(F. Schlosser, P. Delage (Auth.), Peter M. Jarrett
(F. Schlosser, P. Delage (Auth.), Peter M. Jarrett
(F. Schlosser, P. Delage (Auth.), Peter M. Jarrett
Peter M. Jarrett
Department of Civil Engineering,
Royal Military College of Canada,
Kingston, Ontario, Canada
and
Alan McGown
Department of Civil Engineering,
University of Strathclyde,
Glasgow, Scotland, U.K.
Preface ix
INTRODUCTION
PREDICTION EXERCISE
Review of Session on Case Histories and Full Scale Trials ............ 339
RESEARCH NEEDS
PARTICIPANTS
INDEX
Index 635
PREFACE
worked together was quite outstanding and went a long way to ensuring that
the scientific objectives of the workshop were met. The extension of the
working cooperation to the social aspects of life was equally outstanding
leading for example to an impromptu international "soccer" match amongst
participants that showed up some unexpected yet quite prodigious skills.
With such spirit and congeniality it is felt that the additional objectives
of the NATO Scientific Affairs Division for the establishment of inter-
national understanding and cooperation amongst scientists were also ful-
filled at this workshop.
To reach the position where the scientific group could meet in such produc-
tive circumstances and have a book produced from the proceedings demands
considerable effort from a number of people and financial and material
support. The Scientific Director for the meeting and the editors of this
volume wish to acknowledge with sincere thanks the following:
Mr. J.E. DiPietrantonio Qf RMC for his considerable work and skill
in preparing, repal.rl.ng and mounting of drawings in many of the
final edited papers.
1. INTRODUCTION
Soil has been used since the most ancient civilizations as a
construction material. However, due to insufficient tensile strength,
designers had to improve its resistance by using mechanical processes
(compaction, draining), chemical processes (stabilization), or by in-
clusion of resisting elements (reinforcement).
In the past, natural inclusions have most often been used to improve
the mechanical properties of soils and structures straw in clayey
soils to develop construction material; palms or branches at the base
of structures founded on soft soils and so on. It is also interesting to
mention the reinforcement of natural slopes provided by plant roots in
the nature.
During the past twenty years, the important development of Reinfor-
ced Earth, and the concept of reinforced soil as a construction mate-
rial, introduced by its inventor H. Vidal, in the sixties, have
contributed to the birth of a new area of soil improvement soil
reinforcement. This area is based on a generalization of the "reinforced
soil" concept, including various techniques, {or slopes and embankments,
retaining walls, foundations, dams .••
The concept of soil reinforcement is based on the existence of a
strong interaction between the soil and the inclusion. The most common
interaction is friction, but passive pressure may also be mobilized.
Frictional interaction requires good mechanical properties of t~e soil,
particularly in terms of friction angle and drainage; granular soils
are best adapted for this kind of reinforcement. Passive pressure is
generally associated with anchors, which needs more sophisticated in-
clusions. It can be mobilized in all types of soils, including saturated
fine-grained soils, where frictional properties are poor.
The number, type, and arrangement of the inclusions may be quite
variable. Depending on the type of the inclusion, two extreme cases may
be considered (Schlosser et aI, 1983).
1) a "uniform inclusion", for which the soil-reinforcement interac-
tion can develop at any point along the inclusion. In this case, a
relatively high and uniform density of reinforcements will result in a
new composite material called "Reinforced Soil". The behaviour of the
reinforced soil mass can be investigated through laboratory testing
considering a representative sample of the new composite material. This
concept is illustrated in figure 1.
P. M. Jarrett and A. McGown (eds.), The Application of Polymeric Reinforcement in Soil Retaining Structures, 3"'{)5.
© 1988 by Kluwer Academic Publishers.
4
• Composite
Uniform
__- - - - - - - - 1 inclusion
inclusion
I---------------~ Anchor
1---------./
Reinforced Earth wall Ladder wall
I) - REINFORCED SOIL 2L MULTI ANCHORAGE SYSTEM
@,.. PERIODICAL REINFORCEMENT
.. ./ . · . . i
Granular
material Geotextile
--,
© - ISOLATED REINFORCEMENT
~
forcement uniform composite
Density
reinforcement
multi _
periodical re i nforced anchorages
soi I systems
rods, the back-fill material located between this facing and the ancho-
rages may apparently behave as a solid, sustaining small deformations,
as compared to the displacements of the wall. Coyne (1945) describes
some applications of this patent: the work presented in figure 2b is a
dyke submitted to tides, which may even be submerged by big waves
internal seepage may then occur, with 1,50 or 2 m hydraulic gradients,
and the structure behaves as a dam; good rocky material was used as a
fill material between the ties ; this structure supported 0,50 m settle-
ments without any problems, because of good articulations between the
facings. Coyne also mentions a 200 m long quay wall in Brest, and the
construction of 10 to 20 m high retaining structures, such as the side-
walls of an overflow in Mareges, a 10 m high coffer-dam at the same
location, and a 14 m high dam on the river Laurenti (Pyrenees) ; in the
case of dams, the ladder wall may constitute either the up-stream or the
down-stream facing. In all cases, good rocky material is placed in
contact with the ties.
E
o
10
..,.
a Ordinary
backfi II material
@ Concrete
BREAKAGE
.
30
.
20 --- E
-
u
0"3:::: 2
2cml
j
10
(Aluminium
.. foil
2 3 4 5 6
CTI
5300
High CT,
Low 0'2
CD Active zone
® Resis!ant zone r/R
- 700 L-JL..-.L-_ _L-----,-.L-__':-::------,L---:-.L
L--------r--------~ 1 9/0 2/3 1/3 k'O 1/3 2/3 9/01
Backfill material of
good quality
strip_
2.3 Geotextiles
The use of geotextiles in earthworks for reinforcement and separation
at ,the base of an embankment on soft soil started approximately at the
same period as the Reinforced Earth early development. In fact, the
first paper dealing with such an application has been published in 1969
(Vautrain and Puig).
13
JL"
7
_ 6 \
Ribbed reinforcement
.,c
] 5
~
o GRAVEL
04 • T
Y=21 kN/m 3
c
o JL = n.
~ <p =46° (soil/soil)
...
';: 'i' =27,5° (soil/reinforcemen
~2
&.
0.1
_
---"4-
Smooth reinforcement
----e__ tg<p
<t
_ - - -- - - - - - - --:::::!:=-=.--:.::=:-----___ - - tg'i'
O'L----~----2~----~3r-----~4c-----~----~6~---
Height of embankment above the reinforcement - H (m)
(kPa)
. "
.....
100
CT v 1>- PRESSURE CELLS
dr= ~-ANCHOR
50 20cm
ero
Distance d (em)
2816 70 100 200
2.4 Grids
T~first reinforced soil retaining structure uSing grids as
reinforcements was constructed in 1974 on Interstate Route 5, near
Dunsmuir, California (FORSYTH, 1978). One year before in 1973, the
California Transportation Laboratory developed a large direct shear
device in order to test the pullout resistance of different
reinforcement systems (smooth strips, ribbed strips, bars, bar mats).
The purpose was to find a reinforcement system which could enable the
use of granular backfill material containing a large percentage of fine-
grained material. It was found that the best system-was the grid or bar
mat, which provides a relatively linear reinforcement, wisthstanding
large pull-out forces, thanks to the passive thrust mobilized against
the transversal bars. However, compared with the pure frictional
interaction reinforcements, i.e. strips, these bar mats require large
displacements, 5 cm and more, to fully mobilize the pull-out resistance
(CHANG et al., 1977).
Figure 9a shows the welded bar mat tested and used by Cal trans in the
construction of the first "mechanically stabilized embankment" at
Dunsmuir. This structure was approximately 120 m long and consisted of
two walls, 6 m high. The vertical facings were made from rectangular and
long precast concrete elements 3.75 m X 0.6 m X 0.2 m. in size (figure
8c) . The bar mats, which were 1.2 m wide and 3 to 4.5 m in length, were
attached to the facing elements by inversion of the two-bar yoke through
precast holes in the facing panels. These two prethreaded bars were
bolted into position, inducing some interesting prestressing of the
reinforcement. The construction was very similar to Reinforced Earth.
As indicated by Forsyth (1978), it was "anticipated that the bar- mat
mode of reinforcement would have significant economic advantage in
certain areas of the state of California where high quality backfill
material is not readily available".
This argument was considered and put forward by VSL when promoting in
1980 an equivalent retaining system, called Retained Earth, in which the
same , type of welded bar-mat was used.
As shown by Schlosser et al. (1983, 1985), the bar-mat interaction
mechanism is complex and involves both friction along the longitudinal
bars and passive thrust again the transversal bars. For small soil-
reinforcement displacements « 0.5 cm) there is initially a mobiliza-
tion of the friction along the longitudinal bars. For larger displace-
15
.1 I.Gem
FIGURE 9. Type of welded bar mat and panel used by California trans-
portation (Cal trans)
16
FRICTION
---+-
Sand", =40°
Depth =3m
1000
~ Two discs bars
PRESSURE + One disc bars
~+
400
+
200
Head displacement
5 10 15 20 25 8(mm)
100
Complete mobilization
( " h;"""
I GRAVE L I
Displacement 8 (cm)
o'L-------~5------~1~0------~15~----~2~0~----~2~5·
@ PERFORATED
SHEET
@ UNIAXIAL
GEOGRID
(5
0::
- -
....
- -
Roll width
Junction
£:
® BIAXIAL
GEOGRlp
~
c:
(5 Rib
0::
I.. 3.80cm .1
FIGURE 11. Typical geometry of perforated sheet and uniaxial and biaxial
geogrids (Bonaparte et al., 1984)
18
_100
z
-'" Metallic bar mat.
LLI failed by slippage
u
cr 75
0
lJ..
I-
:::>
0
50
..J
..J
:::>
a.
Netlon ER 2 (Tensor)
25 Failed by
tensile break
FIGURE 12. Pull-out force/displacement curves for a metallic bar mat and
a plastic grid
19
Potential failure
i-- surface
® ~ CO ~~
~ ~
r
\
\
\
\,
'\
"'-
JEWELL'S 01 REeT
SHEAR TEST
FIGURE 13. Differences between Jewell's direct shear test and pullout
test
21
Figure 14 shows the first job carried out with this material
(Leflaive et al., 1983). The problem was to repair a slide in a cliff of
chalk and to restore a path at the top of the cliff. About 300 n~ of
Texsol were placed, at a slope angle of 60° and with difficult access
conditions.
Path ta be restored
1
Texsol embankment
Thread reel
6m
CHALK
Placement of the material
6m
o
~200
00
00
w
a:
t-
OO o Non - reinforced
...J
« SLIPPAGE
.Y:f'
~
060"
~ 100
(l> x9(f>
z
~ A 120°
Fibers length : 4cm
Jd: max
200 300
NORMAL STRESS (kPo)
Table II shows cohesion values for reinforced sand, obtained for dif-
ferent types of inclusions (fibers, small plates, continuous filaments),
for the same percentage of inclusion by weight (about 0.2 %). This table
also indicates the different soil-inclusion interactions involved.
Corrugated
membrane
I
I
,.,--,---------
/' I I
I I I
, I I
\ I
\ I
, .... I
Compressive
l
stress
I
I
I
I f+,~--!.'-''---'1 Thoust
\ i
' .... - ..I_-1---------
I
ACTIVE: RESISTANT
ZONE i ZONE
® Behaviour mechanism
L_
~
e
v
\ Soil nailing
/
",_ / Existing wall
I i'
, I: \I \
• I I
FIGURE 17. First soil nailed wall, reinforced earth wall and rock bol-
ted wall
28
]
-~
. I
-
I
3. nailing 4. excavation
Tmax
However, the behaviour of soil nailed walls is not yet very well
understood. In particular, the influence of the construction method
(excavation of steps) hasn't yet been investigated. Due to the great
interest of the technique, several research programs have been initiated
in different countries (Germany, France, USA). The french project CLOU-
TERRE on soil nailing is a 3 years research program : it mainly consists
of 5 to 6 full scale experiments, centrifuge model tests, laboratory and
field research on friction, in order to propose reliable design methods
including an evaluation of the inclusions durability. The aim of the
project is to promote the technique in order to use it for permanent
structures. Nail durability is therefore carefully considered an~ close
attention is given to nails specially designed for preventing metallic
rod or cable corros i on (figure 19).
® " TBH A" NAIL PATE NTED AND DE VELOPED BY SOLRE NFOR
Steel cobles~'---~_.'1P"
•==}
><
w
f--
0
w
(,!)
Woven 50-1000 15- 350
§
c::::::J c=::::J
E
E
0
Q j 1 Extruded
"
og
In
C>
c:
c:
~
• • •
1
.,
p"",~
0
rn ;: 00000
c
'0
E
E 00000
. 00000 >
~ 0 50-700 9-90
·. . . . .
(,!) In
10
c:
0
W
0 00000
.,c:
"in 10
(!)
E
is
.
E
0
10 II}WoId'd bars
10
·.... ..
FIGURE 20. Types and mechanical properties of geotextiles and geogrids
31
-
..-- Sand wi th high strength
inextensible inclusion
Sand with high strength
_ _ _ _ _ .-/- extensible inclusion
b'
"- /
0
~ Sand with low strength
I inextensi ble inclusion
0
:;::
I
I /.~,
/ ' ... :"";':_"":"'::"":''';'':''':
~
II>
II> /
1/ ~--
",'/ Sand alone --,,' ~
- Dense sanl-.":
~
(i) 4
I Sand with low strength
l extensible inclusion
2 4 6 B 10
Axial slrain EI (%)
( Type of )
( In- rein. ! 1-D 2-D 3-D )
( clusion !
(
~~~--~~--------~----------~----------)
! )
( RIGID !Metal strips !Grids (metal- !Fibers (metal- )
( !lic) !lic) )
( )
( !Rigid plastic )
( !strips )
( ! ! )
( SEMI- !Grids (plastic)!Texsol )
( DEFORMABLE !Flexible plas- !Fibers (synthe-)
( !tic strips !tic) )
( ! )
( DEFOR..'1ABLE !Geotextiles )
(
----------~----------~----------~------~---)
Table III - Classification of reinforcement techniques according to the
inclusi~geometry and to its relative deformability
1971 with the construction of the Poitiers wall using Tergal strips.
This wall was a temporary structure, 5 m high and 40 m long. The tergal
strips were attached to cruciform concrete panels. The calculation of
the wall took into account tergal creep and during its service life, the
wall behaved satisfactory. However, it appeared during construction that
tergal strips had to be slightly prestressed in order to prevent exces-
sive lateral displacements of the facing.
In 1981, ten years after its construction, the wall was dismantled
and interesting durability tests were performed on the strip material.
As presented further, it was shown that the plastic fibers had been
degraded and that the mechanical properties of the plastic strips had
decreased.
® U.S. Method
® French method
Corner detai I
~ Detail
showing ~
.....,
.:
"'; ':
':....
':".!: ......:
.
A
rod
settlement
M: .. ..
......... ......,,:
:-
"
Construction
sequence
q
HHHLH
r
compacted sand
t--'~=r-~~~++:~'
I~~'--'" 'h~
L~--r'"--..r-~~---r-~---r---r--.J::..
@ Concrete facing units ® Gobion facing units
Geogrld
Armater sheet
Moltimum
slope
Temporary
wedges Reinforcement
Temporary
clamp
3.6 Websol
This type of soil retaining system has been developed in U.K. by Soil
Structures Limited at the end of the 70s and it is very similar to the
Reinforced Earth technique.
Paraweb is a composite plastic material consisting of polyester yarns
(terylene) coated and protected by a sheet of black polyethylene
(alkathene). The fibers give high strength as well as relatively low
creep characteristics to the material, and the Alkathene sheet protects
it from ultraviolet rays as well as other chemical degradation. Despite
its extensibflity, Paraweb appears to be an interesting plastic material
for soil reinforcement. The Paraweb strips are 2 mm thick and 90 mm wide
(figure 2Sa), and its rupture load ranges between SO and 100 kN.
Facing elements are T-shaped concrete panels (figure 2Sb). The rein-
forcements are formed by a continuous strip which has to be prestressed,
because of the extensibility of Paraweb, in order to avoid displacement
of the facing (figure 2Sc).
John et al. (1983) have performed two full-scale experiments on
Websol reinforced soil walls. The distribution of maximum tensile forces
measured the one corresponding to inextensible reinforcements as steel
strips. In the upper part of the wall, it appears to be close to the Ka
line, suggesting that the lateral deformations of the structure are
large enough to attain the Ka state of stress in the soil. This effect
of the extensibility of the reinforcement strips on the maximum tensile
forces has since been confirmed by several studies ; for instance,
fiberglass reinforced plastic behaves as an inextensible reinforcement
material. Reinforcement extens·ibility has also an influence on the
potential failure line (maximum tensile forces line) ' which is close to
Rankine's plane in the case of extensible reinforcements.
I
\
DETY'
( Q"i\'~"~::E:=S(f«"'THYlE"'1
TER't'LENE (POLYESTER)
CLASSICAL SOLUTION
WITH CONCRETE WALL
} TEXSOL WALL
SOLUTION
i
Natural slope
i,
Concrete
I,
wall I
profile
1.51
SOIL
rc::ol
Highway ~
o90m.1 L
® Texsol wall design method
200
100
THREAD CONTENT
BY WEIGHT (%)
FIGURE 26. Texsol wall solution used for enlarging the french A7 highway
(1985-1986)
41
4.1 Extensibility
4.1.1 Polymers properties
Table IV shows some typical mechanical data for the more current po-
lymers used in geotechnical engineering, as compared to steel proper-
ties. It should be noticed that, for each polymer, both bulk material
and filament properties are given. In fact, filaments are produced by
extrusion of the heated polymer mass, and high tenacity yarns, as des-
cribed in the table, are obtained by controlling the cooling process of
the filament, and by stretching it during extrusion. In such a way,
polymeric chains have a preferential orientation and an increased aniso-
tropy, thus resulting in much higher (up to ten times) mechanical
properties of the material, such as tensile strength or elastic modulus.
(
( Elastic modulus !Tensile strength !Failure elonga)
( ! ! (MFa) (MFa) !tion filament)
( Material !Density! (%) )
( ! ! ! ! )
( ! Bulk !Filament! Bulk !Filament )
( !material! !material! )
( )
( )
(Steel 7,85 200 000 2340 3 )
( )
(Polyester 1,38 2100 13500 60 '800 6-20 )
( 18500 )
( )
(Polypropylene 0,91 1100 7400 35 !380-780 15-25 )
( )
(Polyamide ••• 1,1 2400 12500 40-120 !600-900 12-26 )
( )
(Polyethylene 0,95 800 4300 30 !390-700 10-20 )
(high density 6500 )
( )
Tensile stress
(MPo)
2800
.'F'Iberglass
2000
1600
1200 /Polyester
Nylon
800
400
Extelllion °4
5 10 15 20
moduli, which are, for current polymers, from 10 to 30 times lower than
for metals ; the tensile strength is also lower (2,5 to 5 times),
whereas the deformation at failure is much higher (20-30 % instead of
3 %). When comparing current polymers, it may be seen that polyester
and polyamide, which have higher densities (1,38 and 1,1 respectively)
have higher stiffness and tensile strength than polypropylene and polye-
thylene high density (0,91 and 0,95 respectively).
Figure 27 shows tensile test results for steel, fiberglass, polyester
and polyamide filaments. Whereas failure occurs, for steel, at a defor-
mation of 3.2 % for a tensile strength of 2340 I1Pa, polyester and nylon
fail for much larger deformation (respectively 11 % and 19 % on the
figure), and a tensile strength of about 1000 Mpa.
Baudonnel et al. (1982) took some filaments from current polyester
and polypropylene geotextiles, and elongation tests results are shown in
figure 28.
F
A (N hex)
0,3
)(
~
~
..-< 0,2 PoJypropylene
en
en
0
E
...
0
CI>
c: Polyester
SD A =7.35
LT A =2.45 tJ. • (dtex)
-I
"-
Polypropylene
SD A = 7.1 •,
u. TP A =12.8 0
CI> 0,1-
...
v
~
oL------------,--------____.-__________-.______ _
£ (%)
° 50 100 150
Strain
Dial
gouge
Top clamp
Reinforced
geotextile
Pressure bellows
LH' __-M--.... · Lu bricated
Geotextile --l,;,t----{!l membranes
Bottom clamp
1~f'1J~~,,¥'------ Bottom
clamp
12
10
",.
/'
8 /
e /
"- /
...
z
/
/
/
0 /
« 6 /
9 /
/
Non
/4:::0,.. woven
/ polyester
I
4 /
/
I
/
/
,/
/
/
- - - In(IOOkN/m
soil confined
3)
/
/
/ ------Unconfined
/
/
0
0 10 20 30 40
AXIAL STRAIN (%)
FIGURE 30. Influence of the in-soil confinement and of the fabric struc-
ture on elongation properties (Mc Gown et al., 1982)
47
t Load
I
I
I
I
I
I
I
:E
I
wide strip. The pins cross the fabrics and avoid restraint. Shresta and
Bell (1982) have used this device for testing several geotextiles.
Testing of 200 mID wide strips has been also performed by Murray et al.
(1986), and Richard and Scott (1986).
12r-------------------------~
-----Unconfined in Isolation
10
100
E 8 55
"-
...z 10
/1'"",
c o ,
~
g 6 ",,"
"
/Size of
/ test
, specimens
4
(mm)
400)( 100
2 / 300)( 100
/"~ ~ 200 )( 00
, ,," ~ 100 )( 100
....... ........
10 20 40
AXIAL STRAIN !%)
Leflaive et al. (1982) performed 500 mID wide tests, and proposed a
correction corresponding to the lateral contraction. The corrected
results they obtained compared favourably with results of cylindrical
sleeve test, and in-soil confined tests (McGown et al., 1982). This
approach has been adopted by Cazuffi et al. (1986), Leclerq and Prudon
(1986). Rowe and Ho (1986) also suggest a 500 mID wide value.
Other parameters have been studied. Figure 33 shows the influence of
the strain rate on the maximum tensile strength of some woven fabrics
(Rowe a nd Ho, 1986). In this case, where the constitutive fibers are
directly sollicitated by the tensile test, the effect of strain rate is
import a nt, and the writers suggest a 2 % r a te. Several writers conside-
red the influence of the tensile direction, as compared to warp, weft,
or diagonal direction for woven fabrics, and production direction for
non-woven fabrics. In the later case, Van Leeuven (1977), Leclerq and
49
E
......
...
z 160
:I:
I-
(!)
Z
W
II::
I- 120
If)
w
...J
iii
z
W
I-
::I 80
::l
::I
x
<t
::I
o
40
-r
- - i ]~------------~o~--------~o----
0
0.1 10 100
STRAIN RATE( %/mln) - LOG SCALE
50~-------------+-+~ 200r-------------r-+_
e
.......
z
~ 40r-------------~~-- e
.......
C
z 150r--------------i~--
~
<I:
9 --- Polyester en
::J
~ 30 ~----------j>---/'---+--- --- Polypropylene ..J
...
::J
II.
::J
C
~ 1001--~~-----.~+-+----
::J
II:
z
20r-----~~L----+--- Q
~
2
II:
~ 50r---~~------~~---
1£1
10r---'---+--------+--- c
I
'Y
/
00 500
SURFACE WEIGHT (gl m2 ) SURFACE WEIGHT (g/m 2 )
FIGURE 34. Tensile strength and deformation moduli of polyesters and po-
lypropylene fabrics listed on the sleeve-cylinder apparatus (Paute and
Segouin, 1977)
_ _ _ Polyester
- - - - Polypropylene
~~~~
~~
----------
l--~__~==~~~~~::==------_:~~~~--~~~=-~----------------
-- -- ---
~~
5 .... ----
-~
---------------
0.1
-- -- -- --- 10 100
TIME
1000
109 t (hours)
10000
£{%l
20
~-
~ Q{oQ'i
?o:'1 ____
- I I
----
16
~ Polyoml'd e -----
12
Polyester -----
T
8 ~ I-~
I
i
1
4
t {days I
o
0.001 QOI QI 10 100 1000
I
,
I
I
I min I hour I day
FIGURE 36. Creep of synthetic woven fabrics under prolonged loading (50%
of the breaking strength) (Van Leeuwen, 1977)
ving the polypropylene yarns creep rate under 40 % breaking load, from
1,5 to 0,40 by an adequate treatment. However, for woven fabrics, they
only give creep rates at 20 % breaking load (0,40 - 0,73), which does
not allow a direct comparison to be made.
)
Reference Type of % of breaking Creep rate )
material load ! (% extension for )
!per log10t cycle»
! )
)
(Finnigan (1977) ! Polyester yarn! 41,2 0,192 )
( ! ! )
(Finnigan (1977)!Polyester yarn! 58,8 0,194 )
( ! )
(Greenwood and )
(Myles (1986) !Polyester yarn! 40 0,125 )
( ! ! )
(Greenwood and )
(Myles (1986) !Polyester yarn! 60 0,27 )
( ! ! )
(Van Leeuwen Polyester )
«(1977) woven fabric 50 0,2 )
( )
10
DURATION, HOURS
OL--------.---------.--------.--------.----~--_,--~
0.1 10 100 1000 10000
FIGURE 37. Creep behaviour of tensar SR2 (after Mc Gown et al., 1982)
54
l-
Unconfined I
30 I
In-Isolation ~ I
16
------------ -,... 40 Unconfin .. d I
I
In -Isolal ion-,-
20 /
/
/
;?12 Confining pressure
z 30 ,,/' Confining
z 100 kN/m2
...
~
cc ,"" pressure
,10------
-
4 /
~ IF> 100 kN/m2
...cc 8
IF> ~20 I
>< I
<{ I
I
I
I
10 I
/
/- '\.\.'+-~
, -i>
-s>
Irerram 1000 I 0<::.
Bidim U24
Table VII shows the results of creep tests which were performed on
200 mm wide strips by Shrestha and Bell (1982). Lowest creep was exhi-
bited by polyester resin bonded NW1, and highest creep by polypropylene
staple needle punched NW6. Creep is most sensitive to load levels for
the continuous filament polypropylene geotextile NW3 (Heat bonded) and
NW5 (Needle punched), which show a 3 to 5 fold increase in creep when
the sustained load is doubled. The similar results obtained for these
two fabrics indicate an identical effect of heat bonding and needle
punching regarding creep behaviour. Staple filaments (NW6) seem quite
unfavourable with respect to creep properties. For NW1 and NW6, creep
increased by only 1 % to 2 % for increases in the sustained load levels
55
CREEP STRAIN
(%)
250
200
150
300
100
5O."--_~
TIME (minutes)
O~-----'-------.------.-------.------.-------.-------r--~
0.1 10 100 1000 10000 100000 1000000
CREEP STRAIN
(%)
50
POLYPROPYLENE
J-
40
Typor 3401
30
• .. ".At
Bidim C34
20
• POLYESTER
10 t=::::::~::=~-6-a-i:J:--_-6--6-~"--~~S~t:O:bilenko T 100
TIME (minutes)
OL------.-------.------r------,-------.-------r------,---~
0.1 10 100 1000 10000 100000 1000000
4.3 Durability
As mentionned by many writers, one of the reasons why people started
considering polymers as a possible material for soil reinforcement was
the risks of electro-chemical corrosion encountered when using metallic
inclusions.
In a general manner, the problem of long term durability of any type
of material submitted to given physico-chemical and mechanical
conditions is very much involved, and long term real experiments seem to
be the only reliable solution. Thus, adopting a new technology in which
durability problems are involved is always dangerous, and may lead to
important errors, due to previously unknown phenomena.
For example, in the case of metallic strips, previous m~dium term
experiments developed on stressed stainless steel strips confined in-
soil gave excellent results, which led the Reinforced Earth Company to
adopt this metal for strips. Nowadays, some of the walls built with this
type of strips are presenting some problems due to a rapid and unexpec-
ted corrosion, and have to be repaired. On the contrary, galvanized soft
steel generally shows excellent behaviour, and the numerous corres-
ponding walls, which were erected 15-20 years ago, still behave
satisfactorily.
Paragraph 2.2. of this report described the use of. polymers in
Reinforced Earth Technique. In the case of the wall of Poitiers, strip
samples were taken from the upper part of the wall. According to tensile
tests, a 50 % loss of strength, and a 25 % decrease of the rupture
strain were estimated. Optical and electron micrographs evidenced an im-
portant degradation due to setting (flexion, shearing or crushing
actions). The micrographs also revealed a degradation of the strip sur-
face, corresponding to a "dirty" macroscopic aspect, which was
attributed to a contamination caused by the soil.
In fact, durability problems of polymers, and ageing phenomena are
58
100
GRAB STRENGTH • •+ *[;:, Florida
Arizona
o North California
60
20
WEEKS EXPOSED
8 16 24 32
GENERAL CONCLUSIONS
Polymers are or can be used in numerous types of retaining systems,
ranging from the classical "ladder wall" invented by Coyne to Texsol,
recently developed by Leflaive. Of course, each application is different
and depends on many factors (cost, aesthetic aspect of the structure,
structure deformability, durability, etc ... ).
' As used in reinforcements, polymers provide an alternative to metals,
the latter being so far limited to linear inclusions and grids. However,
their deformability is generally higher than metals and it appears to be
a disavantage. They are not subject to corrosion, but they can be degra-
ded.
Considering wall construction, the use of deformable polymers
61
REFERENCES
ACKNOWLEDGEl1ENTS
The authors acknowledge Mr J. CANOU for his help in preparing the ·En-
glish version of the text.
Prediction Exercise
INTRODUCTION AND RATIONALE FOR PREDICTION EXERCISE
PETER M. JARRETT
Rudolph Bonaparte
Danielle Cazzuffi, Pietro Rimoldi and Roberto Mangiavacchi
Philippe Delmas and Jean-Pierre Gourc
Richard Jewell
Colin Jones
Robert Koerner and Manfred Hausmann
Richard Murray
Gregory Richardson
Jost Studer and B. Graf
The organizers most sincerely thank these gentlemen for their consi-
derable efforts, their spirit in accepting the challenge and for the
friendly, open and cooperative manner in which their discussions were
carried out.
69
P. M. Jarrett and A. McGown (eds.), The Application ofPolymeric Reinforcement in Soil Retaining Structures, 69-70.
© 1988 by Kluwer Academic Publishers.
70
the overall value of the basic test program. There are two keystone papers
for the prediction exercise and three supplementary papers.
During demolition of the test walls, samples of the Tensar SR2 rein-
forcement were recovered and returned to the manufacturer, Netlon Ltd. for
examination and testing to assess the material's resistance to damage
during construction. The third paper by Bush and Swan reports the results
of this work.
1.0 Introduction
1.1 General
Two(2) large-scale model reinforced soil walls were constructed within the soil
retaining wall test facility at the Royal Military College, Kingston, Canada.
The vertical reinforced soil walls were 3 m high and were constructed using:
a) an incremental timber panel facing with a polymer geogrid soil reinforcement, and;
b) a propped (tilt-up) timber panel facing with a polymer geogrid soil reinforcement.
Each wall was subjected to sustained surcharge loadings up to 50 kPa following
construction.
1.2 Objectives
The construction and performance monitoring of the two reinforced soil struc-
tures was undertaken to provide a focal point for discussion at the NATO Advanced
Research Workshop entitled: Application of Polymeric Reinforcement in Soil Retaining
Structures held at the Royal Military College of Canada in June 1987.
The workshop was broadly directed to review the state-of-the-art in polymer rein-
forced soil retaining structures and to critically assess the relationships between design
methodologies, construction practices and the measured performance of these aystems.
In order to stimulate discussion, invited experts were given details of the proposed
reinforced soil wall configurations before construction (Bathurst and Jarrett, 1986)
and asked to make Glass A predictions of the behaviour of the two reinforced soil
structures at various elapsed times after initial construction and after surcharging
pressures were applied to the structures. The results of the prediction exercise were
presented informally and were the centre of discussion for two workshop sessions.
This paper describes the construction of the two test walls and the measured test
results. A summary of the performance predictions offered by Workshop participants
is reported by Bathurst and Koerner in these proceedings.
2.1 General
The RMC Retaining Wall Test Facility was conceived and constructed to provide
a general purpose large-scale test apparatus to examine a variety of reinforced soil wall
systems. The test facility is located within the Dolphin Structures Laboratory of the
Civil Engineering Department at RMC. The principal structural components of the
test facility were completed in November 1985.
71
P. M. Jarrett andA. McGown (eds.), The Application of Polymeric Reinforcement in Soil Retaining Structures, 71-125.
© 1988 by Kluwer Academic Publishers.
72
3.1 General
Schematics of the reinforced soil walls tested are shown on Figures 3 and 4. Both
constructions comprised a high density polyethylene TENSAR SR2 Geogrid-reinforced
sand soil. Reinforcement layers were attached to the plywood bulkhead facing panels
shown on the figures. The principal differences between the structures were the front
panel configurations and details of construction sequence. The incremental facing panel
wall was constructed in four stages with each panel externally supported only until the
sand fill behind the panel was placed and compacted. The facing panel supports on
the propped wall construction, however, were released only after the full height of
sand fill had been placed behind the facing units. Both model Vl{alls were constructed
with a central instrumented section nominally 1 m wide and two 0.7 m wide edge
sections. This construction was adopted to reduce edge-effects on the performance of
the monitored central reinforcement strip and facing unit(s). A foam rubber void filler
was placed along all panel edges in order to prevent the panels from binding during
outward movements. In addition, for the incremental wall test, a layer of compressible
foam rubber was placed at each horizontal panel to panel interface to assist in panel
levelling during construction of the wall.
A principal objective of the current investigation was to examine the behaviour
of these reinforced soil wall systems under working stress conditions. Therefore, the
dimensions of the trial sections were based essentially on the design methods contained
in the U.K. Department of Transport's technical memorandum Reinforced Earth Re-
taining Walls and Bridge Abutments for Embankments, BE 9/78 with one exception,
the grid lengths were limited to 3 metres instead of 5 metres. Preliminary trial tests at
RMC with wrap-around geogrid facings have shown that for the same number of rein-
forcing layers (e.g. four) and a total height offill equal to 3.25 m, 3 metre reinforcement
lengths are more than adequate for stability (Bathurst, 1987).
73
STRUCTURAL
0
4m
0 0 0 MODULE
NUMBER
0 0 0 0
0
0 0 0 0
0 0
0
0 0 0 0 0
0
0 0 0 0
ANCHOR BOLTS
FRONT OF WALL
HSS 152x102x9.5
(
TIMBER BEAM
38 x 184 (TYPJ
t------~*-
SURCHARGE
PANEL
SETTLEMENT
.PLATE ________ 1 ___ _
SAND SURCHARGE 200
REINFORCED SOIL
FRONT PANEL
T E
SR2 LAYER 3
I
SR2 LAYER
2 ""j:af'"'----~--"........"""-t E
SR2 LAYER I
~L~~N~ ~y~ ~ _ _ ... _______ ....
~
l
• STRAIN GAUGE
ErE LEVELLING PAD"
_ PRESSURE CELL
D\ ,
SURCHARGE
PANELS
, 1\
T
SR2 LAYER 3
~~
SR2 LAYER 2 ~
~
"
z
u
1t
~
l
ETE LEVELLING PAD ..:.
• STRAIN GAUGE
_ PRESSURE CELL
Details of wall design and construction were arrived at through collaboration be-
tween the authors and Dr. A. McGown, Dr. K. Andrawes and D. Varney at the
University of Strathclyde, Scotland. Construction details are considered realistic for
actual field structures although some compromises in construction were made to make
the performance· of these structures easier to predict and monitor. For example, facing
panels are not dovetailed or staggered and the facings were arranged without an initial
batter in the vertical direction. Another example is the panel/reinforcement connec-
tions. The number of bolts per grid was limited to four in order to ensure adequate
sensitivity of the bolts as load cells.
76
3.2 Materials
3.2.1 Sand
The soil used was a washed sand with some gravel. This material is composed of
sub-angular to angular quartz and feldspar particles. The grain size distribution for
this material is given on Figure 5. Less than 0.3% of all particles by weight are less
than 0.075 microns.
