Sie sind auf Seite 1von 9

Proceedings of the Institution of

Civil Engineers
Ground Improvement 163
November 2010 Issue GI4
Pages 207215
doi: 10.1680/grim.2010.163.4.207
Paper 900035
Received 12/10/2009
Accepted 21/07/2010
Keywords: embankments/
geotextiles, membranes & geogrids/
piles

Yan Zhuang
Nottingham Centre for
Geomechanics, University of
Nottingham, UK

Ed A. Ellis
University of Plymouth;
formerly University of
Nottingham, UK

Hai-Sui Yu
Nottingham Centre for
Geomechanics, University of
Nottingham, UK

Plane strain FE analysis of arching in a piled embankment


Y. Zhuang

MSc, PhD,

E. A. Ellis

MA, PhD, CEng, MICE

and H. S. Yu DIC, MSc, DPhil, DSc, CEng, FICE, FIEAust

The results of a series of linearly elastic-perfectly plastic


plane strain finite-element (FE) analyses are reported,
investigating the arching of a granular embankment
supported by pile caps over a soft subsoil. The objective
of the research was to contribute further to generic
understanding of arching in a piled embankment, and the
contribution of tensile reinforcement near the base of
the embankment. The stress state in the embankment
and load on the subsoil were examined. The analyses
demonstrate that the ratio of the embankment height to
the centre-to-centre pile spacing is a key parameter in
determining the behaviour. The potential contribution of
tensile reinforcement near the base of the embankment
was also considered. It was found that arching in the
embankment occurred at relatively small settlement of
the underlying subsoil, whereas much more settlement
was required for sag of the reinforcement to have
significant effect in further reducing the stress on the
subsoil.

long been recognised in the study of soil mechanics (e.g.


Terzaghi, 1943). Terzaghi considered shear stress on the
vertical interfaces originating from the rigid supports at either
side of a trapdoor where support was reduced. This approach
has found some application in the analysis of piled
embankments (Russell and Pierpoint, 1997), where the pile caps
act as rigid supports, and the underlying soft subsoil as the
trapdoor (Figure 1). Other approaches for the design of such
structures include a semicircular arch in the granular fill, as
initially proposed by Hewlett and Randolph (1988) (and
developed by Low et al. (1994) and Kempfert et al. (2004)), and
analogy with backfill over a buried pipe (BSI, 1995). Notably
there is not one generally accepted approach for design (Love
and Milligan, 2003). The theories tend to initially consider
plane strain conditions, which are then modified to account for
the three-dimensional nature of most piled embankments
which use a square (or triangular) layout of pile caps in plan.
This paper considers a plane strain analysis as the logical
starting point which can be extended to three dimensions.

NOTATION
a
pile cap width
h
embankment height
K
earth pressure coefficient ratio of horizontal to
vertical effective stress (9h /9v )
K0
initial earth pressure coefficient at rest
Kp
Rankine passive earth pressure coefficient
k
stiffness of the tensile reinforcement
l
span of the tensile reinforcement
N
number of tensile reinforcement layers
s
centre-to-centre spacing of pile caps
ec , em settlement at the surface of the embankment (see
Figure 1)
r
maximum sag of the tenile reinforcement
s
maximum settlement of the subsoil between the pile
caps (see Figure 1)
c
vertical stress on the pile caps
G
vertical stress in the embankment including the effects
of arching
r
vertical stress loading the tensile reinforcement
s
vertical stress on the subsoil
ji
frictional strength at the reinforcementsoil interface.

The above methods consider that there is sufficient tendency for


the soft subsoil to settle such that arching of the embankment
material will occur. They do not, however, specifically link
arching with the amount of support from the subsoil. An
interesting contrast to this is the concept of a ground reaction
curve (GRC) used to determine the load on a plane strain
underground structure such as a tunnel (Figure 2(a)). Figure 2(b)
shows the form of GRC proposed by Iglesia et al. (1999) based
on geotechnical centrifuge tests. p is the support pressure from
the roof of the underground structure to the soil above. It is
presented in non-dimensional form as p* p/p0 , where p0 is
the nominal overburden total stress at the elevation of the roof
derived from the thickness of overlying soil (and any surcharge
at the ground surface). Variation of p* is plotted against
* /B, where  is the settlement of the roof and B is the total
width of the underground structure.

1. INTRODUCTION
The concept of arching of granular soil over an area where
there is partial loss of support from an underlying stratum has
Ground Improvement 163 Issue GI4

Initially p* 1, but as * increases, arching develops in the


soil above the underground structure. A point of maximum
arching (minimum stress on the structure) is reached where
Iglesia et al. consider a parabolic (similar to semicircular) arch
to form in the soil above the structure. However, as *
increases brittle response is observed as p* increases again to
an ultimate state where shear on vertical planes at the edge of
the structure is used to determine p*. The value of * to reach
the point of maximum arching is stated to be about 2 to 6%.

