170
QUEENSLAND UNIVERSITY OF TECHNOLOGY SCIENCE AND ENGINEERING FACULTY Measurement and Monitoring of Naturally Occurring Radioactive Materials for Regulation Submitted by Sami Alharbi (B.App.Sc, M.App.Sc) to the Science and Engineering Faculty, Queensland University of Technology, in fulfilment of the degree of Doctor of Philosophy 2016

Measurement and Monitoring of Naturally Occurring ...use of air conditioners and natural ventilation during this time. This variation was observed less frequently in the case of 220Rn

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

QUEENSLAND UNIVERSITY OF TECHNOLOGY

SCIENCE AND ENGINEERING FACULTY

Measurement and Monitoring of

Naturally Occurring Radioactive

Materials for Regulation

Submitted by Sami Alharbi (B.App.Sc, M.App.Sc) to the Science and Engineering Faculty,

Queensland University of Technology, in fulfilment of the degree of Doctor of

Philosophy

2016

ii

Abstract

Naturally occurring radioactive material (NORM) is a commonly used term to describe

material containing primordial radionuclides. Of these, the radioisotopes in the uranium,

actinium and thorium decay series and potassium-40 are important for radiation protection

considerations. Radiation exposure due to NORM has the potential to cause radiological harm to

humans and the environment. Scientific studies have confirmed the association between

inhalation of short-lived radioactive progeny of radon gas and lung cancer. Radon (Rn) is

released from surfaces of materials containing radium isotopes and accumulated in enclosed

spaces. The measurement of the releases of radon as well as its level in indoor air and the

activity concentration of primordial radionuclides in materials are of particular interest for

controlling the resultant exposure. This thesis investigates such measurements, with an

emphasis on the regulation of NORM.

NORM had been largely unregulated in non-nuclear industries until the mid-1980s. The

current regulation of NORM derives from that developed for artificial radionuclides. However,

exposure pathways and scenarios for NORM differ from that for artificial radionuclides, and

consequently, the implementation of such regulation can create confusion among regulatory

authorities. A comprehensive review of recent developments in the radiation protection system

and the current regulation of NORM has been conducted as part of the present thesis. This

review identifies and discusses the challenges that regulatory bodies and industries may face

when implementing NORM regulation. The review leads to the conclusion that the NORM

industry has to be managed on a case-by-case basis because each industry may have its own

exposure characteristics. The implications of the implementation of newly proposed ICRP and

IAEA approaches for radon regulation are also discussed. The regulation of NORM requires

adequate measurements with sufficient accuracy and reliability. As NORM is a vast subject,

three measurement areas were selected because of their importance in the context of radiation

protection: (i) the measurement of radon activity flux density using the activated charcoal

canister technique; (ii) the concentration of radon and thoron in indoor air; and (iii) the analysis

of NORM in solid materials using a broad energy germanium (BEGe) detector.

The measurement methodology of radon activity flux density using an activated

charcoal canister was revisited. The results showed that the vertical distribution of the

adsorbed 222Rn in the charcoal bed of exposed canisters was not the same as in standard

canisters prepared in the laboratory for calibrating measurement instruments. The distribution

also varied according to exposure times. Another observation was the reduction of the

measured 222Rn activity flux with the exposure time. This reduction was attributed to the back

diffusion effect, where 222Rn escaped from the canister through a canister-soil interface and

iii

possibly through the adsorption of water molecules by the activated charcoal. These factors

place limitations on the accuracy of the activated charcoal canister technique.

The activity concentrations of radon (222Rn) and thoron (220Rn) in various workplaces in

Brisbane, Australia, were measured. Basements and confined areas were found to host higher

222Rn activity concentration, most likely because of poor ventilation systems in such locations.

For 220Rn, the maximum activity concentration was detected in storage rooms. Diurnal variation

in 222Rn concentration was observed in workplaces—low during working hours because of the

use of air conditioners and natural ventilation during this time. This variation was observed less

frequently in the case of 220Rn as the change in its activity concentration during and after

working hours was insignificant. The influence of a number of parameters on indoor 222Rn and

220Rn concentrations was also investigated in this study. Floor levels and flooring types were

found to have the greatest impact on indoor 222Rn levels, while distance from walls and surfaces

have been found to be the parameters that affected indoor 220Rn level the most.

The use of the BEGe detector for the analysis of NORM was also explored in this thesis.

The calibration procedure and sample preparation, as well as the required correction factors,

were described. The performance of the detector was tested using two statistical parameters:

relative bias and u-score. The results of these tests revealed no significant difference between

the measured and the reference activity concentration values. In comparison with a

conventional HPGe coaxial detector, the BEGe detector was superior in detecting peaks at low

gamma energy and intensity, such as 46.5, 50.1, 63.3 and 67.7 keV, of 210Pb, 227Th, 234Th and

230Th, respectively. In addition, a rapid technique of determining the activity concentration of

226Ra through the 186.1 keV peak was successfully performed and compared with the

traditional method, which is through 214Pb and 214Bi peaks. Good agreement was reported for

most samples. This method was limited by the large variations in activity concentration and

density between samples and the calibration standard. Overall, the detector was found to be

suitable for the application of NORM.

The findings presented in this thesis advance the understanding of current measurement

techniques and how they should be carried out to adequately regulate NORM. The review on the

current regulation of NORM identified key challenges that are worth addressing in order to

promote its management.

Keywords: NORM, regulation, recommendation, reference level, measurement, radionuclide,

radon, flux density, activity concentration, equilibrium

iv

Acknowledgements

I would like to thank God for giving me the strength to complete this work. I would like to thank

King Saud bin Abdulaziz University for Health Sciences for granting me a full scholarship to

pursue my higher education.

I wish to thank various people for their contribution to this thesis. First of all I would like to

express my very great appreciation to my principal supervisor, Dr Riaz Akber for his enduring

support and guidance. I would like to extend my thank you to my associate supervisor Dr Jamie

Trapp and to the Radiological laboratory technician Kazuyuki Hosokawa for their assistance

throughout my PhD journey.

Finally, I wish to thank my parents, siblings, friends and my wife, who has never once

complained when I have turned up late, for all the generous support and encouragement.

v

Contents

Abstract ................................................................................................................................................................................. ii Acknowledgements ......................................................................................................................................................... iv Contents .................................................................................................................................................................................v List of Figures .................................................................................................................................................................. viii List of Tables ...................................................................................................................................................................... ix List of Publications ............................................................................................................................................................x

Chapter 1 Introduction .................................................................................................................................................... 1

1.1 Naturally occurring radioactive material (NORM) ....................................................................................... 1

1.2 Research objectives ...................................................................................................................................... 2

1.3 Structure of the research .............................................................................................................................. 3 Chapter 2 Literature Review ......................................................................................................................................... 5

2.1 Introduction .................................................................................................................................................. 5

2.2 Origin and distribution ................................................................................................................................ 5

2.3 Terminology ................................................................................................................................................. 6

2.4 Behaviour of NORM ...................................................................................................................................... 8 2.4.1 Uranium ................................................................................................................................................ 8 2.4.2 Thorium ................................................................................................................................................. 8 2.4.3 Radium .................................................................................................................................................. 9 2.4.4 Radon .................................................................................................................................................... 9 2.4.5 Polonium ............................................................................................................................................. 10 2.4.6 Lead ..................................................................................................................................................... 10

2.5 Exposure to NORM ..................................................................................................................................... 11 2.5.1 Internal exposure .............................................................................................................................. 11 2.5.2 External exposure ............................................................................................................................. 12 2.5.3 Natural background radiation .......................................................................................................... 13

2.6 Regulation of NORM .................................................................................................................................. 13

2.7 Measurement of NORM .............................................................................................................................. 16 2.7.1 Measurement of radon releases ....................................................................................................... 17 2.7.2 Measurement of atmospheric radon concentration ....................................................................... 19 2.7.3 Measurement of NORM in solid media ............................................................................................ 20

2.8 Measurement challenges ........................................................................................................................... 22 2.8.1 Measurement cost ............................................................................................................................. 22 2.8.2 Uncertainties ..................................................................................................................................... 24 2.8.3 The background................................................................................................................................. 25 2.8.4 The equilibrium state ........................................................................................................................ 26

2.9 Conclusion .................................................................................................................................................. 28 References ........................................................................................................................................................ 30

Chapter 3 NORM Regulation: Better understanding for better management .......................................... 51

Abstract ............................................................................................................................................................ 53

vi

3.1 Introduction ................................................................................................................................................ 53

3.2 Overview of the radiological protection system ......................................................................................... 54 3.2.1 A brief history of radiological protection ............................................................................................ 54 3.2.2 Quantitative recommendations .......................................................................................................... 55 3.2.3 The conceptual framework ................................................................................................................. 56 3.2.4 Protection beyond humans ................................................................................................................. 57

3.3 The current system of radiological protection ............................................................................................ 57 3.3.1 Exposure situations ............................................................................................................................. 59 3.3.2 Principles ............................................................................................................................................. 60 3.3.3 Exposure categories ............................................................................................................................ 60

3.4 NORM in the radiation protection context ................................................................................................. 60 3.4.1 Radon in dwellings and workplaces .................................................................................................... 62 3.4.2 NORM industries ................................................................................................................................. 62

3.5 Regulatory issues with NORM .................................................................................................................... 63 3.5.1 Absence of a complete framework ..................................................................................................... 64 3.5.2 The interference between planned and existing exposure situations ................................................ 65 3.5.3 The lack of a homogeneous approach ................................................................................................ 65 3.5.4 Ubiquity of NORM ............................................................................................................................... 66 3.5.5 Source of low dose .............................................................................................................................. 67 3.5.6 Societal attitudes ................................................................................................................................ 68 3.5.7 Measurement challenges .................................................................................................................... 68 3.5.8 Waste and residues ............................................................................................................................. 69 3.5.9 Ethical approach ................................................................................................................................. 70

3.6 Case study .................................................................................................................................................. 71 3.6.1 Indoor 222Rn regulation ....................................................................................................................... 72 3.6.2 New policies and recommendations on indoor 222Rn ......................................................................... 73 3.6.3 Implications of the implementation of the new policies .................................................................... 74 3.6.4 The case of indoor 220Rn ..................................................................................................................... 76

3.7 Conclusion .................................................................................................................................................. 76 References ........................................................................................................................................................................................... 78 Chapter 4 Radon-222 Activity Flux Measurement Using Activated Charcoal Canisters: Revisiting

the Methodology .............................................................................................................................................................. 85

Abstract ............................................................................................................................................................ 87

4.1 Introduction ................................................................................................................................................ 87

4.2 Methods ..................................................................................................................................................... 89 4.2.1 The distribution of 222Rn in the charcoal bed ...................................................................................... 89 4.2.2 The influence of exposure time on the measured activity flux........................................................... 89 4.2.3 Analytical technique ........................................................................................................................... 89

4.3 Results and Discussion ................................................................................................................................ 91 4.3.1 The distribution of 222Rn in the charcoal bed ...................................................................................... 91

4.3.1.1 The uniformity of 222Rn in the charcoal bed ............................................................................ 91 4.3.1.2 Detector efficiency at different distance .................................................................................. 92 4.3.1.3 Canisters with different depth of activated charcoal .................................................................. 93

4.3.2 The influence of exposure time on the measured 222Rn activity flux ................................................. 94 4.4 Conclusion .................................................................................................................................................. 99 References ...................................................................................................................................................... 101

vii

Chapter 5 Radon and thoron concentrations in public workplaces in Brisbane, Australia ............. 103

Abstract .......................................................................................................................................................... 105

5.1 Introduction .............................................................................................................................................. 105

5.2 Materials and Methods ............................................................................................................................ 108 5.2.1 Sampling site ..................................................................................................................................... 108 5.2.2 Radon and thoron measurements .................................................................................................... 109

5.3 Results and Discussion .............................................................................................................................. 111 5.3.1 Radon and thoron concentration measurements............................................................................. 111 5.3.2 Diurnal variation ............................................................................................................................... 115 5.3.3 The influence of the measurement heights ...................................................................................... 117 5.3.4 The influence of floor levels .............................................................................................................. 118 5.3.5 The role of flooring on indoor radon and thoron concentrations .................................................... 119

5.4 Conclusion ................................................................................................................................................ 119 References ...................................................................................................................................................... 121

Chapter 6 A broad energy germanium detector for routine and rapid analysis of naturally

occurring radioactive materials ............................................................................................................................. 124

Abstract .......................................................................................................................................................... 126

6.1 Introduction .............................................................................................................................................. 126 6.1.1 Routine analysis of NORM ................................................................................................................ 127 6.1.2 Rapid measurements of 226Ra ........................................................................................................... 127 6.1.3 Determination of 227Ac using gamma-ray spectrometry .................................................................. 128

6.2 Material and methods .............................................................................................................................. 129 6.2.1 Routine analysis of NORM ................................................................................................................ 129 6.2.2 Rapid measurements of 226Ra ........................................................................................................... 133 6.2.3 Determination of 227Ac using gamma ray spectrometry ................................................................... 133 6.2.4 Sample preparation .......................................................................................................................... 134 6.2.5 Analytical technique ......................................................................................................................... 135 6.2.6 Gamma self-attenuation ................................................................................................................... 137 6.2.7 Coincidence summing ....................................................................................................................... 138 6.2.8 Uncertainties ..................................................................................................................................... 139

6.3 Results and discussion .............................................................................................................................. 140 6.3.1 Routine analysis of NORM ................................................................................................................ 140 6.3.2 Rapid measurements of 226Ra ........................................................................................................... 146 6.3.3 Determination of 227Ac using gamma ray spectrometry ................................................................... 149

6.4 Conclusion ................................................................................................................................................ 151 References ...................................................................................................................................................... 153

Chapter 7 Summary and Conclusion ..................................................................................................................... 157 7.1 Conclusion ................................................................................................................................................ 157 7.2 Future Direction ........................................................................................................................................ 159

viii

List of Figures

Figure 2.1 Uranium and thorium decay series ..................................................................................................... 7

Figure 3.1 The basis of radiological protection regulation. ......................................................................... 58

Figure 3.2 The radiological protection system. ................................................................................................. 58

Figure 3.3. Diagram of the current regulation of NORM. .............................................................................. 61

Figure 4.1 The distribution of 222Rn in the charcoal bed for different exposure times. The

correction factors corresponding to the exposure time are shown in parentheses. ......................... 92

Figure 4.2 Normalised 222Rn activity flux in open-trays plotted against exposure time. ................ 95

Figure 4.3 The measured activity flux versus exposure time for charcoal canister fitted over cans

containing samples. ...................................................................................................................................................... 96

Figure 4.4 The measured activity flux of the samples in the closed-can relative to RAD7 values.

............................................................................................................................................................................................... 97

Figure 4.5 Relative count rate in the charcoal canisters versus the exposure time of samples

taken in a closed-can or open-tray. The dashed line represents a situation referred by Equation

4.3. ........................................................................................................................................................................................ 99

Figure 5.1. A map of the sampling site. ............................................................................................................... 109

Figure 5.2 A box plot showing radon and thoron concentrations across workplaces. The boxes

represent 25–75% of the values with median concentrations represented by the lines inside the

boxes. Whiskers represent the range of the distribution and the dots outside the range of 95%

indicate values as outliers. ....................................................................................................................................... 113

Figure 5.3 Correlation between radon and thoron concentrations ........................................................ 113

Figure 5.4 Histogram showing the distribution of the natural logarithms of radon and thoron

concentrations in workplaces with a normal distribution fitted to the data. Geometric means of

radon and thoron concentrations were 7.4 and 5.8 Bq m-3, with a geometric standard deviation

of 2.2 and 2.0 Bq m-3, respectively. ....................................................................................................................... 114

Figure 5.5 The variation in radon and thoron concentrations during the day. ................................. 115

Figure 5.6 Normalised radon and thoron concentrations during and after work hours. .............. 116

Figure 5.7 Radon and thoron concentrations in the air at 5 cm and 120 cm above the floor for

each sampling site. ...................................................................................................................................................... 117

Figure 5.8 Box-whisker plot showing the overall radon and thoron concentrations at the ground

and breathing zones for all location. ................................................................................................................... 117

Figure 5.9 Box-whisker plot showing radon and thoron concentrations categorised by the floor-

level. .................................................................................................................................................................................. 118

Figure 5.10. Box-whisker plot showing A) radon and B) thoron concentrations characterised by

the flooring type. .......................................................................................................................................................... 119

Figure 6.1 (A) Count per emission versus the gamma energy for the reference materials using

the BEGe detector. (B) A comparison between the detection efficiency of a BEGe detector and an

HPGe coaxial detector. The results are for a calibration standard in petri dish geometry (50 g) in

BEGe and Marinelli Beaker geometry (1000 g) in HPGe-coaxial. ............................................................ 137

Figure 6.2 (A) The measured activity concentration (dots) compared to the certified values

(continuous line) for radioisotopes in the reference materials used in the present study, RGU-1,

RGTh-1, and IAEA-434. The comparison is for uranium, thorium, and actinium radioisotopes.

The uncertainty (2) in the certified values is given as dashed lines and in the measured values

as error bars. (B) Activity ratios of several radionuclides in the uranium and thorium series for

the spiked samples (SS and SG) prepared from the NBL standards. The continuous line

ix

represents a ratio of 1:00 and the error bars are the uncertainty to 2. (C) A comparison of 40K

measured activity concentrations and referenced values for the reference samples SS, SG, and

RGK-1 ................................................................................................................................................................................ 142

Figure 6.3 (A) The 226Ra activity concentration measured in the IAEA standards, RGU-1 and

IAEA-434, using (a) 222Rn progeny gamma lines and the 186 keV peak after correcting it for the 235U contribution, calculated from 235U individual gamma lines at (b) 143, (c) 163, (d) 205 keV,

(e) the combined average from these lines, and (f) from the 238U/235U ratio compared to the

certified values. (B) The 226Ra activity measured using different approaches compared to the

expected values in the spiked samples . ............................................................................................................. 147

Figure 6.4 The activity concentrations of 227Ac versus 235U in various samples examined during

the present study. For most samples, the values lie along the line of equality. Samples such as TI,

TII, and BC show the 227Ac:235U ratio 1. This behaviour is most likely due to extraction of

uranium from these samples. ................................................................................................................................. 151

List of Tables

Table 2.1 Examples of radionuclides in equilibrium ...................................................................................... 26

Table 3.1 The development of dose conversion factor over time ............................................................. 73

Table 4.1 Rn-222 emitting samples used in the present study. ................................................................. 90

Table 4.2 Correction factors of 222Rn activity flux for different charcoal depths in a canister with

respect to that of 2.5 cm depth for one day exposure time. ......................................................................... 94

Table 5.1 Summary of the instrument calibration check using certified sources. ........................... 111

Table 5.2 Radon and thoron concentrations in different locations. ....................................................... 112

Table 6.1 Gamma energy lines of key natural radionuclides (Ekström and Firestone, 2004) .... 132

Table 6.2 Sample descriptions ............................................................................................................................... 135

Table 6.3 Correction factors for self-absorption (fs) and random summing (fR); material

densities (g cm-3) are given between the brackets. A brief description of samples is given in

Table. 2. ............................................................................................................................................................................ 139

Table 6.4 U-scores and relative bias for reference materials .................................................................... 142

Table 6.5 The measured activity concentration of several radionuclides of NORM (Bq kg-1) in

various samples used in the present study. ...................................................................................................... 145

Table 6.6 Ratios of 226Ra activity as measured using different approaches to the ratio obtained

via its progeny............................................................................................................................................................... 149

Table 6.7 Activity ratios of 227Ac to 235U, 238U and 226Ra .............................................................................. 150

x

List of Publications

Alharbi, Sami H and Riaz A Akber. 2014. "Radon-222 activity flux measurement using

activated charcoal canisters: Revisiting the methodology." Journal of

Environmental Radioactivity 129: 94-99.

Alharbi, Sami H and Riaz A Akber. 2015. "Radon and thoron concentrations in public

workplaces in Brisbane, Australia." Journal of Environmental Radioactivity 144:

69-76.

Alharbi, Sami H and Riaz A Akber. 2016. " A broad energy germanium detector for

routine and rapid analysis of naturally occurring radioactive materials." Journal

of Radioanalytical and Nuclear Chemistry. (In Press)

Alharbi, Sami H and Riaz A Akber. 2016. " NORM regulation: better understanding for

better management." (In Preparation)

xi

Statement of Original Authorship

The work contained in this thesis has not been previously submitted to meet

requirements for an award at this or any other higher education institution. To the best

of my knowledge and belief, the thesis contains no material previously published or

written by another person except where due reference is made.

Signature

Date:………………………………

QUT Verified Signature

1

Chapter 1 Introduction

1.1 Naturally occurring radioactive material (NORM)

Naturally occurring radioactive material (NORM) is a widely used term to describe

materials that contain radionuclides of natural origin. The term is usually applied to refer to the

primordial radionuclides 40K, 238U, 235U, 232Th and their radioactive decay products. These

radionuclides are ubiquitous in the environment and they present, with a wide variation of

concentration levels, in raw materials, minerals, products, by-products, equipment, waste and

residues, as well as in many industrial processes. Individual radionuclides of NORM have many

applications in various disciplines, such as environmental science, radiometric dating,

atmospheric studies and erosion and sedimentation studies. Exposure to ionising radiation

emitted by individual radionuclides of NORM can be hazardous to humans and the environment.

The radiation dose rate due to exposure to NORM is generally low in most environments.

However, certain human activities can increase the concentration of NORM and/or alter

exposure conditions. Subsequently, this can give rise to above background radiation dose to

receivers. Such exposure needs to be controlled through regulation to ensure that adequate

protection is given to individuals as well as to the environment.

Since the establishment of radiological protection bodies, most of the regulatory

attention to natural sources has been largely focused on the uranium mining and milling

industry, as it is part of the nuclear fuel cycle. However, over the past three decades, there has

been a growing awareness arising from the recognition of increased levels of NORM in non-

nuclear industries, such as phosphate and petroleum industries. The control of exposure

encompasses both indoor and outdoor environments, whether in dwellings or in workplaces,

including industries involving NORM.

Quantifying the activity levels of individual radionuclides of NORM plays an essential

role in managing radiation exposure. The establishment, implementation and compliance with

any radiological protection regime require accurate measurement. In fact, the measurement of

activity concentration is a prerequisite at all stages of a regulation system, starting from the

screening of radionuclides to periodic monitoring checks. It is also essential for the assessment

of radiation dose received by a certain group of human or non-human biota. However, the

measurement of NORM is associated with difficulties and limitations because of the wider range

of radionuclides involved in great variations in their physical and chemical properties.

Over the past century, radiation protection principles and values have evolved. Any new

radiation protection regulations, standards and monitoring regimes for the protection of people

and the environment from NORM should be developed with the latest knowledge in mind. Old

2

monitoring regimes may also need to be reviewed in light of these new developments. National

regulatory bodies may face challenges related to the implementation of recent radiological

protection concepts in their current system of regulation. A review of the regulation of NORM

would contribute to understand the appropriate ways to control it. On the other hand,

measurements for regulation purposes require standardised methods to ensure that specific

criteria have been met. More research studies are required to develop an understanding of

better ways to manage exposure to NORM and how the measurement of NORM should be

conducted to fulfil regulatory requirements.

1.2 Research objectives

The overall aim of this research project is to develop an understanding of the

measurement, monitoring and regulation of NORM. This aim has a number of subsidiary aims:

- To develop an understanding of the mechanism involved in 222Rn absorption in

activated charcoal and, on that basis, to develop an adequate methodology for its use in

222Rn exhalation measured from materials elevated in uranium series activity

concentrations

- To study the distribution of 222Rn and 220Rn activity concentrations in indoor

environments

- To investigate the suitability of a broad energy germanium (BEGe) gamma spectroscopy

system for quantifying the activity concentration of individual radionuclides of NORM.

The stated aims are accomplished through a set of objectives and sub-objectives:

The first objective is to comprehensively review the latest developments in radiation

protection regarding NORM and then identify the main regulatory issues in establishing,

implementing and complying with radiation protection regimes. The publications to be

considered by this review include those from the International Commission on

Radiological Protection (ICRP), the International Atomic Energy Agency (IAEA) and the

United Nations Scientific Committee on the Effects of Atomic Radiation (UNSCEAR), as

well as those from several nationally-based organisations, such as the Australian

Radiation Protection and Nuclear Safety Agency (ARPANSA) and the Queensland

Radiation Health Unit (as the state-based regulator for radiation protection)

The second objective is to further our knowledge about the measurement of NORM to

demonstrate compliance with national and international requirements for radiation

protection. As NORM is a vast subject, three areas have been selected for investigation:

(1) 222Rn exhalation rate using activated charcoal canisters

(2) Atmospheric 222Rn and 220Rn in workplaces in Brisbane, Australia

(3) The use of the BEGe system for the routine measurement of NORM.

3

These areas were chosen because of their significance in the field of radiation protection

and the availability of instrumentation and devices at QUT.

1.3 Structure of the research

This thesis is presented as a combination of scientific papers, which have been

published, accepted for publication or planned for submission in international peer-reviewed

journals.

A literature review on the measurement and regulation of NORM is presented in Chapter

2. The main purposes of this chapter are to provide a fundamental understanding, examine the

current state of knowledge and identify gaps that require further study in the field of NORM.

This chapter starts with the definition of NORM from the point of view of radiological protection

and a description of its origin, distribution, activity measurement, potential health risks and

regulation. It then highlights the measurement challenges associated with quantifying the

activity concentration of NORM and their potential impact on a radiological protection regime.

The radiological protection system of NORM is critically reviewed in Chapter 3. It

discusses various aspects surrounding the current regulation and highlights key regulatory

issues in establishing and implementing a protection system against exposure to NORM. The

chapter contains a case study on the new ICRP and IAEA recommendations in the regulation of

indoor radon (222Rn) and thoron (220Rn) and the implications of the implementation of these

new recommendations. This chapter is being prepared for submission to a peer-reviewed

journal.

Chapters 4, 5 and 6 investigate the measurement and monitoring methods that are

currently in use to determine the activity concentrations of individual radionuclides of NORM in

different environments. These chapters are presented as published papers. The first paper,

Chapter 4, examines the activated charcoal canister technique for measuring 222Rn activity flux

from surfaces. This method is widely used to establish 222Rn exhalation source terms to model

radon transport and subsequent radiation dose, especially in mine sites and land contaminated

with uranium or 226R. This method basically depends on the mechanism of 222Rn adsorption into

activated charcoal, placed in a canister. The charcoal is then analysed using gamma

spectrometry. This method assumes that 222Rn flux remains constant over time, and that all of

222Rn is adsorbed and retained in the charcoal. Such assumptions may not always be warranted,

which may limit the use of this technique and affect the results. This and some other radiation

detection issues require revisiting the methodology relating to activated charcoal canisters and

reporting its limitations.

The second paper is presented in Chapter 5, and it investigates the level of indoor 222Rn

and 220Rn by measuring and monitoring their activity concentrations in workplaces in Brisbane,

4

Australia. Such measurements become important for regulatory purposes, especially after the

latest findings on the radiation hazards of indoor 222Rn exposure. The measurements cover a

wide spectrum of workplace categories with various building characteristics. The paper also

studies the variations of indoor 222Rn and 220Rn levels and identifies the main parameters that

influence their levels.

The final paper has been accepted for publication and presented in Chapter 6. It

explores the feasibility of using the broad energy germanium (BEGe) detector for determining

the activity concentrations of individual radionuclides of NORM. The use of such a detector for

routine measurements of NORM is of significant advantage for the purposes of complying with

regulations and monitoring the environment. This is because of the large crystal size of the

detector that would enhance the sensitivity for low energy gamma-rays in addition to the ability

to simultaneously measure a large number of radionuclides, which can save time and costs.

Although the detector has been widely used in nuclear physics studies, its application in

environmental radioactivity is limited. The paper establishes a regime for the routine use of the

BEGe detector for quantifying activity levels in various types of samples, including standards,

minerals, soil, tailings and organic materials. It provides details of sample preparations,

geometry selection, efficiency calibration, required corrections and counting time. The study

goes further to examine the rapid analysis of 226Ra in several materials and the determination of

227Ac by gamma spectrometry.

Chapter 7 concludes and summarises the main findings of the project and its

significance and then suggests avenues for further research areas.

5

Chapter 2 Literature Review

2.1 Introduction

In the last few decades, there has been a growing concern about exposure to NORM. This

is after the latest findings on health risks associated with the inhalation of radon and its progeny

and the detection of enhanced levels of NORM in non-nuclear industries (IAEA, 2012; ICRP,

2010; WHO, 2009). Numerous studies have been published describing various aspects related

to NORM, including measurement techniques, exposure characteristics and regulation policies.

This chapter aims to summarise the current state of knowledge relevant to the measurement

and regulation of NORM and highlights some gaps in the current knowledge, to which this

research project makes a unique contribution.

2.2 Origin and distribution

Naturally occurring radionuclides are ubiquitous in the environment and form most of

natural background radiation (UNSCEAR, 2008). They are widespread in the environment and

can be found in soil, sand, rock, minerals, air, water and biota (Al-Masri et al., 2000; Kurnaz et

al., 2007; Zare et al., 2012). Naturally occurring radionuclides can be classified into two

categories: (i) primordial radionuclides, which have existed on the Earth’s crust since it was

formed, (ii) and cosmogenic radionuclides, which have been formed by the spallation process of

the constituents of the atmosphere when they interact with the primary cosmic rays in the

upper atmosphere (Beer et al., 2012).

Primordial radionuclides 40K, 238U, 235U, 232Th and their radioactive decay products are of

the greatest importance in the radiation protection context due to their health hazards and long

half-lives, which are comparable with the age of the Earth. These radionuclides undergo

spontaneous radioactive decay, forming a lighter nuclide and emitting alpha or beta particles

accompanied by gamma radiations. Heavy radionuclides 238U, 235U and 232Th decay through a

chain of radionuclides following different decay modes before reaching a stable state (Figure

2.1). Ideally, radionuclides in each decay chain are assumed to be in a state of equilibrium, in

which the activities of all radionuclides within each decay chain are identical. However, this

equilibrium condition is subject to large disturbances in which activities of radionuclides within

the decay series are no longer equal (Ji et al., 2015; Papadopoulos et al., 2013).

The levels of naturally occurring radionuclides vary widely in the environment.

Typically, they are considered to be low, with global average concentrations of 33, 45 and 420

Bq kg-1 for 238U, 232Th and 40K , respectively (UNSCEAR, 2008). The global population weighted

6

annual effective dose rate is 2.03 mSv (UNSCEAR, 2008). However, in specific areas in the world,

such as Guarapari, Brazil; Kerala, India; Ramsar, Iran; and Yangjiang, China, these levels are

elevated naturally due to the geological and geochemical characteristics of the region (Mohanty

et al., 2004), resulting in a significant effective dose level, as much as 131 mSv per year (Sohrabi

and Esmaili, 2002). These areas are termed High Natural Background Radiation (HNBR) areas

(Hendry et al., 2009). There is another situation where the level of naturally occurring

radionuclides has been technologically altered by certain human activities (Baxter, 1996;

O'Brien and Cooper, 1998; Paschoa and Steinhäusler, 2010). This alteration may be deliberate,

as in uranium mines (USEPA, 2008), or unintentional, as in oil and gas industries (Hilal et al.,

2014).

2.3 Terminology

In the context of radiation protection, NORM is defined as material that is designated by

national law or a regulatory body as being subject to regulatory control because of its

radioactivity, containing no significant amounts of radionuclides other than radionuclides of

natural origin (IAEA, 2007b). The acronym NORM is normally used to refer to the primordial

radionuclides 40K, 238U, 235U, 232Th and their radioactive decay products. Elevated levels of NORM

due to certain human activities, which may alter the physical, chemical or radiological

properties of the radioactive material in a way that increases exposure, are also included in this

term. The term is also used by other international organisations, such as the International

Commission on Radiological Protection (ICRP, 2007a) and the European Commission (EC,

2001).

However, the use of the acronym NORM is not globally adopted to describe condition of

natural radioactive materials subject to human activities. For example, Baxter (1996)

mentioned that the use of the term NORM to describe increased levels of natural radionuclides

in the industry was misleading because it implied that the concentration was as it was in nature.

The author suggested using the abbreviation TENORM (technologically enhanced naturally

occurring radioactive material) instead. Historically, the terminology used to describe elevated

exposure due to natural sources in non-nuclear industries was first coined by Gesell and

Prichard (1975) as technologically enhanced natural radiation (TENR). In the United Sates, both

terms, NORM and TENORM, are used to distinguish between natural levels of naturally

occurring radionuclides in materials and concentrated levels, as a result of human activities

(USEPA, 2008). The Canadian authority has also used both terms, but in such a way to

differentiate between non-nuclear and nuclear industries as NORM and TENORM, respectively

(Canada, 2014). Nonetheless, the IAEA intends to continue using the term NORM in its future

publications as shown in the revision draft of its Safety Glossary 2007 Edition (IAEA, 2015b).

7

Element Uranium series Thorium series Actinium series

Uranium

238U 4.47×109

a

234U 2.45×105

a

235U 7.04×10

8 a

Protactinium

234mPa 1.17

min

231Pa 32760

a

Thorium

234Th 24.1 d

230Th 7.5 ×104

a

232Th 1.41×1010

a

228Th 1.912

a

231Th 25.52 h

227Th 18.72 d

Actinium

228Ac 6.15 h

227Ac 21.77

a

Radium

226Ra 1600 a

228Ra 5.75 a

224Ra 3.66 d

223Ra 11.44 d

Radon

222Rn 3.824 d

220Rn 55.6 s

219Rn 3.96 s

Polonium

218Po 3.05 min

214Po 164 μs

210Po 138.4d

216Po 0.145 s

64%

212Po 0.299

μs

215Po 1.781 ms

211Po 0.516 s

Bismuth

214Bi 19.9

min

210Bi 5.01 d

212Bi 60.55

min

211Bi 2.14

min

Lead

214Pb 26.8 min

210Pb 22.3 a

206Pb Stable

212Pb 10.64

h

36%

208Pb stable

211Pb 36.1 min

207Pb stable

Thallium

208Tl 3.053

min

207Tl 4.77

min

Figure 2.1 Uranium, thorium and actinium decay series

8

2.4 Behaviour of NORM

The behaviour of NORM in the environment is governed not only by the physical and

chemical properties of its individual radionuclides, but also by the type of human activities

taking place (Michalik et al., 2013; O'Brien and Cooper, 1998). NORM contains a wide range of

radionuclides with great variations in their physicochemical properties. These properties allow

certain radionuclides to be released into the surrounding environment and migrate from their

source of origin, causing disruption in the radioactive equilibrium state of the natural series.

The radionuclides released into the environment may establish their own subseries away from

the original series. In the 238U decay series, several subseries, headed by 230Th, 226Ra, 222Rn and

210Pb, have the potential to be separated. For example, 222Rn is gaseous and can easily be

released from the ground and enter buildings and establish its own subseries in indoor air

(Wang and Ward, 2002). The short-lived progeny of 222Rn may result in the presence of another

subseries headed by 210Pb in indoor environments (Chen et al., 2009; Samuelsson and

Johansson, 1994) and outdoor surfaces (Doering et al., 2006).

In industrial practices, the behaviour of NORM can be influenced by the type of industry

and operational process (Baxter, 1996; Cooper, 2005; Michalik et al., 2013). Each industry has

its own characteristics, and thus, the same radionuclide can exhibit quite different behaviour

between one industry and another. For example, 226Ra in the scale of the petroleum industry

(Abo-Elmagd et al., 2010; Jibiri and Amakom, 2011) behaves differently compared to how 226Ra

behaves in phosphogypsum in the phosphate industry (Cañete et al., 2008; Haridasan et al.,

2002). It is therefore important to characterise NORM in each industry in order to control it.

2.4.1 Uranium

Uranium (U) has an atomic number of 92 and is found naturally as three different

isotopes: 238U (99.2740%), 235U (0.7200%) and 234U (0.0055%) (Eisenbud and Gesell, 1997). The

behaviour and occurrence of uranium in the environment depends on oxidation and reduction

processes (Landa, 2007; Swarzenski et al., 2003). The most dominant U species in the

environment are tetravalent U(IV) and hexavalent U(VI) (Sheppard, 1980). The latter forms

soluble complexes with carbonate, oxalate and hydroxide ions and hence can be readily

transported through environmental media (Grenthe, Wanner and Forest 1992; Sheppard et al.

2005; Krishnaswami and Cochran 2011; Anderson, Fleisher and LeHuray 1989). On the other

hand, under reducing conditions, U becomes insoluble and occurs in the tetravalent state with

the tendency to bind with organic materials (Fredrickson et al., 2000; Gascoyne, 1992).

2.4.2 Thorium

Thorium (Th) exists naturally as six isotopes—234Th, 232Th, 231Th, 230Th, 228Th and 227Th.

Thorium isotopes are present in all the natural decay series. Th mainly occurs in the tetravalent

Th(IV) state. It is generally an insoluble element and less mobile (Chabaux et al., 2003;

9

Swarzenski et al., 2003). In aqueous media, the level of pH can play a role in Th solubility (Clark

et al., 2011; Degueldre and Kline, 2007; Langmuir and Herman, 1980; Neck et al., 2002). In near

neutral pH, Th is removed from the dissolved phase onto colloids and particulates (Landa, 2007;

Porcelli and Swarzenski, 2003). At low pH, Th undergoes hydrolysis (Katz et al., 1987) and

becomes more soluble and mobile (Altmaier et al., 2005). In soil, the mobility of Th may be less

affected by pH changes (Hansen and Huntington, 1969).

2.4.3 Radium

Radium (Ra) has four natural radioactive isotopes: 228Ra, 226Ra, 224Ra and 223Ra. Like Th,

Ra exhibits only one oxidation state (Ra2+) and is sensitive to pH changes (Almeida et al., 2004;

El Arabi et al., 2006; Martin et al., 2003; Thiry and Van Hees, 2008). Extensive studies on Ra

behaviour have shown that Ra in water is soluble as chloride, sulfate, carbonate, nitrate and

bromate salts, while it is insoluble as carbonate, iodate, oxalate and sulfate salts (Gao et al.,

2010; Grundl and Cape, 2006; Langmuir and Riese, 1985; Lin et al., 2003; Martin et al., 2003;

Martin and Akber, 1999; Sebesta et al., 1981; Zhang et al., 2014). In soil, Ra is adsorbed by soil

particles, such as clay minerals and organic matter (Landa and Reid, 1983; Shoesmith, 1984;

Totten Sr et al., 2007). However, some Ra can leach from soil or rock into the groundwater,

where it can easily migrate (Cañete et al., 2008; Chalupnik, 2005; Desideri et al., 2008).