The results of direct shear tests and triaxial compression tests carried out at RMC
are given in Appendix A. A summary of test results is given on Table 1. The second
column of friction angles shown on the table are values representing linear failure
envelopes passing through the origin of a T - Un plot. Additional direct shear tests on
the RMC sand are reported by Jewell (1987b) as part of these proceedings and confirm
(secant) peak friction angles interpreted from the RMC direct shear tests.
90
80
(!) 70
Z
en
en 60
«
a..
~ 50
Z
I.&J
~ 40
I.&J
a..
30
20
10
0
200
U.S. SIEVE Nil.
Table 1
4.1 General
The construction sequence adopted for both walls was similar with the exception
of the temporary waling support and release sequence.
Prior to construction of both test configurations, a 250 mm thick blinding layer
was placed and compacted behind the concrete floor levelling pad shown on Figures 3
and 4. The timber facing panels were seated on the concrete levelling pad and initially
restrained between the concrete lip shown on the figures and a wooden waling located
in front of the wall at the same elevation. Subsequently, all sand backfill was placed and
compacted in 125 mm lifts covering the full length of the test facility. A vibrating plate
tamper was the principal means of compaction although a hand-held Kango hammer
with tamper attachment was employed to compact sand in the corners formed by the
facing panE,lls and the test facility walls.
Prior to placing fill over any strip of reinforcement, a light pretensioning load of
about 0.4 kN /m was applied to the geogrid to ensure that the reinforcement was free
of warps. The pretensioning was applied using a bar threaded through the free end of
each strip of reinforcement and attached in turn by a cable to a system of weights at
the back of the test facility. The pretensioning was released after the grid was covered
by 0.5 m of compacted sand.
78
In this construction a total of twelve 0.75 m high timber bulkhead facing panels
were used to construct a reinforced soil wall 3.0 m high. Temporary lateral support
was provided to each row of panels by a pair of horizontal timber walings braced in
turn by a system of wedging plates bolted to the front of the RMC test facility. Each
panel row support was released after the sand backfill had been placed and compacted
to the elevation of the top of the facing panel. The purpose of this form of construction
was to progressively mobilize the inherent tensile capacity of the geogrid reinforcement
as the height of the composite structure was increased.
The construction sequence is illustrated on Figure 6. Also identified on the figure
are critical construction events. A frontal view showing the incremental panel wall
facing units at the end of construction is given on Plate 2. Following construction,
the incremental facing wall was subjected to a series of surcharge loadings using a
sand surcharge or sand surcharge/pressurized airbag combination. The surcharging
schedule is given in Table 2. The design surcharge of 12 kPa was sustained for a period
in excess of 1000 hrs in order to observe creep behaviour under working (i.e design)
loads. Later, a maximum surcharge load of 50 kPa was applied for 500 hrs to observe
further creep deformations in the composite system.
Table 2
Construction Events
..CJ
cd
@ 750 mm uncompacted sand surcharge in place
"00 2 .5
Q)
~
..CJ
Q)
>
0
..CJ 2.0
cd
,::
0
.~
>
Q) 1.5
OJ
0;:1
1.0
I
100 150 200 250
time from start of construction (hrs)
construction sequence for this wall including critical construction events is given on
Figure 7. In a similar manner to the incremental panel wall, the propped wall was
surcharged using the airbag system installed at the top of the RMC test facility. The
surcharging schedule is given in Table 3. In this test the maximum surcharge pressure
of 50 kPa was sustained for 1000 hrs.
Construction Events
1.5
Table 3
o at excavation
moisture
dry density ,Pd (kg/m 3 ) content, w (%)
3.5 -
.... ....t:
P'"
backfill 0 0 0 0 0 0
It) It) CD CD 0 0 0
layers t: ~ t: t: N It) .t-
o 0 0 I I I 0 I
3.0 A &
-
..
f-
4
"0 ~
,..-...
S
'--'
2.5 A 4
..0
cd
.-
rn
Q.)
rn 3
cd
..0
Q.)
2.0 l- • ..a ~ -
?-
0
..0
cd
-" ~
+J
r:::::
0
......
cd
1.5 ~
2
-
?-
Q.)
.-
A 0&
Q.)
.-
.- ....
~ 1.0
1
0.5 ~ 6 ~ -
-
A
o ~ ~ ~ I _I _I ..1- ~
~ during construction
o at excavation
moisture
dry densitY,Pd(kg/m 3 ) content, w (%)
3.5 ,..
•t::
0 0 0 0 0 0
", CD f"- CD C! 0 0
....
It)
t: f"- f"- f"- N ~ ~
I I I I
3.0 !"
S 2.5
'--'
- • .t. ..
-
..D
cd
rJl
..
- -- --e----
Q)
rJl
cd 2.0 ~---- ~~---
..D
•
..
Q)
> .t.
0
..D
cd
~
0
1.5 -
---- - -
.~
--
+"
cd 1 - - - _ . ~-- -~---
>
Q)
Q)
te:
1.0 ~
• A
•
o I I . • •
5.0 Instrumentation
5.1 General
Each wall was instrumented to record the following:
a) horizontal movement of facing panels,
b) longitudinal displacements and strain in the geogrid reinforcement,
c) load at the panel/geogrid connections,
d) distribution of vertical earth pressure below the reinforced block of soil,
e) vertical settlements at the top of surcharge and,
f) temperature in fill.
The instrumentation used to record the data identified above was installed as
construction proceeded and was monitored for the duration of testing including wall
excavation.
5.3.1 General
The displacement of the reinforcement in the longitudinal direction of the model
walls was monitored by extensometers (i.e. by attaching tensioned steel wires to se-
lected locations on the mesh). Movement of the thin steel wires was recorded by
electronic displacement transducers mounted at the back of the test facility. The wires
were protected from the granular fill by passing them through stiff plastic tubes. In the
incremental wall construction, two displacement monitoring points per reinforcement
layer were used. In the propped wall test up to six displacement monitoring points
were used for each strip.
In addition, strain gauges were attached at selected mid-rib locations along the
length of each central reinforcement strip. After several years of experimentation with
different techniques of geogrid strain-gauging, the authors have found that high-strain,
foil-type gauges manufactured by Showa Measuring Instruments Co., Ltd. perform
satisfactorily. These strain gauges (together with proper rib surface preparation and
a two-part RTC epoxy adhesive) are effective at measuring small levels of strain in
the polymer reinforcement while the extensometers are effective at determining large
strains after the strain gauges have failed.
5.3.2 Strain Gauge Calibration
Experience with a variety of TENSAR geogrid reinforcement products has shown
that, in general, strain distribution is not uniform along any reinforcement rib. This
non-uniformity is a consequence of grid geometry and variable material moduli. In
order to make comparable strain measurements between strain gauge readings and
strains deduced from displacements recorded over several grid apertures it was neces-
sary to correlate strain gauge readings against gross average strains in the grid. In this
paper, strains measured over a gauge length consisting of one or more apertures are
called grid strains and strains measured with a gauge mounted at the mid-point on a
rib are called rib strains. It can also be appreciated that grid strain is a more meaning-
ful parameter since it facilitates comparison of mechanical response between different
sheet reinforcement types and is more easily implemented in analytical models.
85
Recent investigations at RMC show that the relationship between grid and rib
strain can vary with geogrid type (e.g. Jarrett and Bathurst, 1987). In the current
investigation the correlation between strain gauge readings giving rib strain and grid
strain was determined from load-controlled in-isolation tests carried out on sections of
SR2 geogrid. A series of increasing static loads was applied rapidly to test specimens
and each load increment was sustained for 24 hrs while strain gauges and grid strains
were recorded electronically. The results of this test program are summarized on Figure
9. For SR2 material in the longitudinal direction, the data shows that grid strains and
rib strains are essentially equivalent up to about 1.8% grid strain. The strain gauges in
this test were arranged over two rows of ribs. Nevertheless, there was some deviation
about the average strain gauge reading at any given grid strain. This variability is
thought to be due to the sensitivity of gauge response to small differences in gauge
positioning, rib dimensions and non-uniform load distribution between ribs.
clamp
SR2 TENSAR Geogrid sample
I I I I I (longitudinal direction)
15 ribs wide by 5 bars long
I
"
clamp
I I
10 strain gauges mounted at
mid-rib locations
2.0r--------------,
NOTES:
6 static load increments
load increment duration = 24 hrs
,--..
maximum load increment = 10 kN/m
~ temperature = 20 ± 2°C
general range
o 2.0
grid strain (%)
surcharge
TOP displacement
potentiometers (2)
3m
PANEL 2
BOTTOM displacement
potentiometers(2}
6m
14
S 12
5
10
6
~------
BOTTOM
~/--40
4 j.50kPa
kPa
2
r-12 kPa surcharge 30 kPa
4000
elapsed time (hrs)
o@
Panel position after movement due to panel #2 bracing removal
Panel position after movement due to panel #3 bracing removal
Panel position 40 min after panel #4 bracing removal
@ Panel position after 12 kPa surcharge for 100 hrs
o Panel position after 12 kPa surcharge for 1000 hrs
@ Panel position after 30 kPa surcharge for 100 hrs
® Panel position after 40 kPa surcharge for 100 hrs
(@) Panel position after 50 kPa surcharge for 100 hrs
® Panel position after 50 kPa surcharge for 500 hrs
3.oor-----T5-----TIo---,,,15~_,~2°r_--172r5----~3TO~__~~~~4,0
geogrid LAYER 4
2.5
I 2.0
~ geogrid LAYER 3
EbO
- - - - -
"
'6
~
.D UI
"0>
.D
1:
'"
bO
'OJ geogrid LAYER 2
"" 1.0
geogrid LAYER 1
10 15 20 25 30 35 40
~ geogrid LAYER 4
°0~----------~1.~0-----------2~.~0----------~3.0
geogrid LAYER 3
1.0 2.0 3.0
~ ","grid LAYER'
°0~----------~1.~0-----------2~.0~--------~3.0
surcharge
LAYER 3, ROW 3
strain gauges
6m
1.0~------------------------------------------~
0.8
r
right
~ 0.6
'£...-=a 12 kPa ,.,w."g•
~
:9 0.4 \. 50 kPa
...
."..------ -------- J..-40 kPa
°0~~~----~IO~O~O~~~~2~OOO~~--L---=3~OO=O=---L---~4~000
Figure 13 also shows that sudden changes in strain gauge response match critical
events in wall construction and the application of increased surcharge loads. It is also
apparent from the figure that the polymer reinforcement continued to deform under
constant surcharge loading.
The average rib strains recorded along the length of each reinforcement layer are
summarized on Figure 14. The results of in-isolation tests described earlier indicate
that rib strains shown on the figures are essentially equivalent to grid strains. A number
of important observations can be made from these figures:
geogrid LAYER 1
0.6
0.5
0 CD end of wall construction (bracing removed from LAYER 4)
® 12 kPa surcharge for 100 hrs
~0.4 ® o 50 kPa surcharge for 500 hrs (end of test)
~
'"
~
';;J 0.3
CD NOTE: strains are those recorded after end of construction
...
+'
(i.e. after panel bracing removed from LAYER 1)
"'
:g 0.2
0.25 m compacted backfill in LAYER 2
0.1
o
0.8 geogrid LAYER 3
end of wall construction(bracing removed from LAYER 4)
0.7 @ 12 kPa surcharge for 100 hrs
® 12 kPa surcharge for 1000 hrs
0.6 o 30 kPa surcharge for 100 hrs
@ 40 kPa surcharge for 100 hrs
_0.5
~
~
@
@ 50 kPa surcharge for 100 hrs
G) 50 kPa surcharge for 500 hrs( end of test)
,,- 0.4 @
.~
NOTE: strains are those recorded
t;
after end of construction
~ 0.3 ?8 (i.e. after panel bracing removed)
0.2
CD
0.1
0.9
0.8
ffi
(i.e. after panel bracing removed)
0.2
geogrid LAYER 4
0.8 NOTE: all strains are those recorded at the end
of test after 500 hrs of 50 kPa surcharge
0.4
0.8
~
'"~
v
O~----~~----~~----~~--~~~--~~----~
0 0.5 1.0
- ........
be geogrid LAYER 2
.S
.:: 0.8
'Cd
'"'"
+"
0.4
o
NOTE: loads all calculated from end of test data after 500 hrs
at 50 kPa surcharge pressure
1>--0 range of load from geogrid strain gauges 0.22 m from panel
cr--o range of load from connection bolt strain gauges
6 estimated load at panel from extrapolation of strain
3.0 distribution
....., ::J:I - - - geogrid LAYER 4
]: 2.5
.£:J
..::lrn
<I)
rn
2.0 - ~ - c.. - - - - - - geogrid LAYER 3
ro
.£:J
<I)
:>
1.5 estimated from strain gauges 0.66 m back from panel
=orE"- - - - - - -
0
~
>:::i
lY - geogrid LAYER 2
0
.~ 1.0
:>
-
<I)
Q)
Outward panel movements with time are plotted on Figure 17. These movements
are referenced to the end of construction corresponding to the panel geometry just
prior to bracing removal. Similar to the results of the incremental wall, the propped
panel wall showed large outward movements at the application of each surcharge load
increment and significant deformation under constant surcharge loading.
Figure 18 shows outward panel movements recorded at selected times during sur-
charging of the propped panel wall. The data shows a progressive rotation of the facing
unit about the toe with a total outward movement at the top of the wall of about 13
rom at the end of the test. As in the incremental panel wall test, the data shows
significant wall movements under constant surcharging load increments.
97
displacement potentiometers
surcharge
!
TOP
3m MID-HEIGHT
BOTTOM
I.. ..I
6m
12
MID-HEIGHT
BOTTOM
2
°O~----------~5~O~O~----------710~O~O~----------~15~O~O--~
Figure 1'7 Panel Movement versus Elapsed Time (Propped Panel Wall Test)
98
3 .0
elevation
geogrid LAYER 1
4 6 8 10 12 14
panel movement (mm)
The propped panel wall was instrumented with up to six extensometers attached
to each reinforcement layer. The horizontal displacements recorded by these devices are
given on Figure 19. Also shown on the figure are outward panel movements recorded
at the end of the 50 kPa surcharging increment.
Figure 19 shows that there was essentially uniform displacement of the grid layers
after the initial 100 hrs of the 12 kPa surcharging increment. For example, in the top
layer this movement was about 2 mm towards the wall which is somewhat less than the
3.3 mm outward movement recorded at the panel facing opposite this reinforcement
layer (Figure 18). This observation suggests that there was some initial slip of the grid
as soil/grid interlock was established during the initial loading increment. At higher
surcharge loads the increased slope of the displacement profiles shows that tensile
straining in the reinforcement layers and load transfer between the grid and soil was
occurring.
99
geogrid LAYER 4
NOTE:
movements are recorded after panel bracing removed
geogrid LAYER 3
geogrid LAYER 1
a) The magnitude of strain recorded at a given distance behind the facing can be
seen to increase with time and surcharge load level.
b) The propagation of strain into the reinforcement can be seen to increase with time
and surcharge load level.
c) In geogrid layers 1, 2 and 3 there is a distinct change in strain gradient at about
0.45 to 0.65 m behind the wall.
The strain distributions for the propped wall show qualitative features in the im-
mediate vicinity of -the wall facing which are different from those recorded for the
incremental wall. In particular, layer 4 shows a progressive increase in geogrid strain
as the distance to the panel facing decreases. At the same location in the incremental
wall there were reduced strain values. Layers 2 and 3 show qualitative features that
fall between the strain profiles recorded for the incremental wall and layer 4 in the
propped wall. The difference in strain response is considered to be due to the rela-
tive vertical compllance in the two wall constructions. In the incremental wall test a
horizontal layer of compressible foam was placed between panel units. After fill com-
paction and surcharging, it was noted that the foam filler compressed leading to an
overall 8hortening of the full wall height by roughly 10 mm. This degree of freedom
was not available in the propped panel wall which was constructed as a single facing
unit. Consequently, the fill moved down with respect to the panel unit during fill
compaction and surcharging and in the process generated additional tensile loading
in the reinforcement close to the geogrid/panel connections. The relative soil/panel
movement can be expected to decrease with distance below the top of the wall and
this accounts for the observation that additional geogrid loading in the vicinity of the
panel connections appears to diminish with lower reinforcement elevation.
Grid strains were also calculated from the results of extenspmeters (i.e. from the
gradients plotted on Figure 19). The results ofthese calculations are given on Figure 22.
The figure shows that these calculated strains under-register the grid strains inferred
from the rib-mounted strain gauges . .Nevertheless, both sets of data show generally
consistent trends with the extensometers tending to give lower-bound estimates of grid
strain. Again, the discrepancy is thought to be due to mechanical compliances in the
steel wire systems.
6.3.4 Load in Panel/Geogrid Connections
A significant number of strain gauges attached to the panel connection bolts were
observed to have failed when the propped wall was excavated. The poor performance
of the strain gauged bolts did not allow panel/geogrid loads to be calculated with
any confidence using this method of instrumentation. Nevertheless, an upper-bound
estimate of panel loads can be determined from the strains recorded in the grid at
strain gauge locations closest to the facing unit. The results of this exercise are given on
Figure 23 and show that the load in the panel connections increases with reinforcement
elevation.
101
surcharge
LAYER 3, ROW 1
strain gauges
6m
1.0r------------------------------------------------,
panel bracing removed
0.8
30 kFa
---.
~
40 kPa
'--' 0.6 right
~
...,...
.~
- - ---- ----
[/J
..0
.;:: 0.4 center
left
0.2
1500
geogrid LAYER 1 NOTE: strains are those recorded after end of construction
(i.e. after panel bracing removed)
0.5
(0 1 hr after bracing removal
0.4 ® 12 kPa surcharge for 100 hrs
o 12 kPa surcharge for 430 hrs
~
';;;' 0.3 o 30 kPa surcharge for 100 hrs
.;;;0:
....
® 40 kPa surcharge for 100 hrs
@
o
tl 0.2 50 kPa surcharge for 100 hrs
.D
.;::
50 kPa surcharge for 1000 hrs (end of test)
0.1
2.5 3.0
.;:: 0.2
.D o 50 kPa surcharge for 1000 hrs (end of test)
0.1
- --
--:::-=:
1.0 1.5 2.0
distance from panel wall (m)
geogrid LAYER 3
0.7
NOTE:
strains are those recorded after end of construction
(i.e. after panel bracing removed)
0.6
0.1
@
®
00 0.5 1.0 1.5 2.0 2.5 3.0
distance from panel wall (m)
0.9
®
0.8 NOTE: strains are those recorded after end of construction
---{i.e. after panel bracing removed)
..-
o 30 kPa surcharge for 100 hrs
.~ 0.5 ® ® 40 kPa surcharge for 100 hrs
'0: @ 50 kPa surcharge for 100 hrs
®
.0
'C
0.4 @ 50 kPa surcharge for 1000 hrs (end of test)
0.3
0.2
0.1
0
--
--
0.5 1.0 1.:1 &.0 2.5 3.0
distance from panel wall (m)
Figure 21 (Cont'd) Summary of Strain Gauge Readings (Propped Panel Wall Test)
104
geogrid LAYER 4
0.8
NOTE: all strains are those recorded at the end
of test after 500 hrs of 50 kPa surcharge
0.4
I I I
- - --
I -'
0.5 1.0 1.5 2.0 2.5 3.0
geogrid LAYER 3
0.8
~0.4
00
geogrid LAYER 1
0.8 l:Jr---6 strain from resistance strain gauges
_ strain from wire connection/rear
~
potentiometer system
00 0.5
distance from panel wall (m)
~ 1.5
.D
o:l
~
- - -0-00 - - - - - - - - - geogrid LAYER 2
o
.~ 1.0
:>
Cl)
Q)
°o~--~~~~--~~--~--~~--~~--~
3.0 6.0 7.0
panel connection load (kN/m)
Figure 23 Geogrid/Panel Connection Loads from Propped Panel Wall Test
The results of earth pressure cell measurements taken at the bottom of the re-
inforced soil block in the incremental wall test were disappointing due to poor cell
sensitivity. The cell response problem was overcome in the subsequent propped wall
test by using different earth pressure cells with a larger diameter. The results of earth
pressure readings taken during the propped panel wall test at the end of construction
and at the end of surcharging are given on Figure 24. The data shown was arrived
at by calibrating each cell response insitu against the initial 2 m depth of fill placed
during wall construction.
In general, all pressure cells at the base of the test facility recorded vertical earth
pressures in the reinforced propped wall test that were below the value calculated from
soil self-weight and surcharge. Portions of this under-registration may be attributed
to pressure cell response, sidewall friction in the test facility and facing/backfill inter-
action for those cells c1oseo" t.o the front of the test facility. The influence of sidewall
friction was' discussed at lengih during the NATO Workshop. As a result of these
discussions, a large-scale unreirJorced propped wall test was undertaken to examine
the contribution of sidewall fr;~ 'ion and facing friction to stability of soil within the
RMC Test Facility. The results 01 this investigation are reported by Bathurst and Ben-
jamin in these proceedings and show that about 10 to 12% of vertical earth pressure
under-registration at the soil base may be due to sidewall friction. Futhermore, the
unreinforced test showed that the effective friction angle between the rigid facing panel
and backfill with the geogrid/panel connections in place is about 33 0 •
106
130
120
au = ,h + q
50 kPa surcharge for 1000 hrs
Q) ~ ......... ,
]
~
90 ( ' , ........
Q) I ~
I-<
::s
rn
I '\,
"~
rn
Q)
I-< I 50 kPa surcharge first applied}.
0..
t
....""§ 70
~
Q)
>
q. \ \ end of wall construction
-~-~~------
'cI' ........................
50 "'C>- _
------0
The results of temperature measurements taken within the reinforced soil mass
showed that the fill temperature never varied from 21 ± 2°C. As a result, the mechani-
cal properties of the SR2 geogrid material established from laboratory in-isolation tests
at 20°C are considered representative of the insitu condition.
Finally, it should be noted that an attempt was made to monitor vertical settle-
ment of the reinforced sand backfill using displacement devices mounted at the top
of the test facility. However, the installation of these devices proved difficult and no
useful data was recovered.
107
7.1 General
The purpose of this paper is to report the results of two large-scale reinforced soil
walls constructed at RMC. The trial walls provided a starting point for discussions re-
lated to the general theme of the Workshop (Le. application of polymeric reinforcement
in retaining structures). This section discusses the RMC test facility and qualitative
features of the model tests which can assist the reader to interpret the test results.
The sand backfill employed in these large-scale tests has a high shear strength.
The results of direct shear tests show that the material has a peak friction angle
between 42° and 53° and possibly as high as 55 or 56° if plane-strain conditions are
considered (Jewell, 1987b). In addition, it was observed during excavation of earlier
wrap-around constructions that the residual moisture content in the sand gave the
material an apparent cohesion great enough to allow a cut face one meter high to
stand unsupported. The inherently high shear strength of the sand backfill is thought
to explain in part why relatively low tensile loads were generated in the reinforcing
elements.
8.0 Conclusions
Two(2) large-scale model reinforced soil walls were constructed within the soil
retaining wall test facility at RMC. The walls comprised an incremental panel wall
construction and a single propped panel wall system. Both walls were constructed
with a Tensar SR2 geogrid as the sand backfill reinforcement. Each wall was subjected
to sustained surcharge loading up to 50 kPa following construction.
The following points summarize some of the important observations made during
the construction and surcharging of these structures.
1) Both large-scale model walls proved to be stable during construction and surcharg-
ing. The low levels of strain developed in the reinforcement rayers and decreasing
rates of panel movement suggest that both walls would have remained stable for
many years under sustained 50 kPa surcharging.
2) The relative vertical compressibility of the facing units in the two tests influenced
the generation of geogrid strains in the vicinity of the wall panels. In the incremen-
tal wall construction, the panel units were better able to settle with the retained
sand backfill owing to a layer of compressible foam placed at each horizontal joint
between panels. In the propped panel wall this degree of freedom was not present
and additional tensile geogrid straining was recorded in the vicinity of the ge-
ogrid/panel connections as the backfill settled under the action of fill self-weight
and surcharging. These observations clearly have important implications to the
design of these structures. Specifically, details of wall type and construction can
lead to additional geogrid loads which are not accounted for directly in current
design methodologies.
3) The history of panel deformation and strain development in the reinforcing layers
was sensitive to the construction technique employed. In the incremental wall,
significant tensile strains were developed during construction. In addition, larger
outward panel movements were recorded for the incremental panel wall than for
the propped wall test.
4) Estimated geogrid/panel connection loads were very low in these large-scale mod-
els. In the incremental wall construction, the trend in geogrid strains close to the
panels suggests that earth pressures on the facing units are very small due to the
reinforcing effect of the geogrid layers.
5) The distribution of vertical earth pressures below the reinforced soil block in the
propped wall test may be markedly reduced close to the wall facing indicating that
significant backfill/facing interaction occurs. A portion of the under-registration
recorded by pressure cells further from the rigid wall facing is considered to be
due to sidewall friction.
109
Acknowledgements
The authors would like to acknowledge the collaboration of Drs. A. McGown and
K. Andrawes of the University of Strathclyde, Scotland in the conceptual development
and pursuit of this testing program. Appreciation is also extended to D. Varney and
D. Swan, graduate students at the University of Strathclyde for assisting in the con-
struction of the walls. Funding to support this liaison between RMC and Strathclyde
was provided through a NATO Collaborative Research Grant. The authors would also
like to thank Netlon U.K. for carrying out the in-isolation tests used to determine the
load-strain-time data for the SR2 reinforcement and for the provision of the material.
Additional support was provided by members of the Civil Engineering Dept. at RMC
including Messrs. J. Bell, D. Wawrychuk, S. Prunster and J. DiPietrantonio. The in-
isolation tests for strain gauge calibrations were performed by Capt. G. Richardson,
graduate student at RMC. Financial support for construction of the test facility and
reinforced soil wall models was provided through the ARP program and the Chief of
Construction and Properties, DND (Canada). Finally the authors would like to thank
Dr. G.W.E. Milligan for his many fruitful discussions while on a sabbatical visit from
Oxford University.
110
References
Appendix A
II
70
""
c..
~
0;,= 67.7 kPa
81
en
en
lJJ
e: 51 54·2
I-
en 40.6
e: «I
-<
lJJ
:J: JI
en
--cr-_ _-----.....--~---...---..----27.'
2Q
__---------_13.5
II
o .5 1 1.5 2 2.5
HORIZONTAL DISPLACEMENT (mm)
~
I-
:z
lJJ
ffi
<.J
-<
.-J
c..
en
~
Cl
.-J
-<
<.J
~
l-
e:
lJJ
>
..II5-+-~-..,....----r----Y-----.---....,....---I
1 1.5 2
HORIZONTAL DISPLACEMENT (mm)
II
ro 70
=
Cl..
en
en
II
tJ.J
a: !III
~
en
a: c
<
tJ.J
:c !II ~_--".--------~---_ 27.1
en
211
__ ----------------_1~
10
0 .5 1 1.5 2 2.5
HORIZONTAL DISPLACEMENT (mm)
~
:z
UJ
:::L
UJ
u
<
...J
Cl..
en
......
CJ .1
...J
<
u
...... .1
~
a:
UJ .4
>
.2
o .5 1 1.5 2 2.5
HORIZONTAL DISPLACEMENT (mm)
;;
C.
e CY,,: 67.7 kPa
en
en
UJ
a::
I-
en 54.2
a::
-<
UJ
:x:
en
2D 27.1
10 13.5
0 2 3 4 6
HORIZONTAL DISPLACEMENT (mm)
e 3.2
..s
I- 2.
Z
UJ
::&:
UJ 2.4
U
-<
....J
....fk
Cl
1.1
....J
-<
....
U
I-
a::
UJ
>
.4
3 4 6 8 9
HORIZONTAL DISPLACEMENT (mm)
80
10
50
.~
0...
C-
oo
40
....,...
00
Q)
...
00
~ '"
Q)
30
'"
20 1.69
0---0 1.78
10 0----0 1.86
eo
normal stress (kPa)
260
240
220
200
...... 180
0
a.
.>I!
160
W
u
z
W
a:: 140
W
u.
!:
0 120
(J)
(J)
W
a:: 100
I-
(J)
------=-
(7"3=30 kPo
..J
80
~
u 0"3= 30 kPo
~
a:: 60
a.
40
20
0
2 4 6 8 10 12 14 16 18 20
-I
~~-2
et:~
I- -3
Wz
~-
:::::><1: AXIAL STRAIN (%l
..Jet:
01-
>(J)
180
- 160
~
.x
- 140
....
.; 120
VI
(7"3 =60kPa
ILl
~ 100
VI
0:: 80
<t
ILl
~ 60
40
20
Appendix B
80r-----1I----~------~----_r----~~-- ___
70
50
40
30 v PEAI< LOAD
20
10
o
o 5 10 15 20 25 30
STRAIN (%)
Figure B.l
-1 +-----~1_--~---~------_+------~
-2+---~--~--~~rr------~------~
.
I~ ; -]
I z:
~
+----t--+-t----J~_4k_------_+------~
39·6
..• -
~ g§
CIt ...
! ~ -1 +-----+-+---+----'\+-------_+---------1
~
-5 .~--~~---r~~-------T-------~
13·2 ~ 28·7
-,+-------4--------r------~-----
-l~------~------~------~----~
LCIfDt ECPiEMlI
I" 1IIt/44 r It.
"IDIN
TEST TEMPERATlH: 20 t1°C 0
45~--~----.-----r---~----.-----r---~----~
40 100h
35
30
-E
~ 25
o
~
oJ
20
15
10
2 6 8 10 11 14 16
STRAIN (0/0)
5
z
«
f!
VI
4
0~
"-
e
"-
z
-3
,x
VI
10% strain
VI
~
u..
u..
......
t-
VI 2
VI
::>
0
z
0
cr;
~
-
u 1
~
a..
l.U
l.U
cr;
U
0
100 10' 10a 10) 10" 105 10·
TIME (hours) [Log Scale]
50
e 4{)
.
10 ( , " ..
,.
~
."
'-
z .-"
--~
~
-"
}) /
Cl ./
/
-I
/
20 /
/
I..
'10
00 2 4 6 8 10 12 14 16 18
STRAIN (%)
1'4
1'2
0:::
...
1,0
0:::
0
.....
u
Lf 0·8
z
....
0
..... 0-6
u
UJ
0:::
0:::
0
u
0"4
W'
.....
:I:
..... 0·2
0 1)1
10° '()' 1)3 1)" 10- 10·
TIME (hours) [Log Scal~]
Figure B.5 100 Hour Isochronous Curve and Time Correction Factor
versus Log Time Relationship for TENSAR SR2
123
APPENDIX C
TABLE C.I
layer 1 0 4 4 4 4 5 5 5
layer 2 6 9 10 11 13 14 14 15
layer 3 11 13 13 16 18 19 20 20
layer 4 14 18 18 21 24 25 26 27
TABLE C.2
layer 1 1 1 1 1 1 2
layer 2 1 1 2 2 3 4
layer 3 2 2 3 4 5 7
layer 4 3 4 6 7 8 10
TABLE C.3
surcharge
duration
strain (%)
TABLE C.4
surcharge
duration
strain (%)
1. INTRODUCTION
A description of the two tests and the test results are presented
in the companion paper in this volume by Bathurst, Wawrychuk and Jarrett.
127
P. M. Jarrett andA. McGown (eds.), The Application of Polymeric Reinforcement in Soil Retaining Structures, 127-171.
© 1988 by Kluwer Academic Publishers.
128
Table 1
Schedule of Prediction Times
for Proposed RMC Trial Walls
3. METHODS OF ANALYSIS
SR2 LAYER 2
~
~
SR2 LAYER I
~l~D~N!.. ~y~ ~
J
• STRAIN GAUGE
L COOC" ETE LEVELLING PAD ~
_ PRESSURE CELL
e- IJ \ r
z
1f:
SR2 LAYER 2
•"
o z
,.; 0
i:!
~
L J
• STRAIN GAUGE
ET • LEVELLING PAO ..:,.
_ PRESSURE CELL
4. PREDICTION METHODOLOGIES
4.1 General
PREDICTOR A
Al . PHILOSOPHY OF METHOD
a and b constants
e:
(0, - 0.) = (A2)
e:R f
L +
E. (0, - o')f
~
Simulations of the construction and loading for the two walls are
shown on Figures Al and A2.
A3.1 , Fill
, ,,
t
i J ! I /4 STEP 1. Elastic Base + weight of Step L
r I I
• r• •
Z
..
r I
.,
,
I
t
't*
a I =
h
A3.3 Reinforcement
A3.4 Facing
The facing was assumed to have the same stiffness as the concrete
footing and to be linear elastic.
Vconcrete = 0.17
E 0.1 X 10 3 kN/m 2
v 0.49
The layer of fill beneath the reinforced soil block was assumed to
be elastic.
v = 0.35
A3.7 Overburden/Surcharge
v = 0.49
134
A4.l General
REFERENCES
2. Konder, R.L. and Zelasko, J.S. (1963). "A Hyperbolic Stress Strain
Formulation for Sands". Proc. 2nd Pan-American Conf. on Soil
Mech. and Founds. Brazil. Vol. 1, pp. 289-324 .
135
PREDICTOR B
REFERENCE
PREDICTOR C
Cl.l ~
The vertical stress distribution at the base of the wall and other
levels was evaluated by considering moment equilibrium about the centre of
the reinforced fill (refer to Figure Cl).
(Cl)
The stress at the front of the wall (of) now requires to be modified to
take account of friction at the facing.
Cl. 2 Step 2
do
-..J!. (C2)
dz
reinforced fill,
p
z
w
~
rnz
C1.3 ~
do
dT
VK ~ + 20 tan ~u (C3)
dx dx v
In the above equation both the earth pressure coefficient (K) and the
interface friction coefficient (tan ~il = ilw) were assumed to be func-
tions of x.
138
dx
I- ~I
ah ah + dah
T
T+dT..- V
--- T
T
(C4)
where 01
0 3
139
2/5 13 2 13.
(K - K - f3,m)/m 2
a w
C1.4 Step 4
F = 1/2Kyz2 (C5)
~m = ~o + !,\Z (C6)
To obtain the lateral stress, the expression for total force is differen-
tiated with respect to z, i.e.
dF K + 12 yz2 ~ (K)
yz dz = °h (Cn
dZ
- sin I/l
K (C8)
+ sin 1/1
Therefore the lateral stress for rotation about the top is obtained from
the following equation:
While for rotation about the base the conventional expression is used:
°h(base) = Ka Yz (CIO)
- sin cJ>m
where K
a + sin cJ>m
C1.5 ~
Cl.6 ~
PREDICTOR D
and to vary linearly with height. The given values for de come from
observations from other large-scale tests. For the propped panel wall a
greater stiffness, with respect to the incremental panel wall, was
assumed, reflected by de = 1% and de = 1.5%. The assumed amount and
distribution of displacements is supported by experimental data cf. (2),
(3). Following the arguments of (4), creep displacements can be neglected
(experience from granular backfill and limit-analysis based design).
c:
-...
o
I f)
distance
(Dl)
K coefficient of earth-pressure
(D2)
Z = Z + Z (D3)
m p
pk
The following assumptions have been used for predicting the earth-
pressure distributions:
- At the face of the wall the lateral and vertical pressures are
assumed to be correlated via the earth-pressure coefficient (cf.
Figure D2).
p = 1234 kN/m 2
REFERENCES
PREDICTOR E
El. INTRODUCTION
(c) We did not know if the value would change for different parts
of the prediction request.
E2.1 Options
For the incremental (Element-) wall the grids are assumed to take
on the resultant of the earth pressure per face element (O.75m times 1 m).
For the single (rigid) panel wall, the earth pressure allocated to
each grid is determined by lines drawn midway between grids and/or the
horizontal boundaries of the backfill.
14 3m
~I
I I uniform
[ , trapezoidal
I Meyerhof
z
Options:
KaKo KaKo~ K
I"I""'""--r-r-~ K rr--r-__
z z z
0.75m
--------
0.375 m
-- ~---
0 .375m
------
0.75m
---- - - - - -
0.375m
-- ~---
0.375m T2
- ----
TI
- per panel
(incremental wall) (rigid wall)
E4.1 Mathematically
E5.1 Procedure
.
(b) Calculate sum of all 4 forces: LFi = FR·
Lh.
