Determination of G-Values For Use in SMOGG Gas Generation Calculations

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 77

Determination of G-values for use in

SMOGG gas generation calculations

To: Radioactive Waste Management Limited


Date: 16 July 2015
From: Amec Foster Wheeler
Your Reference: RWMD/03/037
Our Reference: AMEC/200615/001 Issue 3
Title Determination of G-values for use in SMOGG gas generation calculations

Prepared for Radioactive Waste Management Limited

Your Reference RWMD/03/037

Our Reference AMEC/200615/001 Issue 3

Note that Issue 1 was externally reviewed, Issue 2 was produced in response to
those review comments, and Issue 3 is in response to final comments from RWM.

Confidentiality, This report is submitted by Energy, Safety and Risk Consultants (UK) Limited
copyright & (hereafter referred to as AMEC) in connection with a contract to supply goods and/or
reproduction services.

To minimise our impact on the environment, AMEC uses paper from sustainable
sources

Contact Details Amec Foster Wheeler


Building 150
Harwell Campus
Didcot
Oxon OX11 0QB
United Kingdom

Tel +44 (0)1635 280300

amecfw.com

Name Signature Date


Author(s) Fiona Hunter
Angela Adeogun (Pöyry Energy)
John Dawson 9 July 2015
Jay LaVerne (University of Notre Dame, USA)
Sarah Watson (Quintessa)

Reviewed by
Ben Swift 16 July 2015

Approved by

David Holton 16 July 2015

Transport Flask Photograph courtesy of Magnox Electric Ltd


Submarine Photograph by: Mez Merrill; © Crown Copyright/MOD, image from www.photos.mod.uk. Reproduced with the permission of the Controller of
Her Majesty’s Stationery Office

AMEC/200615/001 Issue 3 Page 2 of 77


Preface
This report has been prepared by AMEC under contract to the Nuclear Decommissioning Authority
(NDA), Radioactive Waste Management Directorate (RWMD) and forms part of an ongoing
programme of research commissioned by NDA RWMD to underpin the long-term safety of a
geological disposal facility for higher-activity radioactive wastes. Before it was published RWMD
became a wholly owned subsidiary of the NDA (on 1 April 2014) called Radioactive Waste
Management Limited.

The report has been reviewed by Radioactive Waste Management Limited. However, references to
NDA and RWMD in the text have been retained as they are appropriate for the period when this
research was being performed. The views expressed and conclusions drawn in this report are
those of AMEC and do not necessarily represent those of the NDA RWMD or Radioactive Waste
Management Limited.

Conditions of publication
This report is made available under the Radioactive Waste Management Limited Transparency
Policy. In line with this policy, Radioactive Waste Management Limited 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
Nuclear Decommissioning Authority (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 2015. All rights reserved.


Bibliography
If you would like to see other reports available from Radioactive Waste Management Limited and
the NDA, a complete listing can be viewed at our website www.nda.gov.uk, or please write to our
Communications department at the address below.
Feedback
Readers are invited to provide feedback to Radioactive Waste Management Limited on the
contents, clarity and presentation of this report and on the means of improving the range of reports
published. Feedback should be addressed to:

Head of Stakeholder Engagement and Communications


Radioactive Waste Management Limited
Building 587
Curie Avenue
Harwell Oxford
Didcot
OX11 0RH
UK

Email: rwmdfeedback@nda.gov.uk

AMEC/200615/001 Issue 3 Page 3 of 77


AMEC/200615/001 Issue 3 Page 4 of 77
Executive Summary
Radiolysis, in particular of water and some organic materials, is a mechanism that needs to be
considered when determining the rates of gas generation from radioactive waste. It is important for
gas generation from Unshielded Intermediate Level Waste (UILW) and Shielded Intermediate Level
Waste (SILW) in a Geological Disposal Facility (GDF), but may also be important for some High
Level Waste (HLW) and Spent Fuel (SF).

RWMD has produced a series of reports that justify data values used as input to the gas generation
calculations undertaken using the Simple Model of Gas Generation (SMOGG) software. To add to
this series, the present study has been commissioned to review data required to calculate gas
generation by radiolysis in relation both to materials present in the UK Radioactive Waste Inventory
(UK RWI) and to the operational and post-closure evolution of a UK Geological Disposal Facility
(GDF). The relationship between the amount of gas, typically hydrogen gas, and the radiation dose
is usually formulated in terms of a G-value (also known as radiation chemical yield or yield). This
report provides an auditable trail for the basis of the G-values used in the SMOGG model to
calculate gas generated by radiolysis and has recommended some changes to the G-values, and
the categories of materials to which they are applied, that are used currently.

The report:
 reviews the current approach to gas generation by radiolysis, describes the use of G-values
and analyses the most recent gas generation results to understand the breakdown of
radiolysis by waste material type;
 analyses the UK Radioactive Waste Inventory for relevant materials that generate gas by
radiolysis to help determine the most sensible waste category groupings;
 critically reviews the available information on G-values using expert input for water,
cementitious materials and organic materials; and
 presents a summary of the findings and their uncertainties and data gaps, including tables of
best estimate and upper bound G-values for relevant waste materials.

A qualitative discussion is presented to understand the effect that using the recommended G-value
data and their material categories would have on gas generation rates used in the Environmental
Safety Case. A summary is provided below.
 Any water is expected to give similar gas generation rates for all types of radiation using best
estimate G-values, compared to the existing calculation of gas generation rates.
 The use of a single best estimate G-value for gas generation by radiolysis of cementitious
grout and its pore water, as opposed to applying a G-value for water for the fraction of water
contained within the cementitious grout, is expected to result in a higher gas generation rate
for all radiation types, compared to the existing calculation of gas generation rates.
 For cellulosic materials, the use of the proposed best estimate G-values for gas generation by
radiolysis, compared to the G-values used in the existing calculation of gas generation, would
be expected to increase the gas generation rate from gamma radiation, reduce the gas
generation from beta radiation, while alpha radiation gas generation rates would be expected
to be similar.
 For plastic wastes, separate G-values have been proposed for several categories of plastic
materials (non-halogenated plastics, halogenated plastics, ion exchange resins and rubbers)
and therefore should provide an improved description of the gas generation rates by radiolysis.
Overall the use of proposed best estimate G-values and the new categories of plastics is
expected to reduce the gas generation rate, compared to the G-values used in the existing
calculation of gas generation.
 The inclusion of polymer encapsulant is an additional category for which a G-value has been
assigned. The proposed G-values for gas generation by gamma and beta radiolysis are small,
and therefore this is expected to be a small addition to the overall gas generation rate.

AMEC/200615/001 Issue 3 Page 5 of 77


 Within the 2013 UK RWI, the category of ‘Other Organics’ accounts for a small proportion
(6 %) of the overall mass of organic wastes. As a result there is expected to be little change to
the overall gas generation rates by radiolysis of ‘Other Organics’.

Overall, the gas generation rates for the proposed G-value data for the proposed material
categories are expected to remain broadly similar, compared to the existing calculation of gas
generation rates. This is based on the expectation that:
 the current categories of ‘water’ and ‘polymers’ (plastic wastes) are the main contributors to
overall gas generation;
 the G-values for ‘water’ will be mainly replaced by the category of ‘cementitious grout and its
pore water’ with an expected increase in gas generation rates, although it is noted that there
may be water associated with the wastes that should continue to use the G-value for (pure)
water, not the G-value for cementitious grout and its encapsulant; and
 the category of ‘polymers’ will be replaced by several individual types of plastic wastes with
their associated general reduction in G-value.

AMEC/200615/001 Issue 3 Page 6 of 77


Contents
1 Introduction 9
1.1 Background 9
1.2 Scope / objectives of work 9

2 Review of current approach to calculation of gas


generation by radiolysis 10
2.1 Description of how G-values are used to calculate gas generation by radiolysis 10
2.2 Calculation of gas generation by radiolysis used by the SMOGG model 11
2.3 Analysis of gas generation assessment results that used the 2007 Derived
Inventory 12
2.4 Summary of key results from the analysis of calculations of gas generation by
radiolysis that used the 2007 Derived Inventory 19

3 Overview of UK radioactive waste inventory data for


relevant materials that generate gas by radiolysis 21
3.1 Water, as reported in the 2013 UK Radioactive Waste Inventory 21
3.2 Organic waste materials, as reported in the 2013 UK Radioactive Waste
Inventory 21
3.3 Discussion of the most sensible categories that are expected to contribute to
gas generation by radiolysis 25

4 Review of G-values for relevant materials 26


4.1 Background information for the review of G-values 26
4.2 Review of G-values for water 28
4.3 Review of G-values for cellulosic materials 30
4.4 Review of G-values for halogenated plastics 32
4.5 Review of G-values for non-halogenated (thermo)plastics 36
4.6 Review of G-values for rubbers 39
4.7 Review of G-values for ion exchange resins 41
4.8 Review of G-values for ‘Other Organics’ 44
4.9 Review of G-values for ‘small organic molecules’ 44
4.10 Review of G-values for encapsulation materials 45

5 Discussion 48
5.1 Recommended G-values 48
5.2 Data gaps and uncertainties in the review of G-values 55
5.3 Requirements for development of the SMOGG tool based on the outcomes from
the review 57
5.4 Demonstration of the effect that filling any knowledge gaps would have on gas
generation rates currently recorded in the Environmental Safety Case 58

6 Summary of the review with suggested G-value data for


use in future gas generation calculations 60

7 References 62
Appendix 1 SMOGG G-values: Treatment in the QPAC model 69
Appendix 2 Additional relevant UK RWI information 75

AMEC/200615/001 Issue 3 Page 7 of 77


AMEC/200615/001 Issue 3 Page 8 of 77
1 Introduction
1.1 Background
Radiolysis, in particular of water and some organic materials, is a mechanism that needs to be
considered when determining the rates of gas generation from radioactive waste. It is important for
gas generation from Unshielded Intermediate Level Waste (UILW) and Shielded Intermediate Level
Waste (SILW) in a Geological Disposal Facility (GDF), but may also be important for some High
Level Waste (HLW) and Spent Fuel (SF).

Assessment of gas generation during transport, storage and disposal of radioactive wastes is
currently carried out using the Simple Model of Gas Generation (SMOGG, version 5.1) software
tool [1, 2]. The tool considers a range of gas generation processes including radiolysis of water and
waste materials. NDA RWMD has produced a series of reports that justify data values used as
input to the gas generation calculations, e.g. [3, 4, 5, 6]. The present study has been commissioned
to review data required to calculate gas generation by radiolysis in relation both to materials
present in the UK Radioactive Waste Inventory (UK RWI) and to the operational and post-closure
evolution of a UK Geological Disposal Facility (GDF). The relationship between the amount of gas,
typically hydrogen gas, and the radiation dose is usually formulated in terms of a G-value (also
known as radiation chemical yield or yield). This report provides an auditable trail for the basis of
G-values that are recommended for use in SMOGG to calculate gas generated by radiolysis.

1.2 Scope / objectives of work


The work has the following objectives:
 On the basis of materials present in the UK RWI, to consider whether the current approach to
radiolytic gas generation in SMOGG, which addresses the radiolysis of water and four organic
compounds, is appropriate;
 On the basis of existing information, to document G-value data relevant to materials present in
the UK RWI, and to recommend G-values for use in SMOGG gas generation calculations;
 Justify whether point values or ranges of G-values most appropriately reflect experimental data
and understanding, and recommend an approach to be followed in any numerical modelling
undertaken as part of an update to the Environmental Safety Case (ESC) regarding G-values
for whole inventory use, or a number of G-values relating to specific waste streams / groups of
waste streams;
 Document any knowledge gaps identified via the above activity, and demonstrate the effect
that filling these knowledge gaps would have with reference to the effect on gas generation
rates presently recorded in the ESC.

The study provides an auditable documentation trail of recommended G-values for use in SMOGG
gas generation calculations.

Our approach has been to present a report that:


 reviews the current approach to gas generation by radiolysis and describes the use of
G-values, and analyses the most recent gas generation results to understand the breakdown
of radiolysis by waste material type;
 analyses the UK Radioactive Waste Inventory for relevant materials that generate gas by
radiolysis to help determine the most sensible waste category groupings;
 critically reviews the available information on G-values, for water, cementitious materials and
organic materials; and
 presents a summary of the findings and their uncertainties and data gaps, including tables of
best estimate and upper bound G-values for relevant waste materials.

AMEC/200615/001 Issue 3 Page 9 of 77


2 Review of current approach to
calculation of gas generation by
radiolysis
2.1 Description of how G-values are used to calculate gas
generation by radiolysis
This section provides an introduction to the current calculation of gas generation by radiolysis that
is applied to UK wastes. It describes how G-values are used to calculate gas generation for
media/materials such as water and organic wastes.

The rate of gas generation (or consumption) from a material depends on:
 the rate of energy emission from radioactive decay;
 the fraction of the emitted energy absorbed by the material; and
 the so-called G-value for gas generation (or consumption) for the given material and type of
radiation.

In particular:
q H = c ∑ Pi Fi Gi (1)
i
where
-1
qH is the rate of gas generation (mol s );
-7 -1 -1
c is a unit conversion constant (= 1.036 10 mol J molecule 100eV);
-1
Pi is the total power output from radioactive decay (J s ) from radiation type i;
Fi is the fraction of energy emitted of radiation type i that is absorbed by the material (–); and
Gi is the so-called G-value (also known as radiation chemical yield or yield), defined as the
number of molecules of gas generated (or consumed) per 100 eV of energy absorbed
-1 -7 -1
from radiation type i (molecule 100eV ; where 1 molecule/100 eV = 1.036 x 10 mol J ).

The factors that will influence the fraction of energy absorbed by a material (i.e. Fi in Equation (1))
include the distribution of radioactivity (in the wastes) relative to the material and the range of
effectiveness of the radiation in the material.

A number of factors will affect the G-value (i.e. Gi in Equation (1)), including:
 The specific composition of the material.
 The Linear Energy Transfer (LET) of the radiation. LET is the linear rate of energy loss by an
ionizing particle travelling through a material. Differences in G-values when a material is
irradiated by different types of radiation can be ascribed to differences in the ways in which
energy is lost (characterised, for example, by LET).
 The dose rate. G-values may exhibit a non-linear dependence on the dose rate.
 The total absorbed dose. As irradiation of a material proceeds, the primary products can
increase to such a concentration in the material that they themselves are irradiated, or react
with some of the free radicals, to form secondary products. If the products are more stable
under irradiation than the parent material, then the measured G-values may decrease as the
absorbed dose increases.
 The atmosphere in which the irradiation occurs. Measured G-values depend on the
atmosphere in which the irradiation occurs, especially whether or not any oxygen is present.
 The temperature. Chemical reaction rates depend on temperature.
 The pressure. An increased pressure can mean more of a gas remains dissolved in the
material being irradiated. This may (slightly) lower G-values as a result of so-called back
reactions.

AMEC/200615/001 Issue 3 Page 10 of 77


2.2 Calculation of gas generation by radiolysis used by the
SMOGG model
This section describes the calculation of gas generation by radiolysis used by the SMOGG
software [1, 2] as an introduction to the analysis of the gas generation results presented in the
following section (Section 2.3). Two types of radiolysis are considered: radiolysis of water and
radiolysis of organic materials, including resins, as follows:
 Radiolysis of water and aqueous solutions
Radiolysis of water produces hydrogen. If tritiated water molecules are present, a
corresponding proportion of the hydrogen released will also be tritiated. The amount of
hydrogen released depends on the chemical constituents of the aqueous phase, but it is not
intended to model this dependence (it can be allowed for by appropriate choice of radiolysis
factors for particular waste streams, taking account of the conditions in those packages).

Water within the waste package can be subjected to radiolysis from alpha, beta, and gamma
radiation. Water external to the package, for example in a backfilled vault or tunnel in a GDF
can also be radiolysed by gamma radiation escaping from a waste package. This is particularly
relevant to some HLW and perhaps SF, and may occur at a time post-closure when there is no
water in the SF or HLW canister to produce hydrogen by internal radiolysis.
 Radiolysis of organic materials
Radiolysis of organic compounds present in wastes can lead to the generation of a variety of
gases. For the purpose of the SMOGG model it is assumed that the bulk gas is hydrogen as
this is likely to be the main gas for many organic materials, although generation of one or
several other gases can be calculated by suitable choice of parameters. A few organic wastes,
notably resins and some small molecules, may contain carbon-14. The SMOGG model
considers four classes of organic material in its model of radiolysis. These classes are
‘cellulose’, ‘polymers’ (taken to include all non-cellulosic plastic substances), ‘oils’ and ‘small
organic molecules’. Note that for the ‘small organic molecules’ class, only the generation of
carbon-14 is calculated, not the bulk (inactive) gas since this can be included within the
radiolysis of cellulose, its precursor.

A simple model is provided in SMOGG for the production of hydrogen from water and some
organic materials by radiolysis. A standard simple model of gas production by radiolysis has been
used, in which the amount of gas produced depends linearly on the radiation energy absorbed by
the material subject to radiolysis (equation (1)).

The G-value depends on the material subjected to radiolysis, and varies with the type of radiation
with similar G-values for beta and gamma radiation often observed (or assumed), although the
values are typically different from those for alpha radiation. The rate of gas production depends on
the rate of energy absorption. The SMOGG model assumes that all of the alpha and beta radiation
produced in a waste package is absorbed within the package based on the distance that these
radiation types are expected to penetrate. The energy absorbed by a particular material in the
package, for example water, is calculated as the product of the total decay energy (for alpha and
beta radiation) produced in the package and the fraction of the wasteform mass consisting of the
material. For gamma radiation, a user specified fraction of the decay energy is assumed to escape
from the package. For the fraction of the decay energy absorbed within the package, the amount
absorbed by the water is calculated as for the other forms of radiation. The fraction that escapes
from the package is assumed to be capable of radiolysing water in the backfill around the package
post-closure, with the proportion of the energy absorbed by the water in the backfill calculated as
the ratio of the mass of water in the backfill associated with the package to the total mass of this
backfill. External radiolysis is assumed only to occur after backfilling and closure of a GDF.

Typically UK ILW/LLW wastes are placed within a vented container and therefore
over-pressurisation as a result of gas generation by radiolysis is not expected to be an issue. The
gas generation by radiolysis that takes place within sealed containers such as those expected for
HLW/SF is expected to be small and therefore over-pressurisation is not expected to be an issue
(see Section 2.3.2).

AMEC/200615/001 Issue 3 Page 11 of 77


In the SMOGG model, hydrogen is assumed to be the gas produced by radiolysis of ‘cellulose’,
‘polymers’ and ‘oils’. For ‘small organic molecules’, only the possibility that carbon-14 containing
gases may be released by radiolysis is considered by SMOGG. Therefore G-values are used by
SMOGG for the release of hydrogen from the first three organic material categories and, as it is
assumed that the carbon-14 containing gases released from ‘small organic molecules’ can occur
as methane and/or carbon dioxide, G-values for the release of both of these gases from ‘small
organic molecules’ are required. The release of carbon-14 from the radiolysis of ‘small organic
molecules’ is calculated on the basis of the proportion of carbon-14 to carbon-12 present in the
organic compound.

Gas generation calculations have also been carried out using the QPAC software [7]. Its approach
to the calculation of radiolysis is similar to the approach applied in SMOGG, with the exception that
for the fraction of the emitted decay energy (Fi) absorbed by a material an alternative approach to
the scaling factor is applied for the radiolysis of water. There is significant uncertainty concerning
the scaling factor that should be used to describe the efficiency of the radiolysis process for water,
especially for the backfill, since the waste package geometry will play a role. The approach taken
by QPAC is to apply a volume fraction rather than the mass fraction applied by SMOGG for
radiolysis of water. A volume fraction is applied in QPAC to make the numerical computation easier
for processes such as variable saturation and creep within an evaporite host rock. The volume
fraction approach taken by QPAC is considered because of the significant differences in the
density of the materials being considered (waste / backfill and water). The difference in results for
radiolysis of water between the approaches taken by SMOGG (mass fraction approach) and QPAC
(volume fraction approach) can be taken as a measure of the uncertainty. Full details of the QPAC
approach are provided in [7] with a summary note of the QPAC approach provided in Appendix 1.

2.3 Analysis of gas generation assessment results that used


the 2007 Derived Inventory
This section considers the proportion of gas generated by radiolysis within a UK geological
disposal facility (GDF) to assess its importance within the overall assessment of gas generation. It
is necessary to understand whether the generation of gas by radiolysis is a significant proportion of
the total gas generation and if so, over which timescale it is significant. For example, a highly
detailed analysis of the types of materials and their G-values that contribute to gas generation by
radiolysis is unlikely to be justified if the rate of gas generation (bulk and radioactive) is very low or
the overall gas generation rate is dominated by other mechanisms.

The results of gas generation calculations that used the 2007 Derived Inventory and applied the
SMOGG model [8] have been analysed as part of this work to understand the breakdown by
radiolytic and non-radiolytic processes. This analysis highlights the contribution of water, and each
of the four organic waste groups currently used in the SMOGG software, to the calculation of gas
generation by radiolysis over both pre- and post-closure timescales. The aim of this analysis is to
identify the most significant contributions to gas generation by radiolysis to provide understanding
of which materials are the most significant sources of radiolysis over different time periods as well
as to understand the overall significance of gas generation by radiolysis.

2.3.1 Examination of G-values applied in the gas generation results that used the
2007 Derived Inventory
The G-values applied in the gas generation results that used the 2007 Derived Inventory are
provided in Table 1 (from [8]). The basis for these data is given in Hoch et al., [9] for hydrogen
production from radiolysis for water, ‘cellulose’, ‘polymers’ and ‘oils’. In short, Hoch et al., [9] based
their chosen G-values on available literature and in some cases, notably for beta radiation, G-value
data were not available and therefore the G-value for gamma radiation for the appropriate material
was used. However Hoch et al., [9] explained that G-values for CO2 and CH4 production by
radiolysis of ‘small organic molecules’ were not identified, and therefore G-values for acetic acid
were applied instead.

AMEC/200615/001 Issue 3 Page 12 of 77


Table 1 Upper bound G-values used in the gas generation calculations that used the
2007 Derived Inventory for the model of gas production from radiolysis in a GDF (from [8]).
G-value (molecules/100eV)
Material Gas
α β γ
Water H2 1.6 0.5 0.5
‘Cellulose’ H2 1.4 6.0 3.2
‘Polymers’ H2 2.7 6.0 6.0
‘Oils’ H2 5.3 5.6 4.6
‘Small organic molecules’ CO2 4.0 5.4 5.4
‘Small organic molecules’ CH4 1.4 3.9 3.9

Examination of the G-values applied in the gas generation calculations that used the 2007 Derived
Inventory [8] suggests that the large contribution to gas generation by radiolysis of ‘polymers’ and
‘oils’ may, in part, relate to the relatively high G-values applied to these organic waste categories
(see Section 2.3.2). Therefore a good understanding of the detailed nature of the organic waste
categories and their associated range of G-values is required and has been considered in
Section 3.2.