2.4.4 Radon

Radon (Rn) is a radioactive gas that is derived from the radioactive decay of Ra in three

natural series as 222Rn, 220Rn and 219Rn. As a noble gas, Rn is chemically inactive, and it does not

precipitate in solid phase, although it can compound with fluorine under exceptional conditions

(Chernick et al., 1962). Rn is transported from subsurface to the ground mostly through

diffusion, which is proportional to the gradient of gas concentration in interstitial spaces;

additionally, advection mechanisms, induced by pressure differences created by meteorological

conditions, also aid Rn transport (Andersen, 2001; Chauhan and Kumar, 2013; Jacob and

Prather, 1990). As Rn is exhaled from the ground into the atmosphere and decays, its short lived

progeny reacts with water vapour and trace gases, forming clusters with diameters from 0.5 to

5 nm, which are then attached to aerosol particles (Michielsen and Tymen, 2007; Porstendörfer,

1994; Porstendörfer and Reineking, 1999). The concentration of Rn is by far much higher in

indoor air, where it is accumulated, than in open air, where it is diluted (UNSCEAR, 2006). In

water, Rn is soluble, and its solubility is sensitive to water temperature and salinity (Schubert et

al., 2012). It decreases with increasing temperature and salinity. The solubility coefficient of Rn

at 0°C is 0.51 and drops to 0.25 at 25°C (Ball et al., 1991). At the water-air interface, Rn tends to

readily escape to the air (Dulaiova and Burnett, 2006; Ongori et al., 2015; Schubert et al., 2012).

10

2.4.5 Polonium

Polonium (Po) is present in nature in seven isotopes: 210Po, 211P, 212Po, 214Po, 215Po, 216Po

and 218Po. Po has several oxidation states, the most predominant being the Po(+IV), which is

relatively more soluble in natural water than other states (Ansoborlo et al., 2012; Lehto and

Hou, 2011). Overall, Po solubility is only slight and its concentration in natural water is lower

than Rn and Ra. In the atmosphere, Po can be found as a result of natural processes, such as the

radioactive decay of Rn isotopes and volcanic eruptions (Gauthier et al., 2000; Su and Huh,

2002), or industrial processes, such as fossil fuel burning and metal smelting (Al-Masri et al.,

2000; Boryło et al., 2012). Po is a volatile element and has a strong tendency to condense on ash

particles (Sahu et al., 2014). Atmospheric Po is brought to the Earth’s surface through dry and

wet deposition (Persson and Holm, 2011). The residence time of Po in the atmosphere varies

from few days up to 75 days, depending on several factors, including location, altitude and

meteorological conditions (Baskaran and Shaw, 2001; Moore et al., 1973; Papastefanou, 2009;

Persson and Holm, 2011). Because of atmospheric fallout, Po is concentrated on the top layer of

the soil (Karunakara et al., 2000) and is present in plant leaves (Al-Masri et al., 2008) as well as

on water surfaces (Tateda et al., 2003). In soil, Po is likely to be adsorbed into clay and organic

colloids, which play a role in its mobility (Aslani et al., 2005; Özden et al., 2013).

2.4.6 Lead

Lead (Pb) comprises eight naturally occurring isotopes, four of which are stable (204Pb,

206Pb, 207Pb and 208Pb) and four of which are radioactive (210Pb, 211Pb, 212Pb, and 214Pb). They are

formed through the radioactive decay of radionuclides in the uranium, actinium and thorium

decay series, except 204Pb, which is non-radiogenic (Dickin, 1997). Pb has two common

oxidation states: Pb(II), which dominates in inorganic reaction, and Pb(IV), which dominates in

organic matter (Casas and Sordo, 2006). Like Po, Pb exists in the atmosphere with a residency

time of 2 to 25 days (Ahmed et al., 2004; Gäggeler et al., 1995; Mudbidre et al., 2014) and can be

carried by aerosol particles over long distances before it is removed from the atmosphere

through dry and wet deposition (Halstead et al., 2000; Yamamoto et al., 2006). Diurnal and

seasonal variations have been observed for atmospheric Pb: being higher in the night than in

the day with maximum level during fall and winter and minimum level during spring and

summer (Sheets and Lawrence, 1999). Pb is precipitated or bound to soil solids, and its mobility

is influenced by soil redox potential, available anions (e.g. carbonate, phosphate and sulfate),

pH, organic contents and presence of cations (Kabala and Singh, 2001; Reddy et al., 1995;

Roussel et al., 2000; Sauvé et al., 1998; Violante et al., 2010; Wang and Benoit, 1996). As the soil

pH increases, Pb forms complexes with organic and inorganic ligands which may enhance Pb

mobility. In aqueous solutions, the solubility of Pb decreases as the pH increases (Dando and

Glasson, 1989; Sauvé et al., 1998).

11

2.5 Exposure to NORM

NORM has the potential to pose significant radiological hazards to humans and the

environment (Vearrier et al., 2009). Exposure pathways by which NORM can cause health risks

may depend on the human activity that alters the level of NORM. Several pathways are involved

in complex scenarios, but they can be grouped into two broad categories: (i) internal exposure;

and (ii) external exposure. The determination of exposure pathways is crucial in terms of

assessing the radiation dose from NORM.

2.5.1 Internal exposure

Internal exposure occurs when radionuclide members of NORM enter the body through

inhalation and ingestion. As radionuclides decay inside the body, various types of ionising

radiations are emitted. The greatest effects are associated with alpha radiation, due to its high

linear energy transfer (LET) in tissues (Cember and Johnson, 2008). The presence of radioactive

gases and metallic radionuclides has been observed in indoor and outdoor environments as well

as in dwellings and workplaces (Arafa et al., 2002; Gründel and Porstendörfer, 2004; Murty et

al., 2010; Zeevaert et al., 2006). Airborne radionuclides can be easily transported for longer

distances since their diffusion coefficient in air is higher than in other material media

(Raghunath and Kotrappa, 1979). Radionuclides in the air readily attach to airborne particles,

such as moisture and dust, and their transport mechanism in air is mainly governed by

advection (Leelőssy et al. 2014). The most frequently detected airborne radionuclides are 210Pb,

210Po, 222Rn and 220Rn and their short-lived progeny (Baskaran, 2011; Jia and Torri, 2007; Murty

et al., 2010).

Most of the attention has been given to exposure to 222Rn in indoor environments (IAEA,

2015a; ICRP, 2014). This is because of available data relating health risks to exposure levels and

the presence of 222Rn in all buildings and closed spaces (Darby et al., 2005; UNSCEAR, 2006).

Inhalation of 222Rn and its short-lived progeny is the second main cause of lung cancer (WHO,

2009). The health risk that is commonly attributed to 222Rn is actually associated with its short-

lived progeny radionuclides 218Po, 214Pb and 214Bi that are produced in air; most of inhaled 222Rn

gas is exhaled from the lungs before it decays, due to its relatively long half-life of 3.8 days

(UNSCEAR, 2006). Based on the review of recent epidemiological studies on the association

between exposure to 222Rn and its progeny and lung cancer, the ICRP has announced that the

risk from exposure due to the inhalation of 222Rn progeny is double that of the current

estimation (ICRP, 2010). These epidemiological studies included residential and occupational

exposures and observed a significant association with the risk of lung cancer at average annual

222Rn concentration of approximately 200 Bq m-3 (ICRP, 2010).

Furthermore, some industrial processes, such as dry milling and smelting processes

generate a large amount of dust contaminated with NORM. This radioactive dust was observed

12

in a number of industries, such as the petroleum industry (Bakr, 2010; Potiriadis et al., 2011),

uranium mining and milling (Shawky et al., 2002), minerals sand extraction (Ballesteros et al.,

2008) and the phosphate industry (Samad et al., 2012). The high volatility of some

radionuclides (e.g. 210Po and 210Pb) allows their release and accumulation in industrial

processes involving heating and smelting (Jia and Torri, 2007). In a zircon refining plant, for

example, the main exposure pathway was inhalation of vaporised 210Pb and 210Po in airborne

dust in the processing plant (Cooper, 2005).

NORM can find its way into the human body through ingestion of water and foodstuff

(Abbady, 2006; Awudu et al., 2012). Doses from these pathways depend on the consumption

and the concentration of NORM in the terrestrial and aquatic environment. Some radionuclides,

such as 228Ra and 226Ra, are more soluble in an aqueous medium (Shuktomova and Rachkova,

2011; Veguería et al., 2002) and, hence, can leach from industrial waste and residue streams

and dissolve into the surrounding water. It has been detected in drinking water (Borio et al.,

2007; da Costa Lauria et al., 2012), groundwater (Almeida et al., 2004; Godoy and Godoy, 2006;

Martin and Akber, 1999), produced water during oil and gas extraction (Al-Masri, 2006;

Barescut et al., 2009; Zhang et al., 2014) and all natural water with different concentrations

(UNSCEAR, 2008). The possible exposure pathways may be direct, by drinking water and

inhaling 222Rn released from water, or indirect, by consuming vegetation irrigated with

contaminated water (Martin and McBride, 2012; Medley et al., 2013).

Natural radionuclides in soil can be transferred to plant and vegetation. They have been

detected in different food categories, including fruit, vegetable, dairy, meat and fish (Froidevaux

et al., 2006). The concentrations of natural radionuclides in foodstuff vary according to the food

category, their concentrations in the soil, their physicochemical properties and the prevailing

conditions (Alharbi and El-Taher, 2013; IAEA, 2010). Many studies have shown that

contamination of edible terrestrial organisms can be caused by the discharge of NORM from

industries (Baker and Toque, 2005; Rajaretnam and Spitz, 2000; UNSCEAR, 2008). For example,

Baker and Toque (2005) showed that radium was absorbed by plants in a contaminated site.

210Po in fly ash discharged from a coal-fired power plant was found in food crops nearby

(Zeevaert et al., 2006). Moreover, an elevated level of 226Ra has been detected in vegetation in

the vicinity of oil and gas industry sites (Rajaretnam and Spitz, 2000). However, it should be

noted that for public exposure due to natural radiation, the ingestion of radionuclides is

generally the least significant exposure pathway (UNSCEAR, 2008).

2.5.2 External exposure

External exposure refers to any exposure resulting from a source outside the body. The

dose from an external pathway is derived mainly from gamma radiation, due to its long

travelling distance and high-penetrating power (Martin et al., 2012). A large number of

13

radionuclides of natural origin are gamma emitters (Gilmore, 2008). The population-weighted

annual effective dose from external exposure due to these radionuclides in indoor and outdoor

environments is estimated to be 0.41 and 0.07 mSv respectively (UNSCEAR, 2008).

In indoor environments, external exposure to gamma radiation originates from

primordial radionuclides in the soil and building materials (IAEA, 2015a). Construction

materials such as marble and granite are known to host a high level of 226Ra, 232Th and 40K

(Beretka and Mathew, 1985; Bou-Rabee and Bem, 1996; Pavlidou et al., 2006; UNSCEAR, 2008;

Yu et al., 2000). Their concentrations depend on the industry type and the raw materials used to

manufacture the building materials. Industrial by-products, such as phosphogypsum and

residues, such as red mud generated in aluminium production, which is known to have NORM

contents, are also used to produce building materials (Cooper, 2005; Somlai et al., 2008). These

materials sometimes contain elevated levels of NORM (UNSCEAR, 2008).

In industries, external exposures to NORM are mainly experienced by workers. For a

given activity concentration of radioisotopes, the geometry of the bulk material, which is shaped

by the industry type, plays a critical role in gamma exposure. In the petroleum industry, gamma

exposure comes from gamma-emitting radionuclides 228Ra and 226Ra and their progeny

accumulated in the sludge and scale inside pipes and equipment (Abo-Elmagd et al., 2010; Hilal

et al., 2014; Jibiri and Amakom, 2011; Landsberger et al., 2013). Nevertheless, external exposure

can be controlled by removing the source or by controlling three parameters: duration of

exposure, distance from the source and use of shielding (Martin et al., 2012).

2.5.3 Natural background radiation

According to UNSCEAR (2000) Report Annex E, people working in uranium mines

receive an annual average dose of 5.0 mSv. This value is above the worldwide average annual

exposure to the natural background radiation, which is 2.4 mSv (UNSCEAR, 2008). Examining

detail about the global distribution of NORM and hence NORM-related industries, the annual

dose is quite uneven. The UNSCEAR (2008) Report discusses regional variations in NORM

radiation dose and also highlights certain areas where the dose rises above the global averages.

Annual doses of a few millisieverts is expected depending on the exposure scenario. The Report

concluded that the full picture of worldwide exposure to NORM industries is far from being

completed.

2.6 Regulation of NORM

Although the presence of natural radionuclides in industries was known from the early

20th century, when radium and radon (or radioactive gases at that time) were detected in crude

petroleum (Burton, 1904; Himstedt, 1904), the radiological protection from natural

radionuclides was a more recent concern. The radiological protection community started to

14

recognise NORM as a serious threat in non-nuclear industries in the mid-1980s, when elevated

levels of NORM were reported in the oil and gas industry (Kolb and Wojcik, 1985; Smith, 1987).

Since then, the acronym NORM began to appear in the literature as well as in radiological

protection recommendations and guidelines.

The current radiological protection system is actually developed for artificial

radionuclides. The concept of exemption, for example, is mainly designed for artificial sources of

radiation in planned exposure situations, where the source of exposure is introduced

deliberately (ICRP, 2007b). Recommendations and policies for NORM are included in general

guidance with other radionuclides with no complete guidance for controlling different

situations of NORM (IAEA, 2014; ICRP, 2007a). This has created complexities in regulating

NORM, because scenarios of exposure to NORM are often quite different from those to artificial

radionuclides (Michalik, 2009). A special case of NORM concerns radon in dwellings and

workplaces, and this is treated separately.

In general, exposure to natural sources, including NORM, is considered to be an existing

exposure situation, as the source already existed when the decision to regulate was made; with

the potential to be subject to the requirements of a planned exposure situation because human

intervention and industrial processes could alter its distribution (IAEA, 2014; ICRP, 2007a).

This can generate confusion among users of the standards (Hedemann-Jensen, 2010). Haridasan

(2015) believed that identifying a situation as either an existing or a planned exposure, and

accordingly using appropriate reference levels or dose constraints, is one of the main radiation

protection challenges. The IAEA (2014) also recognised the interference between the two

situations and stated, “Some exposures due to natural sources may have some characteristics of

both planned exposure situations and existing exposure situations”. To address this problem,

the IAEA (2014) has specified a criterion where exposure to natural sources should be treated

as a planned exposure situation:

(1) If the activity concentration of any radionuclides in the 238U and 232Th decay series

exceeds 1 Bq g-1, or 10 Bq g-1 in the case of 40K

(2) Public exposure due to discharge or waste from the situation outlined above

(3) Exposure due to 222Rn and 220Rn and their decay products in which NORM is

controlled as a planned exposure situation

(4) Exposure due to 222Rn and its decay products where its annual concentration in the

workplace remains above the reference level that is proposed by the relevant

regulatory authority, after applying reduction measures.

The values of activity concentrations, that is, 1 and 10 Bq g-1, were selected as the

optimum boundaries between the ubiquitous unmodified concentrations of the nominated

radionuclides in soil and the activity concentrations in ores, mineral sand, industrial residues

15

and waste (IAEA, 2005). No scenarios were involved in deriving the values of activity

concentrations. According to the IAEA (2004a), these values would be unlikely to deliver to

individuals doses exceeding about 1 mSv in a year. In reality, exposure to NORM involves

complex pathways (O'Brien and Cooper, 1998). An example was given by the ICRP (2007b),

which showed that activity concentrations of approximately 0.2 Bq g-1 of 238U and 232Th in

equilibrium with their progeny in waste rock piles could deliver to the public an exposure dose

of more than 1 mSv per year. This indicates that the activity concentration alone does not

deliver the same exposure dose. Moreover, it has been noted that the 1 Bq g-1 criterion has been

misinterpreted as a limit (as concluded by the Seventh International Symposium on Naturally

Occurring Radioactive Material, China, 2013), although the 1 Bq g-1 criterion is widely accepted

(Haridasan et al., 2015).

There is a wide variation across the globe in defining the scope of the NORM regulation

and implementing the regulation. National and international standards on regulating NORM are

sometimes ambiguous (ICRP, 2007b). There is also a notable discrepancy in managing NORM

industries, such as in the case of radium that is well-regulated when it is encapsulated and left

uncontrolled, as when it is found in residues (Michalik et al., 2013). Such situations can lead to

regulatory challenges in controlling exposures to NORM (Chen, 2015; Haridasan, 2015).

Michalik et al. (2011) argues that one of the most generic problems with NORM is the

identification of circumstances where enhanced levels can be found. The IAEA acknowledged

this fact by publishing a safety report to identify industrial sectors and processes, other than

uranium mines, involving NORM that were likely to require regulatory considerations (IAEA,

2006). Eleven industries have already been identified to have enhanced levels of NORM due to

their processes, including extraction of rare Earth elements; production and use of thorium and

its compounds; production of niobium and ferro-niobium; mining of ores other than uranium

ore; production of oil and gas; manufacturing of titanium dioxide pigments; operation in the

phosphate industry; operation in the zircon and zirconia industries; production of tin, copper,

aluminium, zinc, lead, iron and steel; combustion of coal; and water treatment. There are other

industries that have been reported by researchers where NORM could be present and yet to be

identified by regulatory bodies, including fish hatchery (Kitto et al., 1998), geothermal power

installation (Eggeling et al., 2013) and recycling of metal scraps (Ortiz and Carboneras, 2011).

To date, NORM is only managed in five non-industries, namely, the oil and gas industry,

phosphate industry, the zircon and zirconia industries, the production of rare earths from

thorium-containing minerals, and the titanium dioxide and related industries. Guidelines for

managing NORM in these industries are available (IAEA, 2003c, 2007a, 2011, 2012, 2013d).

General guidance, such as Management of NORM Residues (IAEA, 2013a), are applied to the rest

of the industries. However, the use of a general approach may be inappropriate because the

16

exposure scenario may vary on a case-by-case basis, depending upon the type of industry and

the industrial process (ICRP, 2007b).

The regulation of radon in dwellings and workplaces appears to have more attention

than other radionuclide members of NORM. This is because of its presence at many places of

human occupancy and confirmed radiological hazards (Darby et al., 2005; WHO, 2009).

Guidelines and policies on the management and control of radon exposure have already been

published and applied (IAEA, 2014, 2015a; ICRP, 2014). Radon is normally managed as an

existing exposure situation where a reference level has to be set (IAEA, 2014). Regulatory

authorities are encouraged to establish an action plan for controlling exposure to indoor radon,

which requires conducting a measurement to identify radon-prone areas and set appropriate

reference levels (IAEA, 2015a; WHO, 2009). Reference levels of less than 300 Bq m-3 was

mentioned internationally for indoor radon as in the recent published recommendations (EU,

2013; IAEA, 2014; ICRP, 2014). In two cases, radon exposure is managed as planned exposure

situations: (i) in practices where NORM is managed as a planned exposure situation; (ii) in

workplaces where the concentration of radon exceeds the specified reference level (IAEA,

2014).

There is emphasis on regulatory bodies to consider methodologies for the sampling,

averaging, monitoring and detection of radionuclides when deriving activity concentrations for

regulatory purposes (IAEA, 2005). Errors in measurement can result in errors when including

or excluding materials from a regulation. For example, Lavi et.al. (2004) found that, in gamma

spectrometry, ignoring the interference that occurred between gamma lines of 228Ac and 40K at

1460 keV could lead to errors up to 150% in 40K result. In oil and gas industries, the

measurement of radioactivity in scale inside pipes, pumps, valves and other equipment was

affected by pipe geometry, which represents one of the major issues in managing NORM in

waste stream (ARPANSA, 2005). Lack of an adequate infrastructure for analysing and

interpreting NORM was found to be a problem in managing NORM (Haridasan, 2015).

Therefore, a number of reports have been provided by regulatory bodies to standardise

analytical procedures for the determination of natural radionuclides and to ensure that specific

criteria of accuracy are met (USEPA, 1993; IAEA, 2013c; ISO, 2012).

2.7 Measurement of NORM

A wide range of methods and techniques are available for measuring the activity

concentration of various radioisotopes in materials (Knoll, 2010; Turner, 2007). The selection of

an analytical technique and measurement method is primarily dictated by the purpose for

which the measurement is being conducted (Macášek, 2000). Radionuclides of natural origin

are ubiquitous and present in almost all environments, including soil, sediments, waste,

17

residues, by-products, building materials and rocks (UNSCEAR, 2008). These materials are also

the prime source of radon gas, which is released from them and accumulated in enclosed spaces

(Pasculli et al., 2014; UNSCEAR, 2006). Therefore, measurements of natural radionuclide

members of NORM in solid media as well as measurements of radon released from the ground,

and radon trapped in indoor air are of utmost importance in the context of radiological

protection. Further, the implementation of radiological protection standards and the

demonstration of regulatory compliance rely on such measurements.

Beside radiological assessment, the measurement of individual radionuclides in the

uranium and thorium series has many applications in various disciplines. For example, 210Pb has

been extensively used as an environmental radiotracer in sediment dating (Abril and Brunskill,

2014; Putyrskaya et al., 2015; Sanchez-Cabeza and Ruiz-Fernández, 2012), atmospheric studies

(Brattich et al., 2015; Heijnis et al., 1993; Lamborg et al., 2000), ocean science (Stewart et al.,

2007; Verdeny et al., 2009) and pollution investigation (Cheng and Hu, 2010; Sakan et al., 2015).

2.7.1 Measurement of radon releases

The release of radon from materials to the atmosphere takes place when radon is

emanated from grains into the interstitial space and then transported to the surface, where it

can be released (Nazaroff, 1992). The amount of radon release over a specific area from a

material surface is referred to as exhalation flux density and is given as the activity per unit of

area and time (Bq m-2 s-1). This exhalation is commonly measured by placing a chamber

(accumulator) directly into the sampling area in an upturned position (Ferry et al., 2002; Ielsch

et al., 2002; Schery et al., 1989), or by enclosing the sample materials in an airtight container

(Jonassen, 1983; Samuelsson, 1990; Tufail et al., 2000) using either passive detectors, which do

not require electrical power to operate but require laboratory analysis, or active detectors that

require electrical power to operate (ISO, 2012). The measurement can be distinguished on the

basis of the sampling duration into instantaneous, continuous and integrated (over a short or

long period of time) measurement methods.

Over the past few decades, a number of passive and active techniques have been

developed to measure radon release from surfaces (Cohen and Cohen, 1983; Iimoto et al., 2008;

Kotrappa et al., 1990; Spehr and Johnston, 1983; Wilkening and Hand, 1960). Passive detectors

are usually cheaper and more suitable for integrating measurement, and a large number of

detectors can be deployed over a wide area for a reasonable period of time. On the other hand,

active detectors are expensive and require periodic maintenance, which renders them less

desirable for field measurement.

One of the best known passive techniques is the nuclear track detector (Fleischer et al.,

1965). It has been used for measuring radon flux density in building materials (Al-Jarallah and

Abu-Jarad, 2001; El-Bahi, 2004; Najam et al., 2013) as well as in soil (Abd-Elzaher, 2012; Oufni,

18

2003; Saad et al., 2013) and fly ash samples (Kumar et al., 2005; Mahur et al., 2008). However,

the sensitivity of nuclear track detectors is relatively low, which makes them more suited for

long-term radon concentration averages than for short-term exhalation flux density (IAEA,

2013b). Electret detectors have also been used to measure radon flux density (Kotrappa and

Stieff, 2009; Kovler et al., 2004; Righi and Bruzzi, 2006). Nonetheless, the performance of

electret detectors is affected by both background gamma radiation (Kotrappa et al., 1990;

Usman et al., 1999) and extreme humidity conditions (Mahat et al., 2001).

Activated charcoal canister is another passive technique that has been used extensively

to measure radon exhalation flux density (Bollhöfer and Doering, 2015; Cosma et al., 1996; De

Jong et al., 2005; Doering et al., 2006; Lawrence et al., 2009; Nisti et al., 2014; Petropoulos et al.,

2001; Spehr and Johnston, 1983; Wang et al., 2009). In industrial applications, activated

charcoal canisters were used to estimate the activity from residues and waste sites (Bollhöfer

and Doering, 2015; Dueñas et al., 2007; Nisti et al., 2014). In an unpublished work by Al-Farsi

(2008), radon flux density from petroleum industry waste samples, including sludge and scale,

were measured using activated charcoal detectors as well as an emanometer detector (an active

radon technique). A good correlation was found between the two methods. The activated

charcoal detector was also used for determining radon exhalation from building materials (De

Jong et al., 2005; Petropoulos et al., 2001).

The mechanism of the activated charcoal technique is based on the adsorption of radon

gas on charcoal. When measuring the exhalation flux density, canisters are deployed on the

target area where they are exposed for a few days, typically between one to five days, and then,

the adsorbed radon in the activated charcoal is determined through gamma radiations emitted

from its daughters using gamma spectrometry (Bollhöfer et al., 2006; Countess, 1976; Dueñas et

al., 2007; Lawrence et al., 2009). The technique has many advantages, including low cost,

reusability and simplicity. It is also recommended by some regulatory bodies for regulatory

purposes (ARPANSA, 2008). However, changes in environmental conditions, such as humidity,

temperature and pressure, have the potential to affect the sensitivity of activated charcoal (El

Samman and Arafa, 2002; Iimoto et al., 2008; Ronca-Battista and Gray, 1988; Vargas and Ortega,

2007). Moreover, this technique is based on the assumptions of constant radon flux density over

exposure time and retaining all the radon that entered the canister (Spehr and Johnston, 1983).

Although these assumptions may not always be valid in the real world, it has been widely

accepted for this method.

Radon flux density can also be measured with an emanometer (Whittlestone et al., 1998;

Zahorowski and Whittlestone, 1996). This is an active detector that is based on a flow-through

or accumulator method. It consists of two scintillation cells coupled with two separated

photomultiplier tubes (PMT) connected to a data acquisition and control unit. The two

19

scintillation cells are separated with a delay line enough for the decay of thoron to allow

thoron/radon discrimination (Zahorowski and Whittlestone, 1996). Lawrence et al. (2009) used

an online emanometer and charcoal canisters to independently measure radon exhalation rates

at different sites in tropical northern regions of Australia. Their measurements covered both the

dry and wet seasons, with their large variations in soil and air moisture. The authors reported

obtaining statistically overlapping results with both methods. The emanometer is limited by the

diffusion losses including the build-up of radon/thoron concentrations in the chamber, back

diffusion and leakages at edges of the chamber, which can be significant (Zahorowski and

Whittlestone, 1996).

2.7.2 Measurement of atmospheric radon concentration

Radon presents in both outdoor and indoor environments. In the outdoors, however, it

is readily diluted to low concentration levels, typically around 10 Bq m-3 with a wide range from

1 to 100 Bq m-3 (UNSCEAR, 2006). In a nationwide survey conducted in Japan, the mean outdoor

radon concentration was found to be 6.1 Bq m-3, which corresponded to 40% of the indoor value

(Oikawa et al., 2003). Popic et al. (2011) observed seasonal variations in outdoor radon

concentrations in decommissioned Norwegian iron and niobium mining sites that were rich in

naturally-occurring radionuclides, being higher in the summer, with an arithmetic mean in the

range of 29–82 Bq m-3, than in fall. This result is in contrast to that obtained by Chan et al.

(2010), who found that outdoor radon concentration was lower in the summer season. This

variation in outdoor radon levels was influenced by metrological conditions, such as rainfall,

wind speed, temperature and humidity, which play an essential role in controlling radon

releases from the ground (Akber et al., 1994; Lawrence et al., 2009; Martin et al., 2004).

Nonetheless, time spent outdoors is generally much less than that spent indoors, which makes

outdoor radon less significant in terms of radiological protection.

On the contrary, radon tends to build up in enclosed spaces and when there is less air

volume exchange, such as buildings, tunnels, caves and underground mines. Epidemiological

studies of residential and occupational exposure have shown evidence of a relationship between

indoor radon exposure and lung cancer (Darby et al., 2005; Krewski et al., 2006; Tomasek et al.,

2008; WHO, 2009). These findings have increased awareness of indoor radon exposure,

whether in dwellings or in workplaces (IAEA, 2015a; ICRP, 2014). Controlling such exposure

can significantly reduce its hazards, and the first step to do so is to determine the activity

concentration of indoor radon levels and the parameters that may influence that concentration

(WHO, 2009).

Most indoor radon is driven from the ground through cracks in the floor (UNSCEAR,

2006). The level of radon in the ground is largely influenced by the underlying geology (Moreno

et al., 2008; Sundal et al., 2004). The authors found a significant correlation between the level of

20

indoor radon and the geological composition of the ground (Moreno et al., 2008; Sundal et al.,

2004). Building materials and water supply can also contribute to indoor radon (Cosma et al.,

2013; Najam et al., 2013; Song et al., 2005). Bavarnegin et al. (2013) found that radium content

in building materials used in Ramsar, Iran, a region with high natural background radiation, was

the main source of indoor radon. The contribution of radon in the water supply to indoor air

was much less compared to other sources, and most of the risks arise from the ingestion of

water containing radon (WHO, 2011).

Seasonal and diurnal variations in indoor radon levels have been observed in dwellings

(Bochicchio et al., 2005; Faheem and Mati, 2007; Prasad et al., 2012; Singh et al., 2005). The

highest level was usually obtained during winter, while the lowest was in summer. These

variations were mostly attributed to ventilation systems, such as air-conditioning and heating

systems (Lee and Yu, 2000; Marley and Phillips, 2001). In regions with colder climates, the

heated air was intentionally retained indoors, where radon builds up. If measurements were

made for a period shorter than a seasonal cycle, then some correction factors would need to be

applied to get an annual average (Baysson et al., 2003; Faheem and Mati, 2007; Pinel et al.,

1995). Alternatively, measurement may be conducted during the coldest months (IAEA, 2003b;

NORDIC, 2000). Seasonal variations of indoor radon were observed less in workplaces, and

some have exhibited no seasonal patterns (Inoue et al., 2013; Papachristodoulou et al., 2010;

Tokonami et al., 1996). Instead, diurnal variations were reported more often in workplaces

(Vaupotič et al., 2012; Zahorowski et al., 1998).

Less attention has been given to indoor thoron (220Rn), and the available data are largely

limited (Burghele and Cosma, 2012; Inoue et al., 2013; Vaupotič et al., 2012). This is perhaps

due to measurement difficulties associated with the short half-life of thoron (56 s). It should be

noted that health risks associated with the inhalation of thoron and its decay products are

comparable with those of radon and its progeny (ICRP, 1993; Shang et al., 2005; UNSCEAR,

2000). Moreover, the concentration of thoron in some indoor spaces has been found to be four

times greater than that of radon (Ramola et al., 2012; Yamada et al., 2006). Additionally, the

exhalation flux density of thoron is generally greater than that of radon (Sharma and Virk,

2001). In the indoor environment, building materials represent the main source of thoron gas

(McLaughlin, 2010; Meisenberg and Tschiersch, 2011), and its level is more influenced by the

distance from the source, which includes the walls and floors in buildings (Fujimoto et al., 1994;

Urosevic et al., 2008).

2.7.3 Measurement of NORM in solid media

The radioactivity in a material such as soil, sediment, rock, products, by-products, waste

and residues is often quantified through ionising radiation emitted during the radioactive decay

of the radionuclide within that material. Alpha, beta and gamma spectrometry are most

21

frequently used for such purposes (Bonotto et al., 2009; El-Daoushy and Hernández, 2002;

Martin and Hancock, 1992; Murray et al., 1987; Vesterbacka et al., 2009). Mass spectrometry

has also been employed to measure mass concentration, hence the activity concentration of

long-lived radioisotopes (Roos, 2014). These measurement methods, with the exception of

gamma spectrometry, involve chemical separation, which is destructive and inconvenient

(Bickel et al., 2000). Therefore, gamma-ray spectrometry can be the best choice for quantifying

the activity concentration of natural radionuclides being able to simultaneously detect a large

number of natural radionuclides in a non-destructive manner.

Germanium (Ge) detectors are by far the most common gamma detectors for activity

concentration measurement. They have higher gamma energy resolution, which allows

distinguishing gamma emission for different radioisotopes. The higher atomic number of

germanium compared to the previously-used silicon (Si) detectors also makes it applicable for a

wide range of energy (Gilmore, 2008). Over the last few decades, several types of high-purity

germanium (HPGe) detectors have been developed. These detectors have different crystal

configurations to fit particular applications. The coaxial HPGe detector has been widely utilised

in numerous environmental studies (Alghamdi and Aleissa, 2015; De Corte et al., 2005; El-

Daoushy and Hernández, 2002; Kaste et al., 2006; Lenka et al., 2009; Mauring et al., 2014;

Putyrskaya et al., 2015; Scholten et al., 2013; Siegel, 2013; Van Beek et al., 2010; Yücel et al.,

2010). Lenka et al. (2009) used a coaxial HPGe detector to measure 238U in soil through the 63

keV gamma peak of its daughter radionuclide 234Th. Al-Masri and Suman (2003) measured

radium isotopes from the NORM waste site of an oil and gas industry using a coaxial detector.

Nonetheless, the coaxial detector has relatively lower efficiency at low gamma energy compared

to the well-type and planar detectors (Han and Choi, 2010).

The well-type detector has also been used to determine the activity concentration of

natural radionuclides (Gourdin et al., 2014; Legeleux and Reyss, 1996; Reyss et al., 1995; Van

Beek et al., 2008). Despite the high efficiency of this detector, it suffers from a high coincidence

summing effect due to the 4π solid angle (Sima and Arnold, 1996). This effect can limit the use

of the well-type detector for measuring radionuclides of natural origin because of their complex

decay schemes and complicated spectral interference (Gilmore, 2008). Another type of HPGe

detector is the planar detector with a small and thin crystal shape. Planar detectors showed a

higher sensitivity at low gamma energy (Loaiza et al., 2011) and have the advantage of

successfully detecting natural radionuclides with gamma emissions in that region, such as 210Pb

and 234Th (Al-Masri et al., 2010; Di Gregorio et al., 2007). However, at a higher energy, the

efficiency of conventional planar detectors is lower than that of well and coaxial detectors (Han

and Choi, 2010). One of the HPGe detectors (manufactured by Canberra Industries Inc.) has

developed a planar detector whose sensitivity at a higher gamma energy has been promoted to

22

be comparable with that of the coaxial detector (Canberra, 2013). The new detector is known as

broad energy germanium (BEGe) detector. According to the manufacturer, the BEGe detector

can effectively cover an energy range from 3keV to 3MeV with suitable resolution and efficiency

for NORM radioactivity measurements. It is, therefore, well suited for the application of NORM.

The BEGe detector has been used extensively in nuclear physics (Aalseth et al., 2011;

Agostini et al., 2011; Budjáš et al., 2013; Di Vacri et al., 2009) and medical science (Fantínová

and Fojtík, 2014; Kramer et al., 2009; Liye et al., 2007; Webb and Kramer, 2001). However, the

use of BEGe detectors in environmental studies involving radionuclides of natural origin is still

limited (Condomines et al., 2010; Putyrskaya et al., 2015). Zhang et al. (2009) used the BEGe

detector to accurately determine the activity concentration of 226Ra in soil. The authors

deconvoluted the interference between 226Ra and 235U at 186 keV peak using Aatami software.

The employment of the BEGe detector for routine analysis of natural radionuclides can bring

benefits to the field of NORM, including the detection of low gamma energy emitted from

radionuclides such as 210Pb and 234Th and the determination of radionuclides with low gamma

intensity, such as 230Th. However, like other Ge detectors, there are a number of issues that

need be addressed regarding the use of BEGe detectors in such a field, including efficiency

calibration, geometric effects, coincidence summing and self-absorption effects.

2.8 Measurement challenges

Since the discovery of radioactivity until the present day, there has been significant

development in techniques and instrumentations to quantify the radioactivity of materials. In

the case of NORM, the measurements involve sampling and analysis of a broad spectrum of

radionuclides, which vary in their physicochemical properties in various types of material. In

such cases, difficulties are expected to arise and pose challenges to the field of NORM. These

challenges include, but are not limited to, the cost of analysis, uncertainty of measurement, the

background and the state of equilibrium.

2.8.1 Measurement cost

The cost of the measurement is one of the most important factors in choosing the

appropriate analytical method, in addition to the desired application and detection limit (Tyler,

2014). The appropriate analytical procedure for measuring natural radionuclides of interest

needs to be selected carefully to maximise the benefits and minimise the cost. This cost includes

financial cost, which is required for instrumentation, resource, operation and maintenance and

time required to complete the analysis. NORM radioisotopes' analytical methods generally

involve time and resources and, therefore, are more expensive (Aycik, 2009; L'Annunziata,

2012). NORM has many radionuclides that vary in their chemical and physical properties and

have a ubiquitous distribution within industrial sites, operations, processes, products, by-

23

products, residues and waste (Al-Saleh and Al-Harshan, 2008; Hilal et al., 2014; Kumar et al.,

2005; Palomo et al., 2010; Pontedeiro et al., 2007; Tufail et al., 2006). In each case, specific

radionuclides dominate. The measurement of individual radionuclides is likely to involve

different sampling and analysis techniques due to the differences in the chemical and physical

properties of these materials, which would raise the financial costs of the analysis.

A number of studies have attempted to reduce the cost by reducing the time required for

the analysis (Johnston and Martin, 1997; Palomo et al., 2011; Sengupta et al., 2015; Zapata-

Garcia et al., 2009). Dowdall et al. (2004) measured 226Ra directly from the peak 186 keV by

subtracting the contribution of 235U at the same peak using the activity ratio 235U/238U instead of

waiting for three weeks to achieve secular equilibrium between 226Ra and its daughters 214Pb

and 214Bi, which are usually used to determine the activity concentration of 226Ra. Lin et al.

(2015) developed a method for the rapid and simultaneous detection of alpha and beta

radioactivity in food using liquid scintillation counting. Their method resulted in a reduction of

sample preparation time for a batch of eight samples from three days in traditional methods to

around five hours.