Plane strain FE analysis of arching in a piled embankment

Delivered by ICEVirtualLibrary.com to:


IP: 134.7.248.132
On: Wed, 27 Apr 2011 04:05:57

Zhuang et al.

207

Embankment
ec
Semicircular
arch

s/2

em

Midpoint

Infill
material
beneath
arch

Centreline

Vertical
interfaces

z
(s a)/2

Pile cap
Soft subsoil

Pile

Figure 1. Piled embankment showing potential arching mechanisms (Terzaghi, 1943; Hewlett and Randolph, 1988), and notation for
geometry and settlement () used in this paper

Ground surface

Centreline

Underground
structure

Midpoint

s/2

B
(a)
p*
Rigid support
from half pile cap

Uniform stress
from subsoil s

Figure 3. Typical finite element mesh (h 5 m, s 2.5 m) and


boundary conditions

Ultimate
state

*
Maximum arching
(b)

Figure 2. Ground reaction curve for underground tunnel


(Iglesia et al., 1999): (a) underground structure; (b) ground
reaction curve (GRC)

This paper initially examines arching in the embankment using


the GRC concept, and the resulting stress on the subsoil. The
additional effect of layers of tensile reinforcement near the
base of the embankment is then also considered.

represent lines of symmetry at the centreline of a support (pile


cap), and the midpoint between supports. Hence there was a
restraint on horizontal (but not vertical) movement at these
boundaries. No boundary conditions were imposed at the top
(embankment) surface, and no surcharge was considered to act
here. The bottom boundary represents the base of the
embankment, which was underlain by a half pile cap (width
a/2) on the left, and subsoil (width (s  a)/2) at the right. The
pile cap was assumed to provide rigid restraint to the
embankment, and the vertical stress in the subsoil supporting
the embankment (s ) was used to control the analysis the
subsoil itself was not actually modelled. The pile cap width (a)
was fixed at 1.0 m and the centre-to-centre spacing (s) was 2.0,
2.5 or 3.5 m. The embankment height (h) was varied using
values in the range 1.0 to 10 m.

2. ARCHING IN THE EMBANKMENT


2.1. Analyses
The finite-element (FE) analyses reported here were undertaken
in plane strain using Abaqus Version 6.6. Figure 3 shows a
typical mesh for the embankment, whose vertical boundaries
208

Ground Improvement 163 Issue GI4

The embankment material was assumed to be granular (and


hence with predominantly frictional strength), and modelled
using the linear elastic and Mohr Coulomb (c9, j9) parameters
shown in Table 1. For s 2.5 m the effect of increasing j9 to
408, or increasing the kinematic dilation angle () to 228 was

Plane strain FE analysis of arching in a piled embankment


Delivered by ICEVirtualLibrary.com to:
IP: 134.7.248.132
On: Wed, 27 Apr 2011 04:05:57

Zhuang et al.

Unit weight:
kN/m3

Initial earth pressure


coefficient

Youngs modulus:
MN/m2

Poissons
ratio

Cohesion intercept:
kN/m2

Friction
angle: 8

Kinematic dilation
angle at yield: 8

17.0

0.5

25

0.20

30

Table 1. Material parameters for embankment fill

also considered. The constitutive model was unsophisticated.


However, the study focused on the behaviour at yield and
variation of geometrical parameters rather than (for instance)
precise analysis of pre-yield behaviour.

However it was decided not to introduce such complexity at


this stage. Nevertheless, the analysis does give the stress at
maximum arching, and the displacement required to reach
this point.

The sequence of analysis was straightforward. First the in situ


stresses were specified. Initially s was specified as the nominal
vertical stress at the base of the embankment to give
equilibrium with the in situ stresses, but this value was then
reduced (generally allowing Abaqus to determine increment
size automatically) to mimic loss of support from the subsoil.
The subsoil in question is generally of low permeability, and
thus this process has direct analogy with consolidation of the
subsoil, which causes arching of the embankment material onto
the pile caps as settlement of the subsoil increases with time.

Figure 4 shows results for the standard soil parameters (Table


1), s 2.5 m, and a variety of embankment heights (h). The
highest (10 m) embankment requires the largest displacement
to reach the point of maximum arching, but even here the
normalised displacement is only slightly larger than 1 %.
However, this value was directly related to the soil stiffness
which had been chosen the value was doubled for an
analysis with half the soil stiffness.

2.2. Ground reaction curve


Figure 4 shows normalised GRCs equivalent to Figure 2(b) but
using slightly different notation. Because the analysis was
controlled by reducing s , corresponding settlement of the
subsoil (at the base of the embankment) increased from zero at
the edge of the pile cap to a maximum value at the midpoint
between pile caps (Figure 1). The maximum value at the
midpoint will now be referred to as s , and is thus slightly
different to the definition in Figure 2(a) (where  is constant).
Data points are shown at the values of s generated by the
automatic incrementation in Abaqus. Thus they become more
closely spaced (in terms of s ) towards the end of the analysis
as plasticity is more prevalent and there is more difficulty in
achieving convergence.
The GRC curve was modelled up to the point of maximum
arching. Using displacement (rather than stress) controlled
analyses it was found that at larger displacements a constant
value of s was observed, rather than the subsequent
increase exhibited in Figure 2(b). It was concluded that the
post-maximum stage of the GRC would only be observed in
the FE analysis if brittle soil behaviour was introduced.

