You are on page 1of 50

Review of Data on Gas Migration

through Polymer Encapsulants


Report to NDA - Radioactive Waste Management
Directorate

Your Reference: WBS code WT 15 05 10

Our Reference: SERCO/TAS/000500/001 – Issue 2

Date: September 2008


SERCO/TAS/000500/001 – Issue 2

SERCO/TAS/000500/001 – Issue 2 Page 2 of 50


SERCO/TAS/000500/001 – Issue 2

Title Review of Data on Gas Migration through Polymer


Encapsulants

Customer NDA – Radioactive Waste Management Directorate

Customer reference WBS code WT 15 05 10

Confidentiality, copyright © NDA 2010


and reproduction

Our Reference SERCO/TAS/000500/001 – Issue 2, September 2008

Technical & Assurance Services


Serco
Building 150
Harwell Science and Innovation Campus
Didcot
Oxfordshire
OX11 0RA
Telephone 01635 280415
Facsimile 01635 280305

www.serco.com/assurance

Technical & Assurance Services is part of


Serco Defence, Science and Technology, a division of
Serco Ltd.

In the areas where this work would be performed,


Technical & Assurance Services is certified to
ISO 9001 (2000) and ISO 14001.

SERCO/TAS/000500/001 – Issue 2 Page 3 of 50


SERCO/TAS/000500/001 – Issue 2

SERCO/TAS/000500/001 – Issue 2 Page 4 of 50


SERCO/TAS/000500/001 – Issue 2

PREFACE

This report was prepared under contract to the Nuclear Decommissioning Authority (NDA). It
was submitted to the NDA in January 2008 and summarises work carried out at that time. The
report has been reviewed by the NDA, but the views expressed and conclusions drawn are
those of the authors and do not necessarily represent those of the NDA.

Conditions of Publication

This report is made available under the NDA Transparency Policy. In line with this policy, the
NDA is seeking to make information on its activities readily available, and to enable interested
parties to have access to and influence on its future programmes. The report may be freely
used for non-commercial purposes. However, all commercial uses, including copying and re-
publication, require permission from the NDA. All copyright, database rights and other
intellectual property rights reside with the NDA. Applications for permission to use the report
commercially should be made to the NDA Information Manager.

Although great care has been taken to ensure the accuracy and completeness of the
information contained in this publication, the NDA cannot assume any responsibility for
consequences that may arise from its use by other parties.

© Nuclear Decommissioning Authority 2010. All rights reserved.

Bibliography

If you would like to see other reports available from NDA, a complete listing can be viewed at
our website www.nda.gov.uk, or please write to the Library at the address below.

Feedback

Readers are invited to provide feedback to the NDA on the contents, clarity and presentation of
this report and on the means of improving the range of NDA reports published. Feedback
should be addressed to:

John Dalton,
Head of Communications,
Nuclear Decommissioning Authority (Radioactive Waste Management Directorate),
Curie Avenue,
Harwell Science and Innovation Campus,
Didcot,
Oxon,
OX11 0RH, UK.

SERCO/TAS/000500/001 – Issue 2 Page 5 of 50


SERCO/TAS/000500/001 – Issue 2

SERCO/TAS/000500/001 – Issue 2 Page 6 of 50


SERCO/TAS/000500/001 – Issue 2

Executive Summary
The management of intermediate-level radioactive wastes (ILW) and some low-level
wastes (LLW) in the UK requires their packaging into a form suitable for long-term storage and
eventual disposal. Currently, the majority of packaging requirements can be fulfilled through the
use of cementitious materials to produce a stable wasteform. However, the challenges
presented by some wastes could require the use of organic polymer encapsulants. The
potential advantages offered by polymers include the capability to infiltrate void space,
immobilise fine particulate material and control corrosion of embedded metals. Polymers are
commonly classified into three different categories: thermoplastics, thermosets and elastomers.
Polymers in each category have similar properties that reflect their underlying molecular
structure. Thermoplastics, which often are simply referred to as ‘plastics’, are linear-branched
polymers that can be melted and remoulded with the application of heat. Thermosets, such as
epoxies, are heavily cross-linked polymers, which normally are rigid and cannot be remoulded
when heated. Elastomers are lightly cross-linked polymers, which can be stretched easily to
high extensions and will spring back to their original shape when the stress is released. The
most relevant materials for encapsulation of nuclear waste are the thermosets. To further
support the use of polymer encapsulation, the NDA – RWMD has commissioned research to
better understand and quantify gas migration in thermosetting polymers.

The objective of this study was to review information in the literature on the migration of gases
(radon, hydrogen, helium, methane, carbon dioxide, ethyne, argon and krypton) through
thermosetting polymers that are representative of proposed polymer encapsulants
(e.g. polyester, vinyl ester and epoxy), and to make recommendations for the bulk diffusion
coefficients of these gases in the thermosetting polymers.

The main literature search was conducted using the RAPRA abstract database that covers
academic and commercial literature in the polymer industry. Additional literature searches were
conducted using the British Library On line database, the Web-of-Science Citations database,
Google Scholar and the Energy Citations of the US Department of Energy. The resulting
abstracts from the searches were collated, and fifty of the most relevant papers were ordered.
Each of these papers was reviewed to determine the diffusion properties of the gases through
different polymers.

Permeability coefficients were obtained for a wide range of thermoset, thermoplastic and
elastomeric polymers. The values for the priority gases in the three main classes of
thermosetting resin (epoxy, poly vinyl ester and polyester) are summarised in the table below.
The results obtained for each of the gases indicate that epoxy resins are the least permeable of
the three types of resin. However, it should be noted that the permeability of a resin system
depends on a number of different factors including: the chemical composition of the resin, the
fillers and additives that are added to the resin and the conditions used to cure the resin
(temperature, pressure and time).

SERCO/TAS/000500/001 – Issue 2 Page 7 of 50


SERCO/TAS/000500/001 – Issue 2

The literature search has revealed that although there are considerable data on the diffusion of
gases through thermoplastics there are very few published values for the diffusion of gases
through thermosetting polymers.

Permeability Coefficients† for the Specified Gases through


the Three Main Classes of Thermosetting Resin

Material H2 CO2 He Ar Kr Rn CH4


Epoxy 1.07 0.14–0.24 0.32 0.15 0.11 0.01 —
Poly vinyl ester — 2.03–22.0 16.22 0.3–3.2 2.45 — —
Polyester 10.4–27.1 6.9–22.2 — — — — 0.36–1.39

cm ( STP ) ⋅ cm
3

Permeability coefficients are given in units of × 10 −10 at
cm ⋅ s ⋅ (cmHg )
2

temperatures between 20 and 35°C

SERCO/TAS/000500/001 – Issue 2 Page 8 of 50


SERCO/TAS/000500/001 – Issue 2

Contents
1 Introduction 11
1.1 Objective 11
1.2 Structure of Report 11

2 Background Material 13
2.1 Theory of Diffusion in Polymers 13
2.2 Generation of Gases from Radioactive Wastes 14
2.3 Factors Affecting the Diffusion of Gases through Polymers 15
2.3.1 Polymer chemistry 15
2.3.2 Free volume 16
2.3.3 Porosity and voids 16
2.3.4 Filler particles 17
2.3.5 Temperature 18
2.3.6 Mechanical stress and strain 19
2.3.7 Exposure to mixed diffusants 21
2.4 Measurement Techniques 21
2.4.1 Atmospheric pressure — manometric methods 21
2.4.2 Constant volume methods 21
2.4.3 Constant pressure methods 22
2.4.4 Carrier gas methods 22
2.4.5 Mass spectrometry methods 22
2.4.6 Oxygen transmission using a coulometric sensor 22

3 Permeability Data Obtained from the Literature 23


3.1 Literature Search 23
3.2 Tables of Permeability Data 23

4 Modelling the Diffusion of Gases through Polymers 40


4.1 Use of Fick's Laws 40
4.2 Use of Henry's Law 41
4.3 Software 41

5 Summary of Permeability Coefficients 43

6 References 46

SERCO/TAS/000500/001 – Issue 2 Page 9 of 50


SERCO/TAS/000500/001 – Issue 2

SERCO/TAS/000500/001 – Issue 2 Page 10 of 50


SERCO/TAS/000500/001 – Issue 2

1 Introduction
The Nuclear Decommissioning Agency (NDA), through its Radioactive Waste Management
Directorate (RWMD), has been charged with implementing the UK Government’s policy for the
long-term management of higher-activity radioactive waste by planning, building and operating a
Geological Disposal Facility (GDF). The UK has accumulated a legacy of radioactive waste
from electricity generation, defence activities and other industrial, medical, agricultural and
research activities. Radioactive wastes continue to be produced from these activities.

The management of intermediate-level wastes (ILW) and some low-level wastes (LLW) in
the UK requires their packaging into a form suitable for long-term storage and eventual disposal.
Currently, the majority of packaging specifications to produce a stable wasteform can be fulfilled
through the use of cementitious materials. These are relatively cheap and readily available,
have excellent physical, mechanical and thermal stability and are tolerant of a wide range of
wastes. However, the challenges presented by some wastes may be addressed best through
the use of organic polymer encapsulants. The potential advantages offered by polymers
include the capability to infiltrate void space, immobilise fine particulate material and control
corrosion of embedded metals. Also, polymers may provide compatibility with materials which
are unstable in cement.

A number of polymers have been considered for use in packaging radioactive wastes around
the world, each offering differing advantages (e.g. cost, versatility and enhanced performance).
In the UK, thermosetting polymers such as polyester and vinyl ester styrene have been
investigated.

Gases may be formed by radioactive wastes. To further support the use of polymer
encapsulation, the NDA – RWMD has commissioned this work to better understand and
quantify gas migration in thermosetting polymers.

1.1 Objective
The objectives of this study were:

• to review information in the literature on the migration of gases (i.e. radon, hydrogen,
helium, methane, carbon dioxide, ethyne, argon and krypton) through thermosetting
polymers that are representative of proposed polymer encapsulants (e.g. polyester resins,
vinyl ester resins (including DT900) and epoxy resins); and

• to make recommendations for the bulk diffusion coefficients of the above gases in the
thermosetting polymers.

The emphasis was on gathering data for the high priority gases (i.e. radon, hydrogen and
helium), but, if readily available, data for the lower priority gases (i.e. methane, carbon dioxide,
ethyne, argon and krypton) were also reviewed.

1.2 Structure of Report


The structure of this report is as follows. Section 2 provides background information on the
diffusion of gases through polymers, including: the theory of diffusion in polymers, the gases
which could be generated from radioactive wastes, the factors affecting diffusion of gases
through polymers, and the techniques used to measure diffusion. Section 3 summarises the
permeability coefficients that have been obtained for thermoset, thermoplastic and elastomeric

SERCO/TAS/000500/001 – Issue 2 Page 11 of 50


SERCO/TAS/000500/001 – Issue 2

polymers 1 . Section 4 discusses different approaches to modelling the diffusion of gases


through polymers. Finally, Section 5 gives a summary table of the relevant permeability
coefficients.

