Sei sulla pagina 1di 18

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/223549455

Full-scale field tests on geosynthetic reinforced


unpaved roads on soft subgrade
ARTICLE in GEOTEXTILES AND GEOMEMBRANES FEBRUARY 2006
Impact Factor: 2.38 DOI: 10.1016/j.geotexmem.2005.06.002

CITATIONS

READS

61

302

6 AUTHORS, INCLUDING:
Rudolf Hufenus
Empa - Swiss Federal Laboratories for Mate
114 PUBLICATIONS 197 CITATIONS
SEE PROFILE

All in-text references underlined in blue are linked to publications on ResearchGate,


letting you access and read them immediately.

Available from: Rudolf Hufenus


Retrieved on: 06 January 2016

ARTICLE IN PRESS

Geotextiles and Geomembranes 24 (2006) 2137


www.elsevier.com/locate/geotexmem

Full-scale eld tests on geosynthetic reinforced unpaved roads


on soft subgrade
Rudolf Hufenusa,, Rudolf Rueeggerb, Robert Banjacc, Pierre Mayorc,
Sarah M. Springmanc, Rolf Bronnimannd
a
EMPA, Materials Science and Technology, CH-9014 St. Gallen, Switzerland
Rueegger Systems, Solutions in Geotechnical Engineering, Vonwilstrasse 9, CH-9000 St. Gallen, Switzerland
c
Institute for Geotechnical Engineering, Swiss Federal Institute of Technology, CH-8093 Zurich, Switzerland
d
EMPA, Materials Science and Technology, CH-8600 Duebendorf, Switzerland

Received 16 February 2005; accepted 14 June 2005


Available online 20 October 2005

Abstract
A full-scale eld test on a geosynthetic reinforced unpaved road was carried out, including compaction and trafcking, to investigate
the bearing capacity and its performance on a soft subgrade. The test track was built with three layers of crushed, recycled ll material.
The 1st layer was compacted statically, whereas the 2nd and 3rd were dynamically compacted. The geogrids were instrumented with
strain gauges to measure the short- and long-term deformations and the ongoing formation of ruts was assessed from prole
measurements. The various geosynthetics used for this reinforced unpaved road were found to have a relevant reinforcing effect only
when used under a thin aggregate layer on a soft subgrade. Under such conditions, ruts can form in the subgrade, mobilizing strains and
thus tensile forces in the geosynthetic. The achievable degree of reinforcement depends on the stiffness of the geosynthetic and is limited
by nite lateral anchoring forces.
r 2005 Elsevier Ltd. All rights reserved.
Keywords: Bearing capacity; Full-scale eld test; Reinforcement; Rut formation; Soft subgrade; Unpaved road

1. Introduction
Geosynthetics have been used successfully to reinforce
unpaved roads on soft subgrade for many years. Construction of reinforced temporary roads (Mannsbart et al.,
1999) and bases for heavy machinery (Garcin and Murray,
2003) are examples of short-term usage of the geosynthetic,
where the main goal is to save ll material. In paved roads
(Anderson and Killeavy, 1989; Zia et al., 2001) and railway
tracks (Ashpiz et al., 2002; Izvolt et al., 2001) the adoption
of geosynthetic reinforcement aims at a permanent
improvement of the bearing capacity and the longevity of
the road. Geosynthetics are installed between subgrade and
road to separate or to reinforce. If migration of nes is very
probable, separation is an essential function (Al-Qadi
Corresponding author. Tel.: +41 71 274 7341; fax: +41 71 274 7862.

E-mail address: rudolf.hufenus@empa.ch (R. Hufenus).


0266-1144/$ - see front matter r 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.geotexmem.2005.06.002

et al., 1994; Al-Qadi and Appea, 2003), but with decreasing


bearing capacity of the subgrade, the importance of
reinforcement increases signicantly (Saathoff and Horstmann, 1999).
Numerous eld trials and full-scale laboratory investigations have illustrated that geosynthetics used to reinforce
unpaved roads on soft subgrade facilitate compaction
(Bloise and Ucciardo, 2000), improve the bearing capacity
(Floss and Gold, 1994; Huntington and Ksaibati, 2000;
Meyer and Elias, 1999), extend the service life (Cancelli and
Montanelli, 1999; Collin et al., 1996; Jenner and Paul,
2000; Watts et al., 2004), reduce the necessary ll thickness
(Bloise and Ucciardo, 2000; Cancelli and Montanelli, 1999;
Huntington and Ksaibati, 2000; Jenner and Paul, 2000;
Martin, 1988; Miura et al., 1990), diminish deformations
(Chan et al., 1989; Jenner and Paul, 2000), and delay rut
formation (Cancelli and Montanelli, 1999; Knapton and
Austin, 1996; Meyer and Elias, 1999).

ARTICLE IN PRESS
R. Hufenus et al. / Geotextiles and Geomembranes 24 (2006) 2137

22

Nomenclature
CBR
EV1
EV2

CBR coefcient
Youngs modulus for the 1st plate loading
(MPa)
Youngs modulus for the plate reloading (MPa)

The combination of geosynthetic reinforcement and ll


help to spread the concentrated vertical loads and to inhibit
large deformations and local failures (Su et al., 2002). Two
modes of action can be distinguished (Bourdeau, 1991;
Miura et al., 1990):

Confinement: A vertical load induces lateral forces,


which spread the aggregate particles and thus lead to
local deformations of the ll. Due to frictional interaction and interlocking between the ll material and the
geosynthetic, the aggregate particles are restrained at the
interface between the subgrade and the ll (Jenner and
Paul, 2000). The reinforcement can absorb additional
shear stresses between subgrade and ll (Floss and
Gold, 1994; Meyer and Elias, 1999), which would
otherwise be applied to the soft subgrade (Houlsby
and Jewell, 1990). This improves the load distribution
on the subgrade (Moghaddas-Nejad and Small, 1996)
and reduces the necessary ll thickness. The conning
mechanism does not imply the need for signicant rut
depths to form (Collin et al., 1996; Perkins and Ismeik,
1997), and therefore is also of interest for permanent
paved roads (Sellmeijer, 1990). The effectiveness of the
reinforcement not only depends on the adequate load
transmission to the ll material (via friction and
interlocking), but also is improved by the higher stiffness
of the geosynthetic (Cancelli et al., 1996; Kinney and
Xiaolin, 1995).
Membrane effect (Giroud and Noiray, 1981): If an
unpaved road is pre-rutted during construction, a
geosynthetic reinforcement at the ll-subgrade interface
is distorted and thus tensioned (Meyer and Elias, 1999).
Due to its stiffness, the curved geosynthetic exerts an
upward force supporting the wheel load and thus
improving the bearing capacity (Perkins et al., 1999).
It acts like a tensioned membrane, with the pressure on
the soft subgrade being smaller than the pressure
applied to the ll on the upper, concave side. The
reinforcement, while in tension, spreads the load over a
larger area, leading to a reduction in the settlement
beneath the footing (Ghosh and Madhav, 1994;
Moghaddas-Nejad and Small, 1996). The membrane
effect is predominant for small ll thicknesses (Kenny,
1998) and at low values of shear stiffness of the
granular ll (Ghosh and Madhav, 1994). Signicant rut depths (Perkins and Ismeik, 1997; Watn et al.,
1996) and high stiffnesses of the geosynthetic (Floss
and Gold, 1994) must be provided to initiate the

Evib
h
N
T2%
w
gd

dynamic stiffness (MPa)


road layer thickness (m)
number of standard axle passes
tensile strength at 2% strain (kN/m)
water content (%)
dry density (kN/m3)

membrane effect and thus enhance the bearing capacity


of the footing.
Geosynthetics reinforcing unpaved roads on soft subgrade have been shown to reduce the necessary ll
thickness by approximately 30% (Cancelli et al., 1996;
Cancelli and Montanelli, 1999; Huntington and Ksaibati,
2000; Kenny, 1998; Miura et al., 1990; Perkins et al., 1998;
Watts et al., 2004). Giroud and Noiray (1981) suggested
the following criterion to select the thickness of an
unreinforced unpaved road:
log N

h CBR0:63
,
0:19

(1)

where N is the number of standard axle passes, h the road


layer thickness (m) and CBR is the CBR coefcient.
The empirical approach (1) is valid for Np10000 and a
maximum rut depth of 75 mm, or 40 mm with reference to
the initial level of the pavement, respectively (Jenner et al.,
2002). It is widely applied (Espinoza, 1994; Ingold, 1994;
Koerner, 1997) and has proven satisfactory in practice
(Jenner et al., 2002; Meyer and Elias, 1999; Som and Sahu,
1999).
The impact of reinforcement on an unpaved road on a
soft subgrade is signicant with ll heights less than 0.4 m
only (Collin et al., 1996; Meyer and Elias, 1999; Posposil
and Zednik, 2002). With higher lls, the depth effect of a
(wheel) load generally is too small to mobilize a noticeable
tensile force within the geosynthetic (Gobel et al., 1994).
On the other hand, in unpaved roads the geosynthetic must
be covered by a minimum ll layer of 0.2 m to prevent
damage during trafcking (Hirano et al., 1990; Meyer and
Elias, 1999).
The geosynthetic reinforcement should be placed in the
lower part of the ll height (Jenner and Paul, 2000),
whereas the optimal placement position is dependent on
the subgrade, the ll thickness and the magnitude of the
applied loads. With a soft subgrade and a ll thickness less
than 0.4 m the optimal position lies at the base of the ll
(Cancelli and Montanelli, 1999; Haas et al., 1988; Miura et
al., 1990; Walters and Raymond, 1999). With higher
bearing capacity of the subgrade, increasing ll height or
smaller trafcking loads, the optimal placement position of
the geosynthetic moves upwards to approximately
0.250.35 m below the surface of the ll (Haas et al.,
1988; Moghaddas-Nejad and Small, 1996; Perkins et al.,
1999).