Another way to reduce the cost of the analysis is by using an in situ measurement

technique system (Al-Masri and Doubal, 2013; Aldenkamp et al., 1992; Clouvas et al., 2005;

Tyler, 2014; Wilson et al., 1999). This technique eliminates the need to collect and send samples

to a laboratory for analysis. Povinec et al. (2014) found in situ underwater gamma-ray

spectrometry to be a cost-effective technique for many applications in marine radioactivity. A

comparison between in situ and laboratory analysis of 40K, 208Tl, and 214Bi in seabed samples was

conducted by Androulakaki et al. (2015). They observed a satisfactory agreement between the

two methods with the advantage of in situ measurement in saving time required for laboratory

analysis.

Monte Carlo (MC) simulation is also employed to minimise the cost of NORM

measurement. It has been used to calibrate gamma detectors instead of purchasing expensive

calibration standard sources (Abbas et al., 2002; Britton et al., 2013a; Bronson, 2003; Dejeant et

al., 2014; Guerra et al., 2015; Habib et al., 2014). A sourceless efficiency calibration software

using MC simulation was tested by Bronson (2003) for 13 detectors. The author concluded that

using such a method was more convenient and quicker than the source-based calibration

method. Moreover, the MC simulation was used to determine correction factors required for

some measurements, such as coincidence summing and self-absorption effects (Agarwal et al.,

2011; Britton et al., 2014; Garcıa-Talavera et al., 2001; Huy et al., 2013; Quintana and Montes,

2013). However, the summing effect may be more complicated in samples containing

radionuclides from the uranium and thorium series due to the presence of a large number of

radionuclides that emit several gamma radiations. Additionally, simulation is limited by the lack

24

of precise data from sample geometry and composition, as well as from detector characteristics,

which can be an additional source of error (Britton et al., 2013a). Nonetheless, developing a

cost-effective technique for measuring NORM will remain one of the major challenges of NORM

measurement.

2.8.2 Uncertainties

Radioactive decay is a random process and, therefore, uncertainty is intrinsic in the

analysis of radioactivity. Additionally, systematic errors can also creep into the measurement.

Uncertainties in radioactivity measurement give some indication of the quality of the results

(EURACHEM., 2012). At the same time, they can affect the precision and accuracy of the

measurement of the activity concentration (IAEA, 2003a). Due to its importance in analytical

measurement, a number of guidelines have been published to provide tools for evaluating and

quantifying uncertainty across a wide range of measurement (e.g. EURACHEM., 2012; IAEA,

2004b; ISO, 1995).

The uncertainty associated with the measurement of NORM, in general, can arise from

many components that could be related to the sample itself, including matrix, weighting, water

content, radioactivity and homogeneity; instrumentations, including spectral acquisition,

efficiency calibration and background; and nuclear properties, including radioactive decay,

probability of emission and half-life (Heydorn, 2004; IAEA, 2004b; Xhixha et al., 2013). The

determination of the main source of uncertainty components associated with a particular

measurement is specified by the analytical technique (IAEA, 2004b). For example, in the in situ

gamma measurement of NORM, soil humidity was found to be the main contributor to the total

uncertainty, while in laboratory gamma spectrometry, counting statistics dominated (Al-Masri

and Doubal, 2013). Shakhashiro and Mabit (2009) reported heterogeneity and moisture content

in samples as the largest contributors to the overall uncertainty in their gamma analysis. In

another measurement method, that is, a delayed coincidence counter for the determination of

radium isotopes in water, the associated uncertainties were influenced by the sample volume,

counting time, and elapsed time between sampling and measurement (Garcia-Solsona et al.,

2008).

In order to achieve better statistics when using a detector, such as gamma spectrometry,

long counting time, large sample size and close sample-detector distance are recommended

(Garcia-Solsona et al., 2008; Gehrke and Davidson, 2005; Herranz et al., 2008). However, the

larger sample volume will result in a self-absorption effect (Huy et al., 2013), and the closer

distance to the detector may increase the probability of cascade summing (Garcıa-Talavera et

al., 2001). In addition, the longer counting time will cause the short-lived radionuclides to decay

during the measurement (Abo-Elmagd et al., 2010). The corrections required for these factors

25

are also associated with uncertainties that must be added to the propagated uncertainty of the

measurement.

It should be noted that these uncertainties are associated with the measurement of

radioactivity only. They are essential for any further use of the resulting values. When it comes

to dose assessment, for example, the estimated uncertainties of the radioactivity measurement

must be considered with other sources of uncertainties, such as uncertainties of radionuclide

discharge and transfer to the environment, as well as dose coefficient of intake (Betti et al.,

2004).

2.8.3 The background

Primordial radionuclides from 238U, 235U and 232Th decay series as well as 40K are

ubiquitous and can be found in constructional materials, ambient environments, shielding

materials and even the equipment itself. Hence, a major aspect of the measurement of NORM is

that the radionuclides being measured in the sample are present in the background. They form,

along with cosmic rays, a significant component of the background of the measurement

detectors (Knoll, 2010). Unfortunately, detectors cannot distinguish between radiation

originating from the sample and that from the background. Therefore, the background should

be determined and subtracted from the spectrum (Gilmore, 2008).

Over the last few decades, considerable efforts have been devoted to the reduction of the

background counts of detectors (Alghamdi and Aleissa, 2015; Neder et al., 2000; Niese, 2014;

Povinec, 2008; Semkow et al., 2002). In gamma spectroscopy, lead shielding is usually used to

protect the detector from background radiations. Smith et al. (2008) examined three types of

lead shielding assemblies and found differences in the background radiation in the shielding

material itself. Obviously, low background shielding should be preferred. On the other hand, the

interactions of background radiations with shielding materials can produce a low-energy

bremsstrahlung continuum and induce characteristic X-rays (Alessandrello et al., 1998; Britton

et al., 2013b; Mrđa et al., 2007). Graded shielding, where the internal surface is coated with

elements such as tin, aluminium and copper, was developed for gamma detectors in order to

absorb the lead X-ray and reduce general background radiation (Kramer and Hauck, 2005;

Singh et al., 2011). Moreover, a Compton suppression system was also introduced to reduce

Compton continuum denominated in below 1 MeV of gamma spectrum (Nolan et al., 1994;

Schumaker and Svensson, 2007). In the Compton suppression system, the HPGe detector is

surrounded by an assembly of guard detectors, which will collect the escaping photons from the

HPGe detector.

A further development in reducing background radiation was the introduction of

underground laboratories (Busto et al. 1992; Helbig and Niese 1986; Kaye et al. 1972; Hult et al.

2006; Laubenstein et al. 2004; Povinec, Commanducci and Levy-Palomo 2005; Zeng et al. 2014).

26

Underground laboratories are more efficient in reducing cosmic rays. Nevertheless, such

laboratories are not of great benefit to some detectors, such as gamma detectors because

primordial radionuclides still present in construction materials, and liquid scintillation

detectors because 50% of the background is caused by detectors themselves (Niese, 2014).

Although these solutions are effective in minimising the background in the measurement of

NORM, they may lead to the first challenge, which is cost.

2.8.4 The equilibrium state

In the measurement of radionuclides in the natural decay series, the activity

concentration of a parent radionuclide is often estimated from its daughters and vice versa (El-

Daoushy and Hernández, 2002; Huy and Luyen, 2004; Scholten et al., 2013; Siegel, 2013). Such

measurement is based on the assumption of equilibrium conditions, where the activity

concentrations of the parent and daughter radionuclides are equal in a secular equilibrium or

can be calculated in a transient equilibrium (Debertin and Helmer, 1988). Understanding the

state of equilibrium is vital in terms of dose assessment due to NORM (Lombardo and Mucha,

2008).

Equilibrium conditions are utilised when the radionuclide cannot be directly

determined by the detection method or to overcome difficulties in the measurement. In gamma

spectrometry, for example, the activity concentration of 226Ra is commonly determined through

gamma-rays emitted from its decay products 214Pb and 214Bi after achieving secular equilibrium

(Al-Masri and Aba, 2005; Landsberger et al., 2013; Sartandel et al., 2014). This is more accurate

than using the deconvolution (Saïdou et al., 2008; Zhang et al., 2009) or 238U/235U activity ratio

(Dowdall et al., 2004) when measuring 226Ra directly from its gamma peak at 186.1 keV that

interferes with the 185.7 keV peak of 235U. More examples of radionuclide measurement using

equilibrium are given in Table 2.1.

Table 2.1 Examples of radionuclides in equilibrium

Radionuclide of interest

Measured radionuclide

Typical delay time

Reference

238U 234Th and 234mPa

4 months (Huy and Luyen, 2004; Lenka et al., 2009; Yücel et al., 2009)

226Ra 214Pb and 214Bi 3 weeks (Dowdall et al., 2004; Landsberger et al., 2013; Murray et al., 1987)

228Ra 228Ac 36 hours (Lourtau et al., 2014; Xhixha et al., 2013) 228Th 224Ra

212Pb and 208Tl 3 weeks 2 days

(Awudu et al., 2012; Condomines et al., 2010)

227Ac 227Th and 223Ra 3 months (Köhler et al., 2000; Van Beek et al., 2010)

223Ra 219Rn A minute (Desideri et al., 2008; El Afifi et al., 2006)

27

One of the drawbacks of using equilibrium conditions in the measurement of

radioactivity is the waiting time required to achieve equilibrium, especially for relatively long-

lived daughters. In a well-known example, the measurement of 226Ra through its daughters

required at least three weeks of delay period to attain secular equilibrium (Siegel, 2013). In the

case of the measurement of 238U through 234Th and 234mPa, the delay time is up to four months

(Kaste et al., 2006). Several studies attempted to overcome this issue by developing a shortcut

method (De Corte et al., 2005; Johnston and Martin, 1997; Li et al., 2015; Martin et al., 1995).

For example, Li et al. (2015) studied the possibility of measuring 226Ra through gamma

spectrometry using the 214Pb-ingrowth method without waiting for equilibrium. They reported

good agreement with the results obtained through the traditional method, that is, after

equilibrium.

However, the equilibrium state can be disrupted under certain conditions. This occurs

when members of the radioactive decay series are removed or added to the system, creating a

condition of disequilibrium. This disequilibrium can occur naturally or technologically due to

some human activities, such as in uranium mines, where uranium isotopes are separated from

the rest of the primordial radionuclides in the uranium decay chain (Bollhöfer et al., 2006;

Dejeant et al., 2014). In the environment, disequilibrium is controlled by the behaviour of

individual radionuclides; physicochemical properties play a role in causing leachability and

mobility of radionuclides (Cañete et al., 2008; Rajaretnam and Spitz, 2000; Wang et al., 2012).

This disequilibrium has been largely observed in water (Labidi et al., 2010), sediment (Stewart

et al., 2007) and NORM industrial sites (Abo-Elmagd et al., 2010). Once the equilibrium is

disturbed, it requires days, weeks, months, thousands or even millions of years to be restored,

depending on the half-lives of the radionuclides.

Disequilibrium can be a serious source of error in the measurement of radionuclides in

the decay series (IAEA, 2003a; Mattinson, 2010). Assuming equilibrium, which is not the case in

reality, can lead to an erroneous estimate of the activity concentration of radionuclides of

interest. Charette et al. (2012) attributed the poor agreement in 226Ra gamma measurement

among the participating laboratories of an inter-laboratory comparison experiment on radium

isotopes to insufficient equilibration caused by the escape of radon gas from samples. They

recommended special care to be exercised to ensure the retention of radon gas and to allow

adequate time to establish equilibrium. Charette et al. (2013) studied the influence of a sealing

technique on radon loss from sample containers when measuring 226Ra with gamma

spectrometry. Their results indicated < 6% loss of radon from the sealed container containing

reference materials and found the use of epoxy resin as a sealing material to be the best way to

prevent radon loss.

28

In the field, disequilibrium, especially in the uranium series, has been reported in many

sites, including industrial wastes and residues (Condomines and Rihs, 2006; De Corte et al.,

2005; Gascoyne, 1992; Patra et al., 2013). 226Ra is not expected to be in equilibrium with its

decay products, due to the possibility of radon releases from the surface layer of the soil (Miller

et al., 1994), or due to the leachability of 226Ra from the site (Rajaretnam and Spitz, 2000).

Therefore, the use of in situ measurement at such sites with the assumption of equilibrium can

be misleading. Miller et al. (1994) used 235U peaks at 144 and 163 keV to determine 226Ra in soil

using in situ gamma measurement to avoid the disequilibrium effect due to radon exhalation.

Nuccetelli and Bolzan (2001) attributed the discrepancy between 226Ra activity concentrations

in building materials obtained directly via 186 keV and indirectly from 214Pb using in situ

technique to radon release from building materials.

2.9 Conclusion

This chapter has presented a review of the current literature available on the

measurement, monitoring and regulation of NORM. The regulation of NORM has been given

more attention due to the latest epidemiological findings on the association of lung cancer with

the inhalation of radon and its progeny along with the existence of enhanced levels of NORM in

the oil and gas industry. New research findings in the literature reported the identification of

radon-prone areas and enhanced NORM levels in many non-nuclear industries, including

extraction of rare Earth elements; production of thorium; production of niobium; mining of ores

other than uranium ore; manufacture of titanium dioxide pigments; the phosphate industry; the

zircon and zirconia industries; production of tin, copper, aluminium, zinc, lead, iron and steel;

combustion of coal and water treatment. Each industry has its own exposure characteristics and

pathways. The standards and policies currently employed to regulate NORM are originally

designed for artificial radionuclide. This adoption has caused confusion and has created new

challenges in the radiation protection system across the globe. Some regulatory bodies have

published recommendations and standards, which cover limited aspects of NORM. Current

monitoring and regulation regimes need to be reviewed in light of the most recent knowledge.

Potential regulatory issues that may impact on the development, implementation and

compliance with NORM legislation should be identified.

The core of any NORM regulation is measurement. Three types of measurement are of

utmost importance for the purpose of regulation: radon exhalation flux density from material

surfaces, concentration of radon in indoor air and activity concentration of natural

radionuclides in materials. Radon exhalation flux density is commonly measured using the

activated charcoal canister technique and is recommended to be used for NORM sites by

regulatory bodies (e.g. ARPANSA, 2008). This technique is based on the assumptions that radon

29

exhalation is constant over time and all of the exhaled radon is adsorbed and retained in the

charcoal. While a number of studies have shown the limitations of activated charcoal canisters

due to meteorological conditions, the validity of the stated assumptions is not fully understood.

Therefore, further investigation on the measurement of radon exhalation flux density using the

activated charcoal technique would provide a unique opportunity to examine the set of

assumptions and better understand the potential limitations of such a widely-used technique.

Released radon from materials is either diluted in the open air or accumulated in

enclosed spaces. As radon is readily diluted, its concentration in the open air is of less interest

for radiological protection. More attention has been given to radon in indoor environments.

Epidemiological studies have demonstrated the association between radon exposure and lung

cancer. The first step to control such exposure is to determine the level of radon in the indoor

environment, whether in dwellings or in workplaces. The previously reported data indicate the

dependency of indoor radon on the underlying geology of the region and ventilation systems.

Seasonal and diurnal variations on indoor radon levels have also been observed. However, there

are limited data on indoor levels of radon, as well as thoron, in Australia and other parts of the

world. National and international regulatory bodies encourage conducting measurements of

indoor radon in order to assess radon levels in a country and to establish a reference level as

well as an action plan for controlling radon exposure. Conducting such measurement in

workplaces would be beneficial to assess the levels of radon and thoron in those places and to

investigate the parameters that may influence those levels.

The third important measurement for regulatory applications is quantifying the activity

concentration of natural radionuclide members of NORM in materials. Although alpha, beta and

mass spectrometry were used to measure the radioactivity in materials, these methods involve

chemical separation, which is destructive. Gamma spectrometry is unique in this manner as it is

not destructive and can measure many radionuclides simultaneously. While several types of

gamma spectrometry have been developed, the BEGe detector represents the latest technology

of these detectors with high resolution and good efficiency over a wide spectrum of gamma

energies. This detector has been extensively used in nuclear physics and medical science.

However, in environmental science and NORM, the use of the BEGe detector is very limited. The

employment of such a detector will provide a valuable tool for the routine analysis of NORM.

This will be accomplished with the investigation of the potential use of the BEGe detector in the

field of NORM.

30

References

Aalseth, Craig E, M Amman, Frank T Avignone, Henning O Back, Alexander S Barabash, PS Barbeau, M Bergevin, FE Bertrand, M Boswell and V Brudanin. 2011. "Astroparticle physics with a customized low-background broad energy germanium detector." Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment 652 (1): 692-695.

Abbady, A. 2006. "Level of natural radionuclides in foodstuffs and resultant annual ingestion radiation dose." Nuclear Science and Techniques 17 (5): 297-300.

Abbas, K, F Simonelli, F D’Alberti, M Forte and MF Stroosnijder. 2002. "Reliability of two calculation codes for efficiency calibrations of HPGe detectors." Applied Radiation and isotopes 56 (5): 703-709.

Abd-Elzaher, Mohamed. 2012. "An Overview on Studying 222Rn Exhalation Rates using Passive Technique Solid-State Nuclear Track Detectors." American Journal of Applied Sciences 9 (10): 1653.

Abo-Elmagd, M., HA Soliman, K.A. Salman and NM El-Masry. 2010. "Radiological hazards of TENORM in the wasted petroleum pipes." Journal of Environmental Radioactivity 101 (1): 51-54.

Abril, José-María and Gregg J Brunskill. 2014. "Evidence that excess 210Pb flux varies with sediment accumulation rate and implications for dating recent sediments." Journal of Paleolimnology 52 (3): 121-137.

Agarwal, Chhavi, Sanhita Chaudhury, A Goswami and M Gathibandhe. 2011. "True coincidence summing corrections in point and extended sources." Journal of Radioanalytical and Nuclear Chemistry 289 (3): 773-780.

Agostini, M, E Bellotti, R Brugnera, CM Cattadori, A D'Andragora, A di Vacri, A Garfagnini, M Laubenstein, L Pandola and CA Ur. 2011. "Characterization of a broad energy germanium detector and application to neutrinoless double beta decay search in 76Ge." Journal of Instrumentation 6 (04).

Ahmed, AA, A Mohamed, AE Ali, A Barakat, M Abd El-Hady and A El-Hussein. 2004. "Seasonal variations of aerosol residence time in the lower atmospheric boundary layer." Journal of Environmental Radioactivity 77 (3): 275-283.

Akber, RA, J Pfitzner and A Johnston. 1994. "Wind direction correlated measurements of radon and radon progeny in atmosphere: A method for radon source identification." In Proceedings of the radon and radon progeny measurements in Australia symposium, edited.

Al-Farsi, Afkar Nadhim. 2008. "Radiological Aspects of Petroleum Exploration and Production in the Sultanate of Oman." PhD Thesis, Queensland University of Technology.

Al-Jarallah, MI and F Abu-Jarad. 2001. "Determination of radon exhalation rates from tiles using active and passive techniques." Radiation measurements 34 (1): 491-495.

Al-Masri, MS. 2006. "Spatial and monthly variations of radium isotopes in produced water during oil production." Applied Radiation and Isotopes 64 (5): 615-623.

Al-Masri, MS and A. Aba. 2005. "Distribution of scales containing NORM in different oilfields equipment." Applied Radiation and Isotopes 63 (4): 457-463.

Al-Masri, MS, B Al-Akel, A Nashawani, Y Amin, KH Khalifa and F Al-Ain. 2008. "Transfer of 40 K, 238 U, 210 Pb, and 210 Po from soil to plant in various locations in south of Syria." Journal of Environmental Radioactivity 99 (2): 322-331.

Al-Masri, MS and AW Doubal. 2013. "Validation of in situ and laboratory gamma spectrometry measurements for determination of 226Ra, 40K and 137Cs in soil." Applied Radiation and Isotopes 75: 50–57.

Al-Masri, MS, M Hassan and Y Amin. 2010. "A comparison of two nuclear analytical techniques for determination of 210Pb-specific activity in solid environmental samples." Accreditation and Quality Assurance 15 (3): 163-170.

Al-Masri, MS, S Mamish, Y Budeir and A Nashwati. 2000. "210Po and 210Pb concentrations in fish consumed in Syria." Journal of Environmental Radioactivity 49 (3): 345-352.

31

Al-Masri, MS and H. Suman. 2003. "NORM waste management in the oil and gas industry: The Syrian experience." Journal of Radioanalytical and Nuclear Chemistry 256 (1): 159-162.

Al-Saleh, FS and GA Al-Harshan. 2008. "Measurements of radiation level in petroleum products and wastes in Riyadh City Refinery." Journal of Environmental Radioactivity 99 (7): 1026-1031.

Aldenkamp, FJ, RJ de Meijer, LW Put and P. Stoop. 1992. "An assessment of in situ radon exhalation measurements, and the relation between free and bound exhalation rates." Radiation Protection Dosimetry 45 (1-4): 449-453.

Alessandrello, A, C Arpesella, C Brofferio, C Bucci, C Cattadori, O Cremonesi, Ettore Fiorini, A Giuliani, S Latorre and A Nucciotti. 1998. "Measurements of internal radioactive contamination in samples of Roman lead to be used in experiments on rare events." Nuclear Instruments and Methods in Physics Research Section B: Beam Interactions with Materials and Atoms 142 (1): 163-172.

Alghamdi, Abdulrahman S and Khalid A Aleissa. 2015. "Ultra-low-background gamma spectroscopy for the measurement of environmental samples." Journal of Radioanalytical and Nuclear Chemistry 303 (1): 479-484.

Alharbi, Abdulaziz and A El-Taher. 2013. "A study on transfer factors of radionuclides from soil to plant." Life Science Journal 10 (2): 532-539.

Almeida, RMR, DC Lauria, AC Ferreira and O. Sracek. 2004. "Groundwater radon, radium and uranium concentrations in Regiao dos Lagos, Rio de Janeiro State, Brazil." Journal of Environmental Radioactivity 73 (3): 323-334.

Altmaier, M, Volker Neck, R Müller and Thomas Fanghänel. 2005. "Solubility of ThO2· xH2O (am) in carbonate solution and the formation of ternary Th (IV) hydroxide-carbonate complexes." International Journal for Chemical Aspects of Nuclear Science and Technology 93 (2/2005): 83-92.

Andersen, Claus E. 2001. "Numerical modelling of radon-222 entry into houses: an outline of techniques and results." Science of the Total Environment 272 (1): 33-42.

Anderson, Robert F, Martin Q Fleisher and Anne P LeHuray. 1989. "Concentration, oxidation state, and particulate flux of uranium in the Black Sea." Geochimica et cosmochimica acta 53 (9): 2215-2224.

Androulakaki, EG, C Tsabaris, G Eleftheriou, M Kokkoris, DL Patiris and R Vlastou. 2015. "Seabed radioactivity based on in situ measurements and Monte Carlo simulations." Applied Radiation and Isotopes 101: 83-92.

Ansoborlo, Eric, Philippe Berard, Christophe Den Auwer, Rich Leggett, Florence Menetrier, Ali Younes, Gilles Montavon and Philippe Moisy. 2012. "Review of chemical and radiotoxicological properties of polonium for internal contamination purposes." Chemical Research in Toxicology 25 (8): 1551-1564.

Arafa, W., H. El Samman and A. Ashry. 2002. "Airborne 222Rn concentration in an Egyptian village." Health Physics 83 (1): 105.

ARPANSA. 2005. "Naturally-occurring radioactive material (norm) in Australia: issues for discussion." http://www.arpansa.gov.au/pubs/norm/rhsac_disc.pdf.

ARPANSA. 2008. Management of Naturally Occurring Radioactive Material (NORM), Radiation Protection Series Publication No. 15. Yallambie.

Aslani, MAA, S Akyil, S Aytas, G Gurboga and M Eral. 2005. "Activity concentration of 210 Pb (210 Po) in soils taken from cultivated lands." Radiation Measurements 39 (2): 129-135.

Awudu, AR, A. Faanu, EO Darko, G. Emi-Reynolds, OK Adukpo, DO Kpeglo, F. Otoo, H. Lawluvi, R. Kpodzro and ID Ali. 2012. "Preliminary studies on 226 Ra, 228 Ra, 228 Th and 40 K concentrations in foodstuffs consumed by inhabitants of Accra metropolitan area, Ghana." Journal of Radioanalytical and Nuclear Chemistry 1-7.

Aycik, Gul Asiye. 2009. New techniques for the detection of nuclear and radioactive agents: Springer Science & Business Media.

Baker, AC and C Toque. 2005. "A review of the potential for radium from luminising activities to migrate in the environment." Journal of Radiological Protection 25 (2): 127.

32

Bakr, WF. 2010. "Assessment of the radiological impact of oil refining industry." Journal of Environmental Radioactivity 101 (3): 237-243.

Ball, TK, DG Cameron, TB Colman and PD Roberts. 1991. "Behaviour of radon in the geological environment: a review." Quarterly Journal of Engineering Geology and Hydrogeology 24 (2): 169-182.

Ballesteros, L., I. Zarza, J. Ortiz and V. Serradell. 2008. "Occupational exposure to natural radioactivity in a zircon sand milling plant." Journal of Environmental Radioactivity 99 (10): 1525-1529.

Barescut, J., DØ Eriksen, R. Sidhu, T. Ramsøy, E. Strålberg, KI Iden, H. Rye, K. Hylland, A. Ruus and MHG Berntssen. 2009. "Radioactivity in produced water from Norwegian oil and gas installations–concentrations, bioavailability, and doses to marine biota." Radioprotection 44 (5): 869-874.

Baskaran, M. 2011. "Po-210 and Pb-210 as atmospheric tracers and global atmospheric Pb-210 fallout: a review." Journal of Environmental Radioactivity 102 (5): 500-513.

Baskaran, M and Glenn E Shaw. 2001. "Residence time of arctic haze aerosols using the concentrations and activity ratios of 210 Po, 210 Pb and 7 Be." Journal of Aerosol Science 32 (4): 443-452.

Bavarnegin, E., N. Fathabadi, M. Vahabi Moghaddam, M. Vasheghani Farahani, M. Moradi and A. Babakhni. 2013. "Radon exhalation rate and natural radionuclide content in building materials of high background areas of Ramsar, Iran." Journal of Environmental Radioactivity 117: 36-40.

Baxter, MS. 1996. "Technologically enhanced radioactivity: an overview." Journal of Environmental Radioactivity 32 (1): 3-17.

Baysson, H, S Billon, D Laurier, A Rogel and M Tirmarche. 2003. "Seasonal correction factors for estimating radon exposure in dwellings in France." Radiation Protection Dosimetry 104 (3): 245-252.

Beer, Jürg, K. G. McCracken and R. Von Steiger. 2012. Cosmogenic radionuclides: Springer. Accessed 20 Jun 2012. http://www.qut.eblib.com.au.ezp01.library.qut.edu.au/patron/FullRecord.aspx?p=884863.

Beretka, J and PJ Mathew. 1985. "Natural radioactivity of Australian building materials, industrial wastes and by-products." Health Physics 48 (1): 87-95.

Betti, M, L Aldave De Las Heras, A Janssens, E Henrich, G Hunter, M Gerchikov, M Dutton, AW Van Weers, S Nielsen and J Simmonds. 2004. "Results of the European Commission Marina II Study Part II—effects of discharges of naturally occurring radioactive material." Journal of Environmental Radioactivity 74 (1): 255-277.

Bickel, M, L Holmes, C Janzon, G Koulouris, R Pilviö, B Slowikowski and C Hill. 2000. "Radiochemistry: inconvenient but indispensable." Applied Radiation and Isotopes 53 (1): 5-11.

Bochicchio, F, G Campos-Venuti, S Piermattei, C Nuccetelli, S Risica, L Tommasino, G Torri, M Magnoni, G Agnesod and G Sgorbati. 2005. "Annual average and seasonal variations of residential radon concentration for all the Italian Regions." Radiation Measurements 40 (2): 686-694.

Bollhöfer, A., J. Storm, P. Martin and S. Tims. 2006. "Geographic variability in radon exhalation at a rehabilitated uranium mine in the Northern Territory, Australia." Environmental Monitoring and Assessment 114 (1): 313-330.

Bollhöfer, Andreas and Che Doering. 2015. "Long-term temporal variability of the radon-222 exhalation flux from a landform covered by low uranium grade waste rock." Journal of Environmental Radioactivity 151: 593–600.

Bonotto, Daniel Marcos, TO Bueno, BW Tessari and A Silva. 2009. "The natural radioactivity in water by gross alpha and beta measurements." Radiation Measurements 44 (1): 92-101.

Borio, R., A. Rongoni, D. Saetta, D. Desideri, M.A. Meli and L. Feduzi. 2007. "Natural radionuclides measurements and total dose indicative evaluation in drinking waters of an italian central region." Journal of Environmental Science and Health Part A 42 (11): 1631-1637.

33

Boryło, Alicja, Bogdan Skwarzec and Grzegorz Olszewski. 2012. "The radiochemical contamination (210Po and 238U) of zone around phosphogypsum waste heap in Wiślinka (northern Poland)." Journal of Environmental Science and Health, Part A 47 (5): 675-687.

Bou-Rabee, F and H Bem. 1996. "Natural radioactivity in building materials utilized in the state of Kuwait." Journal of Radioanalytical and Nuclear Chemistry 213 (2): 143-149.

Brattich, E, MA Hernández-Ceballos, G Cinelli and L Tositti. 2015. "Analysis of 210 Pb peak values at Mt. Cimone (1998–2011)." Atmospheric Environment 112: 136-147.

Britton, R, J Burnett, A Davies and PH Regan. 2013a. "Determining the efficiency of a broad-energy HPGe detector using Monte Carlo simulations." Journal of Radioanalytical and Nuclear Chemistry 295 (3): 2035-2041.

Britton, R, JL Burnett, AV Davies and PH Regan. 2013b. "Monte-Carlo based background reduction and shielding optimisation for a large hyper-pure germanium detector." Journal of Radioanalytical and Nuclear Chemistry 298 (3): 1491-1499.

Britton, R, JL Burnett, AV Davies and PH Regan. 2014. "Characterisation of cascade summing effects in gamma spectroscopy using Monte Carlo simulations." Journal of Radioanalytical and Nuclear Chemistry 299 (1): 447-452.

Bronson, FL. 2003. "Validation of the accuracy of the LabSOCS software for mathematical efficiency calibration of Ge detectors for typical laboratory samples." Journal of Radioanalytical and Nuclear Chemistry 255 (1): 137-141.

Budjáš, Dusan, M Agostini, L Baudis, E Bellotti, L Bezrukov, R Brugnera, C Cattadori, A di Vacri, R Falkenstein and A Garfagnini. 2013. "Isotopically modified Ge detectors for GERDA: from production to operation." Journal of Instrumentation 8 (4): P04018.

Burghele, BD and C Cosma. 2012. "Thoron and radon measurements in Romanian schools." Radiation Protection Dosimetry 152 (1-3): 38-41.

Burton, EF. 1904. "XLVIII. A radioactive gas from crade petroleum." The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science 8 (46): 498-508.

Busto, J, D Dassie, F Hubert, Ph Hubert, MC Isaac, C Izac, S Jullian, F Leccia and P Mennrath. 1992. "Ultra low radioactivity measurements in the Frejus Underground Laboratory." Nuclear Physics B-Proceedings Supplements 28 (1): 425-429.

Canada, Health. 2014. Canadian Guidelines for the Management of Naturally Occurring Radioactive Material (NORM). Edited by Canadian NORM Working Group.

Canberra. 2013. "Broad Energy Germanium Detectors (BEGe)." http://www.canberra.com/products/detectors/pdf/BEGe-SS-C40426.pdf (Accessed by 05 November 2013).

Cañete, Socrates Jose P, Lorna Jean H Palad, Eliza B Enriquez, Teofilo Y Garcia and Teresa Yulo-Nazarea. 2008. "Leachable 226Ra in Philippine phosphogypsum and its implication in groundwater contamination in Isabel, Leyte, Philippines." Environmental Monitoring and Assessment 142 (1-3): 337-344.

Casas, José S and José Sordo. 2006. Lead: chemistry, analytical aspects, environmental impact and health effects: Elsevier. (http://reader.eblib.com.au/(S(vssbtyxwm1vjdcjobwsdkfhg))/Reader.aspx?p=274231&o=96&u=Glv179GeyS%2bys746q8UX7Q%3d%3d&t=1452833669&h=5782E95301059CA17360FD4C0A6F63E7B5E1691A&s=22691505&ut=245&pg=1&r=img&c=-1&pat=n&cms=-1&sd=1) (Accessed by 05 October 2013).

Cember, H. and T.E. Johnson. 2008. Introduction to health physics. 4th ed, Medical Physics: McGraw-Hill, Health Professions Division.

Chabaux, F, J Riotte and O Dequincey. 2003. "U-Th-Ra fractionation during weathering and river transport." Reviews in Mineralogy and Geochemistry 52 (1): 533-576.

Chalupnik, S. 2005. "Impact of radium-bearing mine waters on the natural environment." Radioactivity in the Environment 7: 985-995.

Chan, SW, CW Lee and KC Tsui. 2010. "Atmospheric radon in Hong Kong." Journal of Environmental Radioactivity 101 (6): 494-503.

34

Charette, Matthew A, Henrieta Dulaiova, Meagan E Gonneea, Paul B Henderson, Willard S Moore, Jan C Scholten and Mai Khanh Pham. 2012. "GEOTRACES radium isotopes interlaboratory comparison experiment." Limnology and Oceanography: Methods 10 (6): 451-463.

Chauhan, RP and Amit Kumar. 2013. "Study of radon transport through concrete modified with silica fume." Radiation Measurements 59: 59-65.

Chen, J., W. Zhang, D.G. Sandles, R. Timmins and K. Verdecchia. 2009. "210 Pb concentration in household dust: a potential indicator of long-term indoor radon exposure." Radiation and Environmental Biophysics 48 (4): 427-432.

Chen, Jing. 2015. "A discussion on norm guidelines: proposed revisions needed for practical use." Radiation Protection Dosimetry: 1–6.

Cheng, Hefa and Yuanan Hu. 2010. "Lead (Pb) isotopic fingerprinting and its applications in lead pollution studies in China: a review." Environmental Pollution 158 (5): 1134-1146.

Chernick, CL, HH Claassen, PR Fields, HH Hyman, JG Malm, WM Manning, MS Matheson, LA Quarterman, F Schreiner and HH Selig. 1962. "Fluorine compounds of xenon and radon." Science 138 (3537): 136-138.

Clark, Malcolm W, JJ Harrison and TE Payne. 2011. "The pH-dependence and reversibility of uranium and thorium binding on a modified bauxite refinery residue using isotopic exchange techniques." Journal of Colloid and Interface Science 356 (2): 699-705.

Clouvas, A, S Xanthos, G Takoudis, C Potiriadis and J Silva. 2005. "In situ gamma spectrometry measurements and Monte Carlo computations for the detection of radioactive sources in scrap metal." Health Physics 88 (2): 154-162.

Cohen, B.L. and E.S. Cohen. 1983. "Theory and practice of radon monitoring with charcoal adsorption." Health Physics 45 (2): 501-508.

Condomines, M and S Rihs. 2006. "First 226 Ra–210 Pb dating of a young speleothem." Earth and Planetary Science Letters 250 (1): 4-10.

Condomines, M., S. Rihs, E. Lloret and JL Seidel. 2010. "Determination of the four natural Ra isotopes in thermal waters by gamma-ray spectrometry." Applied Radiation and Isotopes 68 (2): 384-391.

Cooper, Malcolm B. 2005. "Naturally occurring radioactive materials (NORM) in Australian industries—review of current inventories and future generation." Report ERS-006 (EnviroRad Services Pty. Ltd.) to the Radiation health and Safety Advisory Council.

Cosma, C, D Ristoiu, A Poffijn and G Meesen. 1996. "Radon in various environmental samples in the Herculane Spa, Cerna Valley, Romania." Environment International 22: 383-388.

Cosma, Constantin, Alexandra Cucoş-Dinu, Botond Papp, Robert Begy and Carlos Sainz. 2013. "Soil and building material as main sources of indoor radon in Băiţa-Ştei radon prone area (Romania)." Journal of Environmental Radioactivity 116: 174-179.

Countess, R.J. 1976. "222Rn flux measurement with a charcoal canister." Health Physics 31 (5): 455-456.

da Costa Lauria, D., E.R.R. Rochedo, M.L.D.P. Godoy, E.E. Santos and S.S. Hacon. 2012. "Naturally occurring radionuclides in food and drinking water from a thorium-rich area." Radiation and environmental biophysics: 1-8.

Dando, Kevin J and Douglas R Glasson. 1989. "Vacuum microbalance studies of lead deposits from natural waters." Thermochimica Acta 152 (1): 87-96.

Darby, S., D. Hill, A. Auvinen, J.M. Barros-Dios, H. Baysson, F. Bochicchio, H. Deo, R. Falk, F. Forastiere and M. Hakama. 2005. "Radon in homes and risk of lung cancer: collaborative analysis of individual data from 13 European case-control studies." Bmj 330 (7485): 223.

De Corte, Frans, Herman Umans, Dimitri Vandenberghe, Antoine De Wispelaere and P Van den Haute. 2005. "Direct gamma-spectrometric measurement of the 226Ra 186.2 keV line for detecting 238U/226Ra disequilibrium in determining the environmental dose rate for the luminescence dating of sediments." Applied Radiation and Isotopes 63 (5-6): 589.

35

De Jong, P., W. Van Dijk, W. De Vries, E.R. van der Graaf and L.M.M. Roelofs. 2005. "Interlaboratory comparison of three methods for the determination of the radon exhalation rate of building materials." Health Physics 88 (1): 59-64.

Debertin, Klaus and Richard G Helmer. 1988. Gamma-and X-ray spectrometry with semiconductor detectors. Amsterdam: North-Holland.

Degueldre, C and A Kline. 2007. "Study of thorium association and surface precipitation on colloids." Earth and Planetary Science Letters 264 (1): 104-113.

Dejeant, Adrien, Ludovic Bourva, Radia Sia, Laurence Galoisy, Georges Calas, Vannapha Phrommavanh and Michael Descostes. 2014. "Field analyses of 238 U and 226 Ra in two uranium mill tailings piles from Niger using portable HPGe detector." Journal of Environmental Radioactivity 137: 105-112.

