You are on page 1of 32

GEOMEMBRANE LINERS FOR RESOURCE

AND ENVIRONMENTAL PROTECTION: ENSURING LONG TERM PERFORMANCE

Ian D. Peggs
I-CORP INTERNATIONAL
Ocean Ridge, Florida, USA

INTRODUCTION

Geomembranes have rapidly become accepted as standard components of


geosynthetic lining systems in landfills, ponds, canals, dams, reservoirs, heap leach
pads, wastewater treatment plants, fish farms, and agricultural facilities. They isolate
leachates generated by waste (municipal, industrial, and hazardous) to prevent
groundwater contamination, they contain valuable product to prevent its loss into the
ground, they prevent rainwater from becoming leachate, they prevent evaporation
and contamination of drinking water, and they make dams waterproof. And they
make very cost-effective, efficient, space-saving barrier systems. But their success has
generated conventions and concepts that are contributing to an increasing number of
performance problems, or failures.

These failures are still a small percentage of the total industry but they receive much
attention that can make some engineers, regulators, and facility owners uncomfortable
in using geomembranes. However, in a new industry they are unavoidable, but
fortunately they provide a great learning opportunity. The real disaster is if we do not
learn from these failures and if we do not put what we learn into practice. The
objective of this paper is to identify the problems that are occurring so that durable,
properly performing geomembrane lining systems can be constructed that will benefit
us all.

THE PHILOSOPHICAL PROBLEMS

There are two fundamental perceptions that are the basis of liner failures, and they are:

1. High density polyethylene (HDPE) is the only acceptable geomembrane


material. This is wrong!
2. Geomembrane materials are commodity items and need only be purchased
on the basis of cost. This is even more wrong!

The acronym HDPE has become synonymous with geomembrane. Engineers


assume it will perform in all applications and regulators approve it easily because other
regulators approve it easily. In many cases regulations require its use, not by
specifying HDPE directly, but by requiring a geomembrane with physical properties that
only HDPE will meet. However, there are several conditions under which HDPE is the
wrong material to use. In fact, not only are there other geomembrane materials with
properties more suited to some applications, but there are many different HDPE
products. Most HDPE geomembrane manufacturers make product using three or four

\\Lara\SharedDocs\site\tech_docs\ensure_longterm_perf.doc
different resins based on availability and cost. And within the available resins there
can be a factor of 500 difference in their long-term mechanical durabilities. Therefore,
all HDPE geomembranes are not created equal. Engineers and regulators are not
typically aware of this.

Geomembranes are sold and bought as commodity items; consequently they are
treated as commodity items. After all they are only thin sheets of plastic that conform
to the profile of the subgrade and only need melting and pressing together to make a
joint!! But should the geomembrane that is used for a hazardous waste liquid lagoon
be made from the same resin, installed the same way, and have the same mechanical
durability as the liner for a golf course pond? I think we would all agree that the former
should be a far more durable material installed with higher quality. Thus, there are
commodity installations and there are custom or specialty applications. Present
geomembrane philosophy must be amended to recognize this, and to apply it in all
sectors materials, design, installation, testing, and construction quality assurance
(CQA).

BASIC DESIGN PHILOSOPHY

There are three fundamental liner designs single, double, and composite liners. Single
geomembrane liners are typically used in non-critical applications. American
philosophy is that a single liner cannot be installed without a few defects that leak.
Thus, for more critical applications, a double lining system is required. Electrical leak
location survey results on installed liners justify this philosophy.

Naturally, there are the same number of leaks in the secondary (lower) liner as there are
the primary (upper) liner, but if leakage through the primary liner is removed from the
space between the two liners there is no hydrostatic head on the secondary liner.
Therefore the lining system does not leak. Leakage will only occur if the leaking liquid
passes over a defect in the secondary liner this is most unlikely. The double lining
system is similar to double hulls on ships all ship hulls leak but the ship does not sink
provided the leaking water is pumped out. If the leaking water is not pumped out from
between the liners, the benefit and costs of a double lining system are totally lost.

For the most critical applications composite lining systems consisting of a


geomembrane and a clay or GCL liner are used. When there is a hole in the
geomembrane the leak only acts on a small area of clay (GCL), provided there is
intimate contact between geomembrane and clay. The permeability of a composite
liner is about three orders of magnitude lower than either a geomembrane or clay liner
alone. However, it is absolutely essential that the geomembrane and clay liner be in
100% contact. If there are wrinkles in the geomembrane or ruts in the clay the benefits
of a composite liner are lost, and the additional cost is wasted.

In the US, geomembranes are typically installed with some slackness (wrinkles) to allow
for contraction in cold weather. The complete geomembrane layer is installed before
the overlying layer is placed. This is acceptable if the liner is covered in the cold

\\Lara\SharedDocs\site\tech_docs\ensure_longterm_perf.doc
weather, when the liner is flat and without stress, but invariably wrinkles are built into the
geomembrane, losing the benefits of a composite system.

In Germany, small areas of the complete lining system are built sequentially. At the
end of the day the soil (drainage stone) that is to be spread on top of the
geomembrane is placed in a berm around the edge of the liner to hold it in place. At
night the geomembrane pulls tight to remove any wrinkles. Early the next morning,
when the geomembrane is still cool and flat, the soil is spread. This ensures complete
intimate contact, but with minimum stress in the geomembrane. The German
approach is technically the better of the two but installation is slower and more costly.

The concern about wrinkles is not only their impact on full contact between
geomembrane and clay, but also relates to the effect of residual stress at creases and
folds on the long-term performance of the geomembrane, particularly when the
geomembrane material is HDPE.

GEOMEMBRANE MATERIALS

HDPE is considered the geomembrane material of choice due to its resistance to many
chemicals and due to its high strength. However, in most cases, such as in municipal
solid waste landfills, leachates are quite weak. Certainly there are concentrations of
specific chemicals in the waste but by the time they have drained down to the
geomembrane and have been diluted by infiltrating rainwater, they are no longer
concentrated. In any case, many of these chemicals are detergents, chlorinated
solvents, and oxidizing acids, which are in fact quite damaging to HDPE.

It should also be noted that the strength of a geomembrane is really of little significance
in liner design. No matter what the strength of the material, no thin geomembrane will
support the weight of several meters of liquid or hundreds of meters of solid waste. The
important parameter is the ability of the geomembrane to deform to accommodate
any deflections or differential settlement in the subgrade and to deform such that the
long-term performance of the geomembrane is not compromised. The philosophy of
liner design should be that the geomembrane simply acts as a barrier and not as a
load-bearing member of the lining system. This is extremely important for HDPE
geomembranes due to HDPEs susceptibility to stress cracking a premature brittle
fracture resulting from a constant stress lower than the yield or break stress of the
geomembrane.

The stress cracking susceptibility of HDPE is its Achilles heel and is essentially the only
parameter that needs to be specified, since all other conventional mechanical and
deformation properties of HDPE geomembranes are independent of the resin used.
However the stress cracking resistance (SCR) of typical HDPE geomembranes can vary
by a factor of 500 or more. Therefore, a grinding gouge at a stressed overheated
seam that will cause failure in 1 year in a geomembrane made with an HDPE resin of
low SCR will not cause failure for 500 years if a high SCR resin is used. Therefore, the
low SCR material may be acceptable for a golf course pond liner but it should not be

\\Lara\SharedDocs\site\tech_docs\ensure_longterm_perf.doc
used to contain hazardous waste liquids. This is an example of the differences
between commodity and specialty applications.

