You are on page 1of 17

Geotextiles and Geomembranes 44 (2016) 656e672

Contents lists available at ScienceDirect

Geotextiles and Geomembranes


journal homepage: www.elsevier.com/locate/geotexmem

Review

A review of the performance of geosynthetics for environmental


protection*
N. Touze-Foltz a, *, H. Bannour c, 1, C. Barral b, 2, G. Stoltz d, 3
a
Irstea, Antony Regional Center, Hydrosystems and Bioprocesses Research Unit, 92761 Antony, France
b
Conservatoire National des Arts et Metiers, 2, rue Cont
e, 75141 Paris cedex 03, France
c
ISSAT, Sousse, Tunisia
d
RECOVER Research Unit, Irstea, 3275 Route C ezanne, CS 40061, 13182 Aix-en-Provence cedex 5, France

a r t i c l e i n f o a b s t r a c t

Article history: This paper, which is based on an Invited Lecture for the 7th International Conference on Environmental
Received 15 April 2015 Geotechnics, gives an updated overview of the properties of transfer of geosynthetic liner materials
Received in revised form used in environmental applications. To begin, the water-retention curves of geosynthetic clay liners
26 March 2016
(GCLs) are discussed, with the focus being on the high temperatures that can be encountered and the
Accepted 15 May 2016
Available online 7 June 2016
concomitant risk of desiccation. Next, an overview is given of quantifying advective transfer through
intact geomembranes (virgin or after exposure on site) and through multicomponent GCLs. Experi-
mental quantification of advective transfer through composite liners is also addressed, whereby
Keywords:
Geosynthetic
geomembranes or the film or coating of a multicomponent GCL is damaged. Finally, based on a
Geomembrane literature review including the most recent data, the discussion turns to the diffusion of organic and
Geosynthetic clay liner inorganic species through virgin and aged geomembranes and GCLs. The synopsis of the most recent
Advection data presented here in terms of elementary transfer mechanisms, either advective or diffusive, should
Diffusion contribute to improving the quantification of transfer through barrier systems. These four topics were
Measurement selected as they correspond to the fields of expertise of the co-authors in which they have been
publishing in the past 20 years.
© 2016 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 657
2. Experimental determination of hydraulic behavior under unsaturated conditions of geosynthetic clay liners . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 658
3. Quantification of advective transfer through geomembranes and multicomponent geosynthetic clay liners . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 659
3.1. Quantification of advective transfer through geomembranes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 659
3.1.1. Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 659
3.1.2. Flow rates for virgin geomembranes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 660
3.1.3. Flow rates for geomembranes after on-site exposure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 660
3.2. Quantification of advective transfer through intact multicomponent geosynthetic clay liners . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 660
4. Quantification of flow rates in composite liners with a GCL and multicomponent GCLS with damaged film or coating . . . . . . . . . . . . . . . . . . . . . . . . 661
4.1. Flow rates and interface transmissivity in composite liners with a geosynthetic clay liner . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 661
4.1.1. Case of virgin geosynthetic clay liners containing sodium bentonite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 661
4.1.2. Influence of geomembrane on flow rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 662
4.1.3. Influence on flow rate of concrete in contact with geosynthetic clay liner . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 663

*
Review of this paper was handled by A. Bouazza.
* Corresponding author. Tel.: +33 1 40 96 60 39.
E-mail addresses: nathalie.touze@irstea.fr (N. Touze-Foltz), hajerbannour@gmail.com (H. Bannour), camille.barral@lecnam.net (C. Barral), guillaume.stoltz@irstea.fr
(G. Stoltz).
1
Tel.: +216 58779292.
2
Tel.: +33 1 40 27 22 91; fax: +33 1 58 80 86 01.
3
Tel.: +33 442666964; fax: +33 442668865.

http://dx.doi.org/10.1016/j.geotexmem.2016.05.008
0266-1144/© 2016 Elsevier Ltd. All rights reserved.
N. Touze-Foltz et al. / Geotextiles and Geomembranes 44 (2016) 656e672 657

4.1.4. Influence on flow rate of composition of bentonite in geosynthetic clay liner . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 663
4.1.5. Influence of bentonite ageing on flow rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 663
4.1.6. Summary of results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 663
4.2. Flow rates in multicomponent geosynthetic clay liners when coating or film is damaged . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 664
5. Diffusion in lining materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 665
5.1. Composition of leachate from landfills . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 665
5.2. Quantifying properties of transfer of chemicals through liner materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 665
5.2.1. Theory of diffusive transfer through geosynthetic clay liners . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 665
5.2.2. Theory of diffusive transfer through geomembranes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 665
5.3. Absorption onto components of geosynthetic clay liners . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 666
5.3.1. Case of organic pollutants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 666
5.3.2. Case of inorganic pollutants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 666
5.4. Diffusive transfer through geosynthetic clay liners . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 666
5.4.1. Diffusion of inorganic species through geosynthetic clay liners . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 666
5.4.2. Diffusion of organic species through geosynthetic clay liners . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 667
5.4.3. Aged GCLs after cation exchange . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 667
5.5. Diffusive transfer through geomembranes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 667
5.5.1. Diffusion of inorganic species . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 667
5.5.2. Diffusion of organic species . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 667
5.5.3. Effect of geomembrane ageing on diffusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 669
6. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 670
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 671

1. Introduction the clay, which contradicts the definitions of clay geosynthetic


barriers. In the following the words GM or GCL will thus be used
Recent years have seen many advances in the understanding of instead of barriers. To clearly distinguish between the various
issues related to the use of geosynthetics such as geomembranes materials, the term “multicomponent GCL” is used systematically in
(GMs) and geosynthetic clay liners (GCLs) as contaminant barriers the following for a GCL that contains a coating or an attached film.
also respectively known as polymeric or bituminous barriers for the The function of a barrier material is to control contaminant
first ones and clay geosynthetic barriers for the latter transport and thereby isolate the contaminants from the environ-
This paper presents an updated version of an Invited Lecture for ment over the long term. To ensure this, methods to quantify
the 7th International Conference on Environmental Geotechnics contaminant transfer through liner materials were developed (some
(Touze-Foltz et al., 2014) and focuses on the performance of geo- very recently). This paper discusses quantification in the laboratory
synthetics for environmental protection (e.g., landfills or mining of the properties of transfer in geomembranes and GCLs. The
applications). The geosynthetics under study are those whose elementary transfer modes focused on herein are advection, which
function is to ensure the lining, in other words, GMs and GCLs. In is the transport of liquid due to a difference in hydraulic head be-
the Recommended Descriptions of Geosynthetics Functions, Geo- tween the two sides of a liner material, and diffusion, which is
synthetics Terminology, Mathematical and Graphical Symbols of transfer of liquid due to different concentrations of a given
the International Geosynthetics Society, GMs are defined as planar, contaminant on the two sides of a liner material. No attempt is made
relatively impermeable, polymeric (synthetic or natural) sheets for here to quantify the performance of barrier systems on the field scale
use in civil-engineering applications. GCLs are defined as an by accounting for the combined effect of advection and diffusion.
assembled structure of geosynthetic materials and low-hydraulic- This aspect was previously discussed extensively by, for example,
conductivity earth material (clay) in the form of a manufactured Rowe (2005, 2007; 2012). Instead, the objective here is to discuss the
sheet to be used in civil-engineering applications. The EN ISO 10318 material properties of geomembranes and GCLs and how they can be
standard (AFNOR, 2006a) defines geosynthetic barriers, which may quantified, with the emphasis on recent innovative developments.
be polymeric, bituminous, or clay geosynthetic barriers (GBR-Cs), The quantified elementary transfer properties can subsequently be
according to which component fulfills the barrier function. A GBR-C used for modeling at the barrier scale to predict contaminant
is defined as factory assembled structure of geosynthetic materials transfer, although this falls outside the scope of this paper.
in the form of a sheet which acts as a barrier. The barrier function is Temperature is a factor that may influence the long-term per-
essentially fulfilled by clay. As explained below, this definition does formance of barrier systems (Rowe, 2005). In waste-containment
not allow the interchangeable use of geosynthetic barriers and facilities, heat may be generated in various ways (Ali et al., 2014).
GCLs. In fact, multicomponent GCLs are now available on the For the biodegradation of waste, the temperature may rise to 60  C
market. The following proposed definitions are currently being under normal operation in municipal solid-waste landfills. Singh
discussed by the ASTM D35 terminology task group and may be and Bouazza (2014) mention that, due to the combined effect of
added in the future to the ASTM terminology standard D4439 (von solar radiation and brine processing, liners containing brines pro-
Maubeuge et al., 2011). Currently, a multicomponent GCL is defined duced from coal seam-gas wells can be exposed continuously to
as a GCL with an attached film, coating, or membrane that de- temperatures close to 70e80  C. One potential consequence of
creases the hydraulic conductivity, protects the clay core, or both. elevated temperatures is the development of thermal gradients
An adhered GCL is a GCL product in which the clay component is across the liner toward the cooler subgrade soil. Such thermal
bonded by adhesion to a film or membrane. A coated GCL is a GCL gradients can create a risk of outward moisture movement and
product with at least one layer that consists of a solidified synthetic possible desiccation of the mineral liner, which may be a GCL (Singh
substance that was applied to the GCL as a fluid (von Maubeuge and Bouazza, 2014).
et al., 2011). In this case, as illustrated in Section 3.2, the coating Thus, to adequately predict transfer through barrier systems,
or attached film may ensure the liner function much more so than understanding the hydraulic behavior under unsaturated
658 N. Touze-Foltz et al. / Geotextiles and Geomembranes 44 (2016) 656e672

conditions of GCLs is of vital importance, because their capability to The difficulty of determining the WRCs of GCLs could arise from a
serve as a barrier to fluids is intimately linked to the uptake of capillary barrier to water continuity developing between the
moisture by the bentonite and can be affected by temperature. bentonite and the water-retention equipment, which could affect
Section 2 of this paper is thus dedicated to providing a synopsis of resulting WRC.
existing data that elucidates the retention curves of GCLs. The effect on the transport characteristics of confining stress
Quantifying advective transfer in geomembranes and multi- applied along a wetting path was investigated by Abuel-Naga and
component GCLs is addressed in Section 3, where an insight is Bouazza (2010), who applied a stress of 50 kPa by using a modi-
given into advective transfer through geomembranes after expo- fied triaxial apparatus that combines a thermocouple psychrometer
sure on site. There, data are presented that were collected from and a relative humidity sensor, and by Beddoe et al. (2011), who
hydraulic applications that can be of interest for using geo- applied a small stress of 2 kPa by using a high-capacity tensiometer
membranes as, for example, landfill covers. and a capacitive relative-humidity sensor. Siemens et al. (2013)
The quantification of advective transfer through a damaged GM, used a numerical simulation done with the SEEP/W finite-
or through a hole in the attached film or coating of a multicom- element software (GeoStudio, 2007) to study how confining
ponent GCL appears in Section 4. This section also discusses the role stresses of 2 and 100 kPa affect water retention in GCLs.
played by the composition of the GM [high-density polyethylene More recently, Bannour et al. (2014) developed a new laboratory
(HDPE) or bituminous], and of the bentonite (sodium or calcium), methodology to determine the WRCs of GCLs under various
by the ageing of the bentonite through cation exchange and stresses. Constant vertical stresses of 10, 50, 100, and 200 kPa,
hydrationedesiccation cycles, and by the use of a multicomponent corresponding to different waste-layer thicknesses, were applied to
GCL rather than a more classical GM-GCL composite liner. GCL specimens placed in controlled-suction oedometers. The suc-
Finally, section 5 summarizes the parameters that, based on tion was set to mimic a wetting path from the initial dry state to
existing data from the literature, govern diffusion through GCLs and zero suction and the capillary barrier effect of the carrier geotextile
GMs for virgin and aged materials. was circumvented by using the vapor-control technique. To obtain
the two largest values of suction (4.2 and 8.5 MPa), they applied the
2. Experimental determination of hydraulic behavior under standard technique of controlling relative humidity by using a
unsaturated conditions of geosynthetic clay liners saturated saline solution. For lower suction (0.1, 0.5, 1, and 2.8 MPa),
they adapted the osmotic technique to control vapor by using
When installed in composite liners at the bottom of landfills, calibrated concentrations of polyethylene-glycol solutions,
GCLs present an as-manufactured water content close to 10%. The ensuring a continuous series of applied suction along the wetting
suction of the bentonite contained in GCLs depends on the ambient path, down to zero suction. Measurements were complemented by
relative humidity and can be as high as 1000 MPa (Beddoe et al., standard saturated oedometer swelling tests with water infiltration
2010). After installation, GCLs typically hydrate from both the at zero suction. The measurements were made on a needle-
liquid flux through defects in the geomembrane and from the punched GCL containing sodium bentonite. The experimental
transfer of vapor and liquid water from the underlying soil through water-retention curves obtained under various confining stresses
the GCL (Azad, 2011; Beddoe et al., 2010). Accurately predicting the allowed the following conclusions to be drawn, which are consis-
hydraulic characteristics of composite liners requires knowledge of tent with conclusions from previous studies:
both the water-retention curve (WRC) of the GCL and the volu-
metric changes of such liners during hydration.  Increasing vertical stress results in a decrease in water uptake
GCLs hydrate under a compressive stress corresponding to the along the wetting path accompanied by a reduction in the
overburden load generated by waste. Assuming a typical depth of swelling capacity and in the saturated hydraulic conductivity of
waste deposits of between 20 and 30 m and using a density of the GCL (Fig. 1).
800e1000 kg/m3 for the waste leads to vertical stresses of up to  In the low-suction range, a GCL hydrating at a high confining
300 kPa applied to GCLs at the bottom of landfills. stress hydrates faster than one at a lower confining stress. At a
As summarized in Table 1, many experimental studies have higher confining stress, GCLs have the benefit of a higher suction
investigated the water-retention properties of GCLs (Daniel et al., where the saturated hydraulic conductivity is achieved and, for
1993; Barroso et al., 2006a; Southen and Rowe, 2007; Abuel- a given suction, their saturated degree is higher than that of an
Naga and Bouazza, 2010; Beddoe et al., 2010, 2011; Hanson unconfined GCL.
et al., 2013). Motivated by the contrasting water-retention  Water retention in bentonites depends only slightly on
characteristics of geotextile and bentonite, these studies bentonite density, which is attributed to the predominance of
focused on determining the WRCs of GCLs under confining stress. physicochemical clayewater interactions.