The ultimate normalised stress was in the range 16 to 20% for


h > 3.5 m, but tended to increase rapidly as h was reduced
below this value. These results will be discussed further in a
broader context below.
2.3. Midpoint profile of earth pressure coefficient
The earth pressure coefficient (K 9h /9v ) plotted on a vertical
profile at the midpoint between piles (the right-hand boundary
of the mesh in Figure 3) gives a good illustration of arching
behaviour. Figure 5 shows the profiles plotted with z  vertical
distance upwards from the base of the embankment (Figure 1),
normalised by s. The profiles do not extend to the top of the
embankment for the higher embankments. Values of 0.5(s  a),
0.5s and 1.5s are highlighted on the z axis; and K K0 and
K Kp 3.0 (taking the standard Rankine passive value and
ignoring the small cohesive element of strength) on the K axis.
Subplots (a) to (c) show results for different values of s.
Referring to Figure 5(b) for (z/s) . 1.5, K K0 , and thus has
not been modified by the formation of the arch. For
embankments where (h/s) . 1.5, K increases with depth for
z/s , 1.5, reaching Kp when z  0.5(s  a). Comparing this
with a semicircular arch (Figure 1), the upper limit of the effect
of arching is about three times higher, but the passive limit is
only reached at the inner radius (and below) the arch, where
the infill material is evidently in a passive plastic state.

s/h

10
08
10
06 15
04

20

h 10 m

25

02

35
50

00
00

02

04

06
08
s/(s a): %

65

80
10

12

14

Figure 4. Ground reaction curves for the standard soil


parameters (Table 1), s 2.5 m, and a variety of embankment
heights (h).

Ground Improvement 163 Issue GI4

Potts and Zdravkovic (2008) report that stable arching occurred


when (h/D) . 3.0 for an infinitely long void (i.e. plane strain).
D is the width of the void, which is equivalent to (s  a) in
Figure 1. For the range of values of (s/a) considered here
normalisation by (s  a) rather than s gives values which are
1.4 to 2.0 times larger. Hence (h/s) . 1.5 is equivalent to
h/(s  a) . 2.1 to 3.0, so that the upper limit is consistent with
(h/D) . 3.0.
For embankments where (h/s) , 1.5 there is increasing
tendency for the highest value of K to occur at the surface of
the embankment, initially giving an S-shaped profile, and
then monotonic reduction in K with depth in the embankment

Plane strain FE analysis of arching in a piled embankment

Delivered by ICEVirtualLibrary.com to:


IP: 134.7.248.132
On: Wed, 27 Apr 2011 04:05:57

Zhuang et al.

209

20

at the pile cap rather than the crown (top of arch) for large s
(Hewlett and Randolph, 1988). This trend is supported for
s 2.0 m, where there would be increased tendency for failure
at the crown. Furthermore K is high at and below z 0.5(s-a)
for this case.

h/s 50
175

15

z/s

125
10
K0

2.4. Ultimate stress on subsoil


Figure 6 shows the ultimate stress on the subsoil (s,ult ),
illustrating variation with (h/s). Subplot (a) shows
normalisation of s,ult by h (as in the GRC), whereas (b) shows
normalisation by s.

075
Kp

05
05(s a)
00

20

Figure 6(a) shows that for (h/s) . 1.5, (s,ult /h) reduces slowly
as h increases, but when (h/s) , 1.5, (s,ult /h) increases
rapidly, tending towards 1.0. This behaviour was previously
noted in Figure 4. The additional data shown here (variation of
s and friction and dilation angles) only show two significant
differences: for s 3.5 m, the value at large h is slightly larger,
and the non-zero dilation angle shows less dramatic increase
when (h/s) , 1.5.

h/s 40, 32, 26, 20

15

z/s

2
K
(a)

h/s 14
10

10
K0

08
Kp 06

05

04

05(s a)

00

2
K
(b)

Figure 6(b) shows lines s h, and s,ult 0.5s. The former


is a condition of no arching, whereas the latter is the nominal
weight of the material below the top of a semicircular arch
(Figure 1). For small (h/s) (s,ult /s) is less than 0.5, and when
(h/s)  0.5 the data converge with the no arching line. At
larger h the behaviour will be compared with two different
predictive approaches as described below.