All HDPEs are susceptible to stress cracking (SC). It is a function of the crystallinity of
the material, its molecular weight, its molecular structure, and the type of comonomer
used. Figure 1 shows the applied stress/break time curve for five different HDPE
geomembranes. At the higher stresses of the shallow slope the breaks are ductile, but
as the stress decreases the slope increases and the breaks are all brittle. Therefore
failure occurs much sooner than expected by extrapolation of the shallow part of the
curve. Hence the requirement to install HDPE without stress in lining systems, particularly
critical lining systems.

Figure 1. Stress rupture curves for five HDPE geomembranes (Hsuan et al, 1992)

Figure 2 shows that stress cracking is accelerated as temperature increases.


Fortunately, in most lining systems, stresses decrease as liner temperatures increase.

\\Lara\SharedDocs\site\tech_docs\ensure_longterm_perf.doc
Figure 2. Effect of temperature on stress rupture curve.

Environmental Stress Cracking (ESC) occurs when SC is accelerated in certain chemical


environments, such as detergents, chlorinated solvents, and oxidizing acids. For
example black liquors at pulp mills and acids in mine solvent extraction facilities can be
a problem. Thus it is very important when performing chemical resistance tests to do a
test under stress to assess the potential for ESC. Most chemical resistance charts do
not present the influence of stress.

SC also becomes a problem when all the antioxidant (AO) additives in an HDPE
geomembrane are consumed. Therefore, when sales people talk about HDPE having
no additives similar to plasticizers in PVC they are correct on one hand - plasticizers are
not needed to make HDPE flexible - but they are incorrect to imply that HDPE contains
no additives that are consumed during service. When all the plasticizer is removed
from PVC the material becomes brittle and cracks, but when all the antioxidant is
removed from HDPE it also becomes brittle and stress cracks. The same thing will also
happen with LLDPE and PP geomembranes when antioxidants and UV stabilizers are
consumed. However, in most cases, sufficient additives are present to more than
provide the expected service life.

Thus, the oxidative induction time (OIT), a measure of the oxidation resistance of HDPE
becomes important, especially where the geomembrane is not covered in service. In
the OIT test a sample from the full thickness of the geomembrane is tested, so the test

\\Lara\SharedDocs\site\tech_docs\ensure_longterm_perf.doc
results provide a performance parameter averaged over the full thickness of the
geomembrane. Clearly, the exposed surface will be oxidized first and may lose its
complete AO additive before the AO at some depth is affected. Therefore, the test
data might show significant AO remaining in the test specimen even though the
surface is completely depleted and susceptible to SC initiation. Therefore, the
durability of an exposed HDPE geomembrane is a function of its OIT then its SCR. Thus,
a material with a low OIT and high SCR may have the same long-term durability as one
with a high OIT and low SCR. Clearly the high OIT and high SCR geomembrane is the
one that should be used for the exposed primary liner of a double geomembrane liner
for a hazardous waste liquid impoundment. In covered applications a high SCR is the
most important parameter.

However, there is one practical concern with OIT in the ASTM D3895 standard OIT test
the specimen is heated at 200oC to monitor the time to degradation. Since the AO
package is made up of several different components providing protection over
different temperature ranges the high temperature performance is not necessarily the
same as would occur at lower operating temperatures in the range of 0 to 80oC
(geomembrane temperatures, not air temperatures). It would, therefore, be possible
to have ample oxidation protection at operating temperatures but to have a low value
of OIT. Of course, the opposite situation is also possible.

Tests are presently being developed to take a thin layer (20 m) of HDPE, representative
of the thin surface layer of the geomembrane, to expose it to a temperature of about
85oC (the maximum experienced in exposed service) for a few hours, and to measure
the change in carbonyl group content a measure of the amount of oxidation. This
more realistic oxidation parameter will then be combined with the SCR to provide an
overall material durability factor (MDF) for exposed HDE geomembrane. The MDF can
then be used to determine whether a specific HDPE geomembrane is more suited for
commodity or specialty applications.

GEOMEMBRANE MANUFACTURING

There are essentially three methods of manufacturing geomembranes, two of which


are used for HDPE. In both these methods all the components are mixed and driven
through a screw extruder to feed the die. In one case the die is a wide flat slit, the
widest of which is about 10 m. In fact, in this case there are two extruders feeding two
conventional flat dies that are linked side by side. Flat extruders can control thickness
to within about 3%.

In the second case the screw feeds a circular die and a tube is extruded vertically. The
tube is drawn (pulled) up about 25 m and is maintained stable by blowing cooling
pressurized air into the tube. Thus, when the tube exits the die it is also increased in
diameter somewhat. The thickness of the tube is controlled, not by changing the die
gap, but by varying the tension and speed at which the tube is pulled upwards.

The tube is flattened and taken over a roll at its maximum height, and is then brought
downwards during which it is slit along one side, opened up, and then rolled up as a

\\Lara\SharedDocs\site\tech_docs\ensure_longterm_perf.doc
single wide sheet. Consequently, these geomembranes have folds running along their
length at the quarter and three quarter widths. Improper opening of the sheet after
cutting can often lead to angled creases centered on the folds. These rolls are
typically 7 m wide with thickness variations between 5 and 10% of nominal.

Polypropylene geomembranes are made the same way.

The third manufacturing method, calendering, is used for PVC and for some reinforced
materials. The plastic mixture of all components is laid as a wide multi-layered ribbon
between two side-by-side rollers through which it is rolled into a sheet, as shown in
Figure 3. The sheet is guided around an inverted L-shaped bank of rolls including a final
pair of rolls which generate the required thickness. The widest calendering rolls are
about 3 m wide. The second pair of rolls may have a machined surface profile that is
imprinted into the surface of the sheet a very fine mesh profile is known as a faille
finish.

Figure 3. Roll arrangements for calendering reinforced geomembrane

The round die method of manufacturing sheet has the advantage that three extruder
dies can be arranged concentrically so that three separate layers can be bonded

\\Lara\SharedDocs\site\tech_docs\ensure_longterm_perf.doc
(laminated) together immediately after extrusion. This allows for several different
modifications to the basic homogeneous geomembrane structure:

1. Chemically resistant HDPE on the outside and more flexible LLDPE on the inside
2. High SCR LLDPE on the outside and HDPE on the inside
3. A white surface layer to keep sheet temperatures low, which will minimize
expansion wrinkling
4. Colored surfaces to suit the environment
5. Conductive lower surface layer to facilitate electrical leak detection as the final
stage of construction quality assurance (CQA)
6. Textures on one or both surfaces

White and colored surfaces preclude the use of carbon black us the UV stabilizer.
Since carbon black is well established as the best UV stabilizer, specifying colored sheet
automatically implies lesser UV resistance. Consequently only black sheet should be
specified for projects requiring the longest exposed service.

Textured geomembranes are made by blowing nitrogen into the mixture(s) in one or
both outer screw extruders. As the material exits the die(s), it expands and explodes to
the outer surface(s) of the sheet roughening the surface like the ocean. It is not easy
to control the minimum thickness of geomembrane and the distribution and uniformity
of the surface texture. The surface layer resin may be somewhat different (higher melt
index) to the core resin in order to make it easier to texture. However care must be
taken that this material does not compromise the performance of the complete
geomembrane. Cadwallader (2001) has shown the use, in the surface layers, of what
appears to be a recycled PE resin that is highly susceptible to stress cracking. Once
the cracks have initiated in the surface layer they appear to propagate quite easily into
the core geomembrane which, under normal circumstances, has a high resistance to
stress cracking. Such products may not have adequate mechanical durability when
used on slopes.

