Вы находитесь на странице: 1из 27

Geotextiles and Geomembranes 16 (1998) 45—71

Shear strength properties of


geomembrane/geotextile interfaces
D. Russell V. Jones!,*, Neil Dixon"
! Golder Associates (UK) Ltd., Landmere Lane, Edwalton, Nottingham NG12 4DG, UK
" Department of Civil and Structural Engineering, The Nottingham Trent University, Burton Street, Nottingham
NG1 4BU, UK

Abstract

This paper presents the results of a series of laboratory tests carried out to investigate the
factors controlling the interface shear strength between geomembranes and geotextiles. Several
factors are identified including polymeric composition of both geosynthetics and cover soils
used above the geotextile. Results of both 300 mm direct shear tests and ring shear tests are
presented and the importance of strain softening behaviour and residual interface shear
strengths are discussed. The effect of interface shear strengths on the stability of a geosynthetic
lined landfill is highlighted. As a result of the many factors influencing interface shear strength,
it is proposed that site specific testing should be carried out for detailed design pur-
poses. ( 1998. Published by Elsevier Science Ltd.

Keywords: Interface shear; Laboratory investigation; Landfills

1. Introduction

The design of modern landfills has developed from the old dilute and disperse
philosophy to state-of-the-art containment facilities. Landfills are now designed on
a fully contained basis to prevent the migration of leachate and uncontrolled escape of
landfill gas. In the UK, guidance is provided on the design of new landfills in Waste
Management Paper 26B (DoE, 1995). The overall approach is based on risk assess-
ment rather than prescription with the onus on the designer to be guided by the state
of knowledge at the time of the design. In order to meet the performance objectives for
a landfill liner, geomembranes are generally used either as a single liner, or forming
part of a composite lining system.

*Corresponding author.

S0266—1144/98/$19.00 ( 1998. Published by Elsevier Science Ltd.


PII S 0 2 66 — 11 4 4( 9 7 ) 10 0 22 — X
46 D.R.V. Jones, N. Dixon/Geotextiles and Geomembranes 16 (1998) 45—71

There has been a move away from the use of sand as a drainage and protection layer
above the geomembrane due to concerns about clogging, and thick, non-woven
geotextiles are increasingly being used to afford the protection to the geomembranes
with gravel layers forming the drainage layer. There is, however, little information
published on the interface shear strength between geomembranes and geotextiles.
This paper presents the results of a series of laboratory tests carried out to
investigate the factors controlling the interface shear strength between geomembranes
and geotextiles. It considers the influence of geomembrane texturing, geotextile
polymer and sheet thickness, together with the cover soil.

2. Background
One of the most critical aspects of landfill design is the side slope lining system,
which usually comprises a combination of natural and geosynthetic materials. Work
carried out in Germany (Brune et al., 1991) demonstrated that the use of sand as
a leachate drainage layer can lead to biological and chemical clogging, and recom-
mended a single sized gravel drainage blanket. This is being increasingly adopted as
the preferred solution, with a geotextile layer used to protect the geomembrane from
damage caused by the drainage layer.
The overall stability of such a system and the integrity of the geosynthetics are both
dependent on the shear strength at the interface between the various materials. There
have been many examples worldwide of landfill slope failures associated with low
shear strength interfaces but most are not reported in the literature for political
reasons. The notorious exception to this is the 1988 Kettleman Hills failure which has
been extensively reported, e.g. (Seed et al., 1988; Mitchell et al., 1990; Byrne et al.,
1992). This failure occurred primarily at the interface between the clay and smooth
geomembrane of the secondary lining system, with sliding in the upper part of the
sideslopes occurring along the primary geomembrane/secondary geotextile interface.
The stability of the lining system can be regarded as ensuring there is no uncon-
trolled slippage between the components of the system. Such slippage may produce
excessive local stressing on the geosynthetic and lead to tearing, or may induce
a global slope failure. Several methods of analysing slopes containing geosynthetics
have been proposed including the limiting equilibrium methods of Giroud and Beech
(1989) and Koerner and Hwu (1991), the limit method proposed by Koerner (1994),
and more recently a displacement compatibility method and a strain compatibility
method developed by Long et al. (1994).
The limiting equilibrium method of Giroud and Beech (1989) has been used
extensively and this approach divides the system into two wedges and balances forces
in the vertical and horizontal directions. The resistance to failure is provided by
mobilised soil resistance at the toe, mobilised interface friction along the bottom of the
potential sliding surface and a mobilised tensile load in the geosynthetics above the
plane of sliding. Since this method provides two equilibrium equations and three
unknowns, an iterative process is required to provide a solution. A major drawback in
this method is that the distribution of tensile loads within the geosynthetic layers
cannot be determined.
D.R.V. Jones, N. Dixon/Geotextiles and Geomembranes 16 (1998) 45—71 47

In the UK, the limit method has been used widely and in this method the tensile
loads within individual components of a layered system are determined by a force
equilibrium procedure which balances forces in a direction parallel to the slope.
A layered system resists shear stresses applied to the top layer of the system by
a combination of shear and tensile load in each structural component. This method
gives high tensile loads in the geosynthetics because of the assumption of no contribu-
tion to stability form either the cover soil buttress or waste buttress.
Two methods have recently been proposed by Long et al. (1994) which satisfy both
force and displacement compatibility. One method approximates a multi layered
system as an axially loaded composite column, while the other models each layer in
the composite system as a column connected by interface elements to the adjacent
column and includes non-linear axial and interface behaviour. The first is a simple
analytical method, while the second is a more rigorous numerical approach for
determining tensile loads in the geosynthetic components.
In back analysing the Kettleman Hills failure, Byrne (1994) identified the import-
ance of strain softening interfaces, i.e. the decrease in interface shear strength with
increasing displacement. He concluded that conventional limiting equilibrium
techniques could give misleading and non-conservative assessments of stability.
Depending on the geometry and the distribution of strains within the lining system,
progressive failure may occur. In such circumstances, using a factor of safety on the
peak shear strength to take into account the mobilised shear strength would not give
a conservative assessment of stability. The work of Byrne (1994) and Long et al. (1994)
has led to the need to characterise the full stress/strain behaviour of interfaces in
laboratory tests, in order to obtain the relevant parameters for use in these more
rigorous analysis techniques.
A modified Bromhead ring shear apparatus has been used by Stark and Poeppel
(1994) to investigate the effect of large displacements on various interface shear
strengths. They state that the most important aspect of landfill stability analysis is the
measurement and selection of the interface shear strength, and also suggest that since
the effect of progressive failure is unknown that a factor of safety greater than unity
should be achieved in design using residual shear strength parameters.
Recent work (Stark et al., 1996) has focused on the geomembrane—geotextile
interface tested in the ring shear apparatus. Five 1.5 mm thick geomembranes and five
non woven geotextiles with mass per unit areas of up to 540 g/m2 have been tested.
While this work is of considerable interest, the geotextiles tested are more applicable
to the North American practice whereas 750 g/m2 geotextiles and above are generally
required for use in European designs. Further, this work does not consider the
influence of the cover soil on the geomembrane—geotextile interface performance.