2.3.2 Analysis of the gas generation results that used the 2007 Derived Inventory
to understand the breakdown by radiolytic and non-radiolytic processes
The results of the gas generation calculations that used the 2007 Derived Inventory [8] have been
analysed to understand the breakdown by radiolytic and non-radiolytic processes. The analysis of
gas generation by radiolysis has been carried out for the results from unshielded intermediate level
waste (UILW), shielded intermediate level waste (SILW) / low level waste (LLW) and depleted,
natural and low enriched uranium (DNLEU) wastes. The gas generation calculations that used the
2007 Derived Inventory are based on a higher strength rock for a UK GDF and different gas
generation rates from some processes may be expected for other rock types; full details of the
modelling assumptions are provided in reference [8].

Consideration has also been given to the generation of gas by radiolysis from high level waste
(HLW) / Spent Fuel (SF) and highly enriched uranium (HEU) / plutonium (Pu), based on the
discussion provided in reference [8]. HLW / SF are expected to have only very low amounts of
water present within the (copper) disposal canisters, and the disposal canisters are likely to remain
5
intact for at least 10 years. In addition, the thickness of the disposal canisters is expected to
reduce any radiolysis of backfill and its pore water within the tunnels and vaults of a GDF. As a
result, it has been assumed that HLW / SF will generate little or no gas by radiolysis over the gas
assessment period. By analogy, based on the assumption that HEU / Pu will be packaged in a
similar manner, in the presence of little water, it has also been assumed that HEU / Pu will
generate little or no gas by radiolysis over the whole of the gas assessment period.

The analysis of the gas generation calculations that used the 2007 Derived Inventory to understand
the breakdown by radiolytic and non-radiolytic processes are shown in Figure 1, Figure 2, Figure 3
and Figure 4 for UILW, SILW/LLW and DNLEU. These figures show the total bulk hydrogen gas
generation rates (from metal corrosion and radiolysis) and the breakdown between the different
radiolysis processes that are expected to contribute to gas generation from UILW, SILW/LLW and
DNLEU. The total hydrogen gas generation rate is included for comparison with the breakdown of
gas generation by radiolysis processes from individual materials.

AMEC/200615/001 Issue 3 Page 13 of 77


1E+7

1E+6
Bulk gas generation rate (m3a-1)

1E+5

1E+4

1E+3

1E+2

1E+1

1E+0

1E-1

1E-2
2040 2090 2140 2190 2240
Year (AD)

1E+7

1E+6
Bulk gas generation rate (m3a-1)

1E+5

1E+4

1E+3

1E+2

1E+1

1E+0

1E-1

1E-2
1000 10000 100000
Year (AD)

Total hydrogen f rom all processes


Hydrogen f rom radiolysis of water
Hydrogen f rom radiolysis of cellulose
Hydrogen f rom radiolysis of polymers
Hydrogen f rom radiolysis of oils

Figure 1 Summary of the bulk gas generation by radiolysis and by all processes for
UILW based on the gas generation results [8] that used the 2007 Derived Inventory.

AMEC/200615/001 Issue 3 Page 14 of 77


1E+2

1E+1
Active gas generation rate (TBq a-1)

1E+0

1E-1

1E-2

1E-3

1E-4

1E-5

1E-6

1E-7
2040 2090 2140 2190 2240
Year (AD)

1E+2

1E+1
Active gas generation rate (TBq a-1)

1E+0

1E-1

1E-2

1E-3

1E-4

1E-5

1E-6

1E-7
1000 10000 100000
Year (AD)

H-3 as H2 by radiolysis of water


H-3 as H2 f rom all processes
C-14 as CH4 by radiolysis of small organic molecules
C-14 as CH4 f rom all processes

Figure 2 Summary of the active gas generation by radiolysis and by all processes for
UILW based on the gas generation results [8] that used the 2007 Derived Inventory.

AMEC/200615/001 Issue 3 Page 15 of 77


1E+7

1E+6
Bulk gas generation rate (m3a-1)

1E+5

1E+4

1E+3

1E+2

1E+1

1E+0

1E-1

1E-2
2040 2090 2140 2190 2240
Year (AD)

1E+7
Bulk gas generation rate (m3a-1)

1E+6

1E+5

1E+4

1E+3

1E+2

1E+1

1E+0

1E-1

1E-2
1000 10000 100000
Year (AD)

Total hydrogen f rom all processes


Hydrogen f rom radiolysis of water
Hydrogen f rom radiolysis of polymers

Figure 3 Summary of the bulk gas generation by radiolysis and by all processes for
SILW / LLW based on the gas generation results [8] that used the 2007 Derived Inventory.

AMEC/200615/001 Issue 3 Page 16 of 77


1E+7

1E+6
Bulk gas generation rate (m3a-1)

1E+5

1E+4

1E+3

1E+2

1E+1

1E+0

1E-1

1E-2
2040 2090 2140 2190 2240
Year (AD)

1E+7

1E+6
Bulk gas generation rate (m3a-1)

1E+5

1E+4

1E+3

1E+2

1E+1

1E+0

1E-1

1E-2
1000 10000 100000
Year (AD)

Total hydrogen f rom all processes


Hydrogen f rom radiolysis of water

Figure 4 Summary of the bulk gas generation by radiolysis and by all processes for
DNLEU based on the gas generation results [8] that used the 2007 Derived Inventory.

AMEC/200615/001 Issue 3 Page 17 of 77


For UILW and SILW/LLW, in addition to bulk hydrogen gas generation, it is recognised that bulk
methane gas is generated from organic degradation of cellulose (i.e. a non-radiolytic process). Bulk
methane gas has not been included in Figure 1 or Figure 3, but it is recognised that this would
account for a proportion of the total bulk gas generated from shortly after closure of the GDF
(at 2150 AD) until the cellulose has fully degraded. For example, for UILW bulk methane accounts
for around 40 % of the calculated total bulk gas generated between 2174 AD and ~2500 AD, and
this drops to ~1 % by 4300 AD. Bulk methane has not been taken into account when comparing
gas generation rates by radiolysis of individual wastes relative to the total bulk hydrogen gas
generation rates, and therefore the relative proportions of gas generation rates by radiolysis of
individual wastes that are discussed in this section are greater than if the bulk methane gas
generation were included in the total bulk gas generation rate.

Note that in the gas generation calculations that used the 2007 Derived Inventory [8], the material
categories available in SMOGG were applied. Therefore the mass of ‘polymers’ was taken to be
the sum of the masses of halogenated plastics, non-halogenated plastics, organic ion exchange
resins, and rubbers, the mass of ‘oils’ was taken to be that of ‘Other Organics’ (as described by the
2007 Derived Inventory), and the mass of ‘cellulose’ was taken to be wood, paper and cotton.

For UILW, the analysis of the calculations for their contribution to gas generation by radiolysis
shows that:
 Bulk hydrogen gas generated by radiolysis of water and radiolysis of ‘polymers’ provides a
significant contribution to the overall bulk hydrogen gas generation in UILW (Figure 1). For
example, the maximum fraction of hydrogen gas generated by all radiolysis processes
compared to the total bulk hydrogen gas generation is calculated to be around 74 % at
3 -1 3 -1
2174 AD (~54 % (588 m a ) from radiolysis of water, and ~18 % (201 m a ) from radiolysis
of ‘polymers’). This occurs after the Magnox wastes and then the aluminium wastes, which
generate gas rapidly, become fully corroded leaving the hydrogen gas generation rate to be
dominated by radiolysis of water, with some contribution from the corrosion of stainless steel
containers and radiolysis of ‘polymers’.
 For UILW, a maximum rate of bulk hydrogen gas generation by radiolysis of around
3 -1
1150 m a is calculated to be generated at 2090 AD (9 % of the total bulk hydrogen gas from
UILW, Figure 1). This is expected during the operational phase, when the last of the packages
are emplaced.
 Over time, there is a decay of the total radioactive inventory in UILW, which leads to a similar
reduction in the rate of bulk hydrogen gas generation by radiolysis (Figure 1). By contrast, the
contribution to the hydrogen gas generation rate by corrosion of stainless steel containers
6
remains relatively constant. Therefore by around 10 years, the fraction of hydrogen gas
generated by all radiolysis processes is calculated to be around 5 % compared to the total bulk
hydrogen gas generation (~4 % from radiolysis of water, and ~1 % from radiolysis of
‘polymers’).
 For C-14 as methane gas from UILW, the generation rate by radiolysis (from small organic
molecules) is calculated to provide only a small contribution of the total C-14 as methane gas
from all processes (Figure 2). For example, a maximum of around 1 % of C-14 as methane
compared to the overall C-14 as methane from all processes is calculated to be observed at
2161 AD. (It is noted that generation of C-14 as methane gas from radiolysis is attributed
waste stream 1A07 (GE healthcare wastes) and is no longer expected to be disposed of in a
GDF.)
 For H-3 as hydrogen gas from UILW, the gas generation rate by radiolysis of water is
calculated to provide a small fraction of the total H-3 as hydrogen gas from all processes
(Figure 2). For example, around 2 % contribution by radiolysis at 2160 AD may be calculated,
although this does not include the solid state diffusion of tritium contribution to the total H-3 as
hydrogen gas (see reference [8] for full details) and hence the contribution due to pure
radiolysis is likely to be smaller than this. (It is noted that by around 2460 AD, a maximum
contribution of H-3 as hydrogen gas (around 30 %) is due to radiolysis of water, although given
the very small amounts of H-3 as hydrogen gas that are calculated to be produced (around
-13 -1
5×10 TBq a ) this is considered insignificant.)

AMEC/200615/001 Issue 3 Page 18 of 77


For SILW / LLW, it is observed that:
 Hydrogen gas generated by radiolysis of water and radiolysis of ‘polymers’ provides a small
contribution to the overall hydrogen gas generation in SILW/LLW (Figure 3). For example, the
maximum fraction of hydrogen gas generated by all radiolysis processes compared to the total
bulk hydrogen gas generation is calculated to be around 0.5 % of the total bulk hydrogen gas
3 -1
from SILW / LLW at 2048 AD (~0.5 m a ). This is mainly due to radiolysis of water, with a
slight contribution from radiolysis of ‘polymers’. Generation of hydrogen by radiolysis of
cellulose and oils is of little significance.
 There are no active gases (C-14 and H-3) generated by radiolysis in SILW / LLW due to the
absence of organic materials that generate C-14 by radiolysis within these waste streams, and
because H-3 as hydrogen gas generated by radiolysis of water is calculated to be insignificant.
 The rate of bulk hydrogen gas calculated to be generated by radiolysis for SILW/LLW
3 -1 3 -1
(maximum of ~0.5 m a ) is low relative to UILW (maximum of 1150 m a ).

For DNLEU, it is observed that:


3 -1
 Hydrogen gas generated by radiolysis of water provides 100 % contribution (up to 14 m a ) of
the overall hydrogen gas generation in DNLEU until 2155 AD when the vaults are fully
resaturated and become anaerobic (Figure 4). After 2155 AD, the anaerobic corrosion of
stainless steel dominates the hydrogen gas generation rate until the final calculation time, with
a smaller, increasing contribution from radiolysis of water. For DNLEU the maximum rate of
3 -1
bulk hydrogen gas by radiolysis is around 40 m a at 100 000 AD (32 % of the total bulk
hydrogen gas generation from DNLEU).
 There are no active gases (C-14 and H-3) generated by radiolysis in DNLEU for similar
reasons as for SILW/LLW (above).

Overall, gas generation from UILW provides the largest bulk hydrogen gas generation rate [8]. Of
this, the gas generation by radiolysis of UILW forms an important contribution.

A summary of the key results from the analysis of calculations of gas generation by radiolysis that
used the 2007 Derived Inventory is presented in the next section.

2.4 Summary of key results from the analysis of calculations


of gas generation by radiolysis that used the
2007 Derived Inventory
Consideration of the calculated bulk and active gases generated by radiolysis, in combination with
the G-values as applied in the gas generation calculations that used the 2007 Derived
Inventory [8], highlights the following points:
 Water and ‘polymers’ are the key media/materials that contribute to bulk gas generation by
radiolysis for UILW, based on the analysis of the gas generation results that used the
2007 Derived Inventory [8] (Section 2.3.2).
3 -1
 Radiolysis of water is calculated to contribute around half (588 m a ) of the total bulk
hydrogen gas generation at 2174 AD for UILW (Section 2.3.2). This occurs after full
resaturation of the GDF, and therefore the mass of water present is based on the assumed
porosity.
3 -1
 Radiolysis of water is calculated to contribute all (14 m a ) of the bulk hydrogen gas
generation rate until 2155 AD for DNLEU, although the amount of bulk hydrogen gas from
3 -1
radiolysis of water is calculated to increase to 40 m a by 100 000 AD (Section 2.3.2).
 The rate of bulk hydrogen gas calculated to be generated by radiolysis for SILW/LLW
3 -1 3 -1
(maximum of ~0.5 m a ) and for DNLEU (maximum of ~40 m a ) is low relative to UILW
3 -1
(maximum of 1150 m a ) (Section 2.3.2).
 HLW/SF and HEU/Pu are expected to generate little or no bulk hydrogen gas by radiolysis due
to the expected absence of water within each container. In addition, only a small amount of

AMEC/200615/001 Issue 3 Page 19 of 77


radiolysis is expected to occur outside the container within the backfill and its pore water due
to the thickness of the disposal canisters being considered (Section 2.3.2).
 The G-values for water (Table 1) that were applied in the gas generation calculations that used
the 2007 Derived Inventory, although provided as an upper bound, are lower than those
applied for the organic materials (Section 2.3.1). An investigation of published G-values for
water is presented in Section 4.2.
 Radiolysis of ‘polymers’ is expected to provide a significant amount of the bulk hydrogen gas
generation rate in UILW based on the analysis of the gas generation results that used the
2007 Derived Inventory (Section 2.3.2). However, the G-values for ‘polymers’ applied in the
gas generation calculations that used the 2007 Derived Inventory (Table 1, [8]) are higher than
those applied for water, for example, which will have a bearing on the relative contribution of
radiolysis of ‘polymers’ to the overall bulk hydrogen gas generation rate. An investigation of
published G-values for ‘polymers’ is presented in Section 4.
 Radiolysis of ‘cellulose’ and ‘oils’ are expected to contribute lower bulk hydrogen gas
generation rates relative to radiolysis of water and polymers based on the analysis of the gas
generation results that used the 2007 Derived Inventory (Section 2.3.2). However, the
G-values for ‘cellulose’ and ‘oils’ are similar to those for ‘polymers’ (Table 1). An investigation
of published G-values for ‘cellulose’ and ‘oils’ is presented in Section 4.
 The generation of C-14 as CH4 by radiolysis forms only a small fraction of the overall C-14 as
CH4, and is only relevant for UILW (Section 2.3.2). Equally, the gas generation calculations
that used the 2007 Derived Inventory apply the carbonation reaction, and therefore no C-14 as
CO2 is released into the GDF.
 The generation of H-3 as H2 by radiolysis is of low significance, and is only relevant for UILW
(Section 2.3.2).

Based on this summary, it is clear that the amounts of water and ‘polymers’ present, and the
composition of the organic materials assigned as ‘polymers’, have a significant bearing on the
calculated gas generation rates. Therefore the UK Radioactive Waste Inventory has been
interrogated in detail to provide further understanding on the amount of water, and the amount and
nature of the organic materials (Section 3). Equally, the assigned G-values for water and ‘polymers’
have a significant bearing on the calculated gas generation rates. G-values for water and relevant
materials are critically reviewed in Section 4.

AMEC/200615/001 Issue 3 Page 20 of 77


3 Overview of UK radioactive waste
inventory data for relevant materials that
generate gas by radiolysis
The amount of water and the amount and composition of ‘polymers’ has a significant bearing on
the calculated gas generation rates by radiolysis as shown by the analysis of the gas generation
results that used the 2007 Derived Inventory (Section 2). The following subsections describe the
information available for water and organic materials in the 2013 UK Radioactive Waste
Inventory [10].

3.1 Water, as reported in the 2013 UK Radioactive Waste


Inventory
The amount of water available for gas generation by radiolysis is (i) initially dependent on the
amount of water present within the conditioned wastes and, (ii) also dependent on water present
after resaturation of a GDF when it is assumed that the pore spaces are completely filled with
water. Large void spaces associated with wastes may be expected to be filled with grout during the
packaging process, although small void spaces may remain unfilled until the GDF becomes
saturated with water. Although there is some limited descriptive information in the UK RWI of the
amount of water within individual wastes, this may not represent the amount of water at the time of
emplacement of the wastes. This is because, for example, the UK RWI describes the bulk density
of the wastes and their associated water which may not necessarily take into account any
conditioning with cementitious grout, or any compression of the wastes to fit into a smaller volume.
For information, an estimation of the pore space associated with the wasteform grout has been
carried out in reference [11].

Using this reasoning, it is clear that the amount of water available to generate gas by radiolysis is
mainly based on the assumption of fully saturated void space after resaturation of a GDF, and
therefore the uncertainty in the amount of water available for gas generation by radiolysis is
dependent on the design of the GDF, as well as the packaging of wastes. It is noted that full
resaturation of the GDF is reasonable for a fractured, hard host rock where the resaturation rate is
not limited by the supply of water from the host rock. However in the case of lower strength
sedimentary rocks (e.g. clays), the rate of resaturation may be limited by the supply of water from
the host rock to a GDF.

The 2013 UK RWI provides limited information about the aqueous content of each waste stream.
Therefore, Table A2 in Appendix 2 provides a list of the waste streams that are presented in
Table 2 with the aqueous content (% weight and total mass) for the whole waste stream as
reported in the 2013 UK RWI. However, as explained above, this information does not provide the
complete picture to understand how much water is available for gas generation by radiolysis after
waste packaging, emplacement of the wastes and closure of a GDF, and therefore this information
is presented as background data only.

3.2 Organic waste materials, as reported in the 2013 UK


Radioactive Waste Inventory
In order to assess the amount of gas generation by radiolysis of organic materials, this section
describes the amount of organic materials that are present in the 2013 UK Radioactive Waste
Inventory (RWI) and analyses the breakdown by individual material.

AMEC/200615/001 Issue 3 Page 21 of 77


The total mass of organic materials reported in the 2013 UK RWI for wastes expected to be
1
present within a geological disposal facility (GDF) is 13,739 tonnes. Organic materials are
reported for the following categories:
 cellulose,
 halogenated plastic,
 non-halogenated plastic,
 rubber (halogenated and non-halogenated),
 ion exchange materials, and
 other organics.

For the majority of waste streams in the 2013 UK RWI a further description of the components of
each of the categories is also given. This information has been used to provide a breakdown for
each category.

Table 2 presents the total masses for each category and components. The waste streams
contributing more than 5 % to the total masses are given with the mass of material in brackets. The
material masses are the sum of reported values in the 2013 UK RWI to the nearest tonne.

Table 2 Organic materials reported in the 2013 UK RWI for wastes expected to be
present within a GDF
Material Name Material Mass Waste Streams Contributing > 5% of the
(tonnes) Material Mass
Cellulose

Paper and cotton 1,336 2D90 (608t), 2D03 (188t), 5B25 (151t), 7A111 (77t),
(amorphous cellulose) 2D83/C (71t)

Wood (crystalline) 1,188 2D06 (614t), 2D90 (405t), 2D116 (92t)

Unspecified 295 2D25 (109t), 2D08 (59t)

Total Cellulosics 2,819 2D90 (1,013t), 2D06 (614t), 2D03 (188t), 5B25 (154t)
Halogenated Plastics
1
PVC 4,401 2D90 (1,216t), 2D03 (799t), 2D06 (614t), 2D116 (413t),
2D83/C (249t), 5B02 (242t), 7A21 (231t)

PVC/PTFE* 94 5C317 (44t), 5C08 (38t), 5C30 (10t)

Unspecified 464 7A111 (318t), 7A117 (74t), 2D39 (51t)

Total Halogenated Plastics 4,960 2D90 (1,216t), 2D03 (799t), 2D06 (614t), 2D116 (413t),
7A111 (318t), 2D83/C (249t), 5B02 (242t), 7A21 (231t)
Non-Halogenated Plastics

Condensation Polymers 956 2D06 (307t), 7A111 (152t), 9G18/C (128t), 2D116 (69t),
9G04/C (59t)

Other Non-Halogenated 1,713


Plastics 2D90 (810t), 2D06 (307t), 2D03 (188t), 5B25 (120t)

Not specified 110 Small contributions over many streams.

Total Non-Halogenated 2,780 2D90 (810t), 2D06 (614t), 2D03 (188t), 7A111 (152t),
Plastics 2D116 (138t), 9G18/C (128t)

1
The inventory for the GDF includes all ILW plus LLW graphite from Magnox and AGRs and LLW Resin from Plant
Decontamination (MODIX) (7D28).

AMEC/200615/001 Issue 3 Page 22 of 77


Material Name Material Mass Waste Streams Contributing > 5% of the
(tonnes) Material Mass
Rubbers

Halogenated 1,514 2D90 (810t), 2D03 (329t), 2D06 (114t), 2D83/C (107t),
7A111 (75t)

Non-Halogenated 209 2D06 (114t), 2F34 (35t), 2D09 (24t), 2F02 (17t)

Not specified 260 2D116 (184t), 2D130 (36t)

Total Rubber 1,982 2D90 (810t), 2D03 (329t), 2D06 (229t), 2D116 (184t),
2D83/C (107t)
Ion exchange resins
1
Styrene-divinylbenzene 163 3S12 (51t), 4B01 (34t), 9C36 (27t), 9G04/C (25t), 7D29
(9t)
1
Phenolformaldehyde 149 9G18/C (40t), 9C02 (22t), 9D28 (12t), 9D27 (10t), 9D29
(9t), 4C03 (9t)

Not specified 108 Small contributions over many streams

Total Ion exchange resins 419 3S12 (51t), 9G18/C (40t), 9C36 (27t), 9G04/C (25t),
4B01 (34t), 9C02 (22t)
Other Organics

Total Other Organics 780 2D90 (203t), 2D27/C (174t), 2D03 (141t), 5B25 (107t),
2D83/C (71t)
Total Organic Materials 13,739
* The mass of material has been determined by reviewing textual descriptions provided for the
2013 UK RWI.

A description of the wastes in each of the categories is given below.

3.2.1 Cellulose
The total mass of cellulose is 2,819 tonnes. The cellulose content is either reported as paper and
cotton (1,336 tonnes), wood (1,188 tonnes) or not specified (295 tonnes).

The major waste stream contributors to the mass of paper and cotton are plutonium contaminated
material (PCM) arising at Sellafield (2D03, 2D90 and 2D83/C) and AWE (7A111), and retrievals
from the ILW shaft at Dounreay (5B25).

The major waste stream contributors to the mass of wood are PCM at Sellafield (2D06 and 2D90)
and decommissioning plant wastes (Processing Plants, Tanks, Silos etc.) (2D116) at Sellafield.

3.2.2 Halogenated plastics


The total mass of halogenated plastics reported is 4,960 tonnes, which is mainly polyvinyl chloride
(PVC). The halogenated plastic content is either reported as PVC (4,401 tonnes); a mixture of PVC
and polytetrafluroethylene (PTFE) (94 tonnes) or not specified (464 tonnes).