Other methods of texturing one or both surfaces of an HDPE geomembrane include


dropping an LLDPE powder onto the hot surface to which it melts and bonds, or by
spraying the surface with hot molten fibrils of PE. However, in the past, small stress
cracks have been noted to form around the edge of the particle where it is welded to
the sheet, which will reduce the mechanical durability of the sheet, particularly if the
texture causes a transfer of stress to the surface of the sheet. Thus, these textures
require a compromise between minimizing the heating and thermal gradients to avoid
stress cracking, and providing sufficient bonding so that the texture cannot be abraded
off the surface. Once again, it is difficult to control the quality and distribution of the
surface texture generated in these two methods.

It appears to me that provided a resin with good SCR is used the nitrogen method
probably produces the most SC resistant textured HDPE geomembrane.

To provide a more uniform and consistent friction-enhancing surface, profiles can be


generated in HDPE geomembranes by a calendering process immediately after
extrusion. Such patterns include conical spikes, stubs, ridges, ridged cells, and swirly

\\Lara\SharedDocs\site\tech_docs\ensure_longterm_perf.doc
profiles. Clearly a profile can be designed for friction increases with different
interfaces. For instance, ridges may be appropriate for sand and clay surfaces, but
they would not be very effective against woven geotextiles. A fine fibrous structure
might work well against a nonwoven geotextile, but would not be very effective
against sand. Therefore, all textured surfaces behave quite differently and it is
absolutely essential to perform direct shear tests with the geosynthetic materials
proposed for use and the actual site soils. Then, if randomly textured material is used,
periodic conformance direct shear tests should be performed to ensure an adequate
texture is being maintained on material delivered to the site.

The use of textured surfaces on both sides of the geomembrane, particularly HDPE
should be carefully considered. Since a design objective is to keep stress out of the
geomembrane (particularly HDPE) it is important that the shear stress on the upper
surface be lower than that on the lower surface. There are engineers who will only use
a texture on the bottom surface and insist on the upper surface being smooth. In this
way, if the layer on top of the geomembrane does move it will slide on the
geomembrane and not tear it. The soil layers on top of the geomembrane can be
reinforced with a geogrid or a high strength geotextile.

GEOMEMBRANE SEAMS

All geomembrane seams should be thermally bonded so they will not peel apart. It
does not make sense to require that thermally bonded HDPE seams do not separate in
a peel test, while allowing chemical seams in materials such as PVC to peel apart
provided the peel force exceeds a minimum value.

Seam stress cracking

There is no question that a double track thermal fusion seam (set up correctly) is
superior to a fillet extrusion seam. However, for maximum durability, the width and
shape of the nip rolls in comparison to the width of the heating wedge should be such
as to eliminate residual stresses at the root of the squeeze-out bead. Residual stresses
can be identified by examining a thin-slice (10-15 m) microsection under a transmitted
light microscope with crossed polarizing filters; in HDPE residual stresses appear brightly
colored.

In susceptible HDPE geomembranes SC typically occurs in the lower sheet just under the
edge of the weld bead of extrusion seams. It occurs at a shallow angle to
perpendicular to the plane of the sheet, probably along the boundary between the
isotropic melted and solidified weld zone and the oriented unmelted sheet material. If
this is a sharp interface with rapid transition in microstructure it will act as a structural
notch defect. However, if the transition is more gradual, such as might be
developed by annealing, the SC susceptibility might be decreased. For this reason, it is
believed that a combined hot air/hot wedge welder will provide the most durable
seams the wedge does the melting and the hot air facilitates a more gradual
temperature gradient at the edge of the wedge.

\\Lara\SharedDocs\site\tech_docs\ensure_longterm_perf.doc
The most common location for stress cracks to occur is at and near the ends of
extrusion beads, such as occur at a short bead repair or at the start/run-out of beads
along seams at patches (Figure 4). It has been proposed that patches be placed on
the underside of the liner so no run-out beads are necessary, or that the start and run-
out beads be done on top of the patch itself. Similarly, there are requirements that no
more than two beads be placed in the same location to avoid overheating the
geomembrane and increasing its susceptibility to SC. Clearly, the most trouble-free
way to avoid SC is to use a geomembrane made with a high SCR HDPE resin so that it
will accommodate the unavoidable abuse that occurs during installation.

Figure 4. Typical stress crack at edge of extrusion bead

\\Lara\SharedDocs\site\tech_docs\ensure_longterm_perf.doc
Seam testing

When testing seams it has been shown that measuring shear and peel strengths
provides no practically useful information on seam bond strength (Peggs 1996) [EG1].
Due to the relatively large seam shear area compared to the small cross sectional area
of the sheet tab through which the seam is stressed, the sheet always breaks before the
weld separates, provided the bond efficiency is more than about 8%. This figure varies
between 8 and 20% depending on geomembrane thickness. This is quite evident
when it is recognized that conventional allowable seam shear and peel strengths
increase with geomembrane thickness despite the fact that what we are attempting to
measure is totally independent of thickness!

Therefore, if seam bond efficiency is only 25%, surely inadequate, we are not able to
challenge the seam enough to be able to determine this. Therefore, there is little
need to measure strengths. The important features of welding are to retain the
ductility of the geomembrane adjacent to the seam and to ensure that the seam will
not peel apart this is the best information we can get from the conventional tests.
Ductility in the shear test is important to demonstrate that the geomembrane has not
been mechanically damaged during preparation for welding and that it has not been
overheated during welding. Separation in the peel test is important to demonstrate
that a minimal degree of bonding has occurred. In HDPE, if peel separation can occur
it is possible that crazing (Figure 5), the forerunner of stress cracking, is initiated in the
separated surfaces. This can lead to significant reductions in the SCR of the remaining
sheet. If this can occur in the laboratory it can occur in the field at wrinkles, at folds,
where ice forms under flaps at the edges of seams, where subgrade slumping occurs,
etc. Yet again, the best way to avoid these problems is to use a material with high SCR
and that means higher than the conventional 200 hr in the ASTM D5397 notched
constant tensile load test. This author has seen SC failures in material with a D5397 SCR
of 240 hr. A minimum figure of 300 hr has recently been incorporated in the GRI-GMB
specification. Clearly, the SCR should reflect the criticality of the installation.

\\Lara\SharedDocs\site\tech_docs\ensure_longterm_perf.doc
Figure 5. Crazing induced in HDPE seam by peel separation

Stress cracking tests can be performed on seams and textured material to confirm that
the respective processes do not adversely affect the SCR of the basic geomembrane
material. Clearly, a notch, as required in ASTM D5397 cannot be used, since this will
result in testing the parent geomembrane at the root of the notch, bypassing the effect
of the surface texture and of welding. Rather, a simple unnotched strip is used, with

\\Lara\SharedDocs\site\tech_docs\ensure_longterm_perf.doc
carefully prepared undamaged edges, tested at 4MPa at 80oC in a surface active
agent (Thomas). Break time must exceed 700 hr.

Occasionally, when peel testing HDPE, the welded interface does not separate but the
geomembrane itself separates, as if delaminating (Peggs, 1985). This phenomenon,
known as separation-in-plane (SIP), is seen more often in PP geomembrane seams
(Figure 6). The SIP can travel part way across a weld track, across the air channel, and
even beyond the second weld track. The cause of this phenomenon is presently being
researched. It has been proposed due to incompatible carrier resins for the carbon
black masterbatch and crystallinity gradients through the thickness of the sheet caused
by differential cooling rates after extrusion. In extreme cases, it has been possible to
delaminate both HDPE and PP geomembranes themselves, as shown in Figure 7.
Therefore, it appears to be a basic geomembrane feature that becomes apparent
when peel testing after welding. It is not a welding phenomenon.