ARTICLE IN PRESS
R. Hufenus et al. / Geotextiles and Geomembranes 24 (2006) 2137

Biaxial geosynthetics are typically used to reinforce


unpaved roads, because well-balanced tensile forces can be
mobilized. Their most important property is the tensile
stiffness at tensile strains between 1% and 2% (Bloise and
Ucciardo, 2000), causing the maximum expected permanent elongation of the geosynthetic (Haas et al., 1988). Due
to settlements, the elongation of the reinforcement
increases with decreasing bearing capacity of the subgrade
(Ghosh and Madhav, 1994).
The reinforcement affects the ll somewhat proportionally to the tensile stiffness of the geosynthetic (Collin et al.,
1996; Gobel and Lieberenz, 1997; Hirano et al., 1990;
Miura et al., 1990), with the soil-geotextile interaction
being a limiting factor (Espinoza, 1994; Palmeira and
Cunha, 1993). In road and railroad applications, the mobilized forces can be limited due to insufcient anchoring.
Thus a very stiff geosynthetic does not necessarily give rise
to an increased reinforcing effect (Bourdeau, 1991).
With stiff geosynthetics, an applied load results in
relatively small elongations and thus minor deformations
of the reinforced ll layer (Chan et al., 1989; Jenner and
Paul, 2000; Meyer and Elias, 1999). On a soft subgrade
(CBRp3), stiff geosynthetics are comparatively more
efcient, but the inuence of the reinforcement tensile
stiffness decreases with increasing bearing capacity of the
subgrade (Cancelli et al., 1996; Cancelli and Montanelli,
1999). The aggregate particles restrain the tensile elements
of the installed geosynthetic and thus stiffen the reinforcement. The degree of stiffening depends on the type of
geosynthetic, with nonwovens being the most susceptible to
connement by interlocking particles (Bauer, 1997).
Due to the fact that only an elongated geosynthetic can
develop forces, the reinforcing effect of a geosynthetic
installed in an unpaved road on soft subgrade often does
not develop until trafcking and some resulting deformation occur (Chan et al., 1989; Huntington and Ksaibati,
2000; Jenner et al., 2002). However, large deformations are
only accepted in unpaved roads. With increasing maximum
rut depth and decreasing bearing capacity of the subgrade,
the impact of the reinforcement on the service life of a road
improves (Cancelli et al., 1996; Cancelli and Montanelli,
1999; Haas et al., 1988). The maximum number of axle
passes required to achieve a given rut depth can be up to 10
times higher on a reinforced unpaved road, compared to
the unreinforced situation (Cancelli and Montanelli, 1999;
Collin et al., 1996; Perkins et al., 1998).
The absence of an accepted design technique explains
why this topic is still being researched despite the initiation
of investigations over 20 years ago (Perkins and Ismeik,
1997). Since the geosynthetic reinforcement interacts with
the soil over the whole width of the unpaved road, the
results of reduced-scale laboratory simulations cannot be
transferred to practice reliably. Currently there are no
incontrovertible indications from laboratory tests of the
inuence that the geosynthetic will have on the performance of the pavement under trafcking (Watts et al.,
2004).

23

2. Experimental
2.1. Concept of field trials
Full-scale eld trials were undertaken in the autumn of
2002 in order to ascertain the effect of geosynthetics on the
load-bearing capacity of an unpaved road on soft
subgrade, which was levelled and prepared in order to
create a track of uniform strength (Hufenus et al., 2004).
Gravelly, angular backll was used for the test track and
compaction and in-service tests were carried out on the ll,
which were reinforced with various geosynthetics. The goal
of the study was to establish the extent to which reinforcing
geosynthetics improve compaction, bearing capacity and
serviceability.
Lessons were learnt from a similar research project
(Schad, 2001) during which weather conditions and
trafcking procedures during construction caused major
variations in the results, which limited assessment of the
results. Surprisingly, in that study no indication of
improved load-bearing capacity was observed due to the
geosynthetic reinforcement (Wilmers, 1999).
An area within a brickworks clay pit in Diessenhofen,
near to Schaffhausen, was available for use as a test track.
The ground consisted of relatively homogeneous clayey silt.
The subgrade is characterized as having a somewhat
irregular bearing capacity, which complicated the interpretation of the results, but also revealed the dependence of
compaction and serviceability (rut formation) on the
subgrade parameters.
The test track was divided into 12 elds (112) of length
8 m, into which one layer of a variety of reinforcing
geosynthetics was placed, and two preliminary test elds
(V1 and V2), where no geosynthetic, or only a separator,
was laid (Fig. 1). The geogrids were partly placed, in
combination with a nonwoven separator underneath.
Three 0.2 m layers (Fig. 2) of relatively poorly compactable
recycled rubble were placed. The 1st layer was compacted
statically (25 kN tandem at roller Bomag BW 120) and the

Fig. 1. Layout of test track with divisions between test elds.

ARTICLE IN PRESS
R. Hufenus et al. / Geotextiles and Geomembranes 24 (2006) 2137

24

2nd and 3rd layers were compacted dynamically with


constant energy (80 kN at roller Bomag Variocontrol BW
177 with overall dynamic compaction control).
The test track was constructed adjacent to an existing
road, with a length of about 130 m. This allowed for
installation from the side, so that the test track was not
subjected to trafc, nor loaded by any equipment prior to
compaction. The subgrade was prepared with a cross slope
(gradient approx. 4%) to allow rain and any seepage water
to run off. Measures were taken such as irrigation and
covering to ensure that the ground was not permitted to
dry out.
Installation, compaction and trafc characteristics were
tested for a section (Zones V1 and V2) that demonstrated
equal or worse conditions compared with elds 112 to
prepare for construction in this area. No geosynthetic
material (Zone V2), or only a separation geosynthetic
(Zone V1), was included at the end of the eld for these
preliminary tests. The truck used for trafcking tests
comprises two single wheel steering axles in the front and
two twin wheel axels in the back. The air pressure of the
30 cm wide truck tires was 8.5 bar.
The progress of the eld trials is listed in Table 1. The
condition of the track and the geosynthetics was monitored
by instruments from installation to removal, using CBR

Fig. 2. Typical cross section through the test track.

measurements (CBR penetrometer), shear vane measurements (Pilcon), specic gravity measurements, static and
dynamic plate load tests, a dynamic falling weight
deectometer (FWD), the overall dynamic compaction
control and the prole measurements (ruts) and strain
gauges on the geogrids.

2.2. Selection of geosynthetics


As a general principle, products that are conventionally
used to reinforce roads were selected (Table 2). These are
biaxial products that are able to withstand approximately
the same force in both directions. Details of the geogrid
rolling width, mesh width and strength mobilized due to
axial tensile strain in the machine (MD) and the cross
direction (XD) are given, together with the respective
number of strain gauges installed (Section 2.5).
Seven different reinforcing geosynthetics were used (nos.
02, 27, 28, 32, 42, 44 and 46), to represent the various
possible raw materials, type and manufacturing process.
Two weaker materials, a nonwoven separating geotextile
(no. 41) and a woven slit tape geotextile (no. 45), were also
included. Nos. 32, 42 and 46 were incorporated with and
without an additional nonwoven separator (no. 40), while
no. 27 was only included in combination with the
nonwoven geotextile. The nonwoven separator no. 40
was installed in the preliminary test eld V1.
The distribution of the geosynthetic samples on the test
track (096 m) is shown in Fig. 3 (grey: grid underlaid with
nonwoven separator), together with the orientation of the
geosynthetic material (arrow direction of production), as
well as the position of the strain gauges and the prole used
to measure the ruts (broken line in Fig. 3 at two locations
per eld). Fill heights and rut depths were measured.

