Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Downloaded from ascelibrary.org by MALAVIYA NATIONAL INSTITUTE OF TECHNOLOGY on 06/22/14. Copyright ASCE.

For personal use only; all rights reserved.

BEARING CAPACITY OF RECTANGULAR FOOTINGS ON


GEOGRID-REINFORCED SAND

By Temel Yetimoglu, ~ Jonathan T. H. Wu, z Member, ASCE,


and Ahmet Saglamer3

ABSTRACT: A study was undertaken to investigate the bearing capacity of rec-


tangular footings on geogrid-reinforced sand by performing laboratory model tests
as well as finite-element analyses. The effects of the depth to the first layer of
reinforcement, vertical spacing of reinforcement layers, number of reinforcement
layers, and the size of reinforcement sheet on the bearing capacity were investi-
gated. Both the experimental and analytical studies indicated that there was an
optimum reinfgrcement embedment depth at which the bearing capacity was the
highest when single-layer reinforcement was used. Also, there appeared to be an
optimum reinforcement spacing for multi-layer reinforced sand. The bearing ca-
pacity of reinforced sand was also found to increase with reinforcement layer num-
ber and reinforcement size when the reinforcement was placed within a certain
effective zone. In addition, the analysis indicated that increasing reinforcement
stiffnessbeyond a certain value would not bring about further increase in the bearing
capacity.

INTRODUCTION

In comparison with other applications of geosynthetic-reinforced soil, for


example, geosynthetic-reinforced soil e m b a n k m e n t s or retaining walls, rel-
atively less emphasis has been placed on reinforced soil beds (i.e., use of
geosynthetics for reinforcing soil foundations). Binquet and Lee (1975a, b)
were among the first to report a systematic study on bearing capacity of
reinforced soil beds. Since then, a n u m b e r of analytical and experimental
studies on the subject have been conducted by several researchers [e.g.,
Bassett and Last (1978); A k i n m u s u r u and A k i n b o l a d e (1981); G u i d o et al.
(1985, 1987); Sridharan et al. (1989); Samtani and S0npal (1989); H u a n g
and Tatsuoka (1990); A b l e d - B a k i et al. (1993); Khing et al. (1993); O m a r
et al. (1993a, b)]. H o w e v e r , in most of the experiments, strip footings were
used, despite the fact that rectangular and square footings are far m o r e
common in practice. A l s o , geotextiles, metals (strips, sheets, or bars), or
fibers (both natural and m a n - m a d e ) were used as reinforcement in most of
these studies. G u i d o et al. (1986) conducted l a b o r a t o r y m o d e l tests on
reinforced soil foundations and c o m p a r e d the p e r f o r m a n c e of geotextile and
geogrid as soil reinforcement. This study showed that the geogrid reinforce-
ment was more effective than the geotextile from the standpoint of im-
proving the bearing capacity of footings on reinforced sand. Jewell et al.
(1984) and Milligan and Palmeira (1987) pointed out that the mechanism
for mobilization of frictional resistance in geogrids was different from that

tRes. Asst., Geotech. Engrg. Dept., Civ. Engrg. Fac., Istanbul Tech. Univ.,
Ayazaga, Istanbul-Turkey; formerly, Visiting Scholar, Univ. of Colorado at Denver~
Denver, CO 80217-3364.
2Prof., Dept. of Civ. Engrg., Univ. of Colorado at Denver, Denver, CO.
3prof., Geotech. Engrg. Dept., Civ. Engrg. Fac., Istanbul Tech. Univ., Ayazaga,
Istanbul-Turkey.
Note. Discussion open until May 1, 1995. To extend the closing date one month,
a written request must be filed with the ASCE Manager of Journals. The manuscript
for this paper was submitted for review and possible publication on October 27,
1993. This paper is part of the Journal of Geotechnical Engineering, Vol. 120, No.
12, December, 1994. 9 ISSN 0733-9410/94/0012-2083/$2.00 + $.25 per page.
Paper No. 7206.

2083
J. Geotech. Engrg. 1994.120:2083-2099.
Downloaded from ascelibrary.org by MALAVIYA NATIONAL INSTITUTE OF TECHNOLOGY on 06/22/14. Copyright ASCE. For personal use only; all rights reserved.

of geotextiles. In geogrid-reinforced soils, the transverse members of the


geogrid together with the longitudinal ribs frictionally interact with the soil,
mobilizing interfacial shear strength opposing the lateral flow. The mode
of frictional interaction is not just the interfacial one but also due to the
passive resistance mobilized by the bearing of soil particles against the lateral
(transverse) elements. They indicated that geogrids generally offer a higher
interfacial shearing resistance than geotextiles.
The present study was undertaken to investigate the bearing capacity of
rectangular footings on geogrid-reinforced sand. Both experimental and
analytical studies were conducted. The parameters investigated include (1)
Depth to the first layer of reinforcement u; (2) vertical spacing of reinforce-
ment layers z; (3) number of reinforcement layers N; (4) size of reinforce-
ment sheet (Br: width of reinforcement sheet in the tests, or Dr: equivalent
diameter of reinforcement sheet in the analyses); and (5) stiffness of rein-
forcement. The symbols of the geometric parameters used in this present
paper are shown in Fig. 1. The experimental study involved performing
more than 100 laboratory load tests. Some of the tests were repeated as
many as four times to assure the repeatability of the results. The analytical
study was conducted by the finite-element method. A finite-element model
that has been verified by a number of researchers for various soil-structure-
interaction problems was used. The analytical results and experimental re-
suits are compared and discussed.