The major contributors to the mass of PVC are PCM at Sellafield and AWE (2D03, 2D06, 2D83/C,
2D90 and 7A21); decommissioning plant wastes (Processing Plants, Tanks, Silos etc.) (2D116) at
Sellafield and low alpha remote handled ILW (5B02) at Dounreay.

There are several operational and decommissioning waste streams originating from RSRL Harwell
described as a mixture of PVC and PTFE. These are solid wastes from glove boxes and other
alpha handling facilities (5C317), wastes from various sites which were packaged for the

AMEC/200615/001 Issue 3 Page 23 of 77


abandoned 1982 sea disposal campaigns (5C08) and operational solid wastes from research
reactors and active cells (5C30).

3.2.3 Non-halogenated plastics


The total mass of non-halogenated plastic is 2,780 tonnes. The non-halogenated plastic content is
either reported as condensation polymers (956 tonnes), ‘other non-halogenated plastics’
(1,713 tonnes) or not specified (110 tonnes).

The major contributor reported as a condensation polymer is Perspex (also called


polymethylmethacrylate, PMMA) from PCM at AWE (7A111). There are 307 tonnes of PCM from
Sellafield (2D06) reported to be a condensation polymer but the type has not been specified.

In addition, a waste encapsulant, the DOW WSB 101 polymer (a modified vinyl ester styrene resin)
from Trawsfynydd (9G04/C and 9G18/C) is described under the category of condensation polymer
under non-halogenated plastics. For the purposes of this report, the G-values for polymer
encapsulants are considered separately.

All the major contributors to the total mass of ‘other non-halogenated plastics’ comprise of
polyethylene. These waste streams are PCM at Sellafield (2D03, 2D06 and 2D90) and retrievals
from the ILW shaft at Dounreay (5B25).

3.2.4 Rubbers
The total mass of rubber is 1,982 tonnes. The total rubber content is either reported as
halogenated rubber (1,514 tonnes), non-halogenated rubber (209 tonnes) or not specified
(260 tonnes).

The major contributor to the total mass of halogenated rubbers is neoprene. The waste streams are
PCM at Sellafield (2D03, 2D83/C, 2D90) and AWE (7A111). There are 114 tonnes of PCM from
Sellafield (2D06) reported to be a halogenated rubber but the type has not been specified.

The RWI does not contain information on the type of non-halogenated rubbers. For the 2013 RWI,
non-halogenated rubber is assumed to consist of the commonly used styrene butadiene
copolymer.

3.2.5 Organic ion exchange materials


The total mass of organic ion exchange materials is 419 tonnes which is a small fraction (3 %) of
the total mass of organic waste materials. The total organic ion exchange material content is either
reported to be a styrene divinylbenzene copolymer (163 tonnes) or a phenol-formaldehyde based
resin (149 tonnes) or not specified (108 tonnes).

The major contributors to the total mass of styrene divinylbenzene copolymers are wastes from
reactor plant decontamination processes at Sizewell B (3S12), Hunterston B (4B01), Dungeness A
(9C36) and Trawsfynydd (9G04/C).

The major contributors to the total mass of phenol-formaldehyde based resins are wastes from
Magnox reactor plant decontamination processes at Trawsfynydd (9G18/C), Hinkley Point A (9D27,
9D28 and 9D29) and Dungeness A (9C02).

The project specification had noted that radioactive isotopes such as C-14 may be present within
ion exchange resins; however there are few data available as yet to quantify the amount. For
example, trace quantities of C-14, U and Pu are expected on the ion exchange resins for waste
stream 3S12 (CVCS Resins and Spent Resins (ILW)) from Sizewell B, but detailed analysis has
not yet been carried out. The content of C-14 present within the other waste streams is either
insignificant or not yet determined. (It is noted that waste stream 1A07 (GE healthcare wastes) that
contains a significant amount of C-14 is expected to be further processes before disposal and
therefore is no longer expected to release C-14 in a GDF.)

AMEC/200615/001 Issue 3 Page 24 of 77


3.2.6 ‘Other organics’
The total mass of ‘other organics’ is 780 tonnes which is a small fraction (6 %) of the total mass of
organic waste materials. The top contributing waste streams are given in Table 2. The majority of
the waste originates from Sellafield (2D03, 2D27/C and 2D90) or retrievals from the ILW shaft at
Dounreay (5B25).

Waste stream 2D27/C contains triphenyl phosphonium bromate encapsulated in cement. There is
insufficient information reported in the RWI to identify and quantify the other major contributors to
‘Other Organics’.

3.3 Discussion of the most sensible categories that are


expected to contribute to gas generation by radiolysis
Based on the information described by the 2013 UK RWI it is suggested that similar categories to
those in the inventory be considered for the review of G-values (Section 4), with additional
categories for encapsulant materials (notably vinyl ester styrene resin that is currently described in
the 2013 UK RWI under non-halogenated plastics, and cementitious encapsulants), and water.

Comparison of the categories of organic inventory data considered in the 2013 UK RWI against
those currently considered by SMOGG is provided (Table 3). Based on this comparison, a final list
of categories for which G-values have been suggested for use in future gas generation calculations
is included as the right hand column in Table 3.

Table 3 Comparison of groups of materials in the 2013 UK RWI, SMOGG and the final
list of categories for which G-value have been suggested for use in future gas generation
calculations.
Categories of relevant Categories of materials Final list of categories
wastes presented in the currently considered by suggested for use in future
2013 UK RWI SMOGG gas generation calculations

Water Water
Cellulose ‘Cellulose’ Cellulose
Halogenated plastic Halogenated plastic
Non-halogenated plastic Non-halogenated plastic
Rubber (halogenated and ‘Polymers’ Rubber (halogenated and non-
non-halogenated) halogenated)
Ion exchange material Ion exchange materials
Other organics* ‘Oils’
‘Small organic molecules’** ‘Small organic molecules’**
Cementitious encapsulant
(water)
and its pore water
Polymer encapsulant
* Other organics comprise 6% of the overall mass of organic wastes described in the 2013 UK RWI
wastes, and given the lack of available G-value information it is suggested they are included within
one of the other categories (see Section 5.1.7 for details).
** G-values are for CO2 or CH4 gases for ‘small organic molecules’

AMEC/200615/001 Issue 3 Page 25 of 77


4 Review of G-values for relevant materials
This section critically reviews available information on G-values. The information on organic waste
materials described in the 2013 UK RWI (Section 3) has been used to determine the groups of
materials to which G-values are assigned. As a result, the following categories of media and
materials have been considered:
 water,
 cellulose,
 halogenated plastic,
 non-halogenated plastic,
 rubber (halogenated and non-halogenated),
 ion exchange materials,
 other organics,
 small organic molecules
 encapsulation materials:
- cementitious encapsulant and its pore water, and
- polymer encapsulant.

Using the information on G-values reviewed in this section, the subsequent sections (Section 5 and
Section 6) discuss and suggest appropriate G-values for water and sensible groupings of materials
for use in future gas generation calculations.

4.1 Background information for the review of G-values


The robust assessment of the G-value for a material and radiation type is not straightforward, but
depends on factors such as the radiation dose rate, the absorbed radiation dose, variation in
material composition, morphology and ageing processes, and the radiation environment
(atmosphere and temperature).

Another important factor is the linear energy transfer (LET ≈ stopping power) of the incident
radiation, which is equivalent to the rate of energy loss per unit path length. Alpha radiation has a
very short range because of its high LET and an external source will deposit its energy rapidly near
the surface of materials, whilst gamma radiation has a low LET and deposits its energy throughout
a material. This variation in range becomes important in the determination of gas generation
because, by definition, it is normally assumed that dosimetry refers to the amount of radiation
energy absorbed per unit mass. As alpha radiation deposits its energy in the surface layer of about
40 microns or so in a particular sample of material, it is important when calculating the gas
produced to account for the energy deposited in the surface layer mass not the mass of the whole
sample, as would be adopted for gamma radiation.

Variations in G-values with LET are due to the local geometry of the energy deposition about the
path of the radiation. Energy deposition and then decomposition of molecules of the medium leads
to the formation of reactive radical species. These radicals then react with each other to give final
products or with the final products themselves to reform molecules of the medium. An increase in
LET leads to an increase in the probability of radical-radical reactions. For instance, the
decomposition of water produces the H atom (and the OH radical). Increasing the LET increases
the probability of H atom reactions with other H atoms to give hydrogen molecules (H2). As a result,
the G-value for H2 in water from gamma radiation is about 0.45 molecules/100 eV (low LET) and it
is about 1.2 molecules/100 eV from alpha radiation (high LET) (see Section 4.2). Any product that
is formed by two transient radical species will increase with increasing LET. A variety of processes
can lead to H2 formation in organic materials and the mechanism of decomposition of the medium
has a strong effect on the variation in gaseous G-values with LET.

In some cases the available experimental data for G-values are sparse, and full details of the
experimental setup are not always available, or in some cases experimental methods have been
superceded by improved best practices and therefore some data (typically older data) may be
viewed as having a larger uncertainty than data for which full experimental details are available.
However, the view has been taken that the inclusion of an indicative G-value for a particular

AMEC/200615/001 Issue 3 Page 26 of 77


material and radiation type may be better than the absence of any data, assuming that the material
is expected to be present in significant amounts within a GDF.

Based on practical experience of carrying out irradiation experiments, it is expected that obvious
degradation of any organic material typically takes place above a total dose of 100 kGy. Note that
one gray (Gy) is the absorption of one joule of energy, in the form of ionising radiation, per kilogram
of matter.

Organic media typically undergo radiolytic decomposition by the breakage of a C-H bond to give
H atoms and a carbon centred radical. The main gaseous product in many organic media is
hydrogen gas, H2. A variety of mechanisms are responsible for hydrogen gas production. In simple
aliphatic compounds hydrogen gas is produced by H atom combination reactions between two
H atoms, and by H atom abstraction reactions with the parent compound (i.e. a hydrogen free
radical is abstracted from a substrate according to the general equation: X• + H—Y → X—H + Y•).
The latter process usually dominates. Aromatic compounds are thought to be radiation resistant
and hydrogen gas yields in gamma radiolysis are generally several orders of magnitude lower in
aromatic compounds than in aliphatic compounds. Yields of hydrogen gas tend to be somewhat
independent of LET in aliphatic compounds, so yields with alpha particles can be assumed to be
similar to those for gamma radiation should no actual data exist. Yields of hydrogen gas in aromatic
compounds tend to increase dramatically with increasing LET and only actual data should be used
for alpha particle radiolysis. The yields are usually measured in an inert atmosphere in order to
minimize side reactions of the H atom precursors to H2.

As discussed above, the primary degradation process in organic materials is the breakage of the
C-H bond. The carbon centered radicals produced in this process can lead to dozens of products
that are hard to determine and quantify. On the other hand, the H atoms most often lead to H2.
Radiation chemistry studies often focus on the production of H2 because its measurement gives
some rough indication of the general radiation sensitivity of the compound. It is noted that small
effects of molecular weight and even cross linking on the yield of gases by irradiation of organic
materials can be found (e.g. [12]). However, the environment often has the strongest effect. For
example, the presence of air or water can influence yield of H2 by scavenging the H atom precursor
responsible for its production (e.g. [13]).

In aqueous systems that contain, or come into contact with, organic molecules there may be an
effect on the measured G-value. For example, there is considerable experimental evidence
(e.g. [14, 15]) of reaction between the hydrogen atom and trace amounts of saturated organic
molecules:

H• + RH → H2 + R•, (where R is a saturated organic molecule)

This reaction results in the scavenging of the H atom and the formation of hydrogen gas, which
would result in an increase in the G-value of water (of around 0.6 molecules / 100 eV) in the
presence of trace amounts of saturated organics. However, it is noted that this yield is small in
comparison to a typical hydrogen gas yield from aliphatic organic molecules, which may be as
much as 5 to 7 molecules / 100 eV (see Section 4.5), and therefore this reaction can generally be
ignored.

In contrast, the reaction of the H atom with trace amounts of unsaturated organic molecules does
not lead to an increase in the G-value of water as no hydrogen gas is generated ([14, 15]):

H• + -CH=CH- → -CH2-CH•-
(where -CH=CH- denotes an organic molecule containing an unsaturated bond).

In some experiments designed to measure G-values of specific organic materials, the radiolysis
products may be systematically removed during irradiation and their quantities measured. In other
experiments, the organic material may be encapsulated in a vessel and the gaseous headspace
analysed away from radiation sources. In the latter case, additional radiolysis of degradation
products might also contribute to the gases in the headspace and, in some cases, organic material
degradation products may influence degradation rates. A typical example is the production of

AMEC/200615/001 Issue 3 Page 27 of 77


HCl(g) from degraded PVC which can autocatalyse further degradation of the PVC material if the
HCl(g) remains in close proximity.

Even the most thorough of experimental studies often consider only a limited set of radiolysis
conditions (total dose and dose rate) and in some cases radiolysis of some secondary degradation
products will be inevitable. However, product release from secondary radiolysis at long times and
at low dose rates is rarely studied, even though it may be important in real systems.

4.2 Review of G-values for water


There has been extensive research on the effects of ionising radiation on water. This is not
surprising as water is the primary coolant for several power generation reactor designs and
changes in water chemistry brought about by radiation can result in the production of oxidising
radical species that can raise concerns about corrosion in reactor materials. The G-values for
hydrogen production from gamma, beta and alpha radiolysis of water are of interest in the study of
radiation effects in waste materials, especially materials that normally contain some water, such as
cements and ion exchange materials.

Hydrogen gas is the only primary gaseous radiolysis product produced from pure water. Oxygen
from water decomposition is mainly incorporated into hydrogen peroxide, which may decompose at
certain surfaces to release oxygen [16]. Much of the information on the production of H2 in water
radiolysis is based on early studies that found an increase in H2 yield with increased LET of the
radiation (see [17] and references within), and therefore alpha radiation typically exhibits larger
G-values for hydrogen in water than gamma radiation. The following reactions that include the
- -
hydrated electron as a reactant (e (aq)) and its precursor, the non-hydrated electron (e (pre)), are
responsible for most of the formation of hydrogen in water, with equation (5) considered to be the
dominant route to molecular hydrogen formation by alpha radiolysis of water [17, 18, 19]:

e (aq) + e (aq) (+ 2H2O) → H2 + 2OH


- - -
(2)
e (aq) + H (+ H2O) → H2 + OH
- -
(3)
H + H → H2 (4)
Dissociative electron recombination:
e (pre) + H2O•+ → H2O*(excited state) → H2 + O꞉
-
(5)
Dissociative electron attachment:
e (pre) + H2O → H2 + O•- e (pre) + H2O (+ H2O) → H- + OH (+ H2O) → H2 + OH• + OH-
- -
(6)

Studies have shown that efficient scavenging of the hydrated electron (and/or its precursor, the
non-hydrated electron) can lead to a decrease in hydrogen production [17, 18]. For example in the
2- 2- 2- - - 2+ 2+
presence of one of a variety of scavengers such as SeO4 , MoO4 , Cr2O7 , NO2 , NO3 , Cd , Cu ,
and H2O2, decreased yields of hydrogen are observed, due to reaction of the scavenger with the
hydrated electron, or its precursor, or both. As a result, hydrogen production in the presence of
high aqueous nitrate concentrations that may be associated with some wastes would be expected
to be considerably lower than in pure water or dilute solutions (e.g. waste streams 5B04 MTR
Raffinate or 5B01 PFR Raffinate or 5B05 DFR Raffinate, all at Dounreay that contain much nitric
acid under oxidising conditions.)

The G-values for hydrogen generation, GH2, from water at room temperature from gamma radiation
are shown in Table 4. A detailed review of the radiolysis of water has been carried out by
Christenson [20], with extensive G-value data for hydrogen generation for radiations of different
LET. Some of the data shown in these tables are reasonably recent (2002 to 2004) and confirm
values measured some time ago (1980’s). Based on Table 4, the mean value of GH2 for gamma
radiation is 0.44 molecules/100 eV. As a comparison, a recent review by Elliot [21] gave
0.45 molecules/100eV as the generally accepted G-value for hydrogen in gamma irradiated,
pH neutral water. G-value data for gamma radiation from a pH range of 3 to 13 give a similar value
of 0.45 molecules/100 eV [22].

However hydrogen production from water by gamma radiation is affected by temperature and
shows a steady increase with increasing temperature. An empirical relationship that describes the
increase has been defined by Elliot [21] and is shown in Figure 5. At temperatures relevant to a

AMEC/200615/001 Issue 3 Page 28 of 77


GDF, the effect of temperature for hydrogen production from water by gamma radiation is small
relative to other uncertainties and therefore need not be considered in any calculation of gas
generation rate for gamma radiation.

For alpha radiation, the G-values for hydrogen generation from water are shown in Table 5. The
generally accepted value for the production of hydrogen by alpha radiation is 1.2 molecules/100 eV
for 5 MeV helium ions, which is consistent with the mean of 1.22 molecules/100 eV from Table 5
(excluding the last two data points). Note that the last two data points, which are the largest
G-value data in Table 5, are from older experiments and are considered to be less accurate, and
therefore represent an upper value. It is also noted that results expected with alpha particles are
generally considered to be equivalent to those found with 5 MeV helium ions [18], which are
obtained from experiments that use helium ions from accelerators and therefore have a relatively
low total error of ±10% [17]. No data are available that consider alpha radiation of water at high
temperatures [20], or at alkaline pH [22].

Note that G-value data that consider the effects of dissolved solutes within cementitious pore water
is considered in Section 4.10.1.

Table 4 G-values for hydrogen generation from water by gamma radiolysis at room
temperature from Christenson [20] (except [33])
GH2 (molecules per 100eV) Data source, [reference number]
0.45 [23]
0.43 [24]
0.42 [25]
0.47 [26]
0.42 [27]
0.43 [28]
0.44 [29]
0.45 [30]
0.44 [31]
0.45 [32]
0.45 [33]

Table 5 G-values for hydrogen generation from water by alpha radiolysis at room
temperature from Christenson [20] (except [33])
GH2 (molecules per 100eV) Data source, [reference number] Comment
1.17 [34]
1.3 [23]
1.28 [35]
1.22 [36]
1.12 [33]
1.7 [37] Considered to be less accurate
1.7 [38] Considered to be less accurate

AMEC/200615/001 Issue 3 Page 29 of 77


GH2

Figure 5 G-values for hydrogen from gamma irradiated water measured as a function of
temperature (from Elliot, [21], and references therein). The ‘accepted’ room temperature
G-value for hydrogen is plotted as a black circle. Note that in the plot legend, the unit of
-3
concentration, m, refers to moles dm .

4.3 Review of G-values for cellulosic materials


The 2013 UK RWI describes paper and cotton or wood as the main cellulosic materials
(Section 3.2.1). Historically GGAS values for cellulosic materials have been difficult to obtain,
however some (typically older) data are available [22] and are shown in Table 6. It is noted that
these data have not been obtained using up-to-date experimental methods and therefore may not
be very accurate. For example, it is rare that G-values for aliphatic compounds for high LET
radiation (i.e. alpha) are significantly lower than those for low LET radiation (i.e. gamma), as shown
in Table 6, although it is noted that the thickness of the alpha particle penetration is not stated in
the experimental detail. If (incorrectly) the total mass of the material has been used to calculate the
G-value then this would give a greater than expected result, although the reported G-value data are
low. Therefore it is suggested that there is significant uncertainty associated with these data, and
they should be used with caution until more experimental data for cellulosic materials are available.

Cellulosic materials are polysaccharide structures containing ester links and alcohol functional
groups. Therefore, when irradiated, even under anaerobic conditions, CO and CO2 are generated,
and hydrogen is not always the dominant gas produced.

It is noted that gas release during irradiation will depend on the structure and type of the cellulose
being irradiated and its environment, despite the generically similar semi-crystalline structure that is
common to all cellulosic materials.

Examination of the two G-values for total gas (mainly hydrogen) for beta radiation (electrons) from
‘American cotton’ shows a substantial range of 6 molecules/100eV and 2 molecules/100eV.
However, it is noted that these G-values have been determined under very different total dose
conditions. The G-value of 6 molecules/100eV has been determined under a low total dose of
1 kGy, while the G-value of 2 molecules/100eV has been determined under a significant total dose
of 250 kGy. Based on practical experience it is inferred that there is a dose dependence of the

AMEC/200615/001 Issue 3 Page 30 of 77


measured G-value, whereby irradiation of a material at a high dose rate will typically lose hydrogen
gas rapidly within an initial short timescale as the surface of the material is irradiated first, and then
the rate of gas production will drop to a lower long-term G-value since the gas must first diffuse out
of the interior of the material before it can escape away from the surface. By contrast, a low total
dose rate may allow a higher G-value for the production of hydrogen gas to continue throughout
the duration of the experiment if the long term gas production rate is not reached. Based on this
assumption, it is suggested that the lower G-value for the production of hydrogen gas by beta
radiation from cellulose is a more typical scenario for expected conditions within a GDF, however
more experimental evidence is required to confirm this hypothesis.

A discussion of G-values and recommendations for cellulose is provided in Section 5 and


Section 6.

Table 6 G-values for cellulosic materials at room temperature from Reference [22].
Cellulose type Radiation Environment G-value Comments
type (molecules
per 100eV)
Mixed alpha oxygen Ggas 0.5 60 % H2, 25 % CO2,
cellulosics, dry depleted 15 % others †
Cheesecloth, alpha oxygen GGAS 1.3 55 % H2, 35 % CO2,
wet depleted 10 % others †
Paper alpha oxygen GGAS 1.5 61% H2, 26% CO2,
depleted 13% CO †
Paper alpha argon GGAS 1.44 †
Mixed alpha air GGAS 0.6 40% H2 †
cellulosics
Cheesecloth, alpha air GGAS 1.4 85% H2 †
wet
Kimwipes alpha air GGAS 0.6 55% H2 †
Kimwipes wet alpha air GGAS 0.4 73% H2 †
Paper tissue alpha air GGAS 0.9 50% H2 †
“American” electron* vacuum GGAS 6 1 kGy, mainly H2
cotton
“American” electron* vacuum GGAS 2.0 250k Gy, mainly H2
cotton
Cotton gamma vacuum GGAS 3.7 35% H2, 43% CO, 22% CO2
Cotton gamma nitrogen GGAS 3.7 25% H2, 50% CO, 25% CO2
Cotton gamma oxygen GGAS 6.2 19% H2, 27% CO, 55% CO2
Cotton gamma air GGAS 5.5 13% H2, 60% CO,
† Thickness of alpha particle penetration of sample not stated.
* These experiments were carried out using electrons. An electron is equivalent to a beta particle.

AMEC/200615/001 Issue 3 Page 31 of 77


4.4 Review of G-values for halogenated plastics
The halogenated plastics that are expected to be contained within the 2013 UK RWI are
polyvinylchloride (PVC) and a mixture of PVC and polytetrafluroethylene (PTFE).

PVC is expected to be in the form of flexible film, where it is used throughout the nuclear industry
for glove-box bagging operations, protective tents and protective suits.