F [_1 - h 1
R 4 r
(E1)
(E2)
149
3m
I -T I +T
~
~
~
c) ideal reinforcement
.~
- T
ttr ··· IT
e) incomplete mobilization f) incomplete mobilization
(triang.) (rect. )
Figuse E5 Various Possible Distribution of Tension and Sh:ar Stresses
F. = a - sh. (E3)
1. 1.
o 2 3
. . . .. . . . . . . . . . . .1. . . . . . . . . . .
.
(m)
0.67
pressure
p = initial overburden stress (including 12 kPa surcharge)
q = additional surcharge
Modulus chosen was the secant modulus for the appropriate stress
level and creep time. For each stress level the appropriate curve was
selected (no correction for "accumulative" creep).
151
PREDICTOR F
F2.1 Soil
On the basis of the results of Direct Shear Box tests and according
to the dry dens ity of the backf ill sand (1790 kg/m 3 = 17.56 kN/m 3 ) an
internal friction angle of 43° was assumed.
where the coefficients aCt) and bet) were obtained from the system:
F2 aCt)
A
(F2)
F2 bet)
B
The input data (FA,£A), (FB'£B) were obtained from the Isochronous
Curves corresponding to FA = 10 kN/m and FB = 30 kN/m.
F2.4 Surcharge
(F3)
Vi = as shown on Figure Fl
- the tensile force versus the distance from the wall face was
represented by a curve, approximated by two straight lines,
exhibiting a maximum at a distance equal t2 xi* from the wall
face (see Figure F2);
r - -
1V5 •
ZI VI Vi
- - - P4 O.9m
--- -
~P3 O.8m
-- - ~P2
H
O.7m
---
---PI O.6m
r{....'\~~A'{(1'" Vf~~,'fo.'t~
Fj max
0.5 m I-.f
F·I
PfaC8j
'----'-=---~- ......- X
Therefore the formulas used for the calculation of the tensile forces in
each inclusion and at the face were:
(Fs)
(F6)
x.*
l.
(F7)
k-l
e: i (Ws k , x, t) {e:[Fi(Ws k , x,t), t=Ol} + ~ j ae: j (F8)
i
(F9)
k-l k-l
~ j M;. =. ~ . {e:[F.(Ws, x),at.l -
J l. J
I J
(FlO)
Another hypothesis (eq. F9) was that the tensile force in the geo-
grids was constant for the entire period of application of each surcharge.
The In-Soil Creep Factor takes into account the difference in creep
behaviour between laboratory and field conditions: without any useful
information on this matter, it was taken as a constant equal to 0.85.
155
50
Ws 40
[kPa] 30 WSj
12~----""'"
n
84
-X
r- I I
~4 ~ ----~-r-~1 +~2+~3
~3--y _____~~ ~1 +~2
~2~ _ _ _ _ _ ~ ~1
~1
L N
J ci(Ws,x,t)dx ~ ~ C •
l.
(W
S
,x., t) • LJ.
J
(Fll)
o 1j
Ai as shown on Figure F4
i N
~ • ~ Ch(Ws,xj,t) • Lj (FI2)
Ih Ij
rex) = Ws + Y • H (F13)
F4. CONCLUSIONS
REFERENCES
1. Berg, R.R., Bonaparte, R., Anderson, R.P. and Chouery, V.E. (1986).
"Design, Construction and Performance of two Geogrid Reinforced
Soil Retaining Walls". Proceedings I I I International Confer-
ence on Geotextiles, Vienna, Austria.
PREDICTORS G, H, I, J
5.1 General
inward outward
T- ____ ~ 1-.-
3.0 measurerl (post-construction deformations only)
jilt
OJ II/
\//\
/ ..
I I
II
I.od' ~- - -1- ~~- ~- --- - -- SR2layer 2
:~!
: Il /
/ .'",{/
r /// \12 kPa surcharge for 100 hrs \
o
i "///~:'f/
-0- - -.0 - / r - - - - - - - - - -- SR2 layer 1
,
G_
\ / \ (total deformations) 1
, . . measured (@ 50 kPa surcharge for 500 hrs) I
'\'I / \. . . . . '" I
L W- -i - - ~~2~a~r] i 9<:- ---d- - - - -1- -
2.0' //
I _1!l9
.§ I I \. 1 I ....( / if
] /.1I' . . '" I
",. /
/ I
I
.,
~ 1.0
I
- aI-. ---/-
'" /
--- ~. .....--.
.....
. .-..t - ~R': 1~'2:
....
:'..-J:> - - -
",-_""_______-, .-_~
_109
I
vt . . ___- -. . . -. . . - - - <>. . /--- . . .
0;-/ ,,/ ,,"""
I SR2 layer 1
6 ~ -/1- - . - - - - - - - - - - - - -
Predicted and measured connection loads at the end of the test are
given on Figure 5. Of all the predictions offered, only the finite-
element model employed by predictor A gave values which were consistent
with the measured values. Other predictions typically over-predicted the
test results.
The RMC test facility at the time of the workshop was only capable
of generating surcharge loadings up to 50 kPa. It was recognized by the
A +200
B 350
C 130
D 1234
E 110
F 365
Predictor Symbol
A 0
B Cl
SR2 layer 4 C 0
D \J
••
2 E 0
F
0 0 0
•
0 G
H
0
2 0 0 0 0
SR2 layer 3
,..--..,
~
'--'
...., Cl
e 0 !!!l 8
>=:
v i
S
v
0
u
J-< 0
.8 0 0 0 0
.....>=:v 2
J-<
SR2 layer 2
0 0 0
.....
>=:
H 0
•
Cl
.....>=:cd
....,J-<
rJJ
• II
0
2 SR2 layer 1
Cl Cl
112 kPa surcharge for 100 hrs I
Predictor Symbol
SR2 layer 4
0 A 0
0 0 B t::.
0 C 0
t::. D 'il
••
4 t::. E 0
t::. F
t::. t::.
•
0 G
0 0
• •
H
• ••
0
2
• •
0 measured
0
0
0 (50 kPa @ 500 hrs)
0
0 0 0 0
,.....-.,
4 SR2 layer 3
8 c
~ t::. t::.
• •
'---"
+" 0
8
~
Q) 0
S 2 measured
.8
Q)
u
I-< • (50 kPa @ 500 hrs)
......
~
Q)
0
I-<
4 0 0 @ ~
•
......
~ 0 SR2 layer 2
......
~
cd
I-<
i • • 0
t::. t::. t::. et::.
• • •
+"
'n 2
measured (50 kPa @ 500 hrs)
0
4
~
• •
t::.
50 kPa surcharge for 1000 hrs
2 0 8 measured (50 kPa @ 500 hrs)
00 0.5
distance from back of panel (m)
r 1 rf-",
3.0
~~ ;,,' range
', of measured
\ \ \ I
U; ~{pa 500 h\ \1
/ values \
2 ,0
@
:§: I "
~
,9
... I \\ \ /
~
I ' , \
\
~
-.; /
1.0
~ \ \ /p 'rJ _0
c?'t \- -
I \
I '
\
'/
'y _r-- \
I / \
t&---; 0- j \
00 2 4 6 8 10 12 14 16 18 20
130
120
cd
0...
~
....
Q)
110
;l
en
en
Q)
....
0..
100
Cdu
'~
....
Q)
> 90
50 kPa @ 1000 hrs
no measured data
80
10 0
researchers at RMC (and by the predictors) that this load level would be
insufficient to fail the trial walls. Nevertheless, the predictors were
invited to submit their estimates of the critical surcharge loading
required to initiate failure. The predicted values fall over a wide range
as shown on Table 2.
rr
outward
1
3.0
measured II /
/
2.0
I II /
I II /
:§: ? I
'<70
1 ?
112 kPa surcharge for 100 hrs I
/
"
.S
10...
I
/
I I /
I / I /
"
?1 /
<l
1.0 d I
0
/
/ / / I
// I I
/i rf
/
20 30 40 60 70 80 90 100
horizontal panel movements (mm)
outward
r
3.0 measured
f /
I
T
I
I /
/
~103
2.0 / /
I /
:§:
"
.S
/' /
J> ,P
/ '
10... / / /
-ll / / /
/ / /
1.0 ? "P ./
"P
I I ./
./
I ./ ./
I I ./
./
./
./
/1 ./
0 if 50 kPa surcharge for 1000 hrs
Predictor Symbol
4 A o SR2 layer 4
C o
o
o•
E
I
J measured
2
0
4
SR2 layer 3
..--.,
~
2
+'
~
Q)
S
Q)
U 0
..s
I-<
4
......
~
Q)
SR2 layer 2
I-< measured
......
~
......
~ 2 0
C'd
I-<
+'
rn
8 0.8 0
0
4
SR2 layer 1
measured
2
112 kPa surcharge for 100 hrs )
00 0.5 1.0
distance from back of panel (m)
Predictor Symbol
A 0
C 0
•
E 0
4 I
0 0 0 J
0 °
0 0 ° @
2 0
0
0 • 0
0
4 0 0 0 0 SR2 layer 3
------
~ 0
0
8
8
•
'--" 0
+" 2 0
>::
Q)
S
Q)
U
I-<
0
...s>::
..... 4 0 0
Q)
I-<
0 SR2 layer 2
0
.....>:: §
.....>::
cd
I-<
2
0°0
• 0
0
+"
[{J
0
4
SR2 layer 1
~
2
8 50 kPa surcharge for 1000 hrs
'} 1 ! f--",
I
/ / "- I
2.0 / I "'- I
"'-
I
~
0
~
1I t---I
/
f ~
\
~23.8
I
1
~ \
I
~H
fI I
50 kPa surcharge for 1000 hrs I
I
~ 27.6
-- -- -- --
___ ___ 1 213
\ I
\I
~
I
~---
I
-- - 1
1
I
~23.5
00 2 4 6 8 10 12 14 16 18 20
load in connections (kN / m)
130
50 kPa surcharge for 1000 hrs
120
"?
P-.
C
~ 110
;::I
[fJ
[fJ
Q)
.....
0..
100
~
.~
...-
.....
Q)
:> 90
measured
80
70
600~----e-----~----------~~--------~
2.0 3.0
distance from back of panel (m)
Figure 10 Vertical Pressure along Base of Reinforced Soil
(Propped Wall Test)
168
A +200
C 101
E 110
F 365
I 540
J 532
Table 3
Predicted Values of Surcharge Load to Initiate
Failure of the Propped Panel Wall Trial
8. DISCUSSION
In general, 1 imi ting equil ibrium methods over-est imated connect ion
loads, facing displacements and strains in the inclusions. However, from
170
an engineering point of view the methods are inherently safe. Perhaps the
most important criticism of these methods is that they are excessively
safe and hence impose an economic penalty when extensible reinforcement is
used in these structures. The finite element approach adopted by predic-
tor A generally gave the most accurate predict ions. Nevertheless, the
model did fail dramatically at some stages in trial wall testing. Speci-
f ically, the model predicted inward facing movements at low surcharge
levels. This observation suggests that despite giving relatively good
estimates of wall performance at the end of each trial wall, much work
remains to be done before FE models (and their "drivers") can be used to
predict the full range of behaviour for these structures.
ACKNOWLEDGEMENTS
The authors would like to thank all those Workshop participants who
gave predictions of the performance of the RMC trial walls. In addition,
the authors would like to express their appreciation to all the partici-
pants for their contribution to the discussions that were held during the
Workshop particularly with regard to interpretation of trial wall behav-
iour and results of analytical models. Finally the authors would like to
thank J. DiP ietrantonio who drafted the figures for this paper and Miss
Martina Lahaie who typed the manuscript.
REFERENCES
8. Jewell, R.A. (1987b). Analysis and Predicted Behaviour for the Royal
Military College Trial Wall, Application of Polymeric Reinforcement
in Soil Retaining Structures, NATO Advanced Research Workshop, Royal
Military College of Canada, June 1987.
AN ASSESSMENT OF THE RESISTANCE OF TENSAR SR2 TO PHYSICAL
DAMAGE DURING THE CONSTRUCTION AND TESTING OF A REINFORCED
SOIL WALL
1. INTRODUCTION
Damage during construction procedures is one of the
factors affecting the physical properties of geotextiles
and related products used to reinforce soil structures.
173
P. M. Jarrett and A. McGown (eds.), The Application of Polymeric Reinforcement in Soil Retaining Structures, 173-180.
© 1988 by Kluwer Academic Publishers.
174
80
70
------
60
E
........ 50
Z
~
Cl 40
«
0- I
AS PRODUCED
30 (average from 8 tests)
--- - - RECOVERED •
(average from 16 tests)
20
Strain rate: 2 %/minute
Temperature: 20o±2°C
10 Specimen size: 5 bars long by 15 ribs wide
5 10 15 20 25 30
STRAIN (%)
Figure 1
AVERAGE LOAD·STRAIN CURVES FROM CONSTANT
RATE OF STRAIN TENSILE TESTS ON TENSAR SR2
175
Table 1: Data from Constant Rate of Strain Testing of As Produced Tensor SR2 Geogrid
used in 'Incremental Panel' Wall.
7
P.2 (Left);
B. 3-7 R.1-17
P.2 (Right);
B.24-28 R.12-2B
55.73
66.59
10.05
17.35
692
692
557
563
As Peak
Load
65.41
. As Peak
Load
20.90
Table 2: Data from Constant Rate of Strain Testing of Damaged Tensor SR2 KeY' P. = Ponel Number
Geogrid used in ' Incremental Panel' Retaining Wall B. = Bors
R. =Ribs
176
Peak load
Number 0/
Specimens Standard Coefficient 0/ Strain at Rupture Strain at
Mean, X' Deviation, Variation an -1 Peak load load Rupture
(kN/m) an-l (kN/m) (%) X (%) (kN/m) (%)
Coefficient 0/ Coefficient of
Mean,X Standard Mean,X Standard
Deviation, Variation Variation
(kN/m) (kN/m) Deviation
a,-l (kN/m) a,-l (%) a -1 (%)
X
a n -1 (kN/m) n X
Table 3: Con~tant Rate 0/ Strain Data from Tensor SR2 Geogrid used in 'Incremental Panel' Retaining Wall.
1week 1 month
---
--- --- 28.7
------ ASPRODUCED
- - - - - RECOVERED
~----------4----------2+------------+---4-------~4-------~--
-2 o 2 3
TEST TEMPERATURE 20 o ± 1°C LOG 10 TIME (hrs) LOADS EXPRESSED
Figure 2 IN kN/m width
COMPARISON OF LOAD-STRAIN-TIME CURVES FOR 'TENSAR' SR2
STRAIN(%) 45
1 hour I. 1
o 5 10 15 20 I I /10hours1
o \
f"'
~;' 100 hours
\
f
r-r-
\ ' 40
\ \ ~ /~ ~
I r1J'
I
\ \ \ / I
1\\\\,\
-1 /1000 hours
~\\\\1
/1
35
i/l, ~
' '-
\
I~1'\\
j ;,; I
30 /;j/
/ I
.I
/Ir VIvi
\ I\ --42.5
\ \\
z I
~ 1--39.6
W
~ 25
o I
\
\ 9 /, j,
!/;
\ 20
\ \ 35.8 'I
r
1 \
-5 f--- 35.1 ...
15
---t-- AS PRODUCED
,/I
28.7
I
-6 r--_
- __ AS PRODUCED ~ -1--'- REC?VERE?
- - - - - - RECOVERED 10
LOADS EXPRESSED IN kN/m width TESTTEMPERATURE: 20'± 1'C
-7
TEST TEMPERATURE: 20 D± 1 DC 5
Figure 3
COMPARISON OF "SHERBY-DORN"
CURVES FOR TENSAR SR2 o
o 2 4 6 8 10 12 14 16
STRAIN (%)
Figure 4
COMPARISON OF ISOCHRONOUS LOAD-STRAIN
CURVES FOR TENSAR SR2
178
Number of Generol
Specime~
Position in Wall
Number Splits Bruises Cuts ~dge Abrasion
Fibri"ation ('Yo ofTotal Area)
Q) ~
"Q)c:
Q) ~ Q)
a.
1
2
P.1 (Centre); B.24-29 R.1-17
P.2 (Centre); B.23-28 R.28-44
0
0
0
0
0
0
. 0
0
10
10
'6~'5
1;;oQ)
3 P.3 (Centre); B.24-29 R.2B-44 0 0 0 0 10
:>00. 4 PA (Centre); B.21-29 R.1-17 0 0 0 0 10
"' ..... '"
Table 4: AmJunt af Visually Assessed Damage from Selected Tensor SR2 Test Specimens.
4. DISCUSSION
The size of the test specimen was chosen so that the
behaviour of the nominal one metre wide 'Tensar' SR2
reinforcing element would be replicated and the influence
of the construction damage would be properly represented.
Extension measurements during creep tests were taken on
the specimen to eliminate possible jaw effects.
179
5. CONCLUSIONS
This project offered the opportunity to examine 'Tensar'
SR2 geogrids used as reinforcing elements of a soil
retaining wall constructed under well defined conditions.
Geogrids recovered from this wall suffered very little
damage in the gravelly sand fill due to the
construction,.testing and recovery processes. There was
no significant change in their physical properties
derived from constant rate of strain tensile testing and
from long-term sustained load tests. The properties of
the "as produced" geogrids can, therefore, be used with
confidence in the predictions of wall behaviour.
ACKNOWLEDGEMENTS
This work was carried out as part of the interaction
between the University of Strathclyde and Netlon Limited
within the S.E.R.C. Teaching Company Scheme.
The authors wish to thank Professor A McGown and Dr K Z
Andrawes of the University of Strathclyde for their
helpful advice during this scheme. Thanks are also due to
the staff of the Royal Military College, under Professor P
M Jarrett and Dr R Bathurst, who assisted in recovery of
the geogrids.
REFERENCES
1.0 Introduction
1.1 General
The use of laboratory models to study the behaviour of soil retaining wall systems is
common in the literature. Typically, experiments are performed in parallel-sided boxes
of finite width with one of the perpendicular faces representing the model wall under
study (e.g. Rowe, 1971). Indeed, the RMC test facility is a rather large example of this
common laboratory approach. Unfortunetly, test facilities of this type can be expected
to introduce some deviation in model response from the behaviour of the same system
with infinite width due to sidewall friction. Strategies to reduce the influence of the
apparatus edge effects include the use of friction reducing surfaces at the sidewalls and
articulated model facings. Both approaches have been employed in the tests reported
in the companion paper by Bathurst, Wawrychuk and Jarrett.
Bransby and Smith (1975) reported the results of small-scale retaining wall tests
carried out using a sand backfill and rigid facings that were allowed to rotate outwards
about the toe. The test results gave a reduction in the active earth pressure coefficient
(Ka) of 14% for a wall height to width ratio H/w = 2 and a sidewall friction angle of
4Jsw = 5.7°.
As a result of discussions held during the NATO Workshop a preliminary inves-
tigation has recently been completed to study the contribution of sidewall friction to
the performance of retaining wall models in the RMC Test Facility. The investigation
comprised the following:
a) Direct shear tests to examine the shear-deformation behaviour of the test facility
sand/polyethylene/plexiglas sidewall interface; •
b) A large-scale unreinforced wall test to determine boundary forces acting on a model
propped wall prior to and at soil failure;
c) A stability analysis of the entire block of unreinforced soil to estimate the contribu-
tion of sidewall friction to the vertical equilibrium of the soil mass after surcharging;
d) An analysis of the stability of the unreinforced test at limiting equilibrium to esti-
mate the contribution of sidewall friction to wedge stability using direct shear test
data.
P. M. Jarrett and A. McGown (eds.), The Application of Polymeric Reinforcement in Soil Retaining Structures, 181-192.
© 1988 by Kluwer Academic Publishers.
182
80
70
60
7p...
50
~
'----'
I-
m
m
Q) 40 Shear Box Size
.....,
I-<
m t:::. 36 cm 2 (Lescoutre, 1986)
I-<
CIl 0 393cm 2 (Jewell, 1987b)
30
Q)
,...q
m 0 36cm 2
0 1m 2
20
10
60 70 80
normal stress an
sand of <P = 53°. Also shown on the figure are the results of direct shear tests carried
out to determine the shear-deformation response of the sidewall surface/soil interface.
The shearing plane in these tests comprised two layers of 0.135 mm thick polyethylene
sheeting over plexiglas with the sheeting left unrestrained. The data from small-scale
tests gives a peak sidewall friction angle of <Paw = 15°. This peak friction angle is lower
than the value <Paw = 20° that was determined from the result of a large (1m x 1m)
direct shear test reported in Bulletin 1 by Bathurst and Jarrett (1986) (see Figure 2).
This test was carried out under very low normal stress and was undertaken to give a
rough. estimate of the sidewall friction angle for prediction purposes and to confirm that
the polyethylene/plexiglas construction did reduce sidewall friction in the test facility.
The figure also shows that an attempt to further reduce sidewall friction by introducing
silicon oil between composite layers was not successful at very low normal stress. Based
on the comments made above it is reasonable to assume that the value <P.w = 20° is an
upper-bound estimate of sidewall friction angle for the trial walls reported by Bathurst
et al. in these proceedings.
183
1.1
1.0
0.9
0.8
7p...
..;.::
........- 0.1 1m X 1m shearbox aQQaratus
><
~
Q)-
~
u
.....
0
'+-<
0.6 6x
.....
o:l
Q)
..d
'"
., 0.5
Cd Fx
Q
0
N
.;::
0
..d 0.4 eand - two
po lyethy lene
1interface
sheets - plexiglas
0.3
0 Test No.1 no lubricant
0.2
• Test No.2 silicone oil between
polyethylene sheets and
polyethylene and plexiglas
0.1
o
05--L~--~-L--L-----------~-------------L~
0.1 0.2 0 .3 0.4 0.5 1.0 1.5
horizontal displacement, 6x (mm)
f..
30
rn
rn
11)
an = 27.1 kPa
10<
+>
rn
10<
III
11)
20
~
rn
:sand/polyethylene/plexiglas Pd = 1.85 (Mg/m 3 )
10
an = 27.2 kPa
00 6 7 8
horizontal movement (mm)
4
1:J sand/sand
i=: 3
S
11)
0
~ ~ 2
S '"0
]~ O~~~~----~---L~----~----~-----L----~----~
~
11)
7 8
>
horizontal movement (mm)
surcharge qo
hydraulic actuator
unreinforced sand
-
backfill 3m
2m
blinding layer \.
------~
O.6m
I .. base slab
6m
- pressure cells
and qb denotes base pressure. It is clear from the figure that losses occur. These
losses are pronounced at the wall facing due to the relatively high wall/soil interaction
developed at this interface (i.e. 6 = 33°) and are diminished further from the facing
panels. A reasonable average R factor is about 0.88 or 0.9 indicating that about 10
to 12% of vertical earth pressure at locations well within the block of soil is lost to
sidewall friction. It is possible that a portion of this under-registration is also due to
cell/soil interaction. It should be noted that the surcharge pressure (qo = 22kPa) in
these calculations has been corrected for the influence of airbag coverage which does not
extend completely to the sides of test facility boundaries when the airbags are inflated.
The reduction in total soil self-weight that is recorded at the base of the test facility
can be denoted by X sw . In addition, there will be a portion of the total surcharge force
that can be expected to be carried by the sidewalls due to sidewall friction. This
component of total vertical force is given the term X . Consider first the contribution
of sidewall friction generated due to soil self-weight: the unit sidewall friction fsw can
be expressed as:
faw = Kswqg tan <Psw (2)
Here qz = "(Z and is the vertical ·stress due to soil self-weight acting at depth z below
the top of the wall and Ksw is the coefficient of sidewall earth pressure. Integration of
equation (2) over both sidewalls gives a total sidewall force due to soil block self-weight
according to:
Xew = Ksw"(H2Ltan<psw (3)
where L is the length of the soil block and H is the soil block height. The additional
sidewall friction generated due to surcharge loading can be accounted for by adopting an
approach similar to that proposed by Jewell (1987b) in this proceedings. The attenuated
vertical stress q" acting at depth z can be described by the relationship:
(4)
where:
C 1_- 2Ksw tan <Psw (5)
W
and w is the width of the soil block. The total surcharge force carried by the test facility
sidewalls becomes:
(6)
Assum:ing that the vertical force acting on the lagging board wall at the back of the test
facility is equivalent to that measured on the front wall, it is possible to calculate the
value of Ksw required to satisfy vertical equilibrium. The results of these calculations
are shown on Figure 5b for <Pew = 15° and 20°. The figure shows that Ksw varies from
about 0.23 to .52 for the range of Rand <Psw considered. This range includes the value
of 0.4 that Jewell (1987b) estimates to be appropriate for the RMC sand under near
plane-strain conditions.
187
0.6 r - - - -..........-----,.....----.----~
0.5
~
rn
~ 0.4
0.3
(8)
where:
2Ksw . (1r
C 2 = --tanr/>swsm - - 0) (9)
w 2
Equations (2), (7) and (8) lead to an expression for the total surcharge force taken by
the sidewalls:
Xq = C 2 wqo tan(~ - 0)
2 i0
H (H - z)e- C2Z dz
The total active force P A acting on the wall can now be expressed as:
(10)
sidewall friction angles <Psw = 15 and 20°. The wall facing/soil friction angle 8 was cal-
culated directly from the measured wall forces at failure. Superimposed on the figure
are the measured ranges of wall forces acting at limiting equilibrium in the large-scale
unreinforced wall test. The Ksw = 0 condition represents the ideal situation of no side-
wall friction. As may be expected, the measured facing loads fall below these values
indicating that sidewall friction does occur. The lateral confinement of the sand soil in
the RMC test facility suggests that the fully-mobilized soil friction angle is somewhat
higher than the value of 53° measured in direct shear tests (e.g. Rowe, 1969). Jewell
(1987a,b) suggests that a value of 55 or 56° may be appropriate. If this is the case,
the percentage reduction in P A due to sidewall friction is of the order of 12 to 18%.
For <p = 55° and sidewall friction angles of <Psw = 15 to 20° the corresponding range for
Ksw is 0.22 to 0.5. This range of values is in agreement with the results of the vertical
stability analysis described earlier and the value of 0.4 proposed by Jewell (1987b).
rz
()
40
cPsw = 20°
cPsw = 15°
1= 18kN/m 3
{) = 33°
35 qo = 22kPa
Z
6 ......
..e cP
«
0... 30 }50
25 }53
}55
20 0 0.3 0.5
Ksw
30
cPsw = 20°
cPsw = 15°
1= 18kN/m 3
25 {) = 33°
qo = 22kPa
~
Z cP
..!<:
20
~
>
« }50
0...
15
measured range of wall force at failure
10 0
0.1 0.2 0.3 0.4 0.5
Ksw
Acknowledgements
The authors would like to acknowledge the contribution of J. Bell (Research As-
sistant) who helped with many aspects of the test program reported here. The authors
would also like to thank J. DiPietrantonio who drafted the figures and Dr. P.M. Jar-
rett for his input during this investigation. Funding for this investigation was provided
through the ARP program and the Chief of Construction and Properties, Dept. of
National Defence (Canada).
192
References
Abstract
Predictions are made for the incremental and propped reinforced
soil walls built at the Royal Military College, Kingston, and
these illustrate the theory and analysis presented in a
companion paper. New ideas are applied to the interpretation
of shear box data to derive parameters for the sand, and
supplementary test data are reported. The predictions are
based on two "bounding" equilibrium states for reinforced soil
walls, and the influence on these of the parameters and
boundary conditions in the trial walls are investigated. The
side wall friction appears to have a major influence: a closed
form analysis is presented for side wall friction so that it
can be taken into account for both the self weight and sur-
charge loading. The predictions are then presented and
compared with the measured data. As all aspects of the wall
behaviour under different loading conditions were predicted
this provides a comprehensive test of the analysis. The find-
ing is that the analysis captures almost all the main features
of the observed behaviour, particularly the patterns of behav-
iour, but also the magnitudes. Further, it provides a natural
link between the results of the two different trial walls.
INTRODUCTION
1.1 Introduction
This is a companion paper to "Reinforced soil wall analysis and
behaviour" by R. A. Jewell published in the same proceedings,
and hereafter referred to as the companion paper. Although all
the theoretical concepts and the analysis are described in the
companion paper additional details about the theory are brought
out by the prediction for the RMC trial wall.
The data for the trial wall were presented in Bulletin 1 by
Bathurst and Jarrett (1986).
193
P. M. Jarrett and A. McGown (eds.), The Application of Polymeric Reinforcement in Soil Retaining Structures, 193-235.
© 1988 by Kluwer Academic Publishers.
194
The (dry) density for the compacted sand in the trial wall is
expected to be 17.6 kN/m3, a relative density of 52%. The bulk
density of the slightly moist sand with the anticipated 3%
moisture content has been taken as 18 kN/m3 for the prediction.
Because the soil is relatively coarse grained the low moisture
content will probably not significantly influence the effective
stresses in the soil. The properties for the dry sand were
used for the prediction.
-- --
~ ~~- ~ ~
v
"1-- OXF Small, RO 150"
La..
0 52 -)to-.
~~~~~
x
-
_~
+_x ~-
-----
~ ~ ~.
C> + "
~ 48
+
~J:
~.~-----
--- • +
Ul
44 ---------------~ --.J!. __
~ •
0:: <40
i5 &tImated critical IItcIte
~'
a.' 36
i 12 kPa can t50 kPa can
2 2.4 3.2 3.11 4 4.4
(I)
13 1.4
~ 1.2
~
~ 0.8
"13
(I) 0.6
• OXF La/oge, 17.5 lcPa
0.4 . OXF l.arge, 32.5 lcPa
~ 0 OXF l.arge, 47 kPa
0.2
~
::t:
I> OXF Small, 47 KPa
(I) 0
0 2 4 6 8 10 12
(0) SHEAR DISPLACEMENT x:mm
E 2.2
E
>.
1.8
~
a 1.04-
~
a.
~
~ 0.6
~ O~~~~~________________________~
2 6 6 10 12
(b)
SHEAR DISPLACEMENT x:mm
Fig. 2. Direct shear tests on the RMC sand completed at Oxford
University.
60
III
50
0'1
Q)
"U
40
z
0
~
0:: 30
lL.
• Direct shear result
lL.
0 A Angle of dilation
w 20 Flow rule, 39 c.l/.
...J
0 Flow rule, 37 c.y.
Z
«
10
0
0 2 4 6 8 10 12 14
SHEAR DISPLACEMENT x:mm
Fig. 3. Direct shear test on the RMC sand with enforced
symmetrical deformation. Interpretation for the plane strain
angle of friction using a flow rule.
The same friction angle was adopted 'for both the 12 kPa and
the 50 kPa surcharge load cases. This was because the
addi tional soil strain under the higher surcharge is
anticipated to cause additional soil shearing resistance to be
mobilised in the sand counterbalancing (to well within the
accuracy of the prediction) the influence of the increased
mean stress level in the soil.
The angle of dilation corresponding with the mobilised angle
of friction is ~~ IS·, from Bolton's flow rule and the
estimated critical state angle of friction.
2.5 Estimated soil shear modulus for compatibility
A simple elastic plastic model for the sand is proposed for
making estimates of strain compatibility. An interpretation
for the direct shear test has been presented to quantify the
relationship between the mobilised soil frictional resistance
and the principal tensile strain. The relevant equations were
derived in the companion paper.
The shear modulus for use with the compatibility curve is
estimated from the stage of the direct shear test between the
critical state stress ratio, marking the end of the
substantial rotation of principal axes at the beginning of the
test, and the selected mobilised plane strain angle of
friction.
202
() L1(t/s) 3 10- 3
E 3 50 = 2 ( Gis) (5)
0·875 m
0·75 m
3m - - - 3 m
Layer 2
0·75m
__JL~La~y:er:I::::::::~~~0.625m
(a) Reinforcement spacing (b) Self weight loading (c) Surcharge loading
Layer 4 9 % 29 %
Layer 3 21 % 25 %
Layer 2 33 % 25 %
Layer 1 37 % 21 %
Total 100 % 100 %
5.1 Introduction
The aim is to predict the reinforcement forces and force
distribution, estimate the corresponding reinforcement strains
and hence the overall elongation in each reinforcement layer
at the end of the loading period. This will be completed for
three mobilised soil strengths covering the likely range for
the trial wall. This will indicate the sensitivity of the
predicted behaviour to the soil strength properties.
206
E
2.6
• 45 degrees
50 degrees
0:::
W 55 degrees
~ 2.2
C>
z
is 1.8
z
:::i
CD
w 1.4
6
CD
~
l-
I
C>
iii 0.6
I
o 2 4 6 8 10
(a) ( b) (c)
(d ) (e) (t)
~ 0.5
LAYER 2 ~ 0.5
LAYER 4
~ 0.4 ... 0.4
a 0.3
iii
Gi 0.3
~ 0.2 ~ 0.2-1---""",-
~
~ 0.1
ft
~ 0.1
a::: 0 0 0.4 0.8 1.2 1.6 2 2.4 2.8 0::: 0 0 0.4 0.8 1.2 1.6 2 2.4 2.8
DISTANCE FROM THE WAlL FACE :m DISTANCE FROM THE WAlL FACE :m
:m H HP :kN :mm
/ ~
0.2
I .
l!; 04
/ !
1 06
~
/
.
~ 0.8 ~
...- ~
1
0 0.2
OUTWARD MOVEMENT
(~ 0.4
0
0 2 •
REINFORCEMENT ELONGATION :mm
10
(A) (8)
!'i
3
... 411. lRUNCATED
411. IDEAl.
0::
w
~
(!) 2
z
5
z
::J
m
~
$!
z
0
~
~
0
0 2 4 6 8 10
REINFORCEMENT ELONGATION :mm
6.1 Introduction
The additional movement from incremental construction
represents a "moving datum" for the reinforcement layers
because they are attached to panels which are aligned over the
previous layer of panels which have already moved outwards.
At any elevation where there is reinforcement in the wall, the
net surveyed outward movement of the face from the vertical
line through the original position of the lowest panels is
made up from (1) the elongation in the reinforcement layer and
(2) the initial position of the unstressed (or lightly
prestressed) reinforcement layer from the vertical line.
6.2 Panel geometry and movement
The trial incremental wall is built from four 0.75 m high
panels each with one reinforcement layer attached, which means
that the incremental displacement will be the saIl1e for each
construction lift. The reinforcement is attached to the panel
at the lower third point, and the layers of panels are not
connected top to bottom. This gives each panel the maximum
freedom of movement.
The actual panel movement in the trial wall is unlikely to be
pure translation because the net soil stress on the panel will
act higher than the lower third position of the equilibrating
reinforcement force. The panel movement is also unlikely to
be pure rotation which would occur if the bottom of the panel
were fixed to the top of the panels in the layer below. Both
these extreme cases are calculated, so that the actual net
movement would be expected to lie somewhere in the range.
6.3 Soil strength and compaction load
Incremental construction movement occurs during ini tial
filling and compaction. The deformations in the soil and the
force in the reinforcement are relatively small. Therefore
the appropriate mobilised angle of friction for the soil
should approximately represent "at rest" conditions or a
little higher. For the RMC sand this would be equivalent to
¢i ~ 35°. A range of values rjJ = 30°.35° & 40° are used below for
comparison.
The other parameter which must be evaluated is the uniform
surcharge equivalent to the compaction and construction
loading. For the trial wall the dead weight of the vibrating
plate compactor is 1 kN/m2. The static equivalent to the
dynamic load depends on the frequency and amplitude of
vibra'tion. Construction loading also comes from people and
equipment moving the fill, and any temporary overfilling. The
213
;3 ;3
g g
~ ~
"~ 2 "~z 2
::J ::J
01 01
I
""-
~
..30_
40_
1lWOIIJJING PNtIl.S
.. 40_
~
z z
~ + ... -
~ + ... -
~ ~ 00
2 4 6 8 10 12 14 4 8 12 16 20 24
OUlWARD MOVEMENT :mm OUlWARD MOVEMENT :mm
(a) (b)
3,---------------____________________,
NO CONSIRUCTION MOVEMENT
• .50. TRUHCA.TED
+ 50. IDEAl.
WITH CONSIRUCTION 1.t0VEM
~ 50. lRUNCATED
2 6 8 10 12 14 16 18 20
FACE MOVEMENT :mm
connection between the face and the rigid base. Finally, the
rough strong base to the wall is assumed not to influence the
force in the lower reinforcement layers.