1
Polymers are commonly classified into three different categories: thermoplastics,
thermosets and elastomers. Polymers in each category have similar properties that reflect their
underlying molecular structure. Thermoplastics, which often are simply referred to as ‘plastics’,
are linear-branched polymers that can be melted and remoulded with the application of heat.
Thermosets, such as epoxies, are heavily cross-linked polymers, which normally are rigid and
cannot be remoulded when heated. Elastomers are lightly cross-linked polymers, which can be
stretched easily to high extensions and will spring back to their original shape when the stress is
released.
The most relevant materials for encapsulation of nuclear waste are the thermosets.

SERCO/TAS/000500/001 – Issue 2 Page 12 of 50


SERCO/TAS/000500/001 – Issue 2

2 Background Material
This subsection provides a summary of some background material relevant to the diffusion of
gases through polymer encapsulants, including: the theory of diffusion in polymers, the gases
which could be generated from radioactive wastes, the factors affecting diffusion of gases
through polymers, and the techniques used to measure diffusion.

2.1 Theory of Diffusion in Polymers


Polymers show a measurable permeability to gases, vapours and other small molecules and the
transport process is generally agreed to involve sorption at the high pressure interface, diffusion
through the solid and de-sorption at the low pressure interface.

For polymers above the glass transition temperature, equilibrium gas sorption has been found to
follow Henry's Law, and:

C = SP (2.1)

where

C is the concentration of gas [mol m-3],


S is the solubility coefficient [mol J-1], and
P is the applied pressure [Pa].

The diffusion process is described by Fick’s first and second laws:

dC
F = −D (2.2)
dx

and

dC d 2C
=D 2 (2.3)
dt dx
where

F is the gas flux [mol m-2s-1],


D is a concentration independent diffusion coefficient [m2s-1],
x is the thickness of the polymer [m], and
t is the time [s].

Fick’s first law and Henry’s Law lead to the expression:

dP
F = −k (2.4)
dx

where k, the permeability coefficient, is the product of the solubility and diffusion coefficients.
The permeability is characteristic of a permeant / polymer system if the permeant does not
interact with the polymer.

SERCO/TAS/000500/001 – Issue 2 Page 13 of 50


SERCO/TAS/000500/001 – Issue 2

If the permeant does interact with the polymer, the sorption of gas is unlikely to follow Henry’s
Law and the value of k may depend on environmental conditions such as pressure. For
practical purposes, the transmission rate Q often is used to describe the transport under a
specified set of conditions:

Q = k × ΔP (2.5)

The permeability coefficient k has been found to have an Arrhenius dependence on


temperature, and:

⎛ E ⎞
k = k 0 exp⎜ − ⎟ (2.6)
⎝ RT ⎠

where

k0 is the pre-exponent constant,


E is the activation energy [J mol-1],
R is the gas constant [J K-1mol-1], and
T is the temperature [K].

k, D and E are experimentally accessible by a range of methods (see Subsection 2.4).


Activation energies for diffusion processes typically are of the order 20 – 40 kJ mol-1, and
depend upon the capability of the polymer to generate holes by molecular motion and upon the
size of the diffusing molecule.

For polymers below the glass transition temperature, careful studies of sorption behaviour show
that the sorption isotherms do not follow Henry’s Law, but rather exhibit the form

S ′bP
C = SP + (2.7)
1 + bP

The second term represents an additional sorption mechanism which follows the Langmuir
isotherm, and so there are two populations of absorbed molecules in equilibrium with each
other. This characteristic complicates modelling of permeability data and has led to the
so-called ‘dual-sorption’ model.

2.2 Generation of Gases from Radioactive Wastes


Gases may be generated from packages of radioactive wastes by the following mechanisms:

• corrosion of metals in wastes and containers, giving hydrogen 2 ;

• radiolysis of water and certain organic materials in the packages, yielding mainly hydrogen;

• microbial degradation of various organic wastes, which give carbon dioxide and methane
as the chief products; and

2 14
C is present in irradiated metals where nitrogen impurities have been converted to
14
C by neutron capture. 14C could therefore be present in metals in elemental form or as metal
carbides. On dissolution, metal carbides could form hydrocarbons such as methane or small
organic molecules such as ethyne.

SERCO/TAS/000500/001 – Issue 2 Page 14 of 50


SERCO/TAS/000500/001 – Issue 2

• radioactive decay, giving helium (i.e. α-particles).

In the absence of oxygen and water, only the radiolysis of organic materials and radioactive
decay produce gases.

In the case of encapsulation in polymer, water may be carried over with the wastes, but would
be displaced from contact with the bulk waste upon addition of the polymer. Since diffusion of
water through the polymer to access the wastes is expected to be very slow, corrosion of the
metal wastes and microbial degradation of the organic wastes would be negligible. Therefore
the only source of gas (mainly hydrogen) would be radiolysis of the polymer.

In addition to the bulk gases above, radioactive gases also may be released from the wastes:

• radioactive isotopes of hydrogen (3H), carbon (14C), argon (39Ar and 42Ar) and krypton (81Kr
and 85Kr) could be released as a result of either corrosion of metals or diffusion out of the
metals 3 ;

• 3
H could be released from tritiated water as a result of the water’s consumption during
degradation of the wastes;

• 14
C could be present in some small organic molecules, and would be released during
microbial degradation;

• gaseous 14C could be released from irradiated graphite (according to some experimental
evidence); and

• radon 4 , of which 222Rn is the isotope of most interest as it has the longest half-life
(3.82 days), produced by radioactive decay in ILW packages.

Therefore the gases that have been considered in this study are radon, hydrogen and helium
(high priority) and methane, carbon dioxide, ethyne, argon and krypton (lower priority).

2.3 Factors Affecting the Diffusion of Gases through Polymers

2.3.1 Polymer chemistry

The chemical composition of a polymer has a strong influence on the solubility and diffusion of
gases through polymers. Polymers with polar groups, such as epoxies, have a strong affinity for
polar gas molecules, including water vapour. Such materials are known to be hygroscopic. In

3
Diffusion is expected to be very slow except for hydrogen.
4
The interest in this gas arises from its radioactivity rather than its volume. Radon is
not present initially in the waste, but is created by the decay of 226Ra. As the half-life of 226Ra is
1600 years, 222Rn will continue to be produced well beyond the operational phase of the
repository and, in the longer term, 222Rn will also be generated by 226Ra arising from the decay
of 238U (half-life 4.5 109 years).
As 226Ra is produced by the decay of 238U, via 234U and 230Th, the time required to
achieve secular equilibrium between the parent and the 226Ra progeny will be governed by the
half-lives of the various radionuclides in the chain and is very long (i.e. on a geological
timescale). Hence, although ‘natural’ uranium deposits may be in equilibrium with 226Ra,
processed and purified materials, such as fuel residues and related wastes, from which 226Ra
will have been separated, will not be at secular equilibrium. As a result, the maximum 226Ra
inventory will occur approximately 105 years after repository closure.

SERCO/TAS/000500/001 – Issue 2 Page 15 of 50


SERCO/TAS/000500/001 – Issue 2

contrast, the uptake of polar species is much lower in non-polar polymers [2]. In polar polymers,
the diffusion coefficients of polar organic molecules can increase with the concentration of
absorbed gas molecules, due to strong interactions between the molecules and polymer chains
that induce structural transformations such as swelling, crazing or partial dissolution of the
polymer [2].

2.3.2 Free volume

Free volume is an intrinsic property of the polymer and arises from the gaps left between
entangled polymer chains. Since the gaps are at the molecular scale, it is not possible to
observe free volume directly. Free volume can be thought of as extremely small-scale porosity,
but free volume pores are dynamic and transient in nature since the size (and existence) of any
individual free volume 'pore' depends on the vibrations and translations of the surrounding
polymer chains. The translation of the polymer chains can open and close 'pores' and
open / close channels between pores, providing 'pathways' for diffusion. The absorption and
diffusion of gas in plastics will depend to a considerable extent on the available free volume
within the polymer. The greater the free volume is, the higher the capacity for absorption and
the higher the mobility of the molecules within the polymer. Free volume depends on the
density and physical state of the polymer.

2.3.2.1 Crystallinity

Crystalline regions in polymers are more ordered than amorphous regions and free volume will
be lower in these regions. It is often assumed that the crystalline region is impermeable and
that the sorption depends only on the volume fraction of amorphous phase [3]. For many
polymer / gas combinations it has been observed that solubility constants are directly
proportional to the volume fraction of amorphous phase [4]. However, there are systems where
the solubilities have been found to be higher than those expected from the volume fraction of
amorphous phase; this has been attributed to the higher probability of denser regions of
amorphous material crystallising preferentially leaving the residual amorphous phase with a
lower density and higher concentration of 'holes' available for absorption.

2.3.2.2 Physical ageing

Physical ageing in glassy polymers occurs when a polymer is cooled to temperatures below the
glass transition temperature Tg. At the elevated temperature, the polymer is in the rubbery state
and the molecular configurations are in equilibrium. When the temperature is cooled through Tg
conformal changes needed to maintain an equilibrium structure are restricted by the reduced
mobility (increased relaxation times) at the lower temperatures. These non-equilibrium
structures have relatively high mobility. However, over time the molecular structure will
reconfigure to an equilibrium structure through the physical ageing process. The effect of
physical ageing is to reduce chain mobility and free volume of the polymer, which reduces the
mobility and solubility of diffusing molecules.

2.3.2.3 Molecular orientation

When a polymer is processed by a method that produces a preferred molecular orientation the
free volume will differ from a random orientated polymer. If the polymer is semi-crystalline,
orientation can cause enhanced crystallisation and, hence, reduced free volume. In amorphous
polymers, orientation can elongate the free volume voids and lead to directional dependence of
mobility, i.e. increased mobility in the direction of drawing and reduced transverse mobility.

2.3.3 Porosity and voids

Pores and voids are on a larger size scale than free volume and are 'permanent' features, not
depending on polymer chain motion. Pores / voids tend to result from the generation of

SERCO/TAS/000500/001 – Issue 2 Page 16 of 50


SERCO/TAS/000500/001 – Issue 2

'defects', e.g. inclusion of air during processing, but can also be generated in service (e.g. stress
generated crazing or chemical swelling). The volume fraction of voids in a sample will depend
on the imposed stress state. Hydrostatic tensile stress will tend to open voids whilst hydrostatic
compressive stresses will close voids.