Table 1
Progess of the eld trials
Layer

Day of eld test

Action

Subgrade

1st28th
29th30th

Adjustment and levelling of subgrade


Geosynthetics and cabling for the strain gauges were laid

Layer 1

31st
31st
35th36th

Test track was laid with 0.25 m loose ballast 8/64 (compacted depth 0.2 m)
Test track was compacted purely statically with 25 kN roller, 34 passes
Static plate load and trafcking test over 1st layer with 130 kN truck: 2 passes for plate load
tests, 2 further passes on Zones V1 and V2 and 6 more on elds 112 (total 48 passes)

Layer 2

43rd44th
44th
48th49th

2nd layer was laid with 0.25 m loose ballast 8/64 (compacted depth 0.2 m)
Dynamic compaction with 80 kN roller, 34 passes
Plate load and trafcking tests over 2nd layer with loaded truck (10 driving passes with 220 kN,
10 passes with 280 kN)

Layer 3

57th
57th
76th78th

3rd layer was laid with 0.25 m loose ballast 0/32 (compacted depth 0.2 m)
Dynamic compaction with 80 kN roller, 34 passes
Plate load and trafcking tests with loaded truck (61 driving passes with 280 kN in total)

Subgrade

78th80th

Ballast cleared to approx. 0.05 m over the subgrade (geosynthetics) with a hydraulic excavator
for all proles, nal excavation by hand shovel

ARTICLE IN PRESS
R. Hufenus et al. / Geotextiles and Geomembranes 24 (2006) 2137

25

Table 2
Geosynthetics used in eld experiment
No.

Field

Type of geosynthetic

Width (m)

Grid (mm)

Strain gauges

Tensile strength at (kN/m)


2%

02
27
28
32
40
41
42
44
45
46

10
9
2
5/6
V1
12
3/4
11
1
7/8

PP slit tape woven


Biaxial extruded PP grid in 5 layers
PVC-coated knitted PET grid
PET at rib grid
PP nonwoven (separation)
PP nonwoven (reinforcement)
PVC-coated knitted PVA grid
PET yarn reinforced PP nonwoven
PP slit tape woven
Biaxial extruded PP grid

5.15
4.50
5.10
4.75
5.00
5.00
5.20
5.20
5.15
3.80

60  60
20  20
32  32

40  40
8.5  8.5

65  65

4
8

12

5%

max.

MD

XD

MD

XD

MD

XD

12
6
9
10
0.2
0.4
12
7.5
2
11

12
10
9
10
0.1
0.3
12
7.5
2
12

30
14
14
20
0.3
0.6
32
22
8
22

30
20
14
20
0.2
0.4
32
22
8
25

65
22
55
30
10
20
40
50
30
30

65
35
55
30
10
20
40
50
30
30

shows the loadstrain curves of the virgin as well as the


installed and excavated geosynthetics, determined according to EN ISO 10319 (1996).
The shape of the loadstrain curves proved to be almost
unaffected by installation damage and subsequently by
trafcking, i.e. even with a small decrease of the ultimate
tensile strength and the elongation at break, the tensile
stiffness remained approximately the same, in accordance
with other studies (Hufenus et al., 2002). Predicted strains
of the geosynthetic reinforcement were far below the
equivalent strain at failure so that installation damage
did not inuence the loadstrain behaviour.

2.3. Soil parameters and environmental conditions

Fig. 3. Test track set-up.

The loadstrain behaviour of the geosynthetics has been


tested either longitudinally or transversely, depending on
the alignment of the samples in the eld trial. Geosynthetics installed transverse to the track have been tested in
machine direction, the rest in cross direction (Fig. 3). Fig. 4

The subgrade was classiable as CM (medium plasticity


silty clay). The water content near the surface was
measured as w 38:6  4:7%. The distribution of particle
sizes can be seen in Fig. 5.
A penetrometer was used to determine the CBR
coefcients at depths of approximately 0.3, 0.45 and
0.6 m before installation and after removal of the ll.
Measurements were carried out for every prole of elds
112 and V1, V2 (Fig. 3) along the axis of track, as well as
at a distance of 0.5 and 1.0 m on the left- and right-hand
sides, respectively. It was assumed that the layers near the
surface have a greater inuence on the bearing capacity and
the deformation behaviour of the subgrade than the deeper
layers. Taking into account that the normal compressive
stress under a plate load decreases progressively with
subgrade depth, a weighted average CBR coefcient was
dened, with weightings of three, two and one according to
depths of 0.3, 0.45 and 0.6 m, respectively. Fig. 6 shows
thus weighted average CBR coefcients determined after
removal of the ll. CBR coefcients higher than 12
(measuring range of the penetrometer exceeded) were
reported as 12.

ARTICLE IN PRESS
R. Hufenus et al. / Geotextiles and Geomembranes 24 (2006) 2137

26

virgin

installed in 5-2

virgin

no. 32

Tensile strength [kN/m]

Tensile strength [kN/m]

installed in 6-2

40
30
20
10
0

installed in 3-2

no. 42
40
30
20
10
0

Strain [%]

virgin
50

Strain [%]
virgin
installed in 7-2

installed in 1-2
40

no. 45

Tensile strength [kN/m]

Tensile strength [kN/m]

installed in 4-2

50

50

40
30
20
10

installed in 7-1
installed in 8-2

no. 46

30
20
10
0

0
0

Strain [%]

Strain [%]

Percent finer by weight [%]

Fig. 4. Loadstrain-curve of the exhumed geosynthetics nos. 32, 42, 45 and 46, installed in elds 6-2, 5-2, 4-2, 3-2, 1-2, 7-1, 7-2, 8-2, respectively.

100
90
80
70
60
50
40
30
20
10
0
0.001

subgrade
fill (layer 1&2)
fill (layer 3)

0.01

0.1

10

100

Particle size [mm]


Fig. 5. Particle size distribution in the subgrade and ll.

The subgrade was not uniform: some zones of soft to


very soft consistency were located next to stiffer areas,
where the bearing capacity and shear strength were also
higher.
Mixing and regrading the subgrade along the entire test
track was not considered to be feasible, so the ground was
scaried and regraded in elds 14 in order to reduce the
bearing capacity to that equivalent elsewhere along the test
track. CBR measurements made after the track had been
removed indicated that the subsoil in elds 14 had

reconsolidated to almost similar CBR values as in the


pre-test soil conditions.
The undrained shear resistance of the subgrade was
measured directly in the eld using a Pilcon shear vane,
where the maximum detectable shear resistance was
124 kPa, corresponding to a CBR value of approx. 34
(Jaecklin and Floss, 1988; Saathoff and Horstmann, 1999).
Post-test measurements of CBR values were used for the
interpretation of the inuence of the ll layers and
reinforcement on rut formation due to compaction and
trafcking.
Loose recycled rubble, consisting primarily of concrete
scrap and secondarily of brickwork scrap (poorly graded
gravel, GP), was used to form the ll layers. The material
was broken down to a maximum grain size of 64 mm, and
the ne portion (with a diameter ofo8 mm) was sieved out.
The particle sizes for layers 1 and 2 then range between
approx. 8 and 64 mm (Fig. 5). Because the proportion of
small particles was limited, the material demonstrated low
sensitivity to changes in the water content, and was
sufciently porous so that meteorological and percolating
water was conducted quickly into the lateral drainage
ditches.
In contrast, ner grained recycled material, with a
particle size of 032 mm (poorly graded sandy gravel, also

ARTICLE IN PRESS
R. Hufenus et al. / Geotextiles and Geomembranes 24 (2006) 2137

27

12
10
CBR

8
6
4
2
V2-2
V2-1
V1-2
V1-1
12-2
12-1
11-2
11-1
10-2
10-1
9-2
9-1
8-2
8-1
7-2
7-1
6-2
6-1
5-2
5-1
4-2
4-1
3-2
3-1
2-2
2-1
1-2
1-1

Pro

file

1
0.5 ]
0
[m
-0.5
-1
ion
t
i
s
Po

Fig. 6. Weighted average CBR coefcients after removal of the foundation.

GP) was used for the 3rd layer, in order to achieve an


improvement in density and hence interlocking, so that less
particle movement in the voids within the material occurs
when the ruts are driven over during trafcking.
The subgrade and 1st ll layer were articially watered
during a very dry period in the middle of August to ensure
that the consistency and hence shear strength of the silty
clay remained relatively constant and the clay did not dry
out. The 2nd and 3rd layers have been levelled off with an
excavator shovel to remove the ruts developed prior to
covering with the top layer. Evaluation of the effect of
crushing by comparing particle size distribution curves
before and after installation, compaction and trafcking
showed that there was no signicant inuence.