Table 1
Published investigations of GCL water-retention curves.

Authors Technique used Confining stress (kPa) Water cycle

Daniel et al. (1993) Thermocouple Psychrometer (SCM)a and vapor equilibrium (MCM) 0 Wetting path
Southen and Rowe (2005) Pressure-plate technique (SCM) 0 Drying
Barroso et al. (2006a) Filter Paper (MCM) 0 Wetting path
Southen and Rowe (2007) Pressure plate (SCM) and pressure membrane extractors (SCM) 0-0.5-3-100 Drying path
Abuel-Naga and Bouazza (2010) Thermocouple psychrometer (MCM) and a capacitive relative humidity sensor (MCM)r 50 Wetting path
Beddoe et al. (2010) High-capacity tensiometers (MCM) and capacitive relative humidity sensors (MCM) 2 Drying path
Beddoe et al. (2011) High-capacity tensiometers (MCM) and capacitive relative humidity sensors (MCM) 2 Wetting/Drying path
Hanson et al. (2013) Pressure plate-Filter paper and relative humidity methods 0 Wetting/Drying path
Bannour et al. (2014) Oversaturated salt solution with forced vapor circulation and polyethylene glycol 0-10-50-100-200 Wetting
solution with forced vapor circulation
a
“SCM” stands for “suction-control method” and “MCM” stands for “moisture-control method.”
N. Touze-Foltz et al. / Geotextiles and Geomembranes 44 (2016) 656e672 659

Fig. 1. Fits made by using van Genuchten's expression in a two-dimensional representation to experimentally determined water-retention curves for various vertical stresses.

 Two well-known equations for WRCs (van Genuchten, 1980; This development gave rise to a French standard and, following
Fredlund and Xing, 1994) correctly fit the data. this, to a European standard (EN 14150; AFNOR, 2006b) for
measuring steady-state liquid flow through a geomembrane under
Additionally, Bannour et al. (2014) proposed two new expres- a hydraulic pressure difference of 100 kPa between opposite sides
sions that include stress effects for WRCs. The corresponding sur- of the geomembrane. The test method and apparatus described in
faces were represented in three-dimensional graphs that show how EN 14150 allow flows to be accurately measured down to
water content depends on stress and suction along a wetting path. 106 m3 m2 d1. The two-part cell described in this standard (see
They also discussed the possible validity of the state-surface Fig. 2) is made of stainless steel because the cell must resist
concept applied to GCLs under stress and concluded that the val- oxidation during long-term immersion. In each part of the cell, a
idity of this concept would likely be confirmed by constant-suction cavity allows hydraulic pressure to be applied. A porous disc placed
compression measurements on hydrated bentonites. in the downstream cavity prevents deformation of the geo-
These results indicate the importance of rapidly confining GCLs membrane. The cell is designed to clamp the specimen with no
once they are installed so as to make them more rapidly operational leaks. No tightening system is required because clamping between
and hydrated in barrier systems. In addition, a resistant bonding flat surfaces is usually sufficient. For bituminous geomembranes, a
structure is also recommended to naturally confine the GCL, even bitumen rubber sealant can be used.
under low stress. The minimum diameter of the measuring chambers is 0.2 m. The
cell is equipped with a liquid inlet on the upstream part, a liquid
outlet on the downstream part, and flushing valves on each part.
3. Quantification of advective transfer through
geomembranes and multicomponent geosynthetic clay liners

3.1. Quantification of advective transfer through geomembranes

3.1.1. Background
A GM is a nonporous medium, which means that the material
contains no voids, but only free spaces of size similar to that of a
solvent molecule. Transport in GMs thus occurs at the molecular
level (Lambert and Touze-Foltz, 2000). However, gases and liquids
can migrate through the intact GMs by an activated diffusion pro-
cess that differs from the liquid convection process that occurs in
the pores of porous soils (Barroso, 2005). Different driving forces
may cause diffusion; for example, concentration, hydraulic, or
temperature gradients. Diffusion has also been shown to occur even
if there is no gradient: this phenomenon is called self-diffusion
(Eloy-Giorni, 1993).
An instrument was first developed in France (Eloy-Giorni, 1993;
Pelte, 1993; Durin et al., 1998; Lambert and Touze-Foltz, 2000) to
quantify flow rates through GMs under the influence of a hydraulic Fig. 2. Photograph of stainless-steel cell and pressure volume controllers of type-B
gradient. device for measuring flow rate in geomembranes.
660 N. Touze-Foltz et al. / Geotextiles and Geomembranes 44 (2016) 656e672

10-5 measured in the range 2.9  105 to 1.5  103 m3 m2 d1 for a
hydraulic head of 0.5 m. One oxidized bituminous GM that was
inflow covered during the 30 years since its installation had a flow rate
outflow identical to that of virgin bituminous GMs, which means that no

Bitumen 3.5 mm
hydraulic evolution was detected. A recent study (Touze-Foltz et al.,
Flow rate (m3/m2/d)

10-6 2015) presents the results obtained from an elastomeric bitumi-

Bitumen 5.5 mm
nous GM and an oxidized bituminous GM after 15 years of exposure
FPP 1 mm

HDPE 1.5 mm
HDPE 1,5 mm
HDPE 1.5 mm
in a pond, during which time neither GM was covered. The flow

HDPE 2.5mm
EPDM 1.5 mm
FPO 2 mm

HDPE 0.75 mm rate through the elastomeric bituminous GM was one order of
Structured HDPE 0.75 mm
FPP 1 mm

magnitude lower than that through the oxidized bituminous GM.

PVC-P 1 mm
FPO 3 mm

PVC-P 1 mm
FPP 1 mm

10-7 The flow rate through the elastomeric bituminous GM, whose
HDPE0.5 mm
FPP 1 mm

bitumen had not been altered, was close to 1.9  106 m3 m2 d1.
Both of these studies concluded by recommending that oxidized
bituminous GMs not be left exposed after installation.
Noval et al. (2015) also recently studied the evolution over 21
10-8 years of an ethylene propylene diene terpolymer (EPDM) GM
installed in a pond in the Canary Islands. After 21 years, the flow
0 2 4 6 8 10 12 14 16 18
rate remained below 106 m3 m2 d1. Further joint studies based
Test number
on samplings from ponds are ongoing between CEDEX (Spain),
Fig. 3. Synopsis of measurements of flow rate through geomembranes made as per EN IFSTTAR (France), and Irstea (France) to better understand the
14150 with Type-B device (from Touze-Foltz and Zanzinger, 2009) (FPP: flexible temporal evolution of the properties, including the hydraulic
polypropylene, EPDM: ethylene-propylene-diene terpolymer, HDPE: high-density properties, of HDPE and plasticized polyvinyl chloride (PVC-P) GMs.
polyethylene, PVC-P: plasticized polyvinyl chloride, FPO: flexible polyolefine).

3.2. Quantification of advective transfer through intact


Volume measurements can be done by using capillary tubes
multicomponent geosynthetic clay liners
(Type-A devices) or pressure-volume controllers (type-B devices).
Fig. 3 summarizes the measurements done with a type-B device.
Multicomponent GCLs have recently been introduced to the
This device allows the application of a constant pressure when
market to seal landfills, dams, dikes, ponds etc. Because these GCLs
measuring the volume and consists of a cylinder in which a piston
combine a GCL and a coating or film, their hydraulic properties fall
slides. An automatically controlled motor applies the required
between GMs and GCLs. Very recently, instruments have been
pressure by moving the piston. Finally, a pressure sensor included in
developed to characterize transfer through these materials. Lucas
the system measures the pressure. Displacement of the piston cor-
(2002) studied hydraulic flows through a multicomponent GCL
responds to a variation of the volume of the liquid. A type-B device
composed of a uniform layer of granular sodium bentonite encap-
uses at least three temperature transducers, with one placed on each
sulated between a slit-film woven geotextile and a virgin staple
pressure-volume controller and one on the cell. Temperature mea-
fiber nonwoven geotextile. Tests in accordance with ASTM D5084
surements are then used to correct variations in volume.
(ASTM, 2010) conducted on this multicomponent GCL estimated a
The validity of a measurement is determined by comparing
hydraulic conductivity of about 5.0  1012 m/s. However, Lucas
upstream and downstream flow rates. Although these values
also suggested the possibility that sidewall leakage occurred during
should be equal in theory, in practice this is rarely the case for GMs.
the tests. Cleary and Lake (2011) estimated the hydraulic conduc-
For flow rates greater than or equal to 106 m3 m2 d1, upstream
tivity of the same multicomponent GCL by using three different
and downstream flow rates are considered to be equal if they differ
types of permeameters. All results, except for the specimens tested
by less than 10% of the flow rate measured on the upstream side. In
in the constant-head, fixed-wall, double-ring permeameter, gave
particular circumstances where testing according to the described
hydraulic conductivities below 1.0  1011 m/s.
test method gives values for a GM that lie below the threshold of
In these studies, the mass per unit area of the coating in the
sensitivity of the test method, then the liquid flow is assumed to be
multicomponent GCL was less than 100 g/m2, which resulted in a
less than 106 m3 m2 d1.
nonuniform coating, so the watertightness was ensured by the clay
core. Touze-Foltz et al. (2012a,b) and Barral and Touze-Foltz (2012)
3.1.2. Flow rates for virgin geomembranes developed a procedure whereby devices from NF EN 14150 (AFNOR,
Most flow rates obtained for virgin GMs are below the threshold, 2006b) designed to measure flow rate through GMs are combined
as can be seen from Fig. 3. In France, the definition of a GM as per the with a rigid-wall permeameter from NF P84-705 (AFNOR, 2008),
standard NF P84-500 (AFNOR, 2013) states that the flow rate which is designed to measure flow rate through GCLs, to measure
measured according to EN 14150 must be less than 105 m3 m2 d1. flow rates through multicomponent GCLs (see Fig. 4). Indeed, the
regular measurement device used in NF P84-705 to measure flow
3.1.3. Flow rates for geomembranes after on-site exposure rate through GCLs is not able to measure small flow rates in
The testing device described by EN 14150 has also been used to multicomponent GCLs. It was thus required to use a measuring
quantify the evolution of the hydraulic properties of GMs used in device for precise measurements of flow rates in GMs combined
hydraulic applications. To the best of our knowledge, no such data with the rigid-wall permeameter.
exist for environmental applications. Data obtained in hydraulic Barral et al. (2014) extended their study to include three GCLs and
applications are thus briefly reported here because they could be five different multicomponent GCLs from several manufacturers.
useful; for example, for landfill covers. Two multicomponent GCLs were coated with a coating with areal
Touze-Foltz et al. (2010a,b) reported the case of six oxidized density below 100 g/m2 and above than 200 g/m2, respectively. Three
bituminous geomembranes. They showed that, when an oxidized multicomponent GCLs were laminated with a film having various
bituminous GM remains exposed for ten to fifteen years, the hy- different thicknesses and bonding modes. The features of the GCLs
draulic performance decreases to the level of protection similar to are presented in Table 2 and the features of the multicomponent
what would be provided by a 1-m-thick clay layer, with flow rates GCLs appear in Table 3. Fig. 5 shows the results obtained for the three
N. Touze-Foltz et al. / Geotextiles and Geomembranes 44 (2016) 656e672 661

flow rate and hydraulic conductivity in regular GCLs. The flow rates
obtained (without any modification of the GCL standard) are pre-
sented in Fig. 6 and range from 1.4  1011 to 2.2  1011 m/s for a
hydraulic head of 0.1e0.6 m. These flow rates are closer to GCL flow
rates than to flow rates in GMs, which are known for virgin GMs to
be around 106 m3 m2 d1. The flow rate through multicomponent
GCLs with an adhesive bounded film thicker than 0.2 mm or with a
coating with a density exceeding 200 g/m2 is closer to the flow rate
in GMs than to the flow rate in GCLs, which implies that, in such
products, flow rate is controlled by the coating or attached film
rather than by the bentonite. Thus these GCLs cannot be considered
as clay geosynthetic barriers. They cannot be presented as GMs
because the flow rate through them exceeds the allowable flow rate
through GMs established in NF P84-500 (i.e., 105 m3 m2 d1).
Thus, the terminology standard for geosynthetics, EN ISO 10318,
lacks the appropriate vocabulary to label some multicomponent
GCLs.