3. Test apparatus and procedure

3.1. 300 mm direct shear apparatus

The 300 mm Direct Shear Apparatus (DSA) used in this study comprises a 305 mm
square top box and a 305 mm]406 mm lower box. It has a maximum travel of
48 D.R.V. Jones, N. Dixon/Geotextiles and Geomembranes 16 (1998) 45—71

102 mm with no loss in area of the shear plane. Normal load is applied using a flexible
diaphragm inflated using compressed air; the pressure being controlled by a precision
regulator. The displacement of the lower box is controlled by a precise DC motor
control system, with the horizontal movement monitored by a Linear Variable
Differential Transformer (LVDT). The horizontal load required to maintain the
chosen shearing rate is measured by a load cell and displayed on a digital transducer
readout. To enable a continuous record of the test a personal computer is used to log
the data.

3.2. Ring shear apparatus

The torsional ring shear apparatus (RSA) used for the experimental programme is
a modified Bromhead ring shear (Bromhead, 1979). The base plate is rotated by means
of a variable speed motor and gearbox through a worm drive. Torque transmitted
along the geomembrane/geotextile interface is reacted by a pair of matched proving
rings bearing on a cross arm. Transducers are placed in parallel with the proving ring
dial gauges, and in series with the vertical dial gauge, for automatic data acquisition.
Recording of the data is achieved using a data logger, and a personal computer is used
to control the data logger and store and manipulate the information.

3.3. DSA testing procedure

The geomembrane samples were cut into rectangles for testing 480 mm]360 mm,
either in the machine direction or at right angles to it. The effect of surface moisture on
the interface shear strength has been described by Seed et al. (1988). All geomembrane
samples were wiped with a hand towel prior to testing. A nylon block was placed in
the lower container to provide a bearing for the geomembrane. The geomembrane
was placed onto the block and clamped to the lower container using four bolts and
a spreader bar.
The geotextiles used were cut either in the machine direction or at right angles to it,
and holes were punched for clamping. The geotextiles were placed on top of the
geomembranes and clamped to the top box. For textured geomembranes, a piece of
thin plastic sheet was placed between geotextile and geomembrane to prevent any
disturbance of the geotextile fibres, and removed prior to lowering the top container
into position.
Screws were used to lift the top box clear of the geomembrane some 0.7 mm to take
into account crushing of the geotextile under the load of the top container. Gravel was
then placed into the top container in two 50 mm layers and compacted by light
tamping using a 5 kg weight. It was considered that this prevented arching in the
container and gave the gravel a similar initial compaction to that experienced on site
after placement.
Tests were carried out on each interface at normal stresses of 25, 50, 100 and
200 kPa with virgin geosynthetic samples used for each of the normal stresses. The
normal stress was held for 5 min and adjusted if required, before shearing commenced.
A shearing rate of 3 mm/min was used for all tests. Displacements and shear forces
D.R.V. Jones, N. Dixon/Geotextiles and Geomembranes 16 (1998) 45—71 49

were recorded at specified time intervals, usually 5 s intervals for the first 25 min to
ensure that the peak shear force was measured, followed by 30 s intervals for the
remainder of the test. Once the maximum displacement of around 100 mm was
achieved, the soil was carefully removed from the container and the geosynthetic
samples unclamped. Once the geomembrane has been removed the failure surface was
inspected and any stretching or damage to either geosynthetic was recorded.

3.4. RSA testing procedure

Annular samples, 100 mm OD and 70 mm ID, were formed from the various
geomembrane sheets by firstly cutting to within 1mm of the final profile with a band
saw, and then trimmed to size by turning on a lathe. As with the direct shear samples,
the surfaces were towel dried prior to testing. All geotextiles were punched from sheets
into the same annular shape as for the geomembrane. The samples of geotextiles were
glued onto wood with care taken to ensure that the glue did not extrude through the
geotextile.
The geomembrane samples were secured to the top platen in two ways. For the
textured geomembranes four small holes were drilled into the ring at locations
corresponding to screw holes in the top plate, counter sunk, and screwed into place.
This method was not suitable for the smooth geomembrane since the screw heads
would affect the surface friction. Consequently, the geomembrane was glued to a thin
wooden ring which had four nails protruding at the locations of the screw holes. This
did not attach the ring to the top platen but prevented slippage during shearing.
A wooden insert was manufactured and placed inside the bottom container, with
the geotextile glued to the wooden ring. When in position, the top of the wooden ring
was 1mm above the top surface of the bottom container to ensure that there was no
friction between the geotextile and the bottom container. The wooden ring was
located in the bottom container by four small nails driven into the ring. These were
placed into the drainage holes in the bottom of the container to ensure the ring, and
hence the geotextile, would rotate as part of the bottom container. A shearing rate of
3 mm/minute and normal stresses of 27, 52, 102 and 201 kPa were used for this series
of tests to enable comparison with the DSA results. The force measured by each of the
proving ring transducers and the vertical displacement were recorded at 10 min
intervals for the duration of the tests. Typical total displacements in the order of 3 m
were achieved.

4. Description of materials

4.1. Geomembranes

Three high density polyethylene (HDPE) geomembranes were used for most of the
testing; one smooth and two textured. Additional tests were carried out using
a smooth polypropylene (PP) geomembrane to compare with the smooth HDPE.
Both smooth geomembranes were manufactured by the blown film process. One of
the textured geomembranes used was manufactured in Germany, with the texturing
50 D.R.V. Jones, N. Dixon/Geotextiles and Geomembranes 16 (1998) 45—71

produced by impinging hot polyethylene particles on the surface of previously


manufactured smooth sheets. The temperature of the sheet and the temperature and
energy of the particles give rise to an adhesion of the particles. The size and density of
the particles controls the degree of texturing. Impingement of the particles is a second-
ary manufacturing operation rather than being simultaneous and the edges of the rolls
are left nontextured to aid with the seaming process.
The second textured sheet used was a coextruded geomembrane manufactured
using a blown film process in the US. It is produced using small extruders in addition
to the main geomembrane core extruder, where these side extruders have the same
polyethylene formulation as the main extruder but also have nitrogen gas added to
the polymer melt. As the melt exits the die, it rapidly cools and the gas expands to the
outer surface where there is least resistance. It eventually bursts leaving a roughened,
or textured surface behind.