Pores, like free volume, offer sites into which molecules can be absorbed and are far less of a
barrier to transport than solid polymer. Pores may also provide sites into which liquids and
vapours can condense and thereby dramatically increase their uptake. A high level of porosity
will increase permeability through increasing both the solubility and the effective diffusion
coefficient. If pores are linked (open pores), then diffusion rates through these channels will
lead to very much greater permeation than if the pores are isolated (closed pores).

2.3.4 Filler particles

Practically every polymeric material contains filler particles introduced to modify their physical
and processing properties. Filler particles can influence the molecular absorption behaviour in
two principal ways. Where the solubility of the filler differs from the polymer the absorption can
be either increased or decreased depending on the relative solubility of the molecule in the
polymer and filler. Most common inorganic filler particles (e.g. glass or carbon fibres, talc, clays,
silica) can usually be considered as impermeable in comparison to the polymer. However,
organic filler particles (e.g. rubber toughening agents, processing agents, natural fibres) may
have higher solubilities than the polymer. The overall effect on the component behaviour will
depend on the concentration of filler and can be characterised through a 'rule of mixtures'
approach provided that the solubility of the diffusant in each phase is known. However, this
approach ignores any role that may be played by the interface between the filler and the
polymer.

If the diffusing molecules have an affinity for the surfaces of the filler particles the interface
between the filler and the polymer may provide absorption sites for the molecules, increasing
the overall solubility. Surface energies of the filler particles and diffusing molecules may be
greater than the polymer, encouraging preferential absorption at the surfaces. There can be a
large increase in solubility of molecular species that strongly interact with the surfaces of
inorganic fillers. For example, inorganic fillers with hydrophilic or polar surface sites would have
a strong affinity for water (and other polar molecules) and substantially increase the overall
solubility of the polymer system. Nano-composite materials show substantially higher moisture
absorption than the base polymer, and, studies on montmorillonite clay nano-composites show
that the extent of absorption depends strongly on the surface properties of the treated nano-
particles [3]. In this situation, the overall effect on absorption is likely to depend strongly on the
surface area of the filler phase.

The presence of filler particles can also affect the diffusion behaviour. Diffusing molecules
would need to work their way around impermeable particles, increasing path lengths and
reducing mass transport rates. Improved barrier properties from fillers would be expected from
the increased lengths of diffusion paths. Some filled polymers, in particular composites, can
exhibit anisotropic diffusion behaviour. This is normally modelled with a single solubility
parameter determined from the saturation concentration and directionally dependent diffusion
coefficients. Several researchers (e.g. [5–9]) have presented three-dimensional finite-element
methodologies for simulating Fickian diffusion, often with temperature-dependent diffusion
coefficients, and predicting moisture diffusion in composite materials.

Particles lead to the presence of internal interfaces within the material that may play a role in the
diffusion process. Strong affinities between surfaces and diffusing molecules may provide a
driving potential for diffusion in addition to concentration gradients (or supply regions where the
local concentration of diffusing molecules far exceeds the solubility limit of the matrix). Strong
interactions between surfaces and diffusing molecules may allow for increased rates of diffusion
along the interfaces. Conversely, strong interactions between particles and surfaces may
reduce permeation rates. Strongly absorbed molecules will need to overcome an energy barrier

SERCO/TAS/000500/001 – Issue 2 Page 17 of 50


SERCO/TAS/000500/001 – Issue 2

to desorb and thus are likely to be far less mobile than molecules dissolved in the polymer
matrix. This mechanism, combined with the increased diffusion path lengths due to the
particles, is supported by observation that the addition of nano-fillers (such as clays) can
significantly increase the solubility of moisture in plastics but reduce the permeability of these
materials to water [3].

2.3.5 Temperature

Temperature has a significant effect on the permeability and diffusion properties of gas
molecules in polymers (e.g. see Figure 2.1 [10]). As the temperature increases, the mobility of
the molecular chains increases and thermal expansion leads to a reduced density. Therefore,
the free volume in the system will increase, leading to an increased solubility. Absorption and
diffusion normally follow Arrhenius behaviour.

The coefficient S is the solubility parameter, which varies with temperature T according to
Equation (2.8):

⎛ − ΔH s ⎞
S = S 0 exp⎜ ⎟ (2.8)
⎝ RT ⎠

The diffusion coefficient D is temperature dependent, and for ideal systems follows an
Arrhenius relationship, Equation (2.9), with an energy barrier to diffusion ED:

⎛ E ⎞
D = D0 exp⎜ − D ⎟ (2.9)
⎝ RT ⎠

These functions suggest predictable relationships between solubility, diffusion, permeation and
temperature. However, phase changes in polymers can significantly change mass transport
properties. For example, a discontinuity is normally observed in the absorption-temperature
curve at the glass transition temperature Tg, with higher absorption into the rubber phase
above Tg. Increased temperatures increase diffusion coefficients as free volume is increased,
and the increased vibrational energies of the polymer chains and permeant molecules increase
the energetic potential for translating between absorption sites. Diffusion coefficients normally
follow Arrhenius behaviour but there may be discontinuous behaviour near phase transitions.
Figure 2.2 shows diffusion coefficients and saturation levels for water in an epoxy [11]. The
highest test temperature (90°C) is slightly greater than Tg (85°C); both diffusion coefficient and
moisture uptake are substantially greater than they are for temperatures below Tg. There are
some circumstances where increasing temperature may reduce solubility, e.g. high
temperatures may reduce condensation of liquid molecules within pores reducing the quantity
that can be absorbed by this mechanism.

SERCO/TAS/000500/001 – Issue 2 Page 18 of 50


SERCO/TAS/000500/001 – Issue 2

Oxygen permeation in PET


9.0E-14 9.0E+06

8.0E-14 8.0E+06
Diffusion Coefficient *m2/s

7.0E-14 7.0E+06

Solubility (Kg/ms.Pa)
6.0E-14 6.0E+06

5.0E-14 5.0E+06

4.0E-14 4.0E+06

3.0E-14 3.0E+06
diffusion coefficient
2.0E-14 solubility 2.0E+06

1.0E-14 1.0E+06

0.0E+00 0.0E+00
0 5 10 15 20 25 30 35 40 45
Temperature (Deg C)

Figure 2.1 The effect of temperature on oxygen permeation in PET [10].

Moisture absorption in epoxy adhesive


8.E-12 8

diffusion coefficient
diffusion coefficient (m2/s)

6.E-12 6
saturation level (%)
saturation concentration

4.E-12 4

2.E-12 2

0.E+00 0
0 20 40 60 80 100
temperature (deg C)

Figure 2.2 The effect of temperature on moisture diffusion in an epoxy [11].

2.3.6 Mechanical stress and strain

In service, many polymer encapsulants will experience stress while the diffusion process is
commencing. Most permeation or mass uptake experiments are carried out free of external

SERCO/TAS/000500/001 – Issue 2 Page 19 of 50


SERCO/TAS/000500/001 – Issue 2

stress, although there may be unquantified internal, residual stresses, arising from processing,
present in the materials. The effect on molecular solubility and mobility will depend on the type
of stress applied. Hydrostatic tensile stress is expected to increase free volume and open
internal voids or crazes, providing additional sites into which molecules can absorb.
Conversely, hydrostatic compressive stress will reduce solubility by closing internal sites.

The results from standard permeation tests may not truly represent the properties of materials in
service. Results from tests performed with applied tensile stresses have been reported in the
literature. Boersma et al. [12] studied the effects of tensile stress applied to plastic films on their
oxygen permeation properties and reported an increase in permeation rate with stress,
e.g. Figure 2.3. Michaeli et al. [13] tested polymer films with and without plasma deposited
barrier coatings at different levels of strain. In both cases gas permeability increased with strain
with much greater effect for the coated material. Unstrained, the coated sample had
permeability around 20% of that of the uncoated sample. At 5–8% strain the differences
between the two samples was insignificant. The large increases in gas permeability in the
coated sample could be correlated with the strains where microscopic cracks begin to form in
the brittle barrier film.

Oxygen Permeation
180

160

140

120
Permeation

polystyrene
100
polysulphone
80 PMMA
LDPE
60 Linear (LDPE)
40 Linear (polystyrene)
Linear (polysulphone)
20 Linear (PMMA)
0
0 10 20 30 40 50 60 70
Stress (MPa)

Figure 2.3 The effect of applied stress on oxygen permeability (cm3.mm.atm-1.day-1 per
m2 of surface) for various plastic films [12].

Combined stress and chemical exposure is known to lead to enhanced degradation of polymers
and polymer structures. Plastics that are stressed in contact with chemicals can experience
environmental stress cracking failures in conditions, which individually, would not be a
problem [14]. The combination of stress and chemical exposure is known to degrade the
mechanical performance of adhesive joints [11, 14–22] far more rapidly than either factor
independently. The effect of absorbed chemicals on stress levels within polymer structures has
been modelled using stress-free Fickian diffusion coefficients and concentration dependent
mechanical properties [5, 6, 8, 18–22] but studies on the effect of stress on diffusion behaviour
are comparatively rare. Crocombe et al. [22] measured the uptake of moisture into stressed
bulk samples of epoxy using mass uptake experiments. They found a positive correlation
between stress and solubility but, unexpectedly, a negative correlation between stress and
diffusion coefficients (see Figure 2.4), although there was a comment that the large
uncertainties in the diffusion coefficient values made it difficult to draw reliable conclusions.

SERCO/TAS/000500/001 – Issue 2 Page 20 of 50


SERCO/TAS/000500/001 – Issue 2

12 4.00E-14

10

diffusion coefficient (m2/sec) .


3.00E-14
8
Mass uptake (%)

6 2.00E-14

4
1.00E-14
2

0 0.00E+00
0 5 10 15 20 0 5 10 15 20
tensile stress (MPa) tensile stress (MPa)

Figure 2.4 The effect of stress on solubility and diffusion coefficients [22].

2.3.7 Exposure to mixed diffusants

Permeation and diffusion tests are generally performed with exposure to a single gas, under
defined, normally constant, conditions. However, real service life is unlikely to be as simple.
For instance, encapsulants may be exposed to mixtures of chemicals contained in the waste.
Alternatively, pre-exposure to one gas may lead to a change in mass transport properties of a
second. The presence of a pre-absorbed species may block sites where the second molecule
may absorb thereby reducing solubility or alternatively increase solubility through swelling the
polymer or disrupting bonds that could block absorption by the second species. For example,
exposure of plastics to aqueous solutions with low pH and other ionic concentrations affects
permeability of other species [23].

2.4 Measurement Techniques

2.4.1 Atmospheric pressure — manometric methods

EN ISO 2556 [24] and ASTM D1434 [25] specify methods for determining gas permeation rates
using pressure changes that are monitored using a manometer. In this method the test gas is
contained within the feed chamber at atmospheric pressure. The chamber on the opposite side
of the sample is evaluated and hermetically sealed. As gas diffuses through the film, the
pressure, as determined by the height of the manometer mercury column, increases at the
low-pressure side. After an initial delay, the pressure-time curve will become linear and the
permeability can be calculated from the rate of pressure increase. This method measures the
total flow of gas through the sample and does not determine permeation of individual
components of mixtures.