In the incremental wall the facing panels do act freely from
one another which is probably as close to the ideal
assumptions as possible. Thus the interaction of the facing
and the base should only influence the force in the lowest
reinforcement layer. The maximum possible reduction in the
reinforcement force in the lower layer due to the rough base
was estimated to be about 20% (section 4.1).
7.2 Side wall friction
The problem always encountered with laboratory and field
trials is at the side boundaries to the structure. Side wall
friction can provide significant stabilising forces in the
soil. Attention was paid to the possible influence of the
side walls for the RMC trial and a measured value of side wall
friction I/J,w = 20' is reported from a large scale shear box test
on the prepared boundary (bulletin 1).
Bransby and Smith (1975) made detailed experimental observa-
tions and presented results from a numerical analysis for the
influence of side wall friction on model retaining wall tests.
An analysis is derived below to evaluate the influence of side
wall friction in the RMC trial wall. A simple closed form
analysis is presented which slightly underpredicts the results
of the numerical solution by Bransby and Smith (1975).
7.3 Side wall friction: analysis for self weight loading
If there is side wall friction then the shear stresses
generated between the side wall and the deforming soil will
act to resist the deformation in the soil. Because there are
additional resisting forces, the side wall forces reduce the
wall pressure required to support the soil, or reduce the
required reinforcement forces in a reinforced soil wall.
Bransby and Smith (1975) show that the side' wall shear
stresses change both the failure mechanisms and the stresses
in the soil. For a simple analysis an assumption can be made
that the critical failure mechanism in the soil is not
changed, and that the side wall friction only acts to alter
the overall force equilibrium. The calculation below is for
the triangular wedge of soil between the most critical plane
and the wall face, the equilibrium of which determines the
maximum reinforcement force.
The most critical plane and the equilibrium force polygon for
the soil wedge are shown in Fig. 12a and b . The required
horizontal force for equilibrium Po is provided by the
reinforcement layers. A side wall force p. acting in the
opposite direction to the movement of the soil wedge changes
the force equilibrium as shown in Fig. 12c and d. For a
simple analysis the soil movement can be assumed to be
parallel to the most critical mechanism.
216
(a) (b)
The side wall force is derived from the normal stress and the
frictional shearing resistance between the sand and the side
wall. The force is found from summing the shearing stress on
the side wall over the cross sectional area of the critical
wedge, as shown in Fig. 12e.
The intermediate stress a 2 has been observed to remain a
constant function of the mean stress s for plane strain
loading,
(6)
(7)
(8)
217
Pa = Ij( . (10)
W V1\.a
Resolving the net side wall force into horizontal and vertical
components, and by symmetry (Fig. 12d and eqn. (10))
( P a ' + P sh J = fK: (12)
(w -?v )
which gives a direct expression for the reduced active earth
pressure coefficient, from equations (11) and (12),
(13)
The components of the net side wall force are, from Fig. 12,
P sh = P s cos ( 45 + ¢ / 2)
P sv = P s sin (45 + ¢ / 2)
The two analyses compare well although the simple closed form
analysis giving a slightly lower reduction in the active earth
pressure, coefficient. One obvious conservatism in the simple
analysis for self weight loading is that it ignores the slight
reduction in the vertical stress in the soil due to !' arching"
between the side walls.
qz/qo
0+0_ _...1.-_--11 1
z/H
(a ) ( b)
(14 )
(15 )
8.1 Introduction
A prediction for the RMC trial wall allowing for the influence
of the side wall friction can now be completed. The
prediction is for the expected plane strain mobilised angle of
friction in the soil ~= 50' (section 2.4), and for the measured
angle of side wall friction ~~= 20'. The prediction is for the
central loading case (12 kPa/100 hours) and for the end of the
test (50 kPa/1000 hours).
For both loading cases the difference between the prediction
with and without the side wall friction is indicated.
An implicit assumption in the analysis for side wall friction
is that it only reduces the magnitude of the required
reinforcement forces for equilibrium and does not change the
location in the soil where these forces are required. This
assumption is likely to be conservative. The side wall
friction is likely to reduce the overall elongation in the.
reinforcement layers more than predicted.
222
Surcharge loading
The same form of presentation can be used for the surcharge
loading. In this case the gross required force is reduced by
more than the 40% due to the apparent earth pressure
coefficient. The reduction of stress in the soil due to the
"arching" over the side walls (section 7.4) gives an
-additional overall reduction in the gross required force of
15%. The resulting reduction in the gross required horizontal
force to support the surcharge load is 49%.
The "arching" of the surcharge loading alters the distribution
of the gross required force between the reinforcement layers.
The distribution between the reinforcement layers of the
reduced gross required force due to surcharge loading is given
in Table 7. The magnitude of the resulting forces in the
reinforcement layers for 12 kPa surcharge loading are also
given. The values can be compared with the results in
Table 4.
Total reinforcement force
The resulting maximum force in the four reinforcement layers
for the central loading case and at the end of the test are
shown plotted in Fig. 15a and b, where they are compared with
the values without side wall friction.
223
i
II>
\ 1
1 \ ~
liO !cPa 1000 HOURS
0 .6 ~0.6 • IIIOCJrtt SIIE 1N.lS
0.2 0 2 4
\ 6 B 10
iii
:>: 0.2
o
.. SllEIIIU...:::noN
2 4 6 8 10
MAXIMUM REINFORCEMENT FORCE :kN/m MAXIMUM REINFORCEMENT FORCE :kN/m
Reinforcement Elevation z oK P 0
:m H HP :kN :mm
The calculation for the central loading case and the truncated
length equilibrium case, and allowing for side wall friction,
is set out in Table B. First the depth of each reinforcement
layer beneath the wall crest is determined and the non
dimensional outward movement read from the relevant chart
(Fig. 8a for this case, but see the companion paper for the
other charts). The maximum force in each layer is recorded in
Table 7, and the reinforcement stiffness and the wall height
are the same for all the layers. The magnitude of the outward
movement can be directly evaluated, Table B.
The predicted set of reinforcement elongations for the two
equilibrium states of ideal length and truncated length are
shown in Figs. 16 and 17. The difference due to the side wall
friction is indicated for the two loading cases.
3
NO SIDE FRICTION
~ • lRUNCAlED SIoAOOTH
0:: ,. IDE'AL SIoAOOTH
W
WITH SIDE FRICTION
~ <> lRUNCAlED ROUGH
t. IDE'AL ROUGH
C,!)
z 2
is MEASURED
z l( PANELS
::J v WIRES
m
w
e;
~
z
0
~
~
w
0
0 2 468 10
TOTAL REINFORCEMENT ELONGATION :mm
3~------------------------------------~
- MEASURED
0+-~~~~~~~~~~~=T=r~T=~
o 2 4 6 8 10 12 14 16 18 20
TOTAL REINFORCEMENT ELONGATION :mm
9.1 Introduction
The analysis for walls involves two idealised equilibrium
states which are expected to bound the actual equilibrium in
the reinforced soil. The data from the measured behaviour of
the trial walls have been added to the figures in this and the
previous section.
9.2 Deformations
The total deformation at the wall face is made up from the
reinforcement elongations and the "moving datum" caused by
incremental construction.
226
Reinforcement elongation
The predicted reinforcement elongations for the central
loading case are shown in Fig. 16. The results for the ideal
length case are also shown, as are the maximum likely
elongations that were predicted assuming no side wall
friction.
The predicted reinforcement elongations at the end of the test
are shown in Fig. 17. Again, the companion ideal length case
is also given, as are the maximum likely elongations
calculated ignoring the side wall friction.
Incremental construction movement
Because the facing panels can either translate or rotate there
is a range of possible movement due to incremental
construction. The predicted range is shown in Fig. 18, based
on a mobilised angle of friction if> = 35' consistent with the
initial filling and compaction of the sand (section 6.3).
3~-----;~~~~------------------,
MEASURED
.. PAt-JEL TRANSLATION
Q PANEL ROTATION
0~---r---,----,----,---,----'---'----1
o 10 20 30 40
INCREMENTAL CONSTRUCTION MOVEMENT :mm
!z 0.4 T."~. !z
w
0.4
~ O~ Ci
(,)
0.3
..-
~ 0.2 VI '"~ 0.2
0.6 0.6
z z
LAYER 2 LAYER 4
~ 0.5
0.4
~
!z
0.5
0.4
!z
w
Ci 0.3 ill 0.3
(,)
ii 0.21-7"<,-._-....
'"z~ 0.2
~
Z 0.1
0.1
iii
ill
~
!:/~
z
LAYER 1 LAYER 3
~ 0.8
!z
w
0.6
2i
- _.
:::!
() 0.4
w
() 0.4
0:: 0::
~
z 0.2 ~ 0.2
Z
iil
0:: V·,
O~~~~~~~~~~~~
iil
0::
O~~~~~~~~~~~~
o 0.4 0.8 1.2 1.6 2 2.4 2.8 o 0.4 0.8 1.2 1.6 2 2.4 2.B
,
DISTANCE FROM THE WAlL FACE :m DISTANCE FROM THE WALL FACE :m
z z
LAYER 2 LAYER 4
~ 0.8 ~ 0.8
+--f-J~=
!z
w
0.6 !z
w
0.6
:::!
w
()
0::
0.4 ~
0::
0.4
~ 0.2 ~
Z 0.2
Z
iil iil
0:: tl!
O~~~~~~~~~~~~
o 0.4 0.8 1.2 1.6 2 2.4 2.8 00 0.4 O.B 1.2 1.6 2 2.4 2.B
DISTANCE FROM THE WALL FACE :m DISTANCE FROM THE WALL FACE :m
§ :3
~
(!)
z 2 PROPPED WALL
az PREDICTED
::J
CD & 1lDL .......
-
~
• lRUNCAlED IIOUGH
MEASURED
~ T
Z
0
~
~ 0
0 2 4 6 8 10
TOTAL REINFORCEMENT ELONGATION :mm
Fig. 21. Predicted face movement for the propped wall under
12 kPa surcharge, compared with the measured movement.
The predicted face movements for the propped wall at the end
of the 12 kPa surcharge loading, and at the end of the test,
are shown in Figs. 21 and 22. The predicted movement is due
only to the reinforcement elongation and is exactly equal to
the equivalent prediction for the incremental wall. The
prediction allows for the measured side wall friction.
230
3
Ii
~
"isz 2 PROPPED WAIl.
PRf.DICIe)
·
•
~ IIE'II._
11IIJNCATED _
lD
I
Z
0
•
II£ASURED
WIllES
~
~
2 4 6 8 10 12 14 16 18 20
TOTAL REINFORCEMENT ELONGATION :mm
Fig. 22. Predicted face movement for the propped wall under
50 kPa surcharge, compared with the measur~d movement.
STRAIN COMPATIBILITY
11.1 Introduction
The aim of this section is to relate the measured soil and
reinforcement properties on a compatibility curve for the
incremental wall (companion paper, section 6.2). Only the
overall compatibility for the whole wall will be considered -
a more detailed analysis would consider equilibrium in each
layer. The influence of the measured side wall friction is
included in the analysis.
E
"-
z 24
-!!
w 20
tJ
'"~
16
~
~ 12
~
0
~ 8
fil
'"5 4
fil'
'" 0
0 0.002 0.004 0.006 0.008
PRINCIPAL TENSILE STRAIN
<40
E
~ 35
w
0 30
D:
f2 25
~
~ 20
~ 15
0
~
10
iii
~ 5
~
0.002 0.004 0.006 0.008 0.01 0.012
PRINCIPAL TENSILE STRAIN
Acknowledgements
The author would like to thank Dr. R. Kaniraj of the Indian
Insti tute of Technology, who carried out four of the direct
shear tests reported in the paper while at Oxford University
on a British Technical Cooperation Award.
The foresight and hard work of Richard Bathurst, Peter Jarrett
and Alan McGown, and their co-workers, was essential in that
they made the prediction symposium possible, designed the
trials to provide the detailed data of wall performance, and
provided the forum for the exchange of knowledge that seems
certain to advance the state of the art.
References
Bathurst, R.J. & Jarrett, P.M. (1986). Class A prediction
exercise for reinforced earth walls. Bulletin No. 1 for NATO
Advanced Research Workshop, Application of Polymeric Rein-
forcement in Soil Retaining Structures, Royal Military
College, Kingston.
Bransby, P.L. & Smith, I.A.A. (1975). Side friction in model
retaining wall experiments. Journal of Geotechnical Engineer-
ing, ASCE GT7, July, 615-632.
Bolton, M.D. (1986). The strength and dilatancy of sands.
Geotechnique 36, No.1, 65-78.
Cornforth, D.H. (1973). Prediction of drained strength of
sands from relative density measurements. ASTM Spec. Tech.
Publ. 523, 281-303.
Jewell, R.A. (1987). Reinforced soil wall analysis and
behaviour. Proc. NATO Advanced Resear.ch Workshop, Application
of Polymeric Reinforcement in Soil Retaining Structures,
Martinus Nijhoff.
Jewell, R.A. & Wroth, C.P. (1987) Direct shear tests on
reinforced sand. Geotechnique 37, No.1., 53-68.
Koerner, R.M. (1968). The behaviour of cohesicmless soils
formed from various materials. PhD Thesis, Duke University.
Soil Mechanics Series No. 16.
Leroueil, S. & Tavenas, F. (1981). Pitfalls of back
analyses. Proc. 10 Int. Conf. Soil Mech. & Fndn. Engng. ,
Stockholm, Vol 1, 185-190.
Stroud, M.A. (1971) . The behaviour of sand at low stress
level in the simple shear apparatus. PhD Thesis, Uni versi ty
of Cambridge.
Case Histories and Full Scale Trials
POLYMERICALLY REINFORCED RETAINING WALLS AND SLOPES
IN NORTH AMERICA
1. INTRODUCTION
This paper presents a summary of reported polymerically
reinforced soil wall and slope projects constructed in North
America. The projects included walls and slopes where the
reinforcement was used to resist lateral earth pressures and
prevent internal failure through the face or toe of the
structure. projects where reinforcement has been used to
provide increased stability against deep seated failure of
embankments were not included. Specific emphasis was placed
on projects that have been instrumented and monitored such
that performance assessments could be made.
General Information
Design Information
Reinforcement Properties and Design Methodology
Instrumentation
2. PROJECT SUMMARIES
The project summaries for each specific review category are
contained in the attached Tables 1 through 4. Table 1
includes general project information and provides references
where additional information may be obtained. Table 2
provides the design requi rements for the specific projects
including height, surcharge loading, and foundation and
backfill characteristics. Table 3 presents information
pertaining to the actual properties of the reinforcement, and
the design methodology used to determine the reinforcement
requirements. Finally, Table 4 presents project
instrumentation and monitoring information. Unfortunately,
only a few cases were reviewed which contained this extremely
valuable and needed information.
In all, summaries of 54 projects are provided in the tables.
The summaries include 13 projects that were at least
239
P. M. Jarrett and A. McGown (eds.), The Application of Polymeric Reinforcement in Soil Retaining Structures, 239-283.
© 1988 by Kluwer Academic Publishers.
240
------------------------------------------------------------------.::----------------------------
Water down Slope failure
10 CP Rail Ontario 1982 repair 4.6 and 5.5 91
Wauwatosa Reta in i ng
12 Wauwatosa WI 1982 wall 8.2 64
Nonwoven Bidim
needlepunched C-28 & C-38 Pol yester
Nonwoven Trev I ra
needlepunched S1115 & S1127 Pol yester
Nonwoven Typer
heatbonded 3401 & 3601 Polypropylene
Wrapped I: 4 Sandy cl ay
slope Woven Mirafi 600x Polypropylene si It backfi II
no facing
Wrapped Nonwoven
gun Ite need I epunched Supac 12NP Pol ypropy lane
googr Id Tenser
uv stabilized Nicolon
treated wood woven Goo I on 500 polypropylene
I to 2 foot
layer spacing
granu lar
bockflll
244
3 .t 6.1 11
6.1 11
6.1 11
2.t7.6 15
2 at 7.6 15
Phi ladelphia
41 Conra i I PA 1987 rai I line 6.7 90
Mlraf I attempt to
1: 1 slope ge09" I d Paragr I d 505/ po I vester determ i ne so i I 26
50S & 5T tabr Ie
S I gnode Interaction
geogrld TNX 5001 & pol vester mechan I sm for
TNX 250 cohesive soils
non woven
geotextlle Qullne polyester
IH:2V
slope
wrap around wI
precast coneret geogrld Tenser SR2 HOPE 36
bk fal se face
Rottvdam stream
43 Rotterd~ NY 1986-87 diversion t!. 5.5 )80
parking lot
Tyler land
46 Res i dence Inn TX 1985 deve lopment 6.1 104
Westbank prOTotype
47A Elliott Plaza British 1985 retaining 2.4 33
Columbia wall
478 2.4 33
Kelowna retaining
48 Smith British 1985 wall 2.4 3D
Col umble
Naramatta Pentkton
49A Road Br i t i sh 1985 1.8 29
Columbia
498 2.4 32
Gorman Westbank
51 Shop British 1986 3.7-4.9 32
Columbia
Borges Vernon
52 Rehabll itatlon British 1986 3.1-4.9 15
Col umbla
CNR Vernon
53 Overpass British 1986 2.4 162
Col umbl a
Kelonna
54 Torhlelm British 1987 1.2-6.1 99
Col umbia
247
full height
concrete geogr j d Tenser SR2 HOPE 36
full height
concrete geogr i d Tenser SR! HOPE 36
masonry
blocks geogr 1d Tensar SR2 HOPE 36
full height
concrete geogrld Tenser SR2 HOPE 36
po I yethy I ene
waf t I eerete geogr i d Tensar SR2 HOPE
waf f I eerete
Tenser SRI
SR2
248
I mater of fill
3.0 0.23 - 0.30 3.0 duel tandem rock
axel loads
311 kN
4 6.7 0.23 4.0 dual tandem rock
whee' load
dolomitic
12 8.2 0.2 - 0.4 3.4 limestone
bedrock
col laps ib Ie
18 0.9 - 6.1 0.3 - 0.9 3.7 12.0 kP. soi Is
soft clay wI
19 6.1 0.3 - 0.6 3.7 12.9 kPa crushed stone 24
on top
dense sand
20 5.3 0.3 3.5 21.0 kPa and gravel at surface 0 30
-~---------------------------------------------------- ------------------------------------------------------
249
clean uniform
concrete 36 15.3 4.5 lean clay
sand
rounded,well
graded clean 35 20.4
sand & grave I
si Ity si ltv
Beaumont clay 20 15 Beaumont clay 20 15
on-site si ltv
clay 24 20
granu I ar 2.4 32
we II graded we II graded
sandy grave I 34 21.0 sandy gravel 30 2).0
250
un i form bedrock 33 -
28 1.5 - 4.3 0.3 - 0.6 clay at 3 to 4.6 m 57 20
at minus
36 0.23 - 0.61 3.8 clay strati tied 0.3 m 50
at minus
37 3.4 - 6.1 0.61 varIes clay strati tied 2.1 m 100 0
weathered
39 3.0 - 7.3 0.15 - 0.30 2.3 - 3.4 rock
gravely at minus
40 6.1 0.38 & 0.76 4.3 3m sand sand un I form 0.5 m 0 39.0
GW 1st 1m
43 5.5 0.3-0.6-1.2 4.4 12 kPa GW clay & silt 45
d w SOIL CO 0° d w SOIL CO 0° d w
kN/m 3 (%) TYPE kP. kN/ml (%) TYPE kP. kN/m 3 (%)
30- 30-
granu lar 35 18.1 granular 35 18.1
75mm minus
sand &- gravel
inorganic clay
15.7 25 sand o 30 -- of low to med
plasticity
5 0
13.3 37 clay ---------- 13.3 --
o 12
weathered
granu lar rock
gravely gravel y
20.2 samd o 39.0 20.2 2 sand o 39 20.2 2
sl It 13 35.5 17.0 19 sl It 13 39.5 17.0 19
2.4mwall
47B 2.4 0.6 & 1.2 2.8 above GP variable o 33.5
fine river
48 2.4 0.6 & 1.2 1.8 SP sand -1m
4.8kPa
50 1.8-4.9 0.6 & 1.2 2.9 of backf III SP 35
B.C.F .S.
51 3.7-4.9 0.6 & 1.2 2.9 L-45 GP 35
GP 32 17.3 -- GP 32 17.3 --
GP 32 17.3 -- GP 32 17.3 --
CASE METHOD OF TENSilE MAXIMUM STRENGTH Ie REOUCTJON FACTORS EXTERNAL INTERNAL STABILITY
No. STRENGTH STREKGlll ELONGAT 100 ---------------------------- 5TABIL I TY
EVALUATION kN/", IS) TlkH/JIIJ (S) DLRAaIU TY CREEP MODEL VERT. STRESS H.STRESS OISTR PULLOUT
"
'00 te$t 12.) o ....r-turnln g wedge overburden 2/3 0
".
Flb.-tell 25_ strip sliding tieback
'00 test 1 1.4 O\I.,.turnlng ",edge overburden Ko 2/3 0
Flb..-tex
600 17.1
Bldl.
C-28 10.7 60
------;j';;;--------:;----------------------------------------------------------------:--------------
C-Ja 18.9
Bldllll Tieback
C-34 greb test (1000 Nl 1J.1 wedge 2/3 0
'.8
,..
Flbertex 2'00 wid e wIdth tST 140 tIeback
9 ------- w- 2oonn wedge
Fl bertex 400 1- 100II1I'I 10.0 14'
Supac 41f' 12.6
" '.1
...
t-l01 ----------------------------------------
Supac 6If' 24.3 60 9.8
'.7 80
Typar }401
16.6
7.7
"
60
10.9
3.2
" 4.2
Typer :5601 12.4
menutecturer sllde/oyerTurn
12 Supac 121f' 25II1II grab
t.ST
(1470 Nl 12.8 short/long Term
slope STo!Iblllty
overburden
'0 pu II out factor
OC sing le rib
1:5 Tensar SR2 W-', L-3 rIbs
t-50".,/1II1n
79 12 15.8
40'
manufacTurer sliding
16 Supac 12'" 25l1li grab
test
11470 IU 12.8 bearing
cap&c:lty
overburden
'0 pu I lout factor
'.7
'"
17 T.n5&111" SR2 glQbeli pullout test
CASE toETHOO OF TEN S ILE ,...... XIMUM STRENGTH @e REDUCTION FACTORS EXTERNAl INTERNAL STA6IL ITY
STRENGTH STRENGTH ELONGATION -------------------"---------- STAB ILITY
EV,lJ.UATION kN/m (~l TikN/m) IS) O~A6ILITY CREEP Po4COEL VERT. STRESS H.STRESS OISTR PULLOUT
wide width
,., t ie-back
21 Typar 3601 w" 200lm,Lz10Omm
t ,. 10:{
12.4
" wedge
slldlng-I.5 tieback
44 Tenser SR2
" 10 29 '.0 1.0 bearing-I.,
ov erturn lng-I.,
wedge Trapezoidal test
slldlng-1.5 tieback
'.0
46 Tenser SR2
" ro 29 1.0 be arlng-2.0
overturn! ng-2.0
wedge Trllgezoldal K. test
54
Tensat' SRI
SR2
IoItg. based on
IS or less
8.' Incorp 5 1 tdlng beat'lng tieback
design cop. overturn wedge Meyer~f .
creep In 120 yrs 16.1 stren gth stlp failure
256
TABLE 4. INSTRUMENTATION
.
_ _ _ _ _~_ _ _ _ C _ _ _ ~_ _ _ _~______I~:_~~_ _ _ _ J::~_ __
RE I NFORCEMENT STRUCTLRE oF WALL OF WALL (differential J
indirectly
calculated observation
bottom 57 Iml
top 152 "'"
slope indicator
slip tube & settlement
devices
survey survey
reference extensoneters I nc I 1nometers reference I nc II nometers
points points
17 ------------------------------------------------------------ ---------------------------
reslstence
strain gauges observational
& Induct.colls
16 ----------------------------------------------------------------------------------------------------
line of mex
strain 18 to 65 II1II
2' frc. v.-tlcal
res I sT.nee
strain gauges obs.. v .... lonsl
& Induct.colls
19 --------------------------------------------------------------------------------------------------
data "i'thin
inconclusive anticipated
range
257
TOTAL STRESSES
INTERNAL LOAD QiANGES Of LOAD PORE PRESSURE TEMPERATURE COI+IENTS
BASE OF FACE OF STRESSES DISTRIBUTION DISTRIBUTION RESPONSE (I f i nstrOOlentction
WALL WALL (soil) IN REINfORCEMEN WITH TIlE BELOW WALL
!:_ _ _ _ _ _ _ _ _ _
~ffected)
___~_____~______~______~______~_____[E ___________
18 SINCO
load cells load cells load cells strain gauges strain gauges res i stance monitoring
thermom probes program
sti I I in
approached approached R2Ink i ne wedge Rank i ne wedge progress
pred i ct press. pred i ct press. def i ned def i ned
at toe of wall at toe of wall
load cells load cell s load cells strain gauges strain gauges monitoring
program
sti II in
approached approached data data progress
pred i ct press. predict press. inconclusive inconclusive
at toe of wall at toe of wall
258
36 -----------------------------------------------------------------------------------------------------
37 -----------------------------------------------------------------------------------------------------
38 ----------------------------------------------------------------------------------------------------
259
TOTAL STRESSES
INTERNAL LOAD CHANGES OF LOAD PORE PRESSURE TEMPERATURE COMMENTS
BASE Of FACE OF STRESSES DISTRIBUTION DISTRIBUTION RESPONSE (It instrumentation
WALL WALL <5011) IN REINFORCEMEN WITH TIME BELOW WALL affected)
load cell at
tie rod location strain gauges strain gauges wall p~rformed
sat i s factor y
----------------------------------------------------------------------------------------------------------- d ur I n9 cr ane
loading
no conclusion no conclusion
plezerneters
@ panel
load celts geogr 1d stra I n gauges stra I n gauges thermocoup les
locations
260
APPROXIMATE NUMBER OF
WALLS CONSTRUCTED
(+ or -30%): 40 10
REINFORCEMENT - GEOGRID
NOTES:
10 15 30 16
a. Tensar SR1, Tensar SR1, SR2, Tensar SR2 and Tensar SR2
SR2, or SR3 and SR3 SR3 and SR3
b. Tensar SS2
7", 14", or 21" 8", 16", 24" , 8" to 4'0" 12" to 24"
or 32"
1. Reinforcing Materials
2. Facing Systems
3. Backfill Materials and Interaction Mechanisms
4. Design Methodology and Factors of Safety
5. Project Costs
3.1 Reinforcing Materials
Of the projects surveyed, nine utilized woven geotextiles,
twelve utilized non-woven geotextiles and thirty-two ·utilized
geogrids. Other. polymeric systems that were used included
heavy duty nylon reinforcing strips and tire sidewalls with
anchor bars. From the introduction of fabric reinforced
retaining structures in North American in 1974 (Project No.
I), non-woven geotextiles were primarily used until the
introduction of geogrids in 1982 (Project Nos. 11 and 15).
Woven geotextiles were found to have had limited use as wall
reinforcement. Their primary application has been in
temporary wall construction.
The current domination of geogrids is probably due to several
factors. Firstly, the Tensar Corporation, the principal
geogrid manufacturer, has provided a strong technical
approach in its promotional efforts, an approach which had
previously been absent from the geotextile manufacturers.
Secondly, the apparent high strengths of geogrids may have
appeared attractive to designers, because they could reduce
the number of layers of required reinforcement.
since facing materials were not required. The New York Route
22 project [No.5) was constructed at a cost of $450/m 2 • The
high cost resulted from the use of expensive crushed stone
backfill ($24/cy). Chassie (1984) noted that the high costs
of Projects No. 4 and 6 may have resulted from the walls
being changed ordered into the project.
1 129
2 127
4 320
5 450
6 312
7 203
8 270
9 134
13 255
15 92
23 46
24 39
27 129
29 96
30 160
32 65
33 109
Q)
~
~ ~60~-------r------~--------~------~
~
~ I reinforced concrete . /
Q) I retaining wall •
S 460 t------------:r-----~-___i
~
C'Q
/ /'
;::j / ~ reinforced concrete
0" / ,
If.l 360 ~_______~_....,..,01:..---,......_ _
cr_ib_w_a_ll_ _ _--1
~ /~./'
~
0..
~
........
~""",.,.
~ • ",.,. metal crib wall
... ....
260 '"""'-.........r::;....._.....r:::;...._ _ _ _ _ _ _ _~,...._I.....,:..-.---~
""0
o ••
• •metallic reinforced soil wall
.....=
..... 160 t----~'ZfIIII""C-----------___I
Cd
~
C+-t
o 60~------~------~------~------~
~
If.l
oCo) 3 6 9 12 I~
height of wall in meters
theoretical Ka line
O"h = Ka = tan 2 (45 - </J/ 2 )
7.0 O"V
,--.., /
cd </J = 36°
~
~
'--"
6.0 I
......
...... I
I
CI)
Co)
~ ~.O I
;:j
rn I
rn
CI) I
~ 4.0 I
I
S
o I
~
"'0 3.0 I
CI)
I
~
;:j I
~ 2.0 I
S Legend
...c= D--O Cell 81H a.3m above bottom
b 1.0 0--0 Cell 85H 1.5m above bottom
tr---6 Cell 89H 2.7m above bottom
20 30 40 ~o 60
O"v calculated from overburden (kPa) ,
For the Glenwood Canyon Walls [9], very small strains were
measured in high elongation needle punched non-woven fabrics,
even though the design methodology would indicated factors of
safety near 1. The results were attributed to the apparent
improvement in stress-strain properties of the geotextiles
under the confined soil condition. This project emphasizes
the need to use confined stress-strain values in predicting
performance of structures constructed with non-woven
geotextiles.
Tolerances
soil weight ± 6%
field measurement ± 20%
\
~
~ field measurements
~.
~,
\<" r trapezoidal distribution
\ .'(
.
,"
\
f.
~
..!c:I trapezoidal distribution
"--'
field measurements
> .'
b
Q) ..,...,' overburden stress bh)
'rn""
:::l
rn
~eyerhOf
Q)
- -[
'""
0..
....... 80 Tolerances type distribution
....
cd
u
-+-"
soil weight ± 6%
field measurement ± 20%
'""
Q)
:>
FIGURE 4 Comparison of Measured Vertical Soil
Pressures to Predicted Values - Tucson
Wall (Case No. 18)
275
II.Y---
\ .
\, . trapezoidal distribution
" .
\
\
IRankilne
atera pressure
~ Meyerhof type distribution
V fi,ld m'M""m~'"
~ / trapezoidal distribution
/
.......
cd
.....u
---r.-~--
/
--
' - Meyerhof type distribution
~
~
Q)
./. Tolerances
> 200 soil weight ± 6%
field measurement ± 20%
FIGURE 6 Comparison of Measured Vertical Soil
Pressure to Predicted Values - Lithonia
Wall (Case No. 19)
276
j.,.o", ...j
I
, .' .'
..
I
T I
1.0",
...L I,~"mrmd Lin, of
t/ Mr.tmurn Tension
I
I •
III III, ,,
~ Ellt.MDm.f.r
(a). Section 1
--
-+- Clamp L()fld Ce 1/
• A".rom,f,r
Tl/erm()C(Jupl,
Obs.rv.d Loc~on of
Mf11timvm Strain
~Om~
,
I
I
I
T 1.1
I.Om
,'1
...l , -ESlimalld LIn, of
I MaX/lflii'" Tension
;'
(b). Section 2
0.8
.... 0.7
~
~ 0.6
II)
...
~
o.e
I.-
0.4
~
~
~
0.3
~ 0.2
~
~ 0.1
0
1.0 I.e 2.0 2.e 3.0. 3.e 4.0 4!5
DISTANCE BEHINO WALL FACE (m)
..... 004
!
~
~ 003
~
"t ·EIII,nSDm.'",
~ Results
~
_-----0
t:) 002
-.J
~
~
~ 0.01
<t
C
:t:
\
1.0 1.5 2.0 2.5 30 3.5 4.0 4!5
DISTANCE BEHIND WALL FACE (m)
10000 10000
~
9000 9000
E 8000 8000
'----
.•" El'dolCO'lCtn.l:._
....... ftlr2'~
7000
E . _ ... 1ter~moncIIs 7000
:::L 6000 "'AftM_~
~""-An. 21 .,.. (~ ~) 6000
c 5000 5000
LEVEL 8 .~
4000 4000
U;
0 3000 3000
,-
u 2000 2000
~ 1000 1000
o~-.--.--.--.--.~,--,~,
0.3 0.6 0.9 1.2 1.S 1.8 2.1 2 "4
10000 - - ElWSoIco.lructioo'l
"·An.r2t~
9000 ·...·Aftr'..-.
E 8000
- . Alt.. aa-.
"." An.- 21 <iQy.I.
Io4dinq
(moqne~( .. a~)
'----
E 7000
.......,
:::L 6000
c 5000
LEVEL 6 "0 .------. '.
b 4000
Ul
3000 '\....,\
0
'-
u 2000
~ 1000
0
2" 4
d.
10000 10000
~
9000 -£ftIIII«~_
_. MtM21
~OOO
.... End ot ccn.tn.octiGn
E 8000 .... Nt. !I mon'h 8000 ....... An.. 21 4ay.I
'----
_ After 5 month,
..... Nt¥~'ClGall'l<J
7000
E 7000 .... £nd ct con:rtruc:.tictI (~Iic
..... NW 21 dG')'ll (~tic: q09"~)
9C'9G)
•
.......,
::l. 6000 6000
c 5000 50PO
LEVEL 3 "0
'- 4000 4000
U; 3000
0 3000
'-
u 2000 2000
~ 1000 1000
o. O~~--.-~--'--'---'-'r-
0 0.3 0.6 0.9 1.2 1.5 1.8 2.1 2.4 o 0.3 0.6 0.9 t .2 1.5 1.8 2.1 2.4
Position from f(lee (m) Position from face (m)
REFERENCES
Ph. DELMAS
Laboratoire Central des Ponts et Chaussees, France
J.C. BLIVET
Laboratoire Regional des Ponts et Chaussees de Rouen, France
Y. MATI CHARD
Laboratoire Regional des Ponts et Chaussees de Nancy, France
1 - INTRODUCTION
P. M. Jarrett and A. McGown (eds.), The Application of Polymeric Reinforcement in Soil Retaining Structures, 285-311.
© 1988 by Kluwer Academic Publishers.
286
"Terre annee n
Geotextile
(b = 40 rrUll, spacing 1,20 m)
Maximum tensile
57,8 kN/m 50 ~ 300 kN/m
strengh
~llowable tensile
atrengh. 26,9 kN/m 5 /"'- 60 kN/m
(permanent use)
Seven month after the facing was uncovered, the structure was
loaded with an embankment 4 m high having a batter of 3 in 2.
- st:r-ain gauge
.'- ......... _,"'".,- settlement meter
......
~
:
~,
,
clayey ........ I..·~"·"~r r!J level bench mark
gravel .......' ..... "'" -iilr''''''''''''-"'t~ "-
old fill
peat
/-7- Highway
T 3.'0 -m.. /
lV
3 /
/
Ct
,/ 0 . 1
,/ /
././ """:•.:::- 0 _ _ 0 , / /
---'
0.70 ......
\~~~\~~'~'i~~~;;;';;-/-"-:;"':';'... . . . . ,. . ." 0
• strain h 1.4 m
€ ~ . .".. 2 ~ 4.0 m
. o h
!t
0.70"""
,
t
\/
~--;-;::~ 3.5'0 __
4.0(/
I
a - a
-€ -€
-a (kN!m)
-€ (%) 0
(%)
0
-
0 0 cr0 €
0
(%)
0(
(leN/",)
0(,= 12.0 IeN/",
-100
£r = 1t r.
o
o s 40
Facing'
Point of measurement 1 2 3 4
Height of embankment 1.4 4.0 1.4 4.0 1.4 400 1.4 4.0
(m)
Geotextile 1 0 0 0 0.5
1.0 o 5 0 0.5
Geotextile 2 0 0 0 005
2.0 0,5 0.5 0,5
The structure was built using one form per layer, supported
by the existing structure. Given the permanent character of
this structure and the state of knowledge when it was built,
the fill material chosen was a crushed limestone aggregate
having a continuous 0/3l.5-mm grading, placed at 95 % of the
normal Proctor optimum, or w = 5 % ; '( d = 20 kN/m 3 • The shear
strength was estimated to be : tf· = 45° ; c' = 0 kPa.
et - et £
-et -£ 0 0
- £
o (kN/rnl
0
(tl
-et (tl
£
(tl
0 0
t pp 77 ,3 18,3 0 - 10,0
z.