4.2. Geotextiles

Three geotextiles commonly used in Europe as protection layers between the


granular drainage layer and the geomembrane were used in the testing programme.
They comprised 750 and 1200 g/m2 PP geotextiles and a 800 g/m2 HDPE geotextile,
all were non-woven and produced by a needle punching technique. In this process
mechanical bonding of the staple fibres is achieved by passing a loose web of fibres
beneath a bank of reciprocating barbed needles which penetrate the full thickness of
the web. As the needle enters the web, it drags some of the staple fibres down into the
body of the web, causing them to interlace with other fibres.

4.3. Soils

Three soils were used in the testing programme. Soil A was a poorly graded
rounded to subrounded medium GRAVEL with occasional fine gravel. Soil B had
a similar grading to Soil A however, it was a crushed stone and hence comprised
subangular particles which enabled an investigation of particle shape on interface
shear strength. Soil C was also a crushed stone but comprised a well graded fine
GRAVEL and coarse SAND with some fine and medium sand. Since this had the
same particle shape as Soil B, the soil was used to investigate the effect of drainage
material particle size on the interface shear strength.

5. Results

5.1. Smooth geomembrane and textured geomembrane

The behaviour of a smooth HDPE geomembrane PP geotextile interface using Soil


B as cover soil above the geotextile is shown in Fig. 1(a). There is an initial increase in
shear stress as soon as displacement starts, with a rapid increase in shear stress with
increasing displacement, followed by a loss of shear stress with further deformation.
D.R.V. Jones, N. Dixon/Geotextiles and Geomembranes 16 (1998) 45—71 51

Fig. 1. (a) Shear stress vs displacement—smooth geomembrane. (b) Shear stress vs displacement—textured
geomembrane.

This strain softening behaviour is evident at all four normal stresses applied. For
smooth geomembranes the peak shear stress occurs typically at displacements of less
than 2 mm, with the shear stress reducing by 20—30% at displacements of around
40—50 mm. Little further change in shear stress with increasing displacement is evident.
52 D.R.V. Jones, N. Dixon/Geotextiles and Geomembranes 16 (1998) 45—71

Fig. 2. Shear stress vs displacement—Soil A cover.

The shear stress vs displacement behaviour of a textured geomembrane against the


same geotextile with Soil B as cover is shown in Fig. 1(b). Much larger displacements
are required to mobilise peak shear stresses, which are significantly higher than for
smooth geomembranes, however a greater decrease in shear stress from peak is
experienced with increasing displacement. Peak shear stresses are reached at displace-
ments of between 5 and 10 mm, depending on the normal stress applied. The textured
geomembrane—geotextile interface exhibits a higher degree of strain softening than the
smooth geomembrane with around 50% reduction of shear strength at displacements
over 50 mm. Both types of textured geomembrane gave similar shear stress vs
displacement curves.

5.2. Cover soils

The three cover soils used in this study all showed similar shear stress vs displace-
ment behaviour for the smooth geomembrane—geotextile interface. The use of sub-
rounded gravel as a cover soil (Soil A), however, increased the peak shear strength of
the textured geomembrane—geotextile interface, see Fig. 2. Peak stresses are higher for
all four normal stresses. The large strain shear stresses are very similar for all three
cover soils, with a high degree of strain softening of around 65%.

5.3. Geotextile thickness/weight

Two polypropylene geotextiles from the same manufacturer were used with weights
of 750 and 1200 g/m2. These are commonly used as geomembrane protectors and
D.R.V. Jones, N. Dixon/Geotextiles and Geomembranes 16 (1998) 45—71 53

were used to assess the influence of geotextile unit weight. The 750 and 1200 g/m2
geotextiles were around 5 and 7 mm thick, respectively. The shape of the shear stress
vs displacement curves is very similar for both geotextiles. The peak shear stress varies
between geomembrane and cover soil used but generally the lower unit weight and
thinner geotextile gives the greatest peak shear stresses.

5.4. Geotextile fibre polymer

The interface shear strength between the smooth geomembrane and the HDPE
geotextile is shown in Fig. 3. The peak shear stresses are reached at small displace-
ments, and are comparable to the peak stresses obtained with the PP geotextile, with
the exception of the 200 kPa normal stress test which is slightly higher. The post peak
behaviour however, is significantly different with very little reduction in shear stress,
and indeed strain hardening occurring under normal stresses of 100 kPa or greater.
The shear stress vs displacement behaviour of the textured geomembrane HDPE
geotextile interface is very similar to the PP geotextiles for both the subangular
gravels used as cover soils. However, the behaviour of this interface when using the
subrounded Soil A is markedly different, see Fig. 4. The flat peak shear stress with
increasing displacement distribution may be attributed to stretching of the HDPE
geotextile; in other words, there is movement at the soil—geotextile interface as well as
the geomembrane—geotextile interface.

Fig. 3. Shear stress vs displacement—smooth geomembrane/HDPE geotextile.


54 D.R.V. Jones, N. Dixon/Geotextiles and Geomembranes 16 (1998) 45—71

Fig. 4. Shear stress vs displacement—textured geomembrane/HDPE geotextile.

5.5. Nylon block cover

The use of a nylon block above the geotextile provides information on the geomem-
brane—geotextile interface shear strength that is independent of the cover soil used.
The shape of the shear stress vs displacement curves for the smooth geomembrane—
PP geotextile interface is not altered by the use of a nylon block. The peak shear
stresses are of the same order as the peaks for both subangular cover soils, which are
both higher than the corresponding subrounded cover soil. However, the shape of the
shear stress vs displacement curves for the textured geomembrane—geotextile interfa-
ces are significantly modified by the use of a nylon block, compare Fig. 5 with
Fig. 1(b). The shear stress increases with displacement at a constant rate for displace-
ments up to around 4 mm, after which there is a distinct change in the shape of the
curves.
It is thought that at this change in gradient of the shear stress vs displacement plot
the effect of the geotextile stretching becomes a dominant factor in the behaviour. The
resulting peak shear stresses are less than when cover soils are used above the
geotextile, and these peaks occur at much higher displacements. The geotextile
stretching was clearly observed and measured on dismantling the apparatus after each
test, with the degree of extension appearing proportional to the measured shear
stresses. The large strain shear stresses were of the same order as the tests with cover
soils. This suggests that the residual interface shear strength is a property of the
geomembrane and geotextile used, and independent of the cover used.
D.R.V. Jones, N. Dixon/Geotextiles and Geomembranes 16 (1998) 45—71 55

Fig. 5. Shear stress vs displacement—nylon block cover.