4. Quantification of flow rates in composite liners with a GCL


and multicomponent GCLS with damaged film or coating

4.1. Flow rates and interface transmissivity in composite liners with


a geosynthetic clay liner

4.1.1. Case of virgin geosynthetic clay liners containing sodium


bentonite
The work done over the past years regarding the characteristics
of GCLs that are part of a composite liner mainly focused on the
situation where the GCL (which contains sodium bentonite) is
located under a hole in a HDPE GM. As indicated by Brown et al.
(1987), the flow through a defect in the GM depends on the con-
Fig. 4. Experimental device for measuring flow rates through multicomponent GCLs. tact between the GM and the underlying soil liner. According to
these authors, if the contact is not perfect, fluid that has migrated
multicomponent GCLs for which the thickness of the film coating through the defect spreads laterally within the gap (i.e., the inter-
exceeded 0.2 mm (M1 to M3, see Table 3). The results show that flow face) between the GM and the underlying soil. The area covered by
rates in multicomponent GCLs are one order of magnitude greater this interface flow is called the “wetted area.” Finally, the liquid
than those usually found for virgin GMs (i.e. 106 m3 m2 d1). migrates into and through the soil liner.
For the two multicomponent GCLs with the light coating (M4) Contact between the GM and the GCL was quantified in terms of
and with perforations by needling (M5), flow rates were measured the flow rate through the composite liner and in terms of interface
according to NF P84-705, which is the standard used to quantify transmissivity. Various situations were tested to evaluate the flow

Table 2
Characteristics of GCLs.

GCL Type of bentonite Cover GTXa Carrier GTX Thickness under 10 kPa (mm) Measured total dry mass per unit
area in specimen (kg/m2)

G1 Sodium/Powder Woven Non-woven 5.7 4.8


G2 Sodium/Granular Non-woven Woven 7.4 6
G3 Sodium/Powder Woven with surface coating and fleece Woven with surface coating 9.2 5.9
a
“GTX” stands for “geotextile.”

Table 3
Characteristics of multicomponent GCLs.

Multi Type of bentonite Cover GTX Carrier GTX Thickness Measured total Bonding type Film or coating Film or coating
component under 10 kPa (mm) dry mass per thickness (mm) measured
GCL unit area of total dry mass per
bentonite in unit area (kg/m2)
specimen
(kg/m2)

M1 Sodium Powder Woven Non-woven 5.2 4.58 Coated 0.4 < ef < 0.7 0.25 < mf < 0.4
M2 Sodium Granular Non-woven Woven 6.6 5.28 Adhered ~0.25 ~0.2
M3 Sodium Granular Slit film woven Non-woven 5.8 4.41 Adhered ~0.25 ~0.2
M4 Sodium Powder Slit film woven Non-woven 5.1 4.13 Coated _a <0.1
M5 Sodium Powder Woven/Surface Woven/Surface coating 9.6 5.7 Stitch-bonded 0.2 < ef < 0.3 ~0.1
coating and fleece
a
Not measurable.
662 N. Touze-Foltz et al. / Geotextiles and Geomembranes 44 (2016) 656e672

10-3
1,E-03

M1; differential pressure 50 kPa M1; average from latest 7 days

upstream flow rate (m3/m2/d)


M2; differential pressure 50 kPa M2; average from latest 7 days

M3; differential pressure 50 kPa M3; average latest 7 days


1,E-04
10-4

1,E-05
10-5

-6
10
1,E-06
0 1 2 3 4 5 6 7 8

Time (Days)

Fig. 5. Synopsis of upstream flow rates obtained for multicomponent GCLs for the latest seven days of measurements with a differential pressure of 50 kPa [adapted from Touze-
Foltz et al. (2012a,b)].

through a smooth GM in contact with a GCL (Harpur et al., 1993; flow rates for both confining stresses are similar under steady-
Barroso et al., 2006b, 2010). Harpur et al. (1993) verified that, un- state-flow conditions (Barroso et al., 2006b).
der steady-state conditions, the most significant fraction of the flow
occurs along the interface between the GM and the cover geotextile 4.1.2. Influence of geomembrane on flow rate
of the GCL, through the cover geotextile, and along gaps between Barroso et al. (2008) investigated the case of a textured HDPE
the cover geotextile of the GCL and the bentonite. A less significant geomembrane in contact with the GCL. Three different textured
amount of fluid percolates through the bentonite and below the geomembranes were used in these experiments. The results show
GCL. Barroso et al. (2006b; 2010) examined how hydraulic head, that the measurements are reproducible: the initial measurements
pre-hydration of the GCL, and confining stress affects the GM-GCL give a larger flow rate for smooth geomembranes than for textured
interface transmissivity. The results show that it is difficult to geomembranes; however, further measurements show that texture
identify general trends for the influence of hydraulic head, pre- has only a small impact on steady-state flow rates. This suggests
hydration, and confining stress on the interface transmissivity. that, in the early phases of such measurements, water flows more
Nevertheless, for flow rate, accounting for both the initial water easily at the interface of smooth GMs. A texture seems to reduce the
content of the specimen and the confining stress appears to be space available at the interface for water flow. However, with time,
important (Barroso et al., 2006b). The effect of confining stress on the sodium bentonite in the GCL swells, resulting in a better contact
flow rate depends on the initial water content of the specimen. In between the GM and the GCL.
fact, the flow rate in pre-hydrated GCLs is about one order of Bannour et al. (2013a) studied the effect of bituminous GMs in
magnitude larger for a confining stress of 50 kPa than for a contact with a GCL instead of a HDPE GM. This practice is not
confining stress of 200 kPa. For non-pre-hydrated specimens, the recommended for environmental applications because doubts exist

1,E-03
10-3
upstream flow rate (m3/m2/d)

1,E-04
10-4

1,E-05 M4; 60 cm; 10 kPa M4; average from latest 7 days


10-5
M5; 60 cm; 10 kPa M5; average from latest 7 days
G1; 60 cm; 10 kPa G1; average from latest 7 days
G2; 60 cm; 10 kPa G2; average from latest 7 days
G3; 10 cm; 10 kPa G3; average latest 7 days
-6
10
1,E-06
0 1 2 3 4 5 6 7 8

Time (Days)

Fig. 6. Synopsis of upstream flow rates obtained for GCLs and multicomponent GCLs for the latest seven days of measurements.
N. Touze-Foltz et al. / Geotextiles and Geomembranes 44 (2016) 656e672 663

about the chemical compatibility of the bituminous GM with of the GCL of the composite liner due to permeation by a highly
leachate for example; however, it works fine for hydraulic appli- concentrated NaCl solution that results in cation exchange.
cations or landfill covers where no environmental stakes are However, the GCL can also be subjected to wetedry cycles due to
involved. Bannour et al. reported no significant difference in terms moisture or temperature gradients generated across the whole
of flow rates compared to the combination of a bituminous GM or a barrier by climatic conditions, especially in landfill covers and
HDPE GM, independently of which bituminous geomembrane face dams. Wet-dry cycles damage GCLs; for example, desiccation with
(rough or smooth) is in contact with the GCL. shrinkage cracks leads to preferential flow paths when the GCL
hydrates (Melchior, 2002). The effect of cation exchange and
4.1.3. Influence on flow rate of concrete in contact with wetedry cycles on the hydraulic performance of GCLs has been
geosynthetic clay liner studied previously and is highly documented, especially as regards
Rowe and Hosney (2014) examined the performance of four landfill covers (Lin and Benson, 2000; Egloffstein, 2001, 2002;
GCLs for use as a hydraulic barrier below concrete-lined sewage- Melchior, 2002; Southen and Rowe, 2005; Benson et al., 2007;
treatment lagoons. Their research was based on a series of Bouazza et al., 2007; Meer and Benson, 2007; Zanzinger and
laboratory-scale measurements designed to detect the change over Touze-Foltz, 2009; Touze-Foltz et al., 2010b; Barral et al., 2012;
a 14 month period in the interface transmissivity, q, between the Benson, 2013). This effect represents the primary mode of degra-
GCLs and a 0.1-m-thick cast-in-place concrete above the GCL. The dation for bentonite in GCLs. In fact, the combination of cation
four GCLs contained either untreated bentonite, polymer-enhanced exchange and wetedry cycles more strongly affects the swelling
bentonite, granular bentonite, or powder bentonite. An increase in capacity of the bentonite and causes a greater increase in the hy-
the wastewater head from 1 to 2.5 m results in a decrease in the draulic conductivity of the GCL than does cation exchange alone, to
interface transmissivity. For the GCL with untreated granular the point that the GCL no longer acts as a hydraulic barrier
bentonite, q ¼ 1.5  1011 m2/s (under 1 m) and decreases by about (Egloffstein, 2001; Melchior, 2002; Meer and Benson, 2007; Benson
one order of magnitude under 2.5 m. The change in interface et al., 2007). In fact, after a number of wetedry cycles, shrinkage
transmissivity of the GCL with polymer-enhanced granular cracks, which occur after desiccation, may not fully heal when the
bentonite is greater than or similar to that for the GCL with un- bentonite hydrates. Cation exchange combined with wetedry cy-
treated granular bentonite. For the GCL with untreated powder cles occurring over the service life of GCLs lead to a significant in-
bentonite and with 1280 g/m2 of bentonite in the cover geotextile, crease (four to five orders of magnitude) in the hydraulic
q ¼ 1.8  1012 and 3.5  1013 m2/s at 1 and 2.5 m head, conductivity of the GCL. This raises the question of how the increase
respectively. in hydraulic conductivity affects the hydraulic characteristics of a
GM-GCL composite liner when the GM covering the GCL has a hole.
4.1.4. Influence on flow rate of composition of bentonite in Bannour et al. (2014) used laboratory measurements to address
geosynthetic clay liner the question of how cation exchange combined with wetedry cy-
The relationship between the composition of the bentonite in cles affects the flow rate and interface transmissivity of a GM-GCL
the GCL (i.e., sodium or calcium bentonite) and flow rates in the GCL composite liner. Three of the GCLs measured were exhumed from
was determined by Mendes et al. (2010), who concluded that the a dam and a fourth GCL was exhumed from a landfill. These
composition of the bentonite and the manufacturing process of the exhumed GCLs had endured cation exchange combined with
GCLs studied do not affect the transmissivity across the GM-GCL wetedry cycles, which had led to an increase in their hydraulic
interface under steady-state flow. They also noticed that, for holes conductivity and a decrease in their swell index. The flow rates of
in the GM with diameters ranging from 4 to 10 mm, the diameter these exhumed GCLs were compared with that of a composite liner
has no significant influence on the flow rate through the GM-GCL containing virgin GCLs: although the increase in hydraulic con-
composite liner. The expansion of the sodium bentonite is effec- ductivity of the GCL renders it permeable as a single liner, steady-
tive in blocking the puncture in the geomembrane, leading to a state flow rates and interface transmissivities for composite liners
significant reduction in flow rate. These results suggest that GCLs containing GCLs that were pre-exposed to cation exchange and
initially containing sodium bentonite, whose hydraulic conductiv- wetedry cycles are of the same order of magnitude as for com-
ity increases due to cation exchange, can maintain low trans- posite liners containing virgin GCLs. Thus, the flow rate through
missivity at the GM-GCL interface and low flow rate through the composite liners containing GCLs that were subjected to cation
composite liner when used in a composite liner. This question will exchange and wetedry cycles is not linked to hydraulic conduc-
be addressed in the next section. tivity, even if the hydraulic conductivity of GCLs exhumed from
field sites has increased by four to five orders of magnitude with
4.1.5. Influence of bentonite ageing on flow rate respect to virgin GCLs. Thus, ageing of GCLs is not a concern when
During its service life in a landfill barrier system, the bentonite they are used in a composite liner.
in the GCL is continuously subjected to cation exchange, whereby
sodium cations, which initially are between the bentonite platelets,
are replaced by multivalent cations that originate either in the 4.1.6. Summary of results
cover or in the bottom liner and that transfer upon contact with the Fig. 7 gives an overview of the various interface transmissivity
leachate or soil liner. Cation exchange leads to a decrease in GCL data obtained from the studies discussed above. All data were ob-
swelling capacity (Lin and Benson, 2000; Barral et al., 2012) and tained under the conditions of geomembrane-GCL contact defined
water absorption (Melchior, 2002) and to an order-of-magnitude by Barroso (2005), which means that the interface transmissivity
increase in hydraulic conductivity compared with virgin GCLs may be related to the hydraulic conductivity kGCL of the GCL as
(Egloffstein, 2001; Benson, 2013). Finally, as pointed out by follows:
Egloffstein (2001), complete cation exchange occurs after one to
log q ¼ 2:2322 þ 0:7155 log kGCL : (1)
two years when the GCL is used in unsaturated conditions. To
simulate this situation, Rowe and Abdellaty (2012, 2013) made In Equation (1), q is the interface transmissivity and kGCL is the
measurements that show that the steady-state flow rate in GM-GCL hydraulic conductivity of the GCL.
composite liners remains similar to that of virgin GCLs containing Recently, Bannour et al. (2015) defined the additional contact
sodium bentonite despite an increase in the hydraulic conductivity condition given by Equation (2) (see Fig. 7) for composite liners
664 N. Touze-Foltz et al. / Geotextiles and Geomembranes 44 (2016) 656e672

Mendes et al. (2010)


-6
10
1E-06 Barroso et al (2006)