E X P E R I M E N T A L TESTS A N D T E S T P R O C E D U R E

A series of model loading tests were conducted inside a cubical steel tank
of 70 cm by 70 cm in plane and 100 cm in depth. The tank was strengthened
by a number of channel-shaped steel beams in both vertical and horizontal
directions to restrain lateral expansion under loads. Static vertical loads

B orD
9

J t
tl
N=I
Z
2
Z
3

L B r or D r
FIG. 1. Geometric Parameters of Reinforced Foundation

2084
J. Geotech. Engrg. 1994.120:2083-2099.
Downloaded from ascelibrary.org by MALAVIYA NATIONAL INSTITUTE OF TECHNOLOGY on 06/22/14. Copyright ASCE. For personal use only; all rights reserved.

were applied using a 100-kN capacity electrical hydraulic pump. Loads trans-
ferred from the pump to a hydraulic jack were carefully recorded by a
proving ring installed between the jack and the test footing. The setup of
the tests is depicted in Fig. 2.
The footing was loaded at a constant loading rate until an ultimate bearing
state was reached. The ultimate bearing state was defined as the state at
which either the load reached a maximum value where settlements continued
without further increase in loads or where there was an abrupt change in
the load-settlement relationship. Settlements of the footing were measured
using two dial gauges situated in diagonal directions. The settlements re-
ported were the average of the two dial-gauge readings, which were nearly
identical until the ultimate bearing state was reached. II~is to be noted that
the tests were conducted under stress-controlled conditions; thus, postfailure
behaviour was not recorded.
Before starting a new test, the sand in the tank (from the previous test)
was removed to a depth of about three times the footing width, and deeper
for multilayer reinforcement tests. Then, the sand was emplaced in lifts.
The lift thickness was dependent on the reinforcement spacing; larger lift
thickness was used for larger spacing. The maximum lift thickness, however,
was 60 ram. For each lift, the sand was compacted by a KANGO vibration

Reaction
Reactionbeam--~ ~ column

I I t tU ~
I
== / ~-~Hydraulic
Oil trans~r p i p e ~ Proving V jack
ring- - ~

Dial
Loading gauge ~ Footmg
rate

On/off adjusting
valve~ valve~ )

E
o

_ Steelplate
II (t = 2.5 mm)
r Sand
Electrical
hydraulic
pump

;I .IL-
i

n n

B = 7O
n

J
- Steel
channel
section

FIG. 2. Bearing Capacity Test Setup

2085

J. Geotech. Engrg. 1994.120:2083-2099.


Downloaded from ascelibrary.org by MALAVIYA NATIONAL INSTITUTE OF TECHNOLOGY on 06/22/14. Copyright ASCE. For personal use only; all rights reserved.

hammer to the prescribed density. To prevent crushing of soil particles and


to provide a more uniform placement density, the vibration was applied
through a wooden plate that had the same area as the tank. On reaching a
prescribed height in the test tank, the top surface was leveled and the footing
was placed on a predefined alignment such that the loads from the jack and
loading frame would be transferred concentrically to the footing.
The test footing was a rectangular steel plate 12.5-mm thick, 101.5-mm
wide and 127-mm long. The bottom surface of the footing was smooth.
Because the width and depth of the test tank are greater than six times
the footing width, the boundary effect on the test results was considered
small. During the tests, it was observed that the extent of soil bulging on
the sides of the footing was less than the footing width, indicating the
boundary effect on the tests was likely to be insignificant.

Soil Properties
A clean, oven-dried, uniform quartz river sand (Yahk6y sand) was used
in the tests. The sand was placed in the test tank at a unit weight of ap-
proximately 17.16 kN/m 3 (a relative density of Dr = 7 0 - 7 3 % ) . Approxi-
mately 10 kN of the sand was seived through BS8 (d = 2 mm) and washed
through BS200 (d = 0.076 mm) and dried in an oven. Some properties of
the sand are given in Table 1.
Three triaxial CD tests were performed on Yallk6y sand samples prepared
at the same density as that of the model loading tests. The results are shown
in Fig. 3.

Reinforcement Properties
The reinforcement used in the tests was a uniaxial geogrid, Terragrid
GS1000 (produced in Turkey). The basic raw material of the geogrid is
polypropylene (per A S T M D-4101). Typical geometric dimensions of the
geogrid are: rib thickness = 0.95 mm, bar thickness = 2.4 mm, rib width
= 5 mm, bar width = 10 mm, aperture length = 80 mm, aperture width

TABLE 1. Properties of Sand


Property Value
(1) (2)
Specific gravity 2.65
Maximum dry density (kN/m3) 17.72
Minimum dry density (kN/m3) 15.87
Maximum void ratio 0.670
Minimum void ratio 0.495
Coarse sand fraction (percent) 3
Medium sand fraction (percent) 47
Fine sand fraction (percent) 5O
Effective grain size Dl0 (mm) 0.15
D6o (mm) 0.35
D3o (mm) 0.20
Coefficient of uniformity C, 2.33
Coefficient of curvature Cc 0:76
c (kPa) ~ 0
d~(degree) a 40
aObtained from direct shear and triaxial compression tests at Dr = 75%.

2086
J. Geotech. Engrg. 1994.120:2083-2099.
Downloaded from ascelibrary.org by MALAVIYA NATIONAL INSTITUTE OF TECHNOLOGY on 06/22/14. Copyright ASCE. For personal use only; all rights reserved.

(a) aj=415kPa

9N
~ 700
I 17 4 = ] 4 0 kPa

3 5 0 ~

0 ~
0 2 4 6 8 10 12 14 16 18 20
Axial Strain, %

(b) ~"
E

>

-1 i i i i i i i i J

0 2 4 6 8 10 12 14 16 18 20
Axial SWain, %
FIG. 3. Triaxial Compression (CO) Test Results for Yalik6y Sand: (a) Axial Strain-
Deviatoric Stress Relationship; and (b) Axial Strain-Volumetric Strain Relationship

= 14 mm. Some of its index p r o p e r t i e s p r o v i d e d by the manufacturer are:


weight = 5 -+ 0.1 N/m 2, roll width = 1 m, tensile strength (per A S T M
D-4595) = 28.6 kN/m, elongation at yield (per A S T M D-4595) = 12%.
FINITE-ELEMENT ANALYSES
The finite-element analyses were conducted following the conclusions of
the experimental tests. A finite-element c o m p u t e r p r o g r a m , D A C S A R (de-

