Sie sind auf Seite 1von 18

Geotextiles and Geomembranes 13 (1994) 263-280

© 1994 Elsevier Science Limited


! Printed in Ireland. All rights reserved
0266-1144/94/$7.00
ELSEVIER

Expanded Polystyrene (EPS) Geofoam: An Introduction to


Material Behavior

J. S. H o r v a t h
Civil EngineeringDepartment, Manhattan College, Bronx, New York 10471, USA
(Received 21 May 1993; accepted 8 November 1993)

ABSTRACT

The new geosynthetic product category of 'geofoam' was proposed in 1992.


It encompasses polymeric and non-polymeric foams that are used in
geotechnical applications. Geofoams perform functions that traditional
geosynthetic products cannot. In addition, geofoams can be used to
complement or enhance the function of other geosynthetics. Thus, geofoams
used either alone or with other geosynthetics offer new, cost-effective solu-
tions to a wide variety of geotechnical problems. This paper provides an
overview of the geotechnically relevant engineering properties of a specific
geofoam material called expanded polystyrene (EPS), which is a type of
"rigid' plastic foam. EPS is the most widely used geofoam material. There
is approximately 30 years experience of using it in geotechnical applica-
tions, primarily as thermal insulation and ultra-lightweight fill (its density
is only about 1% of the density of soil).

INTRODUCTION

Rigid plastic foams were developed around 1950 and have seen consistent
use in geotechnical applications since the early 1960s. However, only recently
has it been proposed to consider such materials geosynthetics (Horvath,
1991) under a new product category called 'geofoam' (Horvath, 1992a).
Other names used previously in the geotechnical literature when referring to
such materials include geoblock, geoboard, geoinclusion, and geosolid.

263
264 J . S . Horvath

There are several reasons for creating the product category of geofoam
at this time:
1. to foster more-widespread recognition and use of the proven
geosynthetic functions and geotechnical applications of these mate-
rials among civil engineers;
2. to improve the understanding of the engineering properties of these
materials using the knowledge and insight gained through the
development of geosynthetics technology; and
3. to encourage research into new geosynthetic functions and
geotechnical applications.
However, recognition of geofoam as a geosynthetic product category will
require broadening of the definition of geosynthetics, as most are limited
currently to products that are planar in shape (Rigo, 1992).

MATERIALS

A summary of potential geofoam materials identified to date is given by


Horvath (1993a). Most geofoam materials are polymeric, although glass
foams are also used. Geofoams are cellular (generally closed cell) in
structure. A gas called a blowing agent is used to create the cellular
structure. Geofoams may be preformed (typically in prismatic shapes)
prior to installation or formed in place. The latter is useful if some irre-
gularly shaped volume requires filling.
There is considerable confusion outside the foam industry concerning
the proper terminology to be used for polymeric foam materials. Specifi-
cally, a colloquial practice has evolved in many countries in which the
registered tradename of one manufacturer's brand of a particular foam
material is used in a generic way, i.e. when another manufacturer's brand
of the same material or even a completely different material is actually
meant. Examples of this incorrect practice are the indiscriminate use of
'styrofoam' in the US and 'styropor' in Germany. As with all geosyn-
thetics, use of these and other tradenames should be limited to those
situations where reference to a specific commercial product is actually
intended and relevant.
Although several materials have been used as geofoams, one has been,
and continues to be, used much more extensively than any other. This
material is expanded polystyrene (EPS), called expanded polystyrol in
Japan and possibly elsewhere. The reasons for the widespread use of EPS
compared to other foams with similar engineering properties are as
follows:
Expanded polystyrene geofoam: Material behavior 265

1. it is available worldwide;
2. it is the least expensive by a significant margin;
3. it is the only polymeric foam that does not use as a blowing agent
CFC, HCFC, or a similar gas linked to the depletion of the Earth's
upper-atmosphere ozone layer; and
4. it does not release formaldehyde, a toxic gas produced for exten-
ded periods (years) by some polymeric foams after their manu-
facture.
Because EPS is the predominant geofoam material used to date, the
remainder of this paper focusses on it.