-7
Barroso et al (2008)
10
1E-07
Barroso et al. (2010)
Interface transmissivity, θ(m2/s) 10-8 Rowe and Abdellaty (2012)
1E-08
Bannour et al. (2013a)
10
-9 Excellent Bannour et al. (2014b)
1E-09

-10
10
1E-10
-11
10
1E-11
GM- GCL (after
10-12 cation exchange and
1E-12 GM- GCL (virgin) wetting drying cycles)
-13
10
1E-13
10-12
1E-12
-11
1E-11
10 -10
1E-10
10 -9
1E-09
10 -8
1E-08
10 -7
101E-07 10-61E-06 10-51E-05

Hydraulic conductivity (m/s)

Fig. 7. Synopsis of transmissivity taken from the literature for GCLs in contact with geomembranes and for GCLs after cation exchange and wet-dry cycles.

containing GCLs whose hydraulic exceeds 1010 m/s. This contact liners with a GCL (Bannour et al., 2013b). Indeed, the results ob-
condition is valid for GCLs pre-exposed to cation exchange and tained in the steady state do fall within the range of values obtained
wetedry cycles and can also be extended to GCLs containing cal- for geomembrane-GCL interfaces.
cium bentonite. Therefore, the GM-GCL contact condition initially To address a limitation in the measurements performed by
given by Barroso (2005) for effective GCLs (i.e., kGCL less than Bannour et al. (2013b) Bannour and Touze-Foltz (2015) extended
1010 m/s) is enhanced and readjusted for all GCLs, whatever their their work by making measurements on the meter scale, so that edge
composition or field history: effects become negligible (see Fig. 8). For all multicomponent GCLs
characterized, the coating or attached film was less than 0.7 mm
log q ¼ 8:5965 þ 0:1476 log kGCL : (2) thick. Steady-state results indicate that the flow rates range from
4.61  1012 to 3.01  1011 m3/s with interface transmissivities
ranging from 1.20  1011 to 7.59  1011 m2/s, which are broadly in
line with flow rates obtained from conventional GM-GCL composite
4.2. Flow rates in multicomponent geosynthetic clay liners when liners. Consequently, when the coating or attached film is damaged,
coating or film is damaged the thickness and rigidity of the coating or attached film appears not
to affect the steady-state flow rate and interface transmissivity,
A study was recently undertaken to determine whether the flow which leads to good contact at the interface. The swell index and
rates through a multicomponent GCL with damaged coating or mass per unit area of bentonite in multicomponent GCLs influence
attached film fall in the range obtained for more classical composite the flow rate when a film is attached (i.e., glued) to the cover

Bannour and Touze-Foltz (2014) GCL 1 (m)


Bannour and Touze-Foltz (2014) GCL 2 (m)
Bannour and Touze-Foltz (2014) GCL 3 (m)
Bannour et al. (2013b) GCL 1 (dm)
Flow rate, Q (m3/s)

Bannour et al. (2013b) GCL 2 (dm)


Bannour et al. (2013b) GCL 3 (dm)

0 200 400 600 800 1000 1200


Time (h)

Fig. 8. Comparison of decimeter- and meter-scale flow-rate dynamics along multicomponent-GCL interfaces.
N. Touze-Foltz et al. / Geotextiles and Geomembranes 44 (2016) 656e672 665

geotextile of the GCL. It is thus important that the mass of bentonite Where C is the concentration in the GCL at depth z and time t, n
in the GCL be sufficient so that the swelling capacity of samples leads is the total porosity of the GCL, De is the effective diffusion coeffi-
to better contact at the interface and thus to better performance of cient, rd is the dry density, and Kd is the sorption coefficient.
the multicomponent GCL. Diffusion coefficients can be estimated by solving Equation (3)
in combination with finite-mass boundary conditions (see Rowe
5. Diffusion in lining materials et al., 2004) for the measurement setup. Note that “finite mass”
refers to the fact that the concentration changes in time at both
5.1. Composition of leachate from landfills boundaries because of mass transfer through the GCL and sampling
for the measurements. The sorption coefficient Kd is first deter-
Besides gaseous emissions, and in particular that of methane mined from batch adsorption measurements (Lake and Rowe,
and trace gases caused by the degradation of organic wastes, 2004; Rowe et al., 2005; Ganne et al., 2008; Ahari et al., 2011).
leachate emissions from landfills pose the main potential long-term Based on knowledge of the sorption coefficients of the various
environmental threat (Kjeldsen et al., 2002). Traditionally, leachate components of a GCL (geotextiles, geotextile fibers in the bentonite,
analyses as a part of regular landfill monitoring have focused on bentonite), an equivalent sorption coefficient Kdeq for the entire
nitrogen content, oxygen consumption measured in the form of GCL may be calculated by using the method suggested by Rowe
biological oxygen demand or chemical oxygen demand, heavy et al. (2005), which accounts for sorption in the geotextile, the
metals, and “classic” persistent organic pollutants such as dioxins bentonite, and geotextiles fibers in the bentonite. An overall
or polychlorinated biphenyls (Van Praagh et al., 2011). diffusion coefficient for the entire GCL can then be calculated for
Metals and metalloids are still recognized as priority pollutants. each experiment and each contaminant, treating the entire GCL as a
In contrast with most organic pollutants, metals and metalloids do homogeneous material.
not degrade in landfills. They are thus maintained in landfills and
are mobilized in the liquid or gaseous phases. The literature reports 5.2.2. Theory of diffusive transfer through geomembranes
considerable concentrations of metals and metalloids in leachate; The diffusion of vapor or aqueous permeants through geo-
for example, Cd appears at a concentration of 0.2e20 mg/L, Cr at membranes occurs in three steps: adsorption, diffusion, and
5e600 mg/L, Mn at 0.01e70 mg/L, and Fe at 0.3e220 mg/L (Pinel- desorption. First, the contaminant is partitioned between the
Raffaitin et al., 2006). source medium and the adjacent geomembrane surface. Second,
Landfills also contain micropollutants with toxic effects (acute driven by chemical potential, the compound diffuses through the
toxicity, genotoxicity, reproductive toxicity, etc.) (Sisinno et al., 2000; geomembrane. Finally, the compound partitions between the outer
Takigami et al., 2002). The presence of organic contaminants in geomembrane surface and the receiving medium (Sangam and
leachate from municipal solid-waste landfills has been clearly Rowe, 2001). When a GM is immersed for sufficient time in a
demonstrated in several countries (Oman and Hynning, 1993; Ahel fluid containing a contaminant of interest, equilibrium is reached
and Tepic, 2000; Robinson et al., 2001; Hiroshi et al., 2002). between the concentration cg in the GM and the concentration cf in
Van Praagh et al. (2011) reviewed the Swedish and international the fluid. These concentrations are related by Henry's law:
literature on analytical techniques and on the results of analysis
and quantification of several compounds or groups of emerging cg ¼ Sgf cf (4)
organic pollutants in leachate. In eleven studies of organic sub-
stances in landfill leachate in various countries, a total of 592 Where Sgf is the partition coefficient, which can be calculated
different compounds were identified. The major fraction of the from batch sorption measurements (Sangam and Rowe, 2001). The
substances identified may be categorized into the following groups diffusion of organic compounds through a GM can be modeled by
(in descending order of number of detections): phenolic com- using Fick's first law:
pounds, aromatic hydrocarbons, heterocyclic substances, carbox- 
f ¼ Dg dcg =dz ; (5)
ylic acids, phthalates, anilines, aliphatic acids, phenoxy acids,
organo phosphorous substances, terpenoids, and triazines, some of Where f is the mass flux or permeation rate per unit area, Dg is
which are used as pesticides. the diffusion coefficient of organic compounds through the geo-
Quantifying the transfer to the surrounding environment of membrane, cg is the concentration of compound in the geo-
organic pollutants from compounds emerging from leachate is thus membrane, and z is the distance parallel to the direction of
of primary importance. diffusion. According to Fick's second law, the change in contami-
nant concentration with respect to time t at any point in the GM is
5.2. Quantifying properties of transfer of chemicals through liner governed by the following differential equation:
materials
vcg v2 cg
¼ Dg 2 (6)
5.2.1. Theory of diffusive transfer through geosynthetic clay liners vt vz
Rowe and Booker (1987) developed a model for predicting the
Desorption is also described by Henry's law. The partition co-
one-dimensional transport of contaminants through soils of finite
efficient into the GM is usually equal to the partition coefficient out
thickness. The model accounts for realistic landfill parameters such
of the GM when the source and receptor fluid are the same
as dynamic surface-boundary concentrations, as is the case for
(Sangam, 2001). Because measuring the concentration of contam-
municipal solid-waste landfills. Lake and Rowe (2004), Rowe et al.
inant inside the GM is difficult, the concentration in the fluids on
(2005), Rosin-Paumier et al. (2011), and Mendes et al. (2013; 2014a)
either side of the geomembrane is used to infer the permeation
used this model to predict the one-dimensional transport of con-
characteristics of the GM. The flux associated with the diffusion
taminants through a saturated GCL for a single reactive solute
process can also be expressed as:
without degradation. The model is based on Equation (3):
dcf
vC v2 C vC f ¼ Pg (7)
n ¼ nDe 2  rd Kd (3) dz
vt vz vt
666 N. Touze-Foltz et al. / Geotextiles and Geomembranes 44 (2016) 656e672

Where Pg is the permeation coefficient or mass-transfer coefficient coefficient Kd. To study the adsorption of phenol onto bentonite,
(m2/s). they assumed a linear isotherm. The results obtained under this
The diffusion coefficient can be obtained from double- assumption range from 2.5 to 2.6 mL/g, which are well within the
compartment tests (Islam and Rowe, 2009; McWatters and Rowe, range given by Richards and Bouazza (2007) (Kd ¼ 1e5 mL/g) and by
2008, 2010; Sangam and Rowe, 2001, 2005; Touze-Foltz et al., Haijian et al. (2009) for phenol (Kd ¼ 1.2e3.3 mL/g). According to the
2012b) or immersion tests (Park et al., 2012). results for geotextiles, the adsorption of chlorophenols increases as
the number of chlorine atoms in the molecule increases. This study
5.3. Absorption onto components of geosynthetic clay liners also shows that the difference between the amount of phenolic
compounds adsorbed by the geotextile and that absorbed by the
5.3.1. Case of organic pollutants bentonite is smaller than for VOCs.
5.3.1.1. Adsorption onto geotextiles. To investigate the possible use
of geotextiles to retain pesticides in agricultural watersheds, 5.3.2. Case of inorganic pollutants
Boutron et al. (2009) measured batch adsorption and desorption of Lange et al. (2004) examined the migration of various metals (Al,
diuron, isoproturon, and azoxystrobin onto and from commercially Fe, Mn, Ni, Pb, Cd, Cu, Zn) through GCLs exposed to a synthetic
available geotextiles. Similar studies were made of the adsorption municipal solid-waste leachate. The GCLs are found to retard the
and desorption of polyamide, polyester, and polypropylene onto migration of the metals, although only under specific pH conditions.
and from polymeric fibers. Polyamide exhibited a high ability to The migration of Mn is the least attenuated. The migration of Al, Fe,
sorb diuron, with little desorption. The adsorption onto poly- and Cu are strongly retarded, so these metals are retained within the
propylene and polyester was less significant but still non-negligible clay. The migration of Ni, Zn, and Cd is moderately attenuated. In
(15% of the initial mass of pesticide in water for isoproturon and addition, Ca may have been responsible for the lack of metal reten-
30% for azoxystrobin). tion of the leachate species. Due to the higher retention at higher pH
Lake and Rowe (2004), Rowe et al. (2005), and Ganne et al. and the release of metals at lower pH, adsorption of hydrolyzed
(2008) addressed the potential for geotextiles to retain volatile species in addition to cation exchange are hypothesized to be the
organic compounds (VOCs) from GCLs. The sorption isotherms are mechanisms that contribute the most to metal retention.
linear for VOCs and the sorption coefficients range from 7 to 20 mL/
g for 1,2-dichloroethane, 79e102 mL/g for trichloroethylene,
5.4. Diffusive transfer through geosynthetic clay liners
20e41 mL/g for benzene, 87e135 mL/g for toluene, 229e248 mL/g
for ethylbenzene, 263e298 mL/g for m&p-xylene, and
5.4.1. Diffusion of inorganic species through geosynthetic clay liners
163e192 mL/g for o-xylene. These values were obtained at 22  C.
Lake and Rowe (2000) showed that, for void ratios ranging from
Rowe et al. (2005) showed that diffusion and sorption depend on
1.5 to 3, the diffusion coefficient of sodium and chloride increases
temperature and that both parameters are lower at 7  C than at
linearly with the bulk-GCL void ratio. The bulk-GCL void ratio was
22  C for benzene, toluene, ethylbenzene, m&p-xylene, and o-
defined by Petrov et al. (1997) as:
xylene.
Ahari et al. (2011) studied the sorption of 13 phenolic com- HGCL  Hs
pounds: phenol, o-cresol, p-cresol, 2-chlorophenol (2-CP), 4- eb ¼ (8)
Hs
chlorophenol (4-CP), 2,4-dimethylphenol (2,4-DMP), 3,4-
dimethylphenol (3,4-DMP), 2,4-dichlorophenol (2,4-DCP), 2,4,6- Where HGCL is the GCL height; and Hs is the height of solids in the
trichlorophenol (2,4,6-TCP), 2,3,5,6-tetrachlorophenol (2,3,5,6- GCL. The height Hs is defined by
TeCP), 2,3,4,6-tetrachlorophenol (2,3,4,6-TeCP), pentachloro-
phenol (PCP), and bisphenol A (BPA). They noticed that the Mbent Mgeo
Hs ¼ þ (9)
adsorption isotherms are nonlinear, contrary to what is observed rs ð1 þ w0 Þ rsg
for VOCs (Rowe et al., 2005; Ganne et al., 2008). Therefore, Ahari
et al. (2011) used the Freundlich model to study adsorption. How- where Mbent is the mass of bentonite per unit area in the GCL, Mgeo
ever, for the sake of comparing with previous results obtained from is the mass of geosynthetics per unit area in the GCL, rs is the
VOCs, they also calculated Kd under the hypothesis of linear sorp- density of bentonite solids, rsg is the density of polypropylene
tion (Ahari et al., 2011). For the compounds studied, all results for geotextile solids; and w0 is the initial water content of the
Kd range from 2.7 to 8.9 mL/g. These values for the sorption coef- bentonite.
ficient remain small compared with those obtained for VOCs. An The diffusion coefficients of sodium and chloride inferred from
important observation is that the amount of chlorophenols sorbed GCL diffusion measurements done with 3e5 g/L solutions decrease
by geotextiles increases as the number of chlorine atoms in the linearly with decreasing final bulk-GCL void ratio. The diffusion
molecule increases. coefficient was shown to depend on the source solution and, upon
significantly increasing the NaCl concentration, the diffusion coef-
5.3.1.2. Adsorption onto bentonite. Lake and Rowe (2004) also ficient inferred also increased. The diffusion coefficients were
studied the potential of bentonites to retain VOCs from GCLs. Very estimated to range from 1  1010 to 2  1010 m2/s.
small partition coefficients are reported in the literature (less than Lange et al. (2009) studied the diffusion of various metals for the
1 mL/g). following four cases where a GCL might serve as an effective barrier
Recently, several studies addressed the potential for sorption of against metals and metalloids: acidic rock drainage, gold-mine
phenolic compounds onto bentonites or organobentonites (Banat tailings, lime-treated mine effluent, and municipal solid waste.
et al., 2000; Yoo et al., 2004; Hameed, 2007; Richards and The averaged diffusion coefficients for Cu, Cd, Zn, Fe, and Ni covered
Bouazza, 2007; Malusis et al., 2010; Ahari et al., 2011). As stated a narrow range from 6.7  1011 to 8.9  1011 m2/s. The diffusion
above, Ahari et al. (2011) obtained nonlinear adsorption curves for coefficients for As, Al, Mg, Mn, and Sr range from 8.0  1011 to
bentonite, so they used the Freundlich model to model bentonite. 1.6  1010 m2/s. The diffusion coefficients of the individual metals
The results obtained are consistent with previous results from Banat did not change significantly upon changing the composition of the
et al. (2000) for phenol. To compare with other published data on solution, which suggests that, although the composition of the
linear adsorption isotherms, they also determined the adsorption solution has some effect on the diffusion coefficient of the metal,
N. Touze-Foltz et al. / Geotextiles and Geomembranes 44 (2016) 656e672 667