2.4. Compaction controls


The dry density gd and the water content w of the 1st,
2nd and 3rd layer of the test track were determined using
the Troxler apparatus. The method has not been calibrated
to other standard tests, such as the sand replacement
method. Thus, the results are only valid for comparitive
purposes. Proctor tests were not carried out due to the high
percentage of coarse grains.
Static plate load tests were carried out to reveal the
Youngs moduli EV1 (1st plate loading) and EV2 (plate
reloading), using a 300 mm diameter device to determine
the deformability and load-bearing capacity of the ll
layers. Measurements were completed for each prole
(Fig. 3) and directly above the strain gauges in elds 2, 5
and 6.
Some dynamic plate load tests (FGSV, 1997) were
carried out to reveal the Youngs moduli of each of the ll
layers. However, performing the test on the unbound
surface of the test track led to unsatisfactory results.

Dynamic compaction control is a technique used to


measure the load-bearing capacity of compacted ground by
analysing the dynamically excited roller (Floss, 2001).
Conclusions can be drawn about the dynamic stiffness Evib
and/or the degree of compaction by measuring and
analysing the acceleration. The depth of measurement for
this is greater than the depth of compaction and in this
particular case reaches the soft subgrade. Consequently the
dynamic compaction results reect primarily the properties
of the subgrade and not those of the ll layer.
2.5. Deformation measurements
The prole measurement used to assess the formation of
ruts on the trafcked ll layers and on the subgrade after
removal of the ll was carried out using a cross bar
developed specially for this eld test. This cross bar rests on
the left and right measuring posts driven in on either side of
the track (Fig. 2), and the distance of the crossbar above
the track is measured with an accuracy of 75 mm.
The geosynthetics have been instrumented with electrical
resistance strain gauges (ERSG) to determine their shortand long-term deformations. Static measurements have
been performed to investigate deformations and stresses
caused by the plate load test. Dynamic measurements were
performed to study the inuence of trafcking on
compaction. The foil strain gauges consisted of a constantan grid on a polyamide lm. The maximum tensile
strain was specied as 5%.
A single component, cold curing adhesive made of
cyanacrylate (Z70), was used to x the strain gauges to the
geogrids. The uneven, coated knitwear was treated with a
two component polyurethane adhesive/ller to atten the
surface and to glue on the strain gauges. The strain gauges
needed to be protected (Bathurst et al., 2002; Springman
and Balachandran, 1994). Protection against ingress of

ARTICLE IN PRESS
R. Hufenus et al. / Geotextiles and Geomembranes 24 (2006) 2137

28

water was achieved by covering with a tough, kneadable


putty that strongly adheres on nearly every material
(AK22). For protection against installation damage, a
rubber foil was loosely attached to the geosynthetic around
the strain gauge without hindering the elongation.
To validate the instrumentation process, i.e. protecting
the strain gauges without inuencing the measurement,
prior to the eld tests, a strip of geosynthetic had been
instrumented and tested in a laboratory simulator for the
installation process according to prEN ISO 10722-1, 2004.
The performance of the installed strain gauges on the
geosynthetic has been compared to an extensometer in a
tensile testing apparatus before and after the simulated
installation, following the standard EN ISO 10319 (1996).
The traverse speed was set to a reduced value of 3 mm/min
to achieve a better temporal resolution.
It was not possible to glue the strain gauges to the slit
tape woven material no. 2. The adhesive led to a stiffening
of the material resulting in a forcestrain relationship,
which is far too steep. Likewise, geogrids could be
instrumented with strain gauges, but not geowovens
(Bathurst et al., 2002). Samples no. 32, 42 and 46 showed
a very good agreement of the force-strain curves. The
installation simulation had neither negative inuence on
the strain gauge nor on the adhesive joint. All strain gauges

Fig. 7. Positioning of the loading plate (static plate load test) and the
transverse strain gauges.

survived the eld test without failure due to the protective


measures taken.
The selection of the samples to be instrumented was
made according to technical practicability (the potential for
attaching the strain gauges without altering the forcedeformation characteristics of the geosynthetic). Four
strain gauges were tted in a line at right angles to the
track axis for each sample under test. Fig. 7 shows their
positions. Since the truck was not externally guided, the
course of the track varied during trafcking. Therefore the
position of the strain gauges could not be chosen to be
precisely below the truck tires.
Fig. 3 shows the position of all strain gauges installed. In
addition to the strain gauges mounted at right angles in
elds 28, four additional strain gauges were afxed in the
direction of travel in eld 7, in order to measure the
longitudinal elongation of the geosynthetic.

3. Results and discussion


3.1. Compaction improvement
The dry density shows no relevant dependency on the
properties of the subgrade and the reinforcement. The
relatively low density of the 1st layer (gd 14.970.2 kN/
m3, w 6.370.4%) and the 2nd layer (gd 14.370.3 kN/
m3, w 7.470.4%) is due to the difculty of compacting
such coarse-grained ll (particle size between 8 and
64 mm), with signicant void space, and also to the lighter
specic weight of the material stemming from the broken
masonry components. Material from the same source was
used for the well-graded 3rd layer, which has been
compacted
to
a
signicantly
higher
density
(gd 16:7  0:1 kN=m3 , w 10:5  0:3%).
Youngs moduli EV1 and EV2 were calculated from the
static plate load test data for the 1st loading and reloading
cycles, respectively (cf. Fig. 8). Because the subgrade was so
soft, it was rarely possible to achieve a maximum initial
load of 0.5 MPa. The 1st loading has been increased in

Profile 3-1

Profile V1-2
0

0
1st loading
unloading
reloading

20

10
Settlement [mm]

Settlement [mm]

10

30
40
50
60

20
30
40

1st loading
unloading
reloading

50
60

0.1

0.2

0.3

0.4

Vertical stress [MPa]

0.5

0.6

0.1

0.2

0.3

0.4

0.5

Vertical stress [MPa]

Fig. 8. Vertical stress-settlement diagrams of the 1st ll layer, for proles V1-2 (without reinforcement) and 3-1 (with reinforcement).

0.6

ARTICLE IN PRESS
R. Hufenus et al. / Geotextiles and Geomembranes 24 (2006) 2137

29

40

Young's modulus of the plate reloading on the 1st layer


Young's modulus of the plate reloading on the 2nd layer
weighted average CBR coefficient of the subgrade

20

Profile

CBR [%]

1-1
1-2
2-1
2-2
3-1
3-2
4-1
4-2
5-1
5-2
6-1
6-2
7-1
7-2
8-1
8-2
9-1
9-2
10-1
10-2
11-1
11-2
12-1
12-2
V1-1
V1-2
V2-1
V2-2

EV2 [MPa]

60

0
Fig. 9. Youngs moduli EV2 of the 1st and 2nd ll layer, compared to the weighted average CBR coefcients of the subgrade.

70
Young's modulus EV2 [MPa]

0.05 MPa steps until a settlement of approx. 50 mm was


achieved for tests on the rst 0.2 m thick layer. Reloading
was carried out with 0.07 MPa steps until the penultimate
stress from the 1st loading cycle was achieved. Consolidation during the test also made it nearly impossible to expect
a settlement change ofo0.02 mm/min in accordance with
the standards.
Fig. 8 shows settlements of 17 and 51 mm, respectively
for a 1st loading cycle to 0.35 MPa, which is typical for the
elds with geogrid (eld 3) and without geogrid (eld V1)
reinforcement. For tests on the top layer, the inuence of
the reinforcement is marginal, since the deformations are
too small to mobilize forces with signicant vertical
components for just one loading, unloading and reloading
cycle.
Fig. 9 shows the Youngs moduli EV2 of the reloading
cycle of the static plate load test in comparison with the
weighted CBR values following the removal of the track.
Signicantly higher CBR values of 68 were measured (in
the area of elds 2 and 3). This inuenced both the
interaction with the 1st ll layer and improved the
compactability of the 2nd and 3rd layers in these elds.
The highest stiffnesses measured for the geogrid reinforcement were without the separation layer in elds 2, 3,
6 and 7. The amount of improvement in the EV2 value
appears to have been reduced when the separation layer
(elds 4, 5, 8 and 9) was present, due to a combination of
lower frictional resistance between the two geosynthetics
and less opportunity for interlocking with the geogrid and
the ll. Values of EV2 for eld 8 that included the extruded
geogrid with the separation layer were particularly poor,
but they probably reect the effect of the lowest CBR
values.
No signicant difference between the remaining elds
and the one representing the unreinforced eld (V2) was
apparent. Typically, similar elasticity moduli were measured after the dynamic compaction of the 2nd layer from

60
50

grid alone, strong slit tape woven


grid with nonwoven separator
nonwoven, weak slit tape woven
not reinforced
R2 = 0.6

40
30
20
10

R2 = 0.2
R2 = 0.6

0
1
3
4
5
7
8
0
2
6
9
10
weighted average CBR coefficient of the subgrade [%]
Fig. 10. Correlation between the Youngs modulus EV2 of the 2nd ll
layer and the weighted average CBR coefcient of the subgrade.

elds 411 (CBRE0.51.5) and elds (V1 and V2)