2087
J. Geotech. Engrg. 1994.120:2083-2099.
Downloaded from ascelibrary.org by MALAVIYA NATIONAL INSTITUTE OF TECHNOLOGY on 06/22/14. Copyright ASCE. For personal use only; all rights reserved.

formation analysis considering stress anisotropy and reorientation), was


used. The finite-element model was originally developed by Iizuka and Ohta
(1987) at the University of Kyoto in Japan. The computer program has been
validated through numerous laboratory element tests (Iizuka and Ohta 1987;
Chou -1992; Chou and Wu 1993; Helwany 1993), model tests (Iizuka and
Ohta 1987; Fukagawa et al. 1990; Helwany 1993), full-scale loading tests
(Chou 1992), and field tests (Iizuka and Ohta 1987; Ohta and Iizuka 1988;
Ohta et al. 1991; Ohta 1991). A detailed description of DACSAR can be
found in Ohta and Iizuka (1986) and Iizuka and Ohta (1987).
The finite-element mesh used for the analyses is shown in Fig. 4. The
geogrid reinforcement and the footing were represented by a series of dis-

L ~D/2 = 6.5cm
C

m~ I
inlln l Ill Ill llllillilll
i Illilln Ill nlgiliHilmli
lnlllllllllllillll
lllllllnllllllllll
llllllllllllnlllll i
llllnlllllllllllll I
lllllllllllllllllil
IIIIIIIIIIIIIlilinl
~ri+L~rirLr~iir~
Illlllmiilmi
Iliniilili
ililiilill
Ililillil me
lillillil in I
~r~)mmmi
m /

H = 85 cm

~r

I
R/2 = 40 cm 'i

FIG. 4. Finite-Element M e s h

2088
J. Geotech. Engrg. 1994.120:2083-2099.
Downloaded from ascelibrary.org by MALAVIYA NATIONAL INSTITUTE OF TECHNOLOGY on 06/22/14. Copyright ASCE. For personal use only; all rights reserved.

crete axi-symmetric shell elements, and the sand was represented by an


assembly of axi-symmetric quadrilateral and triangular elements. In this
study, the analyses were conducted under axi-symmetric conditions with the
soil, reinforcement, loading, and boundary conditions mimicking those of
the experimental tests. In the analyses, the rectangular footing (B = 101.5
mm, L = 127 mm) was treated as an equivalent circular plate (D = 130
ram) of the same footing area.
Vertical loads were applied sequentially in equal increments of 25 kPa.
Each load increment was divided into five steps to achieve better compu-
tational accuracy.
Because the settlement ratios were also small at failure for both unrein-
forced sand and reinforced sands in the tests (i.e., s / B < 5% at failure),
the strains developed in the geogrid reinforcement were likely to be very
small as well. Hence, a constant modulus of elasticity E = 4.9 x 105 kPa
was used for the geogrid in the analyses. The value of E was a secant modulus
determined per ASTM D-4595 at an axial strain of 1%. No offset allowance
was made in the calculation of the modulus. The nominal thickness of the
geogrid was 0.95 mm. The footing was also simulated as a linear elastic
material, since the operative strains were very small. For the footing, the
modulus of elasticity was 2.1 x 108 kPa and the thickness was 12.5 mm.
The Poisson's ratio was 0.3 for both the geogrid and the footing. The stress-
strain-strength behavior of the sand was simulated by the modified Duncan
hyperbolic model (Duncan et al. 1980). The model is capable of simulating
nonlinear, inelastic, stress-dependent material behavior, wherein the prop-
erties of each soil element are calculated in accordance with the minor
principal stress and shear-stress level.
A detailed description of the soil model has been presented by Duncan
et al. (1980). Only a brief description of its soil parameters are given here.
A total of seven parameters (i.e., K, K,r, n, c, +0, A+, and R~) are required
to define the stiffness and strength characteristics, and two additional pa-
rameters (i.e., Kb and m) are required to define the volume-change char-
acteristics. The angle of internal friction +, the stress-dependent tangent
Young's modulus E,, the unloading-reloading modulus E,r, and the bulk
modulus B are expressed in terms of these parameters as
+ = ~bo - A~b log(0"3/Pa) (1)
E, = {1 - [R~(0"1 - e3)/(0"~ - 0"3)r]}eKp~(0"3/P.)" (2)
gur =- gurPa(0"3/Pa) n (3)
B = KvP~(0"JL) m (4)

where 0"3 = minor principal stress; (O" 1 - - 0"3) f = deviatoric stress at failure;
and Pa = a reference pressure usually chosen as the atmospheric pressure.
The values of the modified Duncan hyperbolic soil parameters deduced
from the results of the CD triaxial tests (as shown in Fig. 3) are listed in
Table 2. No attempt was made to adjust the parameters.
It is to be noted that the calculated vertical strains in the soil elements
even when approaching failure were smaller than the strain level at which
shear dilation occurred in the triaxial tests carried out on the sand. There-
fore, the use of the modified Duncan soil model is considered acceptable
in the present study.
2089
J. Geotech. Engrg. 1994.120:2083-2099.
Downloaded from ascelibrary.org by MALAVIYA NATIONAL INSTITUTE OF TECHNOLOGY on 06/22/14. Copyright ASCE. For personal use only; all rights reserved.

TABLE 2. Parameters of Modified Duncan Soil Model for Yahk0y Sand

Parameter Value
(1) (2)
c (kPa) 0.0
~o (degree) 40.2
~+ (degree) 0.5
Rr 0.88
K 672
n 0.57
Kur 806
Kb 817
m 0.35

~ .I.--a- a-. _ _ . _ , a _ _ A _ _ , A

E
E 10
9 Test

E A Analysis
20
I
I

30 i i i i ! I

0 50 100 150 200 250 300 350

Average Contact Pressure, kPa


FIG. 5. Load-Settlement Curves of Unreinforced Sand

RESULTS

Only selected results of the experimental and analytical studies are pre-
sented here. A complete set of the results can be found in Yetimoglu (1993).
To compare load-settlement behavior of reinforced and unreinforced sands,
a dimensionless term B C R (bearing capacity ratio) is defined as (Binquet
and Lee 1975a)
BCR = q/qo (5)
where q0 = average contact pressure of a footing on an unreinforced sand
at a given settlement; and q = average contact pressure of the footing on
a reinforced sand at the same settlement.