Figure 6. SIP in PP seam

\\Lara\SharedDocs\site\tech_docs\ensure_longterm_perf.doc
Figure 7. Delamination in HDPE (Top) and PP (Bottom) geomembrane

At present the Geosynthetic Research Institute is debating with its members whether SIP
is an acceptable mode of peel break in any material, but at a minimum peel force.
However, until it is clearly understood why SIP occurs, I feel that this is an unnecessary
and retrograde step, since it is clearly possible to make both HDPE and PP seams that
do not break in a SIP mode.

Welding temperatures

Project specifications often limit welding to certain ambient temperature ranges, or to


temperatures measured a certain distance above the geomembrane. Neither of
these temperatures has any significance. The only temperature that is important is the
actual temperature of the geomembrane itself. This will be the ambient temperature
when it is cloudy, and a much higher temperature when it is sunny; a black
geomembrane can reach over 80oC in summer sun.

Typically PP geomembranes have a much wider welding window than HDPE and there
are narrower gray areas at each side of this window PP is either welded or it is not.
The weldability of HDPE is a function of the resin and its individual Melt Index. Hence
the reason to specify that the welding rod for extrusion welds be made from the same
resin as the geomembrane. However, this is simply a way of most easily assuring
maximum weld quality, since an experienced installer can successfully join materials of

\\Lara\SharedDocs\site\tech_docs\ensure_longterm_perf.doc
quite different Melt Indices. After all, LLDPE can be successfully joined to HDPE in
coextruded geomembranes, and high molecular weight pipe can be butt-fused to
lower molecular weight pipe.

HDPE can be welded at temperatures below freezing and at temperatures above


75oC, but this should be done with more frequent trial welds and seam tests. And trial
welds should always be done in the environment and geometry in which production
welding is to be done. For instance, trial welds should not be done in the back of a
truck or on a smooth concrete surface when production welds are to be made on
damp soils on a slope, or an a vertical wall.

Textured sheet welding

Care must be taken when welding textured sheets to ensure that they are clean.
Damp soil or bentonite powder trapped in the texture will prevent effective welding.
Consequently, many textured sheets are made with smooth edges to facilitate welding.
However, when round die material is made with smooth edges, the smooth strip along
the extruded tube is not as thick as the textured area so deforms more with the result
that the edges of the sheet are longer than the center of the roll. Thus, the edges may
be rippled (Figure 8), giving the impression that the quality of welding may be
compromised. However, examination of such seams, when made by a competent
welder, shows no problems with weld zone quality, weld symmetry and uniformity, and
residual stresses.

\\Lara\SharedDocs\site\tech_docs\ensure_longterm_perf.doc
Figure 8. Rippled edge of textured geomembrane

There are a few continuing developments in joining HDPE and PP geomembranes


separately and to each other. Electrofusion was initiated in 1987 (Butts) but is only now
being developed by GSE for field use. The welding rod for extrusion welding is
essentially extruded around resistance heating wires. A long length of wire is placed
between the surfaces to be joined, they are ballasted, and power is applied to the
heating wires, the rod melts, and welding occurs. In fact the method is more
appropriate for patches where the rod can be prewelded to the patch, the patch
placed over the defect, a weight placed on it, and current passed through the wires
for a few seconds. Thus patching can be done by local labor. It is also claimed that
such patches can be applied underwater, but this has not yet been adequately
demonstrated.

PE and PP geomembranes can now be given a surface fluorination treatment that


enables them to be adhered to concrete, to themselves, and to each other with an
epoxy adhesive. This again enables joining to be done by local labor rather than by
costly professional welding crews. It also enables geomembranes to be fully bonded
to vertical concrete walls rather than hanging them freely from batten strips at the tops
of the walls. This improves liner durability and leakage performance. The treatment

\\Lara\SharedDocs\site\tech_docs\ensure_longterm_perf.doc
also provides better barrier properties to hydrocarbons the gas tanks in BMW
automobiles are made with a fluorinated HDPE.

LINER PERFORMANCE EXPERIENCES

We learn most about the performance of geomembranes and lining systems when they
fail. We must make the most of these otherwise unfortunate happenings to extract
most benefit from them. They should be investigated independently and lessons
learned communicated to others so similar problems do not happen again. If there is
a crack alongside a seam, an extrusion bead should not be applied to it; the crack
should be cut out for possible future examination and the resulting hole patched.
Examination of a fracture face can show where the crack started, the feature that
caused it, and how it propagated. When a slope failure occurs and the liner is torn,
the tear should be examined to determine if the tear was a result of the slide, or if the
slide was initiated by the tear. This can be very important.

The number of liner failures is increasing. My investigations find that most (~60%) are
due to inadequate design, 20% are due to inadequate installation by installer or
earthwork contractor, 10% are due to material deficiencies, and 10% are due to
miscellaneous features. In most cases, the installer is held responsible. Often this is not
justified.

Conformance testing

HDPE geomembrane was made in Europe and delivered to a site in South Africa.
Conformance testing showed the material to meet all project specifications except
SCR; break time in the ASTM D5397 test was 60 hr instead of the specified minimum of
400 hr. The manufacturer insisted the material was acceptable, but he had not tested
it. The resin manufacturer insisted the SCR had been measured to be about 360 hr.
The material was rejected and new material had to be made in the USA. The
specification was easily met.

It should be noted that the ASTM D1928 standard procedure for casting plaques on
which to do resin tests allows three cooling rates. Each of these rates will generate
different amounts of crystallinity in HDPE. If the cooling rate was too fast the crystallinity
would be lower than in a geomembrane and the SCR would be unrealistically high.

In the owners next project HDPE was being made in Korea, but the manufacturer
could not make all the material in time. Some material was ordered from Europe. This
time conformance testing was done before material was shipped to South Africa. The
Korean material met specifications, the European material did not meet the OIT
specification of 100 minutes at 200oC. Test results showed a value of 101 minutes but
the test was performed at 180oC. At 200oC the result was found to be about 75
minutes. According to ASTM D3895 the test can be performed at 180oC, if specified,
but if no temperature is specified it must be done at 200oC. The engineer was
prepared to reject this material, but the CQA consultant thought this was unnecessary
since the material was not be used exposed. Subsequently the material was accepted

\\Lara\SharedDocs\site\tech_docs\ensure_longterm_perf.doc
for use. In addition, the engineer was not willing to support the generation of data to
support this decision, but required that the installer/manufacturer support the CQA
consultant in generating the information. This is not standard CQA procedure as
usually contracted by the owner to the engineer the engineers responsibility is to
ensure that the owner gets a properly performing installation at the best cost, and that
includes properly dealing with nonconforming materials.

Exposed liner cracking

An HDPE geomembrane was installed in a landfill cell in Florida two steep slopes were
left exposed. After 8 years, cracking occurred in seams, at the side of seams, and
along apex-down folds just below the tops of the slopes. Apex-up folds did not crack.
There was also cracking at a protrusion underneath a patch (Figure 9). Despite the
fact that the SCR of the material was about 240 hr and the OIT just over 100 min, both
meeting project specifications, the surface of the material had oxidized and cracks
had initiated on the surface. The stresses required to initiate the cracks were
contraction stresses (even though temperature changes were relatively small) and wind
induced stresses. Only apex-down folds had cracked because the upper oxidized
surface was in tension when the sheet contracted. In the apex-up folds the oxidized
surface was in compression. Yet again, if this material had had a higher SCR this
cracking probably would not have occurred.