PVC has a relatively simple molecular structure [C2H3Cl]n although the radiolysis of PVC is
complex. The most abundant gas generated during irradiation is normally HCl, and hydrogen
usually accounts for about 10 % of the gas generated [22]. HCl can catalyse the main chain
unzipping degradation mechanism of PVC, which is probably the main reason why the radiolytic
(and thermal) degradation of PVC is so complex. If PVC is irradiated in a confined space, the
released HCl will readily catalyse the degradation process and generate more HCl which may then
attack its local environment. Therefore there is the potential to cause corrosion of metals and
reduce the pH in the local environment. HCl will readily dissolve in water which may enhance
corrosion, despite pH buffering by the cementitious environment present in many wastes. This is a
more widespread issue for a GDF and has been discussed in relation to increased corrosion of
stainless steel drums for certain wastes [39].

In the nuclear industry, PVC is rarely used, if at all, in its original un-plasticised state. It is more
commonly used as a plasticised film which will contain a mixture of additives such as plasticising
agents, fire retardants and stabilisers. The types of additives used by PVC film manufacturers will
vary depending on price, availability and modern attitudes to the toxicity of phthalate plasticisers for
example. The high level of variability in PVC compositions leads to a wide range of radiation
stabilities, as shown in Table 7. Usually, the dominant gas produced is HCl, however the action of
some additives can suppress the generation of HCl, which would allow hydrogen gas to be the
most abundant gas produced for some compositions of PVC.

The reported data in Table 7 are strongly influenced by the composition of the irradiated PVC,
which is usually difficult to report as manufacturers tend not to disclose details of proprietary plastic
formulations. The plasticised and stabilised PVC materials reported in Table 7 are considered to be
typical of those in use within the nuclear industry internationally. However, some work on PVC
materials has been carried out on a plasticised and stabilised PVC film currently used in bagging
operations, and on a plasticised and stabilised PVC film expected to be used in future bagging
operations, in the UK nuclear industry [40]. This shows that the G-values measured on PVC
materials used in the UK nuclear industry are within the range reported internationally.

A clear temperature dependence of the GHCl-value for PVC is observed. A guidance temperature of
less than 50°C for all waste packages following closure is considered, as stated in the Generic
Disposal System Technical Specification [41]. There is significant uncertainty associated with
G-values expected for irradiation of PVC at high temperatures. Also, a small variation in hydrogen
gas yield from alpha radiation with the molar mass of PVC has been observed [12].

Overall, PVC has a complex mixture of radiolytic degradation mechanisms and a strong
dependence of GGAS-values on the specific composition of individual PVC materials and their
additives.

One G-value for production of gas by gamma radiation from PTFE is available, showing that the
expected gas yield is low (GGAS 0.098 molecules/100eV, [42]), and therefore the gas produced by
any mixture of PVC/PTFE wastes would be expected to be dominated by the PVC component for
gamma radiation.

A discussion of G-values and recommendations for PVC and PTFE is provided in Section 5 and
Section 6. It is noted that pure PVC is not expected to be used in the nuclear industry and that the
discussion provided in Section 5 focuses on plasticised and stabilised PVC.

AMEC/200615/001 Issue 3 Page 32 of 77


Table 7 G-values for PVC irradiated at room temperature (unless otherwise stated; grey
text shows data at higher temperature)
Plastic details Radiation Environ G-values Data Comments
type ment (molecules source
per 100eV)
Plasticised alpha air GH2 0.8 [43] PVC from Rocky Flats
PVC GCO2 1.2 Environmental Technology
Site, US – typical of plastics
used in packaging of TRU
wastes.
5
2×10 rad/hr dose rate in
thin layer defined by range
of alpha particle.
11-12 Mrad total absorbed
dose.
Plasticised alpha depleted GHCl 0.2-0.8 [44] Irradiations at different
PVC bag oxygen GH2 0.3-0.7 dose rates at 30°C. Dose
material rate measured in thin layer
GCO2 0.7-1.3
at sample surface.
Plasticised alpha depleted GHCl 0.4 [44] Irradiations at different
PVC sheet oxygen GH2 0.3-0.4 dose rates at 30°C. Dose
rate measured in thin layer
GCO2 0.8-1.1
at sample surface.
Plasticised and alpha depleted GGAS 0.7, [22] Glove box bags from Rocky
stabilised PVC oxygen GH2 0.6 Flats Environmental
Technology Site, US
83% H2 plus CO+CO2(12%)
Thickness of alpha particle
penetration of sample not
stated.
Plasticised and alpha depleted GGAS 0.8, [22] PVC (Pylox) gloves; 30 °C
stabilised PVC oxygen GH2 0.7 85% H2 plus
CH4(2%),CO2(6%),CO(7%)
Thickness of alpha particle
penetration of sample not
stated.
Plasticised and alpha depleted GGAS 6.3, [22] PVC (Pylox) gloves; 70 °C
stabilised PVC oxygen GH2 5.3 85% H2 plus
(at higher CH4(2%),CO2(6%),CO(7%)
temperature) Thickness of alpha particle
penetration of sample not
stated.
Pure PVC alpha de- GGAS 0.45, [12] 100 kGy dose at 1000
aerated 0.41, 0.39 Gy/min. 24 °C
Variation with polymer
molar mass range: 22 000,
-1
47 000, and 99 000 g mol
Local dose rate ~3.6 kGy/s
within thin layer defined by
alpha particle (6.4mm thick
and length 0.023mm).'

AMEC/200615/001 Issue 3 Page 33 of 77


Plastic details Radiation Environ G-values Data Comments
type ment (molecules source
per 100eV)
Pure PVC electron* vacuum GHCl 13 [22] 5 to 20 Mrad total dose;
30 °C
Pure PVC (at electron* vacuum GHCl 23 [22] 5 to 20 Mrad total dose;
higher 70 °C
temperature)
Pure PVC gamma vacuum GGAS 8.4 [45] 10 Mrad total dose
GH2 0.2
GHCl 8.2
Pure PVC gamma vacuum GHCl 4 to 9 [22] 3 Mrad total dose. Only HCl
detectable by measurement
technique
Pure PVC film gamma vacuum GGAS 2.6 [46] 0.2 MGy total dose
produced by GHCl 2.4
precipitation GH2 0.18
from stabilised
polymer
Pure PVC gamma de- GGAS 0.23, [12] 100 kGy dose at
aerated 0.27, 0.25 80 Gy/min. 24 °C
Variation with polymer
molar mass range: 22 000,
-1
47 000, and 99 000 g mol
Pure PVC gamma air GGAS 10.3 [22] 20 Mrad total dose
GHCl 8
GH2 0.1
GCO 1.0
GCO2 1.2
Pure PVC gamma air GGAS 21.6 [47] 1%H2, 85% HCl, 4% CO,
GH2 0.2 10% CO2 in evolved gas
Pure PVC gamma air GHCl 20-27 [48] 2 to 27 Gy/min dose rate;
25 °C
Pure PVC gamma aerated GGAS 0.22 [12] 100 kGy dose at
80 Gy/min. 24 °C
Polymer molar mass of:
-1
22 000 g mol
Pure PVC film gamma air GGAS 6.1 [46] 150 torr O2
produced by GHCl 5.9
precipitation GH2 0.1
from stabilised
polymer
Stabilised PVC gamma vacuum GGAS 2.9 [22] 10Mrad total dose
film GHCl 2.7
GH2 0.2
Stabilised and gamma vacuum GGAS 0.3 [45] 10 to 20 Mrad total dose
plasticised PVC GHCl 0.03
GH2 0.1
GCO 0.1
GCO2 0.1

AMEC/200615/001 Issue 3 Page 34 of 77


Plastic details Radiation Environ G-values Data Comments
type ment (molecules source
per 100eV)
Stabilised and gamma vacuum GGAS 1.4 [47] Three different materials
plasticised GHCl 1.2 studied in this case:
PVC 1 GH2 0.1 PVCl 1
(8%H2, 83%HCl, 5%CO,
3% CO2)
Stabilised and gamma vacuum GGAS 0.7 [47] Three different materials
plasticised GHCl 0.1 studied in this case :
PVC 2 GH2 0.2 PVC 2
(26%H2, 14%HCl, 8%CO,
50% CO2)
Stabilised and gamma vacuum GGAS 1.1 [47] Three different materials
plasticised GHCl 0.1 studied in this case:
PVC 3 GH2 0.2 PVC 3
(15%H2, 8%HCl, 9%CO,
66% CO2)
Plasticised and gamma air GGAS 5.8 [45] 10 Mrad
stabilised PVC GHCl 4.7
film GH2 0.27
Stabilised and gamma air GGAS 5.0 [47] 10Mrad
plasticised PVC GHCl 2.6 (5%H2, 52%HCl, 6%CO,
GH2 0.3 37% CO2)
Stabilised and gamma air GGAS 1.4 [47] 10Mrad
plasticised PVC GHCl 0.2 (15%H2, 15%HCl, 17%CO,
GH2 0.2 51% CO2)
Stabilised PVC gamma air GGAS 1.9 [47] 10Mrad
film GHCl 0.2 (10%H2, 10%HCl, 9%CO,
GH2 0.2 70% CO2)
-1
Plasticised and gamma air GHCl 0.13 & [40] 0.4 Gy hr , 40 °C at 700 Gy
stabilised PVC 0.23 for PVC1** & PVC2***
film respectively
-1
Plasticised and gamma air GHCl 0.06 & [40] 8 Gy hr , 40 °C for
stabilised PVC 0.26 17.5 kGy for PVC1** &
film PVC2*** respectively
-1
Plasticised and gamma air GHCl 181 & [40] 0.4 Gy hr , 70 °C at 700 Gy
stabilised PVC 1.16 for PVC1** & PVC2***
film (at higher respectively
temperature)
-1
Plasticised and gamma air GHCl 9.67 & [40] 8 Gy hr , 70 °C for
stabilised PVC 1.1 17.5 kGy for PVC1** &
film (at higher PVC2*** respectively
temperature)
* This experiment was carried out using electrons. An electron is equivalent to a beta particle.
** PVC1 is a PVC bag material (pink translucent film) currently used as ‘Residues storage
packages’, made by Weston Vinyls Plc in From Somerset UK.
*** PVC2 is a Riflex film (colourless translucent film), proposed for future bagging operations, made
by Stephens Plastics in Corsham, Wiltshire, UK.

AMEC/200615/001 Issue 3 Page 35 of 77


4.5 Review of G-values for non-halogenated (thermo)plastics
Thermoplastics are a group of polymeric materials that can be formed into shape after heating and
return to a solid state after cooling. Non-halogenated thermoplastics include polyethylene (PE,
polythene), polypropylene (PP) and polystyrene (PS). Thermoplastic polymers differ from
thermosetting polymers (e.g. epoxy resins and elastomers such as Neoprene and Hypalon) which
form irreversible bonds during the curing process.

The main non-halogenated (thermo)plastics that are described in the 2013 UK RWI (Section 3.2.3)
are polyethylene and Perspex (PMMA). Perspex and polyethylene have been extensively
examined by gamma and alpha radiation and G-value data for all relevant non-halogenated
(thermo)plastic materials are shown below, in Section 4.5.1 and Section 4.5.2. (Note that the
polymer encapsulant also described within Section 3.2.3 will be discussed under the polymer
encapsulant review (Section 4.10.2).)

It is noted that the G-value data for non-halogenated thermoplastics are shown for room
temperature radiolysis. At higher temperatures, the G-value would be expected to increase [22],
although at slightly increased temperatures, for example 35°C, the expected increase in G-value is
likely to be within the level of uncertainty of room temperature data.

4.5.1 Polyethylene
Polyethlyene is one of the most common plastics in use today because of its ability to be produced
in various forms and morphologies. It is described within the 2013 UK RWI (Section 3.2.3) and is
expected to be present in the form of bottles, containers, pipework and films, typically either as low
density polyethylene (LDPE) or high density polyethylene (HDPE).

As it has a relatively simple structure, [C2H4]n, polyethylene has been studied widely to help
understand basic radiation effects in plastics and a large amount of data describing G-value
measurements is available (Table 8).

The structure of polyethylene is a saturated hydrocarbon and therefore the most abundant gas
released during irradiation is hydrogen, especially under anaerobic conditions. Methane may also
be released in small quantities from polyethylene during radiolysis, typically 2 % of the total gas
produced [22], along with other hydrocarbon gases in negligible quantities.

Oxygen plays an important role in the radiolysis of polyethylene. Oxidation takes place when
polyethylene is irradiated in the presence of oxygen, which allows additional gases such as CO and
CO2 to be produced. A recent study has suggested there are clear differences in observed
proportions of gases generated for similar LDPE samples in the presence of gamma radiation and
this may be attributed to the dose rate at which the sample is irradiated (Table 8, [51]). For an
equivalent total dose of 150 kGy, the ratio of radiolytic CO + CO2 to hydrogen is greater (15 %) at
low dose rate than at higher dose rate (3 %). This result may be accounted for by diffusion-limited
oxidation within the LDPE where the levels of oxidation throughout the material would be expected
to be more extensive at low dose rates since more time would be available for oxygen to diffuse
completely into the material during the irradiation.

Equally, the total dose is observed to have an effect on the G-value (Table 8, [51]). Irradiation of
LDPE samples to 10 MGy (considered to be a possible maximum dose within a GDF over the total
timescale) gives an observed G-value from gamma radiation of 0.8 molecules/100eV, while
-1
irradiation to 150 kGy (at a similar dose rate of 4 kGy hr ) gives an observed G-value from gamma
radiation of 4.0 molecules/100eV. It is suggested that the reason for the difference is due to the
hydrogen to carbon ratio that decreases as a function of total dose in polyethylene and as the dose
increases there is effectively less available hydrogen remaining in the polymer that can be released
radiolytically. Back reactions where an over pressure of hydrogen restricts further release have
also been suggested but discounted in polyethylene in work by Seguchi [49].

A discussion of G-values and recommendations for polyethylene is provided in Section 5 and


Section 6.

AMEC/200615/001 Issue 3 Page 36 of 77


Table 8 G-values for polyethylene at room temperature unless otherwise stated
Plastic Radiation Environment GH2 Data Comments
details type (molecules source
per 100eV)
Polyethylene alpha oxygen 1.7-2.4 [22] Thickness of alpha
(generic) depleted or particle penetration of
absent sample not stated.
Polyethylene alpha nitrogen GH2=3 [43] 2 kGy/hr dose. Only
film GCO2=1.8 thin surface layer
defined by range of
alpha particle
considered.
Polyethylene alpha air GH2=2.5 [43] 2 kGy/hr dose. Only
film GCO2=4.8 thin surface layer
defined by range of
alpha particle
considered.
Polyethylene from
Rocky Flats
Environmental
Technology Site, US –
typical of plastics used
in packaging of TRU
wastes
Polyethylene alpha argon 4.6 [50] 100 kGy dose at 1000
Gy/min. 24 °C
Dose rate measured in
thin layer at sample
surface.
Low density gamma air/vacuum 0.8 [51] Irradiated to 10 MGy at
-1
polyethylene 4 kGy hr
Low density gamma air/vacuum 4 [51] Irradiated to 150 kGy
polyethylene at
-1
4 kGy hr
Low density gamma air/vacuum 1.2 [51] Irradiated to 150 kGy
polyethylene at
-1
40 kGy hr
Polyethylene gamma oxygen 2.8-4.1 [22] Wide range of
(generic) depleted or polyethylene of
absent different density and
morphology
High density gamma vacuum 2.9 [22]
polyethylene
Low density gamma vacuum 3.1-3.5 [22]
polyethylene
Polyethylene gamma air 3.5 [52]
bottle top
(HDPE)

AMEC/200615/001 Issue 3 Page 37 of 77


Plastic Radiation Environment GH2 Data Comments
details type (molecules source
per 100eV)

Polyethylene gamma vacuum 4 [49]


(ultra high
molecular
weight)
Polyethylene gamma air 3-4 [53]
model
compounds
Polyethylene gamma argon 3.3 [50] 100 kGy dose at
200 Gy/min. 24 °C

4.5.2 Other (thermo)plastic materials


G-values for other thermoplastic materials that may be present in the wastes are shown in Table 9.
Of these, Perspex (PMMA) is described as being present within the 2013 UK RWI (Section 3.2.3).

Plastics that contain ester and carbonyl groups such as Perspex, polyethylene terephthalate (PET)
and polycarbonate (PC) tend to generate more CO and CO2 than hydrogen when irradiated. As a
result, the production of gas from these materials is defined in terms of GGAS rather than GH2.

It is noted that the G-values shown for polypropylene are similar to polyethylene, which reflects the
fact that both materials are saturated hydrocarbons. Also, the G-value data for polystyrene from
gamma radiation are significantly lower than the other materials due to the aromatic component of
the material structure that offers protection to the polymer chains during irradiation.

A discussion of G-values and recommendations for other (thermo)plastic materials is provided in


Section 5 and Section 6.

Table 9 G-values for non-halogenated plastics at room temperature


Plastic details Radiation Environ- G-values Data Comments
type ment (molecules per source
100eV)
Polymethylmeth alpha oxygen GGAS 2.0 [22] 23% hydrogen,
acrylate depleted mainly CO and CO2
(PMMA) Thickness of alpha
particle penetration
of sample not
stated.
Polymethylmeth alpha argon GH2 0.62 [50] 100 kGy dose at
acrylate 1250 Gy/min. 24 °C
(PMMA) Dose rate measured
in thin layer at
sample surface
Polymethylmeth gamma vacuum GGAS 4.1 [22] Gases mainly CO
acrylate GH2 0.3 and CO2
(PMMA)
Polymethylmeth gamma argon GH2 0.24 [50] 100 kGy dose at
acrylate 200 Gy/min. 24 °C.
(PMMA)

AMEC/200615/001 Issue 3 Page 38 of 77


Plastic details Radiation Environ- G-values Data Comments
type ment (molecules per source
100eV)

Polypropylene gamma oxygen GGAS 2.4-3.2 [22]


depleted or
absent
Polypropylene gamma oxygen GGAS 2.6 [22]
present
Polystyrene alpha argon GH2 0.15 [50] 100 kGy dose at
1250 Gy/min. 24 °C
Dose rate measured
in thin layer at
sample surface
Polystyrene gamma argon GH2 0.033 [50] 100 kGy dose at
200 Gy/min. 24 °C
Polystyrene gamma air GH2 0.1 [22]
GGAS 0.3
Polystyrene gamma vacuum GGAS<0.1 [22]
Polystyrene gamma air GH2 0.04 [52]
Nylon 6,6 gamma vacuum GH2 0.4 [54] Common nylon
Nylon II gamma vacuum GH2 1.1 [54] Transparent nylon
bag material
Polyethylene gamma vacuum GGAS<0.1 [22]
terephthalate
(PET) polyester
Polyethylene gamma air GGAS 0.3 [22] Gases released
terephthalate (GH2<0.1) mainly CO and CO2
(PET) polyester
Polyethylene gamma argon GH2 0.04 [55] 4000 kGy dose at
terephthalate 100 Gy/min.
(PET) polyester
Polyester resin gamma air GGAS 0.75 [56]
Polysulphone gamma air GGAS 0.01-0.1 [22]
Polycarbonate gamma vacuum GGAS 0.9 [22] 97% CO or CO2
Polyvinyl gamma vacuum GGAS 1.4 [22] 64% hydrogen
acetate

4.6 Review of G-values for rubbers


The rubbers described in the 2013 UK RWI are polychloroprene (neoprene, a halogenated rubber)
and styrene butadiene copolymer (a non-halogenated rubber), see Section 3.2.4. It is expected that
chlorosuplhonated polyethylene (CSPE, or Hypalon) may be within the wastes, although this
rubber has not been explicitly described within the 2013 UK RWI. It is also likely that some amount
of natural or artificial rubber will be present in the form of latex/isoprene gloves. Generally, rubber
materials will have been used for glove-box gauntlets, seals and PPE (personal protective
equipment) or similar.

AMEC/200615/001 Issue 3 Page 39 of 77


The available G-value data for relevant rubbers are summarised in Table 10.

Generally, hydrogen gas is produced by irradiation of rubber materials. However, significant


amounts of HCl may be generated by irradiation of neoprene and CSPE although G-values for
generation of HCl are rarely reported. For example, a recent programme carried out for RWMD has
shown that significant amounts of HCl are produced from irradiated CSPE but the G-value for HCl
could not be determined accurately [51].

A discussion of G-values and recommendations for rubbers is provided in Section 5 and Section 6.

Table 10 G-values for rubbers at room temperature


Plastic details Radiation Environ- GH2 Data Comments
type ment (molecules source
per 100eV, [reference
except where number]
stated)
Neoprene alpha nitrogen GH2 0.4 [43] From Argonne, US.
GCO2 1.6 Typical of glove
materials used in
glove box and
radioactive waste
cleanup operations
2 kGy/hr dose. Only
thin surface layer
defined by range of
alpha particle
considered.
Neoprene alpha air GH2 1.6 [43] From Argonne, US.
GCO2 2.9 Typical of glove
materials used in
glove box and
radioactive waste
cleanup operations
2 kGy/hr dose. Only
thin surface layer
defined by range of
alpha particle
considered.
Neoprene alpha oxygen GGAS 0.03 [22] 95% H2, 3% CO2,
depleted GH2 0.03 1% CO, 1% CH4
Thickness of alpha
particle penetration
of sample not stated.
Neoprene alpha oxygen GGAS <0.1 [22] Thickness of alpha
depleted GH2 <0.1 particle penetration
of sample not stated.
Chlorosulphionated alpha oxygen GGAS 0.15 [22] Thickness of alpha
polyethylene depleted GH2 0.15 particle penetration
CSPE, Hypalon of sample not stated.
Chlorosulphionated alpha oxygen GGAS <0.1 [22] Dry box gloves
polyethylene depleted GH2 <0.1 Thickness of alpha
CSPE, Hypalon particle penetration
of sample not stated.

AMEC/200615/001 Issue 3 Page 40 of 77


Plastic details Radiation Environ- GH2 Data Comments
type ment (molecules source
per 100eV, [reference
except where number]
stated)
Latex gloves alpha oxygen 0.4 [54] Thickness of alpha
depleted particle penetration
of sample not stated.
Isoprene (artificial alpha oxygen 0.7 [54] Thickness of alpha
latex) gloves depleted particle penetration
of sample not stated.
Neoprene gamma vacuum GGAS 0.2 [22] 35% H2, 16% HCl,
GH2 0.1 3% CO, 43% CO2,
3% SO2
Neoprene gamma vacuum GGAS 0.3 [22] 29% H2, 17% HCl,
GH2 0.1 1% CO, 50% CO2,
3% SO2
Neoprene gamma oxygen GGAS 0.6 [22] 6% H2, 7% HCl, 8%
GH2 <0.1 CO, 79% CO2
Neoprene gamma oxygen GGAS 0.7 [22] 17% H2, 9% HCl,
GH2 0.1 9% CO, 58% CO2,
1% CH4, 6% SO2
Chlorosulphionated gamma vacuum 0.3 [22] 50% H2 in evolved
polyethylene gases
CSPE, Hypalon
Chlorosulphionated gamma air 0.3 [22] 50% H2 in evolved
polyethylene gases
CSPE, Hypalon
-1
Chlorosulphionated gamma air/vacuum 0.3 [51] 4 kGyhr
polyethylene
CSPE, Hypalon
-1
Chlorosulphionated gamma air/vacuum 0.1 [51] 45 Gyhr
polyethylene
CSPE, Hypalon
Polyisobutylene gamma air 1.3-1.6 [22]
Polyisobutadiene gamma air/vacuum 0.5 [54]

4.7 Review of G-values for ion exchange resins


The main organic ion exchange resins described within the 2013 UK RWI are a styrene
divinylbenzene copolymer or a phenol-formaldehyde based resin (Section 3.2.5). It is noted that the
ion exchange resins comprise around 3 % of the total mass of organic waste materials in the 2013
UK RWI.