2.4.2 Constant volume methods

ISO method ISO15105-1 [26] is similar to the manometric method except that a differential
pressure transducer is used in place of the manometer. The permeant side of the cell is
effectively at constant volume throughout the test (except for negligible changes due to the
deflection of the transducer diaphragm).

SERCO/TAS/000500/001 – Issue 2 Page 21 of 50


SERCO/TAS/000500/001 – Issue 2

2.4.3 Constant pressure methods

ASTM D1434 [25] provides for a constant pressure measurement of permeation. The basic
system design is the same as the manometric method. The permeant side is maintained at
atmospheric pressure whilst an increased gas pressure is applied to the feed side, typically
2 bar or greater. The diffusion of gas through the sample leads to an increase in volume (at
constant pressure) on the low-pressure side, which is measured by displacement of the
manometer fluid. The volume-time plot is used to calculate gas transmission rates and
permeability coefficients.

2.4.4 Carrier gas methods

Carrier gas methods (e.g. ISO 15105-1 [26]), often referred to as 'dynamic' methods, permit gas
to flow across either side of the sample at equal pressures. The test gas flows in the feeder
side of the permeation cell whilst a carrier gas flows through the diffusant side at a constant flow
rate. The carrier gas sweeps the permeating test gas to a detector. Thermal conductivity or
thermistor sensors can be used as general sensors for most gases.

2.4.5 Mass spectrometry methods

Mass spectrometers can be used as the sensor system in permeation experiments, combining
high sensitivity with simultaneous measurements of different species [27]. In one mass
spectrometer method the sample gas is introduced into a closed gas cell, which is then sealed.
The opposite side of the sample is evacuated, drawing gas molecules through the sample. The
partial pressures of the gases permeating from the gas cell are measured as a function of time
using a mass spectrometer. Since the cell is sealed, the gas pressure within the cell will decay
with time and, hence, the flux of gas permeating will decay with time. The decay of the partial
pressure of each component can be used determine the permeability coefficient of that
component. Since the determination of different species is simultaneous, this method can be
used to study the diffusion of mixtures. This method has high sensitivity to all materials
provided that mass / charge ratios of species or fragments are not the same (e.g. N2 and CO
have an atomic mass of 28 and cannot be distinguished). Mass spectrometry offers very high
sensitivity with accurate quantified detection possible to gas pressures below 10-10 Pa.

2.4.6 Oxygen transmission using a coulometric sensor

ASTM D3895 [28] and ISO 15105-2 [29] are isostatic test methods. Pure oxygen flow in the
feed side of the permeation cell diffuses through the sample and is swept by the pure nitrogen
carrier gas to the solid state sensor, where the oxygen concentration is determined through
current generated by electrochemical reactions.

SERCO/TAS/000500/001 – Issue 2 Page 22 of 50


SERCO/TAS/000500/001 – Issue 2

3 Permeability Data Obtained from the Literature

3.1 Literature Search


The main literature search was conducted using the RAPRA abstracts database
(http://abstracts.rapra.net) that covers academic and commercial literature published in the
polymer industry. Additional literature searches were then conducted using the British Library
On line database (https://www.bl.uk/inside), the Web-of-Science: Science Citations database
(http://scientific.thomson.com/products/wos), Google Scholar (http://scholar.google.co.uk) and
the Energy Citations of the US Department of Energy (http://www.osti.gov/energycitations). In
each case the following keywords were used in the search; Thermoset, Polyester, Vinyl ester,
Epoxy, Radon, Hydrogen, Helium, Gas, Diffusion, Transport, and Permeability. Citation
searches were then also conducted on each relevant paper that was identified using the Web-of
Science: Science Citations database.

The resulting abstracts that were obtained from the searches were collated, and fifty of the most
relevant papers were ordered. Each of these papers was reviewed to determine the diffusion
properties of the gases through the different polymers. The quality of the published data was
then assessed before inclusion in the tables, based upon the reliability of the measurement
techniques, the quoted uncertainty of the results, and the relevance of the materials and test
conditions to nuclear waste encapsulation.

3.2 Tables of Permeability Data


The permeability coefficients that have been obtained for each of the eight priority gases
(i.e. hydrogen, carbon dioxide, helium, argon, krypton, radon, methane and ethyne) are given in
Tables 3.1 to 3.8 5 respectively. Permeability coefficients for the diffusion of other gases and
vapours (such as, oxygen, nitrogen, neon, xenon, toluene, gasoline, methanol and water) are
given in Table 3.9. The results have been divided into the three main types of polymeric
material 6 : thermosetting polymers (such as epoxy, vinyl ester and polyester resins),
thermoplastic polymers (such as polyethylene, nylon and polyvinylchloride) and thermoelastic
polymers (such as neoprene and nitrile rubbers). The most relevant materials for nuclear waste
encapsulation are the thermosetting polymers and these are listed first in the tables. The data
for each material type is then given in ascending order of permeability with the materials with
the best barrier properties appearing at the top of each table.

As has been reported already in Subsection 2.1, permeability coefficients are determined from
both the sorption and the diffusion mechanisms, and in many cases these can change
significantly depending upon the test conditions, in particular temperature. The permeability
coefficients given in the tables have therefore been selected for temperatures as close to 25°C
as possible, and the temperature at which the result was obtained has been included in degrees
Celsius.

5
No permeability coefficients were obtained for ethyne (C2H2); instead Table 3.8 is for
ethylene (C2H4).
6
Polymers are commonly classified into three different categories: thermoplastics,
thermosets and elastomers. Polymers in each category have similar properties that reflect their
underlying molecular structure. Thermoplastics, which often are simply referred to as ‘plastics’,
are linear-branched polymers that can be melted and remoulded with the application of heat.
Thermosets are heavily cross-linked polymers, which normally are rigid and cannot be
remoulded when heated. Elastomers are lightly cross-linked polymers, which can be stretched
easily to high extensions and will spring back to their original shape when the stress is released.

SERCO/TAS/000500/001 – Issue 2 Page 23 of 50


SERCO/TAS/000500/001 – Issue 2

It was also found that a wide range of different units were used in the literature to express the
permeability of the gases. To enable the permeability of the different polymers to be compared
all the values have therefore been converted into the following units:

cm 3 ( STP ) ⋅ cm
× 10 −10 (3.1)
cm ⋅ s ⋅ (cmHg )
2

These are the most commonly used in the literature and are equivalent to the ‘barrer’, which is
commonly used in American literature.

SERCO/TAS/000500/001 – Issue 2 Page 24 of 50


SERCO/TAS/000500/001 – Issue 2

Table 3.1 Permeability† of Hydrogen through Polymers

Thermosets Permeability Reference Temperature


Epoxy resin
1.07 30 25

Polyester resin
10.40 - 27.10 31 30

Thermoplastics Permeability Reference Temperature


Polyoxymethylene (Acetal) 0.02 32 25
Polyvinylidene fluoride (PVDF) 0.32 33 23
0.34 34 25
Polychlorotrifluoroethylene (PCTFE) 0.94 35 25
0.54 - 0.98 33 25
0.39 36 25
0.59 37 25
Polyethylene terephthalate (PET)
0.60 33 25
0.82 35 25
Polypyrrole 1.06 33 25
1.38 36 25
Polyvinyl chloride (PVC)
1.70 - 8.89 35 25
Parylene 0.66 - 1.44 33 25
Ethylene chlorotrifluoroethylene 1.61 - 1.66 33 25
Polyphenylene sulfide 2.51 33 25
2.95 36 25
Polyethylene (HDPE)
2.34 - 3.70 33 25
Polymethylmethacrylate (PMMA) 3.30 38 25
Polyvinylidene chloride (PVDC) 4.75 37 25
Fluorinated ethylene propylene 5.80 - 5.88 33 25
Polyethylene (LDPE) 10.00 38 25
9.60 35 25
Polytetrafluoroethylene (PTFE) 7.86 - 17.86 33 25
24.00 38 25
Polysulfone 10.80 33 23
Polycarbonate 12.00 35 25

cm 3 ( STP ) ⋅ cm

Permeability coefficients are given in units of × 10 −10 with
cm ⋅ s ⋅ (cmHg )
2

temperatures in degrees Celsius

SERCO/TAS/000500/001 – Issue 2 Page 25 of 50


SERCO/TAS/000500/001 – Issue 2

Table 3.1 Permeability† of Hydrogen through Polymers (continued)

Thermoplastics (continued) Permeability Reference Temperature


12.50 36 25
Polystyrene 16.00 38 25
23.30 35 25
Poly(bis-(phenoxy)-phosphazene 23.00 39 30
Poly(propylene glycol) diacrylate 32.00 40 23
Polypropylene 41.00 35 20
Poly(dicyclohexylpolysiloxane) 400.00 41 35
Thermoplastic elastomers Permeability Reference Temperature
Polyisoprene (Butyl rubber) 7.20 35 25
Neoprene 13.60 35 25
Nitrile rubber (27% acrylonitrile) 16.00 32 25
Polybutadiene 41.90 35 25
45.00 32 25
Natural rubber
46.00 37 25
Silicone 269.80 33 25

cm 3 ( STP ) ⋅ cm

Permeability coefficients are given in units of × 10 −10 with
cm ⋅ s ⋅ (cmHg )
2

temperatures in degrees Celsius

SERCO/TAS/000500/001 – Issue 2 Page 26 of 50


SERCO/TAS/000500/001 – Issue 2

Table 3.2 Permeability† of Carbon Dioxide through Polymers

Thermosets Permeability Reference Temperature


Epoxy-amine
0.14 42 20
(cured for 60hr at 60°C)
Epoxy-amine
0.18 42 20
(cured for 3hr at 80°C)
Epoxy-amine
0.24 42 20
(cured for 3hr at 110°C)
Poly vinyl ester
2.03 43 35
(PVAc)
Poly vinyl ester
6.32 43 35
(poly(vinyl benzoate))
Poly vinyl ester
6.73 43 35
(poly(vinyl p-methylbenzoate))
Poly vinyl ester
8.26 43 35
(poly(vinyl cyclohexanecarboxylate))
Poly vinyl ester
22.20 43 35
(poly(vinyl-p-isopropylbenzoate))
Polyester resin
6.91 - 22.2 31 30