Unreinforced Sand
The load-settlement curves o b t a i n e d from the experimental test and finite-
element analysis for the unreinforced sand are shown in Fig. 5. The ultimate
bearing capacity for the unreinforced sand obtained from the test (quit =
316 kPa) and the analysis (quit = 311 kPa) are in excellent agreement.
However, the settlements o b t a i n e d from the analysis at different contact
pressures were much higher than those from the test. The settlement ratio
(settlement/footing width or settlement/footing diameter) for the unrein-
forced sand at failure was a p p r o x i m a t e l y 3% in the test and 16% in the
2090
J. Geotech. Engrg. 1994.120:2083-2099.
Downloaded from ascelibrary.org by MALAVIYA NATIONAL INSTITUTE OF TECHNOLOGY on 06/22/14. Copyright ASCE. For personal use only; all rights reserved.

analysis. This was attributed to the fact that compaction-induced stress was
not accounted for in the analysis.
Lambrechts and Leonards (1978) indicated that the irreversible strain due
to prestressing could increase the reload modulus by more than an order
of magnitude over that of the normally consolidated sand, even though no
residual stresses due to prestressing are extant. The presence of residual
lateral stress can further increase the soil stiffness by an additional factor
of two. Yoshimi et al. (1975), in their study of one-dimensional compress-
ibility of sands under very low confining stresses, also showed that a pre-
stressed sand exhibited more than six times higher soil stiffness than the
normally consolidated sample. Concerning the effect of prestressing on the
ultimate strength, Lade and Duncan (1976), and Lambrechts and Leonards
(1978) indicated that prior stress history, which may influence stress-strain
behavior of a sand very significantly, had no discernible effect on the shear
strength.

Reinforced Sand
In the following, the effects of varying reinforcement configurations on
the BG,R are presented. The findings of the experimental tests are compared
with those of other studies. The results obtained from the experimental tests
and the finite-element analyses are also compared. In addition, the effect
of the reinforcement stiffness examined by finite-element analyses is pre-
sented.
It should be noted that the experimental tests indicated only somewhat
higher settlements at failure for reinforced sands than for unreinforced sand;
however, the ultimate bearing capacity of reinforced sand could be four
times as high as that of unreinforced sand. Table 3 shows a summary of
some measured data for three test groups with different reinforcement con-
figurations. It is seen that the ultimate bearing capacity is mobilized at a
settlement of approximately 3 - 5 % of the footing width (i.e., at settlement
ratio = 3 - 5 % ) for all the unreinforced and reinforced sands, while the
BCR varies from about 1.8 to 3.9.
The model tests conducted by O m a r et al. (1993a, b) and Khing et al.
(1993) also indicated that the settlements at failure for geogrid-reinforced

TABLE 3. Summary of Some Test Results for Reinforced and Unreinforced Sand
Group 1 Group 2 Group 3
u/B = 0.30 u/B = 0.45 u/B = 0.30
zlB = 0.30 z/B = 0.30 z/B = 0.30
B,/B = 4.5 Br/B = 4.5 Br/B = 6.0
s/Ba qu, s/Ba qu, s/Ba qul~
N (percent) (kPa) BCR ~ (percent) (kPa) BCR b (percent) (kPa) BCR b
(1) (2) (3) (4) (5) (6) (7) (8) (9) (10)
0 2.7 316 2.7 316 -- 2.7 316 --
1 3.4 586 1.85 3.1 558 1.77 3.8 579 1.83
2 4.8 790 2.50 3.1 718 2.27 3.4 795 2.52
3 4.8 1002 3.17 3.0 768 2.43 4.9 1081 3.42
4 3.9 1147 3.63 2.8 766 2.42 4.4 1225 3.88
"Ratio of the settlement at failure to the width of footing.
bRatio of the ultimate bearing capacity of reinforced sand to the ultimate bearing
capacity of unreinforced sand.

2091
J. Geotech. Engrg. 1994.120:2083-2099.
Downloaded from ascelibrary.org by MALAVIYA NATIONAL INSTITUTE OF TECHNOLOGY on 06/22/14. Copyright ASCE. For personal use only; all rights reserved.

sands were higher than for unreinforced sand, although the difference in
the ultimate load was more significant. However, Khing et al. (1993) pointed
out that the settlements of strip foundations at failure for geogrid-reinforced
sands and unreinforced sands were comparable if the depth to the first
reinforcement layer was greater than the footing width (i.e., if u > B).

Effect of Depth to First Reinforcement Level


Both the experimental and analytical studies indicated that the effect of
depth ratio on the BCR in single-layer reinforced sands was different from
that in multilayer reinforced sands. The depth ratio is defined herein as the
ratio between u and B (or between u and D).
In single-layer reinforced sand, both the tests and analyses indicated that
there was an optimum value of depth ratio at which the BCR was the highest.
This optimum depth ratio is around 0.3 and appears to be independent of
reinforcement size. The analyses, however, indicated that if the settlement
ratio was higher than 6%, the optimum depth ratio would be somewhat
higher.'Fig. 6 shows a typical relationship between the BCR and the depth
ratio for a single-layer reinforced sand. It is seen that the BCR values
obtained from the tests and analyses are nearly identical for depth ratios
between 0.15 and 0.5. For depth ratios greater than 0.5, the BCR values
obtained from the tests are somewhat smaller than those obtained from the
analyses. It is noted that for depth ratios greater than 1.0, the BCR values
approach a constant.
Both the tests and analyses for the multilayer reinforced sands indicated
that the largest BCR values occurred at a depth ratio of around 0.25. A
typical variation of BCR with depth ratio for a multilayer reinforced sand
is given in Fig. 7. It can be seen that although the BCR value tends to
decrease with increasing depth ratio, the change in BCR is not significant
for a depth ratio less than 0.3. The BCR obtained from the analysis and
the tests are identical up to a depth ratio of 0.3, beyond which the tests
indicate somewhat smaller BCR values than the analyses. Both the tests
and analyses indicated that the change in the BCR became much smaller
when the depth ratio was greater than about 0.9.
It should be pointed out that there has not been a general consensus
regarding the effect of depth ratio on the bearing capacity of footing. Singh