5.6. Polypropylene geomembrane

The smooth PP geomembrane was tested against the 1200 g/m2 PP geotextile with
Soil C as a cover soil. The shape of the shear strength vs displacement curve does not
follow the same strain softening behaviour of the comparable interface using the
HDPE geomembrane, see Fig. 6. The initial steep rise in shear stress is followed by
a small drop, before a gradual increase to peak values. This gradual rise in shear stress
was not observed for the 25 kPa normal stress. The magnitude of the initial peak
stresses are higher than the peaks observed in the HDPE geomembranes, while the
large strain peak values are significantly higher. Very similar results were also
obtained for the same interface using Soil C as a cover soil. The use of a more angular
cover soil increases the interface shear strength of an HDPE geomembrane but makes
very little difference to the interface shear strength when a PP geomembrane is used.

5.7. Large strain interface shear strengths

A modified Bromhead ring shear apparatus (RSA) was used to investigate the large
displacement interface shear strengths between several HDPE geomembranes and
non-woven geotextiles. The normal stresses used corresponded to the range of normal
stresses used in the direct shear tests. A typical shear stress vs. displacement relation-
ship is shown in Fig. 7 for the results of a textured geomembrane against the 750 g/m2
PP geotextile. For the lowest normal stress, there is a rapid increase in shear stress
with displacement to an initial peak, followed by a reduction in shear stress to a near
56 D.R.V. Jones, N. Dixon/Geotextiles and Geomembranes 16 (1998) 45—71

Fig. 6. Shear stress vs displacement—PP geomembrane.

Fig. 7. Shear stress vs displacement—ring shear test.

constant large strain or residual value. This initial peak interface shear strength can be
regarded as an initial ‘bedding-in’ of the geotextile and geomembrane, i.e. a realign-
ment of fibres along the asperities of the textured geomembrane. No significant peaks
were observed for the higher normal stresses. It should be noted that the normal
D.R.V. Jones, N. Dixon/Geotextiles and Geomembranes 16 (1998) 45—71 57

stresses were increased without dismantling the specimens, and hence one pair of
samples was used to produce each complete failure envelope.

6. Discussion

6.1. Interface shear strength envelopes

The interface shear strength can be expressed as a function of the normal stress by
a modified Mohr—Coulomb criterion:
q "a@#p@ tan d@
&
It is assumed that for sliding to occur at an interface, the shear stress has overcome
a frictional resistance p@ tan d@, which is dependent on the effective normal stress p@
acting on the interface and on a friction angle d@, together with a component a@ which
is independent of the normal stress. This component a@ is often called cohesion but is
more usefully regarded merely as an intercept on the shear stress axis, which defines
the position of the failure envelope.
The nomenclature used in this report is as follows:
d@ peak friction angle of the interface
1
a@ peak apparent cohesion of the interface
1
d@ residual, or large strain friction angle of the interface
3
a@ residual, or large strain apparent cohesion of the interface
3
A summary of strength parameters from the testing programme is presented in
Tables 1—3.
For soils, the failure envelope may show some slight curvature, particularly under
low normal stresses. For some geosynthetic interfaces this curvature may be more
pronounced, however, a straight line approximation can still be taken over the stress
range of interest and the interface shear strength parameters determined for that
range. This procedure has been used in this paper to compare the various interface
shear strengths, and the correlation coefficient, R, is reported for each best fit straight
line. The multiple correlation coefficient squared R2 represents the fraction of the total
variation accounted for by the fitted equation. For example, an R value of 0.995 has
an R2 value of 0.990; this indicates that 99% of the total sum of squares of shear
stresses is accounted for by this equation.
Giroud et al. (1993) proposed an hyperbolic expression for geosynthetic—geosyn-
thetic and geosynthetic/soil interface shear strength envelopes, which can be used
when test results give a significantly non-linear shear stress/normal stress relationship.
Second-order polynomial fits can also be used, as described in Section 6.2.

6.2. Tests with cover soils

The peak and residual shear stresses for the smooth HDPE geomembrane—geotex-
tile interface (Fig. 1(a)) are plotted against the normal stress in Fig. 8. The linear
58 D.R.V. Jones, N. Dixon/Geotextiles and Geomembranes 16 (1998) 45—71

Table 1
Peak interface shear strengths-geomembrane/geotextile (cover soil used)

Geotextile Friction Apparent Correlation


Geomembrane (g/m2) (polymer) Cover soil angle (d@ ) cohesion (a@ ) coefficient (R)
1 1
Smooth 750 PP Soil A 7.2 !2.4 0.999
Soil B 7.5 0.3 0.998
Soil C 8.4 !0.4 0.996
1200 PP Soil A 5.3 1.6 0.998
Soil B 8.1 0.9 1.000
Soil C 7.7 !0.4 0.996
800 HDPE Soil A 6.6 0.8 0.998
Soil B 10.1 !0.8 0.997
Soil C 8.6 !0.1 0.999
Impingement textured 750 PP Soil A 28.5 22.9 0.994
Soil B 23.0 2.5 1.000
Soil C 25.6 26.2 0.996
1200 PP Soil A 26.2 21.8 0.998
Soil B 22.3 3.7 0.999
Soil C 25.4 20.3 0.997
800 HDPE Soil A 29.7 10.4 0.999
Soil B 24.5 3.2 1.000
Soil C 29.2 23.3 0.992
Coextruded textured 750 PP Soil A 23.8 9.6 0.999
Soil B 25.5 4.2 1.000
Soil C 24.1 3.9 0.999
1200 PP Soil A 27.4 5.3 0.993
Soil B 26.4 3.2 0.992
Soil C 23.9 2.8 1.000
800 HDPE Soil A 31.6 !1.8 0.977
Soil B 27.9 4.2 0.999
Soil C 24.6 6.6 0.992