F U N C T I O N S A N D A P P L I C A T I O N S OF EPS

The geosynthetic functions of EPS geofoam that have been identified to


date are listed below:
- - thermal insulation
- - lightweight fill
- - compressible inclusion
- - small-amplitude wave damping (ground vibrations and acoustics)
Some, but not necessarily all,of these functions can be provided by
other geofoam materials. Descriptions of these functions, as well as
discussion of actual and potential applications, can be found in Horvath
(1992a, b, 1993c). Because EPS geofoam can perform functions that
traditional geosynthetic products cannot, it offers geotechnical engineers
a new tool that can be used for developing solutions to a wide variety
of problems.

M A N U F A C T U R E OF EPS

P r i s m a t i c b l o c k s

EPS can only be manufactured prior to installation, i.e. it cannot be


foamed on or in the ground. The predominant type of EPS product for
geofoam applications is a prismatic block (simply called a 'block' or
'billet' in the industry) that is a rectangular parallelepiped in shape. This is
a generic (commodity) type of product, as patents on it expired many
years ago. Blocks are produced by a manufacturer called a block molder
using beads of expandable polystyrene from a resin supplier. Expandable
266 J. S. Horvath

polystyrene beads consist of polystyrene and dissolved pentane (the


blowing agent). The beads may also contain a fire-retardant additive.
Beads range from 0.2 to 3 m m in diameter (medium to coarse sand-sized)
but are typically toward the smaller end of the range. Bead size does not
affect the engineering properties of a block. The beads are colorless and
result in a block that is white in color. However, dyes can be introduced
during the manufacturing process to produce a block that is other than
white in color. The reason that this is done is for product identification or
marketing-related purposes.
EPS blocks are manufactured in a two-stage process of pre-expansion
followed by molding. The pre-expansion stage consists of placing the
beads within a container and heating them with steam to between 80 and
l l0°C. Heating softens the polystyrene and vaporizes the pentane.
Expansion of the pentane within the polystyrene produces individual
spheres, each approximately 50 times larger in volume than the original
bead, with each sphere containing numerous closed cells. After allowing
the expanded spheres (called 'pre-puff' in the industry) to cool and stabi-
lize for at least several hours, the second stage of manufacture is
performed. The pre-puff is placed in an enclosed, fixed-wall mold, and the
spheres are simultaneously re-softened and expanded further using steam.
In modern molding equipment, this is done under a partial vacuum. As a
result, a block is formed consisting of numerous particles, each polyhedral
in shape as a result of the original expanded spheres being forced together
during the additional expansion that occurs during the molding stage,
with small or no voids between polyhedra. Each polyhedron retains the
numerous closed cells of the original expanded sphere from which it
evolved. No adhesive is used to join the polyhedra, as the contacts
between them are fused during molding. This structure of individual
polyhedra provides a way to visually differentiate EPS from other foams
that have a homogeneous cellular structure, although the extent to which
each polyhedron is visible in EPS depends to a significant degree on the
size of the beads used initially. This characteristic structure of EPS is also
the reason that it is sometimes referred to colloquially as 'beadboard' or
'molded bead expanded polystyrene'.
After release from the mold, a block is allowed to 'season', typically for
several days. The reason is that some dimensional change (usually
shrinkage but sometimes swelling) may occur after molding at a rate that
decreases rapidly with time. This dimensional change is a result of cooling
and final outgassing of the pentane blowing agent from within the cells. In
most cases, trimming is required after seasoning to produce a block that
meets specified dimensional tolerances. This trimming is done using a
special machine equipped with heated metal wires. In the US, typical final
Expanded polystyrene geofoam: Material behavior 267

dimensions for trimmed blocks are 610-762 mm high by 1220 mm wide by


2440mm long. However, there is increasing use of blocks that are
4880mm long on projects involving lightweight fill applications. Block
dimensions in other countries may vary but are similar.
A seasoned block can also be cut into thinner panels for some geotech-
nical applications such as thermal insulation below roads, above clay
liners in landfills and waste-containment facilities, or adjacent to below-
ground walls. In these applications, geofoam panels that are only
25-100mm thick are usually needed. It is also possible to cut a block into
intricate shapes and this is being done increasingly for architectural
purposes. Precision cutting for thin panels or architectural shapes is also
performed at the molding plant using a machine with heated wires.
However, any cutting of blocks at a job site is generally done using a chain
saw.
Scrap produced as a result o f trimming and cutting a block within the
molding plant is usually recycled by grinding it into small pieces and
blending it with pre-puff to produce new blocks. This recycled material,
called 'regrind' in the industry, typically represents 10-15% of the total
material in a block. The use of regrind is separate from using 'post-
consumer' recycled material.