2.5

6
..~
2.0
A & Test
.~- 1.5 S
zx Analysis
e~

~ 1.0
.r-

m 0.5 I I I I

o.o 0.3 0.6 0.9 1.2 1.5


Depth Ratio, (u/Bor u/D)
FIG. 6. Variation of BCR with Depth Ratio in Single-Layer Reinforced Sand (s/B
= s/D = 2%, Br/B = Dr/D = 4.5)
2092
J. Geotech. Engrg. 1994.120:2083-2099.
Downloaded from ascelibrary.org by MALAVIYA NATIONAL INSTITUTE OF TECHNOLOGY on 06/22/14. Copyright ASCE. For personal use only; all rights reserved.

~" 4.0
L)

6 3.0
t~
9 Test
.'=
r
2.0 Analysis
t~

~0 1.0
t.
Q.)
I I I I
o.o
o.o 00 0.6 0.9 1.2 1.5
Depth Ratio, (u/B or u/D)

FIG. 7. Variation of BCR with Depth Ratio in Multi-Layer Reinforced Sand (s/B =
s/D = 2%, N = 4, z / B = z/D = 0.30, B r/B = DJD = 4.5)

(1988), based on his study of square footings on sands reinforced with mild
steel grids (also called "welded mesh"), indicated that the effect of depth
ratio on the BCR was independent of the number of reinforcement layers.
He indicated that the optimum depth ratio was about 0.25 for both single-
layer and multilayer reinforced sands. Akinmusuru and Akinbolade (1981)
reported that the BCR was the highest at a depth ratio Of approximately
0.5 for square footings on multilayer reinforced sands. In their study, fiat
strips of the rope-fiber material in a grid form was used as reinforcement.
Guido et al. (1986, 1987), using Tensar SS1, SS2, and SS3 geogrids as
reinforcement, reported that BCR decreased with increasing depth ratio for
square footings on multilayer reinforced sands.
The finding of the present study on the effect of depth ratio is similar to
that reported by Singh (1988). The disagreement with the results reported
by Akinmusuru and Akinbolade (1981) and Guido et al. (1986, 1987) is
most likely due to the difference in the material properties and the geometric
dimensions of the reinforcement. Jewell et al. (1984) and Milligan and
Palmeira (1987) showed that the size, shape, and spacing of geogrid bearing
members together with the size of soil particles had a significant effect on
geogrid-soil interaction. They indicated that the ratio of minimum grid ap-
erture dimension (dmi,) to average soil particle size (Ds0) had a significant
effect on the soil-reinforcement direct sliding and bond (pullout) resistance.
The ratio was approximately 50 (14 mm/0.30 mm) in the present study,
while in Singh's study, it was approximately equal to 45 (25.4 mm/0.55 mm).
On the other hand, in the tests conducted by Guido et al., the ratio of drain/
Ds0 was around 190 (28 mm/0.15 mm) for Tensar SS1 and SS2 geogrids,
and 300 (46 mm/0.15 mm) for the Tensar SS3 geogrid; in the tests performed
by Akinmusuru and Akinbolade (1981), the ratio of dmin/Dso varied between
115 (50 mm/0.43 mm) and 350 (150/0.43 mm).
One would expect a geogrid-reinforced sand to exhibit a general shear
failure mode, since the foundation soil would be more rigid than unrein-
forced soil. However, for u / B less than about 0.25, the settlement pattern
resembles that of a typical punching-shear failure. In these tests, small
surface heaving was observed at the edges of the footing. On removal of
the footing from the test tank, the geogrid reinforcements directly below the
2093
J. Geotech. Engrg. 1994.120:2083-2099.
Downloaded from ascelibrary.org by MALAVIYA NATIONAL INSTITUTE OF TECHNOLOGY on 06/22/14. Copyright ASCE. For personal use only; all rights reserved.

footing were visibly deflected downward in a shape conforming to the size


of the footing, suggesting that punching failure had occurred.
Akinmusuru and Akinbolade (1981) and Guido et al. (1987) also reported
the geogrid-reinforced sand exhibited a failure mode that resembles typical
punching-shear failure. Binquet and Lee (1975a), in their study of strip
footings on sand reinforced with aluminum strips, also indicated that failure
was most likely in a punching-shear mode.

Effect o f Vertical Spacing o f Reinforcement Layers


Both the experimental and analytical studies indicated that there was an
optimum value or an optimum range of values for the vertical spacing of
horizontally placed reinforcement layers. The tests conducted here showed
that the optimum vertical spacing was around 0.2B. On the other hand, the
analyses did not show a clearly defined optimum value for vertical spacing.
The maximum BCR values occurred between z/D = 0.2 and z/D = 0.4,
depending on the number of reinforcement layers. Fig. 8 shows a typical
variation of BCR with the normalized vertical spacing of a four-layer rein-
forced sand. Both the tests and analyses indicate the optimum vertical spac-
ing is around 0.2B (or 0.2D). Note that although the BCR values obtained
from the analyses are somewhat different from those obtained from the
tests, their maximum BCR values, corresponding to the optimum vertical
spacing, differ only slightly (3.4 versus 3.0).
Similar to these findings, Singh (1988) indicated that the optimum vertical
spacing of reinforcement layers for square footings on geogrid-reinforced
sands varied between 0.15B and 0.25B, depending on the size of the rein-
forcement. On the other hand, Guido et al. (1986) and Akinmusuru and
Akinbolade (1981) reported that as the vertical spacing increased, the BCR
value would decrease accordingly. Guido et al. (1987) indicated that al-
though the BCR value decreased with increasing vertical spacing for Tensar
SS1, SS2, and SS3 geogrids, the trend of BCR variation with vertical spacing
was different for Tensar SS3 and for Tensar SS1 and SS2 geogrids.
The disagreement with the findings reported by Akinmusuru and Akin-
bolade (1981) and Guido et al. (1986, 1987) is attributed to the difference
in the material properties and the geometric dimensions of the geogrid used
in the present study. As explained earlier, the ratio of dmJDso in their