sorption onto the GCL is the dominant factor controlling the metal publication, had run for 12 years. The receptor concentration in this
mobility. measurement remained below about 0.02% of the source concen-
tration, lying within the range of analytical uncertainty for the
5.4.2. Diffusion of organic species through geosynthetic clay liners chemical analysis. Rowe (2005) also cites a study by August and
The diffusion mechanisms of VOCs were also quantified for Tatsky (1984) that concludes that negligible diffusion of heavy-
virgin GCLs (Lake and Rowe, 2004; Rowe et al., 2005; Ganne et al., metal salts from a 0.5 M acid solution occurs through a HDPE GM
2008) containing sodium bentonite. Lake and Rowe (2004) over a four year measurement period. Based on these results, Rowe
observed no significant increase in the diffusion coefficient for (2012) concluded that an intact GM is an excellent barrier against
bulk-GCL void ratios ranging from 4.1 to 4.8. Their results indicate advective and diffusive migration of inorganic contaminants from a
that bulk-GCL void ratios ranging from 4.1 to 4.8 correspond to low leachate.
normal stress and that the corresponding diffusion coefficients
represent an upper bound for diffusion coefficients. In further 5.5.2. Diffusion of organic species
studies, Rosin-Paumier et al. (2011) reported that the composition August and Tatsky (1984) found that the permeation rates in a
of the bentonite (natural sodium versus calcium-activated) has no HDPE GM of strongly polar penetrant molecules range from
significant impact on the diffusion of VOCs. At 23  C, the diffusion 7  107 m3 m2 d1 for methanol to 9.4  106 m3 m2 d1 for
coefficients for VOCs in bentonite range from 1 to 3  1010 m2/s for trichloroethylene.
bulk-GCL void ratios ranging from 3.7 to 4.8. A number of studies on the use of GMs have focused on the
Mendes et al. (2013) quantified the diffusion of seven phenolic diffusion of VOCs through virgin HDPE GMs (Park and Nibras, 1993;
compounds through GCLs and reported diffusion coefficients for Prasad et al., 1994; Müller et al., 1998; Sangam and Rowe, 2001;
virgin GCLs that range from 5  1011 to 1.3  1010 m2/s for meth- Touze-Foltz et al., 2011; Park et al., 2012) virgin PVC, linear low-
ylphenols and from 5  1011 to 6.3  1011 m2/s for chlorophenols. density polyethylene (LLDPE) GMs with and without a co-extruded
These values fall in the lower range of diffusion coefficients reported ethylene vinyl-alcohol (EVOH) inner core (McWatters and Rowe,
in the literature for the diffusion through virgin GCLs. 2008, 2010; Eun et al., 2014), fluorinated HDPE GMs (Sangam and
Rowe, 2005), and aged HDPE GMs (Rowe et al., 2003; Islam and
5.4.3. Aged GCLs after cation exchange Rowe, 2009). The diffusion coefficients of VOCs in virgin HDPE GMs
Rosin-Paumier et al. (2011) subjected a GCL containing natural are 0.37  1013 and 22.8  1013 m2/s for benzene and DCM,
sodium bentonite to cation exchange by permeating the GCL with a respectively, with partition coefficients ranging from 1.8 to 189. The
synthetic leachate containing a mixture of monovalent and divalent resulting permeation coefficient lies between 1 and 70  1012 m2/s.
cations in proportions representative of what is found in actual Sangam and Rowe (2005) examined how surface fluorination of
landfill leachate. They reported that cation exchange leads to an a HDPE GM affects the diffusion of VOCs. In this study, surface
increase in the hydraulic conductivity of the GCL by a factor 8.5. fluorination consisted of applying elemental fluorine, which ex-
They then studied diffusion in three specimens cut from the sample changes with hydrogen along polymer chains at the surface of a
and subjected to cation exchange and reported an increase in the polyolefin substrate. The partition coefficient remains essentially
diffusion coefficient of VOCs compared with that of virgin GCL the same after surface fluorination; however, surface fluorination
specimens for a bulk-GCL void ratio of 3.9. Trichloroethylene ex- reduces both the diffusion and the permeation coefficients by
hibits the largest increase in diffusion coefficient of the VOCs factors between 1.5 and 4.5, depending on the hydrocarbon
studied: from 1.0  1010 to 2.6  1010. The diffusion coefficients examined.
for dichloromethane (DCM) and dichloroethane (DCA) increase by a McWaters and Rowe (2010) studied diffusive migration of the
factor about 1.4 (from 2.3  1010 to 3.1  1010 m2/s and from aqueous and vapor phases of benzene, toluene, ethylbenzene, and
1.9  1010 to 2.6  1010 m2/s, respectively). This ratio is not as xylenes through a 0.76-mm-thick PVC-P GM and a 0.76-mm-thick
large as the analogous ratio for the increase in the hydraulic con- LLDPE GM. For the PVC GM, diffusion coefficients ranged from 5 to
ductivity of the GCL. In addition, a decrease in the bulk-GCL void 10  1013 m2/s for diffusion from both the aqueous and vapor
ratio from 3.9 to 3 would negate the effect of cation exchange so states. For the LLDPE GM, diffusion coefficients ranged from 2.5 to
that the diffusion coefficient would not increase for DCM and DCA. 5  1013 m2/s. The partition coefficients for PVC ranged from 100
Such a decrease in the bulk-GCL void ratio could occur over the to 1075 with respect to aqueous-phase concentrations. The parti-
lifetime of the landfill as the height of the waste increases. Thus, a tion coefficients for vapor-phase concentrations ranged from 22 to
detrimental effect of cation exchange on the hydraulic conductivity 290. For the LLDPE GM, partition coefficients ranged from 200 to
of a GCL is not necessarily indicative of a detrimental effect on the 475 for aqueous-phase concentrations and from 44 to 123 for
diffusion coefficient of VOCs through GCLs. This result remains to be vapor-phase concentrations. The resulting permeation coefficients
confirmed for other GCLs and other pollutants. thus range from 130 to 750  1012 m2/s for PVC GMs and from 60
For phenolic compounds in aged GCL specimens, diffusion co- to 110  1012 m2/s for LLDPE GMs.
efficients range from 2  1010 to 6.4  1010 m2/s for methyl- McWaters and Rowe (2010) also studied the diffusive proper-
phenols and from 8.1  1011 to 2.5  1010 m2/s for chlorophenols. ties of two coextruded GMs, one with a polyamide inner core and
Due to the high sorption of methylphenols observed in the control the other with an EVOH inner core, and a standard 0.53-mm-thick
cell, only the order of magnitude of these diffusion coefficients is LLDPE GM. The results indicate a significant reduction in mass flux
considered significant (Mendes et al., 2013). Mendes et al. (2014a) through the coextruded GMs compared with the conventional
also quantified the diffusion of bisphenol A after cation exchange LLDPE GM. The EVOH coextruded GM has the lowest permeation
for the same GCL specimen. They obtained a diffusion coefficient of coefficients of about 8  1015 m2/s for diffusion from the aqueous
2.2  1010 m2/s. phase. These permeation coefficients for the EVOH coextruded GM
are upper bounds; actual values may be even lower. The polyamide
5.5. Diffusive transfer through geomembranes coextruded GM has higher permeation coefficients than the EVOH
coextruded GM, ranging from 5 to 8  1014 m2/s from the aqueous
5.5.1. Diffusion of inorganic species phase. The standard LLDPE GM has the highest diffusion co-
Rowe (2005) presented the results of a measurement of the efficients, which range from 2 to 4  1013 m2/s. The partition
diffusion of chloride through a geomembrane that, at the time of coefficients for the EVOH coextruded GM range from 160 to 700 for
668 N. Touze-Foltz et al. / Geotextiles and Geomembranes 44 (2016) 656e672

aqueous-phase concentrations. For the polyamide coextruded GM, of VOCs through virgin HDPE GMs (Sangam and Rowe, 2001; Joo
the partition coefficients range from 120 to 430 and, for the LLDPE et al., 2004; Park et al., 2012).
GM, the partition coefficients range from 180 to 450 for aqueous- The diffusion of polybriominated diphenyl ethers (PBDE) (Saheli
phase concentrations. In summary, the permeation coefficients et al., 2011) is currently under study.
range from 2 to 6  1012 m2/s for the EVOH coextruded GM, from Touze-Foltz et al. (2012b) studied the diffusion of phenol, o-
7 to 22  1012 m2/s for the polyamide coextruded GM, and from cresol, p-cresol, 2,4-xylenol, 3,4-xylenol, 2-chlorophenol, 4-
60 to 200  1012 m2/s for the LLDPE GM. chlorophenol, 2,4-dichlorophenol, 2,4,6-trichlorophenol, 2,3,4,6-
The EVOH coextruded GM thus offers a five-to twelve-fold tetrachlorophenol, 2,3,5,6-tetrachlorophenol, pentachlorophenol,
decrease in the permeation coefficient compared with a 2.0-mm- and BPA through a virgin HDPE GM. The results show that the
thick HDPE GM. partition coefficient is linked to the aqueous solubility and the n-
The question is of the prediction of the evolution of the diffusion octanol-water partition coefficient of the contaminant. This latest
coefficient, the partition coefficient, and the permeation coefficient result is logical because less polar contaminants are less soluble in
as it may not be possible in the future to keep on performing water. No clear link is found with the molecular diameter when all
measurements for the variety of contaminants that can be phenolic compounds are taken into account (especially BPA).
encountered in environmental applications. However, for chlorophenols a strong correlation appears between
Theory to predict the evolution of the diffusion coefficient, the molecular weight and partition coefficient, which may be related to
partition coefficient, and the permeation coefficient as functions of polarity. An analysis of the sole chlorophenols (2-CP, 4-CP, 2,4-DCP,
solubility, octanolewater partition coefficient, and molecular 2,3,6-TCP, 2,3,4,6-TeCP, 2,3,5,6-TeCP, and PCP) suggests that the
diameter of various VOCs is given in the literature for the diffusion partition coefficient is closely linked with the degree of substitution