20
h/s 29, 10

10

14

s 20 m
s 25
s 35
s 25, 40
s 25, 22

08

10

s,ult/h

z/s

15

07
05

06

04

04
02

00

2
K
(c)

00

Figures 5(a) and (c) (s 2.0 and 3.5 m, respectively) show


trends of behaviour which are similar to (b). When s 3.5 m
there is some reduction in K at z 0.5(s  a) for the largest h,
perhaps reflecting an increased tendency for failure of the arch

Eq 3:
s 35

08

25

20 m

Eq 1,
s 25 m
06

s,ult/s

for the lowest h. In fact Kp as indicated on the plots neglects


the small cohesion intercept. Thus the nominal value of 3.0
derived from the frictional strength alone can be exceeded in
the model which also includes a small element of cohesive
strength (Table 1). This is particularly the case when the stress
is small (e.g. near the surface of the embankment or
immediately above the subsoil).

Ground Improvement 163 Issue GI4

2
h/s
(a)

Figure 5. Profiles of earth pressure coefficient (K) on a vertical


profile at the midpoint between piles (z measured upwards
from base of the embankment, Figure 1), showing variety of
embankment heights (h): (a) s 2.0 m; (b) s 2.5 m;
(c) s 3.5 m

210

04

02
No arching
s/s h/s

Legend as
subplot (a)

00
0

h/s
(b)

Figure 6. Normalised stress on subsoil at ultimate conditions


(s,ult ) showing variation with (h/s): (a) normalised by h;
(b) normalised by s

Plane strain FE analysis of arching in a piled embankment


Delivered by ICEVirtualLibrary.com to:
IP: 134.7.248.132
On: Wed, 27 Apr 2011 04:05:57

Zhuang et al.

Low et al. (1994) presented equations for the stress on the


subsoil for plane strain arching, based on the concepts of
failure at the crown of the arch or the pile cap originally
proposed by Hewlett and Randolph (1988). Both Hewlett and
Randolph and Low et al. propose that conditions at the crown
of the arch are always critical for plane strain arching. This is
in contrast to three-dimensional arching (for square pile caps
on a square grid), where conditions at the pile cap are more
likely to be critical for a high embankment, particularly if (s/a)
is large.
The plain strain equation for the stress on the subsoil when
there is failure of the arch at the crown presented by Low et al.
can be written as
1

s (A  AB C)h

N=m2

where the dimensionless terms are as follows


A 1  

Kp 1

; B

s Kp  1
; C B1  
2h K p  2

and  a=s

This equation has identical form to the three-dimensional


equation for failure at the crown of the arch by proposed by
Hewlett and Randolph (in the form it was presented by Love
and Milligan (2003)). However, the numerical factors differ in
the two- and three-dimensional versions.
Prediction of subsoil stress from Equation 1 is shown in Figure
6(b) for s 2.5 m. This gives values which significantly exceed
the corresponding FE results as the embankment height
increases.
Figure 6(b) also shows a simplified version of the plane strain
equation for failure of the arch at the pile cap proposed by
Hewlett and Randolph. Assuming s and c (the vertical stress
on the subsoil and pile cap respectively) to be constant the
plane strain equation of vertical equilibrium is
2

c a s s  a hs

N=m

It is then assumed (from analogy with bearing capacity) that


c K 2p s to give:
3

s h
1

s s K 2p a=s 1  a=s

dimensionless

Hewlett and Randolph consider that c is not constant, but


varies with radius from the centre of the semicircular arch. This
yields a formula which gives similar (but slightly lower) subsoil
stress, but which is considerably more complicated.
Results from Equation 3 are plotted for the three values of s in
Figure 6(b). At larger h Equation 3 shows the correct trend of
behaviour, but tends to overestimate s,ult , particularly as s
reduces. As would be anticipated, increased friction angle or
dilation angle tends to reduce s,ult for a given value of s.
Ground Improvement 163 Issue GI4