\\Lara\SharedDocs\site\tech_docs\ensure_longterm_perf.doc
Figure 9. Cracking (Top) at protrusion under HDPE geomembrane patch (Bottom)

Cast-in liner chemical resistance

An HDPE cast-in liner, about 5 mm thick, was specified for use on the walls in several
concrete basins of a heap leach mine solvent extraction system. Loose
geomembrane was used to line the floors. There were no specifications for SCR or OIT.
The specifications required an HDPE that was compatible with a list of liquids at
specified temperatures. The settler unit contained mixtures of kerosene, naphtha,
oxidizing organics, and sulphuric acid. The engineer was urged to do chemical
resistance tests but declined to do so, arguing that such a system had worked
successfully before so there should be no problems. This was despite the fact that
some chemical resistance tables show incompatibility between PE and naphtha and
kerosene at temperatures close to the maximum specified.

Large gaps between cast-in sheets on the walls that could not be extrusion welded
were capped with loose sheet and welded. Spark tests were performed on all welds
and repairs were made. The basin was filled with water. Leaks were found. The
basin was emptied. Spark testing was performed and additional leaks were located.
The engineer thought the installer had done an inadequate leak survey the first time.
However, leaks were now full of water and were more conductive so it was not
surprising that more were found. The facility was filled with organic solution once
again more leaks were found. Again the engineer was dissatisfied with the installer,
but the organics were less viscous than water so would penetrate even finer passages.

\\Lara\SharedDocs\site\tech_docs\ensure_longterm_perf.doc
Finally the acid (including copper) was added and the temperature raised to the
service temperature - then about 20oC higher. Leaks started to occur as welds along
the edges of the cap strips over the joints started to separate. Welds at corners started
to separate, and at any location where an unsupported sheet was welded to a
supported sheet. The general contractor and liner installer were soundly faulted for a
bad installation and poor welding.

The unsupported sheet was absorbing organics and swelling. The cap strips were
bowing and inducing a peel stress on every millimetre of weld. When seams are
destructively tested it is not unusual to allow one peel test of the five peel and five shear
tests to fail while still accepting the complete weld sample. This might be acceptable
when, as should be designed, the liner is not subject to stress. However, a one in five
failure rate is a 20% failure rate! Therefore, it might not be surprising that when all of
the seam is subjected to a peeling stress in service the most poorly bonded segment
within the unacceptable 20% will fail. Then, when it is repaired, the next weakest
segment will fail. And so on!

But repairs also become more difficult when the surface of the sheet is saturated with
organics and the owner wants to keep down time to a minimum. Sufficient time must
be allowed for the organics to volatilize or be removed, if effective repairs are to be
achieved.

The next problem is that the swelling and rippling stresses build up adjacent to seams,
where the liner has been heated and more of the antioxidants have been consumed.
The sulphuric acid and the oxidizing organics, particularly at operating temperatures
20oC higher than the already high operating temperature, cause additional oxidation
at the surface in those regions and stress cracks are initiated. Once initiated, stress
cracks apparently propagate quite easily. And it has already been described how
stress cracking is accelerated at higher temperatures. Therefore, cracking of the
unsupported liner occurs in and adjacent to the welds. Clearly, the appearance was
of a major welding problem. However, the primary cause of the problem was not the
HDPE or the welding, but the insistence of the engineers in using HDPE in an application
that was totally inappropriate for HDPE.

The engineer was obviously not aware of the performance characteristics of HDPE and
was not prepared to take the advice of those who did know. An assumption was
made that temperatures could be raised higher than those in previous installations
without any changes to the design, and then temperatures were increased even
further. If the engineers were not aware of the performance characteristics of the
liner, the regulating agency engineers would know even less. Clearly the regulators
would rely on the integrity of the engineers presentations a big mistake.

These swelling problems occurred at at least two mine sites, and cost many hundreds of
thousands of dollars in down time, repair time, lost production, testing, consultants, and
resulting legal fees. All essentially because a $10,000 chemical resistance test was not
performed. One of the more significant results of this exercise is that geomembranes
unjustifiably lost many supporters the HDPE liners were replaced by fiberglass.

\\Lara\SharedDocs\site\tech_docs\ensure_longterm_perf.doc
Leaking cast-in liner

A similar thing happened when a double lining system was placed inside a large
concrete basin of a major multinational corporation. When the basin was filled with
water the primary liner leakage collection system was dripping not a constant stream,
but just dripping. The owner requested the installer to make repairs since the liner was
not leak-free. Repairs were made but then the leakage collection system was
constantly running. After a second set of repairs the leakage rate was even higher.
The owner did not understand the philosophy of the double lining system and that all
single liners should be assumed to leak. How difficult would it have been to collect the
few initial drips and recirculate them into the basin!? Again, due to unrealistic
expectations, and totally unnecessarily, geomembranes fell into disfavor the lined
concrete basin was replaced with a stainless steel tank, and at much greater cost.

Geosynthetic clay liners

Here are two examples of geosynthetic clay liners (GCL). In a wastewater treatment
plant lagoon a GCL was specified for use by the engineer. His initial calculations using
the manufacturers specified hydraulic conductivity (HC) showed that the seepage
rate would not meet the States maximum allowable value. He asked the
manufacturer if the calculations were correct the manufacturer said they were. The
engineer then found an older specification with an HC value 20% of the present value.
He asked the manufacturer if the lower value could be guaranteed the manufacturer
said No. The desperate engineer asked what would be the lowest guaranteed value
the manufacturer provided a number and still this was not good enough. The
manufacturer suggested using a GCL incorporating a PE film. The engineer declined
The engineer then remembered that at the wet time of the year the groundwater level
would be higher than the floor of the pond thereby reducing the head on the liner.
This would have been true in the case of a geomembrane but not in the case of the
GCL. With the lower head and the minimum guaranteed HC the seepage rate would
just meet the States maximum allowable value. The pond was built, with very careful
records maintained by the general contractor. On completion, and with a 450 mm
ballast layer on the GCL, the pond was to be filled to a depth of about 3.5 m for a full-
scale hydrostatic test. At just over 2 m the pond was leaking faster than it could be
filled!

Not only was the GCL basically ineffective, the stones in the soil layers on each side of
the GCL were as large as 300 mm, compared to the maximum recommended by the
manufacturer of 25 mm, and the fine fraction in the soil was only 15% the volume
recommended by the manufacturer for uniform pressure confinement of the GCL.
When surveyed and uncovered there were 55 groups of holes in the 1.25 ha GCL.
And the general contractor and manufacturer were blamed for a low quality GCL
badly installed. Ultimately the engineer was found responsible, and the GCL was
replaced with a geomembrane as should have been used in the first case.

In the second case, an unballasted GCL was used to line a decorative pond. A sharp
rock aggregate with about 10% of fines (rather than the recommended minimum of
80%) was used as the subgrade the sharp edges had torn the lower geotextile of the

\\Lara\SharedDocs\site\tech_docs\ensure_longterm_perf.doc
GCL. Batten strips were used to fasten the GCL to peripheral walls, but in many
locations sections of batten strip were missing. Sharp limestone rocks were placed on
the GCL around the periphery of the pond they had punctured the GCL and were
providing calcium ions for cation exchange with the sodium ions in the GCL, thereby
reducing its HC. Transition of the peripheral GCL from batten strips on the walls to
natural materials confinement behind the rocks and soil were not sealed, and were
loose enough to push ones arm between the GCL and the end of the wall. This 750 m2
pond was leaking 136 m3 of water each day. This was a poor design that was badly
installed. The owner paid for the investigation, re-engineering, and installation of a
new lining system.