Thermoplastics Permeability Reference Temperature


Acrylonitrile 0.0006 - 0.0012 33 24
0.0008 44 25
Polyacrylonnitrile (PAN) 0.0018 45 25
0.0035 45 40
Polyvinyl alcohol 0.006 33 24
0.0083 - 0.07 33 23
0.036 46 25
Nylon 6 0.088 35 20
0.090 32 25
0.16 45 30
Nylon amorphous 0.0167 - 0.0578 33 23
0.024 - 0.036 33 24
Polyvinylidene chloride (PVDC)
0.030 35 30

cm ( STP ) ⋅ cm
3

Permeability coefficients are given in units of × 10 −10 with
cm ⋅ s ⋅ (cmHg )
2

temperatures in degrees Celsius

SERCO/TAS/000500/001 – Issue 2 Page 27 of 50


SERCO/TAS/000500/001 – Issue 2

Table 3.2 Permeability† of Carbon Dioxide through Polymers (continued)

Thermoplastics (continued) Permeability Reference Temperature


0.032 34 25
0.048 35 25
Polychlorotrifluoroethylene (PCTFE) 0.048 32 25
0.073 46 25
0.14 - 0.24 33 25
Polyvinylidene fluoride (PVDF) 0.034 33 25
Parylene 0.045 - 1.28 33 25
Nylon 66 0.047 - 0.096 33 23
Polyvinyl fluoride (PVF) 0.065 33 25
0.072 - 0.149 33 25
0.090 32 25
0.11 46 25
Polyethylene terephthalate (PET)
0.15 47 25
0.17 35 25
0.30 48 25
Polyethylene naphthalene (PEN) 0.092 48 25
Polyethylene isophthalate (PEI) 0.10 48 25
0.12 46 25
0.12 - 0.3 33 24
Polyvinyl chloride (PVC)
0.16 35 25
1.60 32 30
0.19 32 25
Polyoxymethylene (Acetal)
0.21 35 25
Polyamide (Trogamid) 0.33 49 30
Polypyrrole 0.34 33 25
0.36 35 25
1.80 45 30
2.00 46 25
Polyethylene (HDPE)
3.00 32 25
3.47 - 4.47 33 25
7.33 50 40

cm 3 ( STP ) ⋅ cm

Permeability coefficients are given in units of × 10 −10 with
cm ⋅ s ⋅ (cmHg )
2

temperatures in degrees Celsius

SERCO/TAS/000500/001 – Issue 2 Page 28 of 50


SERCO/TAS/000500/001 – Issue 2

Table 3.2 Permeability† of Carbon Dioxide through Polymers (continued)

Thermoplastics (continued) Permeability Reference Temperature


0.6 - 14.99 33 25
Polyethylene (MDPE)
10.80 50 39
Ethylene tetrafluoroethylene (ETFE) 1.50 33 25
Silicone 1.67 - 1.79 33 25
ASA 1.67 - 3.55 33 23
Polyester (PBT) 2.12 33 23
SAN 2.39 33 24
3.17 33 25
Polypropylene 9.20 35 30
9.20 32 30
4.2 - 6.59 33 24
10.50 35 25
Polystyrene
11.70 51 25
12.00 32 25
ABS 5.39 - 7.19 33 24
Polysulfone 5.70 33 23
6.00 - 16.19 33 24
10.00 46 25
Polyethylene (LDPE)
12.60 35 25
29.33 50 40
6.64 - 12.59 33 25
6.80 52 25
Polycarbonate
8.00 35 25
8.50 32 25
Polybutylene 7.13 - 8.54 33 22.8
Polyarylate 9.00 53 35
Fluorinated ethylene propylene (FEP) 10.01 33 25
11.70 35 25
Polytetrafluoroethylene (PTFE)
12.70 45 30
16.00 54 28
Polyurethane
17.99 33 25

cm 3 ( STP ) ⋅ cm

Permeability coefficients are given in units of × 10 −10 with
cm ⋅ s ⋅ (cmHg )
2

temperatures in degrees Celsius

SERCO/TAS/000500/001 – Issue 2 Page 29 of 50


SERCO/TAS/000500/001 – Issue 2

Table 3.2 Permeability† of Carbon Dioxide through Polymers (continued)

Thermoplastics (continued) Permeability Reference Temperature


Poly(bis-(phenoxy)-phosphazene 17.00 39 30
Polymethylpentene 83.75 - 91.37 33 23
Polyoctenamer 100.00 54 28
Poly(propylene glycol) diacrylate 130.00 40 23
Polydimethylsiloxane 1300.00 54 28
Poly(dicyclohexylpolysiloxane) 1400.00 41 35
Thermoplastic elastomers Permeability Reference Temperature
Polyisoprene (Butyl rubber) 5.16 35 25
Olefinic thermoplastic elastomers 20.07 - 46.02 33 23
Thermoplastic elastomer vinyl 21.32 - 41.12 33 25
Thermoplastic elastomer polyester 23.68 - 46.05 33 21.5
Neoprene 25.80 35 25
Nitrile rubber (27% acrylonitrile) 30.00 32 25
Thermoplastic elastomer styrenic 35.07 - 37.07 33 23
Thermoplastic elastomer polyamide 42-260 33 23
Thermoplastic elastomer polybutadiene 42.64 - 44.16 33 25
100.00 55 25
Natural rubber 130.00 32 25
153.00 35 25
Polybutadiene 138.00 35 25

cm ( STP ) ⋅ cm
3

Permeability coefficients are given in units of × 10 −10 with
cm ⋅ s ⋅ (cmHg )
2

temperatures in degrees Celsius

SERCO/TAS/000500/001 – Issue 2 Page 30 of 50


SERCO/TAS/000500/001 – Issue 2

Table 3.3 Permeability† of Helium through Polymers

Thermosets Permeability Reference Temperature


Epoxy
0.32 56 25

Poly vinyl ester


16.22 43 35
(poly(vinyl-p-isopropylbenzoate))
Thermoplastics Permeability Reference Temperature
Polyurethane 0.01 54 28
Polyoctenamer 0.01 54 28
Polyisoprene 0.03 54 28
Polydimethylsiloxane 0.30 54 28
Polyvinylidene chloride (PVDC) 0.31 35 34
Nylon 6 0.53 35 25
Polyacrylonnitrile (PAN) 0.53 45 25
Nylon 66 0.90 33 23
1.14 35 25
Polyethylene (HDPE) 1.48 33 23
2.53 50 40
1.31 33 23
Polyvinylidene fluoride (PVDF)
1.87 50 41
2.05 35 25
4.00 57 25
Polyvinyl chloride (PVC) 4.60 32 30
5.03 58 25
9.73 33 25
Polyethylene naphthalene (PEN) 2.12 47 25
2.16 33 25
Polychlorotrifluoroethylene (PCTFE)
6.80 35 25
Polyamide 2.80 54 41
Polyethylene terephthalate (PET) 3.16 47 25
Polyethylene (MDPE) 3.73 50 41
Polyamide (Trogamid) 4.00 49 30
Ethylene tetrafluoroethylene (ETFE) 5.39 33 25

cm 3 ( STP ) ⋅ cm

Permeability coefficients are given in units of × 10 −10 with
cm ⋅ s ⋅ (cmHg )
2

temperatures in degrees Celsius

SERCO/TAS/000500/001 – Issue 2 Page 31 of 50


SERCO/TAS/000500/001 – Issue 2

Table 3.3 Permeability† of Helium through Polymers (continued)

Thermoplastics (continued) Permeability Reference Temperature


6.70 57 25
6.90 59 25
Polyethylene (LDPE)
7.00 38 25
8.53 50 41
Polymethylmethacrylate (PMMA) 7.00 38 25
Polysulfone 11.76 33 23
13.00 52 25
Polycarbonate
15.00 57 25
16.00 38 25
Polystyrene
18.70 35 25
Polyarylate 18.00 53 35
Polypropylene 38.00 35 25
Thermoplastic elastomers Permeability Reference Temperature
Thermoplastic elastomer polyester 4.2 - 20.7 33 21.5
Thermoplastic elastomer polyamide 4.6 - 23.5 33 23
Polyisoprene (Butyl rubber) 8.38 35 25
Nitrile rubber (27% acrylonitrile) 12.00 32 25
Natural rubber 29.00 32 25

cm ( STP ) ⋅ cm
3

Permeability coefficients are given in units of × 10 −10 with
cm ⋅ s ⋅ (cmHg )
2

temperatures in degrees Celsius

SERCO/TAS/000500/001 – Issue 2 Page 32 of 50


SERCO/TAS/000500/001 – Issue 2

Table 3.4 Permeability† of Argon through Polymers

Thermosets Permeability Reference Temperature


Epoxy
0.15 60 25

Poly vinyl ester


0.30 43 35
(PVAc)
Poly vinyl ester
0.61 43 35
(poly(vinyl benzoate))
Poly vinyl ester
0.86 43 35
(poly(vinyl p-methylbenzoate))
Poly vinyl ester
1.56 43 35
(poly(vinyl cyclohexanecarboxylate))
Poly vinyl ester
3.15 43 35
(poly(vinyl-p-isopropylbenzoate))
Thermoplastics Permeability Reference Temperature
3.40 57 25
Polyethylene (LDPE) 3.30 38 25
2.92 50 42
1.69 35 25
Polyethylene (HDPE)
1.02 33 20
0.80 35 25
Polycarbonate 0.55 57 25
0.80 61 35
0.01 35 25
Polyvinyl chloride (PVC)
0.07 57 25
Polyvinylidene chloride (PVDC) 5.80 38 25
Polyacrylonnitrile (PAN) 0.01 45 25
Poly(bis-(phenoxy)-phosphazene 6.70 39 30
Polyethylene (MDPE) 1.47 50 42
Thermoplastic elastomers Permeability Reference Temperature
Chlorosulphonated polyethylene 1.34 62 25
Olefinic thermoplastc elastomers 6.02 - 9.09 33 23
11.00 32 25
Natural rubber
22.80 35 25

cm 3 ( STP ) ⋅ cm

Permeability coefficients are given in units of × 10 −10 with
cm ⋅ s ⋅ (cmHg )
2

temperatures in degrees Celsius

SERCO/TAS/000500/001 – Issue 2 Page 33 of 50


SERCO/TAS/000500/001 – Issue 2

Table 3.5 Permeability† of Krypton through Polymers

Thermosets Permeability Reference Temperature


Epoxy
0.107 60 25

Poly vinyl ester


2.450 43 35
(poly(vinyl-p-isopropylbenzoate))
Thermoplastics Permeability Reference Temperature
Polyacrylonnitrile (PAN) 0.000003 45 25
Polyvinyl chloride (PVC) 0.035 57 25
Polycarbonate 0.500 57 25
Polystyrene 1.000 35 40
Polyethylene (LDPE) 4.500 57 25
Thermoplastic elastomers Permeability Reference Temperature
2.200 62 32
Chlorosulphonated polyethylene
2.380 32 25

cm 3 ( STP ) ⋅ cm

Permeability coefficients are given in units of × 10 −10 with
cm ⋅ s ⋅ (cmHg )
2

temperatures in degrees Celsius

SERCO/TAS/000500/001 – Issue 2 Page 34 of 50


SERCO/TAS/000500/001 – Issue 2

Table 3.6 Permeability† of Radon‡ through Polymers

Thermosets Permeability Reference Temperature


Epoxy
0.013 60 25

Thermoplastics Permeability Reference Temperature


Polyvinyl chloride (PVC) 0.004 57 25
Polycarbonate 0.18 57 25
Polyethylene (LDPE) 15.00 57 25

cm 3 ( STP ) ⋅ cm

Permeability coefficients are given in units of × 10 −10 with
cm ⋅ s ⋅ (cmHg )
2

temperatures in degrees Celsius



Permeability coefficients for radon through the barrier films used in the construction
industry were not obtained during the literature search. However, a comparative
study (conducted jointly by UNAM in Mexico and Oak Ridge National Laboratory)
was found. In this study [63], radon gas detectors were placed inside six
containers that had been coated with different barrier materials. The containers
were exposed to a known concentration of radon, and the gas entering each
container was measured over a period of three months. The amount of gas
absorbed by each of the barrier materials was calculated with respect to an
unprotected container. The results are shown in the table below. Although these
results can be used to compare the different barrier materials, permeability values
were not given, and so the results cannot be directly related to those obtained from
other studies.