(CBRE24). This implies that the improvement in bearing
capacity of the subsoil due to the reinforcement may be
equivalent to an increase in the CBR of DCBR  12.
A correlation between the reloading modulus EV2 on the
second layer and the weighted CBR coefcients of the
subsoil (averaged over the width of the test track) has been
presented in Fig. 10. Although the data is somewhat varied,
general tendencies can be noted.
The ratio of EV2/EV1 from plate loading tests on the 1st
ll layer in elds 411, with reinforcement, was signicantly smaller than values from more robust subsoil (elds
2 and 3) or without any reinforcement (elds 1, 2, V1, V2).
This indicates that the reinforcement on soft ground was
activated and slightly recovered during the unloading.
The recycled ll material was difcult to compact. This
was manifested in the higher values of the ratio of EV2/EV1
(E45) for the large ll layer thickness and the better
quality ground. The values of EV1E22 MPa and
EV2E110 MPa were rather low for the recycled ll material

ARTICLE IN PRESS
30

R. Hufenus et al. / Geotextiles and Geomembranes 24 (2006) 2137

Fig. 11. Results of the overall dynamic compaction control on the 2nd and 3rd layer.

and certainly would not have reached the requirements for


a ll layer of an uncompacted road.
The determination of stiffness Evib deduced from the
dynamic compaction control measurement values from the
4th roller pass over the left and right lane, respectively, on
the 2nd layer (d 0:4 m) and 3rd layer (d 0:6 m) are
plotted in Fig. 11. Only some of the vibration energy from
the dynamic compaction has been effective in compacting
the 1st and 2nd ll layers due to the weak subgrade.
Consequently, the dynamic compaction results reect the
in situ subgrade properties, and hence there are analogies
with the static EV2 values, with the exception of the lack of
increase over elds 57. The higher values shown in Fig. 11
between elds 12 and V1 (chainages 96120 m) and beyond
chainage 136 m were largely due to having placed the rst
two ll layers to a total thickness of 0.6 m rather than
0.4 m, with the inevitable effect on the measured elasticity
moduli.
There appears to be no direct relationship between the
dynamic stiffness of the 3rd layer and the subsoil properties
(Fig. 11) because less energy reaches the natural ground
due to the total layer thickness of 0.6 m. The reductions in
values measured along the edges of eld 11 and 12 are due
to inadequate support on one side of the track, causing
lateral spread during dynamic compaction.
The 3rd layer was found, in general, not to contribute to
the bearing capacity of the test track founded on the stiffer
ground. Marked geosynthetic-specic Evib values were also
not observed except for reaching similar values in elds V1
(with only a separating layer) and V2 (without reinforcement) to the reinforced elds 49 on weaker ground.

3.2. Rut formation


A selection of prole measurement results is shown in
Fig. 12 (prole V1-2, nonwoven separator), Fig. 13 (prole
1-1, slit tape woven, tensile strength 30 kN/m) and Fig. 14
(prole 5-1, at rib grid), where the thickness of the layers
is measured in relation to the initial level of the subgrade
( 0 mm). These examples show the rut formation on an
unpaved road on soft subgrade without reinforcement
(Fig. 12), with a relatively weak geosynthetic (Fig. 13),
and with a comparatively stiff geosynthetic reinforcement
(Fig. 14), respectively.
The 1st layer has been trafcked with an unloaded
130 kN truck. Deep ruts developed in the unreinforced
preliminary test elds V1 and V2 after 4 passes, in spite of
the better bearing capacity of the subgrade, compared to
the elds 311. Field 1, reinforced by relatively weak slit
tape woven, neared the state of failure after 8 truck passes
over the 1st layer.
Despite the relatively high bearing capacity of the
subgrade (CBR 812) in elds 2 and 3 as well as at the
beginning of eld 4, considerable rut formation occurred.
The inuence of the bend in the track resulted in slight
lateral spreading of the wheel loads, explaining the
relatively shallow ruts in the elds 812. Comparatively,
small ruts were found in the elds 6 and 7, reinforced by
bonded geogrids without nonwoven underlay.
The 2nd layer was trafcked 10 times with a 220 kN
truck and 10 times with a 280 kN truck, and the smallest
ruts were found in the reinforced elds 2 and 3. This must
be due to the relatively high bearing capacity of the

ARTICLE IN PRESS
R. Hufenus et al. / Geotextiles and Geomembranes 24 (2006) 2137

3rd layer trafficked


before installation

2nd layer trafficked


after removal

31

1st layer trafficked

800
700
Fill thickness [mm]

600
500
400
300
200
100
0
-100
-200
0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0

Position in profile, from left to right measuring post [m]


Fig. 12. Rut formation without reinforcement (prole V1-2, sample 40).

3rd layer trafficked


before installation

2nd layer trafficked


after removal

1st layer trafficked

800
700
Fill thickness [mm]

600
500
400
300
200
100
0
-100
-200
0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0

Position in profile, from left to right measuring post [m]


Fig. 13. Rut formation with weak reinforcement (prole 1-1, sample 02).

subgrade in this section of the test track, which resulted in a


well-compacted ll. On the other hand, the high bearing
capacity of the subgrade in the unreinforced elds V1 and
V2 did not inhibit the formation of ruts. The smallest ruts
were found again in the elds 6 and 7-1 for the track
sections with low bearing capacity (Fig. 15).
The 3rd layer has been trafcked with a 280 kN truck,
with measurements after 11 and 61 passes. Again, the
smallest ruts were found in the reinforced elds 2 and 3
with high bearing capacity of the subgrade, as well as in the
elds 6 and 7 with the bonded geogrids without nonwoven
underlay. The rut formation was increased in the elds
812, with the stronger subgrade being effective in eld 12.
Due to the load distribution through the granular ll, the
ruts formed on the subgrade surface are shallower and

wider than the track surface above. Considerable deformation of the subgrade was measured after the excavation of
the ll in the poorly reinforced elds 1 (weak slit tape
woven), 12 (strong nonwoven), V1 (separating nonwoven
only) and V2 (no geosynthetic). The low bearing capacity
of the subgrade resulted in relatively deep ruts in the elds
8 and 9 (extruded geogrids with nonwoven underlay), but
did not cause large deformations in the geogrid-reinforced
elds 47. The small ruts formed on the subgrade of the
elds 3, 4, 10 and 11 are partly explained by the
comparably high bearing capacity of the subgrade (elds
3 and 4) and by the lateral spreading of the wheel loads due
to the slightly curved track (elds 10 and 11), respectively.
Fig. 15 compares the mean depths of the left and right
ruts with the Youngs moduli EV2 of the 2nd and 3rd layers

ARTICLE IN PRESS
R. Hufenus et al. / Geotextiles and Geomembranes 24 (2006) 2137

32

3rd layer trafficked


before installation

2nd layer trafficked


after removal

1st layer trafficked

800
700
Fill thickness [mm]

600
500
400
300
200
100
0
-100
-200
0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0

Position in profile, from left to right measuring post [m]

24

-4

20

-8

16
2nd layer after 10 passes
2nd layer after 20 passes
Young's modulus layer 2
average CBR coefficient

-12

3rd layer after 11 passes


3rd layer after 61 passes
Young's modulus layer 3

12

-16

-20

-24

EV2 [10 MPa], CBR [%] respectively

1-1
1-2
2-1
2-2
3-1
3-2
4-1
4-2
5-1
5-2
6-1
6-2
7-1
7-2
8-1
8-2
9-1
9-2
10-1
10-2
11-1
11-2
12-1
12-2
V1-1
V1-2
V2-1
V2-2

Rut depth [cm]

Fig. 14. Rut formation with stiff reinforcement (prole 5-1, sample 32).

Profile
Fig. 15. Rut formation, compared to Youngs moduli EV2 and CBR coefcients.