g 4.0

6 3.0
.r-'~

r162
9 Test
.~
o
2.0
zx Analysis
L~
e~ 1.0
't::
I I I I
0.0
0.0 0.3 0.6 0.9 1.2 1.5
Normalized Vertical Spacing of Reinforcement Layers, (z/B or z/D)
FIG. 8. Typical Variation of BCR with Vertical Spacing of Reinforcement Layers
(s/B = slD = 2%, N = 4, u/B = u/D = 0.30, Br/B = D,/D = 4.5)

2094
J. Geotech. Engrg. 1994.120:2083-2099.
Downloaded from ascelibrary.org by MALAVIYA NATIONAL INSTITUTE OF TECHNOLOGY on 06/22/14. Copyright ASCE. For personal use only; all rights reserved.

studies was quite different from that used here. Also, the arrangement of
geogrid reinforcements (i.e., depth ratio, reinforcement size, number of
reinforcement layers) was not the same.

Effect o f N u m b e r o f Reinforcement Layers


Both the experimental and analytical studies indicated that the BCR value
changed more drastically with the number of the reinforcement layers N
than the other parameters. The BCR increased with increasing number of
the reinforcement layers within a depth of 3B (or 3D) below the footing
base. The rate of increase in BCR was less significant beyond a depth of
1.5B (or 1.5D); in other words, placing geogrid reinforcement beyond the
depth of 1.5B (or 1.5D) would not significantly increase the bearing capacity.
Fig. 9 shows a typical variation of the BCR with the number of rein-
forcement layers. For the results shown, the vertical spacing was a constant,
thus a larger number of layers involved a deeper zone of reinforced sand.
It is seen that the BCR values obtained from the tests and analyses are in
very good agreement for N = 1-4. Both the tests and analysis results
indicated that the BCR increased proportionally with the number of rein-
forcement layers up to N = 4 (corresponding to a reinforcement depth of
approximately 1.5B or 1.5D). Beyond N = 4, the tests indicated that there
was a slight decrease in the BCR, and the analyses indicated a continuous
increase in the BCR although the rate of increase became smaller.
Several researchers have also observed that increasing the number of
reinforcement layers beyond a certain value would not increase the BCR
significantly. Akinmusuru and Akinbolade (1981), Guido et al. (1986, 1987),
Singh (1988), and Omar et al. (1993a, b) reported that the BCR was gen-
erally proportional to the number of reinforcement layers. Guido et al.
(1986, 1987), using reinforcement with a different depth ratio and vertical
spacing from those used in the present study, indicated that the rate of
increase with layer number would decrease beyond three layers.

Effect o f Reinforcement Size


Both the analyses and tests showed that the BCR generally increased
slightly with increasing reinforcement size (B r or Dr). The increase was more

4.0

6
.r~
3.0

9 Test
.~- 2.0 tx Analysis

1.0
"t:::
I I | I I I I
0.0
0 1 2 3 4 5 6 7 8
Number of Reinforcement Layers, (N)
FIG. 9. Variation of BCR with Number of Reinforcement Layers (s/B = s / D = 2%,
u/B = u/D = z/B = z/D -- 0.30, Br/B = D,/D = 4.5)

2095
J. Geotech. Engrg. 1994.120:2083-2099.
Downloaded from ascelibrary.org by MALAVIYA NATIONAL INSTITUTE OF TECHNOLOGY on 06/22/14. Copyright ASCE. For personal use only; all rights reserved.

4.0

.r~
d 3.o
e,,
A" 9 Test
9~- 2.0 zx Analysis
e~

~ 1.0
e-

I I I I I I
c~ o.o
o 1 2 3 4 5 6
Size Ratio of Reinforcement Sheet, (B,/B or D,/D)

FIG. 10. Variation of BCR with Reinforcement Size (s/B = s/D = 2%, N = 4, u/B
= u/D = z/B = z/D = 0.30)

pronounced for reinforcement size ratio (Br/B or Dr~D)up to approximately


4.5, beyond which the BCR remained more or less constant.
A typical variation of BCR with the reinforcement size ratio is shown in
Fig. 10. It is seen that the results of the tests and analyses are in very good
agreement, both indicated that the increase in the BCR became smaller as
the size ratio increased.
Similarly, Omar et al. (1993a, b), based on their study of footings on
geogrid-reinforced sand, indicated that the BCR increased with reinforce-
ment-size ratio up to an effective value of Br/B and remained practically
constant thereafter. It was found that for a given foundation (i.e., B/L),
the effective reinforcement-size ratio increased with decreasing B/L ratio.
In these studies, it was reported that the effective reinforcement-size ratio
was around 8 for strip foundations and 4.5 for square foundations.
Several researchers have also indicated that increasing reinforcement size
beyond a certain value would not increase the BCR significantly. Guido
et al. (1986, 1987) and Singh (1988) advocated that the optimum value of
the reinforcement-size ratio was approximately 2 for square footings on
geogrid-reinforced sand. Fragaszy and Lawton (1984) reported that as re-
inforcement size increased from three to seven times the footing width, the
BCR increased rapidly, beyond which the reinforcement size did not appear
to significantly affect BCR for strip footings on a sand reinforced with strip
reinforcements. On the other hand, Huang and Tatsuoka (1990) indicated
that the bearing capacity of strip footings on sands could increase signifi-
cantly by reinforcing the zone immediately beneath the footing with stiff
short reinforcement layers having only a length equal to the footing width.