regression fits give a peak friction angle of 7.5° and a residual angle of 6.0°, with
negligible apparent cohesion. Both peak and residual shear strength envelopes pro-
vide excellent straight line fits with R values of 0.998 and above. Texturing the surface
of geomembranes results in much higher shear strengths; for geomembrane—geotextile
interfaces however, the type of texturing has little effect on the shear strength
parameters, see Fig. 9(a) and 9(b). Peak interface shear strengths for geomembrane-
geotextile interfaces, together with the correlation coefficients, are summarised in
Table 1 for the tests with gravel as a cover soil.
Considering the interface between the smooth geomembrane and the three geotex-
tiles. The 750 g/m2 geotextile interface has friction angles of between 7.2 and 8.4°
depending on the cover soil used. The highest friction angle calculated was for Soil C,
the subangular fine gravel and coarse sand, whereas the two medium gravels gave very
similar shear strengths. It is likely that the sand particles act through the geotextile
and increase the friction at the interface, and this suggests that for the thinner PP
D.R.V. Jones, N. Dixon/Geotextiles and Geomembranes 16 (1998) 45—71 59

Table 2
Peak interface shear strengths-geomembrane/geotextile (nylon block)

Geotextile Nylon block Friction Apparent Correlation


Geomembrane (g/m2) (polymer) or mean soil angle (d@ ) cohesion (a@ ) coefficient (R)
1 1
Smooth 750 PP N/B 8.4 3.2 0.990
Mean soil 7.7 0.8 0.993
1200 PP N/B 8.2 0.1 0.999
Mean soil 7.0 0.7 0.966
800 HDPE N/B 9.5 0.0 0.999
Mean soil 8.4 !0.1 0.969
Impingement textured 750 PP N/B 22.2 3.9 0.998
Mean soil 25.8 17.2 0.908
1200 PP N/B 20.7 3.2 1.000
Mean soil 24.7 15.3 0.933
800 HDPE N/B 20.4 5.2 1.000
Mean soil 27.8 12.3 0.938
Coextruded textured 750 PP N/B 21.3 7.4 0.994
Mean soil 24.5 5.9 0.996
1200 PP N/B 22.1 5.8 1.000
Mean soil 25.9 3.8 0.986
800 HDPE N/B 21.0 7.7 0.998
Mean soil 28.1 3.0 0.978

geotextile the predominant criterion for shear strength is the grading of the cover soil.
The thicker PP geotextile, 1200 g/m2, has higher interface shear strengths with both
subangular gravel samples than the subrounded as a result of the greater penetration
provided by more angular soils. Since both Soil B and Soil C have similar friction
angles, 8.1 and 7.7° respectively, it suggests that for the thicker geotextiles the primary
criterion for shear strength is the particle shape. The 800 g/m2 HDPE geotextile
interface with the smooth geomembrane was also dependent on the particle shape
with the medium subangular gravel having a friction angle of 10.1 degrees compared
with the medium subrounded gravel at 6.6°. The subangular fine gravel and coarse
sand has a higher shear strength than Soil A at 8.6 degrees but less than the
subangular medium gravel.
For impingement textured geomembrane, higher peak shear strengths were mobil-
ised with Soil A as cover soil for both the PP geotextiles, while the lowest shear
strengths were recorded for the subangular medium gravel, Soil B. Both Soils A and
C had high apparent cohesions when the geomembrane was tested against all three
geotextiles, while Soil B consistently gave significantly lower cohesion. The apparent
cohesion intercept varied from 2.5 to 3.7 kPa for the subangular medium gravel cover
soil, however this was increased to between 20.3 and 27.2 kPa for the subrounded
medium gravel and the subangular fine gravel and coarse sand, with one value lower
at 10.3 kPa.
This finding is significant since any apparent cohesion that could be used for design
purposes would have a large impact on the slope stability, especially during the initial
60 D.R.V. Jones, N. Dixon/Geotextiles and Geomembranes 16 (1998) 45—71

Table 3
Large strain interface shear strengths—geomembrane/geotextile

Nylon block,
Geotextile mean soil Friction Apparent Correlation
Geomembrane (g/m2) (polymer) or ring shear angle (d@ ) cohesion (a@ ) coefficient (R)
3 3
Smooth 750 PP N/B 6.1 1.3 1.000
Ring shear 4.0 1.9 0.999
Mean soil 6.2 !0.2 0.997
1200 PP N/B 6.5 !0.4 0.999
Ring shear 3.0 1.7 0.999
Mean soil 4.8 !0.1 0.981
800 HDPE N/B NSS NSS —
Ring shear 8.0 !0.2 0.987
Mean soil 6.2 !0.3 0.990
Impingement textured 750 PP N/B 17.0 1.1 0.998
Ring shear 11.7 0.8 1.000
Mean soil 10.5 5.8 0.980
1200 PP N/B 14.2 2.1 1.000
Ring shear — — —
Mean soil 10.0 5.8 0.994
800 HDPE N/B 16.4 2.9 1.000
Ring shear 12.5 4.3 1.000
Mean soil 14.4 2.7 0.976
Coextruded textured 750 PP N/B 15.0 4.1 0.997
Ring shear 10.7 1.2 1.000
Mean soil 10.7 4.1 0.998
1200 PP N/B 14.1 4.8 1.000
Ring shear 9.0 0.8 0.985
Mean soil 11.5 3.7 0.991
800 HDPE N/B 18.5 2.8 1.000
Ring shear 17.0 !0.6 0.999
Mean soil 14.2 3.8 0.998

placement of cover soil. From the majority of tests carried out it appears that the
failure envelopes for impingement textured materials against geotextiles exhibit some
apparent cohesion, but the consistently low cohesion for Soil B should serve as
a warning that great care is required when relying on apparent cohesion for design
purposes.
The coextruded textured geomembrane—geotextile interfaces yield similarly high
friction angles to the impingement textured geomembrane, but with lower values of
apparent cohesion. The peak friction angles for the 750 g/m2 geotextile are all very
similar ranging from 23.8 to 25.5°. The thickest geotextile tested, 1200 g/m2, gave
higher friction angles with Soils A and B than with Soil C, suggesting that larger
particle sizes result in higher shear strength for this particular interface.
Several of the interfaces tested have low correlation coefficients and it may be more
applicable to have curved failure envelopes for these interfaces. The impinged textured
geomembrane—750 g/m2 interface with Soil A cover has a correlation coefficient of
D.R.V. Jones, N. Dixon/Geotextiles and Geomembranes 16 (1998) 45—71 61

Fig. 8. Shear stress vs normal stress—smooth geomembrane.