'3 . 00 __
~I
a strain after construction of layer
• strain at end of c~~struction
Point of measurement 1 2 3 4 5 6
Geotextile 2
Geotextile 4
Geotextile 6
.
0.69 0.68 0.87 0.75 0.77
Ei/Ef
,
E. strain of geotextile at the ehd of layer construction
~
initial profile
of cutting
3 .......
-t-t-
'~.
......,u.
3.'1'1'>
-- -- --I -- I
4.S'nl.. .
I
L
.. strain gauge
the geotextiles and the facing grid, together with the low
coefficient of friction of the two products, results in a
relative slippage of about 2 cm. The local decompaction at the
facing leads to local settlements, together with a small local
increase in the tensile forces near the end of the reinforcing
sheet and a slight rotation of the upper retaining structure.
Tensile caracteristics
(NF 38014)
Ci (kN!ro) S- ('1;)
Geotextile
t PP 72.1 13.0
Geogrid
9.1 8.0
G PP PET
,
305
6e~ o.15~
geogrid
1-
'f
.A (em..)
geotextile
3.5"..."...
settlement
strain
Cl~
0.6
0.2-
o
0.6
0.2.
.-.. (10-86)
0-'- Q t PES
1 50
(k Pd.)
2.0
----t--'O
• Jo
I
I
:::0-1
We may conclude from these two examples that, while the use
of soils having a large percentage of fines poses no special
problems, it is necessary to make sure that the placement and
compaction are not likely to engender pore pressures, or that
the geotextile can dissipate them so as to ensure optimum local
soil-geotextile friction.
5 - CONCLUSION
REFERENCES
1. INTRODUCTION
- Aesthetics,
- Maintenance.
313
P. M. Jarrett and A. McGown (eds.), The Application of Polymeric Reinforcement in Soil Retaining Structures, 313-337.
© 1988 by Kluwer Academic Publishers.
314
Because of the high costs of large scale tests, only a few fully instru-
mented and evaluated cases have been reported. For none of the published
experiments limit states could be reached (cf.(13),(15)). The design,
however, is mainly based on limit analysis.
From the foregoing it is clear that there is a need for additional theore-
tical and experimental investigations of the behaviour of geotextile rein-
forced retaining walls.
The aim of this paper is the discussion of available limit analysis design
concepts and their experimental verification. In section 2 for several
reasons mainly the design concept proposed by Gudehus and Schwing (8) will
be presented. The reasons are: Concept is based on model test results, the
most important aspects of stability are incorporated, partial factors of
safety have been calculated with a probabilistic method by these authors.
Section 3 deals with the experimental verification of the presented design
concept, whereas a procedure for test evaluation is discussed, followed by
the application of this procedure to the evaluation of published data from
large scale tests. Finally, in section 3 the proposals for future theore-
tical and experimental investigations are summarized. Section 4 deals with
case studies.
Design steps are discussed with respect to the standard reinforced wall
shown in figure 2.1: cohesionless soil, no surcharge loads. This section
contains the theoretical background needed for evaluation of field data.
failure along distinct slip surfaces, whereas all the other failure modes
(for example zone failure) are called internal.
a) b)
c) d)
English codes assume the Coulomb wedge from figure 2.2b. This kind of
failure can be expected for embankments with local surcharge loads (10).
To the best of our knowledge, none of the discussed failure mechanism has
been verified yet by experiments for geosynthetic reinforced retaining
walls. For reasons of safety (in the sense of the so-called collapse
theorems of the theory of plasticity), however, for kinematic solutions,
the actual failure mechanism should be well approximated by the assumed
one.
Gudehus und Schwing (8) published a paper with model test results for the
failure mechanism. In these tests the stability of standard walls
(cf. figure 2.1) was investigated as a plane strain problem giving regard
to the geometrical and mechanical rules of similarity. The height of these
walls was increased until the collapse of the structure occured. The
observed failure mechanism (cf. figure 2.3) is different from the usually
assumed ones shown in figure 2.2. For walls with no surcharge loads,
cohesionless soil and sufficient internal stability, therefore, a failure
mechanism of 2 rigidbodies under pure translation, can be assumed. More
316
h ~I----:
.1
FIGURE 2.3. Observed failure mechanism (Gudehus and Schwing, (8)).
From model test observations and numerical studies the following knowledge
has been obtained for the failure mechanism shown in figure 2.3 (cf.(8)):
- The slip surface 24 coincides with the ends of the reinforcement. This
observation simplifies the formulation and evaluation of the limit state
equation.
- For the angle ~? (cf. figure 2.3) the Coulomb solution can be assumed
(1J1 = 45° + tp/ti: ; f: friction angle) . This result was derived by nume-
ricaf studies and was experimentally verified.
- The failure mechanism from figure 2.3 contains the following failure
modes as particular cases: overturning, sliding and bearing (vertical
translation). For example, it can be shown with the help of the limit
state equation, that for small Ah.lh. ~ '11.;'_0' • holds, i.e. in
this case sliding is the relevant failure mode for the structure.
- Because of local base failure, the vertical stresses at the base of the
wall are essentially smaller than the overburden-pressure (cf.(8), (3)).
As a consequence of this experimental finding, Gudehus and Schwing (8)
neglect in stability calculations the pull-out resistance of the rein-
forcement near the toe of the wall (up to 10% wall-height). In all other
layers the pull-out resistance is:
(2.1)
where
With the help of the abovementioned findings, the limit state equation can
be easily formulated and evaluated. In this equation the slip-surface in-
clination r~ (cf. figure 2.3) occurs as the only unknown quantity. This
inclination can be derived with the assumption, that the sum of the ten-
sile forces should be maximum along the slip surface ~. This assumption
has been experimentally verified (8). With the help of the abovementioned
limit state equation, a design chart was made available by (8). This
design chart is further discussed in section 2.3.
For design purposes the earth pressure and the maximum tensile force
(related to internal stability) for the reinforcement must be known.
Field data for the earth pressure derived by (10) imply, that the use of
active earth pressure loads leads to conservative design (cf. figure 2.4).
For the upper part of the wall the measured pressures are only a little
greater than the theoretical active pressures; for the lower part of the
wall, however, the measured pressures are up to 50% lower than the active
r
ones.
II
h
IL ax ax
1
\
\
\
ax
0) b) c)
With increasing distance from the lining, the forces in the reinforcement
increase (cf. (16) and figure 2.5), i.e. the maximum tensile forces are
not equal to the earth pressure forces.
318
5.0
4.0
~
3. °
.""
~
W2
..::0.:...O"-~=-~-::-:.::.l-:"""--;;. ____ _
o 1 2 3 -4
Distance from slopf" surface (m)
Firstly, the design earth pressure will be discussed. Most standards re-
commend a linear earth pressure distribution (cf. (10)). Gudehus and
Schwing (8) assume a linear distribution down to 2/3 of the height of the
wall and afterwards a constant value down to the toe (cf. figure 2.6). The
inclination of the linear part of the distribution should be assumed like
in the active case. The distribution from figure 2.6 was chosen as an
approximation of the field data derived by (10). A more sophisticated
approximation of the field data can be derived with the help of the
so-called silo theory (8). By the assumption of an linear distribution,
the pressures are clearly overestimated (cf. figure 2.4). The proposed
distribution of (8) is more realistic, easy to use and,therefore, will be
further discussed in the context of the evaluation of field data (cf.
chapter 3).
A'V<'«'<
h
fAh
For the design of reinforced retaining walls, the partial factors of safe-
ty for the material properties of soil, the friction coefficient between
soil and geotextile, the material properties of the reinforcement and the
surcharge loads are needed.
For the case of external stability, (8) calculated the partial factors of
safety for the standard wall shown in figure 2.1. They assumed the fric-
tion coefficient and angle of friction, occurring in the limit equilibrium
equation, to be statistically independent, random quantities (cf. (8)), for
the assumed probability distributions) and derived the partial safety fac-
tors with the help of the second-moment reliability code. With respect to
the mean values of friction angle and friction coefficient, the following
results were derived:
For the friction angle (of the soil) the partial safety factors
are in the range of 1.1 - 1.2, for the friction coefficient in
the range of 2-4. These ranges for the safety factors correspond
to the range of 30°-40° for the mean values of friction angle of
the soil.
The above discussion holds for the partial factors of safety corresponding
to external stability., For the case of internal stability and deforma-
tions, to the best of our knowledge, similar investigations do not exist.
be prevented (cf. (17) and (16». This holds for cohesion less fill
materials and high strenght reinforcement, characterized by an initially
steep inclination of the stress-strain curve. The creep problem, there-
fore, is mainly relevant for cohesive fill materials and/or continuous
filement needle punched geotextile as reinforcement, which are characte-
rized by a relatively low initial strength.
2.2 Displacements
Displacements are mostly predicted with the help of the above mentioned
empirical estimates. Yamanouchi et. al. (16) published the results of a
FE-analysis predicting the displacements of a geotextile reinforced re-
taining wall under gravity and surcharge loads. For details of the mo-
delling the reader should refer to Yamanouchi et. al. (16). For this
retaining wall, the measured displacements were underestimated by a factor
of 0.5 by the FE-analysis. Creep was not investigated. Qualitatively, how-
ever, some of the derived results show good agreement with the measure-
ments:
321
The case of external stability will be dicussed first. In model tests the
failure mechanism, shown in figure 2.3, was observed. With the help of the
corresponding (to the failure mechanism) limit equilibrium equation, the
design length 1 of the reinforcement layers and the tensile force Z can be
calculated for the given geometry and soil data of a retaining wall. The
design quantities 1 and Z (maximum of Z according to external or internal
stability) are plotted in figure 2.7 versus the friction angle ~ , with~~/~
as contour lines. The design chart (figure 2.7) holds for friction coeffi-
cients (soil/geotextile):
(2.4)
~ Ah _
50 ~ ::::::----.- 11-
-----------
i, -..:::::: ~ 0,20
0,15
0,05/0,10
I
o 1,0 1,5
25 301 35
Ijl
---
i t-----
I--" _
__ 0,10
0,15
0,20
- -::::::::::::::::---
0,08
::--.- I--
0,12
max Z
lh2
FIGURE 2.7. Design of a stande.rd geotextile reinforced retaining wall
(Gudehus and Schwing, (8»: Design chart (cf. figure 2.8 for situtation)
322
h
t~h 4l,Y
To avoid creep the strains caused by the tensile force should be at least
less than 5%. The appropriate value for this strain limit can be gained
from experiments.
3.1.Principles of evaluation
Most of the input and output quantities of the design concept (for example
tensile forces corresponding to limit states) cannot be measured directly
in full scale tests. In addition, in most large scale tests the limit
states which are assumed for design are not reached (cf. (13), (15)). In
the following, therefore, firstly the possibilities of verification of the
design concept are discussed.
3.1.1.Limit states
First of all, the case of external stability will be discussed. For veri-
fication of the design procedure, the following instrumented data and ob-
servations are needed:
The geometry and loads can be easily determined. As mentioned before, how-
ever, l·imi t states can hardly be reached in large scale tests: Because of
large displacements, the loading equipment and instrumentation fail before
the limit state is reached (15). Large scale tests, therefore, are not
very suitable for the determination of the critical geometry and load.
323
More suitable are model tests, which can be used for the systematic inve-
stigation of the behaviour of reinforced retaining walls in limit states
(cf. (8)).
p=500kN/m
E
...
.n.
1 + - - - - - - 3.25m
FIGURE 3.1. Contour lines for the geotextile strains (Wichter et al.,
(15)). Dashed line: expected slip surface.
324
Large scale tests, therefore, are not well suited to the verification of
the modelling of external stability. Large scale tests should be performed
only in combination with model tests.
The earth pressure distribution, which is needed for the design of the
lining and connectors (connection between lining and reinforcement)', will
be discussed first. With the help of force measurements in the connectors,
the earth pressure distribution can be determined with high accuracy
(cf. (10); (8)). Predicted earth pressure distributions, therefore, can
easily be verified in large scale tests.
One of the open questions concerns the basic statistical data needed for
the application of the second moment reliability code. Thus, very little
is known about the probability distribution of the pull-out resistance.
Also, back-calculations of the reliability of existing, well performing
retaining walls have not yet been published. The question concerning par-
tial safety factors, therefore, is closely related to experimental inve-
stigations. Future work on the subject of geotextile reinforced retaining
walls, therefore, should also concentrate on the following topics:
From large scale tests or existing walls mostly only data for displace-
ments are available. Only few data are available for verification of the
force and pressure quantities resulting from the design concept for the
case of internal and external stability (cf. table 3.1).
Authors measured
displace- tensile soil observat'ed remark
ments forces pressures failure
mechanism
Yamanouchi et al. x x - - -
(16)
Caroll and - x x - -
Richardson (3)
For none of the large scale tests listed in table 3.1, data corresponding
to the limit state are available. For some of these tests limit states
should not be reached, the test reported by (15) had to be terminated for
technical reasons before the limit state was reached. Consequently, the
design concept concerning internal and external stability cannot be veri-
fied yet with the help of measurements from large scale tests. On the
326
other hand, however, some of the available data indicate that the ultimate
surface loads are strongly underestimated (cf. (15); (13». This problem
will be discussed in the next section.
The data from the first 3 tests, listed in table 3.1, correspond to ser-
vice states. These data can be used for verification of the predictions
for the earth pressure and maximum tensile force corresponding to internal
stability (cf. section 2.1.2).
3.2.1 Limit state. Several results from large scale tests indicate that
the ultimate surface loads are considerably larger than the predicted ones
(15); (13). The discussion of this contradiction between experiment and
prediction will be based on the experimental and theoretical findings of
Wichter et. al.(15).
Figure 3.2 shows the structural details, the position of the surface load,
and the expected failure surface (cf. chapter 3.1.1) for the retaining
wall. T~e test was §arried out with cohesive fill material: = 21.5°, c = ¢
40 kN/m ,t=20 kN/m . A Stabi1enka 200-type reinforcement was used.
5, 014.37.0.36 P IkN/ml
S:1 ,42.86 .0.46 P
53 ' 85.06 • 0.59 P
R
54 ' 141.16.0.71 P
max 5 0 263.5 .2.12 P
lcm ='OOkN/m
200 IJXJ
At the end of the test, a surface load of P = 500 kN/m was applied, but
the ultimate state was not reached under this load.
Wichter et. al.(15) used the failure mechanism from figure 3.2 to predict
the ultimate surface load. For the surface load a load distribution was
assum~d. Consequently, the total pUll-out resistance (cf. Sl -S4 in figure
3.2) can be calculated as the sum of the following terms: A constant term
due to gravity loads and a term which depends on the surface load. The
formulas for the pull-out resistances are given in figure 3.2. Further-
more, in figure 3.2 G stands for the self-weight of the failure wedge, C
for the cohesion-force on the failure plane and R for the resulting
327
friction force on the failure plane. For the calculation of the ultimate
surface load, the resulting pull-out resistance (max S = 283.5 + 2.12 P;
cf. figure 3.2) was assumed. Formulation of the equilibrium equations and
solving the equations for the unknown surface load P leads to the follo-
wing result: Only for P< 0, i.e. tensile forces as surface loads, equi-
librium can be reached. Wichter et al. (15) gives the following interpre-
tation for this result: The reinforcement layers (for the assumed pUll-out
resistance) at the limit state cannot be pulled out all at once, i.e.
there must be a so-called mechanism of intrinsic resistance.
On the other hand, each of the tensile forces Sl to S cannot exceed the
tensile resistance of the geotextile (200 kN/m). Base~ on the information
in figure 3.2 the pull-out resistance is greatest at position S4. Wichter
et al.(15) conclude from this that the tensile strength is first reached
in position S4 and then assume S4 = constant = 200 kN/m. With this assump-
tion a value of ultimate surface load of P = 83 kN/m can be calculated;
i.e. it is very much smaller than the value observed experimentally
(P = 500 kN/m). In addition, in the test the limit state was not reached.
Thus there is a big contradiction between predicted and observed results.
max S G
c
The solution of problem 1 leads to the following equation for the tensile
force:
(3.1)
l : shear-band inclination.
Using equ. (3.1) with the input-data: C = 200 kN/m, ~ = 55°, ¢ = 22.5°,
P = 500 kN/m (largest applied load in test), G = 90 kN/m (self-weight of
the failure wedge), SR = 156 kN/m can be calculated for the resultand ten-
sile force. This tenslle force is even smaller than the tensile strength
(200 kN/m) of a single reinforcement layer and very much smaller than the
maximum available pUll-out resistance of max S = 1343 kN/m. For the
applied surface load of 500 kN/m, the limit state, therefore, could not be
expected. With the help of the measured strains (cf. figure 3.1), back-
calculation of the resultand tensile force leads to a value of approxima-
tely 100 kN/m. The measured (100 kN/m) and predicted (150 kN/m) tensile
forces are at least not inconsistent.
The results of Berg (1), therefore, are consistent with the experimental
findings of John (10) (cf. chapter 2.1.2). For verification of the assumed
distribution more experimental data are needed.
These data were gained from a retaining wall of 7 m height with a steep
(1:0.2) face. The vertical distance of the golymer reinforcement layers is
~h = 1 m. Soil data: ¢ = 45°; ~= 14.4 kN/m • The measured stresses and
strains are due to gravity loads.
With the help of equ. (2.2) the following prediction for the maximum ten-
sile force can be made:
2
Z = 0.9. 14.4·7·1·tg (45°-45°/2) = 15.6 kN/m.
The maximum measured strain was 0.3 %. The corresponding tensile force can
be expected to be within the range 2.9 kN/m - 6.9 kN/m (cf. 16). A more
accurate back-calculation of forces is for several reasons (small strains,
inaccurate data for strain rates) not possible. The allowable tensile
force (tensile force divided by a factor of safety) is 31.4 kN/m; i.e. the
measured force is considerabley lower than the allowable tensile force.
With Z = 15.6 kN/m from equ. (2.2), the measured force is overestimated by
a factor of 2.3 - 5.3. Reduction of the friction angle - as is common
practice - by a partial factor of safety of 1.5 would lead to a overesti-
mation by a factor 4.2 - 10.1. At least for cases without surface loads it
can be expected that the predicted maximum forces are conservative.
For another large scale test (cf. (3» very low strain values «0.4 %)
are also reported, whereas these values correspond to common service
conditions. Under large surface loads (P = 500 kN/m) Wichter et. al. (15)
measured a maximum strain of 1.5 %. For the last mentioned test cohesive
fill material was used.
For all of these large scale tests the measured strains were markedly
smaller than the ultimate design strains, assumed to prevent creep. The
prediction of maximum tensile force, therefore, should be improved.
On the other hand, with the more sophisticated methods (for example FEM
with accurate constitutive equations for soils and geotextiles, including
nonlinear effects) needed for better prediction of stresses and strains,
only little practical experience has been made and/or the application is
expensive. Thus the application of more sophisticated methods cannot be
330
expected in the near future. For this reason the common design practice
should be improved with the help of experimental investigations. The aim
of such investigations should be a reduction of the vertical spacing of
the reinforcement layers. To attain this goal, reduction factors for the
calculated maximum tensil forces should be defined with the help of ex-
perimental data.
!
1
1
2I :3I 4I em)
Only little experimental data are available for the earth pressure distri-
bution. For verification of assumed distributions, therefore, additional
data from large scale tests are needed. Force measurements in connectors
are best suited for investigations of the earth pressure distribution.
For the derivation of partial safety factors, more theoretical and experi-
mental work has to be done:
4. CASE STUDIES
For the following case studies, 3 typical retaining walls, all located in
Switzerland, have been selected.
For construction of a parking lot in St. Gallen, the retaining wall from
figure 4.2 was chosen as construction method (12). The greatest height of
the wall is 6 m.
The retaining wall from figure 4.3 was used for construction of a new
railroad track (9). The geotextile reinforced wall was designed similar as
an anchored wall. For more details concerning the fill material and the
reinforcement see figure 4.3. The wall is not instrumented.
333
Colbond P450
Sodoco AS 420
cO.7m
a.)
b)
FIGURE 4.1. Temporary road: a) Cross section, b) Road after construction
WendeplatZ
Itt. I 1'-'.'
;C··..
.t:- ' ..0.. ',~. '~',
• •' ~"" ~
• •''''D ".•.", o · ~ .~./
~ N r. 17 v 387.44 -t -I I _ . . ·
evtl. Dlchtungsbahn und DraJnage I eltung
2-3 ~ / zu Oelabscheider :
---.. .. i' - chemisch. resistent
verschweisste Fugen
hohe Dehnbarkeit und Festigke it
hohe Kerbfestigkeit
Wurzel fest
Nagetier-resistent
n I
arbor, I\'
~
Typ HaTe Gewebe Nr. D 00.5}0
Reisakraft longs 51.0 kN/m
Oberlappungen min. 0.30 m
Bahnen zu 7.65 m Liinge
oberate 2 Bohnen 5 + S + 0.60 :. 10.6m
J (Nr. 16 und 17)
Gesamtausmass ca. 13'000 m 2
SchGttmateriai
wtrdichttfer
I siltiger sandiger Kies
in Legen zu max. 40 em im
Endzustand
VerdichllKlg mit schwerer 5 to Vi bra-
tionswalze auf max. 95% Proctor
5 min 90% Proctor in der Randpartie
im Bereich 1 m und rnehr vorn Rand
ME min. = 600 kg/em'
,
,...-ce. Aushublinie
Lagenweise aufgefUJlt
2.UG und verdichtet
5. SUMMARY
In this paper mainly the knowledge, which can be gained from large scale
tests for the design of geotextile reinforced retaining walls, is dis-
cussed. This discussion is based on published design concepts and instru-
mented data. Proposals for future theoretical and experimental work are
made.
On the basis of the above briefly explained design concept, the evaluation
and interpretation of instrumented data is discussed and applied on pub-
lished data. The following results were derived:
- Ultimate loads, derived with the help of rigid body failure mechanisms,
are consistent with observations from large scale tests. Reported incon-
sistents between predicted and measured loads (cf. (15)) can be shown to
be a result of a static misinterpretation.
- For investigations of limit states model tests are better suited than
large scale tests.
The available instrumented data from large scale trials are insufficient
for verification of the design concept. The following additional data are
needed:
REFERENCES
1. Berg, R.R., Bonaparte, R., Anderson, R.P. and Chouery, V.E.: Design
construction and performance of two geogrid reinforced soil retaining
walls. Proceedings of the Third International Conference on Geotex-
tiles, Vienna, Austria, 1986.
16. Yamanouchi, T., Fukuda, N. and Ikegami, M.: Design and techniques of
steep reinforced embankements without edge supportings. Third Inter-
national Conference on Geotexti1es, Vienna, Austria, 1986.
339
P . M .Jarrett and A. McGown (eds.), The Application ofPolymeric Reinforcement in Soil Retaining Structures, 339-340.
© 1988 by Kluwer Academic Publishers.
340
load and that model testing should be used for the systematic
investigation of the behaviour of reinforced walls in limit state and the
determination of the failure mechanism. It is interesting to note that
although the measured vertical stresses at the base in the vicinity of the
wall were significantly smaller than the overburden pressure, it is
accepted that the average value of the vertical stress at the base is
equal to the overburden pressure.
INTRODUCTION
Reinforced soil wall structures comprise a number of components and
materials; the soil backfill, reinforcements, facing units, joints and
connections, foundation for the facing units and the subsoil on which the
structure is built. All are important to the efficient performance of the
structure and should be properly evaluated by appropriate testing.
However, it is the use of polymeric reinforcements which is the specific
point of interest for this Workshop thus this paper deals only with the
evaluation and testing of these. Further, detailed descripticns of test
methods are not included as they are not considered to be the points at
issue, rather it is the choice of which test methods that should be used
that is critical.
The report is written in two parts: the first deals with the purpose of
materials evaluation -and the second provides a recorrmended approach to
materials testing.
1. Design Concepts
The design of any reinforced soil retaining structure requires the study
of both its external and internal stability.
For the purposes of external stability calculations, the retaining
stucture is generally considered to be a rigid body, therefore only tJ:e
properties of the surrounding soil are genera 11 y required to be known,
apart that is fran the mass (weight) of the structure. There is, hcwever,
a growing awareness that the overall deformability of the structure will
influence such factors as base pressure distributions and lateral earth
pressures and that these may in turn influence ttE external stability of
the structure. In such cases the properties of the soil and reinforcement
forming the stucture may have to be taken into account. This is not
343
P. M. Jarrett and A. McGown (eds.), The Application of Polymeric Reinforcement in Soil Retaining Structures, 343-355.
© 1988 by Kluwer Academic Publishers.
344
current practice and is only mentioned here to point out the possiblity of
future developments. Thus external stability will not be discussed further
in this report.
In contrast, a full knowledge of all the materials properties is
certainly necessary for the calculation of the internal stability of
reinforced soil retaining structures. The specific data required for the
reinforcing of materials depends very much on the analytical approaches
taken. These may be classified as follows:
The German, French, PIARC and RILEM papers essentially sUpJX>rted the
Swiss paper and presented testing methods only for geotextiles in-
isolation. However, both the U.K. and U.S.A. papers mentioned that in sane
circumstances a choice rray have to be made between data fran in-soil and
in-isolation testing. The paper from U.K. highlighted this to the greater
extent.
2. Construction practices
The point here is that the analytical approaches used in the design
process soould be vie\\ed only as the theoretical operation leading to an
overall understanding of the stabi l i ty of tiE structure and to its basic
dimensioning. In contrast, practical engineering requires that what
happens during construction must be taken into account as it may affect the
resulting engineered structure in many ways, among which the material
properties are an important aspect. Therefore materials evaluation must
indicate to what extent reinforcing elements may be modified or damaged by
planned or unexpected events during the building of the structure.
'!hus, material testing must give the engineer information on possible
material modifications due to construction procedures, and the tests
required to do so may be quite different from those envisaged for
calculation needs. Theoreticians may sometimes consider these tests to be
of minor importance and include into the Factor of Safety such
uncertainties on the actual material performance. HO\\ever, experience
shows that such material modifications must be fully assessed and
independently allowed for. Failures in structures have often resulted from
material damage due to poor field workmanship or unexpected building
conditions.
Once in the ground the reinforcements may tiEn suffer mechanical and
chemical degradation. For geotextile reinforcements the most obvious risks
are mechanical degradation due to accidental tensioning or tearing,
puncturing and abrasion. In same cases a possible reduction of performance
may also result from loss of permeability due to clogging where the
reinfor~ement is also expected to have a draining effect. Problems related
to joints and connections may also have to be considered from the
construction point of view. It is also possible that degradation may
result from changes imposed on the polymeric constituents of the
reinforcement due to temperature changes, chemical and biological attacks,
exposure to light, fire or even animal attacks. When buried in most soil
347
4. Swrmary
It may be stated that data on materials may interact in many different
ways within the design process and that material evaluation must give to
the Engineer sufficient information to enable him to:
a) check that all the data available are consistent and, thus, reduce the
probability of an error on material identification and properties.
b) refer to existing projects built with similar materials and, thus,
make use of experience gained in the past.
c) dispose of all data required by the reinforcement dimensioning
method(s) to be used.
d) assess possible material modification, degradation and/or misuse due
to construction conditions, and evaluate potential construction
problems.
1. Identification
enable the Engineer to check that the data available is consistent with
previously published data and to allow reference to past experience.
Identification testing is best when it is rapid and easy to control, while
giving the experienced Engineer a good picture of the product and allowing
an assessment of its overall performance with sufficient accuracy for first
stage design.
see Section 2) 1
Puncture Resistance 1
Tear Resistance 1
Abrasion Resistance 1
______________________________________ 1
geotextiles and grids or nets whose structure is aligned along and across
the line of traction,although sample shape has same effect on elongation at
failure.
The constant rate of strain at which the short term tests are carried out
does vary from Country to Country but it is becaning more standardised and
is most often now carried out at 10 per cent per minute. However, a recent
proposal is that the test should bring the test specirren to failure in 2
minutes ~ 10%. These rates of strain are too high for manual recording of
test data and so automatic data recording is generally required.
The effect of the rate of strain during testing on the measured load-
strain behaviour of geotextiles varies with both their polyrrer content and
their structure. The behaviour of wovens, girds and nets is predominantly
controlled by their polyrrer content and the effect of the rate of strain is
significant whereas needle punched (and to a lesser extent melt-bonded)
products, being structure controlled are relatively less sensitive to rate
of strain, particularly at lOt.' strain levels. For the same reasons the
effect of the temperature of testing is significant for wovens, grids and
nets and less important for non-wovens.
2.1.2. Long Term Loading
Once again there is general agreement that unidirectional tensile tests
should be used to determine the long term tensile behaviour of geotextiles,
however, only a limited amount of long term testing has been carried out.
What testing has been done has involved sustained constant loading at
constant temperature on a number of different test specimens. The same
discussion on sample width/length ratio and in-isolation to in-soil
testing, applies to long-term testing as much as to short term testing and
as before the most commonly adopted width/length ratio is 2, although for
woven geotextiles much narrower test specimens are quite often used. Very
few long term in-soil tests have been undertaken. Nevertheless, such tests
on soil-textile composites as have been performed, show that a very
different creep behaviour of the canposite is obtained canpared to the
reinforcing element in-isolation.
Just as important as the test methodology for these long term tests is
the rrethod of presenting the test data. Very often the data is plotted as
strain versus the log of time and a creep coefficient derived fran the
slope of the linear part of the plotted data. This requires that the
initial non-linear part of the plot is ignored, however, this may represent
the majority of the strains developed in the geotextile at that load level.
Thus a material which exhibits very large strains initially but little
strain thereafter, may appear to be a more eff icient soi 1 reinforcement
than a material which has only limited initial strains and modest time
dependent strains. For this reason the total strain from the time of
loading should be measured and included in data presentation. A test
rrethodology and rrethod of data presentation which accanplishes this has
been developed in U.K. between the University of Strathclyde and TRRL,
354
Murray and McGown (1987). It is this method which has been used to present
the data for use in the predictions of the test walls which form the
background to this Workshop.
Only by carrying out such long term tests can the long term tensile
behaviour be correctly established but this type of testing is very time
consuming and expensive. Thus as discussed in Part A it is not expected
that this will be undertaken for routine purposes. Consequently it is most
important that the data obtained from such testing be correlated very
closely with the short term test data which will be routinely obtained in
order that checks on the quality of materials can be properly established.
Recent work suggests that checks on the short term ultimate strength are
not a sufficient measure of material behaviour and that it is the entire
load-strain curve that must be considered. This should be the subject of
future stud ies.
3. Other Properties
Other properties which may be of critical significance to the long term
operational behaviour of reinforcements may be discussed in two groups as
follows:
3.1. Durability
The long term degradation of geotextiles is, as stated previously, one of
the most obvious risks to their satisfactory long term behaviour.
Unfortunately there is almost no agreement on suitable test methods to
establi's h this behaviour. This does not mean that geotextiles will not
perform satisfactorily or that many of the available, but unsuitable, tests
should be adopted until suitable tests are developed. Rather it means that
there is an urgent requirEment for this aspect of materials evaluation and
that new awropriate tests should be developed. Also monitoring of full
scale structures should be urgently undertaken. As a first step in this
355
4. Surrrnary
The testing requirements for the geotextile materials used as soil
reinforcements have been identified but it has been shewn that there are
many areas where no suitable test methods presently exist. It would be
most useful if in time a list of the testing requirements for geotextile
reinforcements made together with the identity of both the available
suitable test methods and test methods that require to be developed.
REFEREOCES
1. MURRAY, R.T. and McGOWN, A.: The Selection of Testing Preeedures for
the Specification of Geotextiles, Pree. II Int. Conf. on Geotextiles,
Las Vegas, Vol. II, pp. 291-296, 1982.
2. MURRAY,.R.T. and McGCWN, A.: Geotextile Test Procedures: Background
and Sustained Load Testing. Application Guide 5, Transport and Road
Research Laboratories, Department of Transport, U. K. 12 pp. 1987.
3. MURRAY, R.T., McGCWN, A., ANDRAWES, K.Z. and SWAN, D.: Testing Joints
in Geotextiles and Geogrids, Pree. III Int. Conf. on Geotextiles,
Vienna, Vol. III, pp. 731-736, 1986.
4. R.I.L.E.M. Committee SM-47 (Synthetic Membranes): Report on Meeting at
Milano, Italy, pp. 10, 1985.
5. Session 2B: International Standards, Proc. II Int. Conf. on
Geotextiles, Las Vegas, Vol. II, pp. 291-333, 1982.
6. STRUD~, J.: Basic Principles Underlying the Swiss Guidelines for the
Use of Geotextiles, Proc. II Int. Conf. on Geotextiles, Las Vegas,
Vol. II, pp. 301-305, 1982.
REVIEW OF SESSION
Leflaive started the session and stated that there are four purposes for
material evaluation. These are a) to check the consistency of information
in order to avoid mistakes in material selection, b) to enable the
engineers to refer to past experience, c) to provide data for calcula-
tions, and d) to enable the prediction of construction problems.
357
P. M. Jarrett and A. McGown (eds.), The Application of Polymeric Reinforcement in Soil Retaining Structures, 357-358.
© 1988 by Kluwer Academic Publishers.
358
which occurred in the facing elements and connections from the early
compressible aluminum facings to present day large concrete units. These
developments, he urged, should be taken into consideration when
identifying the test requirements.
With regard to soil testing, McGown pointed out the importance of lateral
(tensile) strains in reinforced structures and argued that it is therefore
essential to study the lateral strain behaviour of a range of typical
granular soils. He explained that the 10% criterion for tensile strains
in reinforced systems originated from the fact that at about 10% lateral
strain the angle of internal friction reaches its ~cv value. This
aspect of soil testing, particularly at low stresses, has received little
attention so far.
Pullout test
.: .... ' .
'" " .
... .
Soil Reinforcement
I
Reinforcement
~
Shear test for direct sliding resistance
359
P. M. Jarrett and A. McGown (eds.), The Application of Polymeric Reinforcement in Soil Retaining Structures, 359-361.
© 1988 by Kluwer Academic Publishers.
360
References
R.A. Jewell
University of Oxford
Abstract
The paper considers the equilibrium in reinforced soil walls.
Two solutions are presented which are expected to "bound" the
likely state of equilibrium in a wall. The magnitude and
distribution of the reinforcement forces and the soil stresses
are derived analytically, allowing the movement of the wall
face to be calculated. Two non dimensional charts are given
for the calculation of outward movement. The resulting
magnitude of movement predicted by the two equilibrium
solutions is close, so that the likely range of movement in a
wall is well defined by the analysis. Incremental wall
construction causes a "moving datum" for the outward movement
at the wall face due to reinforcement elongation. Thus,
incremental construction results in additional movement at the
wall face. This is the main distinction with propped walls
which do not suffer from incremental construction movement.
An analysis for the incremental movement is given. The
selection of compatible values of mobilised soil shearing
resistance and reinforcement force in a reinforced soil wall
is discussed, and new ideas put forward for the way the soil
is likely to behave in the reinforced zone. The application
of the concepts and the analyses introduced in the paper are
fully illustrated by the prediction for the Royal Military
College trial wall which is presented in a companion paper.
INTRODUCTION
1.1 Design and prediction
The detailed analysis for reinforced soil is complex - like
most soil mechanics - particularly when the soil is reinforced
by relatively extensible polymeric materials. The derivation
of design,methods is more straight forward; simple conserva-
tive assumptions can be made to allow for uncertain material
properties, boundary conditions, compatibility conditions, and
additional generous safety margins can be applied.
365
p, M, Jarrett and A, McGown (eds.), The Application of Polymeric Reinforcement in Soil Retaining Structures, 365-408.
© 1988 by Kluwer Academic Publishers.