Other products

As discussed previously, an EPS block can be cut or shaped to fit a parti-


cular application. However, it is sometimes found more economical to
directly form the EPS into a desired shape rather than cut from blocks.
This is done by placing the pre-puff into custom-shape molds for the
second stage of manufacture. For example, the intricately shaped cushion
packaging found around consumer electronics equipment is generally
manufactured in this way. Some products for geotechnical application,
such as relatively thin panels for thermal insulation or drainage geocom-
posites, are also manufactured in this manner.
Another EPS product exclusively for geotechnical application is manu-
factured by taking the pre-puff and gluing the extended spheres into an
open structure (as opposed to molding) so that the finished product has a
significant porosity and fluid-transmission capability in all directions. This
product, typically produced as panels 1220 mm square and 40 mm or more
thick, is generally used as a drainage geocomposite by placing a geotextile
on one face of the panel. Benefits offered by this product compared to
other drainage geocomposites include its ability to be used for the func-
tions of thermal insulation and compressible inclusion in addition to
providing drainage.
268 J . S . Horvath

BASIC PROPERTIES OF PRISMATIC BLOCKS

Product density

The relative amount of bead expansion during the first stage of manu-
facture is controllable within certain limits. As a result, it is possible to
produce blocks of different densities within certain limits and precision.
This is important, as the geotechnically relevant material properties of
EPS correlate well with the density of the material. Thus, EPS density can
be considered a useful index property of EPS-block geofoam.
EPS blocks can be produced within a range of densities between
approximately 10 kg/m 3 (0.6 lb/ft 3) and 40 kg/m 3 (2.5 lb/ft3). Densities in
the range of 15 kg/cm 3 (1 lb/ft 3) to 30 kg/m 3 (2 lb/ft 3) are most common in
practice, typically available in density increments of approximately 5 kg/
m 3 (0.3 lb/ft3). In most geotechnical applications to date where EPS has
been used as lightweight fill, 20 kg/m 3 (1.25 lb/ft 3) material has been used.

Durabihty

EPS is non-biodegradable, chemically inert in both soil and water, and


provides no nutritive value to living organisms or animals (BASF, 1990).
However, in certain climates burrowing insects such as termites and
carpenter ants have been found to penetrate EPS (and other polymeric
foams as well) for nesting purposes. Within the past few years, an additive to
EPS has been developed in the US that is intended to prevent such infesta-
tions. This additive does not affect the engineering properties of EPS.
Polystyrene is inherently hydrophobic, and the closed-cell structure of
EPS prevents the absorption of water into each expanded polyhedron.
However, water, either in the vapor or liquid phase, can enter the small
voids between the fused polyhedra. This has no effect on mechanical
properties and will not cause any volume change of the EPS. However, the
presence of liquid water will affect the thermal-insulative properties
(BASF, 1990). This is further discussed below.
When exposed to ultraviolet (UV) radiation for an extended period of
time, the surface of EPS will yellow and become somewhat brittle (BASF,
1990). Thus, as with most other polymeric geosynthetics, long-term expo-
sure to UV radiation is undesirable.
There are a few liquids that will dissolve EPS. For geotechnical appli-
cations, the ones of greatest concern are motor-vehicle fuels (gasoline and
diesel fuel). However, in applications such as road embankments where
exposure to spilled fuels is a potential problem, the top surface of the EPS
is easily protected by a geomembrane or other material.
Expanded polystyrene geofoam: Material behavior 269

Polystyrene will melt when exposed to temperatures in excess of


approximately 150°C and is flammable if an ignition source exists (BASF,
1990). Expandable polystyrene beads that contain a flame-retardant
additive (called 'modified beads' in the industry) are available worldwide.
However, use of modified beads in EPS block produced for geotechnical
applications is not routine in all countries, so use of modified beads
should be specified if desired. There is no inherent visual difference
between EPS made from regular and modified beads, nor is there a
difference in engineering properties.