CBR = 0.5

Layer thickness [mm]

800
reinforced
unreinforced

700
600

R2 = 0.96

500

CBR = 1

R2 = 0.92

400

CBR = 2

300
200
100
1

10
100
Number of axle passes

1000

Fig. 16. Number of axle passes achievable without formation of ruts


deeper than 40 mm.

and the average CBR values of the subgrade. The gure


illustrates that the rut formation is primarily a reection of
the values of EV2, the latter being affected by the bearing
capacity of the subgrade and the reinforcement of the ll
layer. The relatively deep ruts in the elds 7-2 and 8 can be
explained by the very low CBR values of the subgrade in
this section. On the other hand, the results of the elds 2
and 4 illustrate that a high bearing capacity with CBR46
increases the longevity considerably. Taking the variations
of the CBR values into consideration, the rut formation is
minimal in the sections reinforced with at rib grids or
extruded grids.
Fig. 16 illustrates the maximum number of truck axle
passes achievable before the rut depth reaches approx.
40 mm, as a function of the ll layer thickness. Hence the
effect of the reinforcement decreases with increasing

ARTICLE IN PRESS
R. Hufenus et al. / Geotextiles and Geomembranes 24 (2006) 2137

thickness of the layer, becoming insignicant beneath a ll


with a thickness exceeding 0.6 m. The transitional straight
lines represent the minimum thickness of an unreinforced
unpaved road as a function of the bearing capacity of the
subgrade, as proposed by Eq (1). They show that the
improvement of this road (approx. 0.4 m thickness)
corresponded to an assumed enhancement of the CBR
coefcient by approx. 0.71.
3.3. Strain development
Fig. 17 shows the results of the static strain measurements under the right wheel track (position 3) during
compacting and trafcking of layers 13. The permanent
deformation was below 1%. Trafcking layer 2 resulted in
a signicant strain increase for all strain gauges attached to
the geogrid in eld no. 8. It is possible that the conning
load provided by a 0.4 m thick cover anchored the
geosynthetic so that more strain was generated under the
wheel. On the other hand, the ll layer was still thin enough
for the build up of lasting ruts, which are still apparent on

Strain [%]

1st layer
compacted

0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0.0
-0.1
-0.2

1st layer
trafficked

2nd layer
compacted

2nd layer
trafficked

3rd layer
compacted

3rd layer
trafficked

field 2
field 3
field 4
field 5
field 6
field 7
field 8

Fig. 17. Strain gauge measurements in position 3 with respect to the


consecutive loadings.

unloading

the subgrade. Compacting and trafcking of layer 3


increased the permanent strain only insignicantly despite
the large number of truck passes.
The static measurements were primarily used to assess
the impact of the plate load test on the geosynthetics
underneath. The load plate could not be placed exactly
above the strain gauges due to the setup of the experiment.
Fig. 18 shows the vertical stress-settlement of the plate load
test on the 1st layer (Section 3.1), as well as the
corresponding tensile and compressive strain of the 4
strain gauges in eld 2 (static, pointwise strain measurement during plate load test) for the prole at position 12 m
(sample no. 28). The negative strains can be explained by a
sidewise raise of the subgrade during the plate load test,
resulting in a compression of the stain gauges xed on the
rear side of the bulging geogrid.
The plate load test on the 1st layer generated similar
strains in the geosynthetic as trafcking (dynamic peak
loads below the immediate load). The deformation
returned to its original value by releasing the plate load
and the permanent pretension remained relatively small
(see Fig. 17). Beginning with the 2nd layer (40.4 m), the
plate load test hardly produced any additional strain,
because the load plate with a diameter of 0.3 m had only
small inuence at the depth of the strain gauge.
The water-saturated subgrade behaved under dynamic
load during trafcking as if it were undrained, which
resulted in settlement below the wheels and heave alongside
them (Giroud and Noiray, 1981). The deformation of the
terrain led to wave-like bending of the geosynthetics. Since
the strain gauges were installed on the bottom side of the
geosynthetics, i.e. below the neutral line, this bending
induced additional strain, which was superimposed. Therefore, the strain gauges at the outer positions had to sustain
considerable compressive strains. In view of the fact that
the location of the plate load tests with respect to the strain
gauges was somewhat arbitrary, the results are biased and
of limited signicance.

reloading

Pos. 1

0.10

10

0.05
Strain [%]

Settlement [mm]

1st loading

20

30

33

Pos. 2

Pos. 3

0.2

0.3

Pos. 4

0.00

-0.05

40

-0.10
0

0.1

0.2

0.3

Vertical stress [MPa]

0.4

0.5

0.0

0.1

0.4

0.5

Vertical stress [MPa]

Fig. 18. Vertical stress-settlement (left-hand side) and corresponding strain gauge measurements (right-hand side) in the prole at position 12 m during
static plate load testing on 1st layer.

ARTICLE IN PRESS
R. Hufenus et al. / Geotextiles and Geomembranes 24 (2006) 2137

34

same order of magnitude as in EN ISO 10319 (1996).


Therefore, it was reasonable to estimate the forces
(Table 3) mobilized by deformation of the geosynthetics
according to the loadstrain diagram in Fig. 4. Large
differences in the strain rate would produce erroneous
prediction, since the stiffness of the geosynthetics depends
strongly on the load speed (Walters et al., 2002).
The maximum strains and corresponding forces had
been partly reached during trafcking of the rst layer, and
partly during compacting and trafcking of layer 2 as the
investigations show. Due to the higher load-bearing
capacity of the subgrade in elds 2 and 3, the maximum
strain was only insignicantly higher than the permanent
strains or forces (Fig. 17). Higher strains and forces had
been measured in eld 4 (knitted grid with nonwoven
underlay), which had a soft subgrade.
The deformation and corresponding forces remained
relatively low in the elds 5 and 6 with at rib grids. The
higher measured strain in the extruded geogrids in elds 7
and 8 could have been caused by local loading. Bathurst
et al. (2002) noted that local pressure caused by larger
stones on the thin extruded elements of the geogrids had

The purpose of the dynamic measurements was to


investigate the dependence of short-term deformations
under the inuence of compacting and trafcking. Fig. 19
serves as an illustration. It shows the strain measured at
position 3 (centre of the right lane) of eld 7 during the
entire duration of the tests. Minimum (min) and maximum
(max) strain of each event (compacting or trafcking) are
given.
The additional mobilized permanent strength, which
results from installation and loading of layers 2 and 3 are
given in Table 3. Compacting with the roller generated only
short-term strain in the geosynthetic, which decayed after
the roller passed and the same occurred with strains in the
driving direction caused by trafcking with trucks.
Permanent strain in the cross direction with the corresponding prestress of the reinforcement built up only if the
subgrade has been deformed correspondingly. The temporary strain generated during roller or truck pass,
exceeded the permanent strain by a factor of two.
The permanent strain of the geogrid was usually below
0.5% and exceeded 1% only in extreme cases. The strain
rates measured during trafcking the ll layer were in the

Day of field test


31th

44th

49th

49th

57th

77th

78th

78th

78th

78th

78th

1.2
1.0

Strain [%]

0.8
max

0.6

compacting 3rd layer


trafficking 2nd layer with 28 t truck

0.2
0.0
-0.2

min

trafficking 3rd layer

0.4

trafficking 2nd layer with 22 t truck


compacting 2nd layer
trafficking 1st layer
compacting 1st layer

Fig. 19. Deformations beneath the centre of the right lane (position 3) in eld 7.

Table 3
Approximate tensile strength induced by ll installation and loading
Field

2, 3
4
5, 6
7
8

Subgrade

Relatively rm
Soft
Soft
Soft
Soft

Geosynthetic

Knitted grid
Knitted grid
Flat rib grid
Extruded grid
Extruded grid

Nonwoven
underlay

With
Without
Without/with
Without
With

Additionally mobilized permanent


strength (kN/m)
2nd layer

3rd layer

3
6
1
3
8

1
3
1
1
1

Permanent
strength (kN/m)

Maximum
strength (kN/m)

7
9
5
6
10

11
17
9
10
18

ARTICLE IN PRESS
R. Hufenus et al. / Geotextiles and Geomembranes 24 (2006) 2137

led to greater local straining. Alternatively, the larger grid


size (Table 2), relative to the soil grading (Fig. 5), will have
permitted dilatation to occur within the opening. The
highest strain and forces have been measured with extruded
geogrids over nonwoven underlay (eld 8).
The formation of ruts in the subgrade can be used to
evaluate the deformation in geosynthetics. The rut was
approximated through the sections A-B, B-C and C-D for
the calculation of the elongation as shown in Fig. 20. The
relative elongation was then estimated from the relative
change of the original length A-D and the stretched length
A-B-C-D.
Average values and the 95% condence interval of
the relative elongation calculated from measurements of
the ruts as well as the permanent strain measured with the
strain gauges (average of positions 14) are shown in
Fig. 21. The comparison shows good agreement. The strain
gauge measurements are slightly higher than those derived
from the rut measurements, with the exception of the at
rib grid (sample 32) in eld 6. A plausible explanation for
the large deformations in eld 8 is that the extruded grid
slid on the nonwoven underlay.
4. Conclusions

35

Typical requirements for the subbase (EV1X11 MPa


and/or CBRX6) were achieved on some test track sections
for which a reinforcing geosynthetic contributed little, even
for very thin ll layers. In this case, problems due to
reaching the bearing capacity were not expected.
4.2. Impact on compaction
Compaction of the 1st and 2nd layers was primarily
affected by the ground properties. The compactability of
thin layers (hp0:5 m) could be improved by inclusion of a
reinforcing geosynthetic if the ground had CBRp3. This
interaction represents a hypothetical improvement in the
ground properties of DCBR  12. Stiff at rib and
extruded grids appeared to have the greatest effect.
Transverse strains in the geosynthetics under the ruts of
depth up to 10 cm were between 0.5% and 1% for layer
thicknesses hp0:5 m, and these values were approximately
doubled during dynamic compaction.
The compactability of the layers hX0:5 m, measured in
terms of the increase in the elasticity modulus, was more
dependent on the compaction properties of the recycling
materials used for the ll than the ground response.
Virtually no further tension was mobilised in the reinforcement during compaction for the third ll layers (hX0:6 m).