366
C1
but the stress ratio in the soil can also be expressed more
generally in terms of the principal stresses to give
(2)
368
(a)
Th~ effect of reinforcement is twofold:
(4)
Critical surface
has greatest
required force
Similar pattern
} of force to
above
competent foundation
There are two surfaces through the soil that separate the
three zones. Firstly, deep in the soil is a surface beyond
which no additional stresses are required from the
reinforcement to maintain equilibrium. This is a locus of
zero required force. Secondly, through the toe of the slope
there is a surface which requires the greatest total
reinforcement force to maintain equilibrium. This is the most
370
Uniform surcharge
{ CT h } required
o ,-.-----r-.......-
I
I
""----
I Available stress
I with uniform
I spacing
I
I
I
,"
A B c o
a..
A B A C E
~.P ~RO
It,::IZ Q
o o 0
b. c. d..
1
H
t
R
Locus of zero
required force
eqn.(8)
Pr/p
Position on reinforcement
Fig. 7. Reinforcement layout and force distribution for the
ideal length case.
(o) (b)
-Lmax-
1 Additional soil
H
loading
ll.PR - Extra required
force
A S s' e"
''-~
1
H
(PR) 08
,
/. ....
(e) (b)
(P R }c's
(c)
Fig. 10. Equilibrium with truncated reinforcement layers
supporting the full active force and with negligible bond
length
A 8 A 8
o E o E
(a) ( b)
Fig. 11. (a) Required reinforcement length for truncated
layers with negligible bond length. (b) Equivalent stress
field in the reinforced soil.
3; 3~
0
~~\
,..... 50 45 40
I
..........
N
~ 0.2
I-
(f)
I
W
0:::
u 0.4 / I / I \
-'
-' // / / I
~
0.6 J J / J / I
/ II I I
:l:
g
w / I
CD
I 0.8 II // I / V /
/ // V / / I
l-
n.
w
~ ~V /
Cl
/
./. ./
o 0.2 0.4 0.6 0.8
~-
/3
-
SoH
'i--
Facing ~
t
--~3
4Reintorcement
Soil
:;-- Soil PRI - PR2
lrt: R =
6L
(0) ( b)
(e)
COLLAPSE LIMITS
WALL DEFORMATION
Reinforced soil
Wi th extensible polymer reinforcement materials the greatest
deformation is likely to develop in the reinforced soil, zones
A and A' Fig. 14. Wi th reinforcement attached to the facing
panels, the elongation in the reinforcement layers should be a
good measure of the lateral deformation in the reinforced
zone.
Between the most critical plane and the wall face, zone A in
Fig. 14, the principal stresses in the reinforced soil are
approximately vertical and horizontal. This zone is likely to
deform • relatively uniformly causing both horizontal and
vertical movement at the face, Fig. 14b.
Behind the most critical mechanism, zone A' in Fig. 14, the
principal stresses in the reinforced soil are inclined from
the vertical, very markedly towards the base of the wall. The
vertical movement in this zone will be much reduced compared
with the horizontal movement.
(a) (b)
I
~ 0.2t-~---r-r~70~~--yL-.~;'--r-~
E
C>:
u 0.4 t-~---t-.HiT-7I--;1'<f--rt----+---+--l--~
~m 0.6 t-~-H7'lt'-r¥--A'-----+-1----+---+---I--1
0.8 t--UAW7f---+--I--+--+-----+---+--I-~
1 ~
o I 0:4 I I I 0:8 I I I :£"
1 1.6 2
OUTWARD MOVEMENT ( ~~
L 0 Y/ / /
V
o 0.2 0.4 0.6 0.8
0
.01 .02 (~
e I'"
I 1\
'"
0.2
I-
UJ
W
u
c::
0.4 ~
::l
~
\ ~
:;= 0.6 1\ !
9
w
CD
~w O.B
\ / •
+
IDEAL SPACING
UNIFORM SPACING
I..t-V
0
5.1 Introduction
The analysis for reinforced soil wall behaviour has been
examined in stages. Firstly, equilibrium in a reinforced soil
wall has been examined for soil with a fixed and constant
angle of friction. Secondly, the state of equilibrium in the
"constructed" wall has been used to deduce the magnitude and
distribution of the reinforcement forces and hence the
elongation in the reinforcement layers.
388
IT
t It-------+--~
z
Most critical
plane
------£:
--- --- a
Fill
construction
(11 )
_( Qs)tan(90-¢)
L 1NC - HINC+-y 2 (12)
..--...
:::c
"-
~ ~ l\
"0l\
N 0.2
'-"
tilw
0:::
0.4
\ ~\
()
..J
..J
«
.3:
3: 0.6 ~
W
0
..J
III
- \', +
• NO INCREMENTAL
DISPLACEMENT
:::c 0.8
TRANSLATIONAL
DISPLACEMENT
I-
a...
~/ <>
)
W ROTATIONAL
CI DISPLACEMENT
6.1 Introduction
There are almost any number of distributions of mobilised soil
frictional resistance and mobilised reinforcement force which
will provide equilibrium in a reinforced soil wall. For
analysis, the problem has to be simplified by assuming that
the mobilised frictional resistance is constant (at least in
specified zones), and this technique has been used to
investigate reinforcement force distributions which would
provide equilibrium.
The two components providing equilibrium in the reinforced
soil, the soil and the reinforcement, are not independent of
one another, however, but linked through strain compatibility.
The mobilised frictional resistance in the soil depends
importantly on the strain which develops in the soil.
Likewise, the mobilised reinforcement force depends on the
strain in the reinforcement. If the unstressed reinforcement
is placed in soil, then the maximum possible strain in the
reinforcement equals the tensile strain in the soil in the
direction of the reinforcement.
Between the most critical plane and the wall face the
principal stress directions are approximately vertical and
horizontal, as indicated in Figs. 8 and 11. In this region
the reinforcement is aligned closely with the direction of
principal tensile strain in the soil.
6.2 Compatibility curve
The link betweeen the mobilised soil frictional resistance and
the mobilised reinforcement force can be shown on a
compatibility curve, Fig. 20 (Jewell, 1985).
In terms of required and available forces, equilibrium can
occur when the reinforcement mobilises sufficient available
force ~o satisfy the requirements for equilibrium, the
required force in the soil. This is where the two curves
intersect, Fig. 20c.
A smaller unit of reinforced soil could also be examined with
a compatibility curve in terms of the local required and
available stresses, rather than overall forces.
The compatibility curve relating overall required and
available forces in Fig. 20c is a direct illustration for
propped wall construction. The starting point of zero tensile
strain and unstressed reinforcement on the left axis is where
the fill and the reinforcement have been placed, but the prop
is pr0viding all the required force. The required force is
transmitted and distributed to the soil through the stiff
facing.
393
:-1obilised A.xia 1
frictional force
resistance
_ :l.esponse to long
term loading
10 (7.) 10 (%)
Tensile strain Tensile strain
a. Soil characteristics b. Reinfo.rcement characteristics
(Isochronous curves)
10 (%)
Tensile strain
c. Compatibility curve
Stress paths
Another difference between propped and incremental wall
construction is illustrated in Fig. 21. As indicated above,
for a propped wall the soil will experience increasing
vertical stress during construction and the horizontal stress
is provided by the stiff, propped facing. Lateral strains
will be small if the propping is stiff, and the stress path RP
will lie close to the K, line, Fig. 21a. On removal of the
prop, the horizontal stress in the soil will reduce as the
reinforced soil moves into equilibrium straining laterally to
a higher st~ess ratio, PO Fig. 21a.
394
(b)
Fig. 21. Stress paths for reinforced soil for propped and
incremental wall construction.
0-8
tis
0-7
0-6
0-5
a
0-10 0-20 L:y
t 0,8
tIs
0·7
0·6 • Critical
state
0-5
and
399
(15)
where dEl and dE3 are the incremental major and minor principal
strains.
tIs tIs
I
I
: sin +p
1/ 2(G/s)
£3
(£1+£3) £1
(a) (b)
( ) LJ(t/s)
E3 = 2(G/s)
(17 )
y = o(x)/ H ( 18)
Ll(t/s)=sin¢p (19)
sin¢ p
G/ s = -----'-- (20)
(o(x)/ H)
(sin¢p-sin¢cv)
G/s=~---"--------""-'- (21)
(o(x)'/H)
The use of this soil model and the equations presented above
is illustrated in the companion paper for the prediction of
strain compatibility in the RMC trial wall.
401
BOUNDARY CONDITIONS
8.1 Introduction
The boundary conditions for a reinforced soil wall are likely
to affect the forces and displacements in the soil and the
reinforcement. In general the boundary conditions for walls
built on competent foundations are likely to reduce the forces
and displacements in the reinforced soil compared with the
assumed ideal boundary conditions.
In this paper it has been assumed that only the reinforcement
provides additional stabilising forces in the soil, and that
at the wall face the reinforcement tensile force is neatly
transferred into a uniformly varying compressive stress on the
soil surface. This may be thought of as ideal facing~ as the
face adds no kinematic restraint and is perfectly free from
interaction with the foundation.
CONCLUSIONS
The main points and conclusions from this paper are summarised
below.
9.1 Equilibrium states
Two possible states of equilibrium have been presented for
reinforced soil walls, together with the analytical equations
describing them.
The ideal length equilibrium case is where the reinforcement
locally provides the required stress in the soil everywhere
that it is needed. The unreinforced soil behind the
reinforced zone does not load the reinforced zone.
405
Acknowledgements
Aspects of the reported study were completed while the author
was supported by the Royal Society/SERC industrial fellowship
scheme. The author is grateful to the Soil Mechanics Group at
the Department of Engineering Science, University of Oxford.,
for stimulating support, to Guy Houlsby for the many
discussions on the behaviour of sand, and to George Milligan
for thoroughly reviewing the manuscript and making many useful
suggestions.
References
Bathurst, R.J. & Jarrett, P.M. (1986). Class A predi ction
exercise for reinforced earth walls. Bulletin No.1 for NATO
Advanced Research Workshop, Application of Polymeric Rein-
forcement in Soil Retaining Structures, Royal Military
College, Kingston.
Bransby, P.L. & Smith, I.A.A. (1975). Side friction in model
retaining wall experiments. Journal of Geotechnical Engineer-
ing, ASCE GT7, July, 615-632.
Hughes, J.M.O., Wroth, C.P. & Windle, D. (1977). Pressureme-
ter tests in sand. Geotechnique 27, 455-477.
Jewell, R.A. (1985). Limit equilibrium analYSis of rein-
forced soil walls. Proc. 11 Int. Conf. Soil Mech. Fdn Engng,
San Francisco, Vol 3, 1705-1708.
Jewell, R.A. (1987). Analysis and predicted behaviour for
the Royal Military College trial wall . Proc. NATO Advanced
Research ~orkshop, Application of Polymeric Reinforcement in
Soil Retaining Structures, Martinus Nijhoff.
Jewell, R.A., Paine N.P. & Woods R.I. (1984). Design methods
for steep reinforced slopes. Proc. Int. Conf. Polymer Grid
Reinforcement, London, 70-81.
Jewell, R.A. and Wroth, C.P. (1987). Direct shear tests on
reinforced sand. Geotechnique 37, No. I, 53-68.
Lambe, T.W. & Whitman, R.V. (1968). Soil Mechanics. John
Wiley, New York .
Stroud, M.A. (1971). The behaviour of sand at low stress
levels in the simple shear apparatus. PhD Thesis, University
of Cambridge.
Symes, M.J.P.R. (1983) . Rotation of principal stresses in
sand. PhD Thesis, Imperial College, London.
Schofield, A.N. & Wroth, C.P. (1968). Critical State Soil
Mechanics, McGraw Hill, London.
408
1. INTRODUCTION
There are many types of reinforcing materials and systems
available for the construction of reinforced soil walls. Of the many
types, the Reinforced Earth system developed by Vidal (1966) in France
has predominated and has provided the basis for most theoretical and
empirical knowledge of the behavior of reinforced soil walls. The
Reinforced Earth system has a number of distinguishing characteristics
that include:
• steel reinforcing elements that have tensile modul i on the
order of 2 x 10· kPa (3 x 10 7 lbs/in2);
• reinforcing elements that are discrete strips, approximately 50
mm (2 in.) wide and 5 mm (0.2 in.) thick; and
• concrete facing (skin) elements that ~an individually undergo
limited translation and rotation in response to movements in
the reinforced fill or settlements of the foundation soils.
More recently, reinforced soil walls have been constructed with
geosynthetic, reinforcement and various facing elements. The two types
of geosynthetics commonly used in reinforced soil wall construction
are geogrids and geotextiles. Geosynthetic soil reinforcement systems
have distinguishing characteristics that include:
• polymer rei nforci ng el ements that have tensil e modul i on the
order of 1 x 10 6 kPa (1.5 x 10 4 lbs/in2);
• reinforcing elements that are continuous or semi-continuous
sheets with thicknesses in the range of 1 to 5 mm (0.04 to
0.20 in.); and
• a variety of possible facing (skin) elements including concrete
panels, timbers, and geosynthetics.
It is clear that the Reinforced Earth system and geosynthetic
reinforcement systems incorporate fundamentally different types of
reinforcing elements: steel strips versus polymer sheets. While the
nature of the reinforcing elements are different, the most commonly
409
P. M. Jarrett and A. McGown (eds.), The Application of Polymeric Reinforcement in Soil Retaining Structures, 409-457.
© 1988 by Kluwer Academic Publishers.
410
used procedures for analysis and design are similar. These procedures
derive largely from research on Rei nforced Earth and other steel
reinforcement systems. There prevails an inherent assumption that
analysis and design procedures developed for steel reinforcement are
also applicable to geosynthetic reinforcement. This assumption is
questionable.
This paper will examine differences in the behavior of soil walls
reinforced with steel strips or other relatively "inextensible"
materials (such as steel grids) and soil walls reinforced with
relatively "extensible" geosynthetic materials (geotextiles or
geog ri ds) • Thi s compari son will be made for the idea 1 case of a
frictionless wall facing, braced during construction, and free to
rotate about its toe after construction. Wall systems and
construction procedures causing deviations from this "ideal" case will
be discussed. Analysis and design procedures for reinforced soil
walls will be reviewed and conclusions will be drawn on the procedures
most appropriate for use with geosynthetic reinforcement. These
procedures will be used to conduct parametric studies and to predict
the behavior of the instrumented geogrid-reinforced soil walls
constructed at the Royal Military College (RMC) in Canada.
2. COMPARISON OF INEXTENSIBLE AND EXTENSIBLE REINFORCEMENT
There are two basic differences between steel and geosynthetic
reinforcing elements:
• as stated in the introduction, the tensile moduli of steel
reinforcement and geosynthetic reinforcement are vastly
different: the ratio, Es/Eg, is on the order of 100 to 1000
(Es = modulus of elasticity of steel and Eg = low-strain
tensile modulus of polymers used in geosynthetics); and
• as shown by design experience, the volume of steel in a steel-
reinforced soil wall is much smaller than the volume of polymer
in a comparable geosynthetic-reinforced wall (0.02% to 0.05%
for steel reinforcement compared to 0.2% to 0.5% for
geosynthetic reinforcement).
The consequences of these differences are discussed below.
2.1 Definitions
McGown et ale (1978) originally defined inextensible and
extensible reinforcements (inclusions) as follows:
• inextensible reinforcements (inclusions) are those that "have
rupture strains which are less than the maximum tensile strains
in the soil without inclusions, under the same operational
conditions"; and
• "extensible reinforcements (inclusions) are those that have
rupture strains larger than the maximum tensile strains in the
soil without inclusions, under the same operational
conditions".
411
(8)
FACING
r - - - -· --- --
~_ _ __ _ _____ ~SUBSEQUENT FILL LEVELS
I /
1------- ---
H
~------- _ EXISTING FILL LEVEL
o
~ SOIL ELEMENT
AT INITIAL
CONDITION A
;;)7 ............. "71" o:::::;c;:::<: OJ
'
(b)
Qa
INITIAL COMPRESSION COMPRESSION
CONDITION (A) LOADING (AC) UNLOADING (SC)
I~1a1
03
O a3
6a3
03 a3
~3
a, a,
t l!. a,
a3 /{a, +60,) = K.
PROPORTIONAL GENERAL INCREMENTAL
LOADING (AB) LOADING
O
J~,a, O~~,O,
~ ..0'3 ~3 "£03
~ ~ ~ ~
a 1 a,
t60, t6a1
(a3+603)/(C11+6a1)=Ko l!.a3 /l!.C1, 4: CONSTANT
(a)
Ka
GENERAL INCREMENTAL
LOADING STRESS PATH
(b)
C C C
I l ~
A •
o 2
( a). 2.'
I I
v--
2.0
T
II v
L...- v- r-t- !:c:: -I-<
i V -r---- t-- to-
I., 11 v
/
"
~~ I ''''-
"'. . , ""
Preshtar hilult
.., ··
b'" If ~ ....... ,.U ,.~
, ~ ,. ~ 11.' ,~ CO"
.... ,.
c-pr........ • .19 '-51 U%"
..
...
1.0
0 c.."'nsioII ........ 0.51 '.51 U'
,."
.~ ~ ,. ~ I~
I. 6
eo.,ressioll ......
~1IIIIDIIIinf:
I.~
1.11
I."
I.~
I.~
1.U .~
,.»
,.~ ...
411f"
"" ,."
.~ O~5 ,~
.04
.'
, .~ .~ It.S1
10."
bt.lISiDn ...... I... 0.55 I.~ , .~ CI'
)( ~ I .• I ." •. U
00
10 12
(b),
(c).
I LKP
-- --
I
Test 3
f,,-r,:--
1\ • I
L1 4
~t:-= ~
../
3
-40L-...L-..l.2----!-3-~4 -2 2
~("'cm')
1
Ka._
_K,
0
*-
-10 -$ 0 +5 +10
HoriIontM (~), f.h
K 0.5
FILL PROPERTIES (a) .
K
</>'- 35°
Q ~m= 20kN/m 3
0.4
0.3
K
•
0.2 0.1 0.2 0.3 0.4 0.5
0 £.h (%)
1 Z(m)
10
2 ( b).
,.. "20
4
C\I
E
.....
Z
.lI: 30
I-
e.
GIS
s:. 40
b
<!h ~KIS'Z
8
eo
10
eo
10
80
Ko = 1 - sin $ (3)
Ss = Es • N • a • b/A (5)
Typical Values
(a>. a 4 to 6 11m
b 50 to 60 11m
N 4
f
Es 3 X 10 7 kPa
E
-
"! Stiffness Pe r Un ita rea
Ss = Es N a b/A
I (A = panel area)
I. 1.3m
1.6 m
Typical Values
(b) • a 1 to 4 11m
b 1m
( N 2 to 4
500 kN/m
b
.1 Jg
Stiffness Per Unit Area
Sg = J g N b/A
Typical Value
( c).
sv = 0.5 m
5g = J g • N • b/A (6)
000 kPa
e
RANGE FOR STEEL
I
I
I
8.= 70,000 kPa
Z(m) 10 I
(9)
(10)
( 11)
2
Sg = 2,000 kP.
j---- Ko LINE REPRESENTS
'ABSOLUTELY'
I IN EXTENSIBLE
I REINFORCEMENT
I
FILL PROPERTIES
I
9\'=311° I
---I-~ RANGE FOR STEEL
lSm=20kN/m3
8- K LINE REPRESENTS 0
• 'ABSOLUTEL Y'
~
EXTENSIBLE
REINFORCEMENT I
f------/-- S. 70 , 000 kPa
10'
Z(m) 12
eh=t!?,
J-H-+-=."*",, STEEL REINFORCEMENT
2 0.5% S. = 70,000 kPa
1.0%
2.0%----!H
I--- Ko LINE
I
4 I
I
e. h = ~\
8
2.0%
1.0%
I------r-- 0.5%
FILL PROPERTIES
10
",'= 35 0
"lf m =20kN/m 3
z(m)
12
obtained from the numerical simulations of the welded wire mesh wall.
The results from the numerical simulations incorporating geogrid
reinforcement indicate 0h = Kao v , where 0v is the overburden pressure
(ov = yz). This result is consistent with Figures 7 and 8.
2.5 Influence of Construction
The preceding analyses were for single elements of reinforced
soil. These analyses can be extended to the behavior of a
geosynthetic-reinforced wall if it is assumed that the reinforcement
does not affect the stress state in the soil (i.e., the soil stresses
and reinforcement tensions are uncoupled).
For the case of an unreinforced wall that is braced during
construction, but free to rotate about its toe after the braces are
removed, the wall fill undergoes Ko loading followed by compress ·ion
unloading (stress path AB + BC in Figure 2a). If the wall face were
frictionless, an active Rankine state would theoretically develop.
This latter case corresponds to that of a flexible, frictionless
cantilever wall rotating about a fixed toe (Figure lOa). If the wall
contains inextensible reinforcement that makes a good bond with the
soil, the large tensile stiffness of the reinforcement will suppress
horizontal soil strains. In the limiting case of an "absolutely"
inextensible reinforcement, the horizontal direction becomes a zero
extension direction and the failure surface is as indicated in Figure
lOb (Bassett and Last, 1978). For the case of extensible
reinforcement, horizontal soil strains should be only partially
suppressed, as indicated by Figures 7 and 8. It is hypothesized that
the failure surface for this case will be close to that shown in
Figure lOa.
For the case of incremental wall construction the situation is
more complex. At depth, the soil will follow some general incremental
loading stress path (Figure 2a), eventually reaching an active plastic
condition. Near the top of the wall, the soil should be in a
subfailure state, with the soil stress path being closer to one of
proportional loading (AB in Figure 2a) than unloading (BC in figure
2a). The failure mechanism for this case is more complex than given
by Figure lOa. Suggestions have been put forth that the fail ure
mechani sm iss imil ar to one for a reta in i ng wa 11 deformi ng about a
hinged crest. There is some experimental support for this suggestion
for inextensible reinforcement systems (Juran and Schlosser, 1978) but
not for extensible reinforcement systems. Additional research is
needed in this area.
2.6 Conclusions for Geosynthetic Reinforcement
It is concluded that currently available geosynthetic materials
are extensible forms of reinforcement. This conclusion is based on an
analysis of the amount of horizontal extension required to induce an
active plastic state in an element of unreinforced soil. The amount
of extension is s.tress path dependent and, for a compacted granular
fill, ranges from less than 0.5% for compression unloading up to 1% to
2% for general incremental loading. The amount of mobilized
geosynthetic reinforcement tension at these low extensions is small.
Theoretically, local yielding of the soil element will occur before
426
( b).
LATERAL EARTH PRESSURE (p")
(aL 500 1000 1500
6 Predicted by FEU
o Meosurt:d
~ g~fJ';~d \~r:e f~~ FEM
A. _ I(Q.Ov"bu'd .... P~e"ure
"b' A D a_ Ko.Bea.inq Preuu.e
GRAVELLY SAND C - I<O·(}v.,rburdcn Prenu'.
to.
o ", 60 - KO·0.,0";"9 P.usun
20'
¢>'~ 40.6°
g,IO
" ~
'.A,S C I\,
FOUNDATION 2 I 20
Su=1100IbO/~! l,lS'OR14'.
lSm=107 Iba/f!
( c).
14 Ft High Wall 20 Ft High Wall
DEFLECTION (in) DEFLECTION (in)
:3 2 1 '0
Jr--,r---'l2:'-'---r1--r-'/O 0 -::; 0
a VSL Bar tlot
o VSL Bar tAat " Welded Wire
• Weldep Wife o Geogrid 2-::;
o geognd o c 2..J o o •
o c. ..J
..: 0 4 -1
o O· 4 3= 0 -1
o lJ.. 0
6 ~
o 6 0
0...
o 0
8 \:i
o 8 I-
0 10~
o 2 0 f-
o 1~
lJ..
0 122
0 o • 0 0
o 0 • 12~ 14B:
0... o •
00' W 0 o • 16~
"'" 14° w
0...
0 o. 180
o o.
20
(b) .
TENSILE
ARC /
X
/
C>( / I--~
PRINCIPALA/!>
TENSIl.E
STRAIN
POTENTIAL .-::==----
FAILURE SURFACES
( a). (b).
RESISTA NT
ZONE
REINFOR CEMENT
L
-I
(c) . (d) .
Figure 11. Common shapes for potent ial failure surfac es for limit
equilib rium stabil ity analys is: (a) straig ht wedge;. (b)
two-part wedge; (c) circle ; and (d) logarithmic spiral
431
(a).
1201 1201 r - - - - . - - - , - - - ,
~ ~
0 o
S S
I----t---t---::r:>'l Rankine ~ 100 to f--..- r..-::..-:...-.-i.:...--r-::::-o--1 Ro nkin e
>-
a o ~F::;:;';:':~
....."
..c:
801
"
..c:
..... 80l:.
"c ~
'c"
c
~
c
0 601 o 60:1;
:'5.
u.J
- - Rank ine '"
UJ
. -.. Coulomb
0 401 ~
«
.....
- - - - Sokolovski
.....
40%
zu.J -·-··lanbu
.-.•.•.. Brinch Honsen
Z
UJ
U
201 _ .. _. log Spira I ~ 201
'"w
"-
UJ
"-
OX
10' 20' 30' 40' 011':-0.--:-::'20:::-.--:3:'::07".--:40'
if> if>
(b). s •.f..
120% r - - - - . - - - , - - - , 120% r-_--._2--,r-_-,
g ~
;
1001. Rankine o 100% Rankine
"~ ~~-:'-~:
~ 80'l. r-:-::~~i'~""'"
~. ·-·-t~
g
..c:
80Y.
.....
~" 60% 1" WI.
c
o __ Rankine o
'" ·-·Coulomb '"
u.J 401', u.J 40%
- - Sokolovski
o 0
«
.....
- - - Jnnbu «
.....
~ 20%
------ Brinch Honsen
zu.J 20%
- " - log Spiral
u U
"""-
u.J '"
UJ
"-
0"'1'':-0':---:20:::'--==30::-'--:40' 01'.
10' 20' 30' 40'
if>
DEFINITIONS:
W= Soil weight
N = Normal force on potential failure surface
S = Shear resistance on potential failure surface
P = Interwedge force
T = Horizontal reinforcement force
Q = Surcharge loads
8 = Angle of potential failure surfaces
~ = Angle of surcharge loads
~ = Angle of tnterwedge force
Figure 13. Two-part wedge model used for reinforced soil wall
stability analysis.
435
Segrestin (1979) point out that the use of a wedge analysis assumes
that all reinforcing elements crossing the failure surface yield
simultaneously. They note that calculated factors of safety based on
Coulomb wedge calculations are larger than those based on local
equil ibrium of individual reinforcing elements. As a result, some
widely used design procedures (e.g., British Dept. of Transport
(1978» incorporate both local and global stability calculations.
Stri ct adherence to 1oca 1 stabi 1 ity requi rements may be overly
restrictive for geosynthetics because: (i) they are much more ductile
than steel; and (ii) they occupy a substantially larger volume in the
reinforced zone (they are in contact with many more soil particles)
than steel. Geosynthetics pick up increasing tensile loads with
increasing tensile strains up to and past the point at which the soil
is yielding. It is reasoned that if local overstressing occurs durjng
the loading process, the soil and reinforcement will deform locally,
increasing the tensile force carried by the reinforcement. Soil shear
stresses generated as a result of differential soil strains should
result in load transfer between reinforcement layers. The relatively
large volume of geosynthetic reinforcement should enhance load
transfer.
Strict adherence to local stability requirements may not be
necessary with geosynthetic reinforcement. However, local stability
can be easily achieved by doing a thorough analysis that assesses the
required reinforcing forces on a wide range of potential failure
surfaces, and then ensuri ng that the rei nforcement di stri buti on
provides the required forces on all surfaces. The procedure described
below for determining the reinforcement force distribution satisfies
local stability requirements.
H 0.5
POTENTIAL
8LIP SURFACE
\ ~_IZ;"b
~-(Z/H).
+ K. ~Z/H)b -(Z/H).]
SLIP SURFACE
T
Z(m) /
/
/
/ ,/
--
! , / !ifILL PROPERT lEa
Til
i .',. 3So
/ ./ 'trw 2 2,0 kN/ra
----
I
./
-----
0 .0
I
' / T lI(kP.'
I
. / ~:o
~
0.0 / 30 .. 0
50 eo
./
~~
10.0
/ //
plots such as those shown in Figure 16. These plots show all the same
features as Figure 15. In addition, they allow rapid evaluation of
Tij anywhere in the fill. The effect on Tij of soil strength ($) and
un1form vertical surcharges is shown in F1gure 16.
Figures 15 and 16 show ~that to satisfy local equilibrium
requirements, tensile forces are required to a distance back from the
wall crest equal to 1.0 to 1.5 times the wall height (X/H = 1.0 to
1.5). In practice, reinforcement lengths are typically on the order
of L/H = 0.8. The wedge of unreinforced soil between the reinforced
zone and zero force line exerts a thrust and overturning moment on the
reinforced soil mass. Calculations show that for L/H = 0.8, $ > 35°,
and q/yH < 0.2, the overtu rni ng moment due to the th rust of the
retained fill is insignificant, as is the calculated vertical stress
increase within the reinforced soil mass due to the overturning
moment. Therefore, calculation of vertical stresses using the
"trapezoidal" or "Meyerhof" distributions is unnecessary. Further,
Jewell (1985b) noted that reinforcement truncation short of the zero
force line causes the locus of maximum reinforcement tensions to move
toward the wall face. While this effect may be significant for steel
strip reinforcement, the effect was found to be insignificant for
geosynthetic reinforcement (due to the short required bond lengths for
geosynthetics compared to steel strips).
4.3 Reinforcement Defonmations
Reinforcement tensile strains, €ij' were calculated from the local
reinforcement tensions as follows:
(14)
where Tij (kN/m) was obtained from Equation 13 and 5g was calculated
as shown in Figure 4. In Equation 14, Tij and 5g are independent
variables (i.e, the reinforcement tensile st1ffness does not influence
the state of stress). The total reinforcement elongation at any
elevation o~(z) was obtained from:
(15)
0
~
g %
~
" ~
"
w
z
~
w
"I
:
-<
o
:
_If
0
II
O .. I.00.,!...
r"
!:
0
~ "a:
~
~
~
0
~
"
0
o
~ ~-------,-------,--------r-------,
" t----T---j--
o
"Io~ • "
:.11
III
o
N 0
-O-c..g
%
)co
13
o
~
"t---r~r-Ti-r~--t-ir
o
..
~ o ~
"o
,0 ~ ~
Fill PROPERTIES
.p ' : 35 0
3
~m = 20kN/m
SgCkPal
100 1,000
FILL PROPERTIES
fI'= 35° 2,000
75 ~m = 20kN/m 3
3,000
50
4,000
8,000
211
2,000
6r • 150mm
4,000
surface areas, and their generally good bond with granular soils,
preclude concern over the length required to prevent pullout for all
but the uppermost layer(s) of reinforcement. With geosynthetics,
there is more concern over reinforcement rupture than pullout.
5.2 5011 and Reinforcement Properties for Design
The · design process necessitates selection of soil and
reinforcement properties for design and the selection of appropriate
factors of safety.
For des i gn, soil strengths and geosynthet i c tens ions shou 1d be
selected at compatible values of strain. Since geosynthetics are
extensible and ductile, it appears both logical and conservative to
444
(a) .
UNREINFORCED
SAND BACKFILL
-7 SR2 GEOGRID
REINFORCED sOIy
SURCHARGEA
PANELS
---jC--------ti --- - II \ -
I-- -SR2 LAY~ 4 ' ::::::::270:
::::::
Te
o
SR2 LAYER 3
SR2 LAYER 2
::::: :: :::: :: y: (
2.4 ...
SAND/AIR PRESSURE BAG SURCHARGE
SURCHARGE
"1\
I\
PANELS
(b).
- j r--
SR2
SR2
T....
U)
"",
~:
SR2 eo
1
oz
SR2
I
• STRAIN GAUGE
_ PRESSURE CELL
weight of the fill was about 18.1 kN/m 3 • The sand shear strength was
measured in 60 mm square di rect shear tests and in conso1 idated-
undrained triaxial tests. Direct shear tests carried out on fill
specimens with an initial dry density of Po = 1.78 Mg/m 3 gave a peak
friction angle ($p)ds = 43 0 • Direct shear tests also gave a value for
the large-strain, constant-volume friction angle of ($cv)ds = 40 0 •
Using the equation by Rowe (1969a) relating plane strain and direct
shear friction angles:
r------~----_.-------r--·----.I------,-----_.I X(m)
o 2 3
o
-- --- ------ 8
GEOGRID
LAYER
-----
____________
- _
:~ «r (kN/m)
----------- -
4-- r-------------------------------~~--------
1.0
-----
3--r------------~~~-------
-----=~
--- -
FILL PROPERTIES
¢"~45°
lI'm= 18.1kN/m 3
:~
o
2.0 -----
2-r--------~~~~--S-U-R-C-H-A-R~G~E~(k~p~a~)-
----- 0
-----12
--- - -50
1- r------"o::...lo..--------------
Z(m)
3.0
q (kPa) q (kPa)
Geogrid
Layer I 12
I 30 I 50
I 12
I 30
I 50
4 I 3.0 I 5.7 I 8.5I 2.3 I 4.5 I 6.7
3 I 4.3 I 6.6 I 9.1 I 3.4 I 5.1 I 7.1
2 I 6.1 I 8.4 I 10.9 I 4.7 I 6.5 I 8.5
1 I 6.5 I 8.5 I 10.5 I 5.0 I 6.6 I 8.2
I $ = 45° I $ = 50°
S, (mm)
30 20 10 o
GEOGRID
LAYER
q(kPa) 50 50 30. 12 0 o 30 50 4
'\ ~12 Y Y i 60
\ \ \
\ \ \ r / \ \
'\\ \ j fLO \\
\ \ \ \
\ \.\ 3
\ \ I
, \ \ \ I .\ \ \ \ \
..
\ \\ \ ~ t
,~\ \ \1
2.0
\\\ \)2
\ I I I
\ I
LOAD DURATION (h,) '\ \
\ I I I '
<>- -
__ - __
- -0 100
1000
"-;\\ \ \
\.\,\\~
\ I
U !
I I
b
1
3.0
Z(m)
&,(mm)
30 20 10 o
GEOGRID
LAYER
q(kPa) 50 50 30 12 30 50
\\ \
4
12 111 \60
\ I '\
\ \ I ,
'\\ \ ( o. \ \ \ \ 3
\ \ I
\'\\\\ ,. \\\\r
\ \ \
\ \\ \ I I
, \ \\ I I I
\\\,
LOAD DURATION (hr)
---
I
Fig'ure 21. Geogrid elongation (&r> and maximum strain (et> for RMC
reinforced soil walls. Results obtained using two-part
wedge analysis.
451
q (kPa)
I 12 I 30 I 50
$ = 45°
I or,max (1l1Il)1 11
I 21
I 30
$ = 50°
lor ,max (1l1Il) I 7
I 13
I 20
REFERENCES
Bacguelin, F. (1978), "Construction and Instrumentation of Reinforced
Earth Walls in French Highway Administration", Proceedings, Symposium
on Earth Reinforcement, American Society of Civil Engineers,
Pittsburgh, pp. 186-201.
Bassett, R.H. and Last, N.C. (1978), "Reinforcing Earth Below Footings
and Embankments", Proceedings, Symposium on Earth Reinforcement,
American Society of Civil Engineers, Pittsburgh, pp. 202-231.
Berg, R.B., La Rochelle, P., Bonaparte, R. and Tanguay, L. (1987),
"Gaspe Peninsula Reinforced Soil Seawall-Case History", Proceedings,
Symposium on Soil Improvement, ASCE Geotechnical Special Publication
No. 12, Atlantic City, pp. 309-328.
Berg, R.R., Bonaparte, R., Anderson, R.A., and Chouery, V.E. (1986),
"Design, Construction and Performance of Two Geogrid Reinforced Soil
Retaining Walls", Proceedings, Third International Conference on
Geotextiles, Vienna, Vol. 2, pp. 401-406.
Bolton, M.D., Choudhury, S.P. and Pang, P.L.R. (1978), "Reinforced
Earth Walls:-A Centrifugal Model Study", Proceedings, Symposium on
Earth Reinforcement, American SOCiety of Civil Engineers, Pittsburgh,
pp. 252-281.
Bonaparte, R. and Berg, R.R. (1987), "Long-Term Allowable Tension for
Geosynthet i c · Rei nforcement", Proceedi ngs, Geosynthet i cs '87, New
Orleans, Vol. I, pp. 181-192.