E N G I N E E R I N G P R O P E R T I E S OF P R I S M A T I C BLOCKS

Application of the currently identified functions of EPS-block geofoam


requires knowledge of the following material-behavior properties:
- - load-deformation, primarily in compression
- - thermal

The remainder of this paper provides an introduction to, and discussion


of, these properties. It should be noted that the properties discussed are
for nominally virgin material made from pre-puff plus the typical 10-15%
of regrind. Preliminary, unpublished test results indicate that EPS block
manufactured using greater proportions of recycled material, especially if
post-consumer in nature, may have significantly different properties,
especially in load-deformation in compression. Discussion of the engi-
neering properties of EPS block manufactured with significant propor-
tions of recycled material is beyond the scope of this paper, but it is likely
to be an important issue in years to come.

L o a d - d e f o r m a t i o n behavior in c o m p r e s s i o n

Short-term loading

Behavior in this mode is generally determined by laboratory testing of


small specimens that are cubic in shape (typically 50 m m in each dimen-
sion) under strain-controlled unconfined axial loading. Strain rates are
rapid, between 1 and 20% per minute, with a rate of 10% per minute most
common. Test conditions are typically those of a laboratory environment
(approximately 23°C and 50% relative humidity). Such tests are referred
to as 'short-term' tests in this paper.
A stress-strain curve from a short-term test of a specimen with a density
close to the commonly used 20 kg/m 3 is shown in Fig. 1 and is qualita-
270 J. S. Horvath

1000

I
reference: BASF (unpublished dafo)
900

800
/
700

v 600

e
L
500

400

oE
(J
300

200

100

0
0 10 20 30 40 50 60 70 80 90 1O0

Compressive Strain (%)

Fig. l. Load-deformation behavior of 21 kg/m3 EPS under short-term unconfined axial


compression loading.

tively typical of the behavior of other EPS densities. Key behavioral


aspects are listed below.
1. Linear-elastic behavior up to between 1 and 2% strain. The elastic
limit increases with increasing EPS density. Some researchers (e.g.
Du~kov, 1991) have reported pseudo-non-linear behavior at small
(< 1%) strain levels that appears to be caused by surface irregula-
rities on test specimens and/or difficulties with the test apparatus
measuring strains accurately and the relatively small load magni-
tudes typical of testing EPS specimens.
2. Yield occurs over a range, not at a single point. However, it is
traditional to identify a single-value 'compressive strength' that is
defined as the compressive stress measured at some arbitrary strain
level. This arbitrary strain level is usually 5 or 10% (the latter is
more common), which corresponds approximately to the end of the
yield range. The difference in compressive strength between these
two definitions is relatively small for a given test specimen.
Compressive strength increases with increasing EPS density.
Expanded polystyrene geofoam: Material behavior 271

3. Post-yield behavior is work-hardening in nature. Stress-strain


behavior in the post-yield range is initially linear. Eriksson and
Tr~ink (1991) found that the slope of this linear post-yield region is
approximately 5% of the slope in the initial elastic range. At larger
strains, behavior becomes non-linear; eventually the cells are
completely crushed, with the behavior reverting to that of the
original polystyrene beads.
Load-deformation behavior is affected by temperature, with compressive
strength (as defined above) increasing with decreasing temperature and
vice versa (BASF, 1987). Over a range in temperatures likely to be
encountered in geotechnical practice (from below 0 to about 45°C), the
variation in strength is approximately d:20% relative to that obtained
under typical laboratory conditions.
Other important behavioral aspects are as follows:
1. The initial tangent Young's modulus in the elastic range (Eti) exhibits
approximately linear correlation with EPS density. Figure 2 shows
results obtained by four research groups. In each case, the results are
plotted for the range in densities over which tests were performed by
that group. The curve labeled 'BASF' represents the mean of data
tabulated in BASF (1991). BASF indicated maximum variations in
Eti of + I . 8 M P a from the mean values, independent of density.
Eriksson and Tr~ink (1991) obtained the following empirical rela-
tionship for the average of data on a plot of log Eti v e r s u s density:

Eti = 0.0097p 2 - 0-014p + 1.8 (1)


where Eti has units of MPa and p is the EPS density in kg/m 3. For a
given density, the maximum variation of Eti from this average curve
was found to be approximately +0.5 MPa at low densities, increas-
ing to +1.0 MPa at higher densities. Magnan and Serratrice (1989)
obtained the following empirical relationship (same units as eqn 1)
for the mean of data on an arithmetic plot of Eti v e r s u s density:

Eti = 0-479p - 2.875 (2)

with a variation in Eti of approximately + 1-5 MPa about the mean,


independent of density. Finally, the line labeled 'van Dorp' is the
median of the range of data scaled from a plot in van Dorp (1988).
The maximum variation of Eti from this median line is approxi-
mately 4-0.5 MPa at low densities, increasing to +1-0 MPa at higher
densities (a range similar to that found by Eriksson and Tr/ink).
2. Using 20 kg/m 3 EPS, van Dorp (1988) found no plastic deformation
272 J. S. Horvath

20

KEY

BASF . . . . . . . . . . . . . . . . . . . . . . . . . . .
Erlksson and TriSnk . . . . . . . .
Mognan and Serrotriee .............................
15 van Dorp
el

:3

10
)-0

C . . " ' ~ °,o° ~~..~


.' "
#,° ~."
o° ~."
~.°

. . . . t . . . . . . . . . i . . . . . . . . . L .........
10 20 30 40
EPS Density ( k g / m '=)

Fig. 2. Correlation between initial tangent Young's modulus and EPS density.

or change in g t i after 2 × 106 cycles of straining between 0 and 1%


at a cyclic-strain rate of l0 Hz.
3. A Poisson's ratio of approximately 0.05 under initial loading can be
inferred from test data given in Eriksson and Tr~ink (1991).

Long-term loading

EPS is a thermoplastic material and time-dependent behavior would be


expected to be a design consideration, as is now recognized for all poly-
meric geosynthetics. This time-dependent behavior includes both creep
(deformation change with time under constant stress) and relaxation
(stress change with time under constant deformation). As discussed by
Koerner et al. (1993), creep and relaxation are essentially reciprocal
phenomena. A review of available data, both published and unpublished,
indicated that there has been a modest amount of creep testing of EPS
(mostly in unconfined compression) but no relaxation testing. This lack of
Expanded polystyrene geofoam." Material behavior 273

relaxation testing is similar to the findings of Koerner et al. (1993) for


other types of geosynthetics.
The shapes of creep curves for EPS in unconfined compression are
qualitatively identical to those of other engineering materials, e.g. as shown
in Fig. 5a in Koerner et al. (1983). Creep data were used to prepare Fig. 3
which shows stress-strain curves for 23-5 kg/m 3 EPS for different durations
of loading up to 10 000 h are typically the longest-duration laboratory creep
tests that are performed. The behavior of EPS at other densities is qualita-
tively similar. Key aspects of time-dependent behavior under sustained load
as inferred from unconfined-compression creep tests are as follows:
1. specimens strained initially up to 0.5% strain exhibit relatively little
additional deformation with time;
2. specimens strained initially to 1% strain exhibit modest additional
deformation (approximately 0.5% additional strain) after 10 000 h;
and

100 i ,, i i i

,' ~ " s h o r t - t e r m " test


90 ; (approximate)

80 /
; load duration (hours):
/
,' 1 10 100 1000 10000
70
Q.
v
6O

5O

40
E
0
¢..)

30

20

10
reference: BASF (unpublished data)
0 i I I I
I 2 3 4 5
Compressive Strain (%)