4.1. Impact on bearing capacity


4.3. Impact of reinforcement on rut formation
A signicant improvement in the bearing capacity of a
ll layer reinforced by a geosynthetic was found to be true
only for thin layers (hp0:5 m) on very weak ground
(CBRp2). The inuence on the bearing capacity for
thicker ll layers, or on stiffer and stronger ground, was
marginal.

Fig. 20. Assessing the rut outline.

1.2
derived from subgrade deformation

Strain [%]

1.0

derived from strain gauges

0.8
0.6
0.4
0.2
0.0

For thin ll layers (h  0:4 m), the rut formation on


weak ground with geotextile reinforcement was signicantly less than without reinforcement, so that either
higher axle loads are possible for reinforced tracks, or for
the same axle loads, reinforced ll layers can be trafcked
by more passes until the same rut depths are reached. Rut
depths should be limited to less than 10 cm for tracks to be
trafcked by trucks up to 400 kN. This eld test demonstrated that this trafcability limit was reached very quickly
for un-reinforced layers with h  0:4 m. Thicker ll layers
will be necessary for heavier transport with greater axle
loads.
The reinforcement reduced the rut depth even for layers
hX0:5 m, as well as the number of trafcking cycles
possible before reaching the maximum allowed rut depth.
However, time and cost limits meant that the number of
trafcking cycles were limited, and so extrapolation is
necessary to represent in-service conditions. It is recommended that some trafcking be carried out before the
completion of the tracks, in order to cause some rutting
and to mobilise tension in the reinforcement. For ground
with good bearing capacity and CBRX3, reinforcement is
only essential to bridge over weak zones.
4.4. Choice of geosynthetic

Field
Fig. 21. Permanent measured geosynthetic strain vs. strain derived from
deformation of the subgrade.

The use of stiffer geosynthetics in the strain ranges of


13% increased the bearing capacity and compactability of
a ll layer on soft ground. The inclusion of a reinforcement

ARTICLE IN PRESS
R. Hufenus et al. / Geotextiles and Geomembranes 24 (2006) 2137

36

is effective for CBRp3 when hp0:5 m. Geosynthetic


tensile strength requirements at 2% strain (T2%), both in
longitudinal and transverse production directions, should
be:
T 2% X8 kN=m

(2)

The stiffness of the geosynthetics within a strain range of


13% was not affected signicantly by the construction
process, and no reduction in the tension capacity arose due
to installation. Tensions rose to 815 kN/m by full
embedment in the granular layer, due to interlocking. It
is unnecessary to specify the use of an extremely stiff
geosynthetic because tensions of only 610 kN/m were
mobilised for heights of ll between 0.2 and 0.5 m.
The effect of the geogrids was found to be reduced when
used in direct combination with a separating layer, because
optimal interlocking with the coarse-grained ll layer was
prevented and the grid was able to slide on the geotextile.
Nonetheless, to prevent mixing between the subsoil and the
ll material, a separating layer should be used and the
geogrid should be laid 5 cm above it within the granular
layer, to improve both shear interaction and the bearing
capacity.
4.5. Benefits
The following benets have been identied as a result of
laying a geosynthetic as a reinforcing layer between the ll
and the subsoil.

gratefully acknowledged. Hastag AG, Probst Maveg AG


and Viagroup SA are thanked for their in-kind contribution to the eld trials. G. Laios contribution towards the
correction of this manuscript is much appreciated.

Reduction of the thickness of the ll layer by E30% for


specied compaction values and bearing capacities,
although a minimum ll thickness of h 0:3 m should
be recommended.
Reduction in the rut formation as a function of the
trafcking, increasing the serviceable life of the track.

Economic advantages of geosynthetic reinforcement lay


primarily in the possibility to reduce the thickness of the ll
layer or to limit the amount of subgrade to be removed.
This saves on the use of granular materials and the amount
of unsuitable material for removal and deposition elsewhere, which has both economical and ecological aspects,
and may be useful for construction tracks laid without
asphalt surfaces.
Acknowledgements
The authors thank the Swiss Federal Roads Authority
for nancial support and for permission to publish the
results. BP Amoco Fabrics, Fritz Landolt AG, Tensar Int.
GmbH, Huesker Synthetic GmbH, Sytec Bausysteme AG,
Polyfelt GmbH. and Naue Fasertechnik GmbH & Co KG
kindly provided the geosynthetics and additional nancial
support. The contribution of G. Feltrin, K. Weingart, R.
Rohr, P. Nater, H. Kung, U. Schrade, P. Barbadoro and
V. Keller in the eld and with the laboratory tests is

References
Al-Qadi, I.L., Appea, A.K., 2003. Eight-year of eld performance of a
secondary road incorporating geosynthetics at the subgrade-base
interface. 82nd Annual Meeting, Transportation Research Board.
Washington, CD-Rom, 21pp.
Al-Qadi, I.L., Brandon, T.L., Valentine, R.J., Lacina, B.A., Smith, T.E.,
1994. Laboratory evaluation of geosynthetic-reinforced pavement
sections. Transportation Research Report 1439, pp. 2531.
Anderson, P., Killeavy, M., 1989. Geotextiles and geogridscost effective
alternate materials for pavement design and construction. Geosynthetics Conference, vol. 2. San Diego, pp. 353364.
Ashpiz, E.S., Diederich, R., Koslowski, C., 2002. The use of spunbonded geotextile in railway track renewal St. PetersburgMoscow.
Seventh International Conference on Geosynthetics, vol. 3. Nice,
pp. 11731176.
Bathurst, R.J., Allen, T.M., Walters, D.L., 2002. Short-term strain and
deformation behavior of geosynthetic walls at working stress conditions. Geosynthetics International 9 (56), 451482.
Bauer, A., 1997. Der Einuss der Verbundwirkung zwischen Boden und
Geotextil auf das Verformungverhalten von bewehrten Steilboschungen. Ph.D. Thesis, Issue 26, Lehrstuhl und Prufamt fur Grundbau,
Bodenmechanik und Felsmechanik, Technical University, Munich.
Bloise, N., Ucciardo, S., 2000. On site test of reinforced freeway with highstrength geosynthetics. Second European Geosynthetics Conference,
vol. 1. Bologna, pp. 369371.
Bourdeau, P.L., 1991. Membrane action in a two-layer soil system
reinforced by geotextile. Geosynthetics Conference, vol. 1. Atlanta,
pp. 439453.
Cancelli, A., Montanelli, F., 1999. In-ground test for geosynthetic
reinforced exible paved roads. Geosynthetics Conference, vol. 2.
Boston, pp. 863878.
Cancelli, A., Montanelli, F., Rimoldi, P., Zhao, A., 1996. Full scale
laboratory testing on geosynthetics reinforced paved roads. International Symposium on Earth Reinforcement. Fukuoka, pp. 573578.
Chan, F., Barksdale, R.D., Brown, S.F., 1989. Aggregate base reinforcement of surfaced pavements. Geotextiles and Geomembranes 8 (3),
165189.
Collin, J.G., Kinney, T.C., Fu, X., 1996. Full scale highway load test of
exible pavement systems with geogrid reinforced base courses.
Geosynthetics International 3 (4), 537549.
EN ISO 10319, 1996. GeotextilesWide-width tensile test. European
Committee for Standardization.
Espinoza, R.D., 1994. Soilgeotextile interactionevaluation of membrane support. Geotextiles and Geomembranes 13 (5), 281293.
FGSV, 1997. Zusatzliche Technische Vertragsbedingungen und Richtlinien fur Erdarbeiten im Strassenbau. ZTVE-StB 94, Forschungsgesellschaft fur Strassen- und Verkehrswesen.
Floss, R., 2001. Verdichtungstechnik im Erdbau und Verkehrswegebau.
Bomag, Boppard, 149 pp.
Floss, R., Gold, G., 1994. Causes for the improved bearing behaviour of
the reinforced two-layer system. Fifth International Conference on
Geotextiles, Geomembranes and Related Products, vol. 1. Singapore,
pp. 147150.
Garcin, P., Murray, H., 2003. Hochfester Verbundstoff zur Stabilisierung
einer Arbeitsplattform auf organischem Untergrund. 8. Tagung
Kunststoffe in der Geotechnik, Munich, pp. 193196.
Ghosh, C., Madhav, M.R., 1994. Reinforced granular llsoft soil
systemmembrane effect. Geotextiles and Geomembranes 13 (11),
743759.