Clearly the concept of commodity and specialty applications for geomembranes also
applies to every sector of the industry there are engineers, installers, contractors,
laboratories, and CQA firms who are only capable of providing commodity services
and there are fewer who are capable of providing the knowledgeable services
required for more critical and custom applications of geomembranes, GCLs, and lining
systems.

QUALITY PROGRAMS

The first stage of a quality program is to ensure that those involved with the project are
experienced and knowledgeable in working with geosynthetics. Clearly this is not as
important for a golf course pond as it is for a hazardous liquid lagoon or a liner on a
steep quarry wall, although it is important to the golf course owner! For those working
with HDPE it is important that they understand the stress cracking phenomenon. For
exposed HDPE the OIT parameter must be understood. For PP, contractors must be
aware of the potential for SIP. For PVC, contractors must be aware of the capabilities if
chemical and thermal seaming and of the significance of plasticizer content. For
GCLs, an understanding of the reason for a uniform and adequate confining pressure is
essential. The initial difficult part of this, of course, is ensuring that the engineer, who
knows more than the owner and regulator does, in fact, understand what is being said.
The only way to confirm this is to have proposals, drawings, and documents submitted
for peer review by an established expert.

Then a quality assurance program becomes quite routine. However, it is most


important to realize the differences between quality control (QC) and quality
assurance (QA). QC is a quality program established by a contractor to ensure the
quality of its own work. Therefore, all parties to a project will have their individual QC
programs. QA is a quality program established typically by an owner to ensure that
the work done by others is done according to the project specifications and related
documents. A QA program, therefore, does not ensure that a perfect structure is built
if a poor design is adopted, the QA program will ensure that that poor design is built.
The most effective QA (only available from experts) will identify problem areas before
construction commences and attempt to have them resolved. It will also ensure that,
within the constraints of the budget and schedule, that work that is done will be to the
highest standard. Clearly, if the installer is not an experienced installer, that standard
will not be as high as it would be for a highly experienced installer. Such standards are

\\Lara\SharedDocs\site\tech_docs\ensure_longterm_perf.doc
established by the budget for the project, and the amount the owner is willing to pay.
If the owners simply want the lowest cost, they will get the lowest quality.

The objective of liner construction QA (CQA) is not only to ensure that what was
designed is built, but if there is a problem with the liner in service that construction
records are adequately comprehensive to enable a full analysis of the construction
process to quickly determine the full extent of the problem. For instance, if a weld
separated, how many seams did the welding machine and/or the operator make
between passing trial welds? If it is a material problem, where was the rest of that roll
used? Therefore, the best CQA personnel are those that have previously been welders
they know what can be done and what cant be done. They also know when quality
can easily be improved. Typically, an engineer does not make a good CQA monitor
the job is not sufficiently creative. And an inexperienced CQA monitor that simply
annoys a good installation crew will be an absolute waste of money and likely result in
a lower quality installation than if the installer had been left alone. To this end a
monitoring program based on the following should be implemented:

1. Review selection of geomembrane material for the project


2. Review specification for material
3. Review project specifications and drawings
4. Review CQA plan and ensure no gaps or conflicting overlaps with other project
documents, including contractors QC programs.
5. Ensure CQA plan describes actions in the event of non-conformance.
6. Review resin manufacturers, geomembrane manufacturers, and installers QC
certificates and programs to ensure there is full traceability of materials, people,
and equipment used, in the event of construction and in-service problems
7. Ensure manufacturer and installer understand and can comply with all project
documents
8. Confirm that all materials arriving on site meet project specifications. When
materials are manufactured overseas it is important to do this before material
leaves the manufacturing plant. Unfortunately, it is not sufficient to rely on QC
certificates.
9. Review proposed panel layout diagrams and liner penetration details
10. Ensure subgrade is suitable for placement of geosynthetics
11. Monitor and record details of all on-site construction activities
12. Identify and record all roll, panel, and seam numbers
13. Ensure trial welding is performed under the conditions of production welding
14. Ensure all failing trial seam results are recorded
15. Monitor all welding
16. Monitor all seam nondestructive and destructive testing
17. Select samples for destructive testing at the independent laboratory and review
test results.
18. Monitor numbering and testing of all repairs
19. Prepare panel layout drawing
20. Monitor placement of first soil layer on top of geosynthetics this is when most
damage to the geomembrane occurs..
21. Perform a geoelectric integrity survey over the complete liner surface.
22. Record all meetings in which problems are identified and resolutions proposed

\\Lara\SharedDocs\site\tech_docs\ensure_longterm_perf.doc
23. Prepare comprehensive final report including copies of all QC certificates,
deployment logs, welding records, test equipment and results, design
modifications, etc.

If CQA is professionally performed, responsible general contractors and installers will not
view CQA as an imposition or a nuisance, but as an unbiased means of helping them
achieve the highest quality and most durable lining system. Under no circumstances
can effective CQA (on behalf of the owner) be performed by the general contractor or
the installer.

LINER TESTING

Conventional testing procedures during liner installation include seam destructive (peel
and shear) testing, and seam air pressure, vacuum box, and spark testing. Obviously it
is most undesirable to cut sample of good double wedge welds out of the seams for
destructive testing then to make repairs with three times the length of an inferior
extrusion weld. The majority of the liner is not tested other than by cursory visual
inspection. This can be singularly ineffective, as outlined below.

The newly installed liner of a concrete basin in a wastewater treatment plant was
leaking. The owner spent nine months raising and lowering water levels, using dyes,
and making visual inspections but could not locate the leak, nor even the level at
which the leak occurred. Within four hours a geoelectric survey found a pinhole leak
in a weld at a floor/wall corner and two parallel holes approximately 20 mm long and 2
mm wide about 15 mm apart on the floor caused by a dropped tool. These holes were
very clear visually after they had been located, but they had not been found during
many previous visual inspections.

Geoelectric leak and integrity surveys

Geoelectric surveys have been commercially routinely performed since the mid-1980s.
The general principle of the technique is based on the geomembrane being an electric
insulator. A potential is applied between an electrode in the water or soil/waste above
the geomembrane and another in the soil or leakage collection system below the
geomembrane. In the ideal situation current will only flow through the leaks we are
trying to find. This assumes that pipe penetrations through the geomembrane, soil
around the edges, or rainwater wetting the edges of the liner do not also provide
pathways for current flow. A survey probe is then use to measure the potential
gradients between the electrodes of the search probe in the water or soil above the
geomembrane. Where there is a leak and a high current density there are also higher
than background potential gradients. Thus the leaks can be pinpointed. Variations of
the technique allow liquid-covered liners (any depth), uncovered liners, and soil-
covered liners to be surveyed. I have outlined (Peggs, 1999) design features that
should be considered if effective geoelectric leak/integrity surveys are likely to be
required.

By wading in shallow water holes of 1 mm or less in diameter can be exactly located.


In deep water, which requires dragging a probe from one side of the pond to the other,

\\Lara\SharedDocs\site\tech_docs\ensure_longterm_perf.doc
the same size of hole can be identified but location is typically to within 500 mm. A
liner covered by 1 m of sand will have a 3 mm diameter hole located to within 150 mm.
Successful surveys have been performed on 640,000 m2 of uncovered LLDPE
geomembrane, and on geomembranes under 5.5 m of heap leach ore (25 mm hole
located), 5 m of municipal solid waste, and 18 m of industrial waste (Peggs and
McEuen, 2001).