0.994 for the best-fit straight line through the peak shear stresses, and is plotted in
Fig. 10(a). By inspection, a curved envelope would seem to be more applicable than
a straight line. A second-order polynomial fit through the four data points relates the
shear stress to the normal stress by
q"12.75#0.8232(p@ )!0.0012(p@ )2
/ /
with a correlation coefficient of 1.000. This curve is plotted in Fig. 10(b) and is visibly
a better fit.

6.3. Nylon block tests

From the above, it can be seen that the particle size and shape of cover soil used in
the test affects the peak shear strength mobilised. A limited series of tests were carried
out using a nylon block instead of cover soil, to try and establish the interface friction
between geomembrane and geotextile without the interlock provided by a soil.
Table 2 compares the results of peak shear strengths calculated from the tests using
the nylon block, and the mean shear strengths from the tests with cover soil. For the
smooth geomembrane tests, there is good correlation between the 1200 g/m2 geotex-
tile, having the lowest interface shear strength, and the 750 g/m2 HDPE having the
highest shear strength using both nylon block and mean soil results. The nylon block
tests consistently gave results around one degree higher than the mean soil result, and
this is considered to be due to a large surface contact between smooth geomembrane
and geotextile when a block is used, whereas for cover soils there is a smaller contact
area with higher contact stress.
62 D.R.V. Jones, N. Dixon/Geotextiles and Geomembranes 16 (1998) 45—71

Fig. 9. (a) Shear stress vs normal stress—impingement textured geomembrane. (b) Shear stress vs normal
stress—coextruded textured geomembrane.
D.R.V. Jones, N. Dixon/Geotextiles and Geomembranes 16 (1998) 45—71 63

Fig. 10. (a) Shear stress vs normal stress—linear regression fit. (b) Shear stress vs normal stress—second-
order polynomial fit.
64 D.R.V. Jones, N. Dixon/Geotextiles and Geomembranes 16 (1998) 45—71

For the impingement textured geomembrane, the value taken as the mean apparent
cohesion is not representative since two values were much higher than the third, and
thus cannot be used in the comparison. The friction angles vary far more than the
smooth geomembrane results with differences of between 3.6 and 7.4 degrees between
nylon block and mean soil results. The mean soil results are higher than the nylon
block results which is the opposite of the smooth geomembrane findings. This
suggests that the effect of the soil above the geotextile is more important when
considering textured geomembranes.
The mean soil results are consistently higher than the nylon block results for the
coextruded textured geomembrane also, with friction angles from 3.1 to 7.1 degrees
higher, for a similar range to the impingement textured geomembrane. It can be
concluded that results from tests carried out with no soil above the geotextile are
likely to give conservative values of peak shear strengths for textured geomembranes.

6.4. Large strain interface shear strengths

The strain softening properties of many geosynthetic interfaces means that the
determination of the large strain shear strength is of particular importance. The large
strain or residual shear strength of geomembrane—geotextile interfaces were obtained
in three different ways; using the DSA with gravel cover soil above the geotextile,
using the DSA with a nylon block above the geotextile, and using the RSA. Table 3
summaries these results.
Firstly, consider the smooth geomembrane—geotextile interface. The large strain
friction angles calculated for the 750 g/m2 interface for the two DSA tests were very
similar while the RSA gave a friction angle 2° lower. Similarly for the 1200 g/m2
geotextile, the RSA gave a lower friction angle than both the nylon block and mean
soil DSA tests. The 800 g/m2 HDPE geotextile exhibits different behaviour since it
was not strain softening when used with the nylon block, and the mean soil tests gave
a lower friction angle than the ring shear testing.
The reason for the anomalous behaviour of the 800 g/m2 HDPE geotextile is
related to its HDPE polymer fibres and the manufacturing process, which produces
a greater propensity for stretching. The strength of a needle-punched geotextile is
related to the amount of entanglement produced by the needling and the inter-fibre
friction. Since the HDPE fibres have a less frictional surface than the PP, the
inter-fibre friction is lower and thus the likelihood of stretching is higher. This is
confirmed by the tensile properties of the two geotextiles, with the HDPE geotextile
having lower tensile strength than the PP geotextile of similar construction. Since the
geotextile is glued to the base in the RSA tests, its ability to stretch is limited, and
hence care should be taken when comparing RSA and DSA results.
The results for the impingement textured geomembrane indicate the residual shear
strength of the interface with both the 750 and 1200 g/m2 materials has a friction angle
ranging from 10 to 17° with apparent cohesion values in the region 1—6 kPa. For both
sets of tests, the mean soil tests gave higher values of apparent cohesion than the nylon
block or RSA tests. This is considered to be due to more ‘‘interlock’’ at the geomem-
brane/geotextile interface when a cover soil is used.
D.R.V. Jones, N. Dixon/Geotextiles and Geomembranes 16 (1998) 45—71 65

The coextruded textured geomembrane tests gave a range of large strain friction
angles from 9 to 15° with apparent cohesion values from 0.8 to 4.8 kPa for the 750 and
1200 g/m2. The RSA results also gave the lowest cohesion values for these tests and
this can be attributed to the very large strains experienced during a ring shear test and
the absence of cover soil above the geotextile. Since the apparent cohesion in the peak
shear strength for a textured geomembrane—geotextile interface is attributed to the
interlock between fibres and asperities, it is likely that this will be destroyed at large
displacements. The 800 g/m2 HDPE geotextile has higher residual interface shear
strength with both textured geomembranes with friction angles ranging from 12.5 to
18.5° and apparent cohesions up to 4.3 kPa. As for the smooth geomembrane
interface, this higher shear strength is considered to be a function of stretching of the
HDPE geotextile in the DSA tests.