Ion exchange resins typically consist of chemically active ion exchange groups attached to the
backbone of a polymer host material. The ion exchange functionality of the resin is controlled by
the type and nature of the chemically active groups that are bound to the host. A wide range of ion
exchange resins is available, with a wide range of active groups and host materials. The host
materials are normally organic polymer materials but they can also be inorganic.

AMEC/200615/001 Issue 3 Page 41 of 77


As ion exchange materials cover a very wide range of different material configurations, their
radiation resistance is expected to vary considerably and therefore radiolytic G-values for gas
generation are expected to vary. In the UK, the ion exchange resins that are likely to be present in
radioactive waste streams fall into four generic classes:
 Cation phenol-formaldehyde resins, either granular or beads;
 Acidic cation resins – acidic ion exchange groups bound to styrene-divinylbenzene (S-DVB)
resin beads;
 Basic anion resins – alkaline ion exchange groups bound to S-DVB resin beads;
 Synthetic Zeolite inorganic resins.

A number of reviews of data describing G-values from gamma radiation for relevant ion exchange
resins are available [57, 58], with details provided in Table 11. In addition G-values for gamma and
alpha irradiation experiments carried out on the same ion exchange resins samples have been
included (Table 11).

Reference [57], describes G-values for the generic classes of ion exchange resin that are expected
to be found in UK waste streams. References [13] and [59] review the radiolysis of S-DVB ion
exchange materials which are mainly encountered in the USA, however these data have been
included in the absence of G-values for alpha irradiation of expected UK ion exchange materials
and therefore provide some relevant G-value data for alpha radiolysis of ion exchange materials,
along with their measured G-values for gamma radiolysis for comparison.

Ion exchange resins in the nuclear industry are often used to remove highly radioactive gamma
emitting isotopes such as caesium-137 and cobalt-60 from aqueous solution. When they are spent
and stored they will “self irradiate”, mostly by gamma radiation (with a smaller contribution from
beta radiation), which provides an additional uncertainty on the experimentally determined
G-values.

Initially, ion exchange resin wastes are likely to be stored under water such that the resin is
immersed and all of the interstitial space between the beads or granules and any porosity is
expected to be filled with water. Reference [57] presents upper-bound GGAS-values for phenol
formaldehyde, cation S-DVB and anion S-DVB resins respectively, based on the assumption that
the resins will either be dry or contain limited amounts of water. However, reference [58]
demonstrates that there is likely to be an effect on the measured G-value for a proprietary S-DVB
+
resin (Dowex S-DVB H form) under different water loadings. For example, GGAS for the Dowex
+
S-DVB H form resin increases with water content from 0.061 molecules/100eV for a dry resin to
2.11 molecules/100eV for resin immersed in water, an increase of around one and a half orders of
+
magnitude. Comparison of the GH2-value of 1.7 molecules/100eV for the Dowex S-DVB H form
resin immersed in water against the typically accepted value of 0.45 molecules/100eV for
water [21] suggests that ion exchange resins and their pore water should be assigned one G-value,
rather than considering the G-values of the fraction of water contained within a mass of (dry) ion
exchange resin. Currently the radiolytic mechanisms that generate gas within a saturated ion
exchange resin are not fully understood but the uncertainty is expected to be bounded by the
available measured G-value data.

References [13, 59] show that the reported G-values for alpha radiolysis for an individual ion
exchange material are significantly higher than those for radiolysis by gamma, and provide
evidence that yields of hydrogen gas are (in this case, slightly) dependent on the amount of water
present and the counter anion.

A discussion of G-values and recommendations for ion exchange resins is provided in Section 5
and Section 6.

AMEC/200615/001 Issue 3 Page 42 of 77


Table 11 G-values for ion exchange media
G-values
Ion exchange Radiation Environ- Data
(molecules per Comments
resin type ment source
100eV)
S-DVB cation
Air/drained/ Assumed dry
(generic) gamma GGAS 0.25 [57]
flooded upper bound value
various forms
S-DVB
Air/drained/ Assumed dry
anion(generic) gamma GGAS 0.6 [57]
flooded upper bound value
various forms
+
S-DVB H form GGAS 0.061
(Dowex 50W gamma Dry GH2 0.026 [58]
x10) GCO2 0.035
+ GGAS 0.085-0.16
S-DVB H form
12%-47% GH2 0.04 to 0.12
(Dowex 50W gamma [58]
water GCO2 0.023
x10)
GCO 0.022
+
S-DVB H form GGAS 0.114
(Dowex 50W gamma Swollen GH2 0.095 [58]
x10) GCO2 0.019
+
S-DVB H form GGAS 2.11
Flooded in
(Dowex 50W gamma GH2 1.7 [58]
water
x10) GCO2 0.41
Phenol-
formaldehyde
Air/drained/ Assumed dry
cation gamma GGAS 0.17 [57]
flooded upper bound value
(generic)
various forms
Inorganic ion
exchange host Air/drained/ Assumed dry
gamma GGAS 0.1 [57]
materials (e.g. flooded upper bound value
Zeolite)
Sample dried under
vacuum then water
Amberlite
loaded.
(quaternary
Dose rate of 63.1
ammonium
alpha Water GH2 0.27 [13] Gy/min.
salt based on
Only thin surface
the S-DVB
layer defined by
backbone)
range of alpha
particle considered.
Room temperature
Reillex Sample dried under
(quaternary vacuum then water
ammonium loaded.
alpha Water GH2 0.19 [59]
salt based on Only thin surface
the S-DVB layer defined by
backbone) range of alpha
particle considered.

AMEC/200615/001 Issue 3 Page 43 of 77


G-values
Ion exchange Radiation Environ- Data
(molecules per Comments
resin type ment source
100eV)
Amberlite
(quaternary
Sample dried under
ammonium
gamma Water GH2 0.07 [13] vacuum then water
salt based on
loaded
the S-DVB
backbone)
Reillex
(quaternary 30°C, 100 Gy/min
ammonium Sample dried under
gamma Water GH2 0.07 [59]
salt based on vacuum then water
the S-DVB loaded
backbone)

4.8 Review of G-values for ‘Other Organics’


The description provided in the 2013 UK RWI for ‘Other Organics’ gives little information on the
types of materials expected under this category except for waste stream 2D27/C that contains
‘Triphenyl phosphonium bromate encapsulated in cement’ and accounts for 22% of the overall
‘Other Organics’ mass. However no G-value information is available for this salt.

Further investigation of the material breakdown of the waste streams that contribute the greatest
mass to this category, notably waste stream 2D90 that contributes 26% of the overall ‘Other
Organics’ mass, would need to be carried out before appropriate G-value information specific to
these wastes can be identified. However, it is noted that the mass of ‘Other Organics’ is small
(about 6 %) relative to the total mass of organic material categories and therefore identification of
G-values for these wastes is considered to be a low priority.

4.9 Review of G-values for ‘small organic molecules’


Some small organic molecules, mainly isosaccharinic acid (ISA), are expected to generate
methane or carbon dioxide gases as a result of radiolysis. However directly applicable G-values for
ISA could not be found. Instead, G-values for acetic acid (CH3COOH) have been presented for the
degradation of small organic molecules to CO2 and CH4 (Table 12). Acetic acid is a small organic
molecule that is expected to be one of the intermediate species formed in the degradation of small
organic molecules to gas.

Table 12 G-values for acetic acid as an analogue for ‘small organic molecules’
G-values
Carboxylic Radiation Environ- Data
(molecules per Comments
acid type ment source
100eV)
GCO2 4.0
GCH4 1.4 Thickness of alpha
Acetic acid alpha vacuum GCO 0.4 [22] particle penetration of
GH2 0.5 sample not stated.
GGAS 7.2
GCO2 5.4
GCH4 3.9
Acetic acid gamma vacuum GCO 0.2 [22]
GH2 0.5
GGAS 10.5

AMEC/200615/001 Issue 3 Page 44 of 77


4.10 Review of G-values for encapsulation materials
This section reviews available G-value data for cementitious and polymeric encapsulation
materials.

4.10.1 Cementitious encapsulants


Many of the wastes expected to be contained within a GDF are expected to be encapsulated within
a cementitious grout matrix. This usually involves intimately mixing the waste with a wet
cementitious encapsulant and allowing the mixture to set, although it is noted that some wastes
such as PCM (plutonium contaminated material) packages, or flood grouted wastes, may not be
intimately mixed in such a manner. The cementitious encapsulant contains pore water which will
generate gas when exposed to radiation. The amount of gas generated may depend on how much
water is retained by the cementitious encapsulant after curing. It is noted that cementitious
encapsulant will be affected by alpha, beta and gamma radiation within a container, while
cementitious backfill (e.g. NRVB) would also be affected by gamma radiation external to a
container. This section primarily focuses on cementitious encapsulant, although it is recognised
that the information may be applicable to cementitious backfill and this is discussed briefly in
Section 5.1.9.1.

Relevant G-value data for cementitious encapsulants are provided in Table 13. Generally, the
observed G-value data are similar, irrespective of water content. For example, examination of
recent data, considered to be carried out under experimental best practice, ([62], Table 6) show
that some cementitious encapsulant samples that are assumed to have low water content (ground
into granules and under vacuum with no obvious surface water, although the samples are expected
to contain some pore water) still generate hydrogen gas and have a G-value for gamma radiation
of around 0.1 molecules/100eV. Compare this with intimately mixed cemented sludge samples that
are expected to contain a high water content but give similar G-values for gamma radiation
(0.13 molecules/100eV, [63]). No data are available for cementitious encapsulant containing other
compounds such as plasticisers.

Note that the higher value of 0.8 molecules/100eV measured by [67] reduces to a lower G-value in
the presence of nitrate, which is consistent with the evidence for water; that is, the presence of
-
scavenger molecules such as NO3 , which react with the hydrated electron and its precursors,
reduce the production of hydrogen gas, and therefore reduce the G-value (see Section 4.2).

Taking into account the evidence that low water content cementitious encapsulants are still able to
produce hydrogen gas, and the fact that scavenger molecules that may be present within any
cementitious pore water can reduce the expected G-value, it is suggested that cementitious
encapsulants and their pore water should be considered as one entity for gas generation
calculations. It is suggested that gas generation by irradiation of cementitious encapsulants is more
complex than the current approach of the gas generation calculations that simply considers the
G-value for the fraction of water contained within a cementitious encapsulant. Further discussion of
this issue is provided in Section 5.4.

Some work has been carried out to understand the mechanisms of the complex radiolytic
processes in cementitious encapsulants and their pore waters for gamma radiation, e.g. [60] that
considers an Ordinary Portland Cement (OPC) in contact with its pore solution.

A discussion of G-values and recommendations for cementitious encapsulants is provided in


Section 5 and Section 6.

AMEC/200615/001 Issue 3 Page 45 of 77


Table 13 G-values for cementitious encapsulant materials irradiated in air unless
otherwise stated
Cementitious Radiation GH2 Data source Comments
encapsulant type (molecules [reference
materials per 100eV) number]
Portland Cement alpha 0.23 [65] Thickness of alpha particle
penetration of sample not
stated.
Envirostone alpha 0.25 [61] Envirostone is a
cementitious filler.
Only thin surface layer
defined by range of alpha
particle considered.
Various grout alpha 0.4 [64] Thickness of alpha particle
formulations penetration of sample not
stated.
OPC/PFA* gamma 0.1 [62] Sample ground into
(vacuum) granules, under vacuum and
no surface water obvious
OPC/PFA* 0.8v/w% gamma 0.1 [62] Sample ground into
superplasticiser (vacuum) granules, under vacuum and
no surface water obvious
OPC/BFS** gamma 0.1 [62]
Portland cement gamma 0.14-0.22 [63] Cured mortar containing
mortar water
Cement solidified gamma 0.11-0.35 [63] Portland cement
LLW
Sludge LLW gamma 0.13 [63] Winfrith samples, sludge
encapsulated in expected to contain high
cement water content
Various grout gamma 0.02 [64]
formulations
Portland Cement gamma 0.1 -0.16 [65]
Cement gamma 0.35 [54]
Portland cement gamma 0.11-0.35 [54]
High alumina cement gamma 0.25-0.29 [54]
OPC/BFS 0.8v/w% gamma 0.1 [62]
superplasticser
Cement grout gamma 0.15 [66]
Cement gamma 0.8 [67] Reduced to very low
G-values after addition of
nitrate
* OPC is ordinary Portland cement, PFA is pulverised fly ash
** OPC is ordinary Portland cement, BFS is blast furnace slag

AMEC/200615/001 Issue 3 Page 46 of 77


4.10.2 Polymeric encapsulants
The polymer encapsulant described in the 2013 UK RWI is the DOW WSB 101 polymer (a modified
vinyl ester styrene resin also known as VES, see Section 3.2.3), however it is expected that other
polymer encapsulants may also be used. Polymer encapsulants can have low permeability and
therefore can provide a barrier to water ingress that may reduce corrosion of metals or reduce gas
escape of e.g. radon gas. They are currently under consideration for the encapsulation of wastes
such as chemically active metals and radon-containing wastes, e.g. [68].

G-values for likely polymer encapsulant materials, including VES, are provided in Table 14.
G-values for other potential encapsulant materials such as epoxy resins, bitumen and asphalt are
also included. No G-value information for alpha radiation is available for any likely polymer
encapsulant material.

A discussion of G-values and recommendations for polymeric encapsulants is provided in


Section 5 and Section 6.

Table 14 G-values for polymeric encapsulants at room temperature.


Polymeric Radiation Environment GH2 (molecules Data Comments
encapsulant type per 100eV) source
Cyanate gamma vacuum 0.18 [69]
ester/epoxy resin
Epoxy resin gamma air/vacuum 0.25 [70] Data effectively
(Alchemie UK) independent of
atmosphere
Epoxy resin (low gamma air/vacuum 0.3 [70] Data effectively
temperature independent of
version Alchemie atmosphere
UK)
Epoxy resin gamma air/vacuum 0.4 [71] Data effectively
(Huntsman UK) independent of
atmosphere
Epoxy Resin gamma Air/vacuum 0.5 [71] Data effectively
(APS Diversified independent of
Technologies atmosphere
USA)
Epoxy resin gamma Air 0.3 (total gas) [56]
Epoxy resin gamma Air 0.1-0.23 [72]
Vinylesterstyrene gamma Air 0.03 [73]
(VES)
-1
Vinylesterstyrene gamma Air/vacuum 0.05 [71] 4 kGy hr
(VES) 10 MGy
-1
Vinylesterstyrene gamma Air/vacuum 0.08 [71] 34 Gy hr
(VES) 150 kGy
Bitumen gamma Air 0.2-0.4(total [56]
gas)
Asphalt gamma Air 0.3-0.6 [73] Various asphalt
compositions

AMEC/200615/001 Issue 3 Page 47 of 77


5 Discussion
This section discusses the G-value data that were reviewed in Section 4 and, based on this
information, suggests best estimate and upper bound G-values for water and sensible groupings of
organic materials. A description of data gaps and uncertainties is provided in Section 5.2, along
with suggestions for development of the SMOGG tool (Section 5.3) and a qualitative discussion of
the effect that using the suggested G-value data would have on the most recent gas generation
assessment (Section 5.4). A summary table of the G-values for each category of materials on
which future gas generation calculations may be performed is presented in Section 6.

5.1 Recommended G-values


Given the range of G-value data typically observed for water and each material type for each
radiation type (Section 4), it was decided that a best estimate G-value and an upper bound G-value
is more appropriate than assigning an individual G-value for each category. Uncertainties in the
experimental data are generally expected to be smaller than the ranges of values measured, so will
also be implicitly accounted for using this approach.

Note that G-values for alpha radiation have been suggested based on the assumption that the
thickness of the penetration of the alpha particle (not the total sample mass) has been used in the
calculation of the observed G-value, even if the penetration thickness has not been reported in the
available experimental details.

For beta radiation for water and for each material type, the recommended G-value has been based
on available data. Therefore, due to a lack of measurements of G-value for beta radiation, in many
cases the G-value for beta radiation is taken to be the same as for gamma radiation (because beta
and gamma radiation have similar LET). However, it is recognised that G-values for beta radiation
in all media will be strongly dependent on the energy of the ionising electron (or positron).

5.1.1 Discussion of G-values to use for water


Initially, water is expected to be present within wastes and any cementitious encapsulants, and it is
expected that a GDF will become fully saturated with groundwater at some time after closure of the
GDF, depending on the geological environment. This section discusses the G-values that may be
applied to pure water, with a further discussion of G-values that may be applied to a cementitious
encapsulant and its pore water provided in Section 5.1.9. It is recognised that most water present
in a GDF cannot be considered ‘pure’ since it is expected to contain some chemical constituents
from ingressing groundwater or be modified by its surrounding environment such as in the case of
a cementitious pore water.

The available G-value information for water is generally comprehensive and consistent. The only
gas produced in the radiolysis of pure water is hydrogen. The generally accepted G-value for the
production of H2 gas for gamma irradiation of pure water is 0.45 molecules/100eV, which is
consistent with the mean of 0.44 molecules/100eV of the data presented in Table 4. There is a
slight expected increase in G-value with increasing temperature within the likely temperature range
(up to 100°C) (see Section 4.2). However, scavenging of the precursors to hydrogen (the hydrated
electron and its precursor the dry electron) can lead to a decrease in hydrogen yields (see
Section 4.2), as could be expected in any pore water contained within waste, cementitious
encapsulant or cementitious backfill. It is suggested that a best estimate G-value for gamma
irradiation of water of 0.45 molecules/100eV is reasonable, with an upper bound of
0.5 molecules/100eV to include any variation due to temperature fluctuation up to 100 °C.
-
The G-value for the production of gas from beta radiation (e ) is rarely measured. Instead it is
considered that since beta and gamma radiation have similar LET, the G-value for beta radiation is
taken to be the same as for gamma radiation by analogy.

A typical G-value for the production of H2 gas for 5 MeV helium ions from pure water is
1.2 molecules/100eV, which is consistent with the mean of the data presented in Table 5. Results
expected with alpha particles are generally considered to be equivalent to those found with

AMEC/200615/001 Issue 3 Page 48 of 77


5 MeV helium ions, which are accelerator based and are considered far easier to perform accurate
experiments (see Section 4.2). Therefore it is suggested that a best estimate GH2-value for alpha
radiation of 1.2 molecules/100eV is reasonable, with an upper bound of 1.7 molecules/100eV
based on the older experimental data presented in Table 5 that are thought to be less accurate.

It is noted that the recommended G-values for radiolytic production of hydrogen from water will be
reduced by the presence of scavengers of the precursors to hydrogen to give a lower G-value. For
example the presence of nitrate significantly reduces the radiolytic formation of hydrogen [17, 18].
As a result, it is recognised that the recommended G-values for water provide an upper bound, and
a full understanding of the chemical mechanisms would be required to reduce the uncertainty.

5.1.2 Discussion of G-values to use for cellulose


Cellulose is expected to be contained within the 2013 UK RWI as paper and cotton, or wood.

The G-value information for cellulosic materials presented in Table 6 is noted to be based on old,
potentially unreliable, measurements and there is significant uncertainty associated with these
data. For example, it is rare for the G-value of total gas from alpha radiation to be lower than that
from gamma radiation because alpha radiation with its higher LET tends to lead to more H atom
combination reactions that are the precursor to hydrogen gas formation. However, the collated data
for cellulose in Table 6 do not exhibit this trend. It is noted G-values for both wet and dry cellulosic
materials are provided in Table 6 for alpha radiation, and show similar results.

Cellulose is expected to generate CO and CO2 gases as well as H2 when irradiated, even under
anaerobic conditions. However, the cellulosic wastes described in the 2013 RWI are not expected
to contain C-14, and therefore the evolution of individual gases from cellulose does not need to be
explicitly modelled since the main issue is the generation of bulk (non-active) gas that acts as a
carrier gas for other gaseous radionuclides. As a result, the G-values presented here are
considered for the total gas generated, not for hydrogen or other individual gases. It is recognised
that any CO2 gas produced is likely to be subsequently removed from the gas phase within a GDF
by the carbonation reaction. Also, it is noted that generation rates of individual gases may be
important when considering the risks of flammable gases such as H2, O2, etc. If such effects are to
be considered, additional experimental data for specific gases will need to be obtained.

Based on the available information for cellulosic materials presented in Table 6, and noting the
uncertainty on the data, it is suggested that best estimate GGAS-value for gamma radiation is
4.8 molecules/100eV based on the mean of the gamma radiation values, with an upper bound of
6.2 molecules/100eV. For beta radiation, a best estimate GGAS-value of 4.0 molecules/100eV
based on the mean of the beta radiation values is suggested, with an upper bound of
6.0 molecules/100eV. For alpha radiation, a best estimate GGAS-value of 1.0 molecules/100eV is
reasonable based on the mean of the alpha radiation values, with an upper bound GGAS-value of
1.5 molecules/100eV. These data should be used with caution until more experimental data for
appropriate cellulosic materials is available.

5.1.3 Discussion of G-values to use for halogenated plastics (PVC and PTFE)
The halogenated plastic wastes described within the 2013 UK RWI that are expected to be present
within a GDF are polyvinyl chloride (PVC) and a mixture of PVC and polytetrafluoroethylene
(PTFE). Within a GDF, it is unlikely that any PVC waste will be in its pure form; instead it is
expected that most PVC will contain significant amounts of plasticising agents, fire retardants and
stabilisers. Equally, the actual composition of the PVC may not be reported in detail by the
manufacturer because of commercial secrecy about proprietary formulations.

The most abundant gas generated by irradiation of PVC is usually HCl, with production of hydrogen
typically around 10% of the total, although the action of some additives can suppress the
generation of HCl gas, which may mean that hydrogen could be the most abundant gas generated.
Assuming a GDF is partly or fully saturated with water, any gaseous HCl(g) produced is likely to
dissolve to form hydrochloric acid (HCl(aq)) and in this case need not be taken into account when
considering the bulk gas produced. It is noted that hydrochloric acid (HCl(aq)) may enhance
corrosion, despite pH buffering by the cementitious environment present in many wastes, and an

AMEC/200615/001 Issue 3 Page 49 of 77


assessment of the impact of HCl, generated by degradation of PVC in supercompacted PCM
waste, on the corrosion of the stainless steel containers is reported in reference [39].