Fig. 3. Time-dependent stress-strain behavior of 23.5kg/m3 EPS in unconfined axial


compression (based on creep-test data).
274 J. S. Horvath

3. there is a rapid transition to a condition of significant additional


deformation with time for specimens strained initially to between
1.5 and 2% strain.
Creep effects increase with decreasing EPS density. In addition, limited
laboratory test data indicate that creep effects increase with increasing
temperature (above standard laboratory conditions) for a given EPS
density. No laboratory creep data were found for temperatures below
standard laboratory conditions, although field observations (all in rela-
tively cool climates) suggest that creep effects decrease with decreasing
temperature (below standard laboratory conditions).
Creep data can be replotted to provide a first approximation of relaxa-
tion (the reason such results are approximate is discussed in Koerner et al.
(1993)). To illustrate this, the data in Fig. 3 were used to create the solid
lines in Fig. 4, which represent compressive stress reduction with time for
different strain levels. Also shown in Fig. 4 as dashed lines are extrapola-

120 , ,

110 "'. 2.0% strain

1 O0

go lII
1.5% strain "
Q.
8o t
..Y
v i
70

60
1.0% strain
P gO
e~

E
O
40

30 0.5?0 strain

20
i
I
10
reference: BASF (unpublished data) 1
0 I I I I ~ I ~
0,0001 0.001 0.01 0.1 1 10 100 1000 10000
Time (hours)

Fig. 4. A p p r o x i m a t e relaxation b e h a v i o r o f 23-5 k g / m 3 E P S u n d e r axial c o m p r e s s i o n


(based on creep-test data).
Expanded polystyrene geofoam: Material behavior 275

tions back to the stresses measured in the typical short-term test at a strain
rate of 10% per minute. It can be seen that relaxation behavior, as infer-
red from creep data, appears to be approximately linear in log time,
especially for smaller strains. However, the preliminary results of uncon-
fined-compression relaxation tests being performed by the present author
suggest that the actual relaxation behavior of EPS is more complex than
this simple inferred behavior, although the general trends shown in Fig. 4
are valid.

Load deformation in other modes

There are three load-deformation modes other than compression in which


EPS specimens are routinely tested: tension, shear, and flexure. In each of
these tests, relatively small specimens are loaded rapidly under strain-
controlled conditions until physical breakage occurs. The calculated
maximum stress within the test specimen at breakage is defined as the
strength in that mode. The tension and flexure tests are particularly useful
in practice as index tests for what is sometimes referred to as EPS 'qual-
ity', i.e. how well the expanded polyhedra were fused during manufacture.
Figure 5 shows the variation of mean tensile, shear, and flexural
strengths with EPS density from data in BASF (1991). Also shown is the
variation in mean compressive strength using the c o m m o n stress-at-10%-
strain definition.
No information was found concerning creep or relaxation behavior of
EPS under tension, shear, or flexural loading.

Thermal conductivity

In applications involving heat flow, the most important engineering


property of EPS is its thermal conductivity. As shown in Fig. 6, the coef-
ficient of thermal conductivity of dry EPS exhibits some variability as a
function of its density and temperature. To put these values in a relative
perspective, the coefficient of thermal conductivity of soil is approximately
20-40 times greater than that of EPS.

DISCUSSION OF E N G I N E E R I N G P R O P E R T I E S

Load-deformation in compression

As with other polymeric geosynthetics, the time-dependent behavior of


EPS under load should be considered in design. Although the creep
276 J. S. Horvath

500
reference: BASF (1991)

?
flexure
• /
400 \/ •

/
/ ,,,"~,,,
/ "'" tension
300 / , o°
Q.
.x
v
~oO° shear
J~
O) •
C

200
, ,..'''''

.....y~ compression(10~ strain)

100 e"" ,~
.~.°~

O.. ~ ' ' ~ ' O ' ' " ~ elastic limit (1% strain)

0 I I I I I
10 15 20 25 30 35 40

EPS Densi~ (kg/m~)


Fig. 5. Correlation between strength (mean values) and EPS density.

behavior of EPS has been studied and understood for many years, it is
not yet fully integrated into routine design practice for geotechnical
applications. Many designs involving EPS-block geofoam under long-
term loading are still based solely on using material that has some
minimum compressive strength defined as the stress at 5 or 10% strain
in a short-term test (Aaboe, 1987), even though such short-duration tests
and strain levels do not provide fundamental insight into behaviour
under long-term loading at lower strain levels. While such a semi-
empirical approach has produced satisfactory results for more than 20
years for EPS used as lightweight fill for highway embankments and
similar applications, specific consideration of deformation with time
under service loads would appear to be a sounder approach that should
be used for all analyses in the future. This is especially true when EPS is
used in load-bearing applications for which there may be no prior
experience and is consistent with recent industry recommendations
(BASF, 1991). One way of doing this would be to use as a basic para-
Expanded polystyrene geofoam: Material behavior 277

50 . . . . . . .