6
1000000
106 1000000
10
Joo et al. (2004) Joo et al. (2004)
Park et al. (2012)
Park et al.(2012)
100000 5 Sangam and Rowe (2001)
105
100000 Phenols (GMB) 10
Phenols (GMB)
VOCs (GMB) VOCs (GMB)
Chlorophenols (v PE Film) 4
10000
10 Chlorophenols (v PE Film)
104
10000 Chlorophenols (a PE Film) Chlorophenols (a PE Film)

1000
10
3

103
1000
2
100
10
Sgf

Sgf

102
100
10

10
1

1 0.1

0.1 0.01
0 0.010.1
0.0010.01 0.1 11 10
10 100
102 1000
10310000
10100000
4 105 106 107
1000000
1000000 -1 0 1 2 3 4 5 6
S (mg/L)
S (mg/l) log Kow

(a) (b)
6
1000000
10 1000000
10
6

Park et al. (2012) Phenols (GMB)


5
Phenols (GMB) VOCs (GMB)
100000
10 1000005 10
VOCs (GMB) Chlorophenols (v PE Film)
Chlorophenols (a PE Film)
Chlorophenols (v PE Film)
4 4
10
10000 Chlorophenols (a PE Film) 10000
10

3
10
1000 103
1000
Sgf

Sgf

2 2
10
100 10
100

10 10

1 1

0.1 0.1
0.45 0.6 0.75 0.9 0 100 200 300
dm (nm) Molecular weight (g/mol)
(c) (d)

Fig. 9. Partition coefficient Sgf as a function of (a) aqueous solubility S, (b) n-octanol-water partition coefficient Kow, (c) molecular diameter dm, and (d) molecular weight for
phenolic compounds and VOCs in diffusion experiments with geomembranes, and for chlorophenols in diffusion experiments with virgin and aged PE films (“v PE Film” and “a PE
Film,” respectively).
N. Touze-Foltz et al. / Geotextiles and Geomembranes 44 (2016) 656e672 669

10 10

1 1

Dg (x 10-12m²/s)

Dg (x 10-12m²/s)
0.1 0.1

0.01 0.01

Park et al. (2012)


Park et al. (2012)
Sangam and Rowe (2001)
Phenols (GMB) Phenols (GMB)
0.001 0.001
VOCs (GMB) VOCs (GMB)
Chlorophenols (v PE Film) Chlorophenols (v PE Film)
Chlorophenols (a PE Film) Chlorophenols (a PE Film)
0.0001 0.0001
0.0010.01
0 0.010.10.1 1 1 1010100
101000
2 1010000
100000
3 10 1000000
4 10 10000000
5 10 6 107 -1 0 1 2 3 4 5 6
SS(mg/l)
(mg/L) log Kow

(a) (b)

10 10

1 1
Dg (x 10-12m²/s)

0.1
Dg (x 10-12m²/s)

0.1

0.01 0.01
Phenols (GMB)

Joo et al. (2004) VOCs (GMB)


Park et al. (2012) Helmroth et al. (2002)
0.001 Phenols (GMB) 0.001
Brandsch et al. (1999)
VOCs (GMB)
Chlorophenols (v PE Film) Chlorophenols (v PE Film)
Chlorophenols (a PE Film) Chlorophenols (a PE Film)
0.0001 0.0001
0.45 0.6 0.75 0.9 0 100 200 300
dm (nm) Molecular weight (g/mol)

(c) (d)

Fig. 10. Diffusion coefficient Dg as function of (a) aqueous solubility, (b) n-octanol-water partition coefficient Kow, (c) molecular diameter dm, and (d) molecular weight for phenolic
compounds and VOCs in diffusion experiments with geomembranes, and for chlorophenols in diffusion experiments with virgin and aged PE films (“v PE Film” and “a PE Film,” respectively).

of chlorine atoms on the phenolic nucleus, a phenomenon that may the parameters obtained for BPA). These trends require further
also be attributed to a difference in polarity of the various chlor- confirmation for other chemical families before attempting to
ophenols studied. predict the diffusion coefficient based only on the molecular
The partition coefficients of phenolic compounds are small diameter (for molecular diameters close to 0.5 nm).
compared with those of VOCs with similar chemical structure (e.g.,
benzene compared with phenol and toluene compared with o- 5.5.3. Effect of geomembrane ageing on diffusion
cresol and p-cresol), which suggests that the partition coefficient Islam and Rowe (2009) examined the effects of aging of HDPE GMs
decreases as the hydroxyl group solubilizes the molecule. on the diffusion and the partition of benzene, toluene, ethylbenzene,
The diffusion coefficient is well correlated with the aqueous and xylenes. For this study, two different 1.5-mm-thick HDPE GMs
solubility and the n-octanol-water partition coefficient of the were aged in the laboratory at 85  C by immersion in a synthetic
phenolic compounds (see Fig. 9). A good correlation with the mo- leachate for up to 32 months. Partition and diffusion measurements
lecular diameter is also obtained when one disregards the results were made at room temperature on both unaged and aged GMs by
obtained for BPA. using a dilute aqueous solution. The diffusion and partition co-
These trends are consistent with those previously obtained for efficients decrease as aging progresses. After aging the GM for 10e32
VOCs. However, the range of obtained for the parameters differs months, the inferred permeation coefficients decreased by 36%e62%.
significantly from that obtained for the VOCs, which means that This decrease in the diffusion, partition, and permeation coefficients
the empirical equations available in the literature that were is related to the increase during aging in GM crystallinity.
derived for VOCs are not valid for predicting the evolution of the These results for aging are consistent with previous results from
permeation coefficient (see Fig. 10) of phenolic compounds, except Rowe et al. (2003), who showed that the permeation coefficient for a
for predicting its evolution with molecular diameter (excluding 14-year-old HDPE GM sampled from a leachate lagoon is between
670 N. Touze-Foltz et al. / Geotextiles and Geomembranes 44 (2016) 656e672

four and five times less than that obtained for unaged HDPE GMs permeation coefficient correlate well with the aqueous solubility and
typical of what is produced today (i.e., no virgin specimen was the n-octanol-water partition coefficient of the chlorophenols, and
available). Measurements were recently performed to elucidate the they also correlate well with the molecular diameter.
effect of GM aging on the diffusion of phenolic compounds through These trends are consistent with previous trends obtained for
GMs. The diffusion of 2,4,6-trichlorophenol, 2,3,5,6- the diffusion of VOCs and phenolic compounds through GMs.
tetrachlorophenol, and pentachlorophenol through two 17-year-old However, the ranges obtained for the parameters differ signifi-
high-density polyethylene geomembranes (HDPE GMs) was studied cantly, so the empirical equations published for VOCs cannot be
(Touze-Foltz et al., 2016). The first was stored indoors away from used to predict the evolution of the partition coefficient of the
direct sunlight whereas the second was exposed for the entire time in chlorophenols studied herein or of the phenolic compounds stud-
a south-facing corner of a pond. Both unexposed and exposed spec- ied by Touze-Foltz et al. (2012b). However, equations previously
imens, although under different exposure conditions, experienced used in the field of PE films for food packaging and that link the
ageing. Partition coefficients ranged from 30 to 190, increasing with diffusion coefficient to the molecular mass (Brandsch et al., 1999;
the number of chlorine atoms in the contaminant. Diffusion co- Helmroth et al., 2002) prove to apply to the data discussed herein
efficients for the unexposed geomembrane ranged from 9  1014 to for GMs (see Fig. 10).
2.4  1013 m2/s and from 9  1014 to 2.8  1013 m2/s for the
exposed geomembrane. These values were compared with those 6. Conclusion
reported in the literature for diffusion of the same phenolic com-
pounds through a virgin HDPE geomembrane from a different The first objective of this review paper is to present the state-of-
manufacturer: the exposure did not noticeably affect the partition the-art measurement techniques for evaluating properties of
coefficients or diffusion coefficients. This result was linked to the transfer of chemicals in lining materials for landfills. The second
temporal evolution of the unexposed geomembrane. Furthermore, in objective is to summarize the current knowledge regarding
spite of their 17-year ageing, both geomembrane specimens still had advective and diffusive transfer parameters for GMs and GCLs,
diffusive characteristics that were consistent with a virgin geo- included multicomponent GCLs.
membrane obtained from a different manufacturer. The retention curves of GCLs are discussed first because they are
Mendes et al. (2014b) made preliminary measurements of the crucial for predicting transfer through barriers, including GCLs that
diffusion of chlorophenols (4-chlorophenol, 2,4-dichlorophenol, contain bentonite with a low water content and that may be
2,4,6-tricholophenol, 2,3,5,6-tetrachlorophenol, and pentachloro- installed in environmental applications and therefore subject to
phenol) through PE films produced from the same base resin as a desiccation. Thus, knowledge of the hydration path is important to
HDPE GM that is commercially available in Europe. The molar mass in adequately predict advective transfer in GCLs or composite liners
number (Mn) and the molecular weight (Mw) of the resin, as deter- containing GCLs. Recently obtained results provide evidence that
mined by “ultrahigh-temperature” gas-phase chromatography, are load has a significant effect on the retention curves of GCLs.
Mn ¼ 44.1 kg/mol and Mw ¼ 136.7 kg/mol. The PE density is 0.911 kg/ Section 3 discusses the quantification of advective flow rates in
m3 and the crystallinity fraction is 51%. The films were fabricated GMs and multicomponent GCLs. Although no dedicated standard
from pellets. Some of the virgin PE film specimens were aged by exists for GCLs, a procedure was recently developed to quantify
thermo-oxidation at 105  C, which leads to the formation of carbonyl advective flow in these materials. The existing methodologies allow
compounds (alcohols, carboxylic acids, etc.) and to an expected in- us to study the potential evolution of the flow rate in GMs during
crease in crystallinity. The final carbonyl concentrations in the their lifetime at a site and to detect a decrease in performance,
oxidized samples were between 0.1 and 0.4 mol/kg, which was ob- which occurs, for example, to oxidized bituminous GMs when left
tained after 300 h of thermo-oxidation (Pons, 2012). For such uncovered. For multicomponent GCLs, the procedure developed,
carbonyl concentrations, the crystallinity fraction ranged from 54% to which will become a French standard, allows us to differentiate
66% according to the calibration curve proposed by Pons (2012) for between the hydraulic performance of the various multicomponent
the same polymer as used in this study. These results for crystallinity GCLs currently on the market. The evidence shows that, when the
fraction should be compared with 51% crystallinity fraction for virgin film or coating is sufficiently thick, it is this component rather than
PE. For PE films, the partition coefficients increase with increasing the bentonite in the GCL that maintains the lining. This brings into
number of chlorine atoms for chlorophenols, whereas the diffusion question the definition of multicomponent GCLs that cannot be
coefficients generally decrease with increasing number of chlorine categorized as caly geosynthetic barriers.
atoms. In addition, the diffusion coefficients decrease with the in- Section 4 discusses the advective flow rate due to holes either in
crease in crystallinity that occurs upon aging. The permeation co- GMs used in association with GCLs or in the film or coating of
efficients, however, increase; contrary to what was reported by Rowe multicomponent GCLs. The thickness of the film or coating was
et al. (2003) when they compared the diffusion of chlorinated and found to have no detectable influence on the meter scale. In addi-
aromatic hydrocarbons through a new modern HDPE GM with that tion, results obtained in the steady state are not significantly
through a 14-year-old HDPE GM. These results show that the parti- impacted by the composition of the GM, its texture, the composi-
tion coefficient is linked to the aqueous solubility and the n-octanol- tion of the bentonite, or bentonite ageing as regards cation ex-
water partition coefficient of the contaminant. In view of the results change and cation exchange combined with hydrationedesiccation
reported by Touze-Foltz et al. (2012a,b), this latest result seems cycles. Based on the data presented herein, the question of GCL
logical because less polar contaminants are less soluble in water. ageing does not appear relevant if the GCL is used in association
Although no clear evidence supports a link with the molecular with a GM in a composite liner.
diameter, for chlorophenols the molecular weight correlates well Finally, the diffusion of organic and inorganic species through
with the partition coefficient; a result that may be related to polarity. GCLs and GM is discussed. Until recently very few species had been
An analysis of the chlorophenols under study (4-CP, 2,4-DCP, investigated. Data now exist on various types of geomembranes,
2,4,6-TCP, 2,3,5,6-TeCP, and PCP) implies that the partition coefficient especially as regards the diffusion of VOCs. A recent study has
is closely linked with the degree of substitution of chlorine atoms on shown that the empirical laws used for predicting the diffusion of
the phenolic nucleusda phenomenon that may also be attributed to VOCs in GMs cannot be extended to other chemical families, such
a difference in polarity between the various chlorophenols studied, phenolic compounds. This situation emphasizes the need for
as noticed by Touze-Foltz et al. (2012b). The diffusion coefficient and further research in this field.
N. Touze-Foltz et al. / Geotextiles and Geomembranes 44 (2016) 656e672 671