Radon Absorbance of Six Barrier Films with Respect to


an Unprotected System [63]

Barrier material Radon absorbance


Cellulose esther wall paper (VINIPLAST) 15%
Vinyl acetate paint (VIN-ECO EXPRESS) 37%
Metal pigment (COMEX) 41%
Epoxy resin (VAR-ECO) 54%
Phenolic resin (COMEX) 58%
Acrylic and amino resins (ECO EXPRESS) 60%

SERCO/TAS/000500/001 – Issue 2 Page 35 of 50


SERCO/TAS/000500/001 – Issue 2

Table 3.7 Permeability† of Methane through Polymers

Thermosets Permeability Reference Temperature


Polyester resin
0.365-1.390 31 30

Thermoplastics Permeability Reference Temperature


Polypyrrole 0.006 33 23
Polyamide (Trogamid) 0.018 49 30
Fluoropolymer (ETFE) 0.120 33 23
Polyurethane 0.210 54 28
Polysulfone 0.225 33 23
Polyethylene (HDPE) 0.863 33 20
Polyethylene (HDPE) 1.227 50 39
Polyethylene (MDPE) 1.600 50 39
Poly(bis-(phenoxy)-phosphazene 1.700 39 30
Polyethylene (LDPE) 5.733 50 41
Polyoctenamer 6.000 54 28
Polyisoprene 9.000 54 28
Polydimethylsiloxane 90.000 54 28
Poly(dicyclohexylpolysiloxane) 350.000 41 35

cm ( STP ) ⋅ cm
3

Permeability coefficients are given in units of × 10 −10 with
cm ⋅ s ⋅ (cmHg )
2

temperatures in degrees Celsius

SERCO/TAS/000500/001 – Issue 2 Page 36 of 50


SERCO/TAS/000500/001 – Issue 2

Table 3.8 Permeability† of Ethylene through Polymers

Thermoplastics Permeability Reference Temperature


Polyethylene (HDPE) 1.690 33 20

cm 3 ( STP ) ⋅ cm

Permeability coefficients are given in units of × 10 −10 with
cm ⋅ s ⋅ (cmHg )
2

temperatures in degrees Celsius

SERCO/TAS/000500/001 – Issue 2 Page 37 of 50


SERCO/TAS/000500/001 – Issue 2

Table 3.9 Permeability† of Other Gases through Thermosets

Thermosets Gas Permeability Reference Temperature


Epoxy-amine
Oxygen 0.128 64 30

Polyester
Oxygen 0.014 65 23
(Unsaturated L1430)
Polyester
Oxygen 0.027 33 25
(Cross-linked)
Polyester
Oxygen 0.041 65 23
(Unsaturated Type 802)
Polyester
Oxygen 1.60 - 5.54 31 30

Polyester
Nitrogen 0.28 - 1.37 31 30

Poly vinyl ester


Neon 6.310 43 35
(poly(vinyl-p-isopropylbenzoate))
Epoxy
Xenon 0.013 60 25

Poly vinyl ester


Xenon 1.580 43 35
(poly(vinyl-p-isopropylbenzoate))
Epoxy
Toluene 0.001 30 37
(Bisphenol-A)
Poly vinyl ester
Toluene 0.001 30 37
(Novolac)
Polyester
Toluene 0.027 30 37
(unsaturated)
Epoxy
Gasoline 0.001 30 37
(Bisphenol-A)
Poly vinyl ester
Gasoline 0.001 30 37
(Novolac)
Polyester
Gasoline 0.013 30 37
(unsaturated)
Epoxy
Methanol 0.067 30 37
(Bisphenol-A)
Poly vinyl ester
Methanol 0.107 30 37
(Novolac)
Polyester
Methanol 0.107 30 37
(unsaturated)

cm 3 ( STP ) ⋅ cm

Permeability coefficients are given in units of × 10 −10 with
cm ⋅ s ⋅ (cmHg )
2

temperatures in degrees Celsius

SERCO/TAS/000500/001 – Issue 2 Page 38 of 50


SERCO/TAS/000500/001 – Issue 2

Table 3.9 Permeability† of Other Gases through Thermosets (continued)

Thermosets Gas Permeability Reference Temperature


Epoxy
Moisture 0.013 30 37
(Bisphenol-A)
Poly vinyl ester
Moisture 0.013 30 37
(Novolac)
Polyester
Moisture 0.053 30 37
(unsaturated)

cm 3 ( STP ) ⋅ cm

Permeability coefficients are given in units of × 10 −10 with
cm ⋅ s ⋅ (cmHg )
2

temperatures in degrees Celsius

SERCO/TAS/000500/001 – Issue 2 Page 39 of 50


SERCO/TAS/000500/001 – Issue 2

4 Modelling the Diffusion of Gases through


Polymers
It is appropriate now to discuss the modelling approach described in Subsection 2.1 (the
'conventional approach') in relation to modern understanding of methodologies that can predict
transport processes in materials.

4.1 Use of Fick's Laws


A fundamental assumption of the ‘conventional approach’ is that mass transport is governed by
Fick's laws. The first law, Equation (2.2), relates the gas flux to its concentration gradient, and
the second law inserts the first law into the continuity equation to obtain Equation (2.3), which is
the classical time-dependent diffusion equation for a one-dimensional problem and for a single
diffusing species. The first law introduces the classical diffusion coefficient D, and there is an
implied assumption that the diffusion of a particular gas species is not affected by the presence
of other species. This is likely to be a reasonable approximation for dilute concentrations of
gases in a polymer.

The flux F of the gas species and concentrations requires clear definition. According to
continuum thermodynamics [66] the diffusive flux vector for the k th species can be defined as:

Fk = ρ k (v k − v ) (4.1)

where

ρk is the mass density of the k th species, having the dimensions of mass per unit
volume,
vk is the mean velocity of the k th species, measured relative to a fixed frame of
reference, and
v is the mean velocity 7 for the system as a whole.

Quite often, a molar based description of transport is used in which the diffusive flux vector for
the k th species is defined by:

Fk∗ = c k (v k − v ′) (4.2)

where

ck is the molar concentration for the k th species, having the dimensions of moles per
unit volume.

7
The velocity v is the barycentric velocity of the system, and it certainly will be non-zero
when considering diffusion in a moving gas stream. The barycentric velocity for n different
n

∑ρ k vk n
species is defined by v = k =1
, where ρ = ∑ ρk is the local mass density of all the
ρ k =1
species in the system.

SERCO/TAS/000500/001 – Issue 2 Page 40 of 50


SERCO/TAS/000500/001 – Issue 2

The velocity v ′ is the mean molar velocity for n different species. It should be noted that the
mean velocities v and v ′ are in general different.

If either of the diffusive fluxes is used in Fick's first law, Equation (2.2), then the value of the
diffusion coefficient will depend upon the definition of the flux. It is thus vital that the definitions
of the flux and concentration must be clearly given when reporting or using values of diffusion
coefficients.

It should be noted that both the flux definitions lead to diffusion coefficients having the same
dimensions (L2T-1).

4.2 Use of Henry's Law


In the ‘conventional approach’, Henry's law, Equation (2.1), is used to relate the 'equilibrium' gas
concentration at a given point to the local pressure P. On substitution of Equation (2.1) in Fick's
first law, Equation (2.2), the diffusive flux is then related to the gradient of the pressure
introducing the permeability coefficient k. Experiments can then be undertaken to measure the
permeability directly.

An examination of the basis for Henry's law reveals that it is defined to be the amount of a given
gas that is dissolved in a given type and volume of fluid and it is directly proportional to the
partial pressure of that gas in the gas mixture above the fluid. Henry's law should only be used
when the system is in equilibrium so that temperature, pressure and concentrations are all
uniform.

At best, Henry's law can provide value(s) for the concentration(s) of the gas species at the
external surface(s), if local equilibrium conditions are assumed. It should not be used to relate
the gas concentration within the polymer to the local pressure. However, if the system has
reached a steady state, where there are uniform concentration and pressure gradients through
the polymer, and a well-defined flux has been established, then the measured permeability will
be a function of the transport coefficients that multiply the concentration gradient and pressure
gradient in a generalised version of Fick's law:

F = − D∇C − k∇P (4.3)

If the concentration gradient term dominates, then the measured permeability is an


approximation to the diffusion coefficient, D. A key issue is the nature of the assumptions that
an experimental method makes to estimate the given permeability.

4.3 Software
Plastics are widely used for making structures and coatings in the civil, chemical and consumer
industries. This is because of their corrosion resistance, low weight and mechanical strength.
Although the polymeric matrix seems to resist chemicals, moisture and chemicals can diffuse
through the external boundary into the material. This may affect the mechanical performance of
the plastic, the material that a plastic coating should protect, and influence the interface
between a plastic and another material leading to leakage that could pollute the environment. A
review of mass diffusion in polymers has been given by Crank and Park [2].

The fact that the process of diffusion is not fully understood leads to problems of interpreting
experimental data and problems of determining whether diffusion is a major factor in the service
life of a structure. Accurate experimental data for transient diffusion in polymer composite
systems is particularly difficult to obtain due to the long saturation times, which can lead to other

SERCO/TAS/000500/001 – Issue 2 Page 41 of 50


SERCO/TAS/000500/001 – Issue 2

types of damage, especially if the material is under tensile load. Techniques for monitoring
water absorption in fibre-reinforced polymer composites have been studied by Broughton and
Lodeiro [9]. A combination of experimental approaches and predictive analysis was considered
as a method of obtaining through-thickness moisture absorption characteristics for polymeric
materials in aqueous environments. A comprehensive review of environmental effects on
polymer composite materials has been conducted by Springer [5]. In particular, analyses and
experimental results for moisture content, absorption and desorption in composites under
transient conditions are presented.