~, 40
I
E
+40 °c.
E

-~ 30
J .... oo .... oo .......

o
0"C ~
/
"6 -40 °C
E
o 20

.~_

o 10

reference: BASF (1978, 1989)


I i I I I I I I

10 20 30 40 50 60 70 80 90 1 O0

EPS Density (kg/m s)


Fig. 6. C o r r e l a t i o n b e t w e e n t h e r m a l c o n d u c t i v i t y a n d E P S d e n s i t y (for d r y EPS).

meter not compressive strength but the elastic limit. Eriksson and Tr~ink
(1991) suggested an elastic limit defined as the compressive stress
measured at a compressive strain of 2% in a typical short-term test. The
present author recommends that a strain level of 1% be used for the
following reasons:
1. it is a somewhat conservative, lower-bound estimate of the elastic
limit regardless of EPS density; and
2. i t is a strain level for which time-dependent effects are relatively
small and negligible in many practical applications.
Using this new, recommended definition of elastic limit as the compressive
stress measured a t 1% compressive strain, the relationship with EPS
density is also shown in Fig. 5. Note that the traditional short-term
unconfined-compression test would still be useful in practice because it
would determine the stress at 1% strain.
278 J. S. Horvath

Thermal conductivity

All geofoam materials absorb liquid water to some degree, and their
coefficient of thermal conductivity will increase in magnitude with
increasing water content. This reduction in thermal efficiency should be
accounted for in design. As noted earlier, the structure of EPS allows
liquid water to accumulate in the small voids that exist between the fused
polyhedra. Both laboratory testing and field observations indicate that the
relative volume of accumulated water is a complex function of the
following:
1. EPS density and confining stress at the time of molding (which
affect the relative volume of inter-polyhedra voids);
2. thickness of the EPS used;
3. the phase of water (vapor or liquid) available on the surface of the
EPS;
4. when only water vapor is available, the thermal gradient across the
EPS (a thermal gradient 'drives' water vapor into the EPS where it
can then condense);
5. when liquid water is available, both the pore pressure and hydraulic
gradient across the EPS; and
6. time.
In addition, it should be noted that the process of water absorption in EPS
is reversible. Long-term field observations have indicated seasonal fluc-
tuations in the water content of EPS geofoam similar to those observed in
the vadose zone of soil.
In typical installations in which the thermal-conductivity function of
EPS is important and positive drainage of liquid water away from the
geofoam is maintained, the water content of EPS can increase only by
diffusion and condensation of water vapor driven by a thermal gradient.
Both laboratory and field tests indicate that the uptake of water in such
cases, even for relatively thin specimens and after a period of several years,
is less than 1% by volume (BASF, 1978). This causes an increase of about
5% in the coefficient of thermal conductivity (BASF, 1978). However, a
more-conservative allowance for a long-term water content in the EPS of
up to 3-5% by volume, which produce increases in thermal conductivity
of between 15 and 25%, is required by some design standards (Horvath,
1993b).
In cases where liquid water is in direct contact with the EPS, liquid
water can enter the inter-polyhedra voids directly. Observation of actual
installations involving long-term submersion of full-size blocks below the
ground water table indicates that the maximum water content of the EPS
Expanded polystyrene geofoam: Material behavior 279

in such cases is approximately 9% by volume (Aaboe, 1987). This would


result in an increase of approximately 50% in the coefficient o f thermal
conductivity (BASF, 1978). Where relatively thin (typically 5 0 m m thick)
panels of EPS are used, such as thermal insulation below roads, somewhat
greater water uptakes of 10-15% are typically allowed for in design. This
would result in an increase of approximately 50-75% in the coefficient of
thermal conductivity. However, even under extreme cases of exposure to
liquid water, EPS remains an efficient thermal insulator compared to soil.
As a result, a recent geosynthetic application of EPS has been as a thermal
insulator against both desiccation and freezing for clay liners used in
waste-containment basins (Horvath, 1993c). Previously, soil layers in the
order o f 1 m thick, which wasted potential landfill volume, were used for
this purpose.