Desideri, D, C Roselli, MA Meli and L Feduzi. 2008. "Analytical methods for the characterization and the leachability evaluation of a solid waste generated in a phosphoric acid production plant." Microchemical Journal 88 (1): 67-73.

Di Gregorio, DE, JO Fernández Niello, H Huck, H Somacal and G Curutchet. 2007. "210 Pb dating of sediments in a heavily contaminated drainage channel to the La Plata estuary in Buenos Aires, Argentina." Applied Radiation and Isotopes 65 (1): 126-130.

Di Vacri, Assunta, Matteo Agostini, Enrico Bellotti, Carla M Cattadori, Alessio D Andragora, Alberto Garfagnini, Matthias Laubenstein, Luciano Pandola and Calin Ur. 2009. "Characterization of Broad Energy Germanium Detector (BEGe) as a candidate for the GERDA experiment." In Nuclear Science Symposium Conference Record (NSS/MIC), 2009 IEEE, edited, 1761-1767: IEEE.

Dickin, Alan P. 1997. Radiogenic isotope geology. Cambridge: Cambridge University Press. Doering, Che, Riaz Akber and Henk Heijnis. 2006. "Vertical distributions of 210 Pb excess, 7 Be

and 137 Cs in selected grass covered soils in Southeast Queensland, Australia." Journal of Environmental Radioactivity 87 (2): 135-147.

Dowdall, M, ØG Selnæs, JP Gwynn and C Davids. 2004. "Simultaneous determination of 226Ra and 238U in soil and environmental materials by gamma-spectrometry in the absence of radium progeny equilibrium." Journal of Radioanalytical and Nuclear Chemistry 261 (3): 513-521.

Dueñas, C., E. Liger, S. Cañete, M. Pérez and JP Bolívar. 2007. "Exhalation of 222Rn from phosphogypsum piles located at the Southwest of Spain." Journal of Environmental Radioactivity 95 (2): 63-74.

Dulaiova, H and WC Burnett. 2006. "Radon loss across the water‐air interface (Gulf of Thailand) estimated experimentally from 222Rn‐224Ra." Geophysical Research Letters 33 (5).

EC. 2001. Practical Use of the Concepts of Clearance and Exemption-Part II: Application of the Concepts of Exemption and Clearance to Natural Radiation Sources, Radiation protection 122. Luxembourg.

Eggeling, Lena, Albert Genter, Thomas Kölbel and Wolfram Münch. 2013. "Impact of natural radionuclides on geothermal exploitation in the Upper Rhine Graben." Geothermics 47: 80-88.

Eisenbud, Merrill and Thomas F Gesell. 1997. Environmental Radioactivity from Natural, Industrial & Military Sources: From Natural, Industrial and Military Sources. San Diego: Academic press.

El-Bahi, SM. 2004. "Assessment of radioactivity and radon exhalation rate in Egyptian cement." Health Physics 86 (5): 517-522.

El-Daoushy, Farid and Francisco Hernández. 2002. "Gamma spectrometry of 234 Th (238 U) in environmental samples." Analyst 127 (7): 981-989.

El Afifi, EM, MA Hilal, SM Khalifa and HF Aly. 2006. "Evaluation of U, Th, K and emanated radon in some NORM and TENORM samples." Radiation Measurements 41 (5): 627-633.

El Arabi, AM, NK Ahmed and K Salahel Din. 2006. "Natural radionuclides and dose estimation in natural water resources from Elba protective area, Egypt." Radiation Protection Dosimetry 121 (3): 284-292.

36

El Samman, H. and W. Arafa. 2002. "Temperature and humididty consideration for calculating airborne 222Rn using activated charcoal canisters." Health Physics 83 (1): 97-104.

EU. 2013. Euratom Basic Safety Standards Directive., Council directive 2013/59/Euratom. EURACHEM. 2012. Quantifying uncertainty in analytical measurement. 3rd ed.

http://www.citac.cc/QUAM2012_P1.pdf (Accessed by 15 September 2014). Faheem, Munazza and N Mati. 2007. "Seasonal variation in indoor radon concentrations in

dwellings in six districts of the Punjab province, Pakistan." Journal of Radiological Protection 27 (4): 493.

Fantínová, K and P Fojtík. 2014. "Monte Carlo simulation of the BEGe detector response function for in vivo measurements of 241 Am in the skull." Radiation Physics and Chemistry 104: 345-350.

Ferry, Cécile, Patrick Richon, Alain Beneito and MarieChristine Robé. 2002. "Evaluation of the effect of a cover layer on radon exhalation from uranium mill tailings: transient radon flux analysis." Journal of Environmental Radioactivity 63 (1): 49-64.

Fleischer, RL, PB Price and RM Walker. 1965. "Solid-state track detectors: applications to nuclear science and geophysics." Annual Review of Nuclear Science 15 (1): 1-28.

Fredrickson, James K, John M Zachara, David W Kennedy, Martine C Duff, Yuri A Gorby, W Li Shu-mei and Kenneth M Krupka. 2000. "Reduction of U (VI) in goethite (α-FeOOH) suspensions by a dissimilatory metal-reducing bacterium." Geochimica et cosmochimica acta 64 (18): 3085-3098.

Froidevaux, Pascal, Tony Dell and Paul Tossell. 2006. "Radionuclides in foodstuffs and food raw materials." In Radionuclide concentrations in food and the environment, edited by Michael Poschl and Leo M.L. Nollet. London: Taylor & Francis Group.

Fujimoto, K, S Kobayashi and H Yonehara. 1994. "Spatial distribution of thoron and radon concentrations in the indoor air of a traditional Japanese wooden house." Health Physics 66 (1): 43-49.

Gäggeler, HW, DT Jost, U Baltensperger, M Schwikowski and P Seibert. 1995. "Radon and thoron decay product and 210 Pb measurements at Jungfraujoch, Switzerland." Atmospheric Environment 29 (5): 607-616.

Gao, Y, W Baeyens, S De Galan, A Poffijn and M Leermakers. 2010. "Mobility of radium and trace metals in sediments of the Winterbeek: application of sequential extraction and DGT techniques." Environmental Pollution 158 (7): 2439-2445.

Garcia-Solsona, E, J Garcia-Orellana, P Masqué and Henrieta Dulaiova. 2008. "Uncertainties associated with 223Ra and 224Ra measurements in water via a Delayed Coincidence Counter (RaDeCC)." Marine Chemistry 109 (3): 198-219.

Garcıa-Talavera, M, JP Laedermann, M Decombaz, MJ Daza and B Quintana. 2001. "Coincidence summing corrections for the natural decay series in γ-ray spectrometry." Applied Radiation and Isotopes 54 (5): 769-776.

Gascoyne, M. 1992. "Geochemistry of the actinides and their daughters." In Uranium-series disequilibrium: applications to earth, marine, and environmental sciences. 2. ed, edited by M Ivanovich and R.S Harmon. Oxford: Clarendon Press.

Gauthier, P-J, M-F Le Cloarec and M Condomines. 2000. "Degassing processes at Stromboli volcano inferred from short-lived disequilibria (210 Pb–210 Bi–210 Po) in volcanic gases." Journal of volcanology and geothermal research 102 (1): 1-19.

Gehrke, RJ and JR Davidson. 2005. "Acquisition of quality γ-ray spectra with HPGe spectrometers." Applied Radiation and Isotopes 62 (3): 479-499.

Gesell, Thomas F and Howard M Prichard. 1975. "The technologically enhanced natural radiation environment." Health Physics 28 (4): 361-366.

Gilmore, Gordon. 2008. Practical gamma-ray spectroscopy. 2nd ed. Chichester: Wiley Godoy, J.M. and M.L. Godoy. 2006. "Natural radioactivity in Brazilian groundwater." Journal of

Environmental Radioactivity 85 (1): 71-83. Gourdin, Elian, O Evrard, S Huon, J-L Reyss, Olivier Ribolzi, T Bariac, O Sengtaheuanghoung and

S Ayrault. 2014. "Spatial and temporal variability of 7 Be and 210 Pb wet deposition

37

during four successive monsoon storms in a catchment of northern Laos." Journal of Environmental Radioactivity 136: 195-205.

Grenthe, Ingmar, Hans Wanner and Isabelle Forest. 1992. "Chemical thermodynamics of uranium.

Gründel, M. and J. Porstendörfer. 2004. "Differences between the activity size distributions of the different natural radionuclide aerosols in outdoor air." Atmospheric Environment 38 (22): 3723-3728.

Grundl, Tim and Mike Cape. 2006. "Geochemical factors controlling radium activity in a sandstone aquifer." Groundwater 44 (4): 518-527.

Guerra, JG, JG Rubiano, G Winter, AG Guerra, H Alonso, MA Arnedo, A Tejera, JM Gil, R Rodríguez and P Martel. 2015. "A simple methodology for characterization of germanium coaxial detectors by using Monte Carlo simulation and evolutionary algorithms." Journal of Environmental Radioactivity 149: 8-18.

Habib, AS, AL Shutt, PH Regan, MC Matthews, H Alsulaiti and DA Bradley. 2014. "Characterization of naturally occurring radioactive materials in Libyan oil pipe scale using a germanium detector and Monte Carlo simulation." Radiation Physics and Chemistry 95: 352-355.

Halstead, Michael JR, Robert G Cunninghame and Keith A Hunter. 2000. "Wet deposition of trace metals to a remote site in Fiordland, New Zealand." Atmospheric Environment 34 (4): 665-676.

Han, J.H. and J.H. Choi. 2010. "Broad Energy HPGe Gamma Spectrometry for Dose Rate Estimation for Trapped Charge Dating." Journal of Analytical Science & Technology 1 (2): 98-108.

Hansen, Richard O and Gordon l Huntington. 1969. "Thorium movements in morainal soils of the high sierra, california." Soil Science 108 (4): 257-265.

Haridasan, PP. 2015. "Managing Exposure to Natural Sources - International Standards and New Challenges." In Proceedings of the seventh International Symposium on Naturally Occurring Radioactive Material (NORM VII), Beijing, China, 22–26 April 2013, edited. Vienna: International Atomic Energy Agency.

Haridasan, PP, M Harikumar, PM Ravi and RM Tripathi. 2015. "Radiological protection against exposure to naturally occurring radioactive material." Radiation Protection and Environment 38 (3): 59.

Haridasan, PP, CG Maniyan, PMB Pillai and AH Khan. 2002. "Dissolution characteristics of 226 Ra from phosphogypsum." Journal of Environmental Radioactivity 62 (3): 287-294.

Hedemann-Jensen, Per. 2010. "Naturally occurring radiation sources: existing or planned exposure situation?" Journal of Radiological Protection 30 (4): 781.

Heijnis, Henk, Julie Ruddock and Peter Coxon. 1993. "A uranium‐thorium dated late Eemian or early midlandian organic deposit from near Kilfenora between Spa and Fenit, Co. Kerry, Ireland." Journal of Quaternary Science 8 (1): 31-43.

Helbig, W and S Niese. 1986. "High sensitivity activation analysis using an underground laboratory for gamma spectrometry." Nuclear Instruments and Methods in Physics Research Section B: Beam Interactions with Materials and Atoms 17 (5-6): 431-434.

Hendry, Jolyon H, Steven L Simon, Andrzej Wojcik, Mehdi Sohrabi, Werner Burkart, Elisabeth Cardis, Dominique Laurier, Margot Tirmarche and Isamu Hayata. 2009. "Human exposure to high natural background radiation: what can it teach us about radiation risks?" Journal of Radiological Protection 29 (2A): A29.

Herranz, M, R Idoeta and F Legarda. 2008. "Evaluation of uncertainty and detection limits in radioactivity measurements." Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment 595 (2): 526-534.

Heydorn, K. 2004. "Evaluation of the uncertainty of environmental measurements of radioactivity." Journal of Radioanalytical and Nuclear Chemistry 262 (1): 249-253.

38

Hilal, MA, MF Attallah, Gehan Y Mohamed and M Fayez-Hassan. 2014. "Evaluation of radiation hazard potential of TENORM waste from oil and natural gas production." Journal of Environmental Radioactivity 136: 121-126.

Himstedt, F. 1904. "Über die radioaktive Emanation der Wasser‐und Ölquellen." Annalen der Physik 318 (3): 573-582.

Hult, Mikael, Werner Preusse, Joël Gasparro and M Kohler. 2006. "Underground gamma-ray spectrometry." Acta Chimica Slovenica 53 (1): 1.

Huy, Ngo Quang, Vo Xuan An, Truong Thi Hong Loan and Nguyen Thanh Can. 2013. "Self-absorption correction in determining the 238 U activity of soil samples via 63.3 keV gamma ray using MCNP5 code." Applied Radiation and Isotopes 71 (1): 11-20.

Huy, NQ and TV Luyen. 2004. "A method to determine 238U activity in environmental soil samples by using 63.3-keV-photopeak-gamma HPGe spectrometer." Applied Radiation and Isotopes 61 (6): 1419-1424.

IAEA. 2003a. Guidelines for radioelement mapping using gamma ray spectrometry data. Vienna: International Atomic Energy Agency.

IAEA. 2003b. Radiation protection against radon in workplaces other than mines, Safety Reports Series 33 Vienna: International Atomic Energy Agency.

IAEA. 2003c. Radiation protection and the management of radioactive waste in the oil and gas industry, Safety reports series; no. 34. Vienna: International Atomic Energy Agency.

IAEA. 2004a. Application of the concepts of exclusion, exemption and clearance, Safety Standards Series No. RS-G-1.7. Vienna: Internat. Atomic Energy Agency.

IAEA. 2004b. "Quantifying uncertainty in nuclear analytical measurements". Vienna: International Atomic Energy Agency.

IAEA. 2005. Derivation of activity concentration values for exclusion, exemption and clearance. Vienna: International Atomic Energy Agency.

IAEA. 2006. Assessing the need for radiation protection measures in work involving minerals and raw materials, Safety Reports Series No.49 Vienna: International Atomic Energy Agency.

IAEA. 2007a. Radiation protection and NORM residue management in the zircon and zirconia industries, Safety Reports Series, No. 51. Vienna.

IAEA. 2007b. Safety glossary : terminology used in nuclear safety and radiation protection. 2007 ed. Vienna: International Atomic Energy Agency.

IAEA. 2010. Handbook of Parameter Values for the Prediction of Radionuclide Transfer in Terrestrial and Freshwater Environments, Technical reports series, no. 472. Vienna.

IAEA. 2011. Radiation protection and NORM residue management in the production of rare earths from thorium containing minerals, Safety Reports Series, No. 68. Vienna.

IAEA. 2012. Radiation protection and NORM residue management in the titanium dioxide and related industries, Safety Reports Series, No. 76. Vienna: International Atomic Energy Agency.

IAEA. 2013a. Management of NORM residues, IAEA TECDOC No. 1712. Vienna. IAEA. 2013b. Measurement and calculation of radon releases from NORM residues, Technical

reports series No. 474. Vienna. IAEA. 2013c. National and regional surveys of radon concentration in dwellings. Vienna:

International Atomic Energy Agency. IAEA. 2013d. Radiation protection and management of NORM residues in the phosphate industry.

Vienna: International Atomic Energy Agency. IAEA. 2014. Radiation protection and safety of radiation sources : international basic safety

standards. Vienna: International Atomic Energy Agency. IAEA. 2015a. Protection of the public against exposure indoors due to radon and other natural

sources of radiation specific safety guide. Vienna: International Atomic Energy Agency. IAEA. 2015b. Safety glossary : terminology used in nuclear safety and radiation protection.

Vienna: International Atomic Energy Agency. http://www-ns.iaea.org/downloads/standards/glossary/iaea-safety-glossary-draft-2016.pdf (Accessed on 05 November 2015).

39

ICRP. 1993. "Protection Against Radon-222 at Home and at Work." ICRP Publication 65. Ann. ICRP 23 (2).

ICRP. 2007a. "The 2007 recommendations of the International Commission on Radiological Protection." ICRP Publication 103. Ann 37 (2-4).

ICRP. 2007b. "Scope of Radiological Protection Control Measures." ICRP Publication 104. Ann 35 (5).

ICRP. 2010. "Lung cancer risk from radon and progeny and statement on radon. ICRP Publication 115." Annals of the ICRP 40 (1).

ICRP. 2014. "Radiological Protection against Radon Exposure. ICRP Publication 126." Ann. ICRP 43 (3) (3).

Ielsch, G, C Ferry, G Tymen and MC Robé. 2002. "Study of a predictive methodology for quantification and mapping of the radon-222 exhalation rate." Journal of Environmental Radioactivity 63 (1): 15-33.

Iimoto, T., Y. Akasaka, Y. Koike and T. Kosako. 2008. "Development of a technique for the measurement of the radon exhalation rate using an activated charcoal collector." Journal of Environmental Radioactivity 99 (4): 587-595.

Inoue, Kazumasa, Masahiro Hosoda, Shinji Tokonami, Tetsuo Ishikawa and Masahiro Fukushi. 2013. "Investigation of radon and thoron concentrations in a landmark skyscraper in Tokyo." Journal of Radioanalytical and Nuclear Chemistry 298 (3): 2009-2015.

ISO. 1995. Guide to the Expression of Uncertainty in Measurement. Geneva: International Organization for Standardization.

ISO. 2012. Measurement of Radioactivity in the Environment — Air: Radon-222 — Part 1: Origin of Radon and its Short-lived Decay Products and Associated Measurement Methods,, ISO Standard 11665-1:2012. Geneva.

Jacob, Daniel J and Michael J Prather. 1990. "Radon‐222 as a test of convective transport in a general circulation model." Tellus B 42 (1): 118-134.

Ji, Young-Yong, Kun Ho Chung, Jong-Myoung Lim, Chang-Jong Kim, Mee Jang, Mun Ja Kang and Sang Tae Park. 2015. "Analytical evaluation of natural radionuclides and their radioactive equilibrium in raw materials and by-products." Applied Radiation and Isotopes 97: 1-7.

Jia, G. and G. Torri. 2007. "Determination of 210Pb and 210Po in soil or rock samples containing refractory matrices." Applied Radiation and Isotopes 65 (1): 1-8.

Jibiri, NN and CM Amakom. 2011. "Radiological Assessment of radionuclide contents in soil waste streams from an oil production well of a Petroleum Development Company in Warri, Niger Delta, Nigeria." Indoor and Built Environment 20 (2): 246-252.

Johnston, A and P Martin. 1997. "Rapid analysis of 226 Ra in waters by γ-ray spectrometry." Applied Radiation and Isotopes 48 (5): 631-638.

Jonassen, N. 1983. "The determination of radon exhalation rates." Health Physics 45 (2): 369-376.

Kabala, Cezary and Bal Ram Singh. 2001. "Fractionation and mobility of copper, lead, and zinc in soil profiles in the vicinity of a copper smelter." Journal of Environmental Quality 30 (2): 485-492.

Karunakara, N, DN Avadhani, HM Mahesh, HM Somashekarappa, Y Narayana and K Siddappa. 2000. "Distribution and enrichment of 210Po in the environment of Kaiga in South India." Journal of Environmental Radioactivity 51 (3): 349-362.

Kaste, James M, Benjamin C Bostick and Arjun M Heimsath. 2006. "Determining 234 Th and 238 U in rocks, soils, and sediments via the doublet gamma at 92.5 keV." Analyst 131 (6): 757-763.

Katz, Joseph J, Glenn T Seaborg and Lester R Morss. 1987. "Chemistry of the actinide elements. Kaye, JH, FP Brauer, RE Connally and HG Rieck. 1972. "Background reductions obtained with

gamma detectors by use of massive cosmic-ray shielding." Nuclear Instruments and Methods 100 (2): 333-348.

Kitto, ME, CO Kunz, CA McNulty, S Covert and M Kuhland. 1998. "Radon measurements and mitigation at a fish hatchery." Health Physics 74 (4): 451-455.

40

Knoll, Glenn F. 2010. Radiation detection and measurement. 4th ed: John Wiley. Köhler, M, S Niese, B Gleisberg, U Jenk and K Nindel. 2000. "Simultaneous determination of Ra

and Th nuclides, 238 U and 227 Ac in uranium mining waters by γ-ray spectrometry." Applied Radiation and Isotopes 52 (3): 717-723.

Kolb, WA and M Wojcik. 1985. "Enhanced radioactivity due to natural oil and gas production and related radiological problems." Science of the Total Environment 45: 77-84.

Kotrappa, P, JC Dempsey, RW Ramsey and LR Stieff. 1990. "A Practical E-PERMTM (Electret Passive Environmental Radon Monitor) System for Indoor 222Rn Measurement." Health Physics 58 (4): 461-467.

Kotrappa, Payasada and Frederick Stieff. 2009. "Radon exhalation rates from building materials using electret ion chamber radon monitors in accumulators." Health Physics 97 (2): 163-166.

Kovler, K., A. Perevalov, V. Steiner and E. Rabkin. 2004. "Determination of the radon diffusion length in building materials using electrets and activated carbon." Health Physics 86 (5): 505-516.

Kramer, Gary H, Barry Hauck and Steve A Allen. 2009. "Recalibration of the cameco mobile lung counter." Health Physics 96 (6): 675-681.

Kramer, Gary H and Barry M Hauck. 2005. "Evaluation of the new graded Z liner in the Human Monitoring Laboratory's lung counter." Radiation Protection Dosimetry 113 (2): 211-213.

Krewski, Daniel, Jay H Lubin, Jan M Zielinski, Michael Alavanja, Vanessa S Catalan, R William Field, Judith B Klotz, Ernest G Létourneau, Charles F Lynch and Joseph L Lyon. 2006. "A combined analysis of North American case-control studies of residential radon and lung cancer." Journal of Toxicology and Environmental Health, Part A 69 (7-8): 533-597.

Krishnaswami, S and J Kirk Cochran. 2011. U-Th series nuclides in aquatic systems. Vol. 13: Elsevier.

Kumar, Rajesh, AK Mahur, D Sengupta and Rajendra Prasad. 2005. "Radon activity and exhalation rates measurements in fly ash from a thermal power plant." Radiation Measurements 40 (2): 638-641.

Kurnaz, Aslı, Belgin Küçükömeroğlu, Recep Keser, NT Okumusoglu, Filiz Korkmaz, Gürsel Karahan and Uğur Çevik. 2007. "Determination of radioactivity levels and hazards of soil and sediment samples in Fırtına Valley (Rize, Turkey)." Applied Radiation and Isotopes 65 (11): 1281-1289.

L'Annunziata, Michael F. 2012. Handbook of radioactivity analysis: Academic Press. Labidi, S, H Mahjoubi, F Essafi and R Ben Salah. 2010. "Natural radioactivity levels in mineral,

therapeutic and spring waters in Tunisia." Radiation Physics and Chemistry 79 (12): 1196-1202.

Lamborg, Carl H, William F Fitzgerald, William C Graustein and Karl K Turekian. 2000. "An examination of the atmospheric chemistry of mercury using 210Pb and 7Be." Journal of Atmospheric Chemistry 36 (3): 325-338.

Landa, Edward R. 2007. "Naturally occurring radionuclides from industrial sources: characteristics and fate in the environment." In Radioactivity in the terrestrial environment, edited by G. Shaw: Elsevier Science.

Landa, Edward R and David F Reid. 1983. "Sorption of radium-226 from oil-production brine by sediments and soils." Environmental geology 5 (1): 1-8.

Landsberger, S, C Brabec, B Canion, J Hashem, C Lu, D Millsap and G George. 2013. "Determination of 226Ra, 228Ra and 210Pb in NORM products from oil and gas exploration: Problems in activity underestimation due to the presence of metals and self-absorption of photons." Journal of Environmental Radioactivity 125: 23-26.

Langmuir, Donald and Janet S Herman. 1980. "The mobility of thorium in natural waters at low temperatures." Geochimica et cosmochimica acta 44 (11): 1753-1766.

Langmuir, Donald and Arthur C Riese. 1985. "The thermodynamic properties of radium." Geochimica et cosmochimica acta 49 (7): 1593-1601.

41

Laubenstein, M, M Hult, J Gasparro, D Arnold, S Neumaier, G Heusser, M Köhler, P Povinec, J-L Reyss and M Schwaiger. 2004. "Underground measurements of radioactivity." Applied Radiation and Isotopes 61 (2): 167-172.

Lavi, N, F Groppi and ZB Alfassi. 2004. "On the measurement of 40K in natural and synthetic materials by the method of high-resolution gamma-ray spectrometry." Radiation Measurements 38 (2): 139-143.

Lawrence, C.E., R.A. Akber, A. Bollhöfer and P. Martin. 2009. "Radon-222 exhalation from open ground on and around a uranium mine in the wet-dry tropics." Journal of Environmental Radioactivity 100 (1): 1-8.

Lee, Thomas KC and KN Yu. 2000. "Effects of air conditioning, dehumidification and natural ventilation on indoor concentrations of 222Rn and 220Rn." Journal of Environmental Radioactivity 47 (2): 189-199.

Leelőssy, Ádám, Ferenc Molnár, Ferenc Izsák, Ágnes Havasi, István Lagzi and Róbert Mészáros. 2014. "Dispersion modeling of air pollutants in the atmosphere: a review." Open Geosciences 6 (3): 257-278.

Legeleux, Françoise and Jean-Louis Reyss. 1996. "228Ra226Ra activity ratio in oceanic settling particles: implications regarding the use of barium as a proxy for paleoproductivity reconstruction." Deep Sea Research Part I: Oceanographic Research Papers 43 (11): 1857-1863.

Lehto, Jukka and Xiaolin Hou. 2011. Chemistry and analysis of radionuclides: laboratory techniques and methodology. Weinheim: Wiley-VCH.

Lenka, P, SK Jha, S Gothankar, RM Tripathi and VD Puranik. 2009. "Suitable gamma energy for gamma-spectrometric determination of 238U in surface soil samples of a high rainfall area in India." Journal of Environmental Radioactivity 100 (6): 509-514.

Li, Qi, Shilian Wang, Yungang Zhao, Shujiang Liu, Yuanqing Fan, Jianfang Shi, Huaimao Jia and Weixiang Yu. 2015. "Determination of 226Ra Activity Using Gamma Spectrometry with 226Ra–222Rn Disequilibrium." Health Physics 109 (2): 113-116.

Lin, Chun-Chih, Tieh-Chi Chu and Yu-Fen Huang. 2003. "Variations of U/Th-series nuclides with associated chemical factors in the hot spring area of northern Taiwan." Journal of Radioanalytical and Nuclear Chemistry 258 (2): 281-286.

Lin, Zhichao, Stephanie Healey and Zhongyu Wu. 2015. "Rapid and simultaneous detection of alpha/beta radioactivity in food by solid phase extraction liquid scintillation counting." Journal of Radioanalytical and Nuclear Chemistry: 1-8.

Liye, Liu, Didier Franck, Loic De Carlan and Li Junli. 2007. "Application of Monte Carlo calculation and OEDIPE software for virtual calibration of an in vivo counting system." Radiation Protection Dosimetry 127 (1-4): 282-286.

Loaiza, P, C Chassaing, Ph Hubert, A Nachab, F Perrot, J-L Reyss and G Warot. 2011. "Low background germanium planar detector for gamma-ray spectrometry." Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment 634 (1): 64-70.

Lombardo, Andrew J and Anita F Mucha. 2008. "Measuring and Modeling Naturally Occurring Radioactive Material: Interpreting the Relationship Between the Natural Radionuclides Present-8516.

Lourtau, AM Carrasco, MP Rubio Montero and M Jurado Vargas. 2014. "Characterization of coal and charcoal by alpha-particle and gamma-ray spectrometry." Radiation Physics and Chemistry.

Macášek, M. 2000. "Isotope dilution and sampling factors of the quality assurance and TQM of environmental analysis." Journal of Radioanalytical and Nuclear Chemistry 246 (3): 709-718.

Mahat, RH, DA Bradley, YM Amin, CY Wong and LD Su. 2001. "The effect of humidity on the accuracy of measurement of an electret radon dosimeter." Radiation Physics and Chemistry 61 (3): 489-490.

42

Mahur, AK, Rajesh Kumar, Meena Mishra, D Sengupta and Rajendra Prasad. 2008. "An investigation of radon exhalation rate and estimation of radiation doses in coal and fly ash samples." Applied Radiation and Isotopes 66 (3): 401-406.

Marley, Frederick and Paul S Phillips. 2001. "Investigation of the potential for radon mitigation by operation of mechanical systems affecting indoor air." Journal of Environmental Radioactivity 54 (2): 205-219.

Martin, AJ, J Crusius, J Jay McNee and EK Yanful. 2003. "The mobility of radium-226 and trace metals in pre-oxidized subaqueous uranium mill tailings." Applied Geochemistry 18 (7): 1095-1110.

Martin, Alan, Sam Harbison, Karen Beach and Peter Cole. 2012. An Introduction to Radiation Protection 6th ed. New York: CRC Press.

Martin, P, GJ Hancock, S Paulka and RA Akber. 1995. "Determination of 227 Ac by α-particle spectrometry." Applied Radiation and Isotopes 46 (10): 1065-1070.

Martin, P, S Tims, B Ryan and A Bollhöfer. 2004. "A radon and meteorological measurement network for the Alligator Rivers Region, Australia." Journal of Environmental Radioactivity 76 (1): 35-49.

Martin, P. and R.A. Akber. 1999. "Radium isotopes as indicators of adsorption–desorption interactions and barite formation in groundwater." Journal of Environmental Radioactivity 46 (3): 271-286.

Martin, Paul and Gary Hancock. 1992. Routine analysis of naturally occurring radionuclides in environmental samples by alpha-particle spectrometry: Supervising scientist.

Martin, Paul and John L McBride. 2012. "Radionuclide Behaviour and Transport in the Tropical Atmospheric Environment." In Tropical Radioecology, edited by J.R. Twining, 59: Elsevier Science.

Mattinson, James M. 2010. "Analysis of the relative decay constants of 235 U and 238 U by multi-step CA-TIMS measurements of closed-system natural zircon samples." Chemical Geology 275 (3): 186-198.

Mauring, Alexander, Torbjörn Gäfvert and Thomas Bandur Aleksandersen. 2014. "Implications for analysis of 226 Ra in a low-level gamma spectrometry laboratory due to variations in radon background levels." Applied Radiation and Isotopes 94: 54-59.

McLaughlin, J. 2010. "An overview of thoron and its progeny in the indoor environment." Radiation Protection Dosimetry 141 (4): 316-321.

Medley, Peter, Andreas Bollhöfer, David Parry and Paul Martin. 2013. "Radium concentration factors in passionfruit (Passiflora foetida) from the Alligator Rivers Region, Northern Territory, Australia." Journal of Environmental Radioactivity 126: 137-146.

Meisenberg, O. and J. Tschiersch. 2011. "Thoron in indoor air: modeling for a better exposure estimate." Indoor Air 21 (3): 240-252.

Michalik, B, S Pepin and N Tsurikov. 2011. "European waste catalogue-A platform for a common approach to NORM and other industrial waste." In Proceedings of the sixth International Symposium on naturally occurring radioactive material, edited. Vienna: International Atomic Energy Agency.

Michalik, Bogusław. 2009. "Is it necessary to raise awareness about technologically enhanced naturally occurring radioactive materials?" Journal of Environmental Monitoring 11 (10): 1825-1833.

Michalik, Bogusław, J Brown and P Krajewski. 2013. "The fate and behaviour of enhanced natural radioactivity with respect to environmental protection." Environmental Impact Assessment Review 38: 163-171.

Michielsen, N and G Tymen. 2007. "Semi-continuous measurement of the unattached radon decay products size distributions from 0.5 to 5nm by an array of annular diffusion channels." Journal of Aerosol Science 38 (11): 1129-1139.

Miller, Kevin M, Peter Shebell and Gladys A Klemic. 1994. "In situ gamma-ray spectrometry for the measurement of uranium in surface soils." Health Physics 67 (2): 140-150.

43

Mohanty, AK, D Sengupta, SK Das, SK Saha and KV Van. 2004. "Natural radioactivity and radiation exposure in the high background area at Chhatrapur beach placer deposit of Orissa, India." Journal of Environmental Radioactivity 75 (1): 15-33.

Moore, HE, SE Poet and EA Martell. 1973. "222Rn, 210Pb, 210Bi, and 210Po profiles and aerosol residence times versus altitude." Journal of geophysical research 78 (30): 7065-7075.

Moreno, V, C Baixeras, Ll Font and J Bach. 2008. "Indoor radon levels and their dynamics in relation with the geological characteristics of La Garrotxa, Spain." Radiation Measurements 43 (9): 1532-1540.

Mrđa, D, I Bikit, M Vesković and S Forkapić. 2007. "Contribution of 210 Pb bremsstrahlung to the background of lead shielded gamma spectrometers." Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment 572 (2): 739-744.

Mudbidre, R, M Baskaran and L Schweitzer. 2014. "Investigations of the partitioning and residence times of Po-210 and Pb-210 in a riverine system in Southeast Michigan, USA." Journal of Environmental Radioactivity 138: 375-383.

Murray, AS, R Marten, A Johnston and P Martin. 1987. "Analysis for naturally occuring radionuclides at environmental concentrations by gamma spectrometry." Journal of Radioanalytical and Nuclear Chemistry 115 (2): 263-288.

Murty, VRK, JG King, N. Karunakara and VCC Raju. 2010. "Indoor and outdoor radon levels and its diurnal variations in Botswana." Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment 619 (1-3): 446-448.

Najam, Laith A, Nada F Tawfiq and Rana Hesham Mahmood. 2013. "Radon Concentration in Some Building Materials in Using CR-39 Track Detector." Nature 1 (3): 73-76.

Nazaroff, William W. 1992. "Radon transport from soil to air." Reviews of Geophysics 30 (2): 137-160.

Neck, Volker, R Müller, M Bouby, M Altmaier, J Rothe, Melissa A Denecke and Jae-Il Kim. 2002. "Solubility of amorphous Th (IV) hydroxide–application of LIBD to determine the solubility product and EXAFS for aqueous speciation." Radiochimica Acta 90 (9-11/2002): 485-494.

Neder, H, G Heusser and M Laubenstein. 2000. "Low level γ-ray germanium-spectrometer to measure very low primordial radionuclide concentrations." Applied Radiation and Isotopes 53 (1): 191-195.

Niese, Siegfried. 2014. "Underground laboratories for low-level radioactivity measurements". In Analysis of Environmental Radionuclides, edited by Pavel P Povinec and M Baxter: Elsevier Science.

Nisti, Marcelo Bessa, Marcia Pires de Campos and Barbara Paci Mazzilli. 2014. "Natural radionuclides content and radon exhalation rate from Brazilian phosphogypsum piles." Journal of Radioanalytical and Nuclear Chemistry 299 (1): 261-264.

Nolan, PJ, FA Beck and DB Fossan. 1994. "Large arrays of escape-suppressed gamma-ray detectors." Annual review of nuclear and particle science 44 (1): 561-607.

NORDIC. 2000. Naturally occurring radioactivity in the Nordic countries - recommendations: The Radiation Protection Authorities in Denmark, Finland, Iceland, Norway and Sweden.

Nuccetelli, Cristina and Chiara Bolzan. 2001. "In situ gamma spectroscopy to characterize building materials as radon and thoron sources." Science of the Total Environment 272 (1): 355-360.

O'Brien, RS and MB Cooper. 1998. "Technologically enhanced naturally occurring radioactive material (NORM): pathway analysis and radiological impact." Applied Radiation and Isotopes 49 (3): 227-239.

Oikawa, Shinji, Nobuyuki Kanno, Tetsuya Sanada, Naoyuki Ohashi, Masaki Uesugi, Kaneaki Sato, Joji Abukawa and Hideo Higuchi. 2003. "A nationwide survey of outdoor radon concentration in Japan." Journal of Environmental Radioactivity 65 (2): 203-213.

Ongori, Joash N, Robert Lindsay and Mashinga J Mvelase. 2015. "Radon transfer velocity at the water–air interface." Applied Radiation and Isotopes 105: 144-149.

44

Ortiz, T and P Carboneras. 2011. "Management of Scrap Metal Contaminated with Radionuclides of Natural Origin." In Naturally Occurring Radioactive Material (NORM VI). Proceedings of an International Symposium, edited.

Oufni, L. 2003. "Determination of the radon diffusion coefficient and radon exhalation rate in Moroccan quaternary samples using the SSNTD technique." Journal of Radioanalytical and Nuclear Chemistry 256 (3): 581-586.

Özden, Banu, Aysun Uğur, Tolga Esetlili, Bihter Çolak Esetlili and Yusuf Kurucu. 2013. "Assessment of the effects of physical–chemical parameters on 210 Po and 210 Pb concentrations in cultivated and uncultivated soil from different areas." Geoderma 192: 7-11.

Palomo, M, M Villa, N Casacuberta, A Peñalver, F Borrull and C Aguilar. 2011. "Evaluation of different parameters affecting the liquid scintillation spectrometry measurement of gross alpha and beta index in water samples." Applied Radiation and Isotopes 69 (9): 1274-1281.

Palomo, M., A. Peñalver, C. Aguilar and F. Borrull. 2010. "Presence of Naturally Occurring Radioactive Materials in sludge samples from several Spanish water treatment plants." Journal of hazardous materials 181 (1): 716-721.

Papachristodoulou, CA, DL Patiris and KG Ioannides. 2010. "Exposure to indoor radon and natural gamma radiation in public workplaces in north-western Greece." Radiation Measurements 45 (7): 865-871.

Papadopoulos, A, G Christofides, A Koroneos, S Stoulos and C Papastefanou. 2013. "Radioactive secular equilibrium in 238 U and 232 Th series in granitoids from Greece." Applied Radiation and Isotopes 75: 95-104.

Papastefanou, Constantin. 2009. "Radon decay product aerosols in ambient air." Aerosol Air Qual. Res 9: 385-393.

Paschoa, A.S. and F. Steinhäusler. 2010. Technologically Enhanced Natural Radiation. Vol. 17: An Elsevier Title. Accessed 19 May 2012. http://www.qut.eblib.com.au.ezp01.library.qut.edu.au/patron/FullRecord.aspx?p=569312.

Pasculli, Antonio, Sergio Palermi, Annalina Sarra, Tommaso Piacentini and Enrico Miccadei. 2014. "A modelling methodology for the analysis of radon potential based on environmental geology and geographically weighted regression." Environmental Modelling & Software 54: 165-181.