As previously noted, for plane strain conditions the equations


proposed by Hewlett and Randolph predict that conditions at
the crown of the arch are always critical. Figure 6(b) confirms
this, with Equation 1 (for the crown) plotting above Equation 3
(for the pile cap). However, the values from Equation 1 appear
overly conservative.
In fact it can be shown that unless the embankment is quite
low Equation 1 predicts higher subsoil stress than the
equivalent three-dimensional equation, which seems
anomalous since a plane strain arch would be expected to be at
least as effective as a three-dimensional arch. The terms ABh
and Ch in Equation 1 are both actually independent of h
(which also appears in the denominator of B and hence C).
Thus as h increases, the term Ah can dominate the equation.
In the three-dimensional version of the equation
A (1  )2( Kp 1) (Love and Milligan, 2003). The factor 2 in
the power term arises since the crown of the arch is assumed to
be supported in both orthogonal horizontal dimensions rather
than only one horizontal direction in plane strain (Hewlett and
Randolph, 1988). If Kp 3.0 and  0.4, then A 0.36 in
plane strain, compared with 0.13 in three dimensions, hence
the plane strain equation can predict higher subsoil stress. In
conclusion, it appears that unless the embankment is quite low,
Equation 1 is overly conservative.
Ellis and Aslam (2009a) gave comparison of physical modelling
data with other predictive methods in a three-dimensional
context. It is shown that for low and medium height
embankments the normalised subsoil stress s /(s  a) is
typically about 1.0 for the Hewlett and Randolph and BS 8006
methods (BSI, 1995) for determination of the subsoil stress
beneath piled embankments. For the range of values of (s/a)
considered here, normalisation by s rather than (s  a) gives
values which are 0.5 to 0.7 times the magnitude. Hence
s /s  0.5 to 0.7. The data in Figure 6(b) are somewhat lower
than this, but broadly consistent. Ellis and Aslam showed that
for higher piled embankments BS 8006 does not predict any
increase in s , whereas the Hewlett and Randolph method
predicts increasing s due to failure of the arch at the pile cap,
particularly when (s/a) is large.
2.5. Settlement at the subsoil and surface of the
embankment
Figure 7(a) shows the maximum value of subsoil settlement at
the midpoint between piles (Figure 1) required to reach
ultimate conditions: s,ult (this value has been estimated by
eye from plots such as Figure 4, and thus is somewhat
subjective). The value has been normalised by the clear gap
between pile caps (s  a) so that it is analogous to * (Figure
2(b)). Variation with (h/s) is shown.
The clearest trend is that the normalised displacement to reach
ultimate conditions increases with (h/s), tending to zero when
(h/s)  0.5, also corresponding to the point of convergence
with the no arching line in Figure 6(b). If there is no arching
then no displacement is required to reach this ultimate
condition. As h increases, arching occurs, and the amount of
stress redistribution from the subsoil to the pile cap increases
with h, thus it is not surprising that the amount of
displacement required to achieve ultimate arching conditions
also increases. This observation is also consistent with the

Plane strain FE analysis of arching in a piled embankment

Delivered by ICEVirtualLibrary.com to:


IP: 134.7.248.132
On: Wed, 27 Apr 2011 04:05:57

Zhuang et al.

211

20

variation with (h/s). This is a measure of differential settlement


at the surface of the embankment, which is of considerable
practical importance. For (h/s) . 1.5 the value is 1.0, indicating
no differential settlement. As h reduces, differential settlement
increases, dramatically so for (h/s) less than about 0.75.

Legend as
subplot (b)

s,ult/(s a): %

15

10

3. THE CONTRIBUTION OF TENSILE


REINFORCEMENT

05

3.1. Analyses
There are relatively well-established methods to account for the
load that a sagging tensile component (e.g. a geotextile or
geogrid) can carry in a two-dimensional situation, for instance
assuming that the deformed shape is a catenary or parabola.

00

h/s
(a)
25
s 20 m
s 25
s 35
s 25, 40
s 25, 40

20

(em/ec)ult

To investigate the contribution of horizontal layers of tensile


reinforcement at or near the base of the embankment, some of
the analyses above were repeated including this component.
Either a single layer of reinforcement was considered, at a
small height 0.10 m above the base of the embankment, or
three layers at heights 0.10, 0.40 and 0.70 m. Again the subsoil
was not modelled, and a uniform vertical stress (s ) was
instead used to support the underside of the embankment and
to control the analysis.

15

10

05

h/s
(b)

Figure 7. Settlement results at the subsoil and surface of the


embankment: (a) ultimate settlement of the subsoil at the
midpoint between piles (s,ult ) normalised by the clear gap
between pile caps (s  a); (b) ratio of the settlement at the
top of the embankment at the midpoint between piles (em )
to the equivalent value at the centreline above the pile cap
(ec )

variation with s, which indicates more displacement as s


increases (for a given h) since this also implies increased
redistribution of load from the subsoil to the pile cap.
The absolute magnitude of s,ult /(s  a) is somewhat smaller
than the value of * of 2 to 6% quoted for maximum arching
by Iglesia et al. (1999), which appears to be relevant to
(h/s)  2 to 5 in the reference. However, the values shown in
Figure 7(a) would vary directly in inverse proportion to the
value of Youngs modulus used in these analyses (e.g. if the
Youngs modulus had been reduced by a factor of 2 to better
simulate the secant modulus to failure then the normalised
displacement would be doubled).
The results showed that the ratio of settlement at the top of the
embankment at the midpoint between piles (m , Figure 1) to
the equivalent value in the subsoil (s ) at the point where
ultimate conditions are reached: (em /s )ult , was typically in the
range 0.450.75. These values seem reasonable: the settlement
at the surface of the embankment is less than beneath the arch.
Figure 7(b) shows the ratio of em to the equivalent value at
the centreline above the pile cap (ec , Figure 1) at the point
where ultimate conditions are reached: (em /ec )ult , showing
212