A C K N O W L E D G E M ENTS

The author is grateful to the m a n y people and their organizations who


generously provided information concerning EPS. Particular gratitude is
due to F. Hohwiller o f BASF A G in Ludwigshafen, Germany, and several
engineers at the Norwegian Road Research Laboratory in Oslo, Norway.

REFERENCES

Aaboe, R. (1987). 13 years of experience with EPS as a lightweight fill material in


road embankments. Publication No. 61, Norwegian Road Research Labora-
tory, Oslo, Norway, p~p. 21-7.
BASF (1978). Styropor~; Technical information - - Building construction;
Highway construction, Insulation at earth level; Pavement insulation:
Guidelines. Document No. TI-1801e, BASF, Ludwigshafen, Germany.
BASF (1987). Strength characteristics of EPS thermal insulation. Technical
Bulletin E-3, Parsippany, NJ, USA.
BASF (1989). Processing of Styropor ~. Document No. HSR 8707e, BASF,
Ludwigshafen, Germany.
BASF (1990). Guidelines for use of EPS board products in building construction.
Technical Bulletin E-I, Parsippany, N J, USA.
BASF (1991). Styropor ® - - Construction; Highway construction/Ground insu-
lation. Technical Information TI 1-800e, BASF, Ludwigshafen, Germany.
Du~kov, M. (1991). Use of expanded polystyrene foam (EPS) in flexible pave-
ments on poor subgrades. In Proceedings of the International conference on
Geotechnical Engineering for Coastal Development - - Theory and Practice.
Port and Harbour Research Institute, Yokohama, Japan, pp. 783-8.
Eriksson, L. & Tr~ink, R. (1991). Properties of expanded polystyrene - - Labora-
tory experiments. Expanded Polystyrene as Light Fill Material; Technical
280 J.S. Horvath

Visit around Stockholm - - June 19, 1991. Internal Publication of the Swed-
ish Geotechnical Institute, Link6ping, Sweden.
Horvath, J. S. (1991). The case for an additional function. IGS News, 7(3), 17-18.
Horvath, J. S. (1992a). New developments in geosynthetics; 'Lite' products come
of age. Standardization News (ASTM), 20(9) 50-53.
Horvath, J. S. (1992b). Dark, no sugar: A well-known material enters the
geosynthetic mainstream. Geotech. Fabrics Rep. (IFAI), 10(7), 18-23.
Horvath, J. S. (1993a). Geofoam geosynthetics: An overview of the past and
future. Geosynth. World, 3(1), 15-17; with corrections, 4(1), 31.
Horvath, J. S. (1993b). Geofoam applications in residential construction. Paper
presented at the National Association of Home Builders 49th Annual
Convention & Exposition, Las Vegas, Nevada, USA, 20 February.
Horvath, J. S. (1993c). Editorial letter. Geotech. Fabrics Rep. (IFAI), 11(1), 4.
Koerner, R. M., Hsuan, Y. & Lord Jr, A. E. (1993). Remaining technical barriers
to obtain general acceptance of geosynthetics. Geotext. Geomemb., 12(1),
1-52; also publ. in Grouting, Soil Improvement and Geosynthetics, ed. R. H.
Borden, R. D. Holtz & I. Juran, Geotechnical Special Publication No. 30,
American Society of Civil Engineers, New York, USA, 1992, pp. 63-109.
Magnan, J.-P. & Serratrice, J.-F. (1989). Propri~t~s m6caniques du polystyrene
expans6 pour ses applications en remblai routier. Bull. liaison Lab. Ponts.
Chauss~es, 164, 25-31.
Rigo, J. M. (1992). Geo what? Terms and definitions worldwide. IGS News, 8(1), 8.
van Dorp, T. (1988). Expanded polystyrene foam as light fill and foundation
material in road structures. Paper presented at the International Congress on
Expanded Polystyrene: Expanded Polystyrene - - Present and Future, Milan,
Italy, 10 May.

Das könnte Ihnen auch gefallen