References Bouazza, A., Jefferis, S., Vangpaisal, T., 2007. Investigation of the effects and degree
of calcium exchange on the Atterberg limits and swelling of geosynthetic clay
liners when subjected to wet-dry cycles. Geotext. Geomembr. 25 (3), 170e185.
Abuel-Naga, H.M., Bouazza, A., 2010. A novel laboratory technique to determine the
Boutron, O., Gouy, V., Touze-Foltz, N., Benoit, P., Chovelon, J.-M., Margoum, C., 2009.
water retention curve of geosynthetic clay liners. Geosynth. Int. 17 (5), 313e322.
omembranes e dictionnaire des termes relatifs aux Geotextile fibres retention properties to prevent surface water nonpoint
AFNOR, 2013. NF P84-500. Ge
contamination by pesticides in agricultural areas. Geotext. Geomembr. 27,
geomembranes.
254e261.
AFNOR, 2008. NF P84-705. Geosynthetic Barriers e Determination of the Swelling, Flow
Brandsch, J., Mercea, P., Piringer, O., 1999. Modeling of additive diffusion coefficients
and Permeability Characteristics of Geosynthetic Clay Liners (GCL) Using an
in polyolefins. In: Risch, S.J. (Ed.), Food Packaging Testing Methods and Appli-
Oedopermeameter d Characterisation Test and Performance Test (English version).
cations, ACS Symposium Series, 753, Washington DC, pp. 27e36.
AFNOR, 2006a. EN ISO 10318. Geosynthetics d Terms and Definitions.
Brown, K.W., Thomas, J.C., Lytton, R.L., Jayawickrama, P., Bhart, S., 1987. Quantifi-
AFNOR, 2006b. EN 14150. Geosynthetic Barriers d Determination of Permeability
cation of Leakage Rates through Holes in Landfill Liners. United States Envi-
to Liquids.
ronmental Protection Agency Report CR810940, Cincinnati, OH, p. 147.
Ahari, M., Touze-Foltz, N., Mazeas, L., Guenne, A., 2011. Quantification of the
Cleary, B.A., Lake, C.B., 2011. Comparing measured hydraulic conductivities of a
adsorption of phenolic compounds on the geotextile and bentonite components
geotextile polymer coated GCL utilizing three different permeameter types. In:
of four geosynthetic clay liners. Geosynth. Int. 18 (5), 322e331.
Han, J., Alzamora, D.A. (Eds.), Geo-frontiers 2011 Advances in Geotechnical
Ahel, M., Tepic, N., 2000. Distribution of polycyclic aromatic hydrocarbons in a
Engineering, pp. 1991e2000.
municipal solid waste landfill and underlying soil. Bull. Environ. Contam. Tox-
Daniel, D.E., Shan, H.eY., Anderson, J.D., 1993. Effects of partial wetting on the
icol. 65 (2), 236e243.
performance of the bentonite component of a geosynthetic clay liner. In: Pro-
Ali, M.A., Singh, R.M., Bouazza, A., Gates, W.P., Rowe, R.K., 2014. Effect of vertical
ceedings of Geosynthetics'93, Vol. 3. IFAI, St. Paul, USA, pp. 1482e1496.
stress on GCL thermal conductivity. In: Proceedings 7 International Congress on
Durin, L., Touze, N., Duquennoi, C., 1998. Water and organic solvents transport
Environmental Geotechnics, 10e14 November, Melbourne, Australia, 8p.
parameters in geomembranes. In: Proceedings 4th International Conference on
ASTM 2010. Standard D5084. Standard Test Methods for Measurement of Hydraulic
Geosynthetics, Atlanta, Georgia, USA, pp. 249e256.
Conductivity Of Saturated Porous Materials using a Flexible Wall Permeameter,
Egloffstein, T.A., 2002. Bentonite as sealing material in geosynthetic clay liners-
ASTM International, West Conshohocken, PA, 2010, DOI: 10.1520/D5084-10,
influence of the electrolytic concentration, the ion exchange and ion ex-
www.astm.org.
change with simultaneous partial desiccation on permeability. Clay Geosynth.
August, H., Tatsky, R., 1984. Permeabilities of commercially available polymeric
Barriers 141e151.
liners for hazardous landfill leachate organic constituents. In: Proceedings In-
Egloffstein, T.A., 2001. Natural bentonites-influence on the ion exchange and partial
ternational Conference on Geomembranes, Denver, USA, pp. 163e168.
desiccation on permeability and self-healing capacity of bentonites used in
Azad, F.M., 2011. Investigation of the Behaviour of Clay Liners at the Base of
GCLs. Geotext. Geomembr. 19, 427e444.
Municipal Solid Waste Landfills. PhD Thesis. University of Sydney, NSW,
Eloy-Giorni, C., 1993. Etude des transferts diffusifs dans les geomembranes :
Australia, p. 203.
me canismes et mesures a  l'aide de traceurs radioactifs. The
se de l'universite
 J.
Banat, F., Al-Bashir, B., Al-Asheh, S., Hayajneh, O., 2000. Adsorption of phenol by
Fourier, Grenoble 1, CEA, p. 174 (in French).
bentonite. Environ. Pollut. 107 (3), 391e398.
Eun, J., Tinjum, J.M., Benson, C.H., Tuncer, B.E., 2014. Volatile organic compound
Bannour, H., Stoltz, G., Delage, P., Touze-Foltz, N., 2014. Effect of stress on water
(VOC) transport through a composite liner with co-extruded geomembrane
retention of geosynthetic clay liners. Geotext. Geomembr. 42, 629e640.
containing Ethylene vinyl-alcohol (EVOH). In: Proceedings Geo-congress 2014,
Bannour, H., Barral, C., Touze-Foltz, N., 2015. Altered geosynthetic clay liners:
ASCE, 23e26 February 2014, Atlanta, GA, USA, 1960e1969.
effect on the hydraulic performance of composite liners. Eur. J. Environ. Civ. Eng.
Fredlund, D.G., Xing, A., 1994. Equations for the soilewater characteristic curve. Can.
19 (9), 1155e1176. http://dx.doi.org/10.1080/19648189.2015.1005161.
Geotech. J. 31, 533e546.
Bannour, H., Touze-Foltz, N., 2015. Flow-rate measurements in meter-size multi-
Ganne, A., Touze-Foltz, N., Mazeas, L., Guenne, A., Epissard, J., 2008. Experimental
component geosynthetic clay liners. Geosynth. Int. 22 (1), 70e77.
determination of sorption and diffusion of organic pollutants through GCLs. In:
Bannour, H., Barral, C., Touze-Foltz, N., 2013a. Flow rate in composite liners including
Proceedings Eurogeo 4, p. 7.
GCLs and bituminous geomembranes. In: International Conference on Geotechnical
GeoStudio, 2007. Version 7. Geo-slope International Ltd, Calgary, Alberta, Canada,
Engineering 2013, ICGE'13, Hammamet, Tunisia, February 21e23, pp. 809e819.
, A., von Maubeuge, K.P., 2013b. Interface p. 17.
Bannour, H., Touze-Foltz, N., Courte
Haijian, X., Yunmin, C., Han, K., Xiaowu, T., Renpeng, C., 2009. Analysis of diffusion-
transmissivity measurement in multicomponent geosynthetic clay liners. In:
adsorption equivalency of landfill liner systems for organic contaminants.
von Maubeuge, Kent P., Kline, J.P. (Eds.), Current and Future Practices for the
J. Environ. Sci. 21 (4), 552e560.
Testing of Multi-component Geosynthetic Clay Liners, ASTM STP 1562,
Hameed, B., 2007. Equilibrium and kinetics studies of 2,4,6-trichlorophenol
pp. 47e61.
adsorption onto activated clay. Colloids Surf. A Physicochem. Eng. Asp. 307
Barral, C., Touze-Foltz, N., Croissant, D., 2014. Flow rate quantification in multi-
(1e3), 45e52.
component geosynthetic clay liners with the oedopermeameter method. In:
Hanson, J.L., Risken, J.L., Yesiller, N., 2013. Moisture-Suction relationship for geo-
Proceedings 10 ICG, September 23e26, Berlin, p. 8.
synthetic clay liners. In: Proceedings of the 18th Conference on Soil Mechanics
Barral, C., Touze-Foltz, N., 2012. Flow rate measurement in undamaged multicom-
and Geotechnical Engineering, Paris, France, pp. 3025e3028.
ponent geosynthetic clay liners. Geosynth. Int. 19 (6), 491e496.
Harpur, W.A., Wilson-Fahmy, R.F., Koerner, R.M., 1993. Evaluation of the contact
Barral, C., Touze-Foltz, N., Loheas, E., Croissant, D., 2012. Comparative study of hy-
between geosynthetic clay liners and geomembranes in terms of transmissivity.
draulic behaviour of geosynthetic clay liners exhumed from a landfill cover and
In: Proceedings of the 7th GRI Conference on Geosynthetic Liners Systems:
from a dam after several years in service. In: Proceedings of the 5th European
Innovations, Concerns and Design. IFAI, St. Paul, MN, pp. 138e149.
Geosynthetics Congress, Valencia, p. 8.
Helmroth, E., Rijk, R., Dekker, M., Jogen, W., 2002. Predictive modelling of migration
Barroso, M., 2005. Fluid Migration through Geomembrane Seams and through the
from packaging materials into food products for regulatory purposes. Trends
Interface between Geomembrane and Geosynthetic Clay Liner. Ph.D. Thesis.
food Sci. Technol. 13, 102e109.
University Joseph Fourier of Grenoble (France) and University of Coimbra
Hiroshi, A., Toshihiko, M., Nobutoshi, T., 2002. Endocrine disrupters in leachate from
(Portugal), p. 215.
two MSW landfills of different waste compositions in Japan. In: Proceedings of
Barroso, M.C.P., Lopes, M.D.G.A., Bergamini, G., 2010. Effect of the waste pressure on
the 2nd Asian Pacific Landfill Symposium, Seoul 2002, pp. 728e736.
fluid migration through geomembrane defects. In: Proceedings 9 ICG, Guaruja,
Islam, M.Z., Rowe, R.K., 2009. Permeation of BTEX through unaged and aged HDPE
Brazil, pp. 959e962.
geomembranes. J. Geotech. Geoenvironmental Eng. 135 (8), 1130e1140.
Barroso, M., Touze-Foltz, N., von Maubeuge, K., 2008. Influence of the textured
Joo, J.C., Kim, J.Y., Nam, K., 2004. Mass transfer of organic compounds in dilute
structure of geomembranes on the flow rate through geomembrane-GCL
aqueous solutions into high density polyethylene geomembranes. J. Environ.
composite liners. In: Proceedings Eurogeo4, Edinburgh, Scotland, p. 8.
Eng. 130 (2), 175e183.
Barroso, M., Touze-Foltz, N., Saidi, F., 2006a. Validation of the use of filter paper
Kjeldsen, P., Barlaz, M.A., Rooker, A.P., Baun, A., Ledin, A., Christensen, T.H., 2002.
suction measurements for the determination of GCLs water retention curves.
Present and long-term composition of MSW landfill leachate: a review. Crit.
In: Proceedings 8ICG, Geosynthetics, Yokohama, Japan, pp. 171e174.
Rev. Environ. Sci. Technol. 32 (4), 297e336.
Barroso, M., Touze-Foltz, N., von Maubeuge, K., Pierson, P., 2006b. Laboratory
Lake, C.B., Rowe, R.K., 2004. Volatile organic compound diffusion and sorption co-
investigation of flow rate through composite liners consisting of a geo-
efficients for a needle-punched GCL. Geosynth. Int. 11 (4), 257e272.
membrane, a GCL and a soil liner. Geotext. Geomembr. 24 (3), 139e155.
Lake, C.B., Rowe, R.K., 2000. Diffusion of sodium and chloride through geosynthetic
Beddoe, R.A., Take, W.A., Rowe, R.K., 2011. Water-retention behavior of geosynthetic
clay liners. Geotext. Geomembr. 18, 103e131.
clay liners. J. Geotech. Geoenvironmental Eng. 1028e1038. November 2011.
Lambert, S., Touze-Foltz, N., 2000. A test for measuring permeability of geo-
Beddoe, R.A., Take, W.A., Rowe, R.K., 2010. Development of suction measurement
membranes. In: Proceedings Eurogeo 2, Bologna, Italy, p. 10.
techniques to quantify the water retention behaviour of GCLs. Geosynth. Int. 17
Lange, K., Rowe, R.K., Jamieson, H., 2009. Diffusion of metals in geosynthetic clay
(5), 301e312.
liners. Geosynth. Int. 16 (1), 11e27.
Benson, C.H., 2013. Impact of subgrade water content on cation exchange and hy-
Lange, K., Rowe, R.K., Jamieson, H., 2004. Metal migration in geosynthetic clay
draulic conductivity of geosynthetic clay liners in composite barriers. Coupled
liners. In: 57th Canadian Geotechnical Conference, Session 3D, pp. 15e22.
Phenom. Environ. Geotech. 79e84.
Lin, L.-C., Benson, C.H., 2000. Effect of wet-dry cycling on swelling and hydraulic
Benson, C.H., Thorstad, P.A., Jo, H.Y., Rock, S.A., 2007. Hydraulic performance of
conductivity of GCLs. J. Geotech. Geoenvironmental Eng. 126, 40e49.
geosynthetic clay liners in a landfill final cover. J. Geotech. Geoenvironmental
Lucas, S.N., 2002. Manufacturing of and the performance of an integrally-formed,
Eng. 133 (7), 814e827.
polypropylene coated geosynthetic clay barrier. In: Zanzinger, H.,
672 N. Touze-Foltz et al. / Geotextiles and Geomembranes 44 (2016) 656e672