Ground Improvement 163 Issue GI4

In Abaqus truss elements can be used to model components


which carry axial load with no bending stiffness. Such
elements were used for the tensile reinforcement in the
analyses. The reinforcement covered the full width of the mesh,
and was restrained against movement in the same way as the
vertical faces of the mesh (Figure 3) since it was subject to the
same conditions of symmetry. A contact model is required to
join the elements used for the soil and the truss elements used
for the reinforcement. A surface to surface contact was used
in these analyses (Zhuang, 2009), allowing movement at the
interface parallel to the truss elements, but not normal to them.
The friction angle of sliding at the interface was assumed to be
either 08 (perfectly smooth) or 208 (rough, but somewhat less
than the friction angle of the soil).
The truss elements are defined in terms of a material stiffness
(E) and a notional cross-sectional area (A). Stiffness of tensile
reinforcement is normally quoted per metre width, for
example, k x MN/m. The cross-sectional area of the truss per
metre width was based on a nominal thickness of 1 mm, and E
was specified to give the required k.
3.2. Results
Figure 8 shows plots similar to the GRCs in Figure 4, but
without normalisation, for s 2.5 m, and h 3.5 m. Figure
8(a) shows results for a single layer of reinforcement with
stiffness k 6 MN/m, whereas Figure 8(b) shows results for
three layers of reinforcement each with stiffness k 2 MN/m.
In both cases the reinforcementsoil interface friction angle
(ji ) was 20o .
Each plot shows three lines. The GRC line shows the effect of
arching in an embankment without reinforcement from the
results shown in Figure 4. The Embankment with
reinforcement line shows results from similar FE analyses with
the addition of tensile reinforcement modelled using truss
elements as described above. The third line GRC + effect of

Plane strain FE analysis of arching in a piled embankment


Delivered by ICEVirtualLibrary.com to:
IP: 134.7.248.132
On: Wed, 27 Apr 2011 04:05:57

Zhuang et al.

where l is the span of the reinforcement.

60
GRC (embankment
without reinforcement)

50
40

s: kN/m2

If G is the vertical stress in the embankment including the


effects of arching, as characterised by the GRC in terms of s

GRC effect of reinforcement


(prediction, Eq 5)
Embankment with
reinforcement (FE analysis)

30

N=m2

s G  r

20
10
0
0

20

40

60

80
s: mm
(a)

100

120

140

The right-hand side of this equation is the line labelled GRC +


effect of reinforcement. If Equation 4 is appropriate then this
line should be similar to the results of the new analyses
Embankment with reinforcement. It has been assumed that
r s and it was found that assuming that l (s a/2) gave
reasonable results. This is an average of the centre-to-centre
pile spacing and the clear gap between pile caps, and possibly
reflects the observation that although the reinforcement only
carries load over the clear gap, it still carries tension (and
strain) over the width of the pile cap.

60
50

GRC (embankment
without reinforcement)
GRC effect of reinforcement
(prediction, Eq 5)

s: kN/m2

40
30

Embankment with
reinforcement (FE analysis)

20
10
0
0

20

40

60

80
s: mm
(b)

100

120

140

Figure 8. Typical results for embankment reinforced with


single or multiple layers of reinforcement for s 2.5 m and
h 3.5 m: (a) one layer of reinforcement (k 6 MN/m,
ji 208); (b) three layers of reinforcement (each k 2 MN/
m, ji 208)

reinforcement is a theoretical comparison line which will be


described below.
For plane strain parabolic deformation of tensile reinforcement
under uniform load Ellis and Aslam (2009b) propose use of the
following simple equation directly relating the vertical stress
carried by the reinforcement (r ) to the maximum sag in the
reinforcement (r )

 r 21

 3
k r
l l

N=m2

where r is determined from Equation 4, and provided there


has been sufficient settlement of the subsoil G s,ult (Figures
4 and 6).

In Figure 8(a) agreement between the theoretical prediction and


the results of the analysis is good. Ultimately the value of s
reduces to zero, when the reinforcement carries the remaining
embankment load at the point of maximum arching on the
GRC (r G s,ult ). However, note that the settlement at this
point considerably exceeds that at which ultimate (maximum)
arching is initially reached.
Referring to Equation 4, if three layers of reinforcement across
the same span all sag by an equal amount then the total stress
carried by them would be equivalent to the sum of the
stiffnesses k. Thus in Figure 8(b) a theoretical prediction based
on the sum of the values (k 3 3 2 6 MN/m) is shown. The
results of the analysis again show reasonable agreement with
the prediction.
As summarised in Table 2, a total of eight FE analyses of
reinforced embankments are reported here, examining
variation of h, s, k, number of layers of reinforcement (N), and
the reinforcementsoil interface friction angle (ji ).
Figures 9(a) and (b) show the maximum settlement at the base
of the embankment (s ; Figure 1), at the point in the analysis

Parameter which is varied

h: m

Embankment height, h
Pile cap centre-to-centre spacing, s
Reinforcement stiffness, k
Reinforcement interface, ji
No. of reinforcement layers, N 3 k

1.0
3.5
10.0
3.5
3.5
3.5
3.5
3.5

s: m

k : MN/m

ji

2.5
2.5
2.5
2.0
3.5
2.5
2.5
2.5

6
6
6
6
6
12
6
332

0
0
0
0
0
0
20
20

Table 2. Summary of analyses presented in Figure 9 and labels for charts. The parameter which is
varied is highlighted in bold, compared to the standard case (h 3.5 m, s 2.5 m, k 6 MN/m,
ji 0)

Ground Improvement 163 Issue GI4

Plane strain FE analysis of arching in a piled embankment

Delivered by ICEVirtualLibrary.com to:


IP: 134.7.248.132
On: Wed, 27 Apr 2011 04:05:57

Zhuang et al.