It is expected that most PVC wastes contained within a GDF will be plasticised and stabilised, and
therefore only G-value data relating to such types of PVC materials has been considered to provide
the best estimate G-value. It is noted that there are GHCl-value data from gamma radiation for two
PVC wastes that may be expected to be contained within a GDF, however these data report a
G-value for the production of HCl(g) at 40 °C, not a G-value for the total bulk gas (Table 7). To
encompass the pre-closure and post-closure phases of a GDF, G-value data are considered for
experiments in all environments (air or vacuum). Therefore using the relevant GGAS-value
information for plasticised and stabilised PVC wastes collated in Table 7 for gamma radiation, and
assuming that any HCl(g) produced will dissolve within available pore water present during both pre-
closure and post-closure phases of a GDF, a best estimate G-value of 1.0 molecules/100eV is
suggested for the generation of bulk gas by gamma radiation based on the mean of available
GGAS-value data (after subtraction of the GHCl-value, and excluding the last four rows of Table 7 that
did not measure any gases other than HCl, or were carried out at high temperature).

A more cautious approach is considered for the upper bound G-value for gamma irradiation of PVC
wastes. An assumption has been made that any HCl(g) produced from PVC wastes will remain as a
gas and will not dissolve in the pore water during the pre-closure and post-closure phases of a
GDF, and that only G-values for plasticised and stabilised PVC wastes in air or under anoxic
conditions are considered. (It is noted that HCl(g) is very soluble in water, and it is unlikely that no
HCl gas will dissolve in the pore water and therefore this assumption is expected to provide an
upper bound G-value.) In addition, only G-value data measured at temperatures below 50°C were
considered to meet the guidance temperature of less than 50°C for all waste packages following
closure, as stated in the Generic Disposal System Technical Specification [74]. Therefore an upper
bound G-value of 5.8 molecules/100eV is suggested for the generation of bulk gas by gamma
radiation based on the maximum of available GGAS-value data for relevant PVC wastes.

Only one relevant G-value for production of HCl gas from pure PVC waste is available for beta
radiation, and therefore it is suggested that by analogy, the best estimate and upper bound
G-values for bulk gas assigned to gamma radiation for plasticised and stabilised PVC wastes are
used.

For alpha irradiation of plasticised and stabilised PVC, a best estimate G-value of
1.3 molecules/100eV is suggested for the generation of bulk gas based on the mean of available
GGAS-value data (after subtraction of the GHCl-value) and assuming that (i) all HCl gas produced
dissolves in pore water, and (ii) data for either air or anoxic environments are considered.

An upper bound G-value for alpha irradiation of plasticised and stabilised PVC of
2.8 molecules/100eV is suggested. This is based on the assumption that any HCl(g) produced from
PVC wastes will remain as a gas and will not dissolve in the pore water during the pre-closure and
post-closure phases of a GDF, and that only G-values for plasticised and stabilised PVC wastes in
air or under anoxic conditions are considered. (As before, it is noted that HCl(g) is very soluble in
water, and it is unlikely that no HCl gas will dissolve in the pore water and therefore this
assumption is expected to provide an upper bound G-value.) In addition, only G-value data
measured at temperatures below 50°C were considered to meet the guidance temperature of less
than 50°C for all waste packages following closure has been considered.

It is noted that some G-value data from gamma irradiation of stabilised and plasticised PVC wastes
are reported at higher temperatures (70 °C) with significantly higher ‘effective’ G-values (e.g. two
orders of magnitude higher). These ‘effective’ G-values are the result of radiolytic processes, and
thermal processes, as well as combined radiolytic and thermal processes and do not represent a
G-value derived solely from generation of gas by radiolysis. It is noted that any future consideration
of radiolysis at higher temperatures should take this into account.

It is noted that the production of HCl(g) by irradiation of PVC will have an impact on its environment
and may affect the pH of the local cementitious grout or cementitious backfill [39].

AMEC/200615/001 Issue 3 Page 50 of 77


For PTFE, a value of 0.1 molecules/100eV for the total gas generated from gamma radiation is
available. This value is within the lower range of possible G-values measured by irradiation of PVC,
and therefore it is suggested that the G-values used for PVC should be applied for any mixture of
PVC/PTFE containing wastes.

5.1.4 Discussion of G-values to use for non-halogenated plastics (typically


polyethylene and Perspex)
The typical non-halogenated plastics expected to be present within a GDF are polyethylene and
Perspex (also called polymethylmethacrylate or PMMA) (Section 3.2.3). (Note that the polymer
encapsulant also described within Section 3.2.3 will be considered within the polymeric
encapsulant discussion in Section 5.1.9.2.)

Polyethylene has been studied widely to help understand the basic radiation effects in polymers,
and therefore a substantial amount of data on its G-values is available. Typically, the most
abundant gas generated is hydrogen gas, especially under anaerobic conditions. Based on the
data collated in Table 8, the best estimate G-value for production of bulk gas by gamma radiation
under either oxic or anoxic conditions is suggested to be 3.1 molecules/100eV based on the mean
of available room temperature data. For the upper bound for gamma radiation, a G-value of
4.1 molecules/100eV is suggested, based on the maximum G-value for gamma radiation from all
reported polyethylene data. In the absence of any data for beta irradiation of polythene, the
G-values for gamma radiation are suggested to be used for beta radiation, by analogy.

The G-values collated in Table 8 for production of bulk gas from polyethylene by alpha radiation
range from 1.7 molecules/100eV to 7.3 molecules/100eV for oxic or anoxic conditions. It is
suggested that the lower range of measured G-values may be due to: (i) a reduction in the amount
of hydrogen gas produced by radiolysis of the polyethylene as a result of a high total dose. For
example, at a high total dose the amount of hydrogen gas produced may be reduced because the
amount of hydrogen available within the polyethylene material has already been depleted (by
analogy with the gamma radiation observation for polyethylene (see Section 4.5.1)); and/or (ii)
under anaerobic conditions, physical changes to the structure of the polyethylene as a result of
radiolysis may reduce the hydrogen diffusion rate (e.g. increased cross-linking of the material may
reduce its gas permeability), although it is noted that the overall effect would depend on the dose
rate and the total dose under the experimental conditions. In addition, for alpha radiation, the
amount of hydrogen diffusing out of the irradiated surface layer may be less than the total amount
of hydrogen gas generated because some gas may diffuse into the bulk material rather than
towards the surface of the material. The experimental dose rate can also have a pronounced effect
on the alpha radiolysis of polyethylene and lower dose rate data are preferred since they are more
similar to expected dose rates within a GDF.

It is suggested that a best estimate G-value of 4.2 molecules/100eV from alpha radiation based on
the mean of the available alpha radiation data (mean of the first four rows in Table 8; note that
equal weighting is given to the upper and lower values in the data range presented in the first row
of Table 8). An upper bound of 7.3 molecules/100eV, for the production of bulk gas from
polyethylene by alpha radiation is suggested, based on the maximum reported GGAS value. The
maximum GGAS value includes a contribution from both H2 and CO2 gases under oxic conditions.

It is noted that the presence and influence of additives contained within polyethylene may be
expected to affect the measured G-value.

There are some data available for Perspex (PMMA) described in Section 3.2.3. The experimental
data for Perspex show that a mixture of gases is produced when irradiated under vacuum or under
an oxygen-depleted environment, including hydrogen, carbon dioxide and carbon monoxide gases
(Table 9). Following the reasoning given for cellulose (Section 5.1.2), G-values for total gas are
suggested rather than for hydrogen gas. Based on the mean of the available G-value data for
Perspex from gamma radiation, a best estimate value of 2.2 molecules/100eV is suggested for total
gas, with an upper bound value of 4.1 molecules/100eV based on the maximum G-value for
gamma radiation for total gas. In the absence of any data for beta irradiation of Perspex, the
G-values for gamma radiation are suggested to be used for beta radiation, by analogy.

AMEC/200615/001 Issue 3 Page 51 of 77


Similarly, based on the G-values from alpha irradiation of Perspex it is suggested that a best
estimate value of 1.3 molecules/100eV for production of total gas should be applied, based on the
mean of available G-value data, with an upper bound value of 2.0 molecules/100eV, based on the
maximum reported G-value for production of total gas. The commonly accepted mechanisms
leading to hydrogen production suggest that the G-value should remain nearly the same or
increase with LET. However, other gases have different mechanisms for their production and it is
possible for their yields to decrease with increasing LET. Because it is rare for G-values from alpha
radiolysis to be less than from gamma radiolysis for aliphatic compounds, further experimental
evidence on the composition of the evolved gases and their G-values from Perspex is required.

G-value data are available for a range of non-halogenated plastics including polypropylene,
polystyrene, Nylon and PET, however the mass of these plastics present in the waste inventory is
relatively low and therefore these have not been considered in more detail here.

5.1.5 Discussion of G-values to use for rubbers


The rubbers or elastomers (a polymer that displays rubber-like elasticity) described in the
2013 RWI are expected to be neoprene (also known as polychloroprene, a halogenated rubber) or
styrene butadiene co-polymer (a non-halogenated rubber known as SBR). The most abundant gas
produced by irradiation of rubber materials is expected to be hydrogen, with a range of other gases
such as CO and CO2. For halogenated rubbers, significant amounts of HCl(g) are expected to be
produced, although the amount is rarely quantified by experimental data. Assuming the waste
package is partly or fully saturated with water, any amount of HCl(g) is expected to dissolve to form
hydrochloric acid (HCl(aq)).

The available G-value data for neoprene are presented in Table 10. To encompass the pre-closure
and post-closure phases of a GDF, G-value data are considered for experiments in either air or
anoxic environments. The available G-value data show that hydrogen (up to 35 %), CO2 (up to
79%) and HCl (up to 17 %) are the most abundant gases produced by gamma irradiation of
neoprene, although it is recognised that the amount of HCl(g) produced is rarely quantified by
experimental data. Therefore the approach to suggest a best estimate value for gamma irradiation
of neoprene has been to use the GGAS-value information for neoprene for gamma radiation, and
assume that any HCl(g) produced will dissolve within available pore water present during both
pre-closure and post-closure phases of a GDF. This best estimate approach gives a G-value of
0.40 molecules/100eV based on the mean of available GGAS-value data (after subtraction of the
fraction of G-value due to HCl(g)). It is also recognised that any CO2 gas produced is likely to be
subsequently removed from the gas phase within a GDF by the carbonation reaction.

A more cautious approach is considered for the upper bound G-value for gamma irradiation of
neoprene wastes. To provide this an extreme assumption has been made that any HCl(g) produced
from neoprene waste will remain as a gas and will not dissolve in the pore water during the
pre-closure and post-closure phases of a GDF. Therefore an upper bound G-value of
0.7 molecules/100eV is suggested for the generation of bulk gas by gamma radiation based on the
maximum of available GGAS-value data for neoprene wastes.

In the absence of any data for beta irradiation of neoprene, the G-values for gamma radiation are
suggested to be used for beta radiation, by analogy.

For alpha irradiation of neoprene, G-value data under oxic and anoxic conditions have been
considered to encompass the pre-closure and post-closure phases of a GDF. The available
G-value data show that both hydrogen and CO2 gases are produced. Based on the mean of
available GGAS-values, it is suggested that a best estimate value is 1.7 molecules/100eV for
production of total gas by alpha radiation, with an upper bound value of 4.5 molecules/100eV
based on the maximum GGAS-value measured for neoprene for alpha radiation. It is also recognised
that any CO2 gas produced is likely to be subsequently removed from the gas phase within a GDF
by the carbonation reaction.

For styrene butadiene co-polymer, no data are available. By analogy with other styrene containing
polymers such as polystyrene or vinylesterstyrene, it is expected that the G-value from gamma
radiolysis will be similarly low because the repeat structure of the polymer includes four benzene

AMEC/200615/001 Issue 3 Page 52 of 77


rings. Aromatic groups more commonly decay back to their ground state rather than decompose to
produce hydrogen gas. As a comparison the GH2-value for gamma irradiation of vinylesterstyrene is
a maximum of 0.08 molecules/100eV and of polystyrene is a maximum of 0.1 molecules/100eV,
and for alpha irradiation of polystyrene one G-value of 0.15 molecules/100eV is observed.

Therefore, it is suggested that by analogy with polystyrene and vinylesterstyrene, a best estimate
G-value for gamma radiation of 0.1 molecules/100eV is applied, with an upper bound G-value of
0.2 molecules/100eV, based on double the best estimate value. In the absence of any data for beta
irradiation of other styrene containing polymers, the G-values for gamma radiation are suggested to
be used for beta radiation, by analogy. For alpha radiation, a best estimate G-value of
0.15 molecules/100eV is applied by analogy with polystyrene. However, aromatic compounds are
known to have surprisingly high yields of hydrogen with high LET and therefore an upper limit
G-value of 4.6 molecules/100eV for alpha radiolysis is suggested in line with the maximum
observed G-value for alpha radiation for hydrogen gas from all organic materials.

5.1.6 Discussion of G-values to use for ion exchange materials


The main organic ion exchange resins described within the 2013 UK RWI are a styrene
divinylbenzene copolymer and a phenol-formaldehyde based resin. It is noted that the ion
exchange resins comprise only around 3 % of the total mass of organic waste materials and
therefore contribute a relatively small amount to the overall generation of gas by radiolysis.

Based on data for styrene divinylbenzene based resins (Table 11), a best estimate GGAS-value for
gamma radiation of 0.21 molecules/100eV is suggested. This is based on a review of styrene
divinylbenzene based resins under dry or limited water conditions, and assumes that ion exchange
resins are more likely to be packaged under dry or limited water conditions than immersed in water
(i.e. based on the mean of the GGAS-values for gamma irradiation of styrene divinylbenzene based
resins except the GGAS datum for the styrene divinylbenzene based resin flooded in water). An
upper bound G-value of 2.1 molecules/100eV for gamma radiation is suggested, based on the
maximum G-value from all styrene divinylbenzene based resins (including those immersed under
water). It is recognised that ion exchange resins may not be completely dry when packaged and
therefore the best estimate and upper bound GGAS-value for gamma radiation reflect the range of
uncertainty.

In the absence of any data for beta irradiation, the G-values for gamma radiation are suggested to
be used for beta radiation, by analogy.

For alpha radiation, only two G-value measurements for styrene divinylbenzene resins are
available based on measurements carried out to investigate ion exchange resins likely to be used
within the US nuclear programme. However, in the absence of other G-value data for alpha
radiation, a best estimate G-value of 0.23 molecules/100eV is suggested based on the mean of the
available two measurements. An upper bound G-value of 0.27 molecules/100eV is suggested
based on the maximum measured G-value for alpha irradiation of styrene divinylbenzene resins.

For phenol-formaldehyde based resins (Table 11), a best estimate GGAS-value for gamma radiation
of 0.17 molecules/100eV is suggested, based on one data point from the review of likely UK ion
exchange resins under dry or limited water conditions. An upper bound G-value of
2.1 molecules/100eV for gamma radiation is suggested, consistent with the styrene divinylbenzene
resin. In the absence of any data for beta irradiation, the G-values for gamma radiation are
suggested to be used for beta radiation, by analogy. There are no available data for alpha
irradiation of phenol-formaldehyde based resin and therefore the best estimate and upper bound
values for styrene divinylbenzene resin are suggested.

5.1.7 Discussion of G-values to use with ‘Other Organics’


There is little descriptive information provided in the 2013 UK RWI on the specific types of material
expected, except for ‘Triphenyl phosphonium bromate encapsulated in cement’ that accounts for
around one fifth of the ‘Other Organics’ mass. However no G-value information is available for this
compound.

AMEC/200615/001 Issue 3 Page 53 of 77


Given that ‘Other Organics’ account for a small proportion (6 %) of the overall mass of organic
wastes, and the lack of available G-value information, it is suggested that the mass of ‘Other
Organics’ should be added to the mass of one of the major categories when carrying out radiolysis
calculations. For example, the ‘Other Organics’ mass could be accounted for within the
polyethylene category of the non-halogenated plastics group, since the polyethylene category is
expected to be a significantly larger mass of waste, with G-values that are typically higher than
other organic waste materials.

5.1.8 Discussion of G-values to use for ‘small organic molecules’


Some ‘small organic molecules’, mainly isosaccharinic acid (ISA), are expected to be present in the
wastes as a result of degradation of cellulosic materials. The ‘small organic molecules’ are
expected to generate methane or carbon dioxide gases as a result of radiolysis. However directly
applicable G-values for ISA could not be found. Instead, G-values for acetic acid (CH3COOH) have
been presented for the degradation of small organic molecules to CO2 and CH4 (Table 12). Acetic
acid is a small organic molecule that is expected to be one of the intermediate species formed in
the degradation of small organic molecules to gas.

It is noted that other (non-radiolytic) processes such as microbial degradation are expected to be
the major source of gas generation, rather than radiolysis of small organic molecules (and
cellulose).

Based on the available information for acetic acid as an analogue for ‘small organic molecules’, it is
suggested that for gamma radiation a best estimate GCO2-value is 5.4 molecules/100eV and a best
estimate GCH4-value is 3.9 molecules/100eV based on the one available experimental datum, with
an upper bound of 10.5 molecules/100eV based on the GGAS-value for acetic acid for gamma
radiation for both CO2 and CH4 gases.

In the absence of any data for beta irradiation, the G-values for gamma radiation are suggested to
be used for beta radiation, by analogy.

For alpha radiation, a best estimate GCO2-value is 4.0 molecules/100eV and a best estimate
GCH4-value is 1.4 molecules/100eV based on the one available experimental datum, with an upper
bound of 7.2 molecules/100eV based on the GGAS-value for acetic acid for alpha radiation for both
CO2 and CH4 gases.

These G-values should be used with caution until experimental data for ISA or other ‘small organic
molecules’ are available.

5.1.9 Discussion of G-values to use for waste encapsulants


Most wastes are expected to be encapsulated, perhaps within a cementitious encapsulant such as
OPC/PFA, or OPC/BFS, or within a polymeric encapsulant such as a vinyl ester resin. In addition,
after emplacement in a GDF, the vaults and tunnels may be backfilled by a cementitious backfill
such as the Nirex Reference Vault Backfill (NRVB). G-values for these materials are considered in
the section below.

5.1.9.1 Cementitious encapsulants


Many of the wastes described in the 2013 UK RWI are expected to be mixed with a wet
cementitious encapsulant and the mixture allowed to set.

Based on the data presented in Table 13, the observed G-value data for cementitious wastes and
their pore water are typically similar, irrespective of water content. Taking into account the evidence
that low water cementitious encapsulants are still able to produce hydrogen gas, and the fact that
scavenger molecules that may be present within any cementitious pore water can reduce the
expected G-value, it is suggested that cementitious encapsulants and their pore water should be
considered as one entity for gas generation calculations. As a result a best estimate G-value for
production of hydrogen gas by gamma irradiation of cementitious encapsulant and its pore water is
suggested to be 0.2 molecules/100eV, based on the mean of available data. An upper bound of

AMEC/200615/001 Issue 3 Page 54 of 77


0.8 molecules/100eV is suggested, based on the maximum measured G-value for gamma
radiation.

In the absence of any data for beta irradiation, the G-values for gamma radiation are suggested to
be used for beta radiation, by analogy.

Based on the available G-values for alpha irradiation of cementitious encapsulant and its pore
water, a best estimate G-value of 0.3 molecules/100eV is suggested for the generation of bulk gas
based on the mean of available data, with an upper bound of 0.4 molecules/100eV.

In the absence of available data for cementitious backfill that is expected to be used to fill the
tunnels and vaults of a GDF, and given the similarity of the range of G-values for cementitious
encapsulants regardless of their composition, it is suggested that these G-value may also be
applied to cementitious backfill and its pore water.

5.1.9.2 Polymeric encapsulants


The polymeric encapsulant described in the 2013 UK RWI is a modified vinyl ester styrene resin
(VES), called DOW WSB 101.

Based on the available G-values for gamma irradiation of VES presented in Table 14, a best
estimate G-value of 0.05 molecules/100eV is suggested for the generation of bulk gas based on
the mean of available data, with an upper bound G-value of 0.08 molecules/100eV based on the
maximum G-value. In the absence of any data for beta irradiation, the G-values for gamma
radiation are suggested to be used for beta radiation, by analogy.

There are no available data for G-values for alpha irradiation of VES. By analogy with other styrene
containing polymers such as polystyrene where there is one reported G-value for alpha irradiation
of polystyrene of 0.15 molecules/100eV, it is suggested by analogy that a best estimate G-value of
0.15 molecules/100eV is applied for alpha radiation. However, aromatic compounds are known to
have surprisingly high yields of hydrogen with high LET and therefore an upper limit G-value of
4.6 molecules/100eV for alpha radiolysis is suggested in line with the maximum observed G-value
for hydrogen gas from all organic materials.

Table 14 also presents G-value data for epoxy resins that may be considered as an alternative
potential future polymer encapsulant; these typically have G-values for gamma radiation of around
an order of magnitude greater than VES.

It is recognised that gas may be generated within polymer encapsulated wastes by radiolytic
processes, or by corrosion if water is available. Any gas generated is expected to diffuse slowly
through the typically low permeability polymer encapsulant. It is assumed that the volume of gas
generated is unlikely to be sufficient to cause over-pressurisation issues.