Patra, A Chakrabarty, S Mohapatra, SK Sahoo, RM Tripathi and VD Puranik. 2013. "Radioactive series disequilibria in underground uranium deposits of the Singhbhum Shear Zone, Eastern India." Journal of Radioanalytical and Nuclear Chemistry 295 (1): 675-683.

Pavlidou, S, A Koroneos, C Papastefanou, G Christofides, S Stoulos and M Vavelides. 2006. "Natural radioactivity of granites used as building materials." Journal of Environmental Radioactivity 89 (1): 48-60.

Persson, Bertil RR and Elis Holm. 2011. "Polonium-210 and lead-210 in the terrestrial environment: a historical review." Journal of Environmental Radioactivity 102 (5): 420-429.

Petropoulos, NP, MJ Anagnostakis and SE Simopoulos. 2001. "Building materials radon exhalation rate: ERRICCA intercomparison exercise results." The Science of the Total Environment 272 (1-3): 109-118.

Pinel, J, T Fearn, SC Darby and JCH Miles. 1995. "Seasonal correction factors for indoor radon measurements in the United Kingdom." Radiation Protection Dosimetry 58 (2): 127-132.

Pontedeiro, EM, PFL Heilbron and Renato M Cotta. 2007. "Assessment of the mineral industry NORM/TENORM disposal in hazardous landfills." Journal of Hazardous Materials 139 (3): 563-568.

Popic, J.M., C.R. Bhatt, B. Salbu and L. Skipperud. 2011. "Outdoor 220Rn, 222Rn and terrestrial gamma radiation levels: investigation study in the thorium rich Fen Complex, Norway." Journal of Environmental Monitoring 14: 193-201.

45

Porcelli, Donald and Peter W Swarzenski. 2003. "The behavior of U-and Th-series nuclides in groundwater." Reviews in Mineralogy and Geochemistry 52 (1): 317-361.

Porstendörfer, J. 1994. "Properties and behaviour of radon and thoron and their decay products in the air." Journal of Aerosol Science 25 (2): 219-263.

Porstendörfer, Justin and August Reineking. 1999. "Radon: Characteristics in air and dose conversion factors." Health Physics 76 (3): 300-305.

Potiriadis, C., V. Koukouliou, S. Seferlis and K. Kehagia. 2011. "Assessment of the occupational exposure at a fertiliser industry in the northern part of Greece." Radiation Protection Dosimetry 144 (1-4): 668.

Povinec, P. 2008. "Low-level gamma-ray spectrometry for environmental samples." Journal of Radioanalytical and Nuclear Chemistry 276 (3): 771-777.

Povinec, Pavel P, Iolanda Osvath and Jean-François Comanducci. 2014. "Underwater gamma-ray spectrometry". In Analysis of Environmental Radionuclides, edited by Pavel P Povinec and M Baxter: Elsevier Science.

Povinec, PP, JF Commanducci and I Levy-Palomo. 2005. "IAEA-MEL's underground counting laboratory (CAVE) for the analysis of radionuclides in the environment at very low-levels." Journal of Radioanalytical and Nuclear Chemistry 263 (2): 441-445.

Prasad, G., T. Ishikawa, M. Hosoda, A. Sorimachi, SK Sahoo, N. Kavasi, S. Tokonami, M. Sugino and S. Uchida. 2012. "Seasonal and diurnal variations of radon/thoron exhalation rate in Kanto-loam area in Japan." Journal of Radioanalytical and Nuclear Chemistry 292 (3): 1385-1390.

Putyrskaya, V, E Klemt, S Röllin, M Astner and H Sahli. 2015. "Dating of sediments from four Swiss prealpine lakes with 210 Pb determined by gamma-spectrometry: progress and problems." Journal of Environmental Radioactivity 145: 78-94.

Quintana, B and C Montes. 2013. "Summing-coincidence corrections with Geant4 in routine measurements by γ spectrometry of environmental samples." Applied Radiation and Isotopes In press.

Raghunath, B. and P. Kotrappa. 1979. "Diffusion coefficients of decay products of radon and thoron." Journal of Aerosol Science 10 (2): 133-138.

Rajaretnam, G. and H.B. Spitz. 2000. "Effect of leachability on environmental risk assessment for naturally occurring radioactive materials in petroleum oil fields." Health Physics 78 (2): 191.

Ramola, RC, GS Gusain, BS Rautela, DV Sagar, G Prasad, SK Shahoo, T Ishikawa, Y Omori, M Janik and A Sorimachi. 2012. "Levels of thoron and progeny in high background radiation area of southeastern coast of Odisha, India." Radiation Protection Dosimetry 152 (1-3): 62-65.

Reddy, Katta Jayaram, L Wang and Steven P Gloss. 1995. Solubility and mobility of copper, zinc and lead in acidic environments: Springer.

Reyss, J-L, S Schmidt, F Legeleux and Ph Bonté. 1995. "Large, low background well-type detectors for measurements of environmental radioactivity." Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment 357 (2): 391-397.

Righi, Serena and Luigi Bruzzi. 2006. "Natural radioactivity and radon exhalation in building materials used in Italian dwellings." Journal of Environmental Radioactivity 88 (2): 158-170.

Ronca-Battista, M. and D. Gray. 1988. "The influence of changing exposure conditions on measurements of radon concentrations with the charcoal adsorption technique." Radiation Protection Dosimetry 24 (1-4): 361-365.

Roos, Per. 2014. "Analysis of radionuclides using ICP-MS". In Analysis of Environmental Radionuclides, edited by Pavel P Povinec and M Baxter: Elsevier Science.

Roussel, Christophe, Catherine Néel and Hubert Bril. 2000. "Minerals controlling arsenic and lead solubility in an abandoned gold mine tailings." Science of the Total Environment 263 (1): 209-219.

Saad, AF, RM Abdallah and NA Hussein. 2013. "Radon exhalation from Libyan soil samples measured with the SSNTD technique." Applied Radiation and Isotopes 72: 163-168.

46

Sahu, SK, M Tiwari, RC Bhangare and GG Pandit. 2014. "Enrichment and particle size dependence of polonium and other naturally occurring radionuclides in coal ash." Journal of Environmental Radioactivity 138: 421-426.

Saïdou, François Bochud, Jean-Pascal Laedermann, MG Kwato Njock and Pascal Froidevaux. 2008. "A comparison of alpha and gamma spectrometry for environmental natural radioactivity surveys." Applied Radiation and Isotopes 66 (2): 215-222.

Sakan, Sanja, Aleksandar Popović, Ivan Anđelković and Dragana Đorđević. 2015. "Aquatic sediments pollution estimate using the metal fractionation, secondary phase enrichment factor calculation, and used statistical methods." Environmental Geochemistry and Health: 1-13.

Samad, MA, MI Ali, D. Paul and SMA Islam. 2012. "An Environmental Impact Study of Jamuna Urea Fertilizer Factory at." Journal of Environmental Science and Natural Resources 4 (2): 27-33.

Samuelsson, C and L Johansson. 1994. "Long-lived radon decay products as a long-term radon exposure indicator." Radiation Protection Dosimetry 56 (1-4): 123-126.

Samuelsson, C. 1990. "The closed-can exhalation method for measuring radon." Journal of Research of the National Institute of Standards and Technology 95 (2): 167-169.

Sanchez-Cabeza, JA and AC Ruiz-Fernández. 2012. "210 Pb sediment radiochronology: an integrated formulation and classification of dating models." Geochimica et cosmochimica acta 82: 183-200.

Sartandel, SJ, SK Jha, SV Bara and RM Tripathi. 2014. "Assessment of 226Ra and 228Ra activity concentration in west coast of India." Journal of Radioanalytical and Nuclear Chemistry 300 (2): 873-877.

Sauvé, Sébastien, Murray McBride and William Hendershot. 1998. "Lead phosphate solubility in water and soil suspensions." Environmental Science & Technology 32 (3): 388-393.

Schery, SD, S Whittlestone, KP Hart and SE Hill. 1989. "The flux of radon and thoron from Australian soils." Journal of Geophysical Research 94 (D6): 8567-8576.

Scholten, Jan C, Iolanda Osvath and Mai Khanh Pham. 2013. "226Ra measurements through gamma spectrometric counting of radon progenies: How significant is the loss of radon?" Marine Chemistry 156: 146-152.

Schubert, Michael, Albrecht Paschke, Eric Lieberman and William C Burnett. 2012. "Air–water partitioning of 222Rn and its dependence on water temperature and salinity." Environmental Science & Technology 46 (7): 3905-3911.

Schumaker, MA and CE Svensson. 2007. "Compton-suppression and add-back techniques for the highly segmented TIGRESS HPGe clover detector array." Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment 575 (3): 421-432.

Sebesta, Ferdinand, Petr Benes, Josef Sedlacek, Jan John and Roman Sandrik. 1981. "Behavior of radium and barium in a system including uranium mine waste waters and adjacent surface waters." Environmental Science & Technology 15 (1): 71-75.

Semkow, TM, PP Parekh, CD Schwenker, AJ Khan, A Bari, JF Colaresi, OK Tench, G David and W Guryn. 2002. "Low-background gamma spectrometry for environmental radioactivity." Applied Radiation and Isotopes 57 (2): 213-223.

Sengupta, Arijit, Renuka H Sankhe and V Natarajan. 2015. "Rapid and non-destructive determination of uranium and thorium by gamma spectrometry and a comparison with ICP-AES." Journal of Radioanalytical and Nuclear Chemistry: 1-6.

Shakhashiro, A and L Mabit. 2009. "Results of an IAEA inter-comparison exercise to assess 137 Cs and total 210 Pb analytical performance in soil." Applied Radiation and Isotopes 67 (1): 139-146.

Shang, B., B. Chen, Y. Gao, Y. Wang, H. Cui and Z. Li. 2005. "Thoron levels in traditional Chinese residential dwellings." Radiation and environmental biophysics 44 (3): 193-199.

Sharma, N and HS Virk. 2001. "Exhalation rate study of radon/thoron in some building materials." Radiation Measurements 34 (1): 467-469.

47

Shawky, S., H.A. Amer, M.I. Hussein, Z. El-Mahdy and M. Mustafa. 2002. "Uranium bioassay and radioactive dust measurements at some uranium processing sites in Egypt—health effects." J. Environ. Monit. 4 (4): 588-591.

Sheets, RW and AE Lawrence. 1999. "Temporal dynamics of airborne lead-210 in Missouri (USA): implications for geochronological methods." Environmental geology 38 (4): 343-348.

Sheppard, Marsha I. 1980. The environmental behaviour of uranium and thorium: Atomic Energy of Canada Ltd., Pinawa, Manitoba. Whiteshell Nuclear Research Establishment.

Sheppard, Steve C, Marsha I Sheppard, Marie-Odile Gallerand and Barb Sanipelli. 2005. "Derivation of ecotoxicity thresholds for uranium." Journal of Environmental Radioactivity 79 (1): 55-83.

Shoesmith, David William. 1984. The behaviour of radium in soil and in uranium mine-tailings: Atomic Energy of Canada Ltd., Pinawa, Manitoba. Whiteshell Nuclear Research Establishment.

Shuktomova, I. I. and N. G. Rachkova. 2011. "Determination of 226Ra and 228Ra in slightly mineralised natural waters." Journal of Environmental Radioactivity 102 (2): 84-87.

Siegel, PB. 2013. "Gamma spectroscopy of environmental samples." American Journal of Physics 81 (5): 381.

Sima, O and D Arnold. 1996. "Self-attenuation and coincidence-summing corrections calculated by Monte Carlo simulations for gamma-spectrometric measurements with well-type germanium detectors." Applied Radiation and Isotopes 47 (9): 889-893.

Singh, Surinder, Rohit Mehra and Kulwant Singh. 2005. "Seasonal variation of indoor radon in dwellings of Malwa region, Punjab." Atmospheric Environment 39 (40): 7761-7767.

Singh, Vishwanath P, DD Veerendra, BN Dileep, Q Sheikh, SS Managanvi, NM Badiger and HR Bhat. 2011. "Background minimisation of HPGe detector by passive graded shielding." International Journal of Low Radiation 8 (4): 313-328.

Smith, AL. 1987. "Radioactive-scale formation." Journal of Petroleum Technology 39 (6): 697-706.

Smith, ML, L Bignell, D Alexiev, L Mo and J Harrison. 2008. "Evaluation of lead shielding for a gamma-spectroscopy system." Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment 589 (2): 275-279.

Sohrabi, M. and AR Esmaili. 2002. "New public dose assessment of elevated natural radiation areas of Ramsar (Iran) for epidemiological studies." In Proceedings of the 5th International Conference on High Levels of Natural Radiation and Radon Areas, Munich, Germany, edited, 15-24: Elsevier.

Somlai, J., V. Jobbágy, J. Kovács, S. Tarján and T. Kovács. 2008. "Radiological aspects of the usability of red mud as building material additive." Journal of hazardous materials 150 (3): 541-545.

Song, Gang, Boyou Zhang, Xinming Wang, Jingping Gong, Daniel Chan, John Bernett and SC Lee. 2005. "Indoor radon levels in selected hot spring hotels in Guangdong, China." Science of the Total Environment 339 (1): 63-70.

Spehr, W. and A. Johnston. 1983. "The measurement of radon exhalation rates using activated charcoal." Radiation Protection in Australia 1 (3): 113-116.

Stewart, Gillian, J Kirk Cochran, Juan-Carlos Miquel, Pere Masqué, J Szlosek, AM Rodriguez y Baena, Scott W Fowler, Beate Gasser and David J Hirschberg. 2007. "Comparing POC export from 234 Th/238 U and 210 Po/210 Pb disequilibria with estimates from sediment traps in the northwest Mediterranean." Deep Sea Research Part I: Oceanographic Research Papers 54 (9): 1549-1570.

Su, Chih‐Chieh and Chih‐An Huh. 2002. "Atmospheric 210> Po anomaly as a precursor of volcano eruptions." Geophysical Research Letters 29 (5): 14-1-14-4.

Sundal, AV, H Henriksen, O Soldal and T Strand. 2004. "The influence of geological factors on indoor radon concentrations in Norway." Science of the Total Environment 328 (1): 41-53.

48

Swarzenski, Peter W, Donald Porcelli, Per S Andersson and Joseph M Smoak. 2003. "The behavior of U-and Th-series nuclides in the estuarine environment." Reviews in Mineralogy and Geochemistry 52 (1): 577-606.

Tateda, Yutaka, Fernando P Carvalho, Scott W Fowler and Juan-Carlos Miquel. 2003. "Fractionation of 210 Po and 210 Pb in coastal waters of the NW Mediterranean continental margin." Continental Shelf Research 23 (3): 295-316.

Thiry, Yves and May Van Hees. 2008. "Evolution of pH, organic matter and 226 radium/calcium partitioning in U-mining debris following revegetation with pine trees." Science of the Total Environment 393 (1): 111-117.

Tokonami, S, J Pan, M Matsumoto, M Furukawa, K Fujimoto, K Fujitaka and R Kurosawa. 1996. "Radon measurements in indoor workplaces." Radiation Protection Dosimetry 67 (2): 143-146.

Tomasek, Ladislav, Agnès Rogel, Margot Tirmarche, Nicolas Mitton and Dominique Laurier. 2008. "Lung cancer in French and Czech uranium miners: radon-associated risk at low exposure rates and modifying effects of time since exposure and age at exposure." Radiation research 169 (2): 125-137.

Totten Sr, Matthew W, Mark A Hanan and Sheri Simpson. 2007. "Natural remediation of marsh soil contaminated by oil-field brine containing elevated radium levels, southern Louisiana." Environmental Geosciences 14 (3): 113-122.

Tufail, M, Sikander M Mirza, Arif Mahmood, AA Qureshi, Yasir Arfat and HA Khan. 2000. "Application of a closed-can technique for measuring radon exhalation from mine samples of Punjab, Pakistan." Journal of Environmental Radioactivity 50 (3): 267-275.

Tufail, M., N. Akhtar and M. Waqas. 2006. "Radioactive rock phosphate: the feed stock of phosphate fertilizers used in Pakistan." Health Physics 90 (4): 361.

Turner, James E. 2007. Atoms, radiation, and radiation protection. 3rd ed: Wiley-VCH. http://onlinelibrary.wiley.com.ezp01.library.qut.edu.au/book/10.1002/9783527616978 (Accessed on 10 November 2015).

Tyler, Andrew N. 2014. "In situ and airborne gamma-ray spectrometry". In Analysis of Environmental Radionuclides, edited by Pavel P Povinec and M Baxter: Elsevier Science.

UNSCEAR. 2000. Sources and effects of ionizing radiation. UNSCEAR 2000 report to the General Assembly with scientific annexes. New York: United Nations.

UNSCEAR. 2006. Sources-to-effects assessment for radon in homes and workplaces: United Nations Scientific Committee on the Effects of Atomic Radiation

UNSCEAR. 2008. Exposures of the public and workers from various sources of radiation, UNSCEAR 2000 report to the General Assembly with scientific annexes New York: United Nations.

Urosevic, V., D. Nikezic and S. Vulovic. 2008. "A theoretical approach to indoor radon and thoron distribution." Journal of Environmental Radioactivity 99 (12): 1829-1833.

USEPA. 1993. Protocol for radon and radon decay product measurements in homes. USEPA. 2008. Technologically Enhanced Naturally Occurring Radioactive Materials From

Uranium Mining, Volume 1: Mining and Reclamation Background, and Volume 2: Investigation of Potential Health, Geographic, and Environmental Issues of Abandoned Uranium Mines. Washington, DC.

Usman, Shoaib, Henry Spitz and Shan Lee. 1999. "Analysis of electret ion chamber radon detector response to 222Rn and interference from background gamma radiation." Health Physics 76 (1): 44-49.

Van Beek, P, M Bourquin, J-L Reyss, M Souhaut, MA Charette and C Jeandel. 2008. "Radium isotopes to investigate the water mass pathways on the Kerguelen Plateau (Southern Ocean)." Deep Sea Research Part II: Topical Studies in Oceanography 55 (5): 622-637.

Van Beek, P, M Souhaut and J-L Reyss. 2010. "Measuring the radium quartet (228 Ra, 226 Ra, 224 Ra, 223 Ra) in seawater samples using gamma spectrometry." Journal of Environmental Radioactivity 101 (7): 521-529.

Vargas, A. and X. Ortega. 2007. "Influence of environmental changes on integrating radon detectors: results of an intercomparison exercise." Radiation Protection Dosimetry 123 (4): 529-536.

49

Vaupotič, J, M Bezek, Norbert Kavasi, Tetsuo Ishikawa, Hidenori Yonehara and Shinji Tokonami. 2012. "Radon and thoron doses in kindergartens and elementary schools." Radiation Protection Dosimetry 152 (1-3): 247-252.

Vearrier, David, John A Curtis and Michael I Greenberg. 2009. "Technologically enhanced naturally occurring radioactive materials." Clinical toxicology 47 (5): 393-406.

Veguería, S.F.J., J.M. Godoy and N. Miekeley. 2002. "Environmental impact in sediments and seawater due to discharges of Ba, 226 Ra, 228 Ra, V, Ni and Pb by produced water from the Bacia de Campos oil field offshore platforms." Environmental Forensics 3 (2): 115-123.

Verdeny, Elisabet, Pere Masqué, Jordi Garcia-Orellana, Claudia Hanfland, J Kirk Cochran and Gillian M Stewart. 2009. "POC export from ocean surface waters by means of 234 Th/238 U and 210 Po/210 Pb disequilibria: a review of the use of two radiotracer pairs." Deep Sea Research Part II: Topical Studies in Oceanography 56 (18): 1502-1518.

Vesterbacka, P, S Klemola, K Salahel-Din and M Saman. 2009. "Comparison of analytical methods used to determine 235U, 238U and 210Pb from sediment samples by alpha, beta and gamma spectrometry." Journal of Radioanalytical and Nuclear Chemistry 281 (3): 441-448.

Violante, A, V Cozzolino, L Perelomov, AG Caporale and M Pigna. 2010. "Mobility and bioavailability of heavy metals and metalloids in soil environments." Journal of soil science and plant nutrition 10 (3): 268-292.

Wang, Edward X and Gaboury Benoit. 1996. "Mechanisms controlling the mobility of lead in the spodosols of a northern hardwood forest ecosystem." Environmental Science & Technology 30 (7): 2211-2219.

Wang, Fan and Ian C Ward. 2002. "Radon entry, migration and reduction in houses with cellars." Building and Environment 37 (11): 1153-1165.

Wang, J, J Liu, L Zhu, JY Qi, YH Chen, TF Xiao, SM Fu, CL Wang and JW Li. 2012. "Uranium and thorium leached from uranium mill tailing of Guangdong province, China and its implication for radiological risk." Radiation Protection Dosimetry 152 (1-3): 215-219.

Wang, N., L. Xiao, C. Li, W. Mei, Y. Hang and D. Liu. 2009. "Level of radon exhalation rate from soil in some sedimentary and granite areas in China." Journal of Nuclear Science and Technology 46 (3): 303-309.

Webb, Joel L and Gary H Kramer. 2001. "An evaluation of germanium detectors employed for the measurement of radionuclides deposited in lungs using an experimental and Monte Carlo approach." Health Physics 81 (6): 711-719.

Whittlestone, S, W Zahorowski and SD Schery. 1998. "Radon flux variability with season and location in Tasmania, Australia." Journal of Radioanalytical and Nuclear Chemistry 236 (1): 213-217.

WHO. 2009. Handbook on indoor radon. Geneva: World Health Organization. WHO. 2011. Guidelines for Drinking-water Quality. 4th ed. Geneva: World Health Organization. Wilkening, Marvin H and John E Hand. 1960. "Radon flux at the Earth‐air interface." Journal of

Geophysical Research 65 (10): 3367-3370. Wilson, Bary W, Norman H Hansen, Chester L Shepard, Timothy J Peters and Frank L Greitzer.

1999. "Development of a modular in-situ oil analysis prognostic system." In International Society of Logistics (SOLE) 1999 Symposium, Nevada, Las Vegas, edited.

Xhixha, G, GP Bezzon, C Broggini, GP Buso, A Caciolli, I Callegari, S De Bianchi, G Fiorentini, E Guastaldi and M Kaçeli Xhixha. 2013. "The worldwide NORM production and a fully automated gamma-ray spectrometer for their characterization." Journal of Radioanalytical and Nuclear Chemistry 295 (1): 445-457.

Yamada, Y., Q. Sun, S. Tokonami, S. Akiba, W. Zhuo, C. Hou, S. Zhang, T. Ishikawa, M. Furukawa and K. Fukutsu. 2006. "Radon–Thoron discriminative measurements in Gansu Province, China, and their implication for dose estimates." Journal of Toxicology and Environmental Health, Part A 69 (7-8): 723-734.

Yamamoto, Masayoshi, Aya Sakaguchi, Keiichi Sasaki, Katsumi Hirose, Yasuhito Igarashi and Chang Kyu Kim. 2006. "Seasonal and spatial variation of atmospheric 210 Pb and 7 Be

50

deposition: features of the Japan Sea side of Japan." Journal of Environmental Radioactivity 86 (1): 110-131.

Yu, KN, ZJ Guan, T Cheung, TTK Cheung and TY Lo. 2000. "Light weight concrete: 226Ra, 232Th, 40K contents and dose reduction assessment." Applied Radiation and Isotopes 53 (6): 975-980.

Yücel, H, AN Solmaz, E Köse and D Bor. 2009. "Spectral interference corrections for the measurement of 238U in materials rich in thorium by a high resolution γ-ray spectrometry." Applied Radiation and Isotopes 67 (11): 2049-2056.

Yücel, H, AN Solmaz, E Köse and D Bor. 2010. "Methods for spectral interference corrections for direct measurements of 234U and 230Th in materials by gamma-ray spectrometry." Radiation Protection Dosimetry 138 (3): 264-277.

Zahorowski, W, S Whittlestone and JM James. 1998. "Continuous measurements of radon and radon progeny as a basis for management of radon as a hazard in a tourist cave." Journal of Radioanalytical and Nuclear Chemistry 236 (1): 219-225.

Zahorowski, W. and S. Whittlestone. 1996. "A fast portable emanometer for field measurement of radon and thoron flux." Radiation Protection Dosimetry 67 (2): 109-120.

Zapata-Garcia, D, M Llauradó and G Rauret. 2009. "Establishment of a method for the rapid measurement of gross alpha and gross beta activities in sea water." Applied Radiation and Isotopes 67 (5): 978-981.

Zare, Mohammad Reza, Mojtaba Mostajaboddavati, Mahdi Kamali, Mohammad Reza Abdi and Mohammad Seddigh Mortazavi. 2012. "235 U, 238 U, 232 Th, 40 K and 137 Cs activity concentrations in marine sediments along the northern coast of Oman Sea using high-resolution gamma-ray spectrometry." Marine Pollution Bulletin 64 (9): 1956-1961.

Zeevaert, T., L. Sweeck and H. Vanmarcke. 2006. "The radiological impact from airborne routine discharges of a modern coal-fired power plant." Journal of Environmental Radioactivity 85 (1): 1-22.

Zeng, Zhi, Jian Su, Hao Ma, Hengguan Yi, Jianping Cheng, Qian Yue, Junli Li and Hui Zhang. 2014. "Environmental gamma background measurements in China Jinping Underground Laboratory." Journal of Radioanalytical and Nuclear Chemistry 301 (2): 443-450.

Zhang, Tieyuan, Kelvin Gregory, Richard W Hammack and Radisav D Vidic. 2014. "Co-precipitation of radium with barium and strontium sulfate and its impact on the fate of radium during treatment of produced water from unconventional gas extraction." Environmental Science & Technology 48 (8): 4596-4603.

Zhang, W., K. Ungar, J. Chen, N. St-Amant and BL Tracy. 2009. "An accurate method for the determination of 226Ra activity concentrations in soil." Journal of Radioanalytical and Nuclear Chemistry 280 (3): 561-567.

51

Chapter 3 NORM Regulation: Better understanding for

better management

Sami H. Alharbi*, Riaz A. Akber

*School of Physical and Chemical Sciences, Queensland University of Technology, 2 George

Street, Brisbane, Qld 4000, Australia

52

Statement of Contribution of Co-Authors for Thesis by Published Paper The authors listed below have certified that:

1. they meet the criteria for authorship in that they have participated in the conception,

execution, or interpretation, of at least that part of the publication in their field of

expertise;

2. they take public responsibility for their part of the publication, except for the

responsible author who accepts overall responsibility for the publication;

3. there are no other authors of the publication according to these criteria;

4. potential conflicts of interest have been disclosed to (a) granting bodies, (b) the editor

or publisher of journals or other publications, and (c) the head of the responsible

academic unit, and

5. they agree to the use of the publication in the student’s thesis and its publication on the

Australasian Research Online database consistent with any limitations set by publisher

requirements.

In the case of this chapter:

Contributor Statement of contribution

Sami Alharbi

Signature:

Date: 31/03/2016

Identified concept for the research paper,

conducted the review, identified the regulatory issues, wrote the

manuscript

Riaz Akber Assisted with editorial process, Aided in concept development,

Reviewed the manuscript.

Principal Supervisor Confirmation

I have sighted email or other correspondence from all Co-authors confirming their certifying

authorship.

Name:

Signature:

Date:

53

Abstract

Naturally occurring radioactive material (NORM) is present in various aspects of human

lives. Radiation exposure resulting from NORM can be hazardous to humans and the

environment. Such exposure may be increased as a result of certain human activities involving

exploration, processing and production of raw materials. The regulation of NORM has been

given more attention since the 1980s, when the enhanced level of NORM was identified in the

petroleum industry. The presence of NORM was also observed in many other industries

involving minerals and raw materials production, including phosphate, zircon and rare earth.

National and international regulatory bodies attempted to control NORM exposure in some of

these industries by applying the radiation protection system that has been designed and

developed primarily for artificial radionuclides. This has created regulatory issues and

challenges that were identified and discussed in this review on the light of best available

knowledge. In addition, the latest findings on radon have confirmed the health risks not only for

workers, but for dwellings residents as well. The implication of the implementation of the

recent ICRP and IAEA recommendations and policies on indoor radon and thoron protection

were also examined in this review.

3.1 Introduction

Radiation protection is a field that has evolved over the last century. It is a dynamic

concept, with its underlying principles and guidelines being regularly revised, reworked and

amended. The main goal of the radiological protection system is to protect humans and the

environment from the harm of ionising radiation, regardless of its size and origin (ICRP, 2007a).

However, NORM has received less attention than other sources of ionising radiation, although

the resultant exposure from NORM can be more significant (Steinhäusler, 2010). Interestingly,

certain human activities involving exploration and production of raw materials have been found

to cause alterations to levels of NORM (IAEA, 2006a), resulting in a potential increase in

radiation exposure to workers, the public and environment. This source of exposure, in most

cases, has not been introduced deliberately and, more importantly in other cases, it is increasing

as new situations are identified. Human and non-human species need to be protected against

such a source of exposure.

Recently, there has been a growing tendency to protect mankind and the environment

against exposure to NORM caused by human activities. This requires a greater focus on the

management of NORM in different practices and its related aspects. The aim of this chapter is to

review the radiological protection system related to naturally occurring radioactive materials

and highlight critical issues associated with NORM that need to be considered when establishing

54

a framework for regulating NORM. The information presented in this chapter has been compiled

from publications provided by the following organisations: ICRP, IAEA, UNSCEAR, WHO, EU,

ARPANSA, and USEPA. Relevant scientific studies and reviews available in the literature have

also been used.

3.2 Overview of the radiological protection system

3.2.1 A brief history of radiological protection

Some literature reviews have described the historical development of radiation

protection organisations worldwide (Clarke and Valentin, 2005; ICRP, 2009b; Walker, 2000). A

summary of these references is outlined in this Section.

The story of ionising radiation began after the discovery of X-rays and radioactivity in

the 1890s. The use of X-ray devices then became widespread throughout medical and scientific

disciplines. It soon became clear, however, that exposure to this form of radiation caused

numerous adverse effects among patients, physicians and scientists. The same phenomenon

was observed after the discovery of radium. At that time, a lack of knowledge regarding the

hazards presented by radiation led to serious illness and death for those handling radium, such

as watch dial painters (ICRP, 2009b; Walker, 2000). In the early years of the 20th century, the

international community recognised the harm caused by exposure to ionising radiation, as well

as the need to protect people and workers from excessive amounts of radiation. Hence, a

number of national and international organisations have been created for the purpose of

radiation protection.

The First International Commission on Radiology was formed in London in 1925 by a

number of radiologists, physicians and scientists. Their hope was to formulate consistent

radiological metrics and measurements to mitigate and control exposure to ionising radiation.

Three years later, the Second International Commission on Radiology established the

International X-Ray and Radium Protection Committee (IXRPC). At that meeting, the first report

on radiological protection was published. The report did not include any quantitative

recommendations; it focused mainly on shielding needed to protect workers from harmful

exposure (IXRPC, 1928). The IXRPC was later renamed, in 1950, the International Commission

on Radiological Protection (ICRP).

With the growing use of radioactive substances and nuclear energy, there was a need to

establish a committee to collect and evaluate scientific information on the effects and levels of

ionising radiation. This was the main reason for creating the United Nations Scientific

Committee on the Effects of Atomic Radiation (UNSCEAR) in 1955. Two years later, the

International Atomic Energy Agency (IAEA) was established in response to the deep fears and

expectations resulting from the discovery of nuclear energy. These international organisations

55

have worked to protect humans and the environment from the harm of ionising radiation, and

to provide a framework to achieve their goal (ICRP, 2009b).

3.2.2 Quantitative recommendations

In 1934, the radiological protection system entered a new era with the introduction of

numerical recommendations for limiting occupational exposure. The IXRPC recommended a

dose limit of one roentgen per week for occupational exposure to X-rays (IXRPC, 1934). This

corresponded to an annual occupational effective dose of 500 mSv. It was a remarkable

improvement for the radiological protection system. A few years later, the IXRPC revised this

figure, believing the dose was too close to the theoretical threshold for adverse effects to occur.

Thus, the limit was reduced to 0.3 roentgen per week, which corresponded to an annual

occupational effective dose of 150 mSv for both X-rays and gamma radiation (ICRP, 1951). The

report included a table of relative biological effectiveness (RBE) and maximum permissible

body burdens for eleven nuclides, including 226Ra.

The 1954 ICRP report contained further improvements to the quantitative

recommendations. Before that report, there was no such limit set for general members of the

public; the ICRP then recommended a maximum permissible dose for the public as a tenth of the

occupational limit (ICRP, 1955). The concept of the critical organ was introduced in the same

report to relate the dose limit to whole body exposure. The numerical values were reduced in

1958 for both occupational workers and members of the public to 5 rem (50 mSv) and 0.5 rem

(5 mSv), respectively. In this recommendation, the dose limits were replaced by the limits for

the accumulative dose equivalent (in rem), which was derived from the RBE.

With advanced developments in the fields of physics and biology, it was realised that the

degree to which ionising radiation affected biological tissues varied with different types of

radiation and the nature of the tissues and organs being exposed. As a result, the effective dose

limit was introduced in 1991 to take into account the type of radiation (radiation weighting

factor) and sensitivity of different organs (tissue weighting factor) exposed. The annual effective

dose limit was adjusted to 20 mSv for occupational exposure and 1 mSv for the general public.

The occupational and public dose limits have remained unchanged since the ratification of

the ICRP recommendations in 2007. However, based on new information, the ICRP changed the

emphasis on the radiation sensitivity of different organs in the human body (ICRP, 2007a).

These changes included increased tissue weighting factors for the breast (due to the higher

breast cancer risk), decreased tissue weighting factors for the gonads (due to the reduced

hereditary risk) and the introduction of new tissue weighting factors for the brain and salivary

glands (due to the increased cancer risk of these two organs).

The impact of changing the tissue weighting factors was investigated in a study conducted

by Lee et al. (2007), which concluded that the effective dose determined in 1991 was influenced

56

more by the gonads than the breasts, and vice versa in the effective dose of 2007. Lee et al.

attributed their findings to the significant changes in the tissue weighting factor for both the

gonads (from 0.20 to 0.08) and the breasts (from 0.05 to 0.12), as per the ICRP 2007

recommendations.

3.2.3 The conceptual framework

In addition to quantitative recommendations, the principles and concepts of radiological

protection have developed. They have become integral to the radiological protection system

because the numerical advice was insufficient for a comprehensive framework. A number of the

current fundamental principles of radiation protection were implicitly stated in the earlier

recommendations. For example, the terms “maximum permissible exposure” and “probable

threshold for adverse effect” were used in the recommendations of 1951 to imply the principle

of the dose limit. Hence, the main principle of radiation protection was to keep all exposure

below the relevant thresholds to prevent deterministic effects and to minimise stochastic

effects. However, the growing need to balance the costs, benefits and detriments in regards to

radiation exposure led the ICRP to introduce the current principles of radiation protection in

1977: justification, optimisation and limitation (ICRP, 1977).

In the 1990 recommendations, the ICRP introduced a new description for human

activities in which it categorised them into practices and intervention (ICRP, 1991). Practices

increase the overall exposure to radiation due to the introduction of new sources or the

modification of exposure, while intervention decreases the overall exposure by influencing

existing sources of exposure. These concepts were replaced by exposure situations in 2007,

when the ICRP distinguished between three types of exposure situations: planned, existing and

emergency (ICRP, 2007a). The three fundamental principles of radiological protection remained

unchanged.

A special concept was proposed for protecting general members of the public and

known as the critical group. This was a group of individuals who received the highest radiation

dose from a particular source for the purpose of compliance with the regulation (ICRP, 1965).

However, the expression “critical group” was found to be confusing because the adjective

“critical” may be understood as meaning crisis, which was not the intended purpose.

Additionally, the word “group” usually refers to a few tens of people and does not reflect the

individual’s dose being assessed (ICRP, 2007a). Therefore, the “critical group” was replaced by

the concept of the “representative person” (ICRP, 2006). The representative person is defined as

an individual who receives a dose that is representative of more highly exposed individuals in

the population. It is expected that regulatory authorities throughout the world will eventually

phase out the use of the concept of the “critical group” in favour of the “representative person”.

57

3.2.4 Protection beyond humans

The radiological protection system was initially designed to protect humans against

ionising radiation. It was thought that the level of safety required for the protection of human

beings was likely to be sufficient to protect other biota (ICRP, 1977). However, it is difficult to

determine whether the environment is adequately protected from the potential impact of

radiation exposure under different circumstances because there are no criteria that address

environmental protection (ICRP, 2003). For that reason, the ICRP has proposed a mechanism to

assess the radiological impact of practice on living organisms, other than humans, which is yet

to be adopted by a number of regulatory organisations worldwide (ICRP, 2009a). This new

approach is based on the concept of Reference Animals and Plants (RAP). This concept has

defined by ICRP as: “A hypothetical entity, with the assumed basic biological characteristics of a

particular type of animal or plant, as described to the generality of the taxonomic level of family,

with defined anatomical, physiological, and life history properties, that can be used for the

purposes of relating exposure to dose, and dose to effects, for that type of living organism” (ICRP,

2009a).

A set of reference animals and plants has been reported and these were grouped at the

family level (ICRP, 2009a). However, certain types of animals belong to species that may be

associated with more than one type of habitat. For example, some frog species may be counted

more in terrestrial environments, while others are counted more in aquatic environments

(Hosseini et al., 2010). Consequently, the correct type of habitat needs to be associated with

each species.

3.3 The current system of radiological protection

The system of radiological protection has been developed as collaboration between

international organisations as illustrated in Figure 3.1. Basic scientific studies, which are the

primary sources of information, are evaluated and compiled by UNSCEAR. The ICRP in turn uses

this information along with other reports provided by national organisations, such as the

Biological Effects of Ionizing Radiation (BEIR) reports, to devise recommendations. These

recommendations are used by the IAEA to produce its Basic Safety Standards (BSS) that are

adopted by national authorities worldwide.

58

Figure 3.1 The basis of radiological protection regulation.

The current system of radiation protection consists of three exposure situations and

three protection principles to protect four individual categories (IAEA, 2014; ICRP, 2007a). This

system was developed based on a combination of three components: scientific knowledge,

practical experience and social and ethical values (Figure 3.2). Recommendations are

incorporated into regulations and standards, and adopted by regulatory authorities in many

countries worldwide.

Figure 3.2 The radiological protection system.