213

Eq 6: s 25 m, k 6 MN/m

Figure 9(b) shows the effect of variation of pile cap spacing (s),
reinforcement stiffness (k), and the number of layers of
reinforcement (N) when h 3.5 m. Corresponding predictions
from Equation 6 are shown for the new values of s and k.
Reducing or increasing s has a corresponding effect on r (see
also Figure 6(b)). When s 2.0 m the data point agrees quite
well with Equation 6, but when s 3.5 m Equation 6 is slightly
conservative in the prediction of sag.

h 10 m

150

h 35 m

100

: mm

h1m
50

0
0

10

20

30

r: kN/m2
(a)
Eq 6: s 35 m, k 6 MN/m
250
200

s 25 m, k 6 MN/m
s 25 m, k 12 MN/m

s 35 m

: mm

s 20 m, k 6 MN/m
150

N3

100
k 12 MN/m
s 20 m

50
0
0

10

20

30

r: kN/m2
(b)

Figure 9. Variation of maximum settlement at base of the


embankment (sag in reinforcement), , with stress on
reinforcement (r ): (a) effect of embankment height (h) and
reinforcement interface friction angle (ji ) for s 2.5 m and
k 6 MN/m: (b) effect of pile cap spacing (s), reinforcement
stiffness (k), and number of layers of reinforcement (N) for
h 3.5 m

where s reached zero. This value was virtually identical to the


corresponding maximum sag in the reinforcement (r ).
Variation of this settlement or sag is plotted with the
corresponding load on the reinforcement (r ), which is the
remaining stress beneath the GRC at the point of maximum
arching (r s,ult ). Hence each analysis is represented by a
single data point.
A number of comparison lines are also shown on Figure 9, all
based on the following equation for maximum sag r in the
reinforcement (rearranged from Equation 4)

 1=3
r l
:
r 0 36l
k

where it is again assumed that l (s  a/2).


Figure 9(a) shows the effect of variation of embankment height
(h) and the reinforcement interface friction angle (ji ) for
s 2.5 m and k 6 MN/m. Increasing embankment height
shows quite significant increase in r (see also Figure 6(b)).
However, this only causes a moderate increase in sag as
implied by Equation 6, and the data show reasonable
agreement with the corresponding prediction line (s 2.5 m,
k 6 MN/m). When h 3.5 m, increasing ji to 208 causes a
slight reduction in r and hence sag.
214

Ground Improvement 163 Issue GI4

When k is increased to 12 MN/m (with s 2.5 m and


h 3.5 m) there is little change in r in comparison with the
corresponding analysis with k 6 MN/m (Figure 9(a)), since
the arching action of the embankment and the corresponding
s,ult are essentially unchanged. The data point shows good
correspondence with Equation 6 for k 12 MN/m, but as
implied by the equation this only gives a modest reduction in
sag.
The data point N 3 shows the result for three layers of
reinforcement (distributed throughout the bottom 0.7 m of the
embankment), with total stiffness 3 3 2 MN/m 6 MN/m, and
ji 208 (s 2.5 m and h 3.5 m). The value of r is again
similar to the equivalent analysis in Figure 9(a) with a single
layer of reinforcement. The result shows slightly more sag than
predicted by Equation 6 based on the total stiffness. The
distributed layers are slightly less effective than a single layer
since the higher layers tend to sag less (and hence carry less
load).
4. CONCLUSIONS
The results of a series of linearly elastic-perfectly plastic plane
strain FE analyses to investigate the arching of a granular
embankment supported by pile caps over a soft subsoil have
been reported. The analyses demonstrated that the ratio of the
embankment height to the centre-to-centre pile spacing (h/s) is
a key parameter.
(a) When (h/s) < 0.5 there is virtually no effect of arching:
ultimate conditions are reached almost immediately (with
very small displacement), relative differential settlement at
the surface of the embankment is very large, and the stress
acting on the subsoil is virtually unmodified from the
nominal overburden stress.
(b) When 0.5 < (h/s) < 1.5 there is increasing evidence of
arching: as (h/s) increases, the displacement required to
reach ultimate conditions increases, relative differential
displacement at the surface of the embankment reduces,
and the stress acting on the subsoil reduces compared to
the nominal overburden stress.
(c) When 1.5 <(h/s) full arching is observed: the
displacement required to reach ultimate conditions
continues to increase. There is no differential displacement
at the surface of the embankment, and the stress acting on
the subsoil is considerably reduced in comparison with the
nominal overburden stress. For a high embankment the
stress state is not significantly affected above a height of
1.5s in the embankment.
Furthermore, it has been shown that up to a critical value of
(h/s) the stress on the subsoil is less than 0.5s, approximately
representing the weight of the material below the top of a
semicircular arch. At higher values of (h/s), conditions at the