Bonaparte, R., Holtz, R.D. and Giroud, J.P. (1985), "Soil
Reinforcement Design Using Geotextiles and Geogrids", Geotextile
Testing and the Design Engineer, American Society for Testing and
Materials, Philadelphia, pp. 69-115.
British Dept. of Transport (1978), "Reinforced Earth Retaining Walls
and Bridge Abutments for Embankments", Technical Memo BE3/78.
Caquot, A. and Kerisel, J. (1956), "Traite de Mecanique des Sols",
Gauthier-Villars, Paris.
Collin, J.G. (1986), Earth Wall DeSign, Ph.D. Dissertation, University
of California, Berkeley, 440 p.
Duncan, J.M. and Chang, C.Y (1970) "Nonlinear Analysis of Stress and
Strain in Soils", Journal of the Soil Mechanics and Foundations
Division, ASCE, Vol. 96, No. SM5, Sep, 1629-1653.
Fang, Y.S. and Ishibashi, I. (1986), "Static Earth Pressures with
Va~ious Wall Movements", Journal of Geotechnical Engineering, Vol.
112, No.3, Mar, pp. 317-333.
Gourc, J.P., Ratel, A. and Delmas, P. (1986), "Design of Fabric
Retaining Walls: The 'Displacement' Method", Proceedings, Third
International Conference on Geotextiles, Vienna, Vol. 2, pp. 289-294.
455
McGown, A., Andrawes, K.Z., Yeo, K.C. and DuBois, D.O. (1984a), "The
Load-Strain-Time Behavior of Tensar Geogrids", Proceedings, Symposium
on Polymer Grid Reinforcement in Civil Engineering, The Institution of
Civil Engineers, London, pp. 11-17.
McGown, A., Paine, N. and Dubois, D.O. (1984b), "Use of Geogrids in
Limit Equilibrium Analysis", Proceedings, Symposium on Polymer Grid
Reinforcement in Civil Engineering, Institution of Civil Engineers,
London, pp. 31-36.
Meyerhof, G.G. (1953), "The Bearing Capacity of Foundations Under
Ecce nt ric and Inc 1i ned Loa ds" , '--P"=ro""'c:..:e:..:e'7d:....:i-:'-'n~g.::.s~,-=-T.:. .h:. :.,',:-;or,-,d=-.;.I:.:..nt~e:..:r...!n.:..;a;-:t:..:.i-70.!.:.nf-!.a1
Conference on Soil Mechanics and Foundation Engineering, Zurich,
Vol. 1, pp. 440-445
Meyerhof, G.G. (1980), "Limit Equilibrium Plasticity ' in Soil
Mechanics", Proceedings, Symposium on Applications of PlastiCity and
Generalized Stress-Strain in Geotechnical Engineering, ASCE, Florida,
pp. 7-24.
Milligan, G.W.E. (1974), "The Behavior of Rigid and Flexible Retaining
Walls in Sand", Ph.D. Dissertation, University of Cambridge, England.
Mill igan, G.W.E. (1983), "Soil Deformations Near Anchored Sheet Pile
Walls", Geotechnigue, Vol. 33, No.1, pp. 41-55.
Mitchell, J.K. (1987), "Reinforcement for Earthwork Construction and
Ground Stabilization", Theme Lecture, Preprint, 'VIII Pan American
Conference on Soil Mechanics and Foundation Engineering, Cartagena,
Aug.
Morgenstern, N.R. and Eisenstein, Z. (1970), "Methods for Estimating
Lateral Loads and Deformations", Proceedings, Specialty Conference on
Lateral Stresses and Earth Retaining Structures, American Society of
Civil Engineers, Ithaca, pp. 51-102.
Murray, R.T. (1984), "Reinforcement Techniques in Repairing Slope
FailtJres", Proceedings, Symposium on Polymer Grid Reinforcement in
Civil Engineering, London, pp. 47-53.
Murray, R.T. (1987), "Factor of Safety ConSiderations Relating to
Reinforced Soil Structures", Preprint, NATO Advanced Research Workshop
on Polymeric Reinforcement in Soil Retaining Structures.
Romstad, K.M., Al-Yassin, A., Hermann, L.R. and Shen, C.K. (1978),
"Stability Analysis of Reinforced Earth Retaining Structures",
Proceedings, Symposium on Earth Reinforcement, Pittsburgh, pp. 685-
713.
Ro~e, P.W., (1969a), "The Relation Between the Shear Strength of Sands
in Triaxial Compression, Plane Strain and Direct Shear", Geotechnigue,
Vol. 19, No.1, pp. 75-86.
Rowe, P.W. (1969b), "Progressive Failure and Strength of a Sand Mass",
Proceedings, Seventh International Conference on Soil Mechanics and
Foundations Engineering, Mexico City, Vol. 1, pp. 341-349.
457
1. INTRODUCTION
- external equilibrium
- critical slip line
- local equilibrium of an isolated reinforcement
H ( a,<p) FIGURE 1:
r.6.H Typical case
459
P. M. Jarrett and A. McGown (eds.), The Application of Polymeric Reinforcement in Soil Retaining Structures, 459-506.
© 1988 by Kluwer Academic Publishers.
460
- Horizontal movement.
2. EXTERNAL EQUILIBRIUM
L = L min (H,f3)
461
* All the methods assume a slip line which is the position of the
maximum tensile stresses in the reinforcement (Juran et al.
-(2)-). This corresponds to the inversion of the directions of
tangential stresses along the soil-geotextile interfaces at the
points of kinematic discontinuity. Tj is the maximum tensile
force in the reinforcement j.
The most general slip line is a concave polygonal slip line. The
class ical procedure of analys is cons ists of cutting the sl iding mass
into n slices. The number of the reinforcing layers cut by the slip
line is represented by r (Figure 6).
Unknowns Eguations
Xi' Zi' e.
l
3(n+1) Xo Zo eO =
° 6
N.
l
n X
n
Z
n
e
n
°
F 1 equilibrium equations 3n
s
T. r
J
464
Janbu's method -(7)- can analyze both circular and polygonal slip
lines: vertical, horizontal, and moment equilibrium of each slice are
satisfied, with the exception of overall moment equilibrium (Ratel
-(8)-). It can thus be used to analyse a polygonal slip line.
LINE
j-J<--_Z
rupture diagram
R R • eatan<l>c
o
o
ASSUMPTIONS:
\. - - mox-~(jsin(<x-~ }d z
.,... -0 T.
AH .
Koter '. equatIon __ rr
4.4 Two-block
lW;
D D c
_____ ~~.T,
a N,
!L--'----
Unknowns Equations
2 equilibrium equations 4
Nab' Tab 2
F
s
468
Unknowns Equations
N equilibrium equations 2
a
F
s
ACTIVE ZONE
FOR LAYER n
ACTIVE Z9NE
FOR LAYER (n·1)
D(n·1)
min
L .
aJ
T . 2 • f (0
n
• tan ¢g ) • dL
aJ
o
Once the slip line has been selected Taj can be calculated.
With Ff fixed, the mobilisable strength in each reinforcement, Tj
mob' can be obtained:
T.
J
F
s
.... Tf
F
s
- ----
To equate Pm ax and the tensile forces in the reinforcements:
P + ~T. = 0
max J
For inclined walls, Jewell et al. -(13)- do the same thing: the
slip line corresponding to the force Pmax .
.II ¢c
-+-+ a = +
4 2
Tf
(same TI for each reinforcement layer)
¢c
-+-+ a < .II +
4 2
/ /
'vi 1 II
-
--------~~-
-P
/,c
Q:'
/
-
R
~ T. ~ min
J
j j
,..- -+-- ~
T J r TENSION IN
REINFORCEMENT
~------~-----~
...... T"=O crt ..........
IT ,"7 }"7~"7
~~ ~~~______-r____U._3____________
ACTIVE RESISTANT
ZONE ZONE
• ~H
H
f ax . dZ p
o
,
This equation is checked in "Standard" method (method C) appl ied to
vertical walls -(18)-
<Pc
since a Ka • Y • (H-z) where Ka tan2 (lI -)
x 4 2
H
f ax . dZ 1. K • Y • H2
2 a
a
in agreem~nt with
(refer to 5.1)
474
p
K • Y • (H - z) where K ~
o
X 1
-yH 2
2
p 1. K • Y • H2
max 2
,
/
"""'""--------"')--..
---=-~ -- --/---=t.--
-W ' T 'J
. . . --~7--+
---- --/~---------
-- --,--.
p -'I~R
L
- .....
1,I
-
'} I
..
T .
local equilibrium: 0xj • ~H ~ min (T I , ~) ~~ TI2
s
T .
global equilibrium: 21.)
F
s
6M
o K [y(H-z) + _z_]
x a (L) 2
y(H-z)
o K
x a 2M
z
1 -
y(H-z)·(L)2
#(-If/
WlTIJJJJ
crn JI
(C)
H
LLIJ (8)
~
I
1//-(/1
IJIDJJ 2. ( A)
DISPLACEMENTS
METHOD
- " -"
sheetj
H
z
SOIL MASS
slip line
passive Zone
I~j<"/2-Wjl t.Zj
aM M
u~I~~""'t~~~l..I2SJ~~J.l
Bj
FIGURE 20: Local behaviour of the geosynthetic layer near the slip
line: anchored membrane -(22)-
478
a
u Aj = v (T aj , La. ) active zone
aJ
P
u Aj = v (T aj , LP . ) passive zone
aJ
* Membrane: This concept has been clarified from the shear test
on a soil-geosynthetic composite, the direction of shear being
perpendicular to the fabric layer (Gourc et al. -(24)-) (Figure
21).
It is considered: K llz./4
s J
B*
j
equilibrium of the membrane: T.
J 2 •
T.
a P
(u Aj + u Aj ) - B*'
J
.~
sini3.
--L
sina.
X.
-1 + 1 +
J B* - (LIZ j - X. )
J J
. cotl a.
J
l 0
• sin a.
J
In the general case, i.e., a polygonal slip line with the appli-
cation of method of slices (Perturbations), the number of equations and
unknowns are as follows:
480
SLIP LINE-
'/'
LAYER j
---- p
Uaj
Displacement Metho d
ond
CARTAGE:smoli displacement
SLIP LINE
LAYER J
Displacement Method \
and
CARTAGE: lorge displocement
Unknowns Eguations
equilibrium of slices 3n
F 1 X Z e 0 3
s 0 0 0
N. ~
n X
n
Z
n
e
n
0 3
III , 112 2 N.
~
N.
~o
(Ill + 112 . f(Si) n
lie 1
4n+2r+7 4n+2r+7
481
8 0,8 I 1°,8
10
:1
I
I ~
~
7 0,8 9 10,8
r
"--- 0,8
'--
6 8 10.8
H-_7 m H
"--- ~
1°,8
5 7
"--- 0,8
"---
4 6 10,8
~ "---
3 0,8 5 }a,8
'--- 4
iO,4
~
10,4
2 1°,8 ~ 3
"-- '---- 2 0,4
1 1°,8 '---- 1
1
.' .•", ~
.'/,'/1
-
L-7.5m
.\ -
L-6m
~z Equilibrium (mm)
60
METHOD:
(E).---.
(A).6--6.
10 30 40
(Figure 24)
!
L
I /
-----~ 1---
..6 8 --I-~
-- --<:i . .----. L--
.
_, ____ ,L..__../__
[t, <DJ I '
H ---- ~-I--I----
-- - ----:,)i--.. -
~--- ~---'"' _....
~=2g.2·
IPgc = (Pb
° 2 =0.
.1H=0.75 m
H(m)
/
/
/ Tr (kN/ml
20
~---17.5
~ _ _ _ _ _ 15
5 /-------125
~--------10
~-----------~
A--------------2.5
L!rr. ;
T;:lkNlml
I-fern)
17.5
/""_ _ _ ,5
f----'25
s
~------10
~--------~
~---------5
~-----------2.5
a'~~:----------:5,----~L{m)
H{m)
T;(kNlml
/ -_ _ _ 12.5
/ -_ _ _ _ _ 10
5 / -_ _ _ _ _ _ _ _ 7.5
; -_ _ _ _ _ _ _ _ 5
}-_ _ _ _ _ _ _ _ _ _ _ 25
a 5 L(m)
'f
. .
retalnlng wa 1 1 s [ F.G =1.5
_design method FR:=1.5
Fs=1
• 1P=25°
... rp=30° ..... .....
• 1P= 40 o
. . . --.----+
"' ,
"' ,
10 ""
"'
"' "'
5 "' • ___ -e
.---~
~ -.-.-.-~-.---~-
90 60 60 90
Tz (kNlm)
40
30
20
10
(]>(. )
10
20
T, (kNlm)
40
30
20
10
10
20
TI (F,=1s)
H = 7.5 _ A ~ = 90°
TI(s: 1)
_ - . p = 60°
I
•
I
I
I
/
/
2
•
I
I
I
I
. ;/ .
A
----.-- -
-_
2 A~
A
Lmin(;;=15)
Lmin(F,:1 )
In paragraph 4.4 , it is seen that the incl inat ion Ii (Figure 24)
must be fixed first.
In the case where 'Pb = <Pgc = tan- 1 (2/3 tan <Pc), the method
of circular sliding overestimates Fs. This is because the slip line
of "two-block method" can follow the weak plane of the wall base, but
not for circular slip line.
I
~ ! !"lM,7X\
! 7
! 7
(2) 0 0 0 ~ =60'
30
'8 qJ.
C
(f)
'C 0 0
(f)
,b
(f)
'C
(j)
'gc
(Pc (f)
'gc
20
10
H{m)
o 5
F 'P.'
c
29.2 ..
Fp Fb
0
F
1.2
21 tJ.
1.2 ..
17.3 • 0
~ IJi i
tJ.
F F
02~ 0 ; rpb=rpc
tJ.
f
'*
1.2 1.2
i ~ .1> a
, _ _ _r!l__ ~------
•
o
0.8 0.8
H(m) H(m)
0 5 10 0 5 10
F F
C)2=0 I"b = rpgc
O~O ; rpb = rpgc
t
~
a
~
1.2 1.2 l:.
tl ~
I .- - - _.i' - -~ _______ _
•
o
0.8 0.8
H(m) H(m)
o 5 10 o 5 10
X (m) Z(m) I F
I Cb -3 9 I 1.08
.... _-- I Cp -3 9 I 1.07
P(kN/m) T:r(kN/m)
lSc 107 11
l Sm 107 11
"Perturbations method"
"Bishop~simplified method"
Z
~=l: "Two-block method"
For a vertical wall, the Sm slip line (for the maximal value of
the thrust Pm) agrees with Rankine wedge. Sm is independant of the
reinforced retaining wall width L.
On the other hand, the Sc slip line (for the critical value of
the thrust Pc is a funct ion of the width L. As a matter of fact, the
computer calculation provides two slip lines for every value of the
width L (Sci, "inferior", and Scs, "superior") and for the same maximal
value of TI'
The larger the L value, the larger is the critical thrust value
until the maximal thrust Pm is reached (superabundant anchorage length
Laj). Also the larger the L value, the smaller is the TI value until
its minimum for Pc = Pm is reached.
F = 1 (6.2.1)
s
492
20 --- T:{ Sc l
f--+ T,{Sml
"'---J.. Pm FIGURE 31 -a-
n
• p.
\•
".--II~I
n "Two-block method": deviation of
0---0 ~ TI values and critical slip line
15
" wit h the choice of thrust (Pm or
Pc) and the retaining wall width.
+,~
I +--- ~
H 4.5 m
=
___ 5 e i
H 4.5 ____ sc:.s
_ _ Sm
L= 3 m
\ ~;=64
\ Pc,=70
\
L:3.5m 1 L= 4 m
\ Pcs=70
~;=70
\
\ \!
L:: 5 m
493
If!: = 29.2'
I
~
'Pgc = rpb
.\
30 ~ =90'
°2=0'
~H=0.75m
= 7.5 L
\
H m
.--.----------
."", Poi with the choice of thrust (Pm or
Pc) and the retaining wall width .
• ___ _____ n H = 7.5 m
20
A __ A __ A ___ .A ____ ....~~
./
~:/
/~ /.-:?"
0.... ./ -'\iiI:;::>''-
r:/ ./
.
1S .'o....
. /" /
./ "" /
/
1/
•
v+---~--~----~--~----~--~--~-
5 6 7 Um)
\ \.
Pci.=-115 '\'. Pd~135
Pes=161
\.. Pc.=153
\
\ \' ...
\ "... "',
''':':'''
'"":~..
Pei=150
Pc.s=150 \. \
"
\ Pei=191
\PQ2194
\
\
\
\
D
~D = 40
'Pg = I"b
Ti(kN/ml f3 =gO'
0,= o·
.1H:O.75m
-,,,,~,, LmfJ1
50
10
2 J 5 6 7 H(ml
o a-- __ FORCsr.sDN/CE:
F = 1. 5
s
F = 1. 5
s
495
tp = 40
'P" ='Pb i - - - - ll ,
~ =90' ~==1~ilH
0,=0'
LlH=0.75 m
Lmin
50
10
2 J 5 6 7 Him)
~~
'~
...... .....
...........
, ..., ..... ," ,
A +- - TERM: ARMEE
"
' .....
5 B {l-.-TRRL
ABeD "b= FR=1.5 F,,=1
C ll ____ STANDARD
o a----FOR£STSERV/CE
F = 1. 5
s
F
s
= 1. 5
496
H 3 m L 3 m y 17.9 kN/m
2
43° <j>
g
= 43° c = 0 c
g
o
For the case shown in Figure 34, the influence of the parameters
up, La' and J on Ta f(uA) is presented three-dimensionally in
Figures 35-37.
0=12 kPa
10 20 30 Ti 20 40 60 ax
(kN) (kPa)
II: \1\
~I:
\\: I
\\t~ \\\ \ \\ \\
\\ \
t\ \ \1 ~
c~rcle( Fs:::;1.03)
x=_6.75
z"" 4.25
\l ~\
0= SOkPa
10 20 30 40 Ti ,I
1/
(kN)
i
W
\\ 1\ \\
l\ ~
IIi doubl~ bloc I
It /[ t
I
T\
~\ \'\ 1\ \\
1\ '. .Ii)
1: coin Ronkin. \
:! 1 J
ii \
I
I \ \ ~ I' x=-8.7S
z .. 5.25
!
V
0= 1QOkPa
A +- TcAAC ARI-IEE
B (t---TRRL
C 6-~---STANDARO
0 . ____ FOREsr.sERVlr.E.
10 20 30 40 Ti 20 40 60
(kN) (Jd'b)
•• fL,/ i
1/
~r
L \\ \
\
\
,,"
double- bloc
I,
Iii \1 \
I
\\
\ \I i/
\
~\. \
\
\
I-
"
\\ \ \ 1i
c~rcle ( Fs_1.04 )
~;-m
\
I
,:1
/ \\ Ii
I
\ I , COin Rankin~
,
)2'
i
I
I
I: I I \\ ,
X
i
FIGURE 33: Comparison of traditional design methods with "two-block
method" and "method of slices" associated with circular slip line
(RMC-Kingston geogrid reinforced earth wall)
498
J
-f-.L-~---_ E!%l
Tensiletest
200
200
J 500 KN/M 5 M 180/~'
P ~
20,78 KPA C
160,-,
7'
UO S:-
Valeur deTA
..--
~
/20
<lJ
0 HBGIJE 189 100 (J1
o
0 170 - 189 ~
-
113 - 132
94 - 113
76 - 94
Ifill 57 - 76
~ 33 - 57
19 38
BELOIJ Ig
.:;~\\\\\\I
-, ,39;'
(- +----------- ""
___ _ .
<~~ ! '4 2 -1 - - - -
J = 500 KrUM ,
Up = 0, 01 11
QJ 2 4 -2 ,-
0p 20,78 KPA
en
TA
2 192
V a le u r d e
-.=.
r- 14 2 242 Q)
o CJJ
8o 1l6UIJ[ JS i'
.J,:) - .1.')c
31</ - ].1;;
«
1-
92 19 2
U
oL
::
i'SO - J I</ 14 2 C
i" 16 - ,'.') f) o
-
212 - 2'10 «
17,1 - 212
92
I- -
1'1; - 17,)
~ Iff ) - 1'1,1
7,; - 110
B[LOI) 76
ri e d in
e o sy n th e ti c b u
~: e b eh av io
u r o f a go f L a
A n ch o railg m as s: in fl u e n c e
sO
Up = 0,01 M 5 M
0p = 20.78 KPA
V a le u r d e T A
CJ
:-
n6(j{J[ 19 1
=
0 17'1 191
0 /,"),')
I ,If)
// .-,
157
1('3 I'll )
:
11)6 Ie,
09 1f)6
=
72 39
~ $') . - ,(
,8
T')
.f;!;
6[LOr,) 38
th e ti c b
u ri e d in
e o sy n
~: ho ra ge b eh av io
in
u r o fn cae go f J
fl u e
A nC o il m as s:
s
500
.1Zj
LOI d~ DEFORMATION:
T~ Tp='fZ.kJ'Pg+C g
J>-- --J> DOUBLE BLOC
Tp •. __ , U =00002S.~
METHODE
____ on DE
'.lOOH PlACEMENT
J=43SkN/M 4 f---J---;!!--~ : p, rz·S4100
: U(m)
~z (m)= 0,0112 J f-.,..!ll!.ll!----!~--l Up
METHODE CARTAGE L.C.?'C
o--o.=lOOH J:-43SkN/M
~z (m)= 0,003 4
Q= 50kPA
x
10 20 30 40 50 60 Ti(kN/M) 1! !
I
I
\ Ii
/ METHODE ~n DEPLACEM!
METHODE CARTAGE
I (x.S,OO, Z.:;OOI
, /
2 up " O.D1Sm
q " 2SkPa
,,
.. I
+ 5
I
I
I
I
101
J.. 1 Lit no) lei
TlkN/ml
20
l Cas nO 1 I2 J ,
!JlkN/ml 500 I 2000 500 500
I fixe
.
I"paremenr' libre fixe fixe
10
1+ • .
Parement
-----
o Iml
Ibl
c K
/
L(m)
IP.: =29.2·
17.5
IP.: =37.4"
S
fP9C = fPb
,
°=
, ~ =90·
I
I
I 2 =0.
5 I .1H 0.75 m
I
I kaO;1/1: (FootS)
I
I k.OjCP;
".-.w
H ~.1 H
'" L
H(m)
5
H ~ 7.5 m
<Pc =21'
4-
I 7
iPgc = rpb H ! 7
O2 =0' / 7 L'.H
~g_--t7
L'.H=0.75 m """,.,,,[}I-' 7
Lmin
3 -- f3 =90'
..0.---.0. f3 =60'
2
o 5 10
7. CONCLUSION
The numerical and graphic methods will provide the engineers with
explicit charts for rapid predesign. However, it is also necessary to
know well the fundamental assumptions in each method to estimate the
design scattering.
ACKNOWLEDGEMENT
REFERENCES
5. Phan, T.L., Segrest in, P., Schlosser, F., Long, N.T. Etude de
la stabilite interne des ouvrages en terre armee par deux
methodes de cercle de rupture. COllOje International sur Ie
Renforcement des Sols. Paris, 1979. ,
13. Jewell, K.A., Paine, N., Wood, R.I. Design Methods for Steep
Reinforced Embankments. Symposium on Polymer Grid Reinforce-
ment in Civil Engineering. 1984.
20. Segrestin, P. Calcul d'un massif en terre armee par les coins
de rupture. International Conference on Reinforced Earth.
Paris, 1979.
23. Gourc, J.P., Ratel, A., Delmas, Ph. Design of Fabric Retaining
Walls: The "Displacements Method". 3rd International Confer-
ence on Geotextiles and Geomembranes .. Vienna, 1986.
ABSTRACT
1. INTRODUCTION
507
P. M. Jarrett and A. McGown (eds.), The Application of Polymeric Reinforcement in Soil Retaining Structures, 507-540.
Crown © 1988.
508
D ~ Ka'YzSVSH . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (1)
R~2BJLL'YzLL . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (2)
i.e R~D
KaSVSH
i.e. LL~------- . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (3)
2JLLB
Now the length to be employed in the final design must have an adequate
factor of safety against adherence failure and the required design
length is therefore given by :
JLL
JL (minimum)~ <--- JL,K~ function(F) ---> K(maximum)~ Ka.F
F
This relation states that in one extreme only the interface friction is
influenced by the safety factor while the other extreme is the full
effect of safety factor being taken only by the soil friction. Between
these extremes both JL and K are affected to some extent.
510
ootan0p
i.e ~M ootan0M~ - - - - - - - - .......................... (4)
FS
KMSVSH
LD~- - - - - - . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (5)
2~MB
now as F
F ----------- . . . . . . . . . . . . . . . . . . . . . . . . . . (8)
F.Ka
and FS~ ..................•.................... (9a)
RM 2BLDJ.tM
i. e. FA
DM SVSHKM
2BLL Ka
and where LL LD/F
SVSH J.tL
............................ (11)
F.Ka
for f o FS ......................... .... ... (9b)
KO
4 .00
.... ....
.... ....
Fs (Eqn 9a)
---~ ..... ---
........ "
1.00
0 .80 "- ..
.. --
-. - - - - __ -. q,p == 30°
tpp,--
U
'"
.§-
!t 0.60
~
:J
- _
tpp ~ 50
--
0
~400
a.
.s
u..
0.40
'0
a
.;:;
'"
a:
0.20 Local stability assessment:
rotation about toe
0 .00
1.00 1.08 1.16 1.24 1.32 1.40 1.48 1.56 1.64
Safety factor of the soil
a·PL
FR -------- ........................... (13)
KM'YzSVSH
i.e. FR
.. oL
'I .'
I
I
T
Sv
1
z ...L
ith reinforcement
.. ,
Nth reinforcement
,....,,0------ L -------1
.. 1
1.00 ,..--r--------------------------...,
Overall stability assessment:
rotation about toe
0.80
<>
'"
~
!S .
0.60
'"
>
..£
u.
'0 0.40
.g
<0
a:
0.20
0.00
0.80 1.20 1.60 2.00 2.40 2.80 3.20 3.60 4.00
Factor of safety of the soil
i=N
TT = 2: T .... (17)
i=l
2 (L - Htan(3) 2 1
i.e. TT ~~B~LH .2(N + 1)[------[(--- + tanf3( '3 + 6N)] ... (18)
Now the disturbing force D = ~~H2, where K takes the value of Ka at the
limiting condition or KM according to the mobilised friction angle. The
specified factor of safety (F) is therefore given by :
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (19)
2B~ij
FOVER = -KM (N
L
+ 1) [il - -'3
JKij
] ......................... (20)
The ratio of FOVER/F has been determined from the values of KM and FS
calculated in the local stability analyses employing Equations 7 and 10
respectively, The results are presented for peak friction angles of
30°, 40° and 50° in Figure 4 for geometries of L/H of 0,6 and 1.6. As
can be seen the results are very similar for these two cases in view of
the fact that the data have been normalised with respect to the limiting
conditions.
It may be noted that the bracketed terms in Equation 21 have very
similar numerical values and a close approximation to the safety factor
against bond failure is given by :
F,K a
FOVER
518
The above equation corresponds to that derived for local bond stability
(Equation 11) and the agreement between the two possible modes of
instability can be explained by the fact that the same relation between
specified factor of safety and mobilised friction angle has been assumed
for both cases. In practice, the mobilised friction angle may vary
throughout the height of the wall according to the magnitude and form of
yielding that takes place. However, the analysis which has been
described above relates to the most commonly assumed case of outward
rotation about the base.
A further assumption involved in the method is that the total force (P)
at the limiting condition conforms to that produced by the usual Coulomb
analysis. The lateral earth pressure is determined by differentiation of
the total force equation and treating Kaz as a variable :
i. e. P = ~Kaz)'z2
dP d
i.e.
dz Kaz)'z + ~)'z2·dz(Kaz)
now
519
i. e.
F - - - - - - - - - - - . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (24)
FStanOo + tanOp
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . (25)
FS - tanOotanOp
(FS.tanBo + tanBp)
where 0MZ tan -1 ( - - - - - - - - - - - - - - - - ) ................ . (27)
(FS - tanBotanBp )
Note that as FS also occurs in the term KMZ, Equation 28a must be solved
iteratively. More rapid convergence is obtained by casting the equation
into the following form for iteration purposes :
- - - - - - - - - - - - - - - - - - - - - - - - - - - - - - . . . . . . . . . . . . . . (2 8b)
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (29)
The values of KMZ are obtained from Equation 26 while FS is derived from
Equation 28 based on the assessment of local bond stability. The
variation of FR and FS over the depth of the structure for a specified
value of F equal to unity is shown plotted against depth factor in
Figure 5. Although the limiting condition corresponding to unit factor
of safety has been specified, this only occurs at the base where
sufficient movement has occurred. Figure 5 shows that the strength of
the soil towards the upper part of the structure is not fully mobilised
by virtue of the fact that FS exceeds unity and also as FR is less than
unity. These results may be contrasted with the corresponding condition
for rotation about the base when the factors of safety are the same. The
influence of wall movement on mobilised friction angle may be seen more
clearly in Figure 6 where the relations between depth factor and
mobiiised friction angle are presented for peak angles of 30°, 40° and
50°.
521
2.00 r-----------------------------"I
1.60
~
~
~
....0 Fs
~
1.20
B
~"
-0
~
~ 0.80 FR
--
::J
co"
u
Fig. 5 Relations between depth factor and F(rupture) for F(spec) of unity
50
r- c\>'P 50°
40 r -
-
'"c
C>
co
c
0
30 r
.;;
~ t--
-0
.~
:0
20 r
0
:2:
1.00 r----------------------------,
0.80
uw
-
~ -.....
0.60 I-
~,.w ...... ~
.3
LL
a
a
.;::;
0.40 t-
r-
--
l=--------------------------~::~~::~-~==-:~::-:~~~ - __
......... Fover
a:'" F
0.20 t-
l Overall stability assessment:
rotation about top
J
0.00 1
L-.._ _-L. 1
__---I_ _ _J I
..I..-_ _--'-_ I
_ _.L..-_ I _ _ _'I - - _ - - - '
_- L
1.00
0.80
uw Top rotation
~
0.60
~,.w
.3
LL
....a
a 0.40
.;::;
a:'"
0.00
1.00 1.08 1.16 1.24 1.32 1.40 1.48 1.56 1.64
Safety factor of the soil
40
Plain strain tests on
medium dense sand
32
Q)
~
cit (ev)
::s'"
Q)
Q)
C> 24
c:
'"o
c:
.;:;
u
:E
"0
16
.~
:.0
o
:2:
o 2 4 6 8 10 12 14 16
Lateral strain (per cent)
100
c: 60
o
.;;;
c: ' Rigid'Reinforcement
~
"0
~
~
::>
40
u
"iii
U
20
o ~----~----~----~----~--__~____~~____~__~
0.00 0.20 0.40 0.60 0.80 1.00 1.20 1.40 1.60
Distance from facing (X/ U
T = m. €R . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (31)
The relation between lateral strain in the soil (€S) and mobilised
friction angle is given by
Now for compatibility, the (average) strain in the reinforcement and soil
must be equal :
dT
dx
Where ~M = f(x)
where A = surface area of reinforcement
where Cl = constant of proportionality
dT
dx
2AavaLnCl
Tx ---------.In.[cos(0KO + n€M(l-x/L»)] + C2
n€M
when x=O ; TX = TM
when x = L TX = 0
o=
_____ !~:!~I_~?~~KO_l_
In[-~?~~~KO-:-~~~~-]
[ cos0KO ]
For the simple case assumed, € varies linearly from the maximum value em
at the facing to zero at L :
o
2B~zaLD (COS00 )
- - - - - - - - -In. ( - - - - - - - - - - - - -) ........................ (37)
n€M (COS(00 + n€M»
The disturbing force corresponds to the lateral force acting over the
area of facing supported by a single element :
R
F
D
Ka (COS00 )
i.e - - - - - - - - - - - - - ln ( - - - - - - - - - - - - - - ) . .. . ............. (38)
F KM·n·€M·tan0L (COS(00 + n.€m»
The calculated factor of safety is plotted versus the depth below the
surface in Figure 11 for peak friction angles of 30°, 40° and '50°. The
reasons for the relatively small factors of safety are as follows :
(1) The mobilised friction angle is less than the limiting value
behind the facing so that the lateral force is greater than the minimum
value corresponding to active earth pressure conditions.
(2) The magnitude of the lateral strain redu~es along the length
of the reinforcement with a corresponding reduction in interface
friction so that less pull-out force is developed.
However, as the limiting conditions approach, greater strains will be
induced and the available soil strength will increase until it is fully
mobilised. The development of such strains will rarely be uniform and
the maximum strength may be achieved at some locations while at others
only a small proportion will have been attained. Thus with increasing
strain the former locations will tend to reduce in strength as post-peak
conditions are developed while elsewhere the strength may be approaching
or at the peak value. It is unlikely, therefore, that with extensible
reinforcement the peak strength will be attained simultaneously at all
locations and although the available resistance in adherence will
improve with increasing strain, it will be usually less than calculated
on the basis of ideal limiting strength values. A prudent approach to
design would thus be to assume that the maximum available friction angle
of the soil corresponds to the ultimate friction angle of the soil 0CV .
• Notwithstanding that the ultimate pull-out resistance is likely
to be less than calculated on the basis of fully mobilised strengths,
the local adherence stability based on conventional design equations
will generally be adequate although the factor of safety at the working
condition will be usually smaller than specified. Of greater
significance will be the effect of the smaller mobilised soil strengths
in relation to potential rupture.
The assessment of local stability against tensile rupture was
based on the same method as described in Section 2.1 whereby the lateral
forces and mobilised friction angles determined for adherence stability
are again employed. It should be noted, however, that such an approach
is not necessary in this case as an independent calculation of local
stability against rupture would produce the same strains and mobilised
frictions and thus highlights the much greater consistency of a method
based on a strain assessment. The relations between depth factor and
calculated factor of safety, for the same values of peak friction angle
529
1.00 r - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - ,
0.80
~
.:::
:Jl
'0 0.60
2
~"
-0
~ 0.40
::l
ro"
u
Local stability assessment:
0.20 strain analysis - bond
0.00 L-_ _-L._ _ _.1.-_ _-..L_ _ _- ' -_ _-..L_ _ _- ' -_ _ _' - -_ _......
1.00
0.80
~
.:::
:Jl
'0 0.60
(;
1:)
~
-0
2l
~ 0.40
ro"
u
0.00
0.00 0.20 0.40 0.60 0.80 1.00 1.20 1.40 1.60
Depth factor (Z/H)
Fig. 12 Relations between depth and calculated factor of safety
530
The safety factor against forward sliding of the wall on the basis of
limiting strengths is given by :
W.tan0p
F .............................................. (39)
and the limiting width of wall (LL) at the point of sliding is obtained
from :
531
Ka· z
............................................. (40)
2tan0p
i.e.
2tan0p
where KM -----------------------
As was the case for internal stability it is apparent that there are
inconsistencies in such an approach as the greater the ultimate bearing
capacity relative to the applied pressure , the smaller the amount of
yielding taking place with less scope for mobilising the shear strength
of the backfill. An alternative method is described which involves an
assumption that the mobilised shear strength is reduced in inverse
proportion to the factor of safety against bearing failure. The
procedure is therefore similar to that employed previously for
533
2.50
External stability:
sliding resistance assessment
2.00
~
~
~
.....a 1.50
B0
~
-0
j!l
'" 1.00
:;
0
"iii
II
0.50
1.00 .-=----------------------------,
cpp = 50° cjl'p = 40° cjl' = 30°
........ '=----.:.:..-=.:--==-~-..:.:~- - - ~~
0.80
uCl)
~ 0.60
~
E
a
E-
LL
.....a 0.40
a
'';:;
0:'"
0.20 External stability:
overturning moment assessment
Now if the shear strength of the soil is not fully mobilised the lateral
earth pressure coefficient increases to KM and the actual factor of
safety against overturning (FMOM) is given by :
Equations 46 and 47 may be solved for FMOH provided the factor of safety
of the soil (FS) is known :
TABLE 1
--
Undamaged
)( reinforcement
0,-
. ---._)( Site damaged
reinforcement
""0
'"o
-'
Strain
7. CONCLUSIONS.
8. ACKNOWLEDGEMENTS.
The work described in this paper forms part of the research programme
of the Transport and Road Research Laboratory and the paper is published
by permission of the Director. The author is particularly grateful to
Mr. I. F. Symons of TRRL for a number of helpful suggestions in the
preparation of this paper.