Modelling has become important for the estimation of properties and the change in properties of
materials when subjected to diffusion by external species. Springer [5] has developed a simple
one-dimensional numerical model based on finite differences and Fickian diffusion [67] for
polymer composite applications. Incropera and De Witt [68] have given expressions for
two-dimensional finite difference solutions for species diffusion with an internal species
generation term. Duncan et al. [69] have reviewed measurement and modelling of permeation
and diffusion in polymers and have developed a three-dimensional numerical model based on
the finite volume method with an aim to model interfacial diffusion. Hedenqvist et al. [70] have
developed a new mass transfer model based on an implicit multi-step integration algorithm to
include concentration-dependent diffusivities, and it has been developed and applied to
multi-layer systems of water diffusing in polyesters. The model was used to predict curves for
diffusivity as a function of water volume as well as predicted water desorption curves. A
comprehensive review of the theoretical background of the modelling of diffusion and
permeation in heterogeneous systems with polymer matrices has been given by Barrer [71]. A
full review of analytical solutions for mass transfer is given by Crank [72].

The increasing development of powerful computers and design software has led to a wider
adoption of simulation tools by industry. The quality of simulations is dependent upon the use
of realistic materials models, reliable software and input data. As a result there have been
many software packages developed using different numerical methods. The software package
Fluent (www.fluent.com) has been successful in solving three-dimensional computational fluid
dynamics (CFD) based on the finite volume (FV) technique [73]. The finite-element
technique [74] has been used extensively, led by ANSYS (www.ansys.com) and COMSOL
(www.comsol.com) for application to multiphysics transport problems.

SERCO/TAS/000500/001 – Issue 2 Page 42 of 50


SERCO/TAS/000500/001 – Issue 2

5 Summary of Permeability Coefficients


An extensive literature search has been conducted to obtain values for the permeability of
gases (i.e. hydrogen, carbon dioxide, helium, argon, krypton, radon and methane) through
thermosetting polymers (e.g. epoxy, vinyl ester and polyester resins) that are representative of
the polymer encapsulants proposed for use in packaging radioactive wastes.

Permeability coefficients were obtained for a wide range of different thermoset, thermoplastic
and elastomeric polymers. The values for the priority gases in the three main classes of
thermosetting resin (epoxy, poly vinyl ester and polyester) are summarised in Table 5.1 below.
Data for other polymers systems and other gases are detailed in Section 3. The results
obtained for each of the gases are consistent, indicating that epoxy resins are the least
permeable of the three types of resin. However, it should be noted that the permeability of a
resin system depends on a number of different factors including: the chemical composition of
the resin [7, 40], the fillers and additives that are added to the resin and the conditions used to
cure the resin (temperature, pressure and time [11]).

Table 5.1 Values in the Literature for the Permeability† of Gases through Thermosets

Material Gas Permeability Reference Temperature


Epoxy
Hydrogen 1.07 49 25

Polyester
Hydrogen 10.4 - 27.1 7 30

Epoxy-amine Carbon
0.14 11 20
(cured for 60hr at 60°C) dioxide
Epoxy-amine Carbon
0.18 11 20
(cured for 3hr at 80°C) dioxide
Epoxy-amine Carbon
0.24 11 20
(cured for 3hr at 110°C) dioxide
Poly vinyl ester Carbon
2.03 40 35
(PVAc) dioxide
Poly vinyl ester Carbon
6.32 40 35
(poly(vinyl benzoate)) dioxide
Poly vinyl ester Carbon
6.73 40 35
(poly(vinyl p-methylbenzoate)) dioxide
Poly vinyl ester Carbon
8.26 40 35
(poly(vinyl cyclohexanecarboxylate)) dioxide
Poly vinyl ester Carbon
22.20 40 35
(poly(vinyl-p-isopropylbenzoate)) dioxide
Polyester Carbon
6.9 - 22.2 7 30
dioxide

cm 3 ( STP ) ⋅ cm

Permeability coefficients are given in units of × 10 −10 with
cm ⋅ s ⋅ (cmHg )
2

temperatures in degrees Celsius

SERCO/TAS/000500/001 – Issue 2 Page 43 of 50


SERCO/TAS/000500/001 – Issue 2

Table 5.1 Values in the Literature for the Permeability† of Gases through Thermosets
(continued)

Material Gas Permeability Reference Temperature


Epoxy
Helium 0.32 8 25

Poly vinyl ester


Helium 16.22 40 35
(poly(vinyl-p-isopropylbenzoate))
Epoxy
Argon 0.15 47 25

Poly vinyl ester


Argon 0.30 40 35
(PVAc)
Poly vinyl ester
Argon 0.61 40 35
(poly(vinyl benzoate))
Poly vinyl ester
Argon 0.86 40 35
(poly(vinyl p-methylbenzoate))
Poly vinyl ester
Argon 1.56 40 35
(poly(vinyl cyclohexanecarboxylate))
Poly vinyl ester
Argon 3.15 40 35
(poly(vinyl-p-isopropylbenzoate))
Epoxy
Krypton 0.11 47 25

Poly vinyl ester


Krypton 2.45 40 35
(poly(vinyl-p-isopropylbenzoate))
Epoxy
Radon 0.01 47 25

Polyester
Methane 0.36 - 1.39 7 30

cm 3 ( STP ) ⋅ cm

Permeability coefficients are given in units of × 10 −10 with
cm ⋅ s ⋅ (cmHg )
2

temperatures in degrees Celsius

The permeability coefficients summarised in Table 5.1 make it possible to calculate the amount
of gas that will penetrate through a polymer in a specific time (dependent on the gas pressure,
the thickness and area of the polymer, and the temperature). For example, from Fick’s first law,
assuming the diffusion coefficient is constant and the sorption isotherm is linear 8 , the
steady-state flux of gas F through a slab is:

ΔP
F = −k (5.1)
l

8
In other words, there is a linear relationship between the external gas pressure and
the corresponding equilibrium concentration within the slab.

SERCO/TAS/000500/001 – Issue 2 Page 44 of 50


SERCO/TAS/000500/001 – Issue 2

where

k is the permeability coefficient,


ΔP is the difference in gas (partial) pressure between the outlet and inlet surfaces,
and
l is the thickness of the slab.

Hence, the total amount of gas Q to have passed through an area A of the slab after time t is:

ΔP
Q = −k At (5.2)
l
Using the permeability coefficients it is therefore possible to compare the volumes of gas that
could escape from different organic polymer encapsulants.

SERCO/TAS/000500/001 – Issue 2 Page 45 of 50


SERCO/TAS/000500/001 – Issue 2

6 References
1. “Generic repository studies. The Nirex phased disposal concept”, Nirex Report N/074,
2003.

2. J Crank and G S Park, “Diffusion in polymers”, Academic Press, London, 1968.

3. G Gorrasi, M Tortora, V Vittoria, E Pollet, B Lepoittevin, M Alexandre and P Dubois, “Vapor


barrier properties of polycaprolactone montmorillonite nanocomposites: effect of clay
dispersion”, Polymer, 44, 2271-2279, 2003.

4. H B Hopfenberg and V Stannett, “The diffusion and sorption of gases and vapours in
glassy polymers, The Physics of Glassy Polymers”, edited by R N Haward, Applied
Science Publishers, London, 1st edition, 1973.

5. G S Springer, “Environmental effects on composite materials”, Technomic Publishing


Company, Westport CT, 1981.

6. G S Springer, “Environmental effects on composite materials”, Volume 2, Technomic


Publishing Company, Lancaster Pa, 1984.

7. Y Yunn-Tzu and K Pochiraju, “Three-dimensional simulation of moisture diffusion in


polymer composite materials”. Polymer, Plastics Technology and Engineering, 42,
737-756, 2003.

8. Y Weitsman, “Stresses in adhesives due to moisture and temperature”, J. Composite


Mater., 11, 378, 1977.

9. W R Broughton and M J Lodeiro, “Techniques for monitoring water absorption in


fibre-reinforced composites”, NPL Measurement Note CMMT(MN)064, 2000.

10. R Auras, B Harte and S Selke, “Effect of water on the oxygen properties of polyethylene
tetraphalate and poly(lactide) films”, J. Applied Polymer Sci., 92, 1790-1803, 2004.

11. R A Gledhill, A J Kinloch and S J Shaw, “A model for predicting joint durability”,
J. Adhesion, 11, 3-15, 1980.

12. A Boersma, D Cangialosi and S J Picken, “Mobility and solubility of antioxidants and
oxygen in glassy polymers. III Influence of deformation and orientation on oxygen
permeability”, Polymer, 44, 2463-2471, 2003.

13. W Michaeli, S Gobel and R Dahlmann, “The influence of strain on the properties of plasma
polymerised permeation barriers”, J. Polymer Eng., 24, 107-122, 2004.

14. A S Maxwell and A Turnbull, “Environmental stress cracking of polymeric materials”,


NPL Report CMMT(A)288, September 2000.

15. M R Bowditch, D Hiscock and D A Moth, “The relationship between hydrolytic stability of
adhesive joints and equilibrium water content”, in Proceedings of Adhesion ‘90, The
Plastics and Rubber Institute, London, 1990.

16. R A Gledhill, A J Kinloch and S J Shaw, “A model for predicting joint durability”,
J. Adhesion, 11, 3-15, 1980.

17. S J John, A J Kinloch and F L Matthews, “Measuring and predicting the durability of
bonded carbon fibre/epoxy composite joints”, Composites, 22, 121-127, 1991.

SERCO/TAS/000500/001 – Issue 2 Page 46 of 50


SERCO/TAS/000500/001 – Issue 2

18. W K Loh, A D Crocombe, M M Abdel Waugh and I A Ashcroft, “Modelling anomalous


moisture uptake and thermal characteristics of a rubber toughened epoxy adhesive”,
Int. J. Adhesion and Adhesives, 25, 1-12, 2005.

19. H O Hambly, J Pan, A D Crocombe and P Megalis, “Diffusion Laws Governing water
uptake of adhesives”, in Proceedings of EURADH '96 – ADHESION '96, Institute of
Materials, 281-286, 1996.

20. D C C Lam, F Yang and P Tong, “Chemical kinetic model of interfacial degradation of
adhesive joints”, IEEE Transactions on Components and Packaging Technologies, 22,
215-220, 1999.

21. W R Broughton and G Hinopoulos, “An improved modelling approach of moisture


absorption in adhesive joints using the Finite Element method”, NPL Report CMMT(A)204,
1999.