In addition to these portable techniques there are several techniques for installing a
series of electrodes underneath the liner during construction that will subsequently
continuously monitor a lining system for leaks, indicating when a leak occurs and
approximately where (Nosko, de Meerleer, Arndt). Usually a portable method is
required to identify the exact location of a leak. General experience shows that only
occasionally do leaks develop after the liner is placed in service.

Leak statistics

Nosko and colleagues (Nosko et al. 1996, Nosko and Touze-Foltz, 2000) have presented
some very useful statistics of liner damage that occurs during construction, which shows
where the emphasis should be during CQA. Table 1 shows that most damage occurs
during placement of the soil layer over the geosynthetics. Most damage is caused by
stones and heavy equipment.

Therefore, in the case of landfills it is essential the CQA program does not finish when
the geomembrane is installed but continues as the geosynthetics are covered.
However, in uncovered pond liners, most of the damage is at seams.

Table 1. Type of damage and when it occurs

When Installation Covering After Covering


Type 24% 73% 2%
Seaming 79%
Stones 17% 60%
Cuts 4%
Stakes 16%
Equipment 16% 64%
Components 27%
Weather 9%

Table 2 shows where most of the damage occurs most is on the floor, not at corners or
penetrations where it might be expected to be. Then Tables 3 through 7 show the
types and frequency of damage at each of these locations.

\\Lara\SharedDocs\site\tech_docs\ensure_longterm_perf.doc
Table 2. Location of damage

Amount of Flat Floor Corner edge Under a Pipe Other


Damage drainage pipe Penetrations

4194 3261 395 165 84 289


100% 77.8% 9.4% 3.9 % 2.0% 6.9%

Table 3. Type and frequency of damage on floor

Type of failure Number of holes %


Stones 2641 81.00
Heavy equipment 430 13.20
Worker 130 4.00
Cuts 33 1.00
Welds 26 0.80
Total 3261 100.00

Table 4. Type and frequency of damage at corners/edges

Type of failure Number of holes %


Stones 234 59.20
Heavy equipment 75 18.90
Worker 14 3.50
Cuts 4 0.90
Welds 69 17.50
Total 395 100.00

Table 5. Type and frequency of damage under drainage pipes

Type of failure Number of holes %


Stones 50 30.30
Heavy equipment 24 14.30
Worker 24 14.50
Cuts 23 13.70
Welds 45 27.20
Total 165 100.00

Table 6. Type and frequency of damage at pipe penetrations

Type of failure Number of holes %


Stones - -
Heavy equipment - -
Worker 7 8.50

\\Lara\SharedDocs\site\tech_docs\ensure_longterm_perf.doc
Cuts 1 0.60
Welds 77 90.90
Total 84 100.00

Table 7. Type and frequency of damage at other locations


(road access, temp. storage, concrete structure, etc.)

Type of failure Number of holes %


Stones 60 20.60
Heavy equipment 125 43.40
Worker 56 19.30
Cuts - 0.00
Welds 48 16.70
Total 289 100.00

Similar statistics for a 640,000 m2 single liner placed on a soft wet peat subgrade are
shown in Table 8. There was an average of 2 holes/ha in this liner that was subjected
to comprehensive effective CQA. However, it should be noted that that the leak
survey was performed as construction progressed so all holes identified by the CQA
team had not been repaired prior to performing the survey.

Table 8. Large single liner leak statistics re. types and sizes of holes.

Size (mm) Punctures Gouges Cut Tears Burns Scrapes Lack Seam
s Of bond
<1 10 1 2 1 1 1
210 28 1 8 7 4 1
1150 7 11 7 2 3 2 1
51100 1 3 1 1 3
101-500 1 1 1 1
501-1 m 1 2
>1m 2 1 1
Unknown 4 3 1 2 1 2

Total 50 16 13 5 8 17 10 12
% total 38.2 12.2 9.9 3.8 6.1 13 7.6 9.2

Infrared spectrometer surveys

Electrical surveys can be performed at the rate of about 1 ha/day but they are not
particularly suited for doing landfill caps. However, if there are leaks in caps they will be
emitting landfill gas (LFG). A portable multichannel analyzer, effectively an infrared
spectrometer, has been developed that will analyze methane, carbon dioxide, and

\\Lara\SharedDocs\site\tech_docs\ensure_longterm_perf.doc
total hydrocarbons several times every second. This equipment was used to sample
LFG from about 150 mm above a 6 ha landfill cap surface while being carried in a truck
at about 15 km/hr. Location was monitored by global positioning system equipment.
The complete site was monitored on parallel tracks approximately 1.75 m apart as
shown in Figure 10 in 1.5 days. This included identifying over 30 leaks then searching
each of these leaks for the exact location of the maximum concentration. It is
estimated that this technique could survey up to 150 ha per day, depending on
topography.

Figure 10. Landfill cap leak survey record

This technology has also been used to locate the source of actual cap leaks on slopes
and near culverts and roads when the surface gas concentration peak could be some
distance form the actual leak. It was found that surface gas concentrations could be
very localized (to within 50 mm) and that the actual leak could be several metres away
from the maximum surface gas concentration.

Development of this technology is underway to apply it to the evaluation of newly-


installed bottom liners, both with and without soil covers, and to large evaporation
pond and heap leach pad liners - the survey time could be decreased to 1% of that
required for geoelectric surveys.

\\Lara\SharedDocs\site\tech_docs\ensure_longterm_perf.doc
Ultrasonic and infrared thermography seam testing

Two methods, ultrasonics and infrared thermography, have been proposed for the
nondestructive evaluation of geomembrane seam bond strength, in order to avoid
cutting holes in seams for destructive peel and shear tests. Ultrasonic measurements of
seam thickness were attempted and terminated many years ago by Schlegel Lining
Technology. They have recently been revived in Region 2 of the United States
Environmental Protection Agency (USEPA). A pulse-echo technique is used to
essentially measure the thickness of the completed seam. Based on German
geomembrane seam analyses (Lders, 1996) the USEPA rationale is that if seam
thickness is between 0.3 and 0.8 mm less than twice the thickness of the
geomembrane, sufficient melting and mixing at the interface has occurred and the
seam is acceptable. Measurements are made every 8 m along the length of the
seam, a considerable improvement on the 150 m sampling frequency for destructive
tests.

However, it is apparent that a reduction in seam thickness can be achieved in two


ways by high pressure and low temperature, or by high temperature and low pressure.
It is doubtful that both methods generate the same quality of weld bonding. Lders
(2000) clearly elaborates that the thickness reduction is a secondary effect of requiring
a minimum depth of melt zone at the interface. Therefore, there must also be some
evaluation of weld zone morphology before finished seam thickness alone can
meaningfully be used as an acceptance criterion.

The transducers used for ultrasonic testing are conventional flat-ended units that require
a couplant to transfer the signal between transducer and geomembrane. This is time
consuming, requires that seams be supported by the subgrade, and means that
extrusion seams, with their rough surface profiles, cannot be tested. Extrusion seams
are of a lower quality than double wedge seams so are more important to test. Non-
contact Airscan ultrasonic transducers are now available that can be traversed
along a seam with a transmitting transducer inducing a signal in the geomembrane on
one side of the seam and a receiving transducer monitoring changes in the signal that
is picked up from the other sheet after the signal has passed through the weld
interface. Thus every millimetre of seam, even on wrinkles and pipe boots, can be
tested - extrusion seams as well.