5.2 Data gaps and uncertainties in the review of G-values


There are several data gaps where further information would aid the assessment of gas generation
by radiolysis, however it is recognised that the benefit of additional data needs to be considered
relative to the overall uncertainties given the current range of G-values, and the expected inventory
mass of the material in question.
 Only limited G-value data have been measured for cementitious encapsulant containing other
compounds such as plasticisers (e.g. [62]), and, given the substantial mass of cementitious
encapsulant expected within a geological disposal facility, it is suggested that G-value data for
relevant cementitious encapsulants containing appropriate plasticisers are obtained.
 Cellulosic materials expected to be present in the waste inventory will be contained within a
broad range of types of material (e.g. wood, paper, cotton) and may be present in a variety of
semi-crystalline structures which will affect the gas release by radiolysis. In addition, the
G-value information for cellulosic materials presented is based on old, potentially unreliable,
measurements and there is significant uncertainty associated with these data. For example,
the two G-values for beta irradiation of ‘American cotton’ show a substantial range that has

AMEC/200615/001 Issue 3 Page 55 of 77


been determined under very different dose conditions and further data under an appropriate
range of dose conditions would help to reduce this uncertainty. Similarly, the G-values
measured for alpha irradiation of cellulose are lower than the G-values for gamma irradiation
of cellulose, which is rare; although it is noted that the thickness of the alpha particle
penetration has not been stated in the experimental detail. The current inventory of cellulose
described by the 2013 UK RWI is around 20 % of the total mass of organic materials (Table 2).
Given its expected mass contained in the 2013 UK RWI, and the reasonably high best
estimate recommended G-values for gamma and beta radiation relative to, for example, water,
it is suggested that further experimental data for relevant cellulosic materials that are expected
to be present within UK wastes should be obtained to reduce the uncertainty.
 G-values from PVC wastes are expected to be very variable due to the very wide range of type
and quantity of any plasticising agents, fire retardants and stabilisers that may be contained
within the polymer. Two G-values for PVC materials that are expected to be destined for a UK
GDF have been measured (Table 7, [40]), but G-values for additional PVC wastes that are in
use and are destined for a UK GDF should be investigated to extend the current
understanding. The recommended upper bound G-value for gamma irradiation of PVC is
suggested as a value 6.0 molecules/100eV, to encompass the uncertainty in G-values of PVC
wastes, although the best estimate value is 1.0 molecules/100eV. The current inventory of
PVC described by the 2013 UK RWI is around 32 % of the total mass of organic materials
(Table 2). Given the uncertainty, the limited G-value data for UK specific PVC wastes, and the
substantial mass of PVC described by the 2013 UK RWI, further investigation of G-values for
specific PVC wastes should be considered. It is noted that any increase in the expected upper
temperature within a GDF would significantly affect the G-value for PVC due to the combined
effects of radiolytic and thermal processes.
 The available G-value data for Perspex (PMMA) for alpha radiation is lower than for gamma
radiation. It is rare for G-values from alpha radiolysis to be less than from gamma radiolysis for
aliphatic compounds and therefore it is suggested that G-value data for Perspex should be
investigated further to reduce the uncertainty. However, it is recognised that Perspex forms a
small proportion (7%) of the overall mass of organic materials (Table 2) and therefore
additional G-value data for Perspex is not a key priority.
 No reliable data have been measured for the non-halogenated rubber, styrene butadiene co-
polymer and therefore an experimental study is suggested. However, by analogy with
radiolysis of other benzene containing compounds, its expected G-value is likely to be low, and
its inventory in the 2013 UK RWI is small (2 %) relative to the overall mass of organic
materials, and therefore it is not a key priority for further investigation.
 No alpha or beta radiation G-value data have been measured for PTFE. The current inventory
of PVC described by the 2013 UK RWI is only around 0.6 % of the total mass of organic
materials (Table 2). Therefore, although not considered to be a key priority for further
investigation, a simple study of radiolysis of PTFE to quantify the amount of HF and H2
released by irradiation is suggested, including consideration of the potential for HF production
within the wastes.
 There are no available alpha radiation G-value data for polymer encapsulants such as VES.
However, further investigation of G-value data for VES should only be carried out when there
is knowledge of the expected mass of VES to be contained within a GDF since its current
contribution relative to other sources of gas generation by radiolysis is expected to be small.
 No G-value data for CO2 and CH4 gases have been measured for small organic molecules,
(mainly expected to be ISA). G-values based on acetic acid for CO2 and CH4 have been
presented as an analogue instead. Although the G-value data for acetic acid are based on only
one measurement, the data are fairly consistent with G-values reviewed for cellulose for the
generation of H2 and CO2 gases (but not CH4 gas). As a result, the G-value data for CO2 and
CH4 gases for small organic molecules should be used with caution until experimental data for
ISA or other ‘small organic molecules’ are available. Based on this lack of appropriate data it is
suggested that an experimental study should be considered to provide G-value data for CO2
and CH4 gases for relevant small organic materials. However, given that small organic
molecules are expected to be formed as degradation products from cellulose, and that other
processes such as microbial degradation, rather than radiolysis, are expected to be the major

AMEC/200615/001 Issue 3 Page 56 of 77


source of gas generation from cellulose and small organic molecules, it is considered that this
is not a key priority for further investigation.

The main uncertainties are:


 Experimental determinations of G-values are typically carried out at much higher dose rates
than the wastes are expected to experience within a GDF. For example, the range of expected
-5 -1
dose rates for a waste stream containing organic materials is between 1.6×10 Gy hr
-1 -3 -1
(minimum) and 1.2 Gy hr (maximum), with a median of 1.2×10 Gy hr , for the external
gamma dose rate at the surface of the waste package at 2040 AD (from the 2013 UK
RWI [10]), This is lower than the dose rates used to obtain most recent experimental data
-1
(recent experiments typically irradiate at around 1 to 10 kGy hr ). Older experimental data are
typically based on irradiations at even higher dose rates.
 Experimental evidence is needed to understand the change in the rate of gas generated by
radiolysis with time. Depending on the material, there is evidence that the generation of gas
may depend on the dose rate (e.g. Section 4.5.1). Although upper bound G-values have been
suggested for each material to encompass the uncertainties, it is recognised that available
measurements of G-values do not provide sufficient data carried out over long timescales at
the low dose rates expected within a GDF.
 It is noted that estimates of gas production can be made according to the composition of
relevant groups of waste material, but reliable data will only come from experimental studies of
specific materials expected to be disposed of within a GDF. Also, gas yields by alpha particle
radiolysis are much more difficult to obtain than for gamma radiation and only very few
laboratories worldwide can perform these studies and therefore alpha particle radiolysis
measurements of relevant materials can be limited in number.
 For alpha radiation, experimental details have been included in Section 4, if available, for the
depth of sample irradiation by alpha particles. However, it is noted that this information is not
always provided and that the use of the whole sample mass, rather than only the thin layer
irradiated by alpha particles would lead to an under-estimation of the reported G-value for
alpha radiation for an individual material.

5.3 Requirements for development of the SMOGG tool based


on the outcomes from the review
Based on the waste categories and corresponding G-values recommended for calculations of gas
generation due to radiolysis, as discussed in Section 5.1, it is clear that the SMOGG tool will
require development to:
 Include separate G-values for each type of radiation for water and each of the different organic
materials listed (see summary in Table 3).
 Assign a single G-value for each type of radiation for cementitious grout and its pore water
regardless of its state of water saturation, rather than the current simple approach that
considers a G-value for the fraction of (pure) water contained within the cementitious
encapsulant. It is suggested that cementitious backfill and its pore water could similarly be
assigned a single G-value for each radiation type.

It is recognised that calculations of gas generation due to radiolysis also depend on the fraction of
the radiolytic energy absorbed by a material (see Appendix 1). In the current SMOGG tool a simple
mass fraction scaling of the total radiolytic energy is used to estimate the absorption by each
material. In practice the energy absorbed will depend on the distribution of the materials within a
waste package. For some calculations for individual packages it may be important to account for
this, but for large-scale calculations (e.g. for a whole vault in a GDF) the simple approach currently
used in the SMOGG tool is considered to be appropriate, so no change to this is recommended.

AMEC/200615/001 Issue 3 Page 57 of 77


5.4 Demonstration of the effect that filling any knowledge
gaps would have on gas generation rates currently
recorded in the Environmental Safety Case
This section qualitatively discusses the effect that using the suggested material categories and
corresponding G-value data (Section 5) would have on gas generation rates presently recorded in
the Environmental Safety Case.

The use of the proposed G-values for gas generation by radiolysis (Section 5, and summarised in
Section 6) is expected to have the following effects:
 Any water is expected to give similar gas generation rates to the existing calculation of gas
generation for all types of radiation using best estimate G-values proposed in this report
(Section 5.1.1), compared to the G-values applied in the gas generation assessment that used
the 2007 Derived Inventory (Table 1). (Note that the gas generation rate from cementitious
pore water (i.e non-pure water) is considered within the next bullet point.)
 The use of a single best estimate G-value for gas generation by radiolysis of cementitious
grout and its pore water is expected to result in a higher gas generation rate for all radiation
types. This suggested approach contrasts with the previous approach used in the gas
generation results that used the 2007 Derived Inventory and considered a G-value for gas
generation by radiolysis of (pure) water, and did not use a G-value for the cementitious solid
encapsulant nor its pore water (i.e. water containing some aqueous species as a result of
interaction with the cementitious solid phases). For example, the gas generation results that
used the 2007 Derived Inventory for a post closure scenario apply a porosity of 0.2 for
cementitious encapsulant with full water saturation, using a G-value for gamma irradiation of
(pure) water of 0.5 molecules/100eV. Therefore, using a mass fraction approach and
-3
assuming a density of around 2000 kg m for the solid phase would give an overall effective
G-value of approximately 0.05 molecules/100eV for gamma radiation for the gas generation
results that used the 2007 Derived Inventory. (Note that in a pre-closure scenario, the previous
approach that used the 2007 Derived Inventory (where limited water is available) would give a
lower overall effective G-value due to the reduced availability of water within the pore space.)
By comparison, the approach suggested as a result of this study uses a single best estimate
G-value for gamma radiation of 0.2 molecules/100eV for a block of cementitious encapsulant
and its pore water, which is therefore higher by a factor of about four than the previous
approach that used the 2007 Derived Inventory.
 For cellulosic materials, the use of the proposed best estimate G-values for gas generation by
radiolysis (Section 5.1.2), compared to the G-values applied in the gas generation assessment
that used the 2007 Derived Inventory (Table 1), would be expected to increase the gas
generation rate from gamma radiation and reduce the gas generation from beta radiation,
while alpha radiation gas generation rates would be expected to be similar.
 For plastic wastes, separate G-values have been proposed for several categories of plastic
materials (non-halogenated plastics, halogenated plastics, ion exchange resins and rubbers)
and therefore should provide an improved description of the gas generation rates by radiolysis.
Overall, the use of proposed best estimate G-values and the new categories of plastic
(Table 3) is expected to reduce the gas generation rate (compared to the G-values applied in
the gas generation assessment that used the 2007 Derived Inventory):

- For halogenated plastics, use of the proposed best estimate G-values given in
Section 5.1.3 (compared to the previously applied G-values, Table 1) is expected to
significantly reduce the gas generation rate for all types of radiation. For halogenated
plastics, use of the proposed upper bound G-values is expected to give similar gas
generation rate for all types of radiation.
- For non-halogenated plastics, use of the proposed best estimate G-values for
polyethylene given in Section 5.1.4 (compared to the previously applied G-values,
Table 1) is expected to give lower gas generation rates for gamma and beta radiation, and
slightly higher gas generation rates for alpha radiation. For polyethylene, use of the
proposed upper bound G-values for polyethylene is expected to give lower gas generation

AMEC/200615/001 Issue 3 Page 58 of 77


rates for gamma and beta radiation, and much higher gas generation rates for alpha
radiation. For Perspex (PMMA), use of the proposed best estimate or upper bound
G-values for Perspex is expected to give lower gas generation rates for all types of
radiation.
- For rubbers, the use of proposed best estimate G-values given in Section 5.1.5 (compared
to the previously applied G-values, Table 1) is expected to significantly reduce the gas
generation rate for all types of radiation. For rubbers, the use of upper bound G-values is
expected to reduce the gas generation rates for gamma and beta radiation, and increase it
for alpha radiation.
- For ion exchange resins, use of the proposed best estimate or upper bound G-values
given in Section 5.1.6 (compared to the previously applied G-values, Table 1) is expected
to significantly reduce the gas generation rate for all types of radiation.
 The inclusion of polymer encapsulant is an additional category for which a G-value has been
assigned. The proposed G-values for gas generation by gamma and beta radiolysis are small,
and therefore this is expected to be a small addition to the overall gas generation rate.
 Within the 2013 UK RWI, the category of ‘Other Organics’ accounts for a small proportion
(6 %) of the overall mass of organic wastes. As a result there is expected to be little change to
the overall gas generation rates by radiolysis of ‘Other Organics’.

Overall, the gas generation rates for the proposed G-value data for the proposed material
categories are expected to remain broadly similar (compared to using the previously applied
G-values, Table 1). This is based on the expectation that:
 The current categories of ‘water’ and ‘polymers’ (plastic wastes) are the main contributors to
overall gas generation;
 The G-values for ‘water’ will be mainly replaced by the category of ‘cementitious grout and its
pore water’ with an expected increase in gas generation rates, although it is noted that there
may be water associated with the wastes that should continue to use the G-values for water,
not the G-values for cementitious grout and its encapsulant, and;
 The category of ‘polymers’ will be replaced by several individual types of plastic wastes with
their associated reduction in gas generation rates (although it is noted that the suggested
G-value for alpha radiolysis of polyethylene is higher and would therefore be expected to
increase the gas generation rate for alpha radiolysis of polyethylene).

AMEC/200615/001 Issue 3 Page 59 of 77


6 Summary of the review with suggested
G-value data for use in future gas
generation calculations
Radiolysis of pure water typically forms hydrogen gas. For radiolysis of organic materials and
cementitious encapsulant, the most common gaseous product is hydrogen; however other gases
such as CO, CO2 or HCl may be formed depending on the composition of the material under
irradiation.

A summary of best estimate G-values is presented in Table 15 for use in future gas generation
calculations. The data should be used to provide more detail and underpinning for the radiolysis
section of the Gas Status report [75], including the table of G-values in its Appendix. In addition, a
summary of upper bound estimates of G-values is presented in Table 16 to provide an indication of
the uncertainities in the G-value data.

Table 15 Suggested best estimate G-values for bulk gas for use in future SMOGG
calculations
GGAS-value
Material Gas (molecules/100eV)
α β γ
(a)
Pure water H2 1.2 As for γ 0.45
(b) (b) (b)
Cellulose H2, CO and CO2 1.0 4.0 4.8
Halogenated plastics
(c) (c) (c)
PVC (and PVC/PTFE mixtures) H2, CO, CO2 1.3 As for γ 1.0
Non-halogenated plastics
Polyethylene H2, CO2 4.2 As for γ 3.1
Perspex (PMMA) H2, CO and CO2 1.3 As for γ 2.2
Rubbers
(c) (c)
Neoprene (halogenated rubber) H2, CO, CO2 1.7 As for γ 0.40
Styrene butadiene co-polymer Bulk gas,
0.15 As for γ 0.1
(non-halogenated rubber) mainly H2
Organic Ion exchange resins
Styrene divinylbenzene H2 0.23 As for γ 0.21
As for γ
(d)
Phenol-formaldehyde based resin H2 0.23 0.17
(b) (b)
Small organic molecules CO2 4.0 As for γ 5.4
As for γ
(b) (b)
Small organic molecules CH4 1.4 3.9
Waste encapsulants
Cementitious encapsulant and its pore water H2 0.3 As for γ 0.2
Polymer encapsulant
(vinyl ester styrene resin, VES, H2 0.15 As for γ 0.05
called DOW WSB 101)
(a) Only H2 is produced from pure water
(b) Data should be used with caution until further experimental evidence is available.
(c) HCl(g) is assumed to dissolve within the pore water and therefore not contribute to the overall
production of bulk gas.
(d) As for styrene divinylbenzene

AMEC/200615/001 Issue 3 Page 60 of 77


Table 16 Suggested upper bound G-values for bulk gas for use in future SMOGG
calculations
GGAS-value
Material Gas (molecules/100eV)
α β γ
(a)
Pure water H2 1.7 As for γ 0.5
(b) (b) (b)
Cellulose H2, CO and CO2 1.5 6.0 6.2
Halogenated plastics
PVC (and PVC/PTFE mixtures) H2, HCl, CO, CO2 2.8 As for γ 5.8
Non-halogenated plastics
Polyethylene H2, CO2 7.3 As for γ 4.1
Perspex (PMMA) H2, CO and CO2 2.0 As for γ 4.1
Rubbers
Neoprene (halogenated rubber) H2, HCl, CO, CO2 4.5 As for γ 0.7
Styrene butadiene co-polymer Bulk gas,
4.6 As for γ 0.2
(non-halogenated rubber) mainly H2
Organic Ion exchange resins
Styrene divinylbenzene H2 0.27 As for γ 2.1
As for γ
(c) (c) (c)
Phenol-formaldehyde based resin H2 0.27 2.1
(b) (b)
Small organic molecules CO2 7.2 As for γ 10.5
As for γ
(b) (b)
Small organic molecules CH4 7.2 10.5
Waste encapsulants
Cementitious encapsulant and its pore water H2 0.4 As for γ 0.8
Polymer encapsulant
(vinyl ester styrene resin, VES, H2 4.6 As for γ 0.08
called DOW WSB 101)
(a) Only H2 is produced from pure water
(b) Data should be used with caution until further experimental evidence is available.
(c) As for styrene divinylbenzene

AMEC/200615/001 Issue 3 Page 61 of 77


7 References
1 B.T. Swift and W.R. Rodwell, Specification for SMOGG Version 5.0: a Simplified Model of Gas
Generation from Radioactive Wastes, Serco Report SA/ENV–0452 (Version 6), 2006.
2 B.T. Swift, SMOGG (Version 6.0), a Simplified Model of Gas Generation from Radioactive
Wastes: User Guide, Serco Report SA/ENV–0511 (Version 6), 2006.
3 N.R. Smart and A.R. Hoch, A survey of steel and Zircaloy corrosion data for use in the
SMOGG gas generation model, Serco Assurance Report SA/ENV 0841 Issue 3, 2010.
4 A.R. Hoch, N.R. Smart and B. Reddy, A survey of reactive metal corrosion data for use in the
SMOGG gas generation model, Serco Assurance Report SA/ENV 0895 Issue 2, 2010.
5 F.M.I. Hunter, C. Leung and A. Adeogun, Emanation coefficients relating to in package
behaviour of radon, AMEC Report AMEC/000142/001, Issue 1, 2014, to be published.
6 J. Small and M. Dutton, Review of research on cellulose degradation and input data for the
Simplified Model of Gas Generation (SMOGG), NNL Report NNL (09) 8870, Issue 3.
7 S. Watson, S. Benbow, P. Suckling, G. Towler, T. Hicks, M. Pekala, P. Towler and S. Saleh,
Assessment of issues relating to pre-closure to post-closure gas generation in a GDF,
Quintessa Report QRS-1378ZP-R1 Version 3, 2012.
8 F.M.I. Hunter and B.T. Swift, An assessment of the generation of GDF-derived gas using the
2007 Derived Inventory, AMEC Report AMEC/006255/001 Issue 3, 2014, to be published.
9 A.R. Hoch, M.C. Thorne, B.T. Swift and F. Bate, Update of the GPA (03) assessment of the
consequences of gas, Serco Report SA/ENV-0948, 2007.
10 Department of Energy and Climate Change and Nuclear Decommissioning Authority, The
2013 UK Radioactive Waste Inventory, NDA Report URN 14D039 NDA/ST/STY (14) 0006,
2014. Available at https://www.nda.gov.uk/ukinventory/.
11 G. Towler, T. Baldwin, A. Paulley and J. Wilson, An initial evaluation of the nature and amount
of voidage associated with an ILW GDF, Quintessa Report ASSIST-1547B-R1 V1.0, 2013.
12 J.A. LaVerne, E.A. Carrasco-Flores, M.S. Araos and S.M. Pimblott, Gas production in the
radiolysis of poly(vinyl chloride), Journal of Physical Chemistry A 112(15), 3345-3351, 2008.
13 A. Baidak, and J.A. LaVerne, Radiation-induced decomposition of anion exchange resins,
Journal of Nuclear Materials 407(3), 211-219, 2010.
14 M. Roder, Aromatic Hydrocarbons. In Radiation Chemistry of Hydrocarbons, G. Foldiak, Ed.,
Elsevier, Amsterdam, 393-468, 1981.
15 I. Gyorgy, Aliphatic Alkanes. In Radiation Chemistry of Hydrocarbons, G. Foldiak, Ed.,
Elsevier, Amsterdam, 61-176, 1981.
16 A. Hiroki and J.A. LaVerne, Decomposition of hydrogen peroxide at water-ceramic oxide
interfaces, Journal of Physical Chemistry B 109, 3364-3370, 2005.
17 B. Pastina, J.A. LaVerne and S.M. Pimblott, Dependence of molecular hydrogen formation in
water on scavengers of the precursor to the hydrated electron, Journal of Physical Chemistry
A 103(29), 5841-5846, 1999.
18 J.A. LaVerne and S.M. Pimblott, New mechanism for H-2 formation in water, Journal of
Physical Chemistry A 104(44), 9820-9822, 2000.

AMEC/200615/001 Issue 3 Page 62 of 77


19 B.C. Garrett, D.A. Dixon, D.M. Camaioni, D.M. Chipman, M.A. Johnson, C.D. Jonah, G.A.
Kimmel, J.H. Miller, T.N. Rescigno, P.J. Rossky, P.J. Rossky, S.S. Xantheas, S.D. Colson,
A.H. Laufer, D. Ray, P.F. Barbara, D.M. Bartels, K.H. Becker, K.H. Bowen, S.E. Bradforth, I.
Carmichael, J.V. Coe, R. Corrales, J.P. Cowin, M. Dupuis, K.B. Eisenthal, J.A. Franz,
M.S. Gutowski, K.D. Jordan, B.D. Kay, J.A. LaVerne, S.V. Lymar, T.E. Madey, C.W. McCurdy,
D. Meisel, S. Mukamel, A.R. Nilsson, T.M. Orlando, N.G. Petrik, S.M. Pimblott, J.R. Rustad,
G.K. Schenter, S.J. Singer, A. Tokmakoff, L-S. Wang, C. Wittig and T.S. ZwierI, The role of
water on electron-initiated processes and radical chemistry: issues and scientific advances.
Chemical Reviews 105, 355-389, 2005.
20 H. Christensen, Fundamental aspects of water coolant radiolysis, SKI Report 16 ISSN 1104-
1374 ISRN SKI-R-06/16-SE. 2006.
21 A.J. Elliot, The reaction set, rate constants and G-values for the simulation of the radiolysis of
light water over the range 20° to 350°C based on information available in 2008. AECL Report
AECL CW-511300-FM-168 Rev.0. 2009.
22 U.S. Department of Energy (DOE), RH-TRU Payload Appendices Revision 2, Appendix 2.1:
Radiolytic G-values for waste materials, Carlsbad Field Office, Carlsbad, New Mexico, 2013.
Available from http://pbadupws.nrc.gov/docs/ML1328/ML13283A186.pdf. Note that the parent
document is ‘The Remote-Handled Transuranic Waste Authorized Methods for Payload
Control (RH-TRAMPAC), available at
http://pbadupws.nrc.gov/docs/ML1328/ML13283A184.pdf.
23 E. Bjergbakke, K. Sehested, O.L. Rasmussen and H. Christensen, Input files for computer
simulation of water, Risö-M-2430, Risö 1984.
24 C.P. Ruiz, C.C. Lin, R.N. Robinson, W.G. Burns and A.R. Curtis, Model calculations of water
radiolysis in BWR primary coolant, Water Chemistry of Nuclear Reactor Systems 5, Vol. 2, p.
131, BNES 1989.
25 G.V. Buxton and A.J. Elliot, High temperature water radiolysis and its relevance to reactor
coolant, Proceedings 1991 JAIF International Conference on Water Chemistry and Nuclear
Power Plants, 283-288, Fukui City, 1991.
26 M.C. Kent and H.E. Sims, The yield of γ-radiolysis products from water at temperatures up to
300°C, Water Chemistry of Nuclear Reactor Systems 6, BNES 1992 and AEA-RS-2302,
Harwell 1992.
27 A.J. Elliot, D.C. Oullette and D.R. McCracken, High-temperature pulse radiolysis facility,
AECL-10667, Atomic Energy of Canada Ltd., Chalk River, Ontario, Canada, 1992.
28 A.J. Elliot, M.P. Chenier and D.C. Ouellete, Temperature of g-values for H2O and D2O
irradiated with low energy transfer radiation, Journal of the Chemical Society, Faraday
Transactions, 89, 1193-1197, 1993.
29 J.A. LaVerne and S.M. Pimblott, Diffusion-kinetic modelling of the electron radiolysis of water
at elevated temperatures, Journal of Physical Chemistry 97, 3291-3297, 1993.
30 G.R. Sunaryo, Y. Katsumura, I. Shirai, D. Hiroishi, and K. Ishigure, Radiolysis of water at
elevated temperatures-I, Irradiation with gamma-rays and fast neutrons at room temperature,
Radiation Physics and Chemistry, 44, 273-280, 1994.
31 A.J. Elliot, Rate constants and g-values for the simulation of the radiolysis of light water over
the range 0-300 °C, AECL Report AECL-11073, 1994.
32 H.C. Christensen, Studsvik Nuclear AB, STUDSVIK/N(K)-0214. , 2002. Cited within
H. Christensen, Fundamental aspects of water coolant radiolysis, SKI Report 16 ISSN 1104-
1374 ISRN SKI-R-06/16-SE. 2006.
33 J.A. LaVerne, Radiation chemical effects of heavy ions, in “Charged particle and photon
interactions with matter”, Eds.: Mozumder, A. and Hatano, Y, M. Dekker, New York 2004.
34 W.G. Burns, and H.E. Sims, Radiation chemistry of high-temperature water, Journal of the
Chemical Society, Faraday Transactions 1, 77, 2803, 1981.

