidation in horizontal direction
Let
Uhs degree of consolidation calculated with settlement data
X
1 f1=ng
j y 2 jð1= nÞþ1 u pore water pressure
a j ð yÞ ¼ (53) um pore pressure immediately before application of mth
j¼0 j!N 2 j 2 j  ð1=nÞ þ 1
instantaneous load
Then uþm pore pressure immediately after application of mth
instantaneous load
ny 1ð1= nÞ us smear zone pore pressure
a 0 ð yÞ ¼ (24b)
n1 u average excess pore pressure
f1=ng j1 u0 initial excess pore pressure
y 2ð j1Þð1= nÞþ1
a j1 ð yÞ ¼ 2ð j1Þ 2ð j  1Þ  ð1=nÞ þ 1
(54) W normalised pore pressure
ð j  1Þ!N Wp normalised pore pressure at preconsolidation pressure
VERTICAL DRAIN CONSOLIDATION 997
v velocity of flow Indraratna, B. & Redana, I. W. (2000). Numerical modelling of
y transformed integration variable vertical drains with smear and well resistance installed in soft
Æ function parameter clay. Can. Geotech. J. 37, No. 1, 132–145.
 non-Darcian radial consolidation parameter Indraratna, B., Bamunawita, C., Redana, I. W. & McIntosh, G.
ªw unit weight of water (2002). Keynote paper: Modelling of prefabricated vertical
 strain drains in soft clay and evaluation of their effectiveness in prac-
@/@t volumetric strain rate tice. Proc. 4th Int. Conf. on Ground Improvement Techniques,
Ł function parameter Kuala Lumpur, 47–62.
k ratio of undisturbed permeability to permeability at drain/ Indraratna, B., Bamunawita, C. & Khabbaz, H. (2004). Numerical
soil interface modelling of vacuum preloading and field applications. Can.
 smear zone parameter for Darcian flow Geotech. J. 41, No. 6, 1098–1110.
r settlement Indraratna, B., Rujikiatkamjorn, C. & Sathananthan, I. (2005).
 total stress Radial consolidation of clay using compressibility indices and
˜ change in total stress varying horizontal permeability. Can. Geotech. J. 42, No. 5,
 average total stress 1330–1341.
9 effective stress Indraratna, B., Rujikiatkamjorn, C., Adams, M., & Ewers, B.
 09 initial effective stress (2010). Class A prediction of the behaviour of soft estuarine soil
 p9 preconsolidation stress foundation stabilised by short vertical drains beneath a rail
track. Int. J. Geotech. Geoenviron. Engng ASCE 136, No. 5,
686–696.
Lekha, K. R., Krishnaswamy, N. R. & Basak, P. (1998).
REFERENCES Consolidation of clay by sand drain under time-dependent
Barron, R. A. (1948). Consolidation of fine-grained soils by drain loading. J. Geotech. Geoenviron. Engng ASCE 124, No. 1,
wells. Trans. ASCE 113, 718–742. 91–94.
Basak, P. & Madhav, M. R. (1978). Analytical solutions of sand Leroueil, S., Magnan, J.-P. & Tavenas, F. (1990). Embankments on
drain problems. J. Geotech. Engng ASCE 104, No. 1, 129– soft clays. Chichester, UK: Ellis Horwood.
135. Liu, M. D. & Carter, J. P. (2002). Structured Cam Clay model. Can.
Berry, P. L. & Wilkinson, W. B. (1969). The radial consolidation of Geotech. J. 39, No. 6, 1313–1332.
clay soils. Géotechnique 19, No. 2, 253–284, http://dx.doi.org/ Mesri, G. & Rokhsar, A. (1974). Theory of consolidation for clays.
10.1680/geot.1969.19.2.253. J. Geotech. Engng ASCE 100, No. 8, 889–904.
Conte, E. & Troncone, A. (2009). Radial consolidation with vertical Rujikiatkamjorn, C. (2006). Analytical and numerical modelling of
drains and general time-dependent loading. Can. Geotech. J. 46, soft clay foundation improvement via prefabricated vertical
No. 1, pp. 25–36. drains and vacuum preloading. PhD thesis, University of Wol-
Ghandeharioon, A., Indraratna, B. & Rujikiatkamjorn, C. (2010). longong, New South Wales, Australia.
Analysis of soil disturbance associated with mandrel-driven Sathananthan, I. & Indraratna, B. (2006). Laboratory evaluation of
prefabricated vertical drains using an elliptical cavity expansion smear zone and correlation between permeability and moisture
theory. Int. J. Geomech. ASCE 10, No 2, 53–64. content. J. Geotech. Geoenviron. Engng ASCE 132, No 7. 942–
Hansbo, S. (1981). Consolidation of fine-grained soils by prefabri- 945
cated drains. Proc. 10th Int. Conf. Soil Mech. Found. Engng, Walker, R. (2006). Analytical solutions for modeling soft soil
Stockholm, 677–682. consolidation by vertical drains. PhD thesis, University of
Hansbo, S. (2001). Consolidation equation valid for both Darcian Wollongong, New South Wales, Australia.
and non-Darcian flow. Géotechnique 51, No. 1, 51–54, http:// Walker, R. & Indraratna, B. (2006). Vertical drain consolidation
dx.doi.org/10.1680/geot.2001.51.1.51. with parabolic distribution of permeability in smear zone.
Indraratna, B. & Redana, I. W. (1995). Large-scale, radial drainage J. Geotech. Geoenviron. Engng ASCE 132, No. 7, 937–941.
consolidometer with central drain facility. Aust. Geomech. 29, Walker, R. & Indraratna, B. (2009). Consolidation analysis of a
103–105. stratified soil with vertical and horizontal drainage using the
Indraratna, B. & Redana, I. W. (1997). Plane-strain modeling of spectral method. Géotechnique 59, No. 5, 439–450, http://dx.
smear effects associated with vertical drains. J. Geotech. Geoen- doi.org/10.1680/geot.2007.00019.
viron. Engng ASCE 123, No. 5, pp. 474–478. Weber, T. M., Plotze, M., Laue, J., Peschke, G. & Springman, S. M.
Indraratna, B. & Redana, I. W. (1998a). Laboratory determination (2010). Smear zone identification and soil properties around
of smear zone due to vertical drain installation. J. Geotech. stone columns constructed in-flight in centrifuge model tests.
Geoenviron. Engng ASCE 124, No. 2, 180–185. Géotechnique 60, No. 3, 197–206, http://dx.doi.org/10.1680/geot.
Indraratna, B. & Redana, I. W. (1998b). Development of the smear 8.P.098.
zone around vertical band drains. Ground Improve. 2, No. 4, Wolfram Alpha LLC. (2011). Wolfram|Alpha. http://www.
165–178. wolframalpha.com (accessed 5 August 2011).

You might also like