Basic scientific studies

UNSCEAR

(Reports)

ICRP

(Recommendations)

IAEA

(Standards)

National authorities

(Regulation)

Radiation Protection System

Principles

Justification

Optimisation

Dose limits

Situations

Planned

Emergency

Exsisting

Categories

Occupational

Medical

Public

Environmental

Values

Experience

Science

59

3.3.1 Exposure situations

The ICRP recognised three situations in which individuals could be exposed to radiation:

planned exposure, existing exposure and emergency exposure. These situations replaced the

previous approach of “practices” and “interventions” used by the ICRP. According to the ICRP,

the new situations are capable of covering the entire range of exposures (ICRP, 2007a).

The planned exposure situations involve the deliberate introduction and operation of

sources; they may result in the rise of both normal exposure (i.e. those expected to occur) and

potential exposure (i.e. those that are not expected to occur during the operation) (ICRP,

2007a). The previous operation, categorised as a “practice”, is included in this situation. In

planned exposure situations, the protection against ionising radiation can be planned in

advance of the installation of the source. The magnitude of exposure from a source in these

situations can be predicted and, hence, actions can be taken to control exposure.

Existing exposure situations are those that already exist when a decision regarding

control has to be taken, including prolonged exposure situations after emergencies. For

example, exposure caused by natural sources (e.g. radon in a dwelling or workplace and NORM).

Cosmic radiation is also considered by the ICRP as an existing exposure situation (ICRP, 2016).

Existing exposure can result from other sources that are related to some human activities (ICRP,

1999). The Chernobyl nuclear plant accident in the Ukraine, for instance, may be initially an

emergency exposure situation, but it resulted in the contamination of the surrounding

environment and the establishment of prolonged exposure (Krolak et al., 2012; Weiss, 2011).

Emergency exposure situations are any unexpected events that require urgent action to

avoid or reduce any undesirable consequences. A recent example is the nuclear reactor accident

in Fukushima, Japan, that occurred in 2011 during the operation of a planned situation. This

situation caused a release of various radionuclides into the environment and, consequently, an

existing exposure situation will be established after the accident has passed (Hirose, 2011; Taira

et al., 2012). Emergency and existing exposure situations require appropriate action to reduce

the exposure if it exceeds a certain reference level that are usually set by the regulatory

authority.

There is a likelihood that more than one exposure situation can exist simultaneously. In

this case, it is important to know that each exposure situation has its own merits and, therefore,

each exposure situation has to be treated separately. The exposure situation could start as a

planned exposure and end up as an existing exposure, for example, through an emergency event

which acts as a transition situation. As an example, Chernobyl started as a nuclear power plant

(planned exposure situation), then an accident occurred leading to an emergency exposure

situation and, eventually, environmental contamination is being controlled as an existing

exposure situation (Krolak et al., 2012; Weiss, 2011).

60

3.3.2 Principles

In 1977, the ICRP introduced the fundamental principles of radiological protection:

justification, optimisation and limitation. Justification is realised when the perceived benefits of

a radiation practice outweigh the harm associated with it. Optimisation aims to keep all

exposure as low as reasonably achievable, with economic and social factors being taken into

account. The third principle is the dose limits that are applied to ensure that individual

exposure does not exceed an acceptable level of risk. These principles ensure that the

deterministic effects are avoided and potential stochastic effects are reduced. The principles of

justification and optimisation are applied in all exposure situations and types. The principle of

dose limit is only applied in planned exposure situations, with an exemption for medical

exposure, while the reference level is used in existing and emergency situations.

3.3.3 Exposure categories

Individuals who actually receive, or will receive, exposure are grouped by the ICRP into

three exposure types: occupational exposure, which refers to all exposure incurred by workers

as a result of their work; medical exposure, which refers to any exposure received by patients as

the result of their treatment or diagnostics; and public exposure, which refers to anyone

receiving exposure that is neither occupational nor medical. Another type of exposure is known

as environmental exposure and it refers to any exposure received by any living organisms, other

than humans.

3.4 NORM in the radiation protection context

The presence of natural radionuclides in industries was recognised in the early 20th

century when radium and radon (or radioactive gas at that time) were detected in crude

petroleum (Burton, 1904; Himstedt, 1904) and in natural gas (Satterly and McLennan, 1918).

However, exposure due to such sources, and NORM in general, did not receive attention from

the radiological protection community until the mid-1980s, when elevated levels of NORM were

reported in oil and gas industry (Kolb and Wojcik, 1985; Smith, 1987). The exception was

uranium mining and milling because this was considered to be part of the nuclear industry. This

omission occurred because exposure to NORM was not considered to be a source of risk to

humans. It was framed in the context of natural background radiation, which was assumed to be

uncontrollable and constant everywhere.

However, this perspective was altered when the concentration of NORM was found to

vary considerably from one place to another depending on the geological and geographical

formation. In addition, some human activity resulted in an elevation or modification of NORM

concentration. Fortunately, exposure in a number of these situations can be controlled, such as

in the case of radon in dwellings and workplaces.

61

The current radiation protection system is applied based on the exposure situation. All

exposure to ionising radiation from any source is subject to the radiological protection system,

regardless of its size or origin (ICRP, 2007b). Exposure due to natural sources of ionising

radiation including NORM is generally considered as an existing exposure situation. However,

exposure due to NORM is subjected to the requirements of planned exposure situations as

follows (IAEA, 2014):

i. Exposure due to material in practices where the activity concentration of any

radionuclide in the uranium or thorium decay chains exceeds 1 Bq g-1, or 10 Bq g-1 for

40K in that material

ii. Public exposure delivered by discharges or arising from radioactive waste of the above

practices

iii. Exposure due to 222Rn and 220Rn and their progeny in workplaces in which occupational

exposure, due to other radionuclides in the uranium or thorium decay chains, is

controlled as a planned exposure situation

iv. Exposure due to 222Rn and its progeny in the workplace where the annual average of

222Rn activity concentration remains above the reference level.

Figure 3.3. Diagram of the current regulation of NORM.

Figure 3.3 shows regulation situations applicable to NORM. The regulation of NORM can

be divided into two headings: radon in dwellings and workplaces, and NORM industries. In

these headings the principle of a graded approach to regulation is adopted if the activity

NORM

Regulated

Existing Exposure Situation

222Rn and220Rn in dwellings and workplaces

Practices 238U,232Th ˂ 1Bq/g and 40K˂ 10Bq/g

Planned Exposure Situation

Practices238U,232Th >1Bq/g and 40K>10Bq/g

Radioactive waste arising from the above practices

222Rn and 220Rn associated with regulated practices

222Rn in workplaces > Reference level

Not regulated

Exposures deemed to be not amenable to control (e.g. 40K

in human body)

62

concentration of any radionuclides in the uranium and thorium decay series exceeds 1 Bq g-1

and 10 Bq g-1 of 40K or the relevant reference level for 222Rn (IAEA, 2014). In that case, more

details have to be considered about the practice, including particular types of operation,

process, material involved and exposure dose to workers and the public (IAEA, 2006a). The

graded approach has four levels of regulatory control, which are exemption, notification,

registration and licensing (IAEA, 2014). Exemption is the basic level of the graded approach in

which the regulatory body may decide not to apply regulatory requirements to a material or

practice. The exemption is granted if a specific criterion, proposed by the regulatory body, is

met. If the regulatory body found the exemption was not the optimum option, the next level in

the graded approach would be required, in which the legal person must submit a formal

notification of the intention of operating a facility to the regulatory body. If the regulatory

authority found that exemption or notification were insufficient for the level of exposure,

authorisation would be required in a form of either a registration or a licence. The difference

between the latter two levels is essentially in the level of regulatory stringency. In registration,

only limited obligations on the legal person may be enough to ensure adequate protection, while

in licensing, enforcement of more stringent exposure control measures are more appropriate

for acceptable levels of protection. However, the use of licensing for practices involving

exposure to NORM is likely to be limited to material with very high activity concentrations

(IAEA, 2013a).

3.4.1 Radon in dwellings and workplaces

Radon exposure in dwellings and workplaces has been given more attention than other

exposure from natural sources. This is because exposure due to radon and its decay products is

the major contributor to public exposure from natural sources (UNSCEAR, 2006) and the risks

of exposure to radon and its decay products have been confirmed by several recent

epidemiological studies (Baysson et al., 2004; Darby et al., 2005; ICRP, 2010; Vacquier et al.,

2009). Moreover, exposure to radon in the indoor environment is controllable as its parameters,

such as entry points and ventilation, can be controlled.

A series of recommendations and policies regarding exposure to radon in dwellings and

workplaces have been published recently (IAEA, 2015; ICRP, 2014; UNSCEAR, 2006; WHO,

2009). More discussion on the recent development of radon exposure regulation is provided in

Section 6 of this chapter.

3.4.2 NORM industries

NORM is present at enhanced levels in a wide range of industrial operations, including

exploration, extraction of raw materials, processing, products, by-products and waste. The main

industries in which NORM has been identified are: the oil and gas industry (Abo-Elmagd et al.,

63

2010; Al-Masri and Suman, 2003; Bakr, 2010; White and Rood, 2001), minerals extraction and

processing industries, including uranium and mineral sand (Ballesteros et al., 2008; Haridasan

et al., 2006; Righi et al., 2005), the bauxite and aluminium industry (Abbady and El-Arabi, 2006;

Jobbágy et al., 2009), the phosphate industry (Beddow et al., 2006; Tufail et al., 2006), iron and

steel production (Brown et al., 2003; Iwaoka et al., 2010), coal mining and coal-fired power

generation (Cevik et al., 2007; Parami et al., 2010), the building materials industry (Aslam et al.,

2012; Turhan et al., 2011) and water treatment (Kleinschmidt and Akber, 2008; Palomo et al.,

2010). Enhancements can be either deliberate, as in uranium mining where a higher ore grade

is sought, or inadvertent, as in petroleum industries where the activity is often associated with

the barite phase in produce water, scale and sludge.

Enhanced levels of NORM may have the potential to pose risks to human and the

environment (Vearrier et al., 2009). However, most exposure from NORM industries relates to

workers. The ICRP has devised a broad framework for radiological protection against NORM

industries, without including details of each type of operation (ICRP, 2007a). NORM industries

are simply regulated based on the intention of modifying or introducing the source. If the source

of exposure is modified intentionally in such a way as to increase the radiation dose, the source

falls under the scope of regulation. The specific details are the responsibility of national and

international regulatory bodies. However, in some situations, the national authorities may lack

experience in dealing with NORM.

The IAEA has published a number of standards for managing NORM in a small number

of industries, including the oil and gas industry (IAEA, 2003), minerals and raw materials

production (IAEA, 2006a) and the phosphate industry (IAEA, 2013b). The concept of

exemptions, clearance and exclusion are preferable for most exposure situations in industries

involving NORM (IAEA, 2005).

3.5 Regulatory issues with NORM

In recent years, there has been a growing need to control exposure caused by NORM.

National and international organisations are aware of the importance of managing exposure to

NORM. Some steps have been taken in this direction so far. However, until now, there has been

no complete vision of the regulation of NORM. A broad approach has been used through the

principle of optimisation utilising the concept of exemptions, clearance and exclusion for most

NORM exposure situations (IAEA, 2005).

The controllability of exposure plays an essential role in determining the scope of

regulation (ICRP, 2007a). Controllable exposure includes those situations in which radiological

protection actions are warranted to reduce exposure, such as dwellings and workplaces where

the radon concentration in air exceeds IAEA, ICRP or a national recommended reference level.

64

In other situations, exposure to NORM is uncontrollable, or is not amenable to control, which

means that exposure cannot be controlled by regulations and protective actions are not feasible.

For example, public exposure due to 40K in the human body is considered to be an uncontrolled

situation and therefore, is excluded from radiological protection legislation.

Regulatory authorities may face a number of issues in attempting to establish a regulation

standard for managing exposure to NORM. Nine such issues have been repeatedly found to be

central components of the regulation of NORM. These issues have a direct influence on the

decision regarding managing exposure to NORM. The issues are listed and discussed as follows:

3.5.1 Absence of a complete framework

There has been no comprehensive framework for regulating NORM in non-nuclear

activities. This is reasonable, as NORM regulation was ignored for a long period of time and

most importantly, it has to be treated case-by-case. The characteristic of NORM exposure varies

depending on the type of industrial activity. Studies have shown that the behaviour of individual

radionuclides of NORM can differ from one practice to another and even from field to field for

the same practice (Haridasan et al., 2015; O'Brien and Cooper, 1998; Xhixha et al., 2013). Dose

assessments and exposure pathways change accordingly across practice. For example, in the

petroleum industry, exposure to NORM is caused by radium isotopes, which dissolve in the

produced water and are brought to the surface, precipitated in sludge and scale (Abo-Elmagd et

al., 2010; IAEA, 2003), while in the phosphate industry, they present more in residue,

phosphogypsum, and tend to leach into groundwater (Beddow et al., 2006; Cañete et al., 2008;

Haridasan et al., 2002; IAEA, 2013b). Thus, the criterion for controlling NORM in the oil and gas

industry is not appropriate for the application to the phosphate industry. For this reason, the

regulation applied to NORM in one industry cannot be generalised to others.

Regulatory bodies have recognised this issue and even in the upcoming NORM

framework currently being prepared by the ICRP, the approach is going to be broad enough to

cover a wide range of industrial sectors involved without entering into detail (Lochard, 2015).

Currently, guidance for managing NORM has been issued for a limited number of industries,

including oil and gas, phosphate, rare earths, titanium and zircon and zirconia industries (IAEA,

2003, 2007a, 2011, 2012, 2013b). For other industries where NORM has been identified, (e.g.

geothermal power installation and recycling of metal scraps), guidance is still unavailable.

Meanwhile, general guidelines (e.g. IAEA 2004, 2006a, 2013b) may be applied. The management

of NORM, therefore, should start with identifying industries in which elevated concentrations of

natural radionuclides can be found. It should also characterise exposure, practice and

radionuclide behaviour, as well as resolve waste and residue issues.

65

3.5.2 The interference between planned and existing exposure situations

In the recent ICRP situation-based approach, two situations have been applicable to

exposure from NORM: planned exposure situations and existing exposure situations (ICRP,

2007a). They correspond to the practice and interventions of the previous approach (ICRP,

1991), which are still being used by many national organisations. By definition, exposure to

natural sources is an existing exposure situation. However, if exposure is directly related to the

work, the protection requirements for occupational exposure in planned exposure situations

may be applied. When applying a radiation protection system, each exposure situation is treated

separately, and the radiological protection system is applied according to the exposure

situation.

Sometimes it is difficult to identify whether exposure due to NORM is an existing or

planned exposure situation (Haridasan, 2015; Hedemann-Jensen, 2010). This issue has been

mentioned by the IAEA, which has admitted that in some circumstances the distinction between

exposure situations may be vague (IAEA, 2014). In its recent International Basic Safety Standard

(IBSS), the IAEA (2014) declared that “…some exposures due to natural sources may have some

characteristics of both planned exposure situations and existing exposure situations”. Even in

the previous approach, i.e., practice and intervention, this issue was observed. In waste disposal,

the ICRP found that distinguishing between practices and intervention is not always easy (ICRP

1997).

In situations where the boundary between planned and existing exposure situations is

unclear, it is the national regulatory authority's responsibility to decide which exposure to

NORM should be considered as planned, and which should be considered as an existing

situation (ICRP, 2007b). Most national regulatory authorities around the world, however, may

lack experience and may have not yet adopted the new situations-based approach, with the

practice and intervention-based approach still being used. The regulatory authorities in these

countries may face challenges when transferring their regulation systems to the new approach

and when deciding which exposure situation is applicable to NORM. This may result in

confusion for these national regulatory authorities.

3.5.3 The lack of a homogeneous approach

One of the major challenges facing regulatory authorities is the inconsistency in the

protection system against exposure to NORM. Indeed, the ICRP has recognised the variability of

the regulation of exposure to NORM in its recommendation. The ICRP (2007b) declared that

there is a need for an international consensus on including or excluding exposure to NORM from

the scope of regulations. The main cause of this inconsistency may be the absence of guidance

on the protection against exposure to NORM. Furthermore, the definitions of the scope of the

66

NORM regulations vary across countries, which means that what is regulated in one country

may be exempted in other countries.

Another possible cause of the heterogeneous regulation of NORM could be the rapid

progress in the radiological protection system, which may prevent national authorities keeping

pace with new policies. Consequently, the latest standards are likely to be adopted in some

countries sooner than others. The language and desire to be consistent can also be effective

factors in this regard (Rochedo and Lauria, 2008). The terms and phrases used in the

recommendations may be interpreted or translated incorrectly. For example, the reference level

may be interpreted as a limit, which is not what is meant by the ICRP.

Inconsistency appears mostly in the variations in the clearance and exemption levels of

radionuclides in the natural decay chains. Within Australia, for example, the exemption level of

226Ra is 10 kBq in Queensland, while in the bordering state, New South Wales, it is 40 kBq

(ARPANSA, 2005). There is also inconsistency in the terms used to describe NORM in the

policies. For example, in the US, the acronym TENORM is used for enhanced level of NORM

industries (USEPA, 2008), while in IAEA publications, TENORM is included in the acronym

NORM (IAEA, 2007b).

The absence of a homogeneous approach to regulate NORM may have economic, social

and political impact. The international trading of raw materials has been affected by this lack of

a harmonised approach (Wymer, 2007). For instance, some industries may migrate to countries

or territories with less strict regulatory systems. Consequently, workers may move to a new

place or lose their jobs, which in turn would lead to social and financial difficulties.

Nevertheless, much effort has been recently devoted to achieving a homogeneous

system to regulate NORM across international and national jurisdictions. At the international

level, the IAEA has published a general safety guide that is directly applicable to NORM: on the

Classification of Radioactive Waste (IAEA, 2009), in addition to a safety guide on the application

of the concepts of exclusion, exemption and clearance (IAEA, 2005). The ICRP is preparing a

guideline for NORM, which will be available in the near future. On the national level, in Australia

for example, the Australian Radiation Protection and Nuclear Safety Agency (ARPANSA) has

recently established the National Directory for Radiation Protection to provide a uniform

framework for radiation safety, including NORMs to be adopted by all jurisdictions across

Australia (ARPANSA, 2013). More efforts are still required to overcome this problem.

3.5.4 Ubiquity of NORM

Naturally occurring radionuclides are ubiquitous in all materials and, to some extent, all

materials are NORM. It is present naturally in sand, minerals, ores, products, by-products,

commodities, foodstuff, air, water and species' bodies. This nature of NORM can create unaware

67

complex exposure situation, which can deliver a radiation dose to human and non-human

species. According to Steinhäusler (2010), the lack of awareness of the presence of NORM by

many industries has resulted in the release of millions of tons of NORM into the environment

without control. It may last for a long time post-operation and the resultant contamination can

extend to the surrounding environment and cause a prolonged exposure situation.

Cumulative evidence has shown a concentrated level of natural radionuclides in vegetables in

the vicinity of coal-fired stations (Baeza et al. 2012; Papp et al. 2002) as well as around

petroleum industry facilities (Rajaretnam and Spitz, 2000). The challenge that regulatory bodies

may face is the determination of which industrial process can enhance the concentration of

NORM, and which of them should fall within the scope of regulation.

3.5.5 Source of low dose

The current radiation protection system is based on the linear no-threshold (LNT)

model. This model is used to extrapolate the cancer risks associated with the exposure dose.

According to this model, the carcinogenic risk is linearly proportional to the radiation dose. This

means that even a very low dose ˂ 1 mSv can carry a potential risk (BEIR, 2006). In contrast,

evidence has shown that a low level of radiation can be beneficial, as the radiation hormesis

model predicts (Feinendegen, 2014).

The LNT model adequately describes the dose response at a high radiation dose;

however, at low doses, it is controversial. Statistical errors in epidemiological and experimental

carcinogenesis studies are dominant at low (˂ 10 mSv) and very low (˂ 1 mSv) doses, as well as

the dose rate (ICRP, 2005). These errors have raised doubts about the LNT model for estimating

cancer risk at low doses. The BEIR (VII) report has recognised the validity of the LNT model, and

has concluded that even the smallest dose has the potential to cause an increase in cancer risk

(BEIR, 2006).

On the other hand, the French Academy of Sciences reported that the biological

mechanisms at low doses are different from those at high doses. The academy concluded that

the use of the LNT model to assess the risk of doses below 20 mSv is unjustified and,

consequently, should be discouraged (Tubiana et al., 2005). Apparently, the LNT fails to

estimate the cancer risk at the low dose and the low dose rate region (Cohen, 2012).

Comprehensive reviews of recent findings on the effect of low dose levels have been published

(Averbeck, 2009; Calabrese and O'Connor, 2014; Dauer et al., 2010; ICRP, 2005; Tubiana et al.,

2006).

What is apparent is that the effects of low-level radiation doses are not fully understood.

The result may cause doubts about the current and future regulation of exposure due to NORM,

68

which may be considered as a source of low dose exposure. Solving the dilemma of dose

response in the low dose zone will reflect positively on the regulation of NORM.

3.5.6 Societal attitudes

The goal of the radiation protection system is to protect people and the environment

from the harm of ionising radiation, regardless of its size and origin (ICRP, 2007a). However,

natural radionuclides have received less attention than artificial radionuclides. This is perhaps

because of the lack of public awareness of the risks that exposure to natural sources may cause.

People’s perceptions about the risks from radiation are usually associated with artificial

radionuclides, which are well-regulated and contribute little to overall public exposure (Lee,

1992), although exposure to workers and members of the public from natural sources

considerably exceeds that from nuclear technologies (Steinhäusler, 2010). These perceptions,

usually voiced loudly, have led the radiological protection community to focus more on artificial

radionuclides. This attitude may have historical roots; developed with time from occupational

exposure due to X-rays and 226Ra to the arrival of nuclear power (ICRP, 2007b). From a public

point of view, nuclear power and weapons are related to artificial radionuclides, while natural

radionuclides are related to natural background radiation. In addition, the potential of a

catastrophic accident, such as a nuclear accident, does not exist in the case of NORM.

Societal attitudes have a large influence on the final decision of the level of radiation

safety (ICRP, 2007a). They are related to the principles of justification and optimisation (ICRP,

1977). The influence of societal attitudes can be shown clearly in the variability of decisions by

regulatory bodies regarding their definitions of the scope of NORM regulation from one region

to another (ICRP, 2007b). Thus, in many countries, NORM is not regulated by nuclear

commission bodies but by local jurisdictions or environmental-related bodies. For instance, in

the US, NORM is regulated by the Environmental Protection Agency (EPA), while man-made

radionuclides are regulated by the Nuclear Regulatory Commission (NRC). Nevertheless, the

goal of radiation protection needs to be achieved regardless of the source of radiation, whether

natural or artificial (ICRP, 2007b).

3.5.7 Measurement challenges

In the regulation context, measurements are necessary to verify compliance with the

requirements of protection and safety. Measurements are performed during, pre- and post-

operation phases for several reasons and applications, including risk assessment and periodic

monitoring checks of sources.

As mentioned earlier, NORM contains a range of radionuclides that vary in their

chemical and physical properties. They also interact with the ambient environment and can

migrate to other areas (Michalik et al., 2013; Swarzenski et al., 2003). Therefore, quantifying the

69

activity concentrations of individual radionuclides of NORM and estimating the risk assessment

are not straightforward tasks. They involve several complex parameters and exposure

pathways. Modelling and simulations are usually employed for estimating the risk assessment

for NORM (O'Brien and Cooper, 1998; Pontedeiro et al., 2007).

Measurement is an important factor that is considered when establishing a regulation

standard (IAEA, 2014). The current available techniques and associated difficulties are reflected

in the regulatory standards. Using radon regulation as an example, the actual health risks are

caused by its progeny; however, reference levels are usually given in terms of radon gas

concentrations as a surrogate for radon progeny. The reason is related to the measurement

simplicity and cost-effectiveness of radon gas over the progeny measurement (WHO, 2009). On

the other hand, the data on thoron gas and its progeny is scarce because of their measurement

difficulties. Regarding its impact on the regulation, thoron receives less attention than radon,

even though the health risk of thoron progeny is comparable with that of radon progeny (Shang

et al., 2005).

The measurement and management of NORM are interrelated. The improvement of the

measurement techniques of NORM will have a positive impact on its regulation. National

authorities may need to provide procedures and protocols for the measurement techniques to

assist operators and ensure the accuracy of results. For example, the USEPA has published

protocols that provide a technological guidance for measuring radon and its decay products in

homes (USEPA, 1993).

3.5.8 Waste and residues

NORM, in most industries, presents accidently and therefore, it is mainly concentrated in

waste and residues, which are the main pathways of the release of NORM from practice to the

environment (O'Brien and Cooper, 1998). NORM in waste and residues can have a negative

impact on humans and the environment (Michalik et al., 2013; Steinhäusler, 2005). It can

interact with the surrounding environments and individual radionuclides may leach and

transport causing contamination to the environment (Rajaretnam and Spitz, 2000). Once

radionuclides are released or leached from waste sites, it becomes difficult for them to be traced

in the environment.

Numerous studies have reported contamination of the surrounding environment due to

the discharge of NORM from industries (Boryło et al., 2012; Cañete et al., 2008; Michalik, 2008;

Paschoa, 1998; Pontedeiro et al., 2007; Thiry and Van Hees, 2008). Zeevaert et al. (2006)

observed 210Po in food crops near a coal-fired power plant. They attributed the presence of 210Po

to the atmospheric discharge of fly-ash from that facility. Rajaretnam and Spitz (2000) also

detected a concentrated level of 226Ra in vegetation in the surrounding area of a petroleum

70

industry site. The IAEA has acknowledged this issue by publishing a number of guidelines on

NORM which focus mainly on the management of waste and residues (IAEA, 2003, 2007a, 2012,

2013b).

Industrial waste and residues are usually generated in large quantities in different forms

and, most importantly, they contain long-lived radionuclides. The activity concentrations of

these radionuclides are found to be elevated in several industries (Abo-Elmagd et al., 2010;

Chanyotha et al., 2012; Hilal et al., 2014). The management of such waste and residues is one of

the major challenges facing the radiological protection community. Regulatory bodies strongly

encourage recycling and reusing of residue rather than disposing it of as waste (IAEA, 2006b).

This would depend on the type of residue, the rate of generating it, location of the facility and

market conditions (Haridasan, 2015). An example of recycling residues is the use of

phosphogypsum, which is a residue of a phosphate fertiliser processing plant, as a by-product to

manufacture building materials (Cooper, 2005; Somlai et al., 2008). However, utilising residues

requires large spaces for storage and may also need special treatment, which may raise an

economic issue.

The recycling strategy is important to be considered for managing NORM waste and

residues. However, if recycling is not feasible, then the residues have to be treated as waste

(IAEA, 2013a). A graded approach is recommended to regulate and manage radioactive waste in

practices involving NORM (IAEA, 2006a, 2013b). This will require characterising NORM in that

industry and ensuring that the waste management plan is safe, technically optimal and cost-

effective (IAEA, 2013a). There are situations where the regulation of NORM waste is unclear,

such as waste containing both NORM and artificial radionuclides in a mix. In such situations, the

rules designed and developed for one category of radioactive waste cannot be easily applied to

another (Michalik et al., 2013).

3.5.9 Ethical approach

In 1977, the ICRP introduced the three fundamental principles of the radiological

protection system: justification, optimisation and limitation. These principles represent the

ethical foundation of the radiation protection system. The justification principle ensures that

the benefit outweighs the harm and the optimisation principle aims to keep all exposure as low

as reasonably achievable under the prevailing circumstances. These two principles correspond

to utilitarian ethics and the third principle, the dose limit, satisfies deontological ethics

(Hansson, 2007). The justification also represents a virtue ethic. In utilitarianism, actions are

judged by their overall consequences. The positive and negative effects of actions are weighted.

The benefits or the good of actions must be maximised and outweigh the harm. On the other

71

hand, deontology focuses on the rightness or wrongness of actions themselves, rather than their

consequences.

The ethical dimensions of radiological protection have been proposed mainly for man-

made radionuclides or planned exposure situations, in which the source is introduced

intentionally. On the other hand, NORM exposure is adventitious in most practices, and the

source is not deliberately introduced. Moreover, some industries can cause dissemination of

NORM radiation in consumer products, which may raise indoor exposure, such as in building

materials with elevated uranium and thorium content (Kovler, 2009).

An important ethical issue for the radiological protection system is the naturalistic

fallacy, in which ethical problems are reduced to purely scientific problems, as in the case of

natural background radiation (Shrader-Frechette and Persson, 2002). The author argues that

background radiation should not be used as a criterion for ethically acceptable exposure levels.

An alternative ethical dimension may be required to be implemented within, to the

protection from exposure to NORM. One suggestion is to introduce a pragmatic and flexible

approach to radiological protection against NORM (Cool, 2015; Loy, 2015). The pragmatic

approach is already applied in the nuclear industry, especially for waste, which may pose long-

term risks for humans and the environment (Blowers, 2010; Shrader-Frechette, 2013). In

addition, flexibility is expected to be recommended in the future ICRP guideline being prepared

specifically for NORM exposure, and it will apply to the application of dose limits in the planned

exposure situation (Lecomte, 2015).

3.6 Case study

Radon is a special case of NORM and is treated separately because of its possibility to

accumulate, capability to harm and feasibility to control. These have raised the awareness of

radiation exposure due to radon among international and, to a lesser extent, national

organisations. Consequently, a number of recommendations, standards and guidelines have

been published and others will be available in the future (IAEA, 2015; ICRP, 2014; WHO, 2009).

The publications provide a framework for managing and controlling exposure due to radon gas

and its progeny with the goal of protecting people against radon exposure.

There are three isotopes of radon: 222Rn, 220Rn and 219Rn derived from the three natural

decay series 238U, 232Th and 235U, respectively. The most significant source of radon exposure is

delivered by 222Rn due to its relatively long half-life (3.82 days) compared with 220Rn (55.6

seconds) and 219Rn (3.96 seconds). The exposure dose from the latter isotope is negligible due

to its short half-life and low abundance of 235U in nature.

72

3.6.1 Indoor 222Rn regulation

The current regulation of indoor 222Rn distinguishes between two exposure situations:

existing and planned (IAEA, 2014). Generally, indoor 222Rn in dwellings and workplaces is

managed as an existing exposure situation because it is a natural source of radiation and existed

when the decision of the regulation was made. In such a situation, workers are treated in the

same way as general members of the public and the protection from exposure is achieved by the

use of reference levels and optimisation. National authorities are responsible for setting an

appropriate reference level for 222Rn concentrations in indoor environments (IAEA, 2015). If the

concentration of indoor 222Rn exceeds the proposed reference level, intervention actions may be

needed. In workplaces, if for any reason 222Rn concentration after the intervention remains

above the reference level, then the requirements of occupational exposure in a planned

exposure situation have to be applied to workers (IAEA, 2014). Another situation in which

exposure to 222Rn is considered as a planned exposure is in a workplace where other

radionuclides in the uranium decay chains are managed as a planned exposure situation.

National authorities are encouraged to establish a radon action plan for controlling

222Rn exposure to the public in indoor environments. Recommendations and guidance have

been provided by international organisations in this regard (IAEA, 2014, 2015; ICRP, 2010,

2014; WHO, 2009). The 222Rn action plan is suggested to include the establishment of an

appropriate reference level, the identification of radon prone areas and the implementation of

actions to reduce 222Rn exposure in existing and new buildings (IAEA, 2015). The national

authority may add other components to the action plan according to the situation.

The reference level is the key parameter in controlling indoor 222Rn. Several reference

levels have been recommended for the purpose of protection against 222Rn exposure in

dwellings and workplaces. It is suggested to set reference levels based on an annual effective

dose of around 10 mSv per year (IAEA, 2015). However, for practical reasons, the reference

level is provided in terms of radon concentration. The conversion between the two quantities

requires a dose conversion factor to calculate the effective dose. It is usually associated with the

assumption of a 0.40 equilibrium factor for 222Rn indoors and annual occupancy factors of 7000

hours indoors or 2000 hours at work. Table 3.1 shows the development of the dose conversion

factors over the past few decades. These factors were derived using either the dosimetric

approach, which is based on the Human Respiratory Tract Model (HRTM) developed by the

ICRP (ICRP, 1994), or epidemiological approach, which is based on a direct comparison to the

risk to uranium miners (ICRP, 1993). Most recently, the ICRP reported that it intended to use

the dosimetric model again (ICRP, 2010).

Reference levels of less than 300 Bq m-3 are recommended internationally for 222Rn in

indoor environments, as reported in the latest recommendations of ICRP, IAEA, WHO and EC.

73

On a national level, some countries have not yet updated their 222Rn regulations, while other

countries’ indoor 222Rn levels are yet to be regulated. In many countries, 222Rn regulation is

limited to workers in a planned exposure situation (e.g. mines). In Australia, for example, a

reference level of 200 and 1000 Bq m-3 is recommended by ARPANSA for dwellings and

workplaces, respectively (ARPANSA, 2002). However, public exposure to indoor 222Rn is not

under the scope of regulation in any jurisdiction.

The tools for controlling indoor 222Rn exposure are preventive measures for new

buildings and corrective actions for existing buildings (WHO, 2009). The aim of these actions is

to prevent 222Rn ingress into facilities and reduce the current elevated 222Rn level in indoor

environments. There are some circumstances in which the regulatory authority decides when

the intervention actions are mandatory or voluntary (IAEA, 2015). It is always mandatory to

take corrective actions when the annual dose due to exposure to 222Rn in dwellings and other

buildings is greater than 100 mSv (IAEA, 2015). Co-operation with other governmental

organisations may require ensuring the effective implementation of an indoor 222Rn control

plan. The emanation of 222Rn from building materials should be controlled by limiting 226Ra

content in these materials (WHO, 2009). The release of 222Rn from the water supply may also

need to be controlled by following the approach provided by WHO (2011).

Table 3.1 The development of dose conversion factor over time

Factor Up-until Publication 65 Publication 103

Publication 115

Publication 126

1993 (ICRP, 1993) (ICRP, 2007a) (ICRP, 2010) (ICRP, 2014) Risk approach Dosimetric Epidemiological Epidemiological Dosimetric Dosimetric Protection level (mSv y-1) Action

level (10 ) Action level (3–10 )

Reference level (10)

Reference level (10)

Reference level (10)

Dose Conversion factor (nSv (Bq h m-3)-1)*

15 9 9 Under revision Under revision

222Rn level (Bq m-3)

Dwelling 200–600 600 300 300 Workplace 1500 500–1500 1000–1500 1000 300

* Based on UNSCEAR calculations in 1982, 1988 and 2000 taking into account radon progeny in terms of equilibrium equivalent concentration (EEC) which represents the concentration of 222Rn gas, in equilibrium with its progeny that would have the same potential alpha energy as the existing non-equilibrium mixture.

3.6.2 New policies and recommendations on indoor 222Rn

In view of the latest findings, relevant international organisations have revised and

updated their guidelines and policies for protection against 222Rn exposure (EU, 2013; IAEA,

2015; ICRP, 2014). The revised guidelines are focused more on public exposure due to 222Rn

indoors, because occupational exposure due to 222Rn in planned exposure situations is well-

regulated. The reference levels of 222Rn in dwellings and workplaces have been reduced, with

slight variations across international organisations. The ICRP recommends a reference level

ranging between 100 and 300 Bq m-3 for all buildings, without distinguishing between dwellings

and workplaces (ICRP, 2014). These replaced the previously recommended levels of 300 and

74

1000 Bq m-3 for dwellings and workplaces, respectively (ICRP, 2010). The IAEA has adopted the

same approach as an alternative to its main approach for indoor 222Rn policy. The IAEA

recommends a reference level that does not exceed an annual average 222Rn concentration of

300 Bq m-3 for dwellings as well as workplaces with a high occupancy factor for members of the

public. In workplaces with a low occupancy factor, the reference level of 222Rn concentration

should not exceed 1000 Bq m-3 (IAEA, 2015). The European Commission has reduced its

reference levels from 200 and 500 Bq m-3 to 100 and 300 Bq m-3 for both dwellings and

workplaces, respectively (EU, 2013).

The new policies emphasise the regulator’s role in providing information on indoor

222Rn levels in buildings. The information includes the average activity concentrations and

associated risks of 222Rn in dwellings and other buildings with a high occupancy factor (IAEA,

2015). This may be accomplished by conducting a national survey. It may also include

information about 222Rn testing techniques and protocols (ICRP, 2014). This information may

then be used to establish and implement an action plan for controlling public exposure due to

indoor 222Rn. Another new policy for controlling indoor 222Rn is the involvement of individual

responsibility, the so-called “legal responsible person” (ICRP, 2014). It is a responsibility of a

person of a building towards the users of that building. It could be stakeholders towards

occupants, an employer towards employees or house owners towards tenants. This legal

responsibility should be addressed and reflected in the indoor 222Rn protection strategy plan.

Although the health risk of indoor 222Rn exposure in smokers is higher than non-

smokers (Darby et al., 2005; Leenhouts, 1999), the new policies of the 222Rn protection system

do not distinguish between the two groups (ICRP 2014). Instead, national authorities are

encouraged to consider the conjunction of 222Rn policy with other health policies, such as non-

smoking and indoor air quality (IAEA, 2015), considering that ceasing smoking is likely to

reduce the health risks of 222Rn (Bochicchio, 2008; Méndez et al., 2011).

3.6.3 Implications of the implementation of the new policies

National authorities are encouraged to take steps to adopt the new recommendations on

indoor 222Rn. Despite the benefits of protecting humans against exposure to indoor 222Rn, the

implementation of the new 222Rn policies and recommendations may challenge national

authorities. Financial and practical implications are the main concerns of this section.

The cost implications of the new policies may affect the corresponding governmental

body as well as stakeholders and residents. The relevant authorities are required to undertake a

national survey in order to determine 222Rn exposure to the population and set an appropriate

reference level. The cost of such a survey depends on several factors, including the size of the

population, the measurement technique, staff training and sampling duration. Individuals,

75

whether stakeholders, employers or residents, may also be required to conduct measurements

for indoor 222Rn levels in premises under their responsibility.

The reduction of the reference level in the new recommendations means that more

buildings would be above the proposed reference level. Intervention actions in certain buildings

may be required to reduce the 222Rn level below the reference level. Some cost-effective

measures for existing and new buildings were highlighted in the WHO (2009) Report. These

extra costs can affect the decision of the legal responsible person or organisation regarding

complying with the requirement of 222Rn regulation and, consequently, influence the efficient

implementation of a protection strategy against indoor 222Rn.