However, infrared thermography (IRT) offers a more rapid and direct method of
evaluating seam bond strength and internal flaws nondestructively (Peggs et al, 1996).
This method requires the surface of the weld to be heated about 10oC and then to
monitor the surface temperature a second or two later as heat is conducted into the
geomembrane. A thermogram is shown in Figure 11. Heat is rapidly absorbed at good
seams (blue) and is not absorbed where the bonding is poor (red). Voids and grains of
sand in the seam can clearly be seen, as can blockages in the center air channel.
Even changes caused by the thermal cycling of the hot wedge can be identified. This
is far more effective than performing peel and shear tests.

\\Lara\SharedDocs\site\tech_docs\ensure_longterm_perf.doc
Figure 11. IR thermogram of double track seam.

With appropriate artificial Intelligence (AI) it is projected that seams could be surveyed
at about 10 km/hr, with flaws being analyzed in real time and with the location and
type of critical flaws being identified by a spot of spray paint on the liner. A hard copy
(videotape) of every millimetre of the seam would be available for further examination.
Such technology is available now for special investigations, but needs the development
of the AI for routine CQA purposes.

GEOMEMBRANE LINER LIFETIME

In general geomembrane materials of all types are providing very satisfactory service,
but we still do not have much more than 40 years experience with PVC, 25 years with
HDPE and 8 years with PP and GCLs. However, despite this tremendous success, there
have been liner installations that have not lasted 8 years or even 6 months, and there
are some that have not worked on commissioning. These are typically due to a lack
of knowledge by at least one party to the project. This is, unfortunately, to be
expected in a new, growing industry, in which those with appropriate knowledge do
not have the resources to educate the large number of people entering the industry.
Thus, while the level of knowledge is increasing at the leading edge, the overall level of
knowledge of those involved in the complete industry may, in fact, be decreasing!
However, if a geomembrane-based lining system is properly designed, the correct
material selected, the material (and its grade) is properly specified, it is properly
installed, knowledgeable CQA is performed, it is properly tested, and the facility is

\\Lara\SharedDocs\site\tech_docs\ensure_longterm_perf.doc
properly operated, it is felt that a buried geomembrane should provide a service life of
400 years or more. Figures of 700 years have been proposed for HDPE.

SUMMARY

Geomembranes are a very cost-effective way of building durable resource and


environmental protection barrier systems, but only if it is recognized that there are
clearly commodity and specialty applications. The commodity philosophy applied to
specialty applications is responsible for the increasing number of failures. It is even
responsible for failures in some commodity applications.

The problem is not, typically, with the materials themselves, but with a lack of
knowledge of the performance characteristics of the different geomembrane materials
by design engineers, regulators, and CQA personnel.

HDPE is an excellent geomembrane material. It is the predominant geomembrane


material and will remain the predominant geomembrane, but it is not suited to all
applications. Its susceptibility to stress cracking must be recognized and factored into
its specification a higher resistance being required for more critical applications or for
construction on sites where the material will suffer more abuse.

Similarly, more critical liner applications require the input of experienced design
engineers, manufacturers, installers, CQA personnel and testing laboratories. The
additional cost is simply a worthwhile insurance policy. After all, the use of
geosynthetics in the first place provides a large economic advantage over
conventional engineering practice so a 43% saving in cost rather than a 45% saving is of
little significance. Consider the $10,000 cost of a chemical resistance test or for a final
geoelectric integrity survey compared to $1,000,000 or more for a major failure. There
is no comparison.

Therefore, first evaluate the technical competence of all project parties, then consider
the costs. Under no circumstances should costs be first, particularly in the more critical
and difficult applications.

REFERENCES

ASTM D3895 The Method of Oxidative-Induction Time of Polyolefins by Differential


Scanning Calorimetry, ASTM Annual Book of ASTM Standards, Vol 08.02, West
Conshohockern, PA, USA.

ASTM D1928-90, Standard Practice for Preparation of Compression-Molded


Polyethylene Test Sheets and Test Specimens, ASTM Annual Book of ASTM Standards,
Vol 08.02, West Conshohockern, PA, USA.

Butts, E.O., Lord Jr., A.E., Electrical Methods of Seaming of Geomembrane Sheet,
Proceedings of Geosynthetics 91, IFAI, Roseville, MN, USA, pp 425-438.

\\Lara\SharedDocs\site\tech_docs\ensure_longterm_perf.doc
Cadwallader, M.W., 2001, Textured HDPE Geomembrane Variability Effects on
Constant Load Stress Crack Testing, Geosynthetics Conference, IFAI, Roseville, MN,
USA, pp 847-858.

Hsuan, G.Y., Koerner, R.M., Lord Jr., A.E., The Notched Constant Tensile Load (NCTL)
Test to Evaluate Stress Cracking Resistance, Proceedings of the 6th GRI Seminar
MQC/MQA and CQC/CQA of Geosynthetics, 1992, GRI, Philadelphia, PA, USA, pp 244-
256.

Lders, G., 1998, Assessment of Seam Quality and Optimization of the Welding Process
of HDPE Geomembranes, Sixth International Conference on Geosynthetics, IFAI,
Roseville, MN, USA, pp 337 344.

Lders, G., 2000, Quality Assurance in Hot Wedge Welding of HDPE Geomembranes,
Proceedings of the Second European Geosynthetics Conference 2000, AGI-IGS,
Bologna, Italy, pp 591-596.

Nosko, V., Andrezal, T., Gregor, T., & Ganier, P., 1996, SENSOR Damage Detection
System (DDS) the Unique Geomembrane Testing Method, Proceedings of
Geosynthetics: Applications, Design and Construction, Balkema, Rotterdam, pp 743-
748.

Nosko, V., Touze-Foltz, N., 2000, Geomembrane Liner Failure: Modelling of its Influence
on Contaminant Transfer, Proceedings of the Second European Geosynthetics
Conference 2000, AGI-IGS, Bologna, Italy, pp 557-560.

Peggs, I.D., 1985. Why Quality Control? A Graphic Case History, Geotechnical Fabrics
Conference, Cincinnati, OH, USA.

Peggs, I.D., 1996, A Reassessment of HDPE Geomembrane Seam Specifications,


Geosynthetics: Applications, Design and Construction, Balkema, Rotterdam, pp 693-
696.

Peggs, I.D., 1999. Mobile Geoelectric Liner Integrity Surveys: Planning Ahead,
Proceedings of Geosynthetics 99, IFAI, Roseville, MN, USA, pp 627-634.

Peggs, I.D., 2001, Three Challenging Electrical Integrity/Leak Surveys on Uncovered


and Deep Waste-Covered Liners, Geosynthetics Conference 2001, IFAI, Roseville, MN,
USA, pp 245-262.

Peggs, I.D., Miceli, G.F., McLearn, M.E., 1994, Infrared Thermographic Nondestructive
Testing of HDPE Geomembrane Seams: A Feasibility Study, 5th International Conference
on Geotextiles, Geomembranes and Related Products, Singapore, pp 941-944.

Thomas, R.W., Woods-DeSchepper Beth, Stress Crack Testing of Unnotched HDPE


Geomembranes and Seams, Proceedings of the 7th GRI Seminar Geosynthetic Liner
Systems: Innovations, Concerns and Designs, GRI, Philadelphia, PA, USA, pp 116-125.

\\Lara\SharedDocs\site\tech_docs\ensure_longterm_perf.doc

You might also like