22. A D Crocombe, J Pan and H O Hambly, “The coupling between stress and moisture in
adhesives loaded in a wet environment”, Proceedings of EURADH '96 – ADHESION '96,
Institute of Materials, 217-222, 1996.

23. T Yoshizawa, Y Shinya, K J Hong and T Kajiuchi, “pH- and temperature-sensitive


permeation through polyelectrolyte complex films composed of chitosan and
polyalkyleneoxide-maleic acid copolymer”, J. Membrane Sci., 241, 347-354, 2004.

24. BS EN ISO 2556: 2001, “Plastics – determination of the gas transmission rate of films and
thin sheets under atmospheric – manometric method”, 2001.

25. ASTM D1434 – 82, “Standard test methods for determining gas permeability characteristics
of plastic film and sheeting”, 1982.

26. BS ISO 15105-1: 2002, “Plastics – film and sheeting – determination of gas transmission
rate – Part 1: differential-pressure method”, 2002.

27. H Norenberg, T Mityamoto, Y Tsukahara, G D W Smith and G A D Briggs, “Mass


spectrometric estimation of gas permeation coefficients for thin polymer membranes”,
Rev. Sci. Instruments, 70, 2414-2420, 1999.

28. ASTM D3895 – 95, “Standard test method for oxygen gas transmission rate through plastic
film and sheeting using a coulometric sensor”, 1995.

29. BS ISO 15105-2: 2003, “Plastics – film and sheeting – determination of gas transmission
rate – Part 2: equal-pressure method”, 2002.

30. S van der Wal, “Composites agency”, http://www.composite-agency.com, December 2007.

31. J Zhang, G S Sun and X H Hou, “Gas-permeability in an aromatic polyester”,


Macromolecules, 26, 7176-7181, 1993.

32. W J Roff and J R Scott, “Fibres, films, plastics and rubbers – handbook of common
polymers”, Butterworths, 1971.

33. L K Massey, “Permeability properties of plastics and elastomers”, Plastics Design Library,
New York, 2003.

34. D W Brubaker and K Kammermeyer, “Flow of gases through plastic membranes", Ind. Eng.
Chem., 45, 1148-1152, 1953.

SERCO/TAS/000500/001 – Issue 2 Page 47 of 50


SERCO/TAS/000500/001 – Issue 2

35. M J Guest and J Dawson, “Permeation of non-interacting gases in polymeric materials – a


review of recent literature”, AERE Report G3023, 1984.

36. K Toi, K Takeuchi and T Tokuda, “Isotope effect in the diffusion of hydrogen and deuterium
in polymers”, J. Polym. Sci. Polymer Physics Ed., 18, 189-198, 1980.

37. R H Steinmeyer and J D Braun, “Hydrogen isotope permeation in elastomer materials”,


176-186, in Proceedings “Radiation Effects and Tritium Technology in Fusion Reactors”,
USERDA, 1976.

38. R S Barton, AERE Report M599, 1960.

39. C J Orme, J R Klaehn and F F Stewart, “Gas permeability and ideal selectivity of poly
[bis-(phenoxy)phosphazene], poly [bis-(4-tert-butylphenoxy)phosphazene], and poly
[bis-(3,5-di-tert-butylphenoxy)(1.2)(chloro)(0.8)phosphazene]”, Polymer, 44, 6773-6780,
2003.

40. N P Patel, C M Aberg, A M Sanchez, M D Capracotta, J D Martin and R J Spontak,


“Morphological, mechanical and gas-transport characteristics of cross-linked
poly(propylene glycol): homopolymers, nanocomposites and blends”, Polymer, 45,
5941-5950, 2004.

41. D P Dworak, H Lin, B D Freeman and M D Soucek, “Gas permeability analysis of


photo-cured cyclohexyl-substituted polysiloxane films”, J. Applied Polymer Sci., 102,
2343-2351, 2006.

42. C Damian, M Escoubes and E Espuche, “Gas and water transport properties of epoxy-
amine networks: Influence of crosslink density”, J. Applied Polymer Sci., 80, 2058-2066,
2001.

43. T Hirose, K Mizoguchi, K Terada, “Gas transport in poly(vinyl-p-isopropylbenzoate):


comparison and correlation with some other poly(vinyl esters)”, J. Applied Polymer Sci., 58,
1031-1040, 1995.

44. M Salame, “Transport phenomena through polymer films”, J. Polymer Sci., Polymer
Symposia No. 41, 1, Interscience, 1973.

45. S M Allen, M Fujii, V Stannett, H B Hopfenberg and J L Williams, “The barrier properties of
polyacrylonitrile”, J. Membrane Sci., 2, 153, 1977.

46. H B Hopfenberg, “Permeability of plastic films and coatings to gases, vapours and liquids”,
Polymer Science and Technology, 6, Plenum Press, London, 1974.

47. E A McGonigle, J J Liggat, R A Pethrick, S D Jenkins, J H Daly and D Hayward,


“Permeability of N-2, Ar, He, O-2 and CO2 through biaxially oriented polyester films –
dependence on free volume”, Polymer, 42, 2413-2426, 2001.

48. L Abis, G Floridi, E Merlo, R Po and C Zannoni, “Investigation on the dynamics of aromatic
polyesters by means of high resolution solid state CPMAS 13C NMR”, J. Polymer Sci.
Part B: Polymer Physics, 36, 1557-1566, 1998.

49. J Espeso, A E Lozano, J G de la Campa and J de Abajo, “Effect of substituents on the


permeation properties of polyamide membranes”, J. Membrane Sci., 280, 659-665, 2006.

50. B Flaconneche, J Martin and M H Klopffer, “Permeability, diffusion and solubility of gases
in polyethylene, polyamide 11 and poly(vinylidene fluoride)”, Oil & Gas Science and
Technology – Revue de l’IFP, 56, 261-278, 2001.

SERCO/TAS/000500/001 – Issue 2 Page 48 of 50


SERCO/TAS/000500/001 – Issue 2

51. Y. Kang, K. Araki, K. Iwamoto and M. Seno, “Preparation and gas permeability of polymer
blend membranes of polystyrene and poly[1,1,1-tris (trimethylsiloxy) methacrylate
propylsilane]”, J. Applied Polymer Sci., 27, 2025-2032, 1982.

52. L A Pessan, WJ Koros, J C Schmidhauser and W D Richards, “Gas-transport properties of


polymers based on spirobiindane bisphenol”, J. Polymer Sci. Part B: Polymer Physics, 33,
487-494, 1995.

53. M R Pixton and D R Paul, “Gas-transport properties of polyarylates based on


9,9-Bis(4-Hydroxyphenyl) Anthrone”, Polymer, 36, 2745-2751, 1995.

54. P Tremblay, M M Savard, J Vermette and R Paquin, “Gas permeability, diffusivity and
solubility of nitrogen, helium, methane, carbon dioxide and formaldehyde in dense
polymeric membranes using a new on-line permeation apparatus”, J. Membrane Sci., 282,
245-256, 2006.

55. A Tanioka, A Oobayashi, Y Kageyama, K Miyasaka and K Ishikawa, “Effects of carbon filler
on sorption and diffusion of gases through rubbery materials”, J. Polym. Sci. Polymer
Physics Ed., 20, 2197-2208, 1982.

56. T Ogasawara, Y Ishida, T Ishikawa, T Aoki and T Ogura, “Helium gas permeability of
montmorillonite/epoxy nanocomposites”, Composites Part A: Applied Science and
Manufacturing, 37, 2236-2240, 2006.

57. H G Hammon, K Ernst and J C Newton, “Noble gas permeability of polymer films and
coatings”, J. Applied Polymer Sci., 21, 1989-1997, 1977.

58. P Tiemblo, E Saiz, J Guzman and E Riande, “Comparison of simulated and experimental
transport of gases in commercial poly(vinyl chloride)”, Macromolecules, 35, 4167-4174,
2002.

59. R W Macdonald and R Y M Huang, “Permeation of gases through modified polymer films;
permeation and diffusion of helium, nitrogen, methane, ethane, and propane through
gamma-ray crosslinked polyethylene”, J. Applied Polymer Sci., 26, 2239-2263, 1981.

60. R B Carr, P E Coyle, J M Barnett, R C Berlo, C S McCaleb and J K Prono, “Lawrence


Livermore Laboratory Energy and Technology Review”, Lawrence Livermore
Laboratory Report UCRL-52000-75-5, 1975.

61. W J Koros, A H Chan and D R Paul, "Sorption and transport of various gases in
polycarbonate", J. Membrane Sci., 2, 165-190, 1977.

62. L F Epstein, T C Hall Jr. and S E Mills, “Preliminary gas permeation measurements on
plastics for use as concrete containment vessel liners”, Nucl. Eng. and Design, 8, 345-359,
1968.

63. G Espinosa, J I Golzarri and R B Gammage, “Comparative studies of polymer materials as


radon protection coatings”, Nucl. Trcks Radiat. Meas., 22, 329-330, 1993.

64. C Damian, E Espuche, M Escoubes, “Influence of three ageing types (thermal oxidation,
radiochemical and hydrolytic ageing) on the structure and gas transport properties of
epoxy-amine networks”, Polymer Degradation and Stability, 72, 447-458, 2001.

65. S Pauly, “The radiation-resistance of thermoset plastics. 3. Oxygen permeation


experiments”, Radiation Physics and Chemistry, 39, 269-272, 1992.

SERCO/TAS/000500/001 – Issue 2 Page 49 of 50


SERCO/TAS/000500/001 – Issue 2

66. R B Bird, W E Stewatr and E N Lightfoot, “Transport phenomena”, John Wiley and Sons,
1960.

67. A E Fick, “Ueber diffusion”, Ann. Physik und Chemie, 94, 59, 1855.

68. F P Incropera and D P De Witt, “Fundamentals of heat and mass transfer”, Third Edition,
John Wiley and Sons, 1981.

69. B C Duncan, J Urquhart and S J Roberts, “Review of measurement and modelling of


permeation and diffusion in polymers”, NPL Report DEPC-MPR 012, 2005.

70. M S Hedenqvist, M Ohrlander, R Plamgren and A-C Albertsson, "Multi-layer modeling of


diffusion of water in acrylamide-grafted aliphatic polyesters", Polymer Engineering and
Science, 1998.

71. R M Barrer “Diffusion and permeation in heterogeneous media”, 165-217, in “Diffusion in


Polymers”, ed. J Crank and G S Park, Academic Press, 1968.

72. J Crank, “The mathematics of diffusion”, Second Edition, Clarendon Press, 1975.

73. T Barth and M Ohlberger, “Finite volume methods: foundation and analysis”,
Encyclopaedia of Computational Mechanics, John Wiley and Sons, 2004.

74. T R Hsu, “The finite element method in thermomechanics”, Allen & Unwin Inc., 1986.

SERCO/TAS/000500/001 – Issue 2 Page 50 of 50

You might also like