In 2010, the ICRP announced that the nominal risk coefficient for 222Rn exposure should

be double the value previously reported and thus, new dose conversion factors will be

published in the near future. Subsequently, the cases in which corrective action is mandatory,

when the annual effective dose due to 222Rn concentration is greater than 100 mSv, are likely to

increase. In addition, some workplaces that may previously have been below the reference level

may now exceed it, if it is reduced in response to the new dose conversion factors.

Practically, the implementation of the new policies on 222Rn protection may cause some

confusion to the regulatory authorities. This is because the new policies are structured on a

situations-based approach of a radiological protection system, which replaced the

practice/intervention classification. Some countries are still regulating ionising radiation based

on the previous approach. The main confusion could arise from the fact that two situations are

applicable to indoor 222Rn, namely existing and planned exposure situations. This problem has

had a negative impact on the implementation of the standard and this has been recognised by

some authors (Chen, 2015; Lochard, 2015).

Another confusing issue is the level of occupancy factor mentioned by IAEA- BSS (2014).

Exposure to indoor 222Rn in buildings with a high occupancy factor for the public are treated as

dwellings, while buildings with a low occupancy factor are treated as workplaces, with a

correspondingly higher reference level and potential implementation of occupational exposure

to workers. There is no clear definition of when occupancy is deemed to be high or low, except

for some examples, such as schools attended by students and offices attended by visitors.

Instead, the IAEA recommends that the national authority decides which buildings should be

considered as having a high occupancy factor for general members of the public.

In recent years, there have been a number of changes in indoor 222Rn policies. These

changes may not be adopted by some authorities, which could create a heterogeneous pattern of

222Rn management across nations and states. National regulatory authorities need to keep pace

with the latest changes to avoid such situations.

76

3.6.4 The case of indoor 220Rn

Indoor 220Rn has received less attention than 222Rn due to the very short half-life of 220Rn

compared with 222Rn. However, the risk presented by the inhalation of 220Rn and its decay

products is considered to be equal if not greater than that of 222Rn because 50 % or more of the

total potential alpha energy concentration (PAEC) is caused by 220Rn progeny (Shang et al.,

2005). Therefore, determining the activity concentration of 220Rn alone may not reflect the

actual risk level.

The level of indoor 220Rn gas drops significantly as it travels away from walls or its

origin (Meisenberg and Tschiersch, 2011). It can lose more than 99 % of its value in the indoor

environment at a distance of 1 m from the wall surface (McLaughlin, 2010). However, 220Rn

progeny may remain in indoor air at a considerable level causing potential exposure to the

occupants.

On the other hand, the data available for 220Rn is limited, perhaps because of the

measurement difficulties associated with its short half-life (Tokonami, 2010; Zhang et al., 2010).

There is also lack of epidemiological studies on the exposure of 220Rn. These factors have a

negative impact on the current regulation of 220Rn. Perhaps it is for these reasons, 220Rn is not

regulated in most parts of the world, although the IAEA has included it in its recent guidelines

(IAEA, 2015). The IAEA recommends indoor 220Rn to be regulated only if necessary, when the

activity concentration of 220Rn in indoor air is high, but no value is provided. The control of

indoor 220Rn can be accomplished by controlling 232Th in building materials. The management of

indoor 220Rn is identical to that of 222Rn in indoor environments (IAEA, 2014). The same

procedure of developing an action plan and establishing a reference level for 222Rn is followed

for 220Rn with differences in dose conversion factors. The current dose conversion factor (DCF)

derived for 220Rn is 40 nSv (Bq h m-3)-1 for equilibrium equivalent concentrations (EEC) as

proposed by UNSCEAR (2000).

3.7 Conclusion

Protection against NORM is an ongoing concern for regulatory authorities. Exposure to

NORM is currently regulated with some ambiguity and confusion with the absence of a

comprehensive framework. It is considered as an existing exposure situation but also can be a

planned exposure situation. There are nine regulatory issues and challenges related to NORM

that have been highlighted in this review, including, heterogeneous approach, ubiquities nature,

wastes management, social attitudes, measurement methods and ethical values. These issues

created gaps in the current radiological protection system. Addressing these issues will not only

be crucial for characterising NORM exposure in different human activities, it will be essential for

improving the radiological protection system. The diversity of human activities involving NORM

77

and the complexity of exposure pathways dictate continuing efforts to formulate a framework

for protecting the public, workers and the environment. The framework should be broad

enough to accommodate a variety of industrial activities with enhanced levels of NORM,

including exploring, extraction, processing, production and disposal. New recommendations

and policies are currently underway with the aim of providing a clear and implementable

approach to regulating NORM.

78

References

Abbady, A.G.E. and AM El-Arabi. 2006. "Naturally occurring radioactive material from the aluminium industry—a case study: the Egyptian Aluminium Company, Nag Hammady, Egypt." Journal of Radiological Protection 26: 415.

Abo-Elmagd, M., HA Soliman, K.A. Salman and NM El-Masry. 2010. "Radiological hazards of TENORM in the wasted petroleum pipes." Journal of Environmental Radioactivity 101 (1): 51-54.

Al-Masri, MS and H. Suman. 2003. "NORM waste management in the oil and gas industry: The Syrian experience." Journal of Radioanalytical and Nuclear Chemistry 256 (1): 159-162.

ARPANSA. 2002. Recommendations for limiting exposure to ionizing radiation (1995) and national standard for limiting occupational exposure to ionizing radiation (republished 2002), Radiation Protection Series No.1. Yallambie

ARPANSA. 2005. "Naturally-occurring radioactive material (norm) in Australia: issues for discussion." http://www.arpansa.gov.au/pubs/norm/rhsac_disc.pdf.

ARPANSA. 2013. National Directory for Radiation Protection, Radiation Protection Series Publication No. 6. YALLAMBIE: Australian Radiation Protection and Nuclear Safety Agency.

Aslam, Muhammad, Rahmat Gul, Tauseef Ara and Manzur Hussain. 2012. "Assessment of radiological hazards of naturally occurring radioactive materials in cement industry." Radiation Protection Dosimetry 151 (3): 483-488.

Averbeck, D. 2009. "Does scientific evidence support a change from the LNT model for low-dose radiation risk extrapolation?" Health Physics 97 (5): 493.

Bakr, WF. 2010. "Assessment of the radiological impact of oil refining industry." Journal of Environmental Radioactivity 101 (3): 237-243.

Ballesteros, L., I. Zarza, J. Ortiz and V. Serradell. 2008. "Occupational exposure to natural radioactivity in a zircon sand milling plant." Journal of Environmental Radioactivity 99 (10): 1525-1529.

Baysson, H., M. Tirmarche, G. Tymen, S. Gouva, D. Caillaud, J.C. Artus, A. Vergnenegre, F. Ducloy and D. Laurier. 2004. "Indoor radon and lung cancer in France." Epidemiology 15 (6): 709.

Beddow, H, S Black and D Read. 2006. "Naturally occurring radioactive material (NORM) from a former phosphoric acid processing plant." Journal of Environmental Radioactivity 86 (3): 289-312.

BEIR. 2006. Health risks from exposure to low levels of ionizing radiation: BEIR VII Phase 2, National Research Council . Committee to Assess Health Risks from Exposure to Low Level of Ionizing Radiation. Washington: National Academies Press.

Blowers, Andrew. 2010. "Why dump on us? Power, pragmatism and the periphery in the siting of new nuclear reactors in the UK." Journal of Integrative Environmental Sciences 7 (3): 157-173.

Bochicchio, Francesco. 2008. "The radon issue: considerations on regulatory approaches and exposure evaluations on the basis of recent epidemiological results." Applied Radiation and Isotopes 66 (11): 1561-1566.

Boryło, Alicja, Bogdan Skwarzec and Grzegorz Olszewski. 2012. "The radiochemical contamination (210Po and 238U) of zone around phosphogypsum waste heap in Wiślinka (northern Poland)." Journal of Environmental Science and Health, Part A 47 (5): 675-687.

Brown, S., D. Collier and P. Kenny. 2003. Radionuclides in the Iron Sinter Plant circuit BHP, Port Kembla: ANSTO report to Illawarra Radionuclide Investigations Committee.

Burton, EF. 1904. "XLVIII. A radioactive gas from crade petroleum." The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science 8 (46): 498-508.

Calabrese, Edward J and Michael K O'Connor. 2014. "Estimating risk of low radiation doses-a critical review of the BEIR VII Report and its use of the linear no-threshold (LNT) hypothesis." Radiation Research 182 (5): 463-474.

79

Cañete, Socrates Jose P, Lorna Jean H Palad, Eliza B Enriquez, Teofilo Y Garcia and Teresa Yulo-Nazarea. 2008. "Leachable 226Ra in Philippine phosphogypsum and its implication in groundwater contamination in Isabel, Leyte, Philippines." Environmental Monitoring and Assessment 142 (1-3): 337-344.

Cevik, U., N. Damla and S. Nezir. 2007. "Radiological characterization of Cayirhan coal-fired power plant in Turkey." Fuel 86 (16): 2509-2513.

Chanyotha, S, C Kranrod, N Chankow, R Kritsananuwat, P Sriploy and K Pangza. 2012. "Natural radionuclide concentrations in processed materials from thai mineral industries." Radiation Protection Dosimetry 152 (1-3): 71-75.

Chen, Jing. 2015. "A discussion on norm guidelines: proposed revisions needed for practical use." Radiation Protection Dosimetry: 1–6.

Clarke, R. and J. Valentin. 2005. "A history of the International Commission on Radiological Protection." Health Physics 88 (6): 717.

Cohen, Bernard L. 2012. "The cancer risk from low level radiation." In Radiation Dose from Multidetector CT, 61-79: Springer.

Cool, DA. 2015. "Review of the ICRP system of protection: the approach to existing exposure situations." In Proceedings of the Second International Symposium on the System of Radiological Protection edited: Annals of the ICRP 44 (1S).

Cooper, Malcolm B. 2005. "Naturally occurring radioactive materials (NORM) in Australian industries—review of current inventories and future generation." Report ERS-006 (EnviroRad Services Pty. Ltd.) to the Radiation health and Safety Advisory Council.

Darby, S., D. Hill, A. Auvinen, J.M. Barros-Dios, H. Baysson, F. Bochicchio, H. Deo, R. Falk, F. Forastiere and M. Hakama. 2005. "Radon in homes and risk of lung cancer: collaborative analysis of individual data from 13 European case-control studies." Bmj 330 (7485): 223.

Dauer, L.T., A.L. Brooks, D.G. Hoel, W.F. Morgan, D. Stram and P. Tran. 2010. "Review and evaluation of updated research on the health effects associated with low-dose ionising radiation." Radiation Protection Dosimetry 140 (2): 103.

EU. 2013. Euratom Basic Safety Standards Directive., Council directive 2013/59/Euratom. Feinendegen, LE. 2014. "Evidence for beneficial low level radiation effects and radiation

hormesis." The British journal of radiology. Hansson, Sven Ove. 2007. "Ethics and radiation protection." Journal of Radiological Protection 27

(2): 147. Haridasan, PP. 2015. "Managing Exposure to Natural Sources - International Standards and New

Challenges." In Proceedings of the seventh International Symposium on Naturally Occurring Radioactive Material (NORM VII), Beijing, China, 22–26 April 2013, edited. Vienna: International Atomic Energy Agency.

Haridasan, PP, M Harikumar, PM Ravi and RM Tripathi. 2015. "Radiological protection against exposure to naturally occurring radioactive material." Radiation Protection and Environment 38 (3): 59.

Haridasan, PP, CG Maniyan, PMB Pillai and AH Khan. 2002. "Dissolution characteristics of 226 Ra from phosphogypsum." Journal of Environmental Radioactivity 62 (3): 287-294.

Haridasan, PP, PMB Pillai, AH Khan and VD Puranik. 2006. "Natural radionuclides in zircon and related radiological impacts in mineral separation plants." Radiation Protection Dosimetry 121 (4): 364.

Hedemann-Jensen, Per. 2010. "Naturally occurring radiation sources: existing or planned exposure situation?" Journal of Radiological Protection 30 (4): 781.

Hilal, MA, MF Attallah, Gehan Y Mohamed and M Fayez-Hassan. 2014. "Evaluation of radiation hazard potential of TENORM waste from oil and natural gas production." Journal of Environmental Radioactivity 136: 121-126.

Himstedt, F. 1904. "Über die radioaktive Emanation der Wasser‐und Ölquellen." Annalen der Physik 318 (3): 573-582.

Hirose, K. 2011. "2011 Fukushima Daiichi nuclear power plant accident: Summary of regional radioactive deposition monitoring results." Journal of Environmental Radioactivity

80

Hosseini, A., NA Beresford, JE Brown, DG Jones, M. Phaneuf, H. Thørring and T. Yankovich. 2010. "Background dose-rates to reference animals and plants arising from exposure to naturally occurring radionuclides in aquatic environments." Journal of Radiological Protection 30: 235.

IAEA. 2003. Radiation protection and the management of radioactive waste in the oil and gas industry, Safety reports series; no. 34. Vienna: International Atomic Energy Agency.

IAEA. 2005. Derivation of activity concentration values for exclusion, exemption and clearance. Vienna: International Atomic Energy Agency.

IAEA. 2006a. Assessing the need for radiation protection measures in work involving minerals and raw materials, Safety Reports Series No.49 Vienna: International Atomic Energy Agency.

IAEA. 2006b. Fundamental Safety Principles, Safety Standards Series No. SF-1. Vienna: International Atomic Energy Agency.

IAEA. 2007a. Radiation protection and NORM residue management in the zircon and zirconia industries, Safety Reports Series, No. 51. Vienna.

IAEA. 2007b. Safety glossary : terminology used in nuclear safety and radiation protection. 2007 ed. Vienna: International Atomic Energy Agency.

IAEA. 2009. Classification of radioactive waste : safety guide. Vienna: International Atomic Energy Agency.

IAEA. 2011. Radiation protection and NORM residue management in the production of rare earths from thorium containing minerals, Safety Reports Series, No. 68. Vienna.

IAEA. 2012. Radiation protection and NORM residue management in the titanium dioxide and related industries, Safety Reports Series, No. 76. Vienna: International Atomic Energy Agency.

IAEA. 2013a. Management of NORM residues, IAEA TECDOC No. 1712. Vienna. IAEA. 2013b. Radiation protection and management of NORM residues in the phosphate industry.

Vienna: International Atomic Energy Agency. IAEA. 2014. Radiation protection and safety of radiation sources : international basic safety

standards. Vienna: International Atomic Energy Agency. IAEA. 2015. Protection of the public against exposure indoors due to radon and other natural

sources of radiation specific safety guide. Vienna: International Atomic Energy Agency. ICRP. 1951. "International recommendations on radiological protection: Revised by the

International Commission on Radiological Protection at the Sixth International Congress of Radiology, London, July 1950." British Journal of Radiology. doi: 10.1016/s0074-2740(28)80013-4.

ICRP. 1955. "1954 Recommendations of the International Commission on Radiological Protection." BJR

ICRP. 1965. "Report by Committee 4 of the International Commission on Radiological Protection. Principles of Environmental Monitoring related to the Handling of Radioactive Materials." ICRP Publication 7 (11).

ICRP. 1977. "Recommendations of the ICRP." ICRP Publication 26 Ann. ICRP 1. (3) (3). ICRP. 1991. "1990 Recommendations of the International Commission on Radiological

Protection." ICRP Publication 60. Ann. ICRP 21 (1-3) (1-3). ICRP. 1993. "Protection Against Radon-222 at Home and at Work." ICRP Publication 65. Ann.

ICRP 23 (2). ICRP. 1994. "Human respiratory tract model for radiological protection " ICRP Publication 66.

Ann. ICRP 24. ICRP. 1999. "Protection of the Public in Situations of Prolonged Radiation Exposure." ICRP

Publication 82. Ann 29 (1-2). ICRP. 2003. ICRP Publication 91: A Framework for Assessing the Impact of Ionising Radiation on

Non-human Species. Vol. 33: (3). ICRP. 2005. "Low-dose extrapolation of radiation-related cancer risk." Annals of the ICRP 35 (4):

1.

81

ICRP. 2006. " Assessing dose of the representative person for the purpose of radiation protection of the public and the optimisation of radiological protection." Annals of the ICRP 36: 3.

ICRP. 2007a. "The 2007 recommendations of the International Commission on Radiological Protection." ICRP Publication 103. Ann 37 (2-4).

ICRP. 2007b. "Scope of Radiological Protection Control Measures." ICRP Publication 104. Ann 35 (5).

ICRP. 2009a. "Environmental protection: transfer parameters for reference animals and plants." Annals of the ICRP 39 (6): 1.

ICRP. 2009b. "The History of ICRP and the Evolution of its Policies." Annals of the ICRP 39 (1): 75-110.

ICRP. 2010. "Lung cancer risk from radon and progeny and statement on radon. ICRP Publication 115." Annals of the ICRP 40 (1).

ICRP. 2014. "Radiological Protection against Radon Exposure. ICRP Publication 126." Ann. ICRP 43 (3) (3).

ICRP. 2016. "Radiological Protection from Cosmic Radiation in Aviation." ICRP Publication 132. Ann. 45 (1): 1–48.

Iwaoka, K., K. Tagami and H. Yonehara. 2010. "Natural radioactivities in iron and nickel ores imported into Japan and the dose assessment for workers handling them." Journal of Radiological Protection 30: 613.

IXRPC. 1928. "International recommendations for X-ray and radium protection." Radiology 23: 682-685.

IXRPC. 1934. "International recommendations for X-ray and radium protection." Radiology 23: 682-685.

Jobbágy, V., J. Somlai, J. Kovács, G. Szeiler and T. Kovács. 2009. "Dependence of radon emanation of red mud bauxite processing wastes on heat treatment." Journal of hazardous materials 172 (2): 1258-1263.

Kleinschmidt, R. and R. Akber. 2008. "Naturally occurring radionuclides in materials derived from urban water treatment plants in southeast Queensland, Australia." Journal of Environmental Radioactivity 99 (4): 607-620.

Kolb, WA and M Wojcik. 1985. "Enhanced radioactivity due to natural oil and gas production and related radiological problems." Science of the Total Environment 45: 77-84.

Kovler, Konstantin. 2009. "Radiological constraints of using building materials and industrial by-products in construction." Construction and Building materials 23 (1): 246-253.

Krolak, E., G. Golub and K. Barczak. 2012. "Caesium-137 and potassium-40 in selected oxbow lakes of the border Bug River more than 20 years after the Chernobyl accident." Water International 37 (1): 75-85.

Lecomte, JF. 2015. "Application of the Commission’s recommendations to naturally occurring radioactive material." Annals of the ICRP 44 (1 suppl): 188-196.

Lee, C., E.Y. Han and W.E. Bolch. 2007. "Consideration of the ICRP 2006 revised tissue weighting factors on age-dependent values of the effective dose for external photons." Physics in Medicine and Biology 52: 41.

Lee, TR. 1992. "The public's perception of radon." Radiation Protection Dosimetry 42 (3): 257-262.

Leenhouts, HP. 1999. "Radon-induced lung cancer in smokers and non-smokers: risk implications using a two-mutation carcinogenesis model." Radiation and environmental biophysics 38 (1): 57-71.

Lochard, J. 2015. "Application of the Commission’s recommendations: the 2013–2017 Committee 4 programme of work." Annals of the ICRP 44 (1 suppl): 33-46.

Loy, J. 2015. "What should a radiation regulator do about naturally occurring radioactive material?" In Proceedings of the Second International Symposium on the System of Radiological Protection edited: Annals of the ICRP 44 (1S).

McLaughlin, J. 2010. "An overview of thoron and its progeny in the indoor environment." Radiation Protection Dosimetry 141 (4): 316-321.

82

Meisenberg, O. and J. Tschiersch. 2011. "Thoron in indoor air: modeling for a better exposure estimate." Indoor Air 21 (3): 240-252.

Méndez, David, Omar Alshanqeety, Kenneth E Warner, Paula M Lantz and Paul N Courant. 2011. "The impact of declining smoking on radon-related lung cancer in the United States." American journal of public health 101 (2): 310.

Michalik, B. 2008. "NORM impacts on the environment: An approach to complete environmental risk assessment using the example of areas contaminated due to mining activity." Applied Radiation and Isotopes 66 (11): 1661-1665.

Michalik, Bogusław, J Brown and P Krajewski. 2013. "The fate and behaviour of enhanced natural radioactivity with respect to environmental protection." Environmental Impact Assessment Review 38: 163-171.

O'Brien, RS and MB Cooper. 1998. "Technologically enhanced naturally occurring radioactive material (NORM): pathway analysis and radiological impact." Applied Radiation and Isotopes 49 (3): 227-239.

Palomo, M., A. Peñalver, C. Aguilar and F. Borrull. 2010. "Presence of Naturally Occurring Radioactive Materials in sludge samples from several Spanish water treatment plants." Journal of hazardous materials 181 (1): 716-721.

Parami, V.K., S.K. Sahoo, H. Yonehara, S. Takeda and L.L. Quirit. 2010. "Accurate determination of naturally occurring radionuclides in Philippine coal-fired thermal power plants using inductively coupled plasma mass spectrometry and [gamma]-spectroscopy." Microchemical Journal 95 (2): 181-185.

Paschoa, Anselmo S. 1998. "Potential environmental and regulatory implications of naturally occurring radioactive materials (NORM)." Applied Radiation and Isotopes 49 (3): 189-196.

Pontedeiro, EM, PFL Heilbron and Renato M Cotta. 2007. "Assessment of the mineral industry NORM/TENORM disposal in hazardous landfills." Journal of Hazardous Materials 139 (3): 563-568.

Rajaretnam, G. and H.B. Spitz. 2000. "Effect of leachability on environmental risk assessment for naturally occurring radioactive materials in petroleum oil fields." Health Physics 78 (2): 191.

Righi, S., M. Andretta and L. Bruzzi. 2005. "Assessment of the radiological impacts of a zircon sand processing plant." Journal of Environmental Radioactivity 82 (2): 237-250.

Rochedo, Elaine RR and Dejanira Lauria. 2008. "International versus national regulations: Concerns and trends." Applied Radiation and Isotopes 66 (11): 1550-1553.

Satterly, J and JC McLennan. 1918. "The radioactivity of the natural gases of Canada." Trans Royal Soc (Canada) 3: 153-160.

Shang, B., B. Chen, Y. Gao, Y. Wang, H. Cui and Z. Li. 2005. "Thoron levels in traditional Chinese residential dwellings." Radiation and environmental biophysics 44 (3): 193-199.

Shrader-Frechette, Kristin. 2013. "Environmental injustice inherent in radiation dose standards". In Social and Ethical Aspects of Radiation Risk Management, edited by Deborah Oughton and Sven Ove Hansson. Amsterdam: Elsevier

Shrader-Frechette, Kristin and Lars Persson. 2002. "Ethical, logical and scientific problems with the new ICRP proposals." Journal of Radiological Protection 22 (2): 149.

Smith, AL. 1987. "Radioactive-scale formation." Journal of Petroleum Technology 39 (6): 697-706.

Somlai, J., V. Jobbágy, J. Kovács, S. Tarján and T. Kovács. 2008. "Radiological aspects of the usability of red mud as building material additive." Journal of hazardous materials 150 (3): 541-545.

Steinhäusler, F. 2005. "Radiological impact on man and the environment from the oil and gas industry: Risk assessment for the critical group." Radiation Safety Problems in the Caspian Region: 129-134.

Steinhäusler, Friedrich. 2010. "Geohazards due to technologically enhanced natural radioactive wastes." Acta Geophysica 58 (5): 922-938.

83

Swarzenski, Peter W, Donald Porcelli, Per S Andersson and Joseph M Smoak. 2003. "The behavior of U-and Th-series nuclides in the estuarine environment." Reviews in Mineralogy and Geochemistry 52 (1): 577-606.

Taira, Y., N. Hayashida, S. Yamashita, T. Kudo, N. Matsuda, J. Takahashi, A. Gutevitc, A. Kazlovsky and N. Takamura. 2012. "Environmental contamination and external radiation dose rates from radionuclides released from the fukushima nuclear power plant." Radiation Protection Dosimetry 151 (3): 537-545.

Thiry, Yves and May Van Hees. 2008. "Evolution of pH, organic matter and 226 radium/calcium partitioning in U-mining debris following revegetation with pine trees." Science of the Total Environment 393 (1): 111-117.

Tokonami, S. 2010. "Why is 220Rn (thoron) measurement important?" Radiation Protection Dosimetry 141 (4): 335.

Tubiana, M., A. Aurengo, D. Averbeck, A. Bonnin, B. Le Guen, R. Masse, R. Monier, AJ Valleron and F. De Vathaire. 2005. "Dose-effect relationships and estimation of the carcinogenic effects of low doses of ionizing radiation." Int J Low Radiat 2 (3/4): 134-151.

Tubiana, M., A. Aurengo, D. Averbeck and R. Masse. 2006. "Recent reports on the effect of low doses of ionizing radiation and its dose–effect relationship." Radiation and environmental biophysics 44 (4): 245-251.

Tufail, M., N. Akhtar and M. Waqas. 2006. "Radioactive rock phosphate: the feed stock of phosphate fertilizers used in Pakistan." Health Physics 90 (4): 361.

Turhan, S., IH Arikan, H. Demirel and N. Gungor. 2011. "Radiometric analysis of raw materials and end products in the Turkish ceramics industry." Radiation Physics and Chemistry 80 (5): 620–625.

UNSCEAR. 2006. Sources-to-effects assessment for radon in homes and workplaces: United Nations Scientific Committee on the Effects of Atomic Radiation

USEPA. 1993. Protocol for radon and radon decay product measurements in homes. USEPA. 2008. Technologically Enhanced Naturally Occurring Radioactive Materials From

Uranium Mining, Volume 1: Mining and Reclamation Background, and Volume 2: Investigation of Potential Health, Geographic, and Environmental Issues of Abandoned Uranium Mines. Washington, DC.

Vacquier, B., A. Rogel, K. Leuraud, S. Caer, A. Acker and D. Laurier. 2009. "Radon-associated lung cancer risk among French uranium miners: modifying factors of the exposure–risk relationship." Radiation and environmental biophysics 48 (1): 1-9.

Vearrier, David, John A Curtis and Michael I Greenberg. 2009. "Technologically enhanced naturally occurring radioactive materials." Clinical toxicology 47 (5): 393-406.

Walker, J.S. 2000. Permissible dose: a history of radiation protection in the twentieth century: University of California Press. Accessed 5 Nov 2011. http://site.ebrary.com.ezp01.library.qut.edu.au/lib/qut/docDetail.action?docID=10057084

Weiss, D. 2011. "Distribution pattern of artificial radionuclides in the Baltic Sea in the special event of the Chernobyl fallout." Isotopes in Environmental and Health Studies 47 (3): 254-264.

White, G.J. and A.S. Rood. 2001. "Radon emanation from NORM-contaminated pipe scale and soil at petroleum industry sites." Journal of Environmental Radioactivity 54 (3): 401-413.

WHO. 2009. Handbook on indoor radon. Geneva: World Health Organization. WHO. 2011. Guidelines for Drinking-water Quality. 4th ed. Geneva: World Health Organization. Xhixha, G, GP Bezzon, C Broggini, GP Buso, A Caciolli, I Callegari, S De Bianchi, G Fiorentini, E

Guastaldi and M Kaçeli Xhixha. 2013. "The worldwide NORM production and a fully automated gamma-ray spectrometer for their characterization." Journal of Radioanalytical and Nuclear Chemistry 295 (1): 445-457.

Zeevaert, T., L. Sweeck and H. Vanmarcke. 2006. "The radiological impact from airborne routine discharges of a modern coal-fired power plant." Journal of Environmental Radioactivity 85 (1): 1-22.

84

Zhang, L., J. Wu, Q. Guo and W. Zhuo. 2010. "Measurement of thoron gas in the environment using a Lucas scintillation cell." Journal of Radiological Protection 30: 597.

85

Chapter 4 Radon-222 Activity Flux Measurement

Using Activated Charcoal Canisters: Revisiting the

Methodology

Sami H. Alharbi*, Riaz A. Akber

*School of Physical and Chemical Sciences, Queensland University of Technology, 2 George

Street, Brisbane, Qld 4000, Australia

Journal of Environmental Radioactivity 129 (2014), 94-99.

http://dx.doi.org/10.1016/j.jenvrad.2013.12.021

86

Statement of Contribution of Co-Authors for Thesis by Published Paper The authors listed below have certified that:

1. they meet the criteria for authorship in that they have participated in the conception,

execution, or interpretation, of at least that part of the publication in their field of

expertise;

2. they take public responsibility for their part of the publication, except for the

responsible author who accepts overall responsibility for the publication;

3. there are no other authors of the publication according to these criteria;

4. potential conflicts of interest have been disclosed to (a) granting bodies, (b) the editor

or publisher of journals or other publications, and (c) the head of the responsible

academic unit, and

5. they agree to the use of the publication in the student’s thesis and its publication on the

Australasian Research Online database consistent with any limitations set by publisher

requirements.

In the case of this chapter:

Radon-222 Activity Flux Measurement Using Activated Charcoal Canisters: Revisiting the

Methodology

Contributor Statement of contribution

Sami

Alharbi

Signature:

Date: 31/03/2016

Conducted experimental work, prepared samples of analysis,

analysed and interpreted the data, wrote the manuscript

Riaz Akber Original idea, Assisted with editorial process,

helped in experimental design, reviewed the manuscript.

Principal Supervisor Confirmation

I have sighted email or other correspondence from all Co-authors confirming their certifying

authorship.

Name:

Signature:

Date:

87

Abstract

The measurement of radon (222Rn) activity flux using activated charcoal canisters was

examined to investigate the distribution of the adsorbed 222Rn in the charcoal bed and the

relationship between 222Rn activity flux and exposure time. The activity flux of 222Rn from five

sources of varying strengths was measured for exposure times of one, two, three, five, seven, 10,

and 14 days. The distribution of the adsorbed 222Rn in the charcoal bed was obtained by

dividing the bed into six layers and counting each layer separately after the exposure. 222Rn

activity decreased in the layers that were away from the exposed surface. Nevertheless, the

results demonstrated that only a small correction might be required in the actual application of

charcoal canisters for activity flux measurement, where calibration standards were often

prepared by the uniform mixing of radium (226Ra) in the matrix. This was because the diffusion

of 222Rn in the charcoal bed and the detection efficiency as a function of the charcoal depth

tended to counterbalance each other.

The influence of exposure time on the measured 222Rn activity flux was observed in two

situations of the canister exposure layout: (a) canister sealed to an open bed of the material and

(b) canister sealed over a jar containing the material. The measured 222Rn activity flux

decreased as the exposure time increased. The change in the former situation was significant

with an exponential decrease as the exposure time increased. In the latter case, lesser reduction

was noticed in the observed activity flux with respect to exposure time. This reduction might

have been related to certain factors, such as absorption site saturation or the back diffusion of

222Rn gas occurring at the canister-soil interface.

Keywords: 222Rn; activity flux; activated charcoal canisters

4.1 Introduction

Radon (222Rn) is a natural radioactive gas produced in rocks and soil by the

disintegration of its parent nuclei radium (226Ra). A small fraction of the produced 222Rn can find

its way to the surface and can be released from the ground. The activity flux of 222Rn released

from the ground is commonly measured by placing a chamber (accumulator) directly into the

sampling area in an upturned position (Ferry et al., 2002; Ielsch et al., 2002; Schery et al., 1989)

or by enclosing the sample materials in an airtight container (Jonassen, 1983; Samuelsson,

1990; Tufail et al., 2000). In both situations, the exhaled 222Rn from the sample surface enter the

chamber and the concentration of 222Rn increases. This concentration can be measured using

passive or active detectors. Passive devices, such as alpha track detectors and activated charcoal

canisters do not require electrical power to function, whereas active devices, such as

88

scintillation cell and semiconductor detectors require electricity to operate. In both cases, the

flux density is then deduced using a proper formula.

Over the past few decades, a number of techniques have been developed to measure

222Rn activity flux from surfaces. A commonly and extensively used technique for measuring

222Rn utilises activated charcoal packed in canisters. It can be used to measure 222Rn

concentration in air (Cohen and Cohen, 1983; George, 1984; Scarpitta, 1992) as well as the 222Rn

activity flux as it escapes from the ground (Bollhöfer et al., 2006; Countess, 1976; Lawrence et

al., 2009; Spehr and Johnston, 1983; Wang et al., 2009) and building materials (De Jong et al.,

2005; Petropoulos et al., 2001). This technique is based on 222Rn adsorption in the charcoal,

which affords many advantages including: low cost, reusability and simplicity.

Since Countess (1976) published a paper on measuring 222Rn using an activated charcoal

canister, numerous studies have been conducted about the applications and limitations of this

method (El Samman and Arafa, 2002; Iimoto et al., 2008; Ronca-Battista and Gray, 1988; Vargas

and Ortega, 2007). These studies have shown that meteorological conditions, including

humidity, temperature and atmospheric pressure have the greatest influence on 222Rn

adsorption in activated charcoal. Most of these studies focused on the measurements of 222Rn

concentration in the air whereas only a few studies have been done on activity flux.

When 222Rn activity flux is measured, canisters are deployed on the target area and

exposed for a few days, typically between one to five days (Bollhöfer et al., 2006; Countess,

1976; Dueñas et al., 2007; Lawrence et al., 2009). The adsorbed 222Rn in the activated charcoal

is commonly measured through the gamma radiation detection of its short-lived progeny nuclei.

A sodium iodide (NaI(Tl)) detector is usually used for this purpose. The detection efficiency of

the gamma detector is determined by counting a standard canister, which is generally prepared

by adding a known amount of 226Ra to the charcoal and mixing it uniformly into the charcoal bed

(Bollhöfer et al., 2006; Spehr and Johnston, 1983). This may not always reflect the actual

situation as the distribution of the adsorbed 222Rn in the measured canister may not necessarily

have the same distribution as in the standard canister.

The 222Rn adsorption rate in this method is assumed to be constant over the exposure

period. Although this assumption may not be always valid in the real world, it has been widely

accepted for this method. This paper aims to visit two aspects of the current methodology of

measuring 222Rn activity flux using activated charcoal canisters. The first is the distribution of

222Rn in the charcoal bed. The second is the influence of exposure time on the measured value of

222Rn activity flux.

89

4.2 Methods

A total of 293 individual measurements of the 222Rn activity flux were carried out for five

222Rn sources (Table 4.1). Of the canisters, 180 were used to examine the distribution of 222Rn in

the charcoal bed and 77 of them were used to investigate how the time influenced the measured

222Rn activity flux. All measurements were made under a controlled environment in the

laboratory. The relative humidity and temperature were 40%–60% and 22oC–25oC,

respectively. The canisters were cylindrical with a height of 0.060 m, a diameter of 0.070 m and

an open face area of 0.0038 m2. Each canister was filled with 25g of activated charcoal (Chem-

Supply, Australia). Prior to use, the canisters were annealed for 24 hours at 110oC to drive off

any residual 222Rn in the charcoal and then sealed with a polyethylene lid and adhesive tape.

4.2.1 The distribution of 222Rn in the charcoal bed

The charcoal bed in each canister was divided into six layers that could be sectioned and

counted separately for estimating the vertical distribution of the adsorbed 222Rn. Each layer was

filled with 5g of activated charcoal providing about 5 mm of layer thickness separated by a

metal mesh and secured with a wire spring. The canisters were exposed to different222Rn

sources for one, three, seven, 10, and 14 days. At the end of the sampling time, each layer in the

canister was transferred into a separate container and measured. The reason for counting each

layer separately was to maintain the same geometry throughout. The counting system used for

this purpose was a 2" × 2" NaI (Tl) detector. This setup is further described in Section 4.2.3.

4.2.2 The influence of exposure time on the measured activity flux

Activated charcoal canisters are integrating devices. When 222Rn activity flux is

measured, the obtained values actually represent the average activity flux over the exposure

time. In this procedure, the measured activity flux from the surfaces was examined as a function

of the exposure time. Two sets of experiments were performed, in the first, canisters were

placed directly onto the sources that were hosted on a tray. The tray had a diameter of 40 cm

and a depth of 5 cm and could host seven canisters simultaneously. Canisters were pressed to a

depth of 1 cm into the material to prevent gas leakage. In the second set, canisters were firmly

sealed over sources hosted in polystyrene cylindrical containers of 10 cm height and 6.3 cm in

diameter. The thickness of the materials inside the container was 8 cm. The open-tray method

may have more accurately represented field situation, but the material depth was only 3 cm.

Canisters in both sets were exposed for one, two, three, five, seven, 10, and 14 days.

4.2.3 Analytical technique

After the exposure, canisters were resealed and given four hours to allow the ingrowth

of 222Rn progenies 214Pb and 214Bi. The activity of the adsorbed 222Rn on the charcoal was

determined by detecting gamma rays that were emitted from its short-lived progenies 214Pb and

90

214Bi using a 2" × 2" NaI (Tl) detector. Detector efficiency was 11% at the selected region of

interest (143 - 740 keV), which covered the salient peaks 214Pb at 242, 295, and 352 keV, as well

as 214Bi at 609 keV. The detector was housed in a lead castle to reduce background counts which

were obtained by counting an unexposed canister. The activity flux of 222Rn was inferred using

the following formula (Spehr and Johnston, 1983):

𝐽 =𝑅. 𝜆2. 𝑡𝑐 . 𝑒𝜆.𝑡𝑑

𝜀. 𝑎. (1 − 𝑒−𝜆.𝑡𝑐). (1 − 𝑒−𝜆.𝑡𝑒) (4.1)

where J is the measured 222Rn activity flux (Bq m−2 s−1), R is the count rate (s-1) after background

subtraction, λ is the decay constant (s-1) for 222Rn, ɛ is the efficiency of the NaI(Tl) detector

(cps/Bq), a is the surface area covered by the charcoal canister (m2), tc is the counting time, te is

the exposure time, and td is the delay time between the end of the exposure until the start of the

counting. For uniformity in units, tc, te and td were all measured in seconds. This formula is

based on two basic assumptions. The first assumption is that all of the 222Rn that enters the

canister is adsorbed by the charcoal and retained in it. The second assumption is that the 222Rn

activity flux from the surface is constant over the exposure period.