6.5. Cover soil effects

A geotextile used as a geomembrane protector will invariably have granular


material placed on it as cover. The effect of such a granular layer on the interface shear
strength has been investigated in this paper. For a smooth geomembrane—geotextile
interface finer particle size increase the contact area between the geomembrane and
geotextile and hence the shear strength. Similarly when a nylon block is used as cover
there is an increase in interface shear strength due to the block spreading the normal
stress throughout the surface of the geomembrane, i.e. lower contact stresses. The use
of a block as cover does give higher interface shear strength than when soils are used
and therefore, when assessing the likely interface shear strength mobilised on site, it is
proposed that site specific soils should be used in the laboratory investigation.
Soil A and C gave the highest interface shear strength between textured geomem-
branes and geotextiles, with Soil B giving significantly lower apparent cohesions.
Due to the surface asperities, the predominant factor in mobilising interface shear
strength between a textured geomembrane and a geotextile does not seem to be the
contact area, as is the case for the smooth geomembrane. Apparent cohesion is
a function of the interlock between asperities and fibres and the use of Soil B and the
nylon block as cover mobilised only a fraction of the apparent cohesion mobilised by
Soils A and C.
The use of nylon block as a cover in the textured geomembrane against geotextile
tests consistently gave lower interface shear strength than the tests with a soil cover.
Thus the use of a block would give a conservative assessment of the shear strength of
this interface.

6.6. Geotextile effects

This investigation has revealed little difference between the interface shear strength
behaviour of a 750 and a 1200 g/m2 PP geotextile produced by the same manufac-
turer. The calculated friction angle of the heavier geotextile was consistently 1—2°
lower than the lighter geotextile and probably relates to the ability of individual
particles to penetrate into the geotextile and hence act through it. A thicker, stronger
66 D.R.V. Jones, N. Dixon/Geotextiles and Geomembranes 16 (1998) 45—71

Table 4.

Polymer Weight (g/m2) Tensile strength (kN/m)

PP 750 24 MD (machine direction)


65 XMD (cross machine direction)
PP 1200 32 MD
90 XMD
HDPE 800 17 MD
32 XMD

geotextile will afford more protection to the geomembrane, and hence restrict the
penetration of the soil particles.
The peak shear stresses mobilised for HDPE geotextiles are comparable to interface
shear stresses mobilised using PP geotextiles. However, the post peak behaviour of the
HDPE geotextile was significantly different and is attributed to stretching of the
geotextile. The propensity to stretching is related to the fibre needling and the tensile
strength of the geotextile. Results of wide width tensile tests carried out by the
geotextile manufacturer in accordance with BS 6906: Part 1:1991 can be summarised
as in Table 4.
The direction of the geotextile sample during testing, whether PP or HDPE, was
investigated further using the 750 g/m2 PP geotextile and revealed that the weaker
machine direction samples stretched significantly more than the stronger cross ma-
chine direction samples. Fig. 11(a) and 11(b) show the shear stress vs. displacement
plots for a textured geomembrane against a 750 g/m2 geotextile with the geotextile
tested in the cross machine direction and machine direction, respectively. When tested
in the cross machine direction there is a certain amount of stretching between the first
and second peaks, whereas the machine direction continues to stretch for most of the
displacement.

6.7. Comparison of peak and residual strengths

The loss of interface shear strength with displacement can be quantified by consid-
ering the ratio of residual to peak friction angle, D, where

tan d@
D" r
tan d@
1
For the smooth geomembrane—geotextile interfaces the 750 g/m2 geotextile gave
D values ranging from 0.72 to 0.87 and the 1200 g/m2 geotextile gave values from 0.65
to 0.79. This suggests that the use of a heavier, thicker geotextile results in greater
strain softening effects with a mean D value of 0.7 compared with a mean of 0.8 for the
thinner geotextile. Both textured geomembranes showed similar degrees of strain
softening with both PP geotextiles with D values ranging from 0.32 to 0.49 and an
average value of 0.4.
D.R.V. Jones, N. Dixon/Geotextiles and Geomembranes 16 (1998) 45—71 67

Fig. 11. (a) Shear stress vs displacement—geotextile cross machine direction. (b) Shear stress vs displace-
ment—geotextile machine direction.

Using a nylon block cover has little effect on the smooth geomembrane—geotextile
interface shear strength reduction with a D value of 0.76. For the textured geomem-
brane—geotextile interface tests using the nylon block cover the mean D value is 0.68;
much higher than the mean values with cover soil. This is because generally the nylon
68 D.R.V. Jones, N. Dixon/Geotextiles and Geomembranes 16 (1998) 45—71

block results have low peak strengths and high residual strengths, thus the strain
softening effect is not as pronounced.
It is suggested that the D values could be used at preliminary design stage to
estimate the residual friction angle based on knowledge of the peak friction angle from
the literature. For this purpose, D values of 0.7 and 0.4 could be used for smooth and
textured geomembrane—geotextile interfaces, respectively. It should be noted however
that this approach does not take into account any apparent cohesion that the
interface may have.

6.8. Stability analysis

A rigorous assessment of the stability of a landfill lined with geosynthetics needs


an understanding of the interface shear strengths between the various layers. Since
the interfaces will invariably be strain softening, the anticipated deformation at
each interface is required to establish the mobilised shear stresses. For this reason the
full shear stress—displacement curve is required for the level of anticipated normal
stress.
The loading conditions will vary throughout the design life of the landfill with the
following main considerations:

f Short-term loading immediately post-construction with the relatively low normal


stress due to protection/drainage materials only.
f Intermediate loading as waste placement proceeds.
f Long-term loading under full height of waste and the effects of waste settlement.

For less detailed analysis, at outline design stage when actual materials to be
used are usually not known, a more simplistic approach is widely used. In such
cases, the interface shear strength is considered in terms of the friction angle
and apparent cohesion. The choice of whether to use the peak values, residual
values or factored values to take into account the mobilised shear strength
becomes critical. In some instances residual values may be appropriate on the side
slope where large displacements are anticipated, used together with peak values on the
base.
The BS 6906 approach, which is currently also adopted in the draft European
Standard, to assessing the interface shear strength requires that the apparent
cohesion is ignored if positive. This assumption can have a significant effect in
that the shear stress for any particular normal stress will be quoted as being
lower than measured. In particular, residual interface shear strength envelopes
such as the example shown in Fig. 10a would be drastically reduced. It is possible
that the failure envelope may curve to the origin at very low normal stresses, in which
case ignoring the apparent cohesion will result in overconservative results. In the
authors’ opinion, both friction angles and apparent cohesions should be measured
and reported but should only be used in analysis within the range of normal stresses
tested.
D.R.V. Jones, N. Dixon/Geotextiles and Geomembranes 16 (1998) 45—71 69