Koerner, R.M., Gartung, E. (Eds.), Proceedings of the International Symposium Rowe, R.K., Lake, C.B., Mukunoki, T., 2005. BTEX diffusion and sorption for a geo-
Clay Geosynthetic Barriers, Nuremberg, Germany, pp. 227e232. synthetic clay liner at two temperatures. J. Geotech. Geoenvironmental Eng. 131
McWaters, R.S., Rowe, R.K., 2010. Diffusive transport of VOCs through LLDPE and (10), 1211e1221.
two coextruded geomembranes. J. Geotech. Geoenvironmental Eng. 136 (9), Rowe, R.K., Quigley, R.M., Brachman, R.W.I., Booker, J.R., 2004. Barrier Systems for
1167e1177. Waste Disposal Facilities, second ed. E&F Spon, p. 587.
McWatters, R.S., Rowe, R.K., 2008. Transport of volatile organic compounds through Rowe, R.K., Sangam, H.P., Lake, C.B., 2003. Evaluation of an HDPE geomembrane
PVC and LLDPE geomembranes from both aqueous and vapour phases. Geo- after 14 years as a leachate lagoon liner. Can. Geotech. J. 40, 536e550.
synth. Int. 16 (6), 468e481. Saheli, P.T., Rowe, R.K., Rutter, A., Brachman, R.W.I., 2011. Sorption and diffusive
Malusis, M.A., Maneval, J.E., Barben, E.J., Shackelford, C.D., Daniels, E.R., 2010. In- transport of PBDE through an HDPE geomembrane. In: Han, J., Alzamora, D.A.
fluence of adsorption on phenol transport through soil-bentonite vertical bar- (Eds.), Proceedings Geo-frontiers 2011 Advances in Geotechnical Engineering,
riers amended with activated carbon. J. Contam. Hydrol. 116 (1e4), 58e72. pp. 1141e1151.
Meer, S.R., Benson, C.H., 2007. Hydraulic conductivity of geosynthetic clay liners Sangam, H.P., 2001. Performance of HDPE Geomembrane Liners in Landfills Appli-
exhumed from landfill final covers. J. Geotech. Geoenvironmental Eng. 133 (5), cations. PhD Thesis. University of Western Ontario, London, Ontario, p. 426.
550e563. Sangam, H.P., Rowe, R.K., 2005. Effect of surface fluorination on diffusion through a
Melchior, S., 2002. Field studies and excavation of geosynthetic clay barriers in high density polyethylene geomembrane. J. Geotech. Geoenvironmental Eng.
landfill covers. Clay Geosynth. Barriers 321e330. 131 (6), 694e704.
Mendes, M.J.A., Touze-Foltz, N., Gardoni, M., Mazeas, L., 2014a. Quantification of Sangam, H.;P., Rowe, R.K., 2001. Migration of dilute aqueous organic pollutants
diffusion of bisphenol A in a GCL after contact with a synthetic leachate. In: through HDPE geomembranes. Geotext. Geomembr. 19 (6), 329e357.
10th International Conference on Geosynthetics, Berlin, September 21e25, Siemens, G., Take, W.A., Rowe, R.K., Brachman, R., 2013. Effect of confining stress on
p. 8. the transient hydration of unsaturated GCLs. In: Proceedings of the 18th Con-
Mendes, M.J.A., Touze-Foltz, N., Gardoni, M., Mazeas, L., 2014b. Diffusion of phenolic ference on Soil Mechanics and Geotechnical Engineering. Paris, France,
compounds through polyethylene films. Geosynth. Int. 21 (2), 137e150. pp. 1187e1190.
Mendes, M.J.A., Touze-Foltz, N., Gardoni, M., Ahari, M., Mazeas, L., 2013. Quantifi- Singh, R.M., Bouazza, A., 2014. Modelling of GCL moisture and heat migration in a
cation of the diffusion of phenolic compounds in a virgin GCL and in a GCL after composite liner subjected to high temperatures. In: Proceedings 7 International
contact with a synthetic leachate. Geotext. Geomembr. 38, 16e25. Congress on Environmental Geotechnics, 10e14 November, Melbourne,
Mendes, M.J.A., Touze-Foltz, N., Palmeira, E.M., Pierson, P., 2010. Influence of Australia, p. 7.
structural and material properties of GCLs on interface flow in composite liners Sisinno, L.S., Oliveira-Filho, E.C., Dufrayer, M.C., Moreira, J.C., Paumgartten, J.R.,
due to geomembrane defects. Geosynth. Int. 17 (1), 34e47. 2000. Toxicity evaluation of a municipal dump leachate using zebrafish acute
Müller, W., Jakob, R., Tatsky-Gerth, August, H., 1998. Solubilities, diffusion and tests. Bull. Environ. Contam. Toxicol. 64 (1), 107e113.
partitioning coefficients of organic pollutants in HDPE geomembranes: exper- Southen, J.M., Rowe, R.K., 2007. Evaluation of the water retention curve for geo-
imental results and calculations. In: Proc. 6th International Conference on synthetic clay liners. Geotext. Geomembr. 25 (1), 2e9.
Geosynthetics, Atlanta, IFAI, Vol. 1, pp. 239e248. Southen, J.M., Rowe, R.K., 2005. Laboratory investigation of geosynthetic clay liner
Noval, A.M., Blanco, M., Farcas, F., Aguiar, E., Castillo, F., Touze-Foltz, N., 2015. Long- desiccation in a composite liner subjected to thermal gradients. J. Geotech.
term performance of EPDM geomembrane in El Boquero n reservoir. Geosynth. Geoenvironmental Eng. 131 (7), 925e935.
Int. 21 (6), 387e398. Takigami, H., Matsui, S., Matsuda, T., Shimizu, Y., 2002. The Bacillus subtilis rec-
Oman, C., Hynning, P.A., 1993. Identification of organic compounds in municipal assay: a powerful tool for the detection of genotoxic substances in the water
landfill leachates. Environ. Pollut. 80 (3), 265e271. environment. Prospect for assessing potential impact of pollutants from stabi-
Park, M., Benson, C.H., Edil, T.B., 2012. Comparison of batch and double compart- lized wastes. Waste Manag. 22 (2), 209e213.
ment test for measuring volatile organic compound transport parameters in Touze-Foltz, N., Mendes, M., Farcas, F., Maze as, L. 2016. Diffusion of phenolic
geomembranes. Geotext. Geomembr. 31, 15e30. compounds through two high-density polyethylene geoemembranes after 17
Park, J.K., Nibras, M., 1993. Mass flux of organic chemicals through polyethylene years under different exposure conditions. Geosynth. International, Ahead of
geomembranes. Water Environ. Res. 65, 227e237. Print, pp. 1e13, Published online:March 17, 2016. DOI: http://dx.doi.org/10.
Pelte, T., 1993. Etude the orique et expe rimentale de la fonction e tanche ite
 et du 1680/jgein.16.00004.
comportement thermique des ge omembranes. The se de doctorat de l'universite Touze-Foltz, N., Farcas, F., Benchet, R., 2015. Evaluation of the Ageing of two Bitu-
J. Fourier, Grenoble 1, p. 253 (in French). minous Geomembranes of Different Natures after 15 years in Service. Geo-
Petrov, R.J., Rowe, R.K., Quigley, R.M., 1997. Selected factors influencing GCLs hy- synthetics 2015, Feb 15e18, Portland, Oregon, USA, pp. 142e151.
draulic conductivity. J. Geotech. Geoenvironmental Eng. 123 (8), 683e695. Touze-Foltz, N., Bannour, H., Barral, C., Stoltz, G., 2014. Evaluating the performance of
Pinel-Raffaitin, P., Ponthieu, M., Amouroux, D., Mazeas, L., Donard, O.F.X., Potin- geosynthetics for environmental protection. In: Proceedings 7th International
Gautier, M., 2006. Evaluation of analytical strategies for the determination of Congress on Environmental Geotechnics, Melbourne, Australia, pp. 81e110.
metal concentrations to assess landfill leachate contamination. J. Environ. Touze-Foltz, N., Barral, C., Croissant, D., von Maubeuge, K., 2012a. Flow Rate Mea-
Monit. 8, 1069e1077. surement in Multi-component Geosynthetic Clay Liners. Current and Future
Pons, C., 2012. Durabilite  des geomembranes en polye thyle
ne haute densite  uti- Practices for the Testing of Multi-Component Geosynthetic Clay Liners. STP
es dans les installations de stockage de de
lise chets non dangereux. PhD. Thesis. 1562, 2013. http://dx.doi.org/10.1520/STP156220120088. Available online at:
Universite  Paris-Est, Paris, France, p. 200. www.astm.org.
Prasad, T.V., Brown, K.W., Thomas, J.C., 1994. Diffusion coefficients of organics in Touze-Foltz, N., Ahari, M., Mendes, M., Barral, C., Gardoni, M., Maze as, L., 2012b.
high density polyethylene (HDPE). Waste Manag. Res. 12, 61e71. Diffusion of phenolic compounds through an HDPE geomembrane. Geotech.
Richards, S., Bouazza, A., 2007. Phenol adsorption in organo-modified basaltic clay Eng. J. SEAGS AGSSEA 43 (4), 11.
and bentonite. Appl. Clay Sci. 37, 133e142. Touze-Foltz, N., Rosin-Paumier, S., Maze as, L., Guenne, A., 2011. Diffusion of volatile
Robinson, H.D., Knox, K., Van Santen, A., Tempany, P.R., 2001. Compliance of UK organic compounds through an HDPE geomembrane. In: Han, J., Alzamora, D.A.
landfills with EU pollution emissions legislation: development of a reporting (Eds.), Geo-frontiers 2011 Advances in Geotechnical Engineering,
protocol. In: Proceedings Sardinia 2001, Eighth International Waste Man- pp. 1121e1130.
agement and Landfill Symposium, pp. 21e30. Cagliari, Italy, 1e5 October Touze-Foltz, N., Croissant, D., Farcas, F., Royet, P., 2010a. Quantification of oxidized
2001. bituminous geomembranes ageing through hydraulic testing. In: Proceedings
Rosin-Paumier, S., Touze-Foltz, N., Maze as, L., Guenne, A., 2011. Quantification of 9th International Conference on Geosynthetics, Guaruja, Brazil, pp. 753e756.
volatile organic compounds diffusion for virgin geosynthetic clay liners and for Touze-Foltz, N., Croissant, D., Rosin-Paumier, S., Pirrion, T., Ouvry, J.F., 2010b. Perfor-
a GCL after contact with a synthetic leachate. J. Geotech. Geoenvironmental Eng. mance of a GCL in a landfill cover after six years in service. In: 3rd International
137 (11), 1039e1046. Symposium on Geosynthetic Clay Liners, Würzburg, Germany, pp. 83e91.
Rowe, R.K., 2012. Short- and long-term leakage through composite liners. The 7th Touze-Foltz, N., Zanzinger, H., 2009. Wasserdichtheit von Kunst-
Arthur Casagrande Lecture Can. Geotech. J. 49 (2), 141e169. stoffdichtungsbahnen e Erfahrungen mit der Prüfung. Geotech. Spec. Issue
Rowe, R.K., 2007. Advances and remaining challenges for geosynthetics in geo- 2009, 125e132.
environmental engineering applications, 23rd Manuel Rocha Lecture Soils van Genuchten, M.T.H., 1980. A closed-form equation for predicting the hydraulic
Rocks 30 (1), 3e30. conductivity of unsaturated soils. Soil Sci. Soc. Am. J. 44 (5), 892e898.
Rowe, R.K., 2005. Long-term performance of contaminant barrier systems, 45th Van Praagh, M., Torneman, N., Johansson, M., Ingelstedt Frendberg, L.E., Heander, E.,
Rankine Lecture Geotechnique 55 (9), 631e678. Johansson, A., 2011. Emerging organic contaminants in leachate e review and
Rowe, R.K., Abdellaty, K., 2013. Leakage and contaminant transport through a single risk assessment. In: Proceedings Sardinia 2011, Thirteenth International Waste
hole in the geomembrane component of a composite liner. J. Geotech. Geo- Management and Landfill Symposium S. Margherita di Pula, Cagliari, Italy; 3e7
environmental Eng. 139, 357e366. October 2011. CISA Publisher, Italy, p. 8.
Rowe, R.K., Abdellaty, K., 2012. Modeling contaminent transport through composite von Maubeuge, K.P., Sreenivas, K., Pohlmann, H., 2011. The New Generation of geo-
liner with a hole in the geomembrane. J. Geotech. Geoenvironmental Eng. 49, synthetic clay liners. In: Seminar Geosynthetics India’ 11 and an Introductory
773e781. Course on Geosynthetics, 22e24 September 2011 Chennai (Tamil Nadu), India, p. 8.
Rowe, R.K., Booker, J.R., 1987. An Efficient Analysis of Pollutant Migration through Yoo, J., Choi, J., Lee, T., Park, J., 2004. Organobentonite for sorption and degradation
Soil. Numerical Methods for Transient and Coupled Problems. In: Lewis, R.W., of phenol in the presence of heavy metals. Water, Air, Soil Pollut. 154, 225e237.
Hinton, E., Bettess, P., Schrefler, B.A. (Eds.). Wiley, New York, pp. 13e42. Zanzinger, H., Touze-Foltz, N., 2009. Clay geosynthetic barriers performance in
Rowe, R.K., Hosney, M.S., 2014. Hydraulic Conductivity and Interface Transmissivity landfill covers. In: GeoAfrica 2009 Cape Town 2e5 September 2009,
of GCLs below Poured Concrete (Submitted to Geosynthetics international). pp. 175e184.

You might also like