Effect of Reinforcement Stiffness


Effect of the reinforcement stiffness on the BCR was investigated only
by the finite-element analysis. The axial stiffness per unit width of rein-
forcement is equal to the product of E (modulus of elasticity) and t (thickness
of reinforcement). In the finite-element analyses, the reinforcement stiffness
was varied by choosing different elasticity modulus E while keeping its
thickness a constant (t = 0.95 ram).
The stiffness of reinforcements that have been used in practice falls over
a wide range. In the present study, the reinforcement stiffness was varied
2096

J. Geotech. Engrg. 1994.120:2083-2099.


Downloaded from ascelibrary.org by MALAVIYA NATIONAL INSTITUTE OF TECHNOLOGY on 06/22/14. Copyright ASCE. For personal use only; all rights reserved.

o" 7 9 s/D = 2%
4.
t, s/D = 4%
5 .t3
o s/D = 6%
-t*
4 + s/D = 8%
L) 9 s/D = 10%

I
0 1000 2000 3000 4000
Stiffness of Reinforcement, k N / m

FIG. 11. Variation of BCR with Reinforcement Stiffness ( N = 3, u/D = z/D = 0.30,
Dr/D = 4.5)

between 15 kN/m and 4,000 kN/m. Typical variations of BCR with the
reinforcement stiffness for different s/D ratios (varying between 2 and 10%)
are shown in Fig. 11. It is seen that increasing reinforcement stiffness beyond
an axial stiffness of approximately 1,000 kN/m would not result in significant
increases in the BCR. Fig. 11 also indicated that the BCR increased pro-
portionally with the s/D ratio for the different axial stiffness values.

SUMMARY AND CONCLUSIONS

A study was undertaken to investigate the bearing capacity of rectangular


footings on geogrid-reinforced sand. The effects of the depth to the first
layer of reinforcement, vertical spacing of reinforcement layers, number of
reinforcement layers, size of reinforcement sheet, and stiffness of reinforce-
ment on bearing capacity were investigated. This paper presents the results
obtained from laboratory tests and finite-element analyses, and a compar-
ison of these results. The study brings forth the following conclusions.
Both the analyses and tests clearly indicated that the bearing capacity of
rectangular footings could be increased significantly by incorporating geo-
grid reinforcement at strategic elevations in the foundation soil. However,
the model tests indicated that the settlement at failure may not be affected
significantly by the geogrid reinforcement. The reinforcement configuration,
that is, the depth to the first layer of reinforcement, the vertical spacing of
reinforcement layers, the size of reinforcement sheet, and especially the
number of reinforcement layers can have a very significant effect on the
bearing capacity of the reinforced foundation.
For single-layer reinforced sand, there is an optimum embedment depth
for the first reinforcement layer at which the bearing capacity is the highest.
The tests indicated that the optimum embedment depth was approximately
0.3 of the footing width. The analyses indicated that the optimum depth
would be somewhat larger for settlement ratios (settlement/footing width)
greater than 6%. For multilayer reinforced sand, the highest bearing ca-
pacity occurs at an embedment depth of approximately 0.25B. For multilayer
reinforced sand there is an optimum vertical spacing of reinforcement layers.
The optimum spacing for the reinforced sand investigated is between 0.2B
and 0.4B.
The bearing capacity of reinforced sand increases significantly with r e -
2097
J. Geotech. Engrg. 1994.120:2083-2099.
Downloaded from ascelibrary.org by MALAVIYA NATIONAL INSTITUTE OF TECHNOLOGY on 06/22/14. Copyright ASCE. For personal use only; all rights reserved.

inforcement size and reinforcement layer number within a certain effective


zone. For the conditions investigated, the extent of the effective zone lies
approximately within 1.5B from both the base and edges of the footing.
Increasing reinforcement stiffness beyond a certain value would only re-
sult in small increases in the bearing capacity of reinforced sand. For the
conditions investigated, that value is 1,000 kN/m.
It should be pointed out that since the influence of foundation size and
the scale effects on the bearing capacity of reinforced soil foundations have
not been investigated fully, the behavior of actual foundations is not well
known. Hence, further studies are needed to establish more accurate design
criteria for reinforced soil foundations.

ACKNOWLEDGMENTS

The first author is a N A T O Science Scholarship Fellow selected by the


Scientific and Technical Research Council of Turkey (TOBtTAK). He would
like to thank T O B I T A K for their support. He also wishes to express his
sincere gratitude to his host university, the University of Colorado at Den-
ver, for providing him with the facilities to continue this research endeavor.