7. Conclusions

This paper describes a series of direct shear and torsional ring shear tests on
geomembrane—geotextile interfaces. From the results of laboratory testing, the follow-
ing conclusions can be drawn:

f The shape of the shear stress vs displacement curves for the smooth geomembrane
are inherently different from the textured geomembrane—geotextile interface. The
smooth geomembrane peak interface shear stresses are mobilised almost immedi-
ately at very low displacements, usually less than 2 mm. For textured geomem-
branes, peak stresses are obtained at displacements of between 5 and 10 mm for the
normal stress ranges considered.
f Peak interface shear strengths are significantly higher for textured geomembranes,
however, these interfaces have a greater degree of strain softening than the smooth
geomembrane—geotextile interfaces.
f Smooth polypropylene geomembranes have higher interface shear strengths
against geotextiles than high-density polyethylene geomembranes. Further, these
interfaces did not undergo strain softening in the small number of tests carried
out.
f Cover soil particle shape and grading influence the interface shear strengths. For
smooth geomembrane—750 g/m2 geotextile higher shear strengths were obtained
using a well graded sand and gravel, since the fine particles have an influence
through the geotextile. The 1200 g/m2 geotextile interface resulted in higher shear
strengths when a subangular gravel cover soil was used.
f The interface shear strengths for textured geomembrane—geotextile systems depends
both on the cover soil used and the type of textured geomembrane. In particular, the
impingement textured geomembrane recorded significant apparent cohesion except
when a subangular medium gravel was used as cover soil.
f Some textured geomembrane—geotextile interface shear strength envelopes can be
described more accurately by non-linear failure envelopes. A second-order poly-
nomial expression has been demonstrated to be applicable.
f A nylon block can be used in the direct shear apparatus as a replacement for a cover
soil in order to estimate interface shear strengths, but caution is required. For
smooth geomembranes the nylon block results are higher than results using cover
soils above the geotextile. For textured geomembranes, it is considered that the use
of a nylon block cover is a conservative approach.
f Torsional ring shear tests gave lower residual interface shear strengths than the
direct shear tests. These results are considered to be lower bound values since
the effect of cover soils are not modelled. Further, once the ring shear has
rotated through 360° the geotextile fibres have undergone realignment along the
geomembrane asperities. The observed subsequent reduction in shear strength has
been attributed to polishing and fibre loss effects. This is an extreme case and
possibly does not model the field condition where the random nature of most
geomembrane texturing would result in continuing fibre tangle even after large
displacements.
70 D.R.V. Jones, N. Dixon/Geotextiles and Geomembranes 16 (1998) 45—71

f The strain softening behaviour of the geomembrane—geotextile interface can be


described by the D value. This ratio was found on average to be 0.7 and 0.4 for
smooth and textured geomembrane—geotextile interfaces respectively.
f Due to the influence of the cover soil on the geosynthetic interfaces, site-
specific direct shear tests should be carried out to ensure cost effective design
solutions.
f The full stress—displacement curves at the anticipated normal stresses, such as
those presented in this paper, are required to enable rigorous analysis of lining
systems.

Acknowledgements

The results presented in this paper were funded by Golder Associates (UK) Ltd.
and the EPSRC as part of the Teaching Company Scheme (Grant Ref. GR/H85731).
Mr Russell Jones was employed as the Teaching Company Associate on the above
programme, and carried out the testing with the help of Mr Robert Stickney whose
help is gratefully acknowledged. Materials used in the testing programme were
provided by Geofabrics and GSE Environmental, and thanks are due to these
suppliers.

References

British Standards Institution, 1991. BS 6906, Method of test for Geotextiles, Part 8 Determination of
sand-geotextile frictional behaviour by direct shear.
Bromhead, E. N., 1979. A simple ring shear apparatus. Ground Engineering 12 (5), 40—44.
Brune, M., Ramke, H.-G., Collins, H.-J. and Hanert, H. H., 1991. Incrustation processes in drainage
systems of sanitary landfills. Proc. Sardinia 91, 3rd Int. Landfill Symp. Cagliari, Sardinia, Italy,
pp. 999—1035.
Byrne, R. J., Kendall, J. and Brown, S., 1992. Cause and mechanism of failure, Kettleman Hills Landfill
B-19, Phase IA. Stability and performance of slopes and embankments-II. Geotechnical Special
Publication No. 31. in: R. B. Seed, R. W. Boulanger (Eds.), pp. 1188—1215.
Byrne, R. J., 1994. Design issues with strain-softening interfaces in landfill liners. Proc. Waste Tech ’94,
Charleston, South Carolina.
Department of the Environment, 1995. Waste Management Paper 26B, Landfill Design, Construction and
Operational Practice, HMSO, London.
Giroud, J.-P., Beech, J. F., 1989. Stability of soil layers on geosynthetic lining systems. Proc. Geosynthetics
’89, San Diego, USA, 35—46
Giroud, J.-P., Darrasse, J., Bachus, R. C., 1993. Hyperbolic expression for soil—geosynthetic or geosyn-
thetic—geosynthetic interface shear strength. Geotextiles and Geomembranes 12, 275—286.
Koerner, R. M., 1994. Designing with Geosynthetics, 3rd. Ed., Prentice-Hall, Englewood Cliffs, NJ.
Koerner, R. M., Hwu, B. L., 1991. Stability and tension considerations regarding cover soils on geomem-
brane lined slopes. Geotextiles and Geomembranes 10, 335—355.
Long, J. H., Gilbert, R. B., Daly, J. J., 1994. Geosynthetic loads in landfill slopes: Displacement compatibil-
ity. Journal of Geotechnical Engineering 120 (11, ASCE), 2009—2025.
Mitchell, J. K., Seed, R. B., Seed, H. B., 1990. Stability considerations in the design and construction of lined
waste repositories. Geotechnics of waste fills theory and practice ASTM STP 1070, Landva & Knowles,
Eds., ASTM, Philadelphia.
D.R.V. Jones, N. Dixon/Geotextiles and Geomembranes 16 (1998) 45—71 71

Seed, R. B., Mitchell, J. K., Seed, H. B., 1988. Slope stability failure investigation: landfill unit B-19, Phase
IA, Kettleman Hills, California. Report No. UCB/GT/88-01. Department of Civil Engineering, Univer-
sity of California, Berkeley.
Stark, T. D., Poeppel, A. R., 1994. Landfill liner interface shear strengths from torsional-ring-shear tests.
Journal of Geotechnical Engineering 120 (3, ASCE), 597—615.
Stark, T. D., Williamson, T. A., Eid, H. T., 1996. HDPE geomembrane/geotextile interface shear strength. J.
Geotech. Eng. 122 (3, ASCE), 197—203.

Вам также может понравиться