538
9. REFERENCES.
Bishop,A.W. (1955) The use of the slip circle in the stability analysis
of slopes.Geotechnique 5,pp 8 - 17 ,London
LIST OF SYMBOLS.
1. INTRODUCTION
The beneficial effect of incorporating tensile inclusions
within a soil mass is well recognized and has been demonstrated
by the successful construction of numerous reinforced soil
walls using facings ranging from relatively rigid full face
concrete panels to a flexible geotextile "skin", and reinforce-
ment ranging from relatively stiff steel strips or meshes to
geotextile sheets of low stiffness. Design methods for rein-
forced-soil walls ("7hich are discussed in detail in other
papers at this workshop) are typically based on limit equili-
brium calculations "Thich do not explicitly consider deforma-
tions or interaction beh7een the inclusion and the soil.
In the case of Reinforced Earth (R) walls, there is now
twenty years of empirical evidence to suggest that the approxi-
mate method of analysis, in conjunction with the normally
adopted soil properties and safety factors, provides a safe de-
sign which gives acceptably small deformations under working
conditions. This may also be the case for some geotextile and
geogrid reinforced walls designed using current practice, how-
ever the "Tide range of facings, backfill materials and proper-
ties of the reinforcement leads one to question the generality
of the simplified methods of analysis that are being proposed
for the design of reinforced soil walls with geosynthetic in-
clusions. Questions that can be raised include:
- what is the effect of reinforcement extensibility;
- what is the effect of using different facings and construc-
tion techniques;
- what is the-effect of soil-facing-reinforcement interaction
both during construction and subsequently (over the life of
the structure);
- under what circumstances might one expect strain-softening
within the soil mass and what influence do the properties
of the reinforcement have on strain softening;
- what is the "Factor of Safety" of a reinforced soil wall?
In prinClple, these questions could all be answered by the
construction and monitoring of a large number of full scale
field t~st walls. Unfortunately, the cost of performing and
adequately monitoring a sufficiently large number of full sc~le
walls is so large that it is not practical to perform a detail-
ed experimental study. "Numerical experiments" or simulations
provide an alternative and more cost effective means of perform-
ing such a study.
Finite Element techniques have the potential to allow us to:
541
P. M. Jarrett and A. McGown (eds.), The Application of Polymeric Reinforcement in Soil Retaining Structures, 541-553.
© 1988 by Kluwer Academic Publishers.
542
(E/p ) = K(aIP )n (1 )
a a
where E is the Young's modulus of the soil, a is the minor
principal stress or the mean stress depending on the details of
the formulation, P a is atmospheric pressure and K and n are the
material parameters. This non-linearity is included in the
"hyperbolic" model and can also be readily included in non-
linear elastic-plastic models (eg. Rowe, 1986). It should be
recognized that the modelling of "yield" implicit in Eq. 1 is
only approximate and is not appropriate for situations where
there may be cyclic loading (eg. see Zylynski et al., 1978).
However, there is considerable evidence to suggest that this
approach can provide reasonable results for problems involving
monotonic loading (as is generally the case in modelling wall
construction) .
Non-linear elastic (hyperbolic) models can be expected to
provide acceptable results at low stress levels (eg. when there
is a large "factor of safety"), however since they are based on
elastic theory they can not correctly model plastic failure and
plastic strains within the soil mass (it is noted that the use
of a cohesion- intercept c and friction angle ¢ in a hyperbolic
model does not imply that the model is a plasticity model--eg.
see Duncan, 1980). Numerous plasticity formulations have been
proposed in the literature. The simplest of these involves
Mohr-Coulomb failure surface and a non-associated flow rule.
This model has been successfully applied in the analysis of
geotextile reinforced embankments (eg. Rowe, 1982; Rowe, 1984;
Rowe et al., 1984; Rowe and Soderman, 1984). This form of ana-
lysis can be readily modified to include the consideration of
a non-linear failure envelope commonly encountered with granular
materials (see Rowe et al., 1982). These models can be expected
to model the soil behaviour up to arid including failure. By
examining the results of studies performed using this class of
model it is possible to assess the magnitude of the strains to
be expected prior to collapse of the structure and hence to
make some initial assessment of potential significance of
strain 90ftening. However, this class of model is not suitable
for modelling strain-softening behaviour and indeed the model-
ling of localization and strain softening even for unreinforced
granular materials requires considerable additional research.
5. FOUNDATION-WALL INTERACTION
One of the advantages of a reinforced soil wall over conven-
tional wall systems is that it should be more tolerant of de-
formations and stresses induced by some yielding in the founda-
tion, thereby allowing construction of walls on less than ideal
sites. Unfortunately, conventional methods of analysis (ego
see Gourc et al., 1987) can not provide insight regarding the
effect of foundation movements on the stresses and deformations
of the wall. The finite element method is ideally suited for
modelling the foundation-reinforced soil wall interaction which
would occur when there is yielding in the foundation soils.
Modelling of this interaction will, however, require the use
of a constitutive model that models plastic strains using a
consistent plasticity formulation which can take account of the
influence of rotations in principal stress directions which
will occur near the toe of the wall. As a prerequisite, the
finite element procedure adopted in these calculations should
be capable of accurately predicting bearing capacity collapse
loads and should be calibrated against relevant published bear-
ing capacity solutions (ego Davis and Booker, 1973).
551
6. TIME EFFECTS
An important concern in the analysis of walls reinforced with
geosynthetics is the effect of the time dependent characteris-
tics of the reinforcing material. In principle, finite element
methods are well suited to modelling creep/relaxation in both
the reinforcement and the soil (the latter being of particular
importance if cohesive backfills are used). Numerous visco-
elastic-plastic finite element formulations have been published
in the literature however, as yet, these techniques have not
been applied to a comprehensive study of reinforced soil wall
systems. Hodelling of this time dependent behaviour is an im-
portant challenge but one that can not be fully met until there
is good quality field (or model scale) time dependent test data
which can be used for validating the finite element calcula-
tions.
7. CONCLUSION
In this paper we have attempted to review the application of
the Finite Element technique to the analysis of reinforced soil
walls. There are still many unanswered questions regarding the
behaviour of reinforced soil structures and the finite element
method provides a very useful tool which can be used to help
anS\oler these questions. However, it has also been emphasized
that there are many different types of finite element analyses
and that considerable care must be exercised in both the selec-
tion of the particular finite element formulation to be adopted
(eg. in the choice of constitutive model for the soil; modelling
of the interface, etc.) and in the detailed application of the
technique (eg. choice of elements, distribution of elements,
construction simulation, etc.). There is also considerable
scope for additional research in developing or adapting techni-
ques for modelling time dependent interaction between the
various components of the reinforced soil system as well as for
developing techniques which model strain softening and locali-
zation (a major problem in itself) within reinforced soil
systems.
ACKNOWLEDGEMENTS
This review forms part of a general programme of research
into reinforced soil and geosynthetics being conducted by the
Geotechnical Research Centre with funding from the Natural
Sciences and Engineering Research Council of Canada under grant
A1007.
552
REFERENCES
12. Naylor DJ: A Study of r.e. Wall Allowing Strip Slip. Proc.
ASCE Symposium on Earth Reinforcement, Pittsburg, pp. 618-
643, 1978.
13. Naylor DJ, Richards H: Slipping Strip Analysis of Reinforc-
ed Earth. Int. J. for Numerical and Analytical Methods in
Geomechanics, Vol. 2, pp. 343-366, 1978.
14. Romstad KM, Herrmann LR, Shen CK: Integrated Study of Rein-
forced Earth - I. Theoretical Formulation. ASCE, J. Geotech
Eng. Div., Vol. 102, No. GT5, pp. 457-471, 1976.
15. Rowe RK, Lo KY, Tham L: The Analysis of Tunnels and Shafts
in ' Dense (Oil) Sands. Proceedings of the Fourth Inter-
national Conference on Numerical Methods in Geomechanics,
Edmonton, pp. 587 - 596, 1982.
16. Rowe RK: The Analysis of an Embankment Constructed On a
Geotextile. Proc. 2nd Int. Conf. on Geotextiles, Las Vegas,
2, pp. 677-682, 1982.
553
The chairman, Rowe, began the session by posing a number of questions that
require answers when considering analysis and design and which the various
speakers would address during the session. They included the following:
555
P. M. Jarrett and A. McGown (eds.), The Application of Polymeric Reinforcement in Soil Retaining Structures, 555-556.
© 1988 by Kluwer Academic Publishers.
556
Rowe presented his contribution which dealt with the general methodology
for deve'loping a successful finite element analysis for reinforced soil
walls. He suggested that finite element analysis was sufficiently complex
that it is unlikely to become a day to day design tool but is likely to be
used in the development of parametric studies and design charts. One
point of discussion of importance to most analyses was the need to
simulate the construction process and thus avoid the "switch on gravity"
approach in which gravitational forces are applied only to an idealized
complete wall model.
In summary during this and most other sessions there was general agreement
that <Pcv was the best and safest strength parameter for general design
usage. Limit equilibrium analyses still provide an acceptable simple
method, for general design and that there is hope to improve them through
better comprehension of ~ the aspects involved. This includes the soil.
the polymer and their interaction. Finite element analysis can be a most
valuable aid in developing parametric studies and making the best use of
the limited amount of case study information available. Formal contribu-
tions to this session were provided by Scott. Floss and Milligan and they
follow this review.
FINITE ELEMENT ANALYSIS OF REINFORCED SOIL
RlgldFoundatJon
-.
I ,.".,.. RemrI SM,
-- i
~\
-
--~ .
l-tii~·i"I· '
--
AIInfCIrc:ement a.r..nt--+--"
...
.! 55
\1 < zseel '--'-'-'---'.:.~ I!~
c \ '. """
o
~\~:.................:::::::...:.:::::.:::,:.:.....:::::,:::..... .... ,
II
.c I
e ~
w50 --.. lio 120
• I
, i i i
: :,;, TO
i!l
~
• M_4""
"'H'~'.'~'" • E
'ii"H'~'ii';;; ~ -H-:'-';'--
ii·H;·i~' i
H .. IO,..
'i"'H~rom i
:;~ / ' ------- ' ~ E
- +... g:f
:l~
Vl
~ 3m '
/' /
//,\" '\
3 m ....... 0.0"0111"'- ~ V / ...... \\.\
./",-:::::= ~1 , /'
J" .5
~'" ,/ '
.. ... \'\\\". '8
-'_/ ,p:-::!i. -/ -'/ ...... , ... -.3
g - ~ \, .
,~:::::::::---- ,/,.;'
."..- -'- -- - ......... ,"'S-'
-::::::::::"~""""""~,::,"''', ,," ' \\';;1 a
i;;'
-----
.. ,.:3:::-::;:-::=:7:::::::;:::::;:=·····..·····
t5 IV
!8
o
a \ 25 20 '5 10
Ohdonce from Foce of Embankment. m
0
561
P. M. Jarrett and A. McGown (eds.), The Application of Polymeric Reinforcement in Soil Retaining Structures, 561-567.
© 1988 by Kluwer Academic Publishers.
562
E, Ez
'15'0
Smox Smox
, --- Go
z
S' ---+- Ze -+
<D active zone Z
B' =B-Z·e E,<Ez(tono.OJ ~o = Yt. 'l'iz
+
!llllill
Tz=N· ton y;
\: reinforcement
FIGURE 3 Block model for steep slope with two elements, and
showing approach of tensile force distribution in the failure
zone
567
569
P. M. Jarrett and A. McGown (eds.), The Application of Polymeric Reinforcement in Soil Retaining Structures, 569-570.
© 1988 by Kluwer Academic Publishers.
570
References
i' = ¥
shear strain = Z() Sec.
volumetric strain = -.l.() tiu-.. r= v
----______~------~~--_r~~--~c.
horizontal strain -/E"
1. INTRODUCTION
The modern forms of reinforced soil have evolved from an
understanding and introduction of effective .construction techniques,
the first of which was developed by Vidal (1966). The first
structures used steel strip reinforcements arid it was not until the
early 1970' s that polymeric reinforcements were used successfully
with vertical faced structures. Prior to this, high density
polythene grids had been used in the construction of railway
embankments in Japan, Yamamoto (1966).
This paper considers the practical aspects of the construction of
vertical or near vertical reinforced soil structures formed using
polymeric reinforcements. Included are details relating to
construction systems and techniques, economics and specifications.
Although there are numerous other possible construction arrangements,
those detailed here have all been used in practice. Where possible
the potential advantages displayed by individual systems or materials
are described.
2. CONSTRUCTION METHODS
Reinforced soil structures must be of a form in keeping with the
assumed idealization and analysis, however, the theoretical form of
the structure may be quite different from the economical prototype,
and attention should be paid to the method of construction throughout
the design process.
Speed of construction is usually essential to achieve economy and
in part this may be achieved by the simplicity of the construction
technique. Construction techniques compatible with the use of soil
as a constructional material are required. The use of soil,
deposited in layers to form the structure, results in deformations
within the soil mass caused by gravitational and compaction forces.
These deformations result in the reinforcing elements positioned on
discrete planes moving together and being tensioned as the layers of
soil separating the planes of reinforcement are compressed vertically
and expand laterally. Construction techniques capable of
accommodating this internal consolidation and straining of the fill
are required.
Failure to accommodate the compression, particularly at the face,
may result in loss of serviceability or even rupture of
reinforcements or connections, whilst restriction of lateral
expansion of the fill may prevent the tension in the reinforcements
developing fully.
573
P. M. Jarrett and A. McGown (eds.), The Application of Polymeric Reinforcement in Soil Retaining StructlUes, 573-611.
© 1988 by Kluwer Academic Publishers.
574
(b)
(e)
101
I.)
Facing unit
Temporary wedges
Reinforcement
.Iev.tion concrete
foating
Con'll'lIction sequence - Step 1
Step I: Casl rOOlinl and dninaae ,approximately
I SO JC 300 mm) with top surface lnocl. Erect half' panels. Erect
fuU s.ize panek and rll temporary wed,es to creale horizontal
pp between units. Clamp .djacmt units IOJCthcr and prop
fim rows from front.
clamps to mlintlin
.lign~ent =~f :~~!
I
·IT--~'~__L~______~_.
1Omoo"'" _ \
elev.tion
vertical poles
reinforcement
nep 1
SKlion
1.' elellJlion
, concrete footing
Sr,p J
Cast Coolin! approximately nODx 1SO) mm 'With lOP surface Intel ErClCt ~cnK:.al reinforcement pOles.
[,oct halt uniL Platt porous dr.lina,c pipe.
Erect first full units. PUtt no fines concrele.
step 2
Ibl
section A-A stchon 8-8
COmpaC1101ytr of
hit 'el
SIt/, 1
Note: Rcin(orcancnl poIittoD is al mid·hei,pn of facin, umt
Any fonn or rclnJ'Ota1Dml mlJ be lUCId,
Com~ctioD or the rill dost 10 the rlee is restricted 10 s.m.all plan. so u not 10 dillon the- ficin,
Spt'td of c:onSlrl,IC'tMJn is dC'pC"ndnsl upoa the speed or platLnJ Ibe fill
A bokf raein, is uwally used to dis,utle any incolULutorio Of dsslonlons c.au~ durin, conllNC1ion
Construction can be Itopped.1 .n,.lnrl or position .. ilbout .ny fcar (or the wet,. or Ihe \!Work en
b~
~T"~
~~- -;
1/:--
I ~
\
~ m'n50mm
galvanised
!tHltublng
nut
Detail 'A'
Slag, J:
Cast fooling (appro. 1SO x 300 mm) plus upsland. Elocl and
prop facing panels
concrete footln9
/ Slagt 2:
Construct drainage.
Place fill layers and compacL
When fin level with top of first pair of connecting lugs attach
first level of reinforcement.
It!'mpor.lIy folding wwgt!' Continue filling.
When filling is complete, or when sufficient fill has be.n
fbi stage 1 placed to stabilise facing lemove props and folding ... edges.
------'\
------1
de-tail 'A'
(jf
limit of rein-
M24 bolt forced soil
masonry facework
~
mass concrete
baCking to
masonry
reinforced-concrete
footing
. - ~para~:e~etween
"'~"",,,,--- fiilll and drainage layer
reinforcement laid on
j These layers
only
necessary if
PFA is used
This structure has a maximum height of 6.1 metres and forms part of a
material handling platform supporting 450kN dumptrucks. A cross
section through the wall is shown in Fig.9. The design and
construction of the wall has been described by Berg et al (1986).
2.2.3 Incremental Construction with Discrete Panels.
The most common form of reinforced soil structure is the telescope
method using incremental panels, Fig. 3. A number of proprietary
systems are based upon this technique, including the familiar
Reinforced Earth Company's system which uses a reinforced concrete
cruciform facing unit and the Georgia Department of Transportation
GASE method, which uses steel grid reinforcement. The method has
been used with polymeric reinforcement in the United Kingdom, the
Middle East, and recently in Japan, using Tee shaped and 'Z' shaped
concrete facing panels.
The Soil Structures International Limited 'Websol' Fig.4 A number
of the most sucessful proprietary methods which uses polymeric
reinforcement. This is based on a precast concrete facing and tape
reinforcement, Paraweb, formed from ten lanes of high tenacity
polyester fibres encased in a sheet of low density polythene. The
Paraweb reinforcement, which is connected to the facing panels by
toggles, is laid using an entire roll at a time, typically 100m.
long. This is acheived by lacing the reinforcement between the
facing and a restraining member laid in the fill, Figure 10. In some
structures layers of , Terram' geotextile are laid on top of the
reinforcement, Kempton et al (1985). Recently, an anchored earth
system using paraweb has been introduced in Austria. This system is
now being used in the United Kingdom, Lazlo et al (1995).
Geogrid reinforcement has also been used with the incremental
construction system. Pigg and McCafferty (1985) have described an
application in the UK, whilst Berg et al (1986) have detailed the use
of geogrid reinforcement with the Georgia GASE system. Recently
geogrid has been used with the Japanese 'z' panel, Yamanouchi (1986).
2.2.4 Full Height Panels. The use of full height panels is often
an attractive form of construction for walls of limited height «10
metres). However, since the possibility of articulation by either
the concertina or telescopic methods is lost, consideration must be
given to p~oviding sliding connections, Figure 7.
In the United Kingdom a number of agencies have successfully used
full height panels for highway and industrial projects.
Consolidation of the reinforced fill has been accommodated by use of
sliding connections between the reinforcement and facing panel.
Three polymeric reinforcing materials have been used namely,
"Fibretain" glass reinforced plastic stips, "Tensar" geogrids and
"Paraweb" tape. The system of full height panels was also used for
the construction of the first Anchored Earth structure constructed on
the Otley by Pass in Northern England (Jones et aI, 1985).
In the United States some full height panel wallsconstructed with
rigidly connected steel strip reinforcement have exhibited severe and
unacceptable distortion of the facings on account of there being no
provision for relative movement between the reinforcement and panels
as the fill settles and deforms laterally. Crushing of steel facing
panels has also been observed in Japan on Anchored Earth structures
when no provision has been made for movement of the reinforcing
anchors relative to the facing. Observations on full height
582
U ~~"'M' "
r~ ~~""'r. ~;.~~:::'.. c
L - 3.6m
I
1.3m
Figure 9
-
.
I 11-
II
I
I I I , I
[
II
,
ij
, ,. , ,3~ 11
II
JlJ ~
667661
Figure 10
583
- G,.~;!.{dtail
/ Public footpath
I
POlystyrene
MHWST
--.lL
Membrane
1/ MHWNT
--lL
40
Facing panel
Plan detail
of facing
panel and joints
Geogrids
Figure 12
585
, =
Figure 13
586
Figure 14
FINE
MESH
REINFORCEKENT
SCAFfOLD FPJ,.....,E
AND BOARDS I!I~--==~f----
Figure 15
REINFORCEMENT
VERTICAL TUBE
LEf"T IN PLACE
Figure 16
587
Figure 17
Figure 18
geogrid
~.
turf lining
Figure 19
589
t-----It--...~,
CONCU'7E
(F." .... ,
Figure 20
Figure 21
590
TABLE 1
-- - - ---
TABLE 2
APPLICATIONS
* Secondary Importance
** Important
*** Very Important
593
74;',
,
I
~'
j
!€;==~~= ___________";I f~bric
reinforcement
IT",.m Rf 12)
~~~==~~------------------~,
I
I
random
masonry
facing
concrete 150Ld.
ellSs 30/20 porous drain Figure 22
'" 10 B
panel
closure
(rnn)
""
J 0
"" '00 300
tine
"0 \
Figure 23
(days)'OO
'"
O~o~~-7,o~--,~o~--~~--~~--~
(0) (bl
st.p3 ,"d 1 /~ ..
~-----
stl!p4
soil structures
Figure 26
598
3. ECONOMICS
The primary advantage gained from the use of reinforced soil
structures may be the improved idealisation which the concept
permits; thus, structural forms which would have been impossible to
contemplate become feasible and economic. It is generally accepted
that the use of reinforced soil walls or bridge abutments will
produce significant savings; costs of (30-50 per cent) relative to
conventional construction.
The savings produced through the use of reinforced soil are
influenced by a range of factors, not least of which is the level of
competition available. Without competition the cost of reinforced
soil is heavily influenced by the cost of the conventional structure
alternative, and savings can be reduced in line with what the market
will bear. Without competition from alternative soil reinforcement
systems costs may be 70 to 80 per cent of conventional structures.
Before embarking upon a reinforced soil structure design an
estimate of the possible cost is required, the most reliable of which
are based upon previous work. The base cost of any reinforced soil
structure is very difficult to determine, in some cases it may only
be determined by the use of a form of direct labour contract.
However, the major factors which influence costs for any practice can
be identified, Table 3
Table 3
,. reinforcerrent
. ..
50
50
of of
soil fill total
total
cost cost
,.
,. facing -elenents
,.
.+-----r----+----~,----,,-----
o 10'~ 20
Figure 27
100 , - - - - - - - - - - - ,
econany 7.
100
80
percentage cost 75
of reinforced soil 60
relative to
reinforced concrete
40
cantilever walls
25
20
o+-----~----~----~--~
10 15 20 o 10 15
height of structure (m) height of structure (m)
Figure 28
603
%
Materials Facing 21
Faeing moulds 4
Vertical reinforcements 4
Horizontal rcinforcemcnu 32
Drainage 2
Others 4
67
labour Site clearance I
Retaining wall II
Drainage 5 7
17
PlancJ.and Site clearance I
operatives Retaining wall 10
Drainage 5
16
100
H 3m
~I ~
--------~
2 /
""J 1 --:2J 1
,,-/ 2
~3m
\'olume of earth fill required in reinforced earth and "inforccd concrete retaining ".lIs
4. SPECIFICATIONS
Many civil engineering projects worldwide are undertaken by
contract. The system requires that the client and his appointed
representatives have the means with which to adequately design and
properly specify their needs and that the Contractor is fully aware
of these requirements and their implications at the time of tender.
The Specification describes how aspects of the work shall be
executed, referring when appropriate to approved Codes of Practice
and technical guidelines.
The use of polymers as reinforcing materials for soil structures
raises a number of problems with the development of specifications.
Unlike established materials such as concrete and steel, the
properties and characteristics of polymer materials are less well
understood. In addition, because of their relatively recent
introduction and the range of available polymeric materials, standard
tests to evaluate their properties are only now emerging.
For these reasons, the development of Codes of Practice, design
guidelines and contract documentation, for use with polymeric
materials in reinforced soil structures, have tended to divide into
two categories - those for methods and those for materials, indeed in
some countri~s only the former is presently underway.
S02 -emission
.just - emission
labour - construction
Figure 31
607
Concrete 0.6~ m2 /m 3
Steel 5 m /ton
DUST EMISSION
5. DISCUSSION
The number of reinforced soil structures using polymeric
reinforcement is growing. The rate of growth is likely to increase
as the technical and economic advantages of the technique become more
widely appreciated.
From current usage it is possible to identify areas where
improvements can be made, and which, if introduced would assist both
designers and contractors, and also lead to further economies.
5.2 ECONOMICS
A better understanding of the economics of reinforced soil structures
is required. Advances in this important area would be greatly
assisted by:
a) The pUblication of case studies which detail actual costs,
including background data.
b) Studies of the long term and maintenance costs, including
further consideration of the durability of polymeric materials.
c) Identification of the factors which influence the economics
of reinforced earth soil structures in different countries.
5.4 FUTURE. Likely future developments will corne from two areas.
The first is through _ the introduction of new advanced polymeric
materials, such as the Directionally Structured Fabrics (D.S.F.)
which have recently been announced.
The second area of future development will result from a general
growth of t.he reinforced soil technique. This will produce more
diverse applications. It is already possible to discern the
development of different "styles" in reinforced soil technology based
upon need and frequently influenced by local topology and local
practice. An example of the latter is the development of the
Norwegian Trondblock walling system to incorporate polymeric grid.
610
REFERENCES
1. Berg, R. R., Boneparte R., Anderson R.P, and Chouery V.E, (1968).
"Design Construction and Performance of Two Geogrid Reinforced
Soil Retaining Walls". Proc. 3rd Int. Conf. on Geotextiles,
Vienna.
2. Cantelli R., and Munfakh G., (1986). "Geotextile Walls in
Mountainous Terrain". Proc. 3rd. Int Conf. Geotextiles,
Viennna.
3. Department of Transport (1978). Reinforced Earth Retaining Walls
and Bridges Abutments for Embankments. Tech. Memo. (Bridges)
BE 3/78, London.
4. Douglas,G.E., (1981). "Design and Construction of Fabric-Rein-
forced Retaining Walls by New York State". Trans Research
Record 872.
5. Findlay, ~.W. (1978). "Performance of a Reinforced Earth
Strucutre at Granton". Ground Engineering 2, No.7, 42-44.
6. Forsyth R.A., (1987). "Alternative Earth Reinforcements". ASCE
Spring Convention, Pittsburg.
7. Fukuoka M., (1986). "Fabric Retaining Wall with Multiple
Anchors". Proc. 3rd Int. Conf. Geotextiles, Vienna.
8. Fukuda N, 'Yamanouchi T, and Miura N., (1986). "Comparative
Studies on Design and Construction of Steep Reinforced
Embankment". Proc. 3rd Int. Conf. Geotextiles, Vienna.
9. Godfrey K.A. Jnr. (1984). "Retaining Walls: Competition or
Anarchy" . ASCE Civil Eng. Dec.
10. Jones, C.J.F.P. (1978). "The York Method of Reinforced Earth
Construction". ASCE, Spring Convention, Pittsburg.
11. Jones C.J.F.P. (1985). "Earth Reinforcement and Soil
Structures". Butterworths, England.
12. Jones C.J.F.P. (1987). "Practical Construction techniques for
Retaining Structures using Fabrics and Geogrids". 2nd Int.
Conf. Geotextiles, Las Vegas.
13. Jones C.J.F.P., Murray R.T., Temporal J. and Mair R.J. (1985).
"First Application of Anchored Earth". XI ISSMFE San
Francisco, August.
14. Jones C.J.F.P., Jamison W, and Garner D. (1987). "Design
Construction and Economics of Dewsbury Retaining Walls". In
print.
15. Jones C.J .F.P., and Edwards L.W. (1980). "Reinforced Earth
Structures Situated on Soft Foundations". Geotechnique, June.
16. Kempton G.T, Entwistle, R.W. and Barclay M.J. (1985). "An
anchored fill harbour wall using synthetic fabrics". Proc.
Inst. Civil Eng. Part 1, vol.78 April pp 327-347.
17. Laboratoire Central des Ponts et Chausees (1979). "Les Onulages
en terre armee - Recommendation et regles d'art". Paris.
18. Leary R.M., and Klinedinst G.L., (1984) "Reinforced Wall
Alternatives". FHWA, Washington.
19. Perrier, H., Blivet, J-C, and Khay, M, (1986). "Experimental and
,Actual use of Geotextile Reinforcement of a Slope". 3rd Int.
Conf. Geotextiles, Vienna.
20. Pigg D.Z., and McCafferty W.R., (1985). "The Design and
Construction of Reinforced Soil Retaining Wall at Low
Southwick, Sunderland". Polymer Grid Reinforcement, Thomas
Telford Ltd.
611
21. Saran, S., Talwar D.V., and Prakash, S. (1979). "Earth pressure
distribution on retaining walls with reinforced earth
backfill". C.R. Col. Int. Renforcement des Sols, Paris.
22. Templeman J., and Jones C.J.F.P. (1979). "Soil Structures using
high tensile plastic grids". U.K. patent No. 7941627.
23. Varley W.R., (1979). "Aspects of Alternative Tendering". BCSA
Nat. Struct. Steel Conf. Part 2, London.
24. Vidal, H (1978). "The development and future of Reinforced
Earth". Keynote address. ASCE, Spring Convention, Pittsburg.
25. Vidal, H. (1978). "La terre armee". Annales de L'Institut
Technique du Batiment et des Travaux Publics, 19, Nos. 223-224,
July-August, France.
26. Yamanouchi, T (1986). Private Communication.
REVIEW OF SESSION
Jones opened this session with a general appreciation of the many factors
affecting the economics of geosynthetic construction. Apart from
regional market force aspects, which in many cases control cost and
methodology, he pointed out that working practices may be the dominant
cost factor. The actual cost of the reinforcement is in most cases a
small proportion of the total cost of the structure and as such may have
little influence on overall economics. He then discussed the concept of
using an "Ecology Audit" to establish comparable base costs for different
forms of reinforced soil. An ecology audit sums the overall cost to
society of a particular form of construction especially with respect to
the energy used throughout the process. It therefore avoids to a great
extent market factors and other artificial variables. Such an audit
indicates that geosynthetic reinforcement is a cost effective means of
construction. Despite this economic advantage designers are still hesi-
tant to use this form of construction due primarily to lack of experi-
ence. Therefore the keys to market growth are Specifications and Codes of
Practice for geosynthetic reinforced soil as such standards relieve the
"tension" from the design situation by providing set terms of reference.
Work is progressing in a number of countries on these standards.
The problems of bidding for jobs, where the fill to be used is tightly
specified to allow steel reinforcing to be used, when in fact much
cheaper fills could be safely used with geosynthetics . This problem
ment ioned by Delmas, Bonaparte, McGown and Christopher pointed once
again to the need for Specifications and Codes specifically for
geosynthet ics. Jones pointed out that the maj ori ty of the large
613
P . M. Jarrett and A. McGown (eds.). The Application of Polymeric Reinforcement in Soil Retaining Structures, 613-615.
© 1988 by Kluwer Academic Publishers.
614
number of case studies that he had presented could in fact not have
been constructed using steel reinforcement due to the corrosive nature
of many of the fills.
In summation this session stressed the urgent need for Standard Codes and
Specifications directed at geosynthetic reinforced soils. A major benefit
from such standards will be the allowable use of cheaper, lower quality
fills with geosynthetic reinforcement than are presently allowed with
metallic reinforcement. At present many authorities in the absence of
615
1. INTRODUCTION
This chapter is intended to identify the areas which require further work
and research in order to improve the understanding of geosynthetic -r ein-
forced soil behaviour. From such enhanced comprehension should come the
establishment of safe yet more economic design methods and further
advances of the construction technology for such structures. The material
presented is based primarily on the extensive discussion which took place
in the final session of the workshop and which involved all those partici-
pating. In addition various speakers and authors during the workshop
identified topics requiring further study. These too have been included.
The statement therefore represents to some extent the product of the
research workshop as it records the questions still unanswered and felt by
the participants to be of importance.
(1) There is an urgent need to def ine the environmental working condi-
tions for geosynthetic materials in the field. The following soil
descriptions are needed for a complete range of soils and conditions
likely to arise in backfills:
619
P. M. Jarrett and A . McGown (eds.), The Application of Polymeric Reinforcement in Soil Retaining Structures, 619-625.
© 1988 by Kluwer Arlldemic Publishers.
620
The first two conditions are needed in order to define the parameters
for which the durability of the polymeric materials should be assessed.
The third aspect will set representative temperature ranges to be
considered in design and polymer testing. Of special interest is the
possibility of. high temperatures directly behind facing elements.
(i) Chemical
(ii) Biological
( iii) Ultra violet
(iv) Ageing under stress
(v) Construction damage
(vi) Temperature and humidity
(3) Communications with Polymer and Textile specialists: The range and
complexity of polymeric materials available as reinforcement is continu-
ally increasing. Civil engineers researching the use of these materials
must learn more about their most basic properties and methods of manufac-
ture. There is therefore a serious need to develop communications and
cooperative research with polymer and textile specialists especially with
regard to the durability problem and the development and design of new
products. Please note that this is not the usual pious platitude but a
very real and serious need! Civil engineers must have dialogue with
specialists who can assist them in dealing with such questions as the
effect of chemicals on the behaviour of different polymers, what environ-
mental conditions should be avoided, is biological attack most likely
621
(4) Testing composite materials can mean testing samples of soils con-
taining polymers with each sample considered as a unit. The results
obtained are not divided between the constituent materials but refer to
the unit as a whole. Such an approach was used in the early years of the
development of soil reinforcement, and although many researchers have
abandoned this technique in favour of testing the constituent materials
separately, the composite testing approach may still offer advantages in
certain situations. Recently there has been a revival of this type of
testing due to the introduction of new methods of soil reinforcement using
randomly oriented polymers as in Texsol and the mesh element technique.
It seems that a large scale triaxial apparatus is most suitable for the
determination of the properties of the composites. However, other types
of testing &hou1d be explored and the parameters relevant to design should
be identified.
3. TESTING SYSTEMS
(i) Stresses and strains in the soil both horizontally and ver-
tically
( ii) Stresses and strains in the reinforcement
( iii) Forces at the reinforcement to facing connection
(iv) Vertical earth pressures to confirm any variation from uni-
form overburden pressure
(v) Movements of the wall facing
(vi) Temperatures throughout the , backfill
4. DESIGN
(5) Distinction between the design of a safe structure and the prediction
of its behaviour in terms of deformations and stresses.
(5) Stiffness of the reinforced soil mass, of the facing elements and of
the foundation soil.
5. CONSTRUCTION
PARTICIPANTS
LIST OF PARTICIPANTS
Durability, 19, 32, 33, 57-60, 346, Internal stability, 317-319, 433,
354, 593, 620 508-530
ageing, 308, 309, 620
biological, 620 Joints, 108, 355, 444, 574-595, 613
chemical, 309, 620
construction damage, 173-180, Ladder walls, ~-6
292, 309, 535, 620
ultra-violet, 58, 620 Limit states, 314-316, 326-328,
381, 382, 427
Dynamic behaviour, 620, 622
Paraweb, 38, 581
Earth pressures
Meyerhof distribution, 135, 146, Participants, 628-632
273-275, 423
vertical, 106, 143, 146, 162, Prediction symposium
167, 186-187, 227, 273, 324, class A predictions, 130-157
330, 475 comparison to measurements,
157-167
Ecology audit, 603, 605-607 discussion, 168-170
RMC tests, 71-125
.
Economics, 268-270, 599-605, 609
Properties
Extensibility of reinforcement, 31, geogrids, 118-122, 173-180, 203-
42, 286, 287, 409-453, 507- 205, 401, 447
508 geotextiles, 45-57
polymer, 41-45, 51-52, 620-621
External stability, 315-317, 443, soil, 392-400, 411-415, 619-620
460-461, 530-534 testing, 346-355, 619-621
Facing movements, 88, 89, 96-98, Pullout tests, 16, 18, 268, 355,
141, 156-158, 164, 229, 230, 359-361, 621
290, 294, 320, 330-331, 382-
391, 593-595 Reinforced earth, 3, 7-12, 32, 285-
287, 429, 581
Facings, 267, 418, 625
armateer, 35 Reinforced soil, 3
concrete, 12, 35, 38, 426, 581
contained facings, 587 Reinforcement types, 30-33, 266
fabric, 33, 587 fiberglass, 5, 32, 581
fiberglass, 36 fibers, 19-24, 38-41, 340
637