APPENDIX. REFERENCES
Abled-Baki, S., Raymond, G. P., and Johnson, P. (1993). "Improvement of the
bearing capacity of footings by a single layer of reinforcement." Proc., Geosyn-
thetics '93 Conf., Canada, 407-416.
Akinmusuru, J. O., and Akinbolade, J. A. (1981). "Stability of loaded footings on
reinforced soil." J. Geotech. Engrg., ASCE, 107(6), 819-827.
Bassett, R. H., and Last, N. C. (1978). "Reinforcing earth below footings and
embankments." Syrup. on Earth Reinforcement, ASCE, New York, N.Y. 202-
231.
Binquet, J., and Lee, K. L. (1975a). "Bearing capacity tests on reinforced earth
slabs."./. Geotech. Engrg., ASCE, 101(12), 1241-1255.
Binquet, J., and Lee, K. L. (1975b). "Bearing capacity analysis of reinforced earth
slabs." J. Geotech. Engrg., ASCE, 101(12), 1257-1276.
Chou, N. N. S. (1992). "Performance of geosynthetic reinforced soil walls," PhD
thesis, University of Colorado, Boulder, Colo.
Chou, N. N. S., and Wu, J. T. H. (1993). "Effects of foundation on the performance
of geosynthetic-reinforced soil walls." Proc., Geosynthetics '93 Conf., Canada,
189-201.
Duncan, J. M., Byrne, P., Wong, K. S., and Mabry, P. (1980). "Strength, stress-
strain and bulk modulus parameters for finite element analyses of stresses and
movements in soil masses." Rep. No. UCB/GT/80-O1, Dept. of Civ. Engrg., Uni-
versity of California, Berkeley, Calif.
Fragaszy, R. J., and Lawton, E. (1984). "Bearing capacity of reinforced sand
subgrades." J. Geotech. Engrg., ASCE, 110(10), 1500-1507.
Fukagawa, R., Fahey, M., and Ohta, H. (1990). "Effect of partial drainage on
pressuremeter test in clay." Soils and Found., 30(4), 134-146.
Guido, V. A., Biesiadecki, G. L., and Sullivan, M. J. (1985). "Bearing capacity of
a geotextile reinforced foundation." Proc., llth Int. Conf. Soil Mech. and Found.
Engrg., San Francisco, Calif., 1777-1780.
Guido, V. A., Dong, K. G., and Sweeny, A. (1986). "Comparison of geogrid and
geotextile reinforced earth slabs." Can. J. Geotech. Engrg., 23(1), 435-440.
Guido, V. A., Knueppel, J. D., and Sweeny, M. A. (1987). "Plate loading tests on
geogrid-reinforced earth slabs." Proc., Geosynthetics '87 Conf., New Orleans, 216-
225.
Helwany, M. B. (1993). "Long-term soil-geosynthetic interaction in geosynthetic-
reinforced soil structures,"PhD thesis, University of Colorado, Boulder, Colo.
2098
J. Geotech. Engrg. 1994.120:2083-2099.
Downloaded from ascelibrary.org by MALAVIYA NATIONAL INSTITUTE OF TECHNOLOGY on 06/22/14. Copyright ASCE. For personal use only; all rights reserved.

Huang, C. C., and Tatsuoka, F. (1990). "Bearing capacity of reinforced horizontal


sandy ground." Geotextiles and Geomembranes, 9(1), 51-82.
Iizuka, A., and Ohta, H. (1987). "A determination procedure of input parameters
in elasto-viscoplastic finite element analysis." Soils and Found., 27(3), 71-87.
Jewell, R. A., Milligan, G. W. E., Sarsby, R. W., and Dubois, D. (1984). "Inter-
action between soil and geogrids," Syrup. Polymer Grid Reinforcement in Cir.
Engrg., London, England, 18-30.
Khing, K. H., Das, B. M., Puri, V. K., Cook, E. E., and Yen, S. C. (1993). "The
bearing capacity of a strip foundation on geogrid-reinforced sand." Geotextiles and
Geomembranes, 12(4), 351-361.
Lade, P. V., and Duncan, J. M. (1976). "Stress-path dependent behavior of cohe-
sionless soil." J. Geotech. Engrg., ASCE, 102(1), 51-68.
Lambrechts, J. R., and Leonards, G. A. (1978). "Effects of stress history on de-
formation of sand." J. Geotech. Engrg., ASCE, 104(11), 1371-1387.
Milligan, G. W. E., and Palmeira, E. (1987). "Prediction of bond between soil and
reinforcement." Proc., Int. Conf. Prediction and Performance in Geotech. Engrg. ,
Calgary, Alberta, 147-153.
Ohta, H., and Iizuka, A. (1986). Manual of Dacsar FEM program. Kanazawa Uni-
versity, Japan.
Ohta, H., and Iizuka, A, (1988). "Soil-structure interaction related to actual con-
struction sequences." Proc., 6th Int. Conf. Numerical Methods in Geomechanics,
Swoboda, ed., A. A. Balkema, Rotterdam, The Netherlands, 2043-2050.
Ohta, H., Iizuka, A., Mitsuhashi, Y., and Nabetani, M. (1991). "Deformation anal-
ysis of anisotropically consolidated clay foundation loaded by 5 embankments."
Proc., 7th Int. Conf. Comp. Methods and Adv. In Geomechanics, Cairns, Beer,
Booker, and Carter, eds., A. A. Balkema, Rotterdam, The Netherlands, 1017-
1022.
Ohta, H. (1991). "Performance of soft clay foundation under construction." Proc.,
1st Asian Young Geotech. Engrs. Conf., Bangkok, Thailand, 17-35.
Omar, M. T., Das, B. M., Yen, S. C., Puff, V. K., and Cook, E. E. (1993a).
"Ultimate bearing capacity of rectangular foundations on geogrid-reinforced sand."
Geotech. Testing J., 16(2), 246-252.
Omar, M. T., Das, B. M., Puri, V. K., and Yen, S. C. (1993b). "Ultimate bearing
capacity of shallow foundations on sand with geogrid reinforcement." Can. J.
Geotech. Engrg., 30(3), 545-549.
Samtani, N. C., and Sonpal, R. C. (1989). "Laboratory tests of strip footing on
reinforced cohesive soil." J. Geotech. Engrg., ASCE, 115(9), 1326-1330.
Singh, H. R. (1988). "Bearing capacity of reinforced soil beds," PhD thesis, Indian
Institute of Science, Bangalore, India.
Sridharan, A., Murthy, B. R. S., Bindumadhava, F., and Vasudevan, A. K. (1989).
"Model tests on reinforced soil mattress on soft soil." Proc., 12th Int. Conf. on
Soil Mech. and Found. Engrg., Rio de Janeiro, Brazil, 1765-1768.
Yetimoglu, T. (1994). "The bearing capacity of rectangular footings on geogrid-
reinforced sand," PhD thesis, Technical University of Istanbul, Ayazaga, Istanbul,
Turkey (in Turkish with an English Summary).
Yoshimi, Y., Kuwabara, F., and Tokimatsu, K. (1975). "One-dimensional volume
change characteristics of sands under very low confining stresses." Soils and Found.,
15(3), 51-60.

2099
J. Geotech. Engrg. 1994.120:2083-2099.

You might also like