AMEC/200615/001 Issue 3 Page 63 of 77


35 N. Bibler, Curium-244 α-radiolysis of nitric acid. Oxygen production from direct radiolysis of
nitrate ions, Journal of Physical Chemistry, 78, 211-215, 1974.
36 K. Lundgren, H. Wijkström and G. Wikmark, Recent developments in the LwrChemradiolysis
code, International Conference on Water Chemistry of Nuclear Reactor Systems, San
Francisco 2004.
37 M. Lefort, Radiation chemistry, Annual Review of Physical Chemistry 9, 123, 1958.
38 G.H. Jenks, Effects of reactor operation on HFIR coolant, ORNL-3848, Oak Ridge National
Laboratory, 1965.
39 B.T. Swift, G.M.N. Baston, J. Dawson, A. Craven and N.R. Smart, Effect of chloride on internal
drum corrosion for supercompacted PCM waste, Serco Report SERCO/TAS/002067/001,
2009.
40 J. Dawson, Radiation effects in PVC bag materials: Measurement of HCl Evolution, Serco
report SERCO/TAS-1011 Issue 03, 2008.
41 Nuclear Decommissioning Authority, Geological Disposal: Generic Disposal System Technical
Specification, NDA Report NDA/RWMD/044, 2010.
42 M. Celina, H. Kudoh, T.J. Renk, K.T. Gillen and R.L. Clough, Pulsed ion beam irradiation
effects on surfaces of polymeric materials, Polymers for Advanced Technologies 9(1), 38-44,
1998.
43 D.T. Reed and M.A. Molecke, Generation of volatile organic compounds by alpha particle
degradation of WIPP plastic and rubber material, Argonne National Laboratory
report ANLI/CMT/CP-79852, 1993.
44 D.T. Reed, J. Hoh, J. Emery, S. Okajima and T. Krause, Gas production due to alpha particle
degradation of polyethylene and polyvinyl chloride, Argonne National Laboratory Report ANL
97/7, 1998. Available at http://www.osti.gov/scitech/biblio/303944
45 E.A. Hegazy, T. Seguchi and S. Machi, Radiation-induced oxidative degradation of poly(vinyl
chloride), Journal of Applied Polymer Science, Volume 26 (9) 2947–2957, 1981.
46 A.H. Zahran, E.A Hegazy and E. Eldin, Radiation effects on poly (vinyl chloride)—I. gas
evolution and physical properties of rigid PVC films, Radiation Physics and Chemistry,
Volume 26(1) 25-32, 1985.
47 K. Arakawa, T. Seguchi and K. Yoshida, Radiation-induced gas evolution in chlorine-
containing polymer. poly(vinyl chloride), chloroprene rubber, and chlorosulfonated-
polyethylene, International Journal of Radiation Applications and Instrumentation. Part C.
Radiation Physics and Chemistry, Volume 27(2), 157-163, 1986.
48 C. Decker, Oxidative degradation of poly(vinyl chloride), Journal of Applied Polymer Science.
Volume 20(12) 3321–3336, 1976.
49 T. Seguchi, Mechanisms and kinetics of hydrogen yield from polymers by irradiation, Nuclear
Instruments and Methods in Physics Research B, Volume 185, 43, 2001.
50 Z. Chang, and J.A. LaVerne, Hydrogen production in the heavy ion radiolysis of polymers. 1.
polyethylene, polypropylene, poly(methyl methacrylate), and polystyrene, Journal of Physical
Chemistry B 104(45), 10557-10562, 2000.
51 J. Dawson, NAPL Generation from the Radiolysis of Polyethylene and Chlorosulphonated
Polyethylene Elastomer (Hypalon®), AMEC report AMEC/PPE/2834/002, 2013.
52 K. Saito, Gas generation by gamma-ray irradiation to polyethylene, polystyrene and cellulose
resin, JAEA report 2011-043, 2010.
53 T. Sasuga, Radiation effects on polymer materials. Ionizing radiation induces degradation or
improvement?: (2) Gas evolution by irradiation, IAEA report, IAEA INIS REF ISSN 0286-8873,
2005.

AMEC/200615/001 Issue 3 Page 64 of 77


54 J.R. Green, Progress report for the enhancement of radcalc: Isotope database, gamma
absorption fractions, and G(H2) values, Westinghouse Hanford Company report WHC-SD-TP-
RPT-014, 1994.
55 R. Aliev, R. Navarro-Gonzalez and R. Medina, A comparative radiation degradation of some
aromatic polyesters, Polymer Bulletin 57(4), 499-504, 2006.
56 D.C. Philips and S.G. Burnay, Irradiation effects in polymers, edited by D.W. Clegg and
A.A. Collyer, Elsevier Applied Science, 1991.
57 M. Sari and A. Craven, A review of G-values from radiolysis of ion exchange media, Nuclear
Technologies report NT/722500694/R821 Issue 1, 2010.
58 K.K.S. Pillay, A review of the radiation stability of ion exchange materials, Journal of
Radioanalytical and Nuclear Chemistry, Articles, Volume 102(1), 247-268, 1986.
59 K. Enomoto, J.A. LaVerne, L. Tandon, A.E. Enriquez, J.H. Matonic, The radiolysis of
poly(4-vinylpyridine) quaternary salt ion exchange resins, Journal of Nuclear Materials 373(1-
3), 103-111, 2008.
60 P. Bouniol, E. Bjergbakke, A comprehensive model to describe radiolytic processes in cement
medium. Journal of Nuclear Materials 372, 1-15, 2008.
61 Idaho National Engineering and Environmental Laboratory, TRUPACT-II Matrix Depletion
Program. Final Report INEL/EXT-98-00987, Rev. 1, 1999. Available at
http://infohouse.p2ric.org/ref/09/08069.pdf .
62 K. Chandler, M. Constable, J. Dawson and S. Morgan, Superplasticiser irradiation trials;
gamma irradiation testing of grout and aqueous samples containing ADVA cast 550, WMT
report WMT(08)P111, 2009.
63 B. Siskind, Gas generation from low-level radioactive waste: concerns for disposal,
Brookhaven National Lab report BNL NUREG 47144. 1992. Available at
http://www.osti.gov/scitech/biblio/5601498.
64 L.R. Dole and H.A. Friedman, Radiolytic gas generation from cement-based waste hosts for
DOE low-level radioactive wastes, Symposium on the Effects of Radiation on Materials,
Seattle, Washington, June 23-25, 1986. Available at
http://www.iaea.org/inis/collection/NCLCollectionStore/_Public/18/001/18001687.pdf
65 K. Kroth, E. Barnert, P.H. Brucher, H. Lammertz, D. Niephaus, Formation of hydrogen and
radioactive gases in waste packages with cemented intermediate level waste, Waste
Management Symposia conference proceedings 1990, volume 2, 363-368, 1990. Available at
www.wmsym.org/archives/1990/V2/52.pdf
66 D.J. Lee and C.R. Wilding, Cement Wasteform properties Waste Management Symposia
conference proceedings 1989, volume 2, 319-326, 1989. Available at
www.wmsym.org/archives/1989/V2/53.pdf.
67 G.L. Bykov, A.V. Gordeev, T.K. Yurik, B.G. Ershov, Gas formation upon γ-irradiation of cement
material, High Energy Chemistry, Volume 42(3) 211-214, 2008.
68 Nuclear Decommissioning Authority, Assessment of modifications to pre-treatment and
packing of remaining radium wastes: summary of assessment report, 2012. Available at
http://www.nda.gov.uk/documents/biblio/.
69 A. Idesaki, H. Uechi, S. Matsuda and H. Kishi, Irradiation effect of gamma rays on cyanate
ester/epoxy resins (2), JAEA Takasaki Annual Report 2010, JAEA-Review, 2011.
70 J. Dawson, "Spot-Check" ageing tests on a candidate polymer encapsulant - low temperature
alchemix epoxy resin, AMEC report AMEC/PPE/3344/001, 2013.
71 J. Dawson, G. M.N. Baston, M.M. Cowper and T.A. Marshall, The effects of gamma radiation
and thermal ageing on the stability of candidate polymeric encapsulant materials, Serco
Report SERCO/TAS/002008/001 Issue 02, 2010.

AMEC/200615/001 Issue 3 Page 65 of 77


72 N.J. Simon, A review of irradiation effects on organic-matrix insulation, National Institute of
Standards and Technology report NISTIR 3999, 1993. Available at
http://www.iaea.org/inis/collection/NCLCollectionStore/_Public/35/058/35058183.pdf
73 M.D. Naughton and P.J. Robinson, Radwaste radiolytic gas generation literature review,
Electric Power Research Institute report EPRI NP-5977, 1988. Available at
http://www.epri.com.
74 Nuclear Decommissioning Authority, Geological Disposal: Generic Disposal System Technical
Specification, NDA Report NDA/RWMD/044, 2010.
75 Nuclear Decommissioning Authority, Geological Disposal: Gas Status Report, NDA Report
NDA/RWMD/037, 2010. Available at http://www.nda.gov.uk/documents/biblio/.

AMEC/200615/001 Issue 3 Page 66 of 77


Appendices

Contents
Appendix 1 SMOGG G-values: Treatment in the QPAC model
Appendix 2 Additional relevant UK RWI information

AMEC/200615/001 Issue 3 Page 67 of 77


AMEC/200615/001 Issue 3 Page 68 of 77
Appendix 1 SMOGG G-values:
Treatment in the QPAC model

Contents
SMOGG G-values: Treatment in the QPAC model.

AMEC/200615/001 Issue 3 Page 69 of 77


AMEC/200615/001 Issue 3 Page 70 of 77
SMOGG G-values: Treatment in the
QPAC Model
Sarah Watson
QRS-1517C
September 2013

A1 Introduction
This appendix provides information about the implementation of radiolysis in the recent QPAC gas
generation modelling [7]. It covers the implementation of radiolysis in QPAC and some of our
thoughts concerning the approximations in the model. Where the appendix refers to ‘as in
SMOGG’, this should be taken to indicate either SMOGG default or the value used in the GPA
Update.

A2 Implementation in QPAC
The QPAC implementation of radiolysis is based on the SMOGG implementation. However, the
coupled nature of the QPAC model, and in particular the emphasis on cases in which either the
backfill or the waste is partially saturated, resulted in some differences in treatment.

QPAC, like SMOGG, has the ability to calculate gas generation for whatever scale of system the
user chooses, from a single package to an entire GDF. The calculations carried out for RWMD
considered each disposal vault to be a single well-mixed volume but the QPAC model could
consider a discretised vault if required.

A2.1 Processes
The radiolysis processes included in the QPAC model are the same as the processes included in
SMOGG, namely:

• radiolysis of water to produce H2;


• radiolysis of polymers to produce H2;
• radiolysis of oils to produce H2;
• radiolysis of cellulose to produce H2;
• radiolysis of ISA (small organics) to produce CH4 and CO2.

The equations used are summarised below.

Process Schematic reaction


Radiolysis of water α. Water → H2
where α (-) is a ‘radiolysis waste water consumption factor’
parameter taken from the SMOGG work (set to 1).

Polymer Radiolysis (polymer) → H2


Polymer is assumed not to be consumed in the reaction

Oil Radiolysis (oil) → H2


Oil is assumed not to be consumed in the reaction

Cellulose Radiolysis (cellulose) → H2


Consumption of cellulose is captured in the rates assigned to
the hydrolysis etc reactions

AMEC/200615/001 Issue 3 Page 71 of 77


Process Schematic reaction
Small Organics Radiolysis to CO2 ISA → 6.CO2

Small Organics Radiolysis to CH4 ISA → 6.CH4

Where C-14 or H-3 are present in the source material, the fraction of the generated gas that is
radioactive is the same as in the source material.

The basic process underlying the radiolysis calculations is the same as in SMOGG. The radiation
energy released from the waste is calculated at each timestep from decay of the initially specified
inventory. G-values, taken from the SMOGG documentation, are then used to calculate the gas
generation rate for each source material.

G-values taken from SMOGG documentation:


-1 -1 -1
Substrate Gas G (α) mol J G(β) mol J G(γ) mol J

Water H2 1.7E-7 5.2E-8 5.2E-8

Cellulose H2 1.5E-7 6.2E-7 3.3E-7

Polymers H2 2.8E-7 6.2E-7 6.2E-7

Oils H2 5.5E-7 5.8E-7 4.8E-7

ISA CO2 4.1E-7 5.6E-7 5.6E-7

ISA CH4 1.5E-7 4.0E-7 4.0E-7

As in SMOGG, it was assumed that all alpha and beta radiation is contained within the waste
packages. A fraction of the gamma radiation penetrates the waste container to cause radiolysis of
water in the backfill or buffer during the post-closure period. For SILW, UILW and DNLEU this
fraction was taken to be 10%, as in SMOGG, and for HLW/SF/Pu/HEU this fraction was taken to be
0.1% based on a scoping calculation of the typical shielding likely to be provided by a waste
container.

A2.2 Water
Radiolysis of water was treated slightly differently in the QPAC results presented in the NDA
report [7] to the way it appears to be treated in SMOGG. The basic implementation was the same
as in SMOGG – determine the radiation energy emitted from the waste, multiply by the appropriate
G-factor with water being consumed in the reaction and then multiply by a scaling factor. The
differences arise from the way in which the scaling factor used to estimate the fraction of the
energy that affects the water is calculated. It should be noted that the SMOGG implementation was
also coded into QPAC for use in basic testing.

In QPAC, the model tracks the water content of the waste and the backfill or buffer separately and
allows exchange between the waste and surrounding material at a rate defined by the properties of
the waste packages and the backfill. For UILW, the model also tracked packages containing
Magnox or aluminium wastes separately from the bulk waste to allow for the fact that rapid
corrosion processes within these packages might result in them drying out during the operational
period or post-closure when water supply is limited; these packages were assumed to see the
average radiation field. Thus in the UILW vaults (the most complicated case) there are four
different media in which radiolysis can occur (water in packages containing waste, Mx, Al, and
backfill).

There is significant uncertainty concerning the scaling factor that should be used to describe the
efficiency of the radiolysis process, especially for the backfill where waste package geometry will

AMEC/200615/001 Issue 3 Page 72 of 77


play a role. The scaling factor used in QPAC is a volume fraction, rather than a mass fraction as
used in SMOGG. Using a volume fraction made implementation easier when considering variable
saturation and creep of an evaporite host rock. It was also considered that a volume fraction was
possibly a better measure to use than a mass fraction because of the difference in the densities of
the materials being considered (waste/backfill and water). In practice, using a volume fraction
generally results in higher gas generation rates.

A2.3 Oils and Polymers


Radiolysis of oils and polymers was treated exactly as in SMOGG. This was in part because the
available inventory information was not sufficient to allow the rate at which the polymer or oil
substrate is consumed by radiation to be determined. As a result the gas production rate is
overestimated at late times.

A2.4 Cellulose
Radiolysis of cellulose was treated in the same way as in SMOGG. Cellulose was not consumed in
the radiolysis reaction, although this could easily have been coded in. Instead, it was assumed that
the decrease in the mass of cellulose available for degradation reactions had already been
captured in parameters assigned to the cellulose degradation reactions.

A2.5 ISA (Small Organics)


The QPAC model considers ISA to potentially be present in the inventory at the start of the
calculation and to be generated by the degradation of cellulose. It can be degraded by a range of
different reactions depending on the prevailing conditions (see below) and by radiolysis.

Process Schematic chemical reaction

Aerobic degradation of ISA ISA + 6.O2 → 6.CO2 + 6.Water

5.ISA + 24.NO3 → 30.CO2 + 12.N2 + 18. Water + 24.OH


- -
Oxidation of ISA by nitrate

ISA + 2.SO4 → 6.CO2 + 2.S + 4.H2 + 2. Water


2- 2-
Oxidation of ISA by sulphate

Methanogenic degradation of ISA ISA → 3.CO2 + 3.CH4

The QPAC model considers all of the processes that can occur under the prevailing conditions
according to the reaction rate that has been assigned to the process.

A2.6 Calculation Strategy


The QPAC calculations were carried out in two parts: a detailed calculation for the operational
period; and a calculation that included a less detailed calculation for the operational period and the
post-closure evolution. This strategy was necessary to cope with the lower strength sedimentary
and evaporite host rock cases in which different vaults are closed and so no longer ventilated at
different times; for the higher strength host rock the strategy was a convenience that was useful for
managing the very large amounts of output data produced.

For the operational period, the calculation was carried out for a single ‘vault’. In the case of the
higher strength host rock, radiolysis is effectively treated in the same way as in the SMOGG
calculations because the system can be approximated as a single large vault to which waste is
added.

For the lower strength sedimentary and evaporite host rocks the detailed QPAC calculation for the
operational phase was carried out assuming inventory for the average (mid-time) vault so that the
other processes that are coupled to radiolysis through availability of water and ISA see some sort
of average water consumption. A more accurate estimate of the component of the gas generation
due to radiolysis (and the scaling of radioactive species) was substituted during the post-
processing stage when the output from multiple vaults was added together. This updated estimate

AMEC/200615/001 Issue 3 Page 73 of 77


uses the correct energy due to decay and the average water availability. Although approximate, it
was judged that the uncertainties introduced were likely to be small compared with other
uncertainties in the calculations.

For the post-closure part of the calculations, the calculation is carried out for a single representative
vault or tunnel. The following assumptions were made with regard to the inventory used for
radiolysis:

 higher strength host rock: SILW/UILW/DNLEU – the calculation covers the entire disposal area
and all vaults enter the post-closure phase at the same time. Thus the inventory at closure
(2150) is used as a starting point.
 higher strength host rock: HLW/SF/Pu/HEU – tunnels are closed as soon as they have been
filled. To be conservative, the inventory at the date of closure of the first tunnel is used.
 lower strength sedimentary and evaporite host rocks: SILW/UILW/DNLEU – in these cases the
disposal vaults are closed as soon as they have been filled so the appropriate initial inventory
assumption for the radiolysis calculations is different for each vault. The average vault (i.e. the
middle one in the closure sequence) is used to provide the inventory for the post-closure
radiolysis calculation, and in the operational analysis that forms the initial part of the post-
closure run.
 lower strength sedimentary and evaporite host rocks: HLW/SF/Pu/HEU – tunnels are closed
as soon as they have been filled. To be conservative, the inventory at the date of closure of
the first tunnel is used.

A3 Further Thoughts
The largest uncertainty in the radiolysis calculations is almost certainly associated with the scaling
factor used to estimate the fraction of the energy that affects each substrate. This factor will
depend on the contents of each waste package, in particular the distribution of radionuclides
relative to the distribution of substrate, and on geometry. However, the relationship between
substrate and inventory could also be captured in the definition of the G-values (and perhaps it
already is). At the vault-level considered in the SMOGG and QPAC calculations it is difficult to see
how to justify a more complex derivation of the scaling factor than the mass and volume scalings
used in QPAC and SMOGG.

The fact that radiolysis does not consume oils, polymers, and to some extent cellulose, is an
inconsistency and is conservative in terms of the radiolysis gas generation rate. However, overly
conservative gas generation rates for processes not linked to water availability can have a
significant effect in coupled calculations for systems where water availability is low because the
excess gas generated by radiolysis may result in the amount of corrosion, and possibly
radionuclide release from metals, being underestimated. For this reason, we considered
implementing consumption of polymers and oils in QPAC but concluded that we did not have the
input data available to make any attempt plausible.

AMEC/200615/001 Issue 3 Page 74 of 77


Appendix 2 Additional relevant UK
RWI information

Contents

The 2013 UK RWI provides limited information about the aqueous content of each waste stream.
Therefore, Table A2 provides a list of the waste streams that are presented in Table 2 with the
aqueous content (% weight and total mass) for the whole waste stream as reported in the 2013 UK
RWI. However, this information does not provide the complete picture to understand how much water
is available for gas generation by radiolysis after waste packaging, emplacement of the wastes and
closure of a GDF, and therefore this information is presented as background data only.

AMEC/200615/001 Issue 3 Page 75 of 77


AMEC/200615/001 Issue 3 Page 76 of 77
Table A2 The aqueous content for waste streams presented in Table 2 (from the
2013 UK RWI).
Waste Waste Stream Description Aqueous Content
Stream
% weight Tonnes

2D03 Plutonium Contaminated Materials; Drums <1 47


2D06 Plutonium Contaminated Materials; Crates and Filters <1 71
2D08 Magnox Cladding and Miscellaneous Solid Waste 40.6 2,381
2D116 Miscellaneous Plants Initial/Interim Decommissioning: 0 -
Processing Plants, Tanks, Silos etc.
2D130 Miscellaneous Plants Initial/Interim Decommissioning: 0 -
Processing Plants, Tanks, Silos, etc. (PCM)
2D27/C Encapsulated Floc from Effluent Treatment 0* -
2D39 Miscellaneous Beta/Gamma Waste Store 0 -
2D83/C Encapsulated Plutonium Contaminated Materials 0* -
2D90 Plutonium Contaminated Materials; Drums <1 203
2F02 Plutonium Contaminated Materials; Drums <1 1.7
2F34 Plutonium Contaminated Materials; Drums <1 1
3S12 CVCS Resins and Spent Resins (ILW) 30 22
4B01 Ion Exchange Resin and Sand 0** -
4C03 Pond Water Filtration Resin 0** -
5B02 Low Alpha RHILW Not Estimated
5B25 ILW Shaft (Contents) Not Estimated
5C08 ILW Concrete Lined Drums <0.01 0.2
5C30 Harwell Remote Handled ILW <0.01 0.8
5C317 Harwell Contact Handled ILW Drums <0.01 0.05
7A111 Decommissioning Waste PCM ILW 0 -
7A117 Decommissioning Waste Uranium Contaminated ILW 0 -
7A21 Operational ILW Plutonium Contaminated 0 -
7D29 Intermediate Level Waste Resin from Plant 25 3
Decontamination (MODIX)
9C02 PWTP Ion Exchange Material 30 10
9C36 Ion Exchange Resin from Ponds 30 38
9D27 Ion Exchange Material 66 20
9D28 Ion Exchange Material 67 25
* It is noted that these are conditioned wastes and therefore they are expected to have water
present in the cementitious grouts although this is not stated.
** These waste streams are expected to contain some water although this is not stated.

AMEC/200615/001 Issue 3 Page 77 of 77

You might also like