Plane strain FE analysis of arching in a piled embankment


Delivered by ICEVirtualLibrary.com to:
IP: 134.7.248.132
On: Wed, 27 Apr 2011 04:05:57

Zhuang et al.

pile cap are critical and Equation 3 can be used to


conservatively estimate the stress on the subsoil.
Alternatively, the data can be normalised by the clear spacing
(s  a), for example, Ellis and Aslam (2009a). For the range of
values of (s/a) considered here normalisation by (s  a) rather
than s gives values which are 1.4 to 2.0 times larger. Hence the
critical height for full arching (h/s) 1.5 reported here
corresponds to h/(s  a) in the range 2.1 to 3.0. The lower limit
of this range is consistent with the equivalent value h/
(s  a) 2.0 reported by Ellis and Aslam (2009a), based on the
results of centrifuge tests for three-dimensional arching. The
upper limit is consistent with an equivalent value h/D 3.0 for
arching over a plane strain void reported by Potts and
Zdravkovic (2008).
Further analyses considered the effect of tensile reinforcement
at the base of the embankment. It has been shown that a
simple relationship such as Equation 4 can be used to predict
the contribution of the reinforcement as it sags with reasonable
accuracy for a wide range of pile cap and embankment
geometries.
ACKNOWLEDGEMENTS
Yan Zhuang has been financially supported by a Dorothy
Hodgkin Postgraduate Award and an EPSRC PhD Plus
Fellowship at the University of Nottingham. This support is
gratefully acknowledged.
REFERENCES
BSI (British Standards Institution) (1995) BS 8006 1995. Code
of practice for strengthened/reinforced soils and other fills.
BSI, Milton Keynes, UK.
Ellis EA and Aslam R (2009a) Arching in piled embankments:
comparison of centrifuge tests and predictive methods part
1 of 2. Ground Engineering June: 3438.

Ellis EA and Aslam R (2009b) Arching in piled embankments:


comparison of centrifuge tests and predictive methods
part 2 of 2. Ground Engineering July: 2831.
Hewlett WJ and Randolph MF (1988) Analysis of piled
embankments. Ground Engineering April: 1218.
Iglesia GR, Einstein HH and Whitman RV (1999) Determination
of vertical loading on underground structures based on an
arching evolution concept. Proceedings of the 3rd National
Conference on Geo-Engineering for Underground Facilities.
ASCE, Reston, VA, USA, pp. 495506.
Kempfert HG, Gobel C, Alexiew D and Heitz C (2004) German
recommendations for reinforced embankments on pilesimilar elements. Proceedings of the 3rd European
Geosynthetics Conference. Technische Universitat Munchen,
Germany, pp. 279284.
Love J and Milligan G (2003) Design methods for basally
reinforced pile-supported embankments over soft ground.
Ground Engineering March: 3943.
Low BK, Tang SK and Choa V (1994) Arching in piled
embankments. ASCE Journal of Geotechnical Engineering
120(11): 19171938.
Potts VJ and Zdravkovic L (2008) Finite element analysis
of arching behaviour in soils. Proceedings of the 12th
International Conference of the International Association
for Computer Methods and Advances in Geomechanics,
Goa, India. Indian Institute of Technology, Mumbai,
India, pp. 36423649.
Russell D and Pierpoint N (1997) An assessment of design
methods for piled embankments. Ground Engineering
November: 3944.
Terzaghi K (1943) Theoretical Soil Mechanics. Wiley, New York,
USA.
Zhuang Y (2009) Numerical Modelling of Arching in
Piled Embankments including the Effects of Reinforcement
and Subsoil. PhD thesis, University of Nottingham,
UK.

What do you think?


To discuss this paper, please email up to 500 words to the editor at journals@ice.org.uk. Your contribution will be forwarded to the
author(s) for a reply and, if considered appropriate by the editorial panel, will be published as discussion in a future issue of the
journal.
Proceedings journals rely entirely on contributions sent in by civil engineering professionals, academics and students. Papers should be
20005000 words long (briefing papers should be 10002000 words long), with adequate illustrations and references. You can
submit your paper online via www.icevirtuallibrary.com/content/journals, where you will also find detailed author guidelines.

Ground Improvement 163 Issue GI4

Plane strain FE analysis of arching in a piled embankment

Delivered by ICEVirtualLibrary.com to:


IP: 134.7.248.132
On: Wed, 27 Apr 2011 04:05:57

Zhuang et al.

215

Das könnte Ihnen auch gefallen