ARTICLE IN PRESS
R. Hufenus et al. / Geotextiles and Geomembranes 24 (2006) 2137
Giroud, J.P., Noiray, L., 1981. Geotextile-reinforced unpaved road design.
Journal of the Geotechnical Engineering Division, ASCE 107 (GT9),
12331254.
Gobel, C., Lieberenz, K., 1997. Beeinussung des Tragverhaltens von
Schichtsystemen durch Geokunststoffe. 5. Tagung Kunststoffe in der
Geotechnik, Munich, pp. 6167.
Gobel, G.H., Weisemann, U.C., Kirschner, R.A., 1994. Effectiveness of a
reinforcing geogrid in a railway subbase under dynamic loads.
Geotextiles and Geomembranes 13 (2), 9199.
Haas, R., Walls, J., Carroll, R.G., 1988. Geogrid reinforcement of
granular bases in exible pavements. Transportation Research Report
1188, 1927.
Hirano, I., Itoh, A., Itoh, M., Kawahara, S., Shirasawa, M., Shimizu, H.,
1990. Test on trafcability of a low embankment on soft ground
reinforced with geotextiles. Fourth International Conference on
Geotextiles, Geomembranes and Related Products, vol. 1. Den Haag,
pp. 227232.
Houlsby, G.T., Jewell, R.A., 1990. Design of reinforced unpaved roads for
small rut depths. Fourth International Conference on Geotextiles,
Geomembranes and Related Products, vol. 1. Den Haag, pp. 171176.
Hufenus, R., Ruegger, R., Flum, D., Jaecklin, F., Brinkmann, A., Zeiter,
P., Sterba, I., 2002. Anforderungen an Geokunststoffe mit den
Aufgaben Bewehren und Schutzen. Forschungsbericht 1004 (VSS
1999/124), Bundesamt fur Strassen, Bern, 153 pp.
Hufenus, R., Ruegger, R., Weingart, K., Springman, S.M., Mayor, P.,
Banjac, R., Bronnimann, R., Feltrin, G., 2004. Reinforcing foundation
layers on soft subgrade. Third European Geosynthetics Conference,
vol. 1. Munich, pp. 255260.
Huntington, G., Ksaibati, K., 2000. Evaluation of geogrid-reinforced
granular base. Geotechnical Fabrics Report, January/February,
pp. 2228.
Ingold, T.S., 1994. The Geotextiles and Geomembranes Manual. Elsevier,
Oxford, 610pp.
Izvolt, L., Turinic, L., Baslik, B., 2001. Geogrid reinforced subgrade
intstead of traditional solutions in the railway track foundation.
Geosynthetics Conference. Portland, pp. 2336.
Jaecklin, F.P., Floss, R., 1988. Methode zur Bemessung von Geotextilien
im Strassenbau auf besonders weichem Untergrund. 1. Tagung
Kunststoffe in der Geotechnik, Hamburg, pp. 6976.
Jenner, C.G., Paul, J., 2000. Lessons learned from 20 years experience of
geosynthetic reinforcement on pavement foundations. Second European Geosynthetics Conference, vol. 1. Bologna, pp. 421425.
Jenner, C.G., Watts, G.R.A., Blackman, D.I., 2002. Trafcking of
reinforced, unpaved subbases over a controlled subgrade. Seventh
International Conference on Geosynthetics, vol. 3. Nice, pp. 931934.
Kenny, M.J., 1998. The bearing capacity of a reinforced sand layer
overlying a soft clay subgrade. Sixth International Conference on
Geosynthetics, vol. 2. Atlanta, pp. 901904.
Kinney, T.C., Xiaolin, Y., 1995. Geogrid aperture rigidity by in-plane
rotation. Geosynthetics Conference, vol. 2. Nashville, pp. 525537.
Knapton, J., Austin, R.A., 1996. Laboratory testing of unpaved
roads. International Symposium on Earth Reinforcement. Fukuoka,
pp. 615618.
Koerner, R.M., 1997. Designing with Geosynthetics, fourth ed. PrenticeHall, Englewood Cliffs, NJ 761pp.
Mannsbart, G., Magnus, M., Risse, J., 1999. Baustrassen auf geokunststoffbewehrtem PolsterErfahrungen mit der EBGEO. 6. Tagung
Kunststoffe in der Geotechnik, Munich, pp. 257259.
Martin, D., 1988. Die Trennfunktion der Geotextilien in ungebundenen
Verkehrswegebefestigungen. 1. Tagung Kunststoffe in der Geotechnik,
Hamburg, pp. 7786.
Meyer, N., Elias, J.M., 1999. Dimensionierung von Oberbauten von
Verkehrsachen unter Einsatz von multifunktionalen Geogrids zur

37

Stabilisierung des Untergrundes. 6. Tagung Kunststoffe in der


Geotechnik, Munich, pp. 261268.
Miura, N., Sakai, A., Taesiri, Y., Yamanouchi, T., Yasuhara, K., 1990.
Polymer grid reinforced pavement on soft clay ground. Geotextiles and
Geomembranes 9 (1), 99123.
Moghaddas-Nejad, F., Small, J.C., 1996. Effect of geogrid reinforcement
in model track tests on pavements. Journal of Transportation
Engineering 11/12, 468474.
Palmeira, E.M., Cunha, M.G., 1993. A study on the mechanics of
unpaved roads with reference to the effects of surface maintenance.
Geotextiles and Geomembranes 12 (2), 109131.
Perkins, S.W., Ismeik, M., 1997. A synthesis and evaluation of
geosynthetic-reinforced base layers in exible pavementspart 1.
Geosynthetics International 4 (6), 549604.
Perkins, S.W., Ismeik, M., Fogelsong, M.L., Wang, Y., Cuelho, E.V.,
1998. Geosynthetic-reinforced pavementsoverview and preliminary
results. Sixth International Conference on Geosynthetics, vol. 2.
Atlanta, pp. 951958.
Perkins, S.W., Ismeik, M., Fogelsong, M.L., 1999. Inuence of
geosynthetic placement position on the performance of reinforced
exible pavement systems. Geosynthetics Conference, vol. 1. Boston,
pp. 253264.
Posposil, K., Zednik, P., 2002. Geosynthetics impact recognition on soil
bearing capacity in the geotechnical laboratory testing eld. Seventh
International Conference on Geosynthetics, vol. 1. Nice, pp. 419421.
Saathoff, F., Horstmann, J., 1999. Geogitter als Bewehrung in ungebundenen mineralischen SchichtenTeil 1. Strassen- und Tiefbau (9),
1622.
Schad, H., 2001. Erhohung der Tragfahigkeit ungebundener Tragschichten uber nicht ausreichend tragfahigem Erdplanum durch Bewehrungslagen aus Geokunststoffen. Forschungsbericht 05.105G951, BAST,
Bergisch Gladbach.
Sellmeijer, J.B., 1990. Design of geotextile reinforced paved roads and
parking areas. Fourth International Conference on Geotextiles,
Geomembranes and Related Products, vol. 1. Den Haag, pp. 177182.
Som, N., Sahu, R.B., 1999. Bearing capacity of a geotextile-reinforced
unpaved road as a function of deformationa model study.
Geosynthetics International 6 (1), 117.
Springman, S.M., Balachandran, S., 1994. Performance of a woven
geotextile-reinforced retaining wall in the centrifuge. Fifth International Conference on Geotextiles, Geomembranes and Related
Products, vol. 1. Singapore, pp. 251254.
Su, Q., Cai, Y., Zhou, H.B., 2002. Geogrid- and geocell-reinforced sand
blanketlarge-scale model test and the ability to reduce deformation.
Seventh International Conference on Geosynthetics, vol. 1. Nice,
pp. 427430.
Walters, D.L., Raymond, G.P., 1999. Monotonic loading of geogridreinforced nite depth granular material. Geosynthetics Conference.
vol. 1. Boston, pp. 265278.
Walters, D.L., Allen, T.M., Bathurst, R.J., 2002. Conversion of
geosynthetic strain to load using reinforcment stiffness. Geosynthetics
International 9 (56), 483523.
Watn, A., Sognen, H., Emdal, A., 1996. Improvement of bearing capacity
for trafc areas on soft subsoillarge scale laboratory testing. First
European Geosynthetics Conference. Maastricht, pp. 467472.
Watts, G.R.A., Blackman, D.I., Jenner, C.G., 2004. The performance of
reinforced unpaved sub-bases subjected to trafcking. Third European
Geosynthetics Conference, vol. 1. Munich, pp. 261266.
Wilmers, W., 1999. Geotextilien und Geogitter unter Tragschichten. 6.
Tagung Kunststoffe in der Geotechnik, Munich, pp. 251255.
Zia, N., Khan, A.A., Fox, P.J., 2001. Pavement subgrade stabilization
using geogrid reinforcement. Geosynthetics Conference. Portland,
pp. 437450.

Potrebbero piacerti anche