The results were given relative to the measured values of the activity flux after one day

of exposure. Equation (1), in that case, led to a relationship between the measured activity flux

after one day (J1) and n days (Jn) as

𝐽𝑛

𝐽1=

𝑅𝑛(1 − 𝑒−𝜆.𝑡𝑒1)

𝑅1(1 − 𝑒−𝜆.𝑡𝑒𝑛) (4.2)

where R1 and Rn are the count rates after the background subtraction (s-1) for one and n days of

exposure, respectively and te1 and ten are the exact exposure times for nominally one and n days

of exposure (s). Equation (2) assumed that the counting times tc1 and tcn were equal—1000

seconds during the present work. The delay times td1and tdn for one and n days of exposure

were the same for each sample (about four hours). Reported uncertainties represent one

standard deviation calculated from the statistics of counting.

In addition to the activated charcoal canisters and gamma count rate, an independent

method was also used to measure 222Rn activity flux from the five sources. This method was

based on the use of a solid-state alpha detector incorporated into two RAD7 systems

(DURRIDGE, 2000). The results of this independent method (Table 4.1) were compared with

those obtained from the activated charcoal canisters.

Table 4.1 Rn-222 emitting samples used in the present study.

Sample ID Description Method of use Expected activity flux(mBq m−2 s−1)a O – 1 Rock Closed-can 4528 ± 94 O - 24 Rock Closed-can 1427 ± 112 O - 31 Rock Closed-can 1551 ± 106 T Uranium tailing Closed-can, Open-tray 43 ± 1, 14.7 ± 0.1 B Uranium tailing Closed-can, Open-tray 313 ± 98, 102 ± 49 a: Measured using a calibrated RAD7 device.

91

4.3 Results and Discussion

4.3.1 The distribution of 222Rn in the charcoal bed

4.3.1.1 The uniformity of 222Rn in the charcoal bed

As mentioned in Section 4.2.1, after the exposure different 5 mm depth sections of the

total 30 mm (30 g) charcoal bed were removed from the canister. Each section was then

separately counted under the same source-detector geometry. The distribution of 222Rn in the

charcoal bed obtained is shown in Figure 4.1. The results were for five different exposure times

covering the range from one to 14 days. All results have been plotted relative to the value in the

section closest to the 222Rn source, which was furthest from the canister base as they have been

exposed in an inverted position. This experiment was performed using four sources of 222Rn

covering activity flux ranging from nearly 45 to 4500 mBq m−2 s−1. The values in Figure 4.1 were

the average for all sources. However, the level of the activity flux within our experimental range

appeared not to affect the relative distribution of 222Rn in the charcoal bed.

The distribution of 222Rn in the charcoal bed appeared less uniform at short exposure

periods. For one day of exposure, the closest layer to the canister base had 27 ± 1 % of the

amount of 222Rn that was in the layer furthest from the canister base. This difference became

smaller with the increase of exposure time. Most likely, as the exposure time increased more

222Rn nuclei were infused gradually into the deeper layers of the charcoal bed. Our findings

were supported by Uroševic et al. (1999) who theoretically simulated the time dependency of

the adsorbed 222Rn activity in the thickness of the charcoal layer. They found that the adsorbed

222Rn activity in the charcoal approached equilibrium as time proceeded.

The detection efficiency of the gamma detector might have been affected by the

distribution of 222Rn in the charcoal bed, especially for short exposure times, where the degree

of uniformity was low. This is because the geometry of 222Rn in the measured canister may not

be similar to that of the standard canister, which has a uniform distribution of 222Rn across the

charcoal bed. For one day of exposure, most of the 222Rn accumulated in the surface layer that

faced the source (Figure 4.1). However, after the exposure and during the counting period this

layer became the furthest from the detector and was therefore counted with less efficiency.

92

Figure 4.1 The distribution of 222Rn in the charcoal bed for different exposure times. The correction factors corresponding to the exposure time are shown in parentheses.

4.3.1.2 Detector efficiency at different distance

The effect of the low degree of uniformity of 222Rn in the charcoal bed on the detection

efficiency required investigation to obtain a realistic value of the activity flux. For this purpose,

detector efficiency at different heights was first checked using two certified point sources of

133Ba and 137Cs in the activity proportion of 0.64:1.00 housed in an empty canister. This

combination gave a spectrum that was comparable to our canisters within our region of interest

(143–740 keV).

Detector efficiency, as a function of distance from the canister base and the data of

Figure 4.1, had been used to work out a correction factor between a uniform distribution of

222Rn in the calibration standard and that which actually occurs due to 222Rn infusion during

exposure. The correction factor was obtained by the differences in the counting geometries

between the uniformly distributed standards and measured canisters. This correction factor is

shown in the legend of Figure 4.1. Its numerical value was 1.14 and 1.05 for exposure periods of

one day and three days, respectively, and then it levelled off to1.02 for 14 days.

93

Interestingly, the phenomenon of the reduction in detection efficiency as a function of

sample height above the detector and the adsorption of 222Rn in the charcoal bed tended to

counterbalance each other, at least for the setup we used. Consequently, the correction factor

that was due to the difference between the 222Rn distribution in the standard and exposed

canister for different days became smaller. The highest value of the correction factor was for the

first day of exposure due to the lower degree of uniformity of the 222Rn distribution in the

charcoal bed in this period as indicated in Figure 4.1. A correction factor of this magnitude is

likely to be well within other sources of uncertainty in the 222Rn measurement.

The numerical values of the activity flux given by Equation (1) should be divided by this

correction factor to obtain the expected value. Care should be exercised to generalise the use of

the correction factor given in this report because its value may change with the detector

characteristics, shape of the canister, and the properties of the activated charcoal material.

4.3.1.3 Canisters with different depth of activated charcoal

A number of charcoal canisters are simultaneously deployed during most applications of

activity flux measurements. Some differences in the material quantity may exist between these

cups. Differing amounts of activated charcoal in a canister corresponds to a different geometry,

which may require a correction factor. On the basis of the results presented in Figure 4.1, this

factor is expected to be negligible when the difference in the amount is small. The following

experiment was carried out to observe the effect of a dramatic difference of charcoal amount in

the cups.

A set of canisters were filled with five, 10, 15, 20, 25, and 30 g of activated charcoal

corresponding to bed depths of 0.5, 1.0, 1.5, 2.0, 2.5, and 3.0 cm, respectively. Five such sets

were prepared and exposed to different 222Rn sources as described in Table 4.1. This

experiment was, however, limited to one day exposure time which, as shown in Figure 4.1, gave

the most dramatic differences in the adsorbed 222Rn across the charcoal layers. The results

given in Table 4.2 were normalised to 2.5 cm charcoal bed depth because this is the depth we

normally use. Variation with respect to charcoal amount was significant, but little variations

were observed with respect to the 222Rn source strength within the investigated range of the

activity flux (45 - 4500 mBq m−2 s−1).

In Table 4.2, the charcoal amount correction factor is a value by which the activity flux

given in Equation (1) should be divided if charcoal depth is different from the calibration

reference (2.5 cm in our case). This correction factor was for one day exposure in the field. In

light of the observations given in Figure 4.1, the correction factors for large exposure times are

expected to be lesser in magnitude. As shown in Table 2, the amount of the adsorbed 222Rn in

the 0.5 cm layer was double that of the 2.5 cm layer. This indicates that the thin layer of charcoal

94

bed is counted with higher detection efficiency. Scarpitta (1996) investigated the dependence of

222Rn adsorption and desorption as a function of the depth of an activated charcoal bed.

Scarpitta (1996) reported that the thin layer of charcoal bed has higher 222Rn adsorption

coefficient. However in our case, the difference in the counting geometry as a function of

charcoal depth might also have played a role.

Table 4.2 Correction factors of 222Rn activity flux for different charcoal depths in a canister with respect to that of 2.5 cm depth for one day exposure time.

Activated charcoal mass Charcoal bed depth (x) Correction factor (g) (cm) 5 0.5 2.1 ± 0.3 10 1.0 1.6 ± 0.2 15 1.5 1.4 ± 0.2 20 2.0 1.2 ± 0.1 25 2.5 1.0 30 3.0 0.9 ± 0.2

4.3.2 The influence of exposure time on the measured 222Rn activity flux

The activity flux of 222Rn over the open-tray of the material was measured for different

exposure times using two independent 222Rn measurement systems: RAD7 detector and

activated charcoal canisters filled to 2.5 cm depth and exposed for one, two, three, five, seven,

10, and 14 days. The results of this experiment were plotted in Figure 4.2. These results were

corrected for the uniformity distribution of 222Rn in the charcoal bed using the correction

factors given in Figure 4.1. For the purpose of comparison, the measured values of the 222Rn

activity flux have been normalised to the value obtained at one day of exposure. The measured

activity flux using activated charcoal canisters appeared to decrease as the exposure time

increased. The value after 14 days of exposure was about half of that obtained after one day. The

relationship between the measured activity flux and the exposure time was

Jn /J1= (1.00 ± 0.03). e− (0.05 ± 0.01) n−1 where n≥1

where J1 and Jn represented the 222Rn activity flux after one day and n days of exposure time,

respectively. The data fit was obtained with R2 = 0.88.

95

Figure 4.2 Normalised 222Rn activity flux in open-trays plotted against exposure time.

Through a broad comparison with the results of RAD7, the measured 222Rn activity flux

appeared to agree with those obtained using activated charcoal canisters for exposure periods

between one to seven days, given an average of around four days as may be seen in Figure 4.2.

There might be multiple factors involved in this situation, causing the difference

between the two methods. The measurements of the 222Rn activity flux were not carried out

simultaneously with the charcoal canisters and the RAD7 detector. Although the measurements

were conducted under laboratory conditions, temperature and humidity may have influenced

the results for the two measured systems. The effect of temperature and humidity on the 222Rn

activity flux measurement methods had been reported by a number of authors (Iimoto et al.,

2008; Jha et al., 2000; Lawrence et al., 2009).

Interestingly, the reduction in the activity flux with respect to canister exposure time

was less when the samples were held in an airtight can (closed-can method as described in

Section 4.2.2) Figure 4.3. The values of the 222Rn activity flux that were measured after 14 days

of exposure remained at 80% or more of those obtained after one day of exposure. In Figure

4.3, the 222Rn activity flux that was measured with activated charcoal canisters was normalised

against the values obtained after one day of exposure. The influence of the exposure time on the

96

measured activity flux was small, as statistically indicated by the nearly flat slopes of the linear

fit of the activity flux versus exposure time in Figure 4.3. The observed values demonstrated

that not all decreases in the measured activity flux with exposure time were statistically

significant. Comparing the results with the values obtained by the RAD7, the values of the

activated charcoal agree within 25 % of those from RAD7 throughout the exposure time Figure

4.4.

Figure 4.3 The measured activity flux versus exposure time for charcoal canister fitted over cans containing samples.

97

Figure 4.4 The measured activity flux of the samples in the closed-can relative to RAD7 values.

To explore the influence of exposure time on the measurement of the 222Rn activity flux,

count rates were plotted versus the exposure time for the closed-can and open-tray samples in

Figure 4.5. Assuming continued and constant 222Rn activity fluxes from the sample and no loss of

222Rn by other processes than radioactive decay results in a trend of the normalised count rate

in the charcoal canisters shown by the dashed line in Figure 4.5 and described by equation (4.3).

𝑅𝑛

𝑅1=

(1 − 𝑒−𝜆.𝑡𝑒𝑛)

(1 − 𝑒−𝜆.𝑡𝑒1) (4.3)

The observed values in Figure 4.5 tended to level off well below this trend, particularly

for the open-tray samples. For the closed-can, it may be so that the 222Rn activity concentration

in the air gap between the sample and the charcoal could have kept on building; therefore the

increase in the count rates are closer to the expected values. On the other hand, the build-up of

222Rn initially continued in the same fashion under the canister in the open-trays. However,

after three days or so of exposure, the build-up did not occur as expected; indicating the loss of

222Rn from the canister due to a process other than radioactive decay.

We suggested that the divergence between the open-tray and closed-can behaviours

was related to the differences in the design of these two methods. Placing the canister directly

98

to the sample surface did not isolate it completely from the ambient environment because the

horizontal extent of the sample material underneath the canister was in direct contact with the

atmosphere. Therefore, a back diffusion of 222Rn from the canisters into the soil gas below is a

possibility. However, the 222Rn concentration in the charcoal canister grew at a lower than

expected rate, hence only a portion of 222Rn might have diffused back into the soil gas

underneath and possibly into the air.

The possibility of such back diffusion was also proposed by Mayya (2004) who analysed

the effect of the presence of an inverted cup on the diffusion of 222Rn from the soil underneath in

a two-dimensional framework. Mayya’s study demonstrated that the use of an inverted cup can

underestimate the measured activity flux due to the back diffusion phenomena that occurs at

the soil-cup interface. According to Mayya (2004), the underestimation of the activity flux

would require a correction factor that depends on the time period of the measurement,

dimensions of the cup and soil properties. The effect of back diffusion on the measured 222Rn

activity flux had also been mentioned in a number of other radon detectors, such as RAD7

(Tuccimei et al., 2006), Lucas cell (Aldenkamp et al., 1992) and alpha track detectors (Faheem,

2008).

It should be noted that the presence of activated charcoal in the canisters may mitigate,

but not eliminate the effect of back diffusion due to the ability of the charcoal to adsorb 222Rn

gas. In this case, another factor, adsorption of moisture by activated charcoal, might have also

contributed to the observed reduction in the measured 222Rn activity flux with exposure time in

the open-tray method. The adsorption of moisture by activated charcoal is known to reduce

222Rn adsorption in the charcoal (Ronca-Battista and Gray, 1988; Vargas and Ortega, 2007).

Water molecules might have been picked up from the ambient air through the sample material

during the exposure. This could explain the slight reduction in the measured activity flux in the

closed-can set. This is because enclosing the sample with charcoal canister in an airtight

container can prevent water molecules from the ambient air getting inside the container during

the exposure period. Therefore, the adsorption of 222Rn in the charcoal bed remained close to

the expected value as shown in Figure 4.5.

99

Figure 4.5 Relative count rate in the charcoal canisters versus the exposure time of samples taken in a closed-can or open-tray. The dashed line represents a situation referred by Equation 4.3.

4.4 Conclusion

The method of measuring 222Rn activity flux using activated charcoal canisters was

revisited through lab measurements using samples with activity flux ranging between 45 and

4500 mBq m-2 s-1. Two aspects were studied: (a) the distribution of 222Rn in the charcoal bed

and (b) the influence of the exposure time on the measured activity flux. The distribution of

222Rn in the charcoal bed was found to become more uniform with longer exposure times. A

correction factor was introduced by the differences in the counting geometries between the

uniformly distributed standards and measured canisters with non-uniform distribution. Its

value for one day of exposure, which recorded the lowest degree of uniformity, was small at

14 % or less. This was because the lower uniformity due to the diffusion of 222Rn in the charcoal

bed and the detection efficiency as a function of the charcoal depth appeared to counterbalance

each other. The values of the correction factors in this study may vary somewhat depending on

the type of detection system used and the shape of the charcoal canister.

100

The influence of the exposure time on the measured 222Rn activity flux using activated

charcoal canisters was observed in two different sets of experiments: (a) canisters introduced

to an open bed of the material and (b) sealed over a jar containing the material. The measured

activity flux decreased as the exposure time increased from one to 14 days. The reduction was

significant in the open-bed samples, while it was slight in canisters sealed over the jars

containing the same material. The reduction in the measured activity flux with exposure time

may be attributed to the back diffusion of 222Rn from the canister through canister-soil interface

and possibly to the adsorption of water molecules by the activated charcoal. The measurement

of the activity flux over the time period between one to seven days appeared to agree with the

values obtained using the RAD7 detector.

The activated charcoal canister is a useful method for routine measurement of 222Rn

activity flux. The statistical uncertainty due to counting is not a limiting factor in obtaining

results. Even for our lowest strength sample (about 45 mBq m-2 s-1) and for exposure time as

short as 1 day, one standard deviation uncertainty due to statistics of counting has been 10%.

However, limitations of the methodology are due to other factors, some of which have been

discussed in this paper and summarised above.

Acknowledgements

Sami Alharbi is grateful for the financial support provided by King Saud bin Abdulaziz

University for Health Sciences during PhD study. Many thanks go to Kazuyuki Hosokawa for

laboratory technical support.

101

References

Aldenkamp, FJ, RJ de Meijer, LW Put and P. Stoop. 1992. "An assessment of in situ radon exhalation measurements, and the relation between free and bound exhalation rates." Radiation Protection Dosimetry 45 (1-4): 449-453.

Bollhöfer, A., J. Storm, P. Martin and S. Tims. 2006. "Geographic variability in radon exhalation at a rehabilitated uranium mine in the Northern Territory, Australia." Environmental Monitoring and Assessment 114 (1): 313-330.

Cohen, B.L. and E.S. Cohen. 1983. "Theory and practice of radon monitoring with charcoal adsorption." Health Physics 45 (2): 501-508.

Countess, R.J. 1976. "222Rn flux measurement with a charcoal canister." Health Physics 31 (5): 455-456.

De Jong, P., W. Van Dijk, W. De Vries, E.R. van der Graaf and L.M.M. Roelofs. 2005. "Interlaboratory comparison of three methods for the determination of the radon exhalation rate of building materials." Health Physics 88 (1): 59-64.

Dueñas, C., E. Liger, S. Cañete, M. Pérez and JP Bolívar. 2007. "Exhalation of 222Rn from phosphogypsum piles located at the Southwest of Spain." Journal of Environmental Radioactivity 95 (2): 63-74.

DURRIDGE. 2000. RAD7 radon detector-User manual: DURRIDGE Co, USA. El Samman, H. and W. Arafa. 2002. "Temperature and humididty consideration for calculating

airborne 222Rn using activated charcoal canisters." Health Physics 83 (1): 97-104. Faheem, M. 2008. "Radon exhalation and its dependence on moisture content from samples of

soil and building materials." Radiation Measurements 43 (8): 1458-1462. Ferry, Cécile, Patrick Richon, Alain Beneito and MarieChristine Robé. 2002. "Evaluation of the

effect of a cover layer on radon exhalation from uranium mill tailings: transient radon flux analysis." Journal of Environmental Radioactivity 63 (1): 49-64.

George, A.C. 1984. "Passive, integrated measurement of indoor radon using activated carbon." Health Physics 46 (4): 867-872.

Ielsch, G, C Ferry, G Tymen and MC Robé. 2002. "Study of a predictive methodology for quantification and mapping of the radon-222 exhalation rate." Journal of Environmental Radioactivity 63 (1): 15-33.

Iimoto, T., Y. Akasaka, Y. Koike and T. Kosako. 2008. "Development of a technique for the measurement of the radon exhalation rate using an activated charcoal collector." Journal of Environmental Radioactivity 99 (4): 587-595.

Jha, S., AH Khan and UC Mishra. 2000. "A study of the 222Rn flux from soil in the U mineralised belt at Jaduguda." Journal of Environmental Radioactivity 49 (2): 157-169.

Jonassen, N. 1983. "The determination of radon exhalation rates." Health Physics 45 (2): 369-376.

Lawrence, C.E., R.A. Akber, A. Bollhöfer and P. Martin. 2009. "Radon-222 exhalation from open ground on and around a uranium mine in the wet-dry tropics." Journal of Environmental Radioactivity 100 (1): 1-8.

Mayya, YS. 2004. "Theory of radon exhalation into accumulators placed at the soil–atmosphere interface." Radiation Protection Dosimetry 111 (3): 305-318.

Petropoulos, NP, MJ Anagnostakis and SE Simopoulos. 2001. "Building materials radon exhalation rate: ERRICCA intercomparison exercise results." The Science of the Total Environment 272 (1-3): 109-118.

Ronca-Battista, M. and D. Gray. 1988. "The influence of changing exposure conditions on measurements of radon concentrations with the charcoal adsorption technique." Radiation Protection Dosimetry 24 (1-4): 361-365.

Samuelsson, C. 1990. "The closed-can exhalation method for measuring radon." Journal of Research of the National Institute of Standards and Technology 95 (2): 167-169.

Scarpitta, S.C. 1992. "Optimum 222Rn-adsorbing activated charcoals." Health Physics 62 (6): 576-580.

102

Scarpitta, S.C. 1996. "A new beaded carbon molecular sieve sorbent for 222Rn monitoring." Health Physics 70 (5): 673-679.

Schery, SD, S Whittlestone, KP Hart and SE Hill. 1989. "The flux of radon and thoron from Australian soils." Journal of Geophysical Research 94 (D6): 8567-8576.

Spehr, W. and A. Johnston. 1983. "The measurement of radon exhalation rates using activated charcoal." Radiation Protection in Australia 1 (3): 113-116.

Tuccimei, P., M. Moroni and D. Norcia. 2006. "Simultaneous determination of 222Rn and 220Rn exhalation rates from building materials used in Central Italy with accumulation chambers and a continuous solid state alpha detector: Influence of particle size, humidity and precursors concentration." Applied Radiation and Isotopes 64 (2): 254-263.

Tufail, M, Sikander M Mirza, Arif Mahmood, AA Qureshi, Yasir Arfat and HA Khan. 2000. "Application of a closed-can technique for measuring radon exhalation from mine samples of Punjab, Pakistan." Journal of Environmental Radioactivity 50 (3): 267-275.

Uroševic, V., D. Nikezic, S. Vulovic and M. Kojic. 1999. "Optimization of radon measurements with active charcoal." Health Physics 76 (6): 687-691.

Vargas, A. and X. Ortega. 2007. "Influence of environmental changes on integrating radon detectors: results of an intercomparison exercise." Radiation Protection Dosimetry 123 (4): 529-536.

Wang, N., L. Xiao, C. Li, W. Mei, Y. Hang and D. Liu. 2009. "Level of radon exhalation rate from soil in some sedimentary and granite areas in China." Journal of Nuclear Science and Technology 46 (3): 303-309.

103

Chapter 5 Radon and thoron concentrations in public

workplaces in Brisbane, Australia

Sami H. Alharbi*, Riaz A. Akber

*School of Physical and Chemical Sciences, Queensland University of Technology, 2 George

Street, Brisbane, Qld 4000, Australia

Journal of Environmental Radioactivity 144 (2015), 69-76.

http://dx.doi.org/10.1016/j.jenvrad.2015.03.008

104

Statement of Contribution of Co-Authors for Thesis by Published Paper The authors listed below have certified that:

1. they meet the criteria for authorship in that they have participated in the conception,

execution, or interpretation, of at least that part of the publication in their field of

expertise;

2. they take public responsibility for their part of the publication, except for the

responsible author who accepts overall responsibility for the publication;

3. there are no other authors of the publication according to these criteria;

4. potential conflicts of interest have been disclosed to (a) granting bodies, (b) the editor

or publisher of journals or other publications, and (c) the head of the responsible

academic unit, and

5. they agree to the use of the publication in the student’s thesis and its publication on the

Australasian Research Online database consistent with any limitations set by publisher

requirements.

In the case of this chapter:

Radon and thoron concentrations in public workplaces in Brisbane, Australia

Contributor Statement of contribution

Sami Alharbi

Signature:

Date: 31/03/2016

Identified concept for the research paper,

conducted experimental work, analysed and interpreted the data,

wrote the manuscript

Riaz Akber Assisted with editorial process, aided in concept development, and

reviewed the manuscript.

Principal Supervisor Confirmation

I have sighted email or other correspondence from all Co-authors confirming their certifying

authorship.

Name:

Signature:

Date:

105

Abstract

Radon and thoron are radioactive gases that can emanate from soil and building

materials, and it can accumulate in indoor environments. The concentrations of radon and

thoron in the air from various workplace categories in Brisbane, Australia were measured using

an active method. The average radon and thoron concentrations for all workplace categories

were 10.5 ± 11.3 and 8.2 ± 1.4 Bq m-3, respectively. The highest radon concentration was

detected in a confined area, 86.6 ± 6.0 Bq m-3, while the maximum thoron level was found in a

storage room, 78.1 ± 14.0 Bq m-3. At each site, the concentrations of radon and thoron were

measured at two heights, 5 cm and 120 cm above the floor. The effect of the measurement

heights on the concentration level was significant in the case of thoron. The monitoring of

radon and thoron concentrations showed a lower radon concentration during work hours than

at other times of the day. This can be attributed to the ventilation systems, including the air

conditioner and natural ventilation, which normally operate during work hours. The diurnal

variation was less observed in the case of thoron, as the change in its concentration during and

after the working hours was insignificant. The study also investigated the influence of the floor

level and flooring type on indoor radon and thoron concentrations. The elevated levels of radon

and thoron were largely found in basements and ground floor levels and in rooms with concrete

flooring.

Keywords: Radon; Thoron; Workplaces; Indoor

5.1 Introduction

Radon (222Rn) and thoron (220Rn) are radioactive gases formed by the decay of radium

isotopes within the 238U and 232Th series. Radon and thoron emanate from the soil, rocks and

construction materials, and tend to accumulate in enclosed spaces, such as buildings, tunnels

and underground mines. The inhalation of radon and thoron gases, as well as their progenies, is

recognised as a potential cause of lung cancer (WHO, 2009) and considered to be the major

contributor to public exposure from natural sources (UNSCEAR, 2000).

Indoor radon and thoron are of particular concern due to the possibility of their

concentrations being elevated to a considerable level. In some environments, indoor radon was

found to be greater than outdoors by a factor of up to 50 (Porstendorfer et al., 1978; Song et al.,

2005). In general, in cities such as Brisbane, Australia, people tend to spend more time indoors,

whether in dwellings or workplaces, increasing their exposure to indoor sources. A

considerable amount of time is often spent at the workplace, and exposure due to radon and

thoron in some workplaces can be significant (Gillmore et al., 2000). Fortunately, exposure due

to radon and thoron in indoor environments is feasible to control (IAEA, 2003; ICRP, 1993,

2014; UNSCEAR, 2006).

106

Radon and thoron concentrations can be measured using passive or active methods

(Alharbi and Akber, 2014; IAEA, 2013; Laiolo et al., 2012). Passive detectors integrate the

average radon and thoron concentrations over one season or longer while active detectors

provide continuous radon and thoron monitoring during the day. However, workplaces are only

occupied for a certain time period on specific days of the week; thus, the average concentration

over the working hours may differ from that obtained over the total time. In this case, using

passive detectors may not be appropriate, as most of these detectors are incapable of

monitoring radon and thoron concentrations selectively during work hours.

Generally, the levels of indoor radon and thoron vary widely in time and space

(Bochicchio et al., 2005; Faheem and Mati, 2007; Prasad et al., 2012; Singh et al., 2005), being

higher in the winter and lower in the summer. This seasonal pattern is mainly attributed to

ventilation systems, including air conditioning and heating systems, which play essential roles

in changing radon levels in indoor environments (Lee and Yu, 2000; Marley and Phillips, 2001).

A correction factor for such variations has been proposed (Baysson et al., 2003; Faheem and

Mati, 2007; Pinel et al., 1995); alternatively, the measurement may be taken only during the

coldest months, as recommended by some authorities (IAEA, 2003; NORDIC, 2000). The

seasonal trend of indoor radon, however, has been less observed in workplaces, and some have

exhibited no seasonal patterns (Inoue et al., 2013; Papachristodoulou et al., 2010; Tokonami et

al., 1996). Instead, the diurnal variation has been found more often in workplaces (Vaupotič et

al., 2012; Zahorowski et al., 1998).

Another factor that can affect the level of indoor radon is the underlying geology. A

significant correlation has been found between the level of indoor radon and the geological

composition of the ground (Moreno et al., 2008; Sundal et al., 2004). Radon in the soil gas can

enter buildings through cracks in the floor by way of a pressure-driven mechanism. The

relatively long half-life of radon (3.8 d) gives it enough time to travel, which can be clearly seen

in multi-story buildings where radon levels were found to be higher in the basement and on the

ground level (Papachristodoulou et al., 2010; Shaikh et al., 2003). The influence of the

underlying geology is less pronounced in indoor thoron, possibly due to its short half-life (56

seconds). Instead, thoron is more influenced by the distance from the source, which includes the

walls and floors in buildings (Fujimoto et al., 1994; Urosevic et al., 2008).

There are other parameters that might influence indoor radon and thoron levels, but

with less significance. These include meteorological conditions (Miles, 2001), the buildings’

characteristics (Gunby et al., 1993) and the occupants' habits1. These factors can contribute in

different degrees to the radon and thoron variations in indoor environments.

1 Such as keeping opening windows and doors, introducing floor and wall covering using fans or air conditioning. etc.) —Note: footnotes were added to the thesis, not present in the published paper.

107

During the past few years, there has been an increasing awareness of the exposure of

radon and thoron in workplaces other than mines (IAEA, 2003, 2011; UNSCEAR, 2006; WHO,

2009). These workplaces are often located underground, such as in subways and tunnels, or in

aboveground places, such as offices, schools, shops and factories. Mines that are open to tourists

are also included in these areas. Workers and members of the public may be adventitiously

exposed to radon and thoron in these workplaces. Recently, the ICRP has recommended a

reference level in the range of 100 to 300 Bqm-3 for radon concentration in buildings without

distinguishing between dwellings, workplaces and mixed-use buildings (ICRP, 2014).

A broad spectrum of workplaces have been surveyed worldwide for indoor radon

concentrations (Inoue et al., 2013; Martín Sánchez et al., 2012; Solomon et al., 1996; Vaupotič et

al., 2012). In Spain, a wide range of workplaces was measured, including offices, shops, spas,

caves, museums, etc. (Martín Sánchez et al., 2012), and the average radon concentration

exceeded 400 Bq m-3 in 16% of the measured workplaces. In a study conducted in Greece by

Clouvas et al. (2007), the indoor radon concentrations in schools were found to be higher than

in other workplaces.

On the other hand, the available data on indoor thoron is largely limited, especially for

workplaces (Burghele and Cosma, 2012; Inoue et al., 2013; Vaupotič et al., 2012). This is

perhaps due to the measurement difficulties associated with its short half-life; additionally,

there is no desire to regulate indoor thoron in most countries. It is worth noting that the health

risks associated with the inhalation of thoron and its decay products are considered to be equal

if not greater than those of radon and its progeny (ICRP, 1993; Shang et al., 2005; UNSCEAR,

2000). Moreover, the concentration of thoron in some indoor spaces was found to be four times

greater than that of radon (Ramola et al., 2012; Yamada et al., 2006).

In Australia, a very limited number of workplaces, other than those for which the

exposure is controlled as a planned exposure situation, were surveyed for indoor radon

concentrations (Solomon et al., 1992; Solomon et al., 1996; Zahorowski et al., 1998). A radon

level of 6330 Bq m-3 was detected in Australian show caves, giving off a yearly effective dose of

up to 5.9 mSv for tour guides (Solomon et al., 1996). An action level of 200 and 1,000 Bq m-3 of

radon in dwellings and workplaces, respectively, was proposed by the Australian Radiation

Protection and Nuclear Safety Agency (ARPANSA) based on the Australian National Health and

Medical Research Council (NHMRC) recommendation, jointly with the National Occupational

Health and Safety Commission (NOHSC) (NHMRC, 1995).

Universities are educational premises where a wide range of workplaces can be found with

a high occupancy factor for members of the public. Radon and thoron levels in universities may

pose potential long-term health risks to the full-time staff and students who may extend their

work until late, to conduct research. This paper aims to assess the levels of indoor radon and

108

thoron across a wide spectrum of workplace categories at the Queensland University of

Technology.

5.2 Materials and Methods

5.2.1 Sampling site

The measurements of indoor radon and thoron were conducted at different sites at the

Queensland University of Technology (QUT) Gardens Point campus. The campus is located in

the central business district of Brisbane city, the state capital of Queensland, Australia, at a

latitude of 27°29ʹ South and a longitude of 153°2ʹ East (Figure 5.1). The campus consists of a

wide range of multi-story buildings designed for mixed-use purposes. These buildings are

constructed of concrete, brick, metal and cement, and they are occupied by students, staff

members and visitors during the daytime. Currently, the University has 4,349 staff members

including academic, technician and administrative staff. All buildings are air-conditioned to

adjust the temperature in the summer and winter during work hours.

The study area is underlain by basement rocks that consist mainly of meta-sedimentary

and metamorphic formation covered by pale, shallow and stony soil (Willmott et al., 2012).

Brisbane has a subtropical climate and the rain is distributed between the summer and autumn

months.

The sampling sites have been chosen to cover a broad range of workplace categories.

The examined workplaces in this study include offices, classrooms, research rooms,

laboratories, a library, main halls, confined areas, basements, storage areas and a car park.

These places are normally occupied by professional academic staff members and students.

109

5.2.2 Radon and thoron measurements

The study was conducted at the selected workplaces throughout the cooler months,

between June and November 2013. All measurements were carried out during the weekdays

(Monday through Friday) using a solid-state alpha detector RAD7 designed by Durridge, USA.

The radon and thoron concentrations were measured on several floors of the multi-

story buildings. At each sampling site, the radon and thoron levels were monitored for two days

at the ground (5 cm) and at the breathing levels (120 cm). The reason for measuring the

concentrations at two heights was to compare the variations in radon and thoron levels with the

distance2. The radon and thoron concentrations were monitored simultaneously and

continually over the examination period, and the average radon and thoron concentrations over

the work hours, as well as over the whole day, were obtained.

Aside from the measurement of radon and thoron concentrations, the data has been

handled in such a way as to investigate various parameters that may influence the radon and 2 Distance: height above the floor surface—Note: footnotes were added to the thesis, not present in the published paper.

" Brisbane

Figure 5.1. A map of the sampling site.

110

thoron concentrations in indoor environments. The effect of the floor level on the concentration

of radon and thoron has been studied. The influence of flooring type on indoor radon and

thoron levels was investigated.

The detector used in this study, Rad7, has an internal cell of a 0.7-litre volume

hemisphere with a solid-state, ion-implanted, planar, silicon alpha detector installed in the

centre. The inner side of the internal cell is coated with an electrical conductor. A high voltage is

applied across the internal cell to create an electric field, which directs positive charges toward

the centre, where the detector sits. The sampled air is pumped through a 0.45 µm filter, which

excludes radon and thoron progeny, into the internal cell with a flow rate of 650 ml/min. In the

cell, radon and thoron decay into polonium isotopes by emitting detectable alpha particles,

which in turn are converted into electrical signals. A spectrum of alpha energy in the range of 0–

10 MeV is produced and grouped into eight windows (A – H). Windows A and C are used to

determine the 222Rn level by collecting alpha energies from 218Po (6.0 MeV) and 214Po (7.69

MeV), respectively. Windows B and D are used to determine the 220Rn through 216Po (6.78 MeV)

and 212Po (8.78 MeV) levels. The rest of the windows on the detector are not applicable to our

measurements. A drying tube containing desiccant (CaSo4) is used to maintain the relative

humidity of the RAD7 below 10%, as its sensitivity is degraded at high humidity levels.

The RAD7 was set to a 3-hour cycle time to obtain better statistics. Prior to each

measurement, the instrument was purged for 10 minutes to remove any remaining radon or

thoron gas from the internal cell. During the sampling period, the values of radon and thoron

are monitored and printed out every 3 hours. The data is then transferred from the RAD7 to the

computer using the Capture software supplied by the manufacturer. The software allows the

values to be corrected for the humidity and the spillover of the counts that occur between the

adjacent windows of the RAD7.

The instrument was calibrated by the manufacturer for radon, with an overall accuracy

of about 5%; however, thoron sensitivity is considered by the manufacturer to be half of the

radon sensitivity. For verification purposes, the RAD7 was also calibrated using the radon and

thoron flow through type sources (Pylon Model Rn-1025 and TH-1025, Pylon Electronics Inc.,

Canada). The radon and thoron concentrations from the Pylon source were predicted using a

formula provided by the manufacturer. The measured radon and thoron concentrations from

the Pylon sources with the RAD7 are within 10% of the calculated concentrations (Table 5.1).

As a quality control procedure, the measurements of radon and thoron concentrations in

three locations were repeated twice in different days using the same detector and setup. There

were no significant differences between the repeated measurements of radon and thoron

concentrations (p = 0.570 and p = 0.880; respectively) in the three locations.

111

Table 5.1 Summary of the instrument calibration check using certified sources.

Isotopes Flow rate (m3 min-1) Concentration (Bq m-3) Ratio (%)

Measured Calculated

222Rn 0.001 2343 ± 158 2158 ± 86 1.09 ± 0.11

220Rn 0.001 38800 ± 1100 40457 ± 3237 0.96 ± 0.10

5.3 Results and Discussion

5.3.1 Radon and thoron concentration measurements

The average radon and thoron concentrations from each sampling site are given in Table

5.2. The averages are worked out for 5 and 120 cm heights combined for the entire two days

period of measurements. The radon concentration ranged from 0.7 to 86.6 Bq m-3, with an

arithmetic mean of 10.5 ± 11.3 Bq m-3. The maximum radon level was detected in a confined

room on the ground floor, and the minimum level was found in an office on the third floor. A

comparison of radon concentrations across all sites for both the floor level and 120 cm above

the floor is shown in Figure 5.2. A one-way analysis of variance (ANOVA) revealed a significant

difference (p < 0.05) in radon concentrations across the workplaces. A post hoc multiple

comparison test indicated that the radon concentration in basements and in the confined room

were significantly higher than in other locations.

All the values for the radon concentrations were below the reference level (200 Bq m-3)

recommended by the Australian authority for such workplaces (NHMRC, 1995). The reported

radon concentration is comparable to the average indoor radon concentration in Queensland

homes, 6.3 ± 4.4 Bq m-3 (Langroo et al., 1991), and it is lower than the typical global indoor

radon level, 40 Bq m-3, (UNSCEAR, 2000).

The level of thoron varied between 0.6 and 78.1 Bqm-3, with an arithmetic mean of 8.2 ±

3.3 Bq m-3. This average is comparable to that of radon. However, a poor correlation (R2 =

0.015) has been observed between their concentrations, as shown in Figure 5.3. Thoron

concentration was significantly higher in the basements and storage rooms. Interestingly, the

lowest thoron concentration was observed in the confined room where the radon concentration

was the highest. It is worth noting that no ventilation system was operating in the confined

room at the time of conducting the measurement. This implies that the thoron concentration

build-up is less influenced by the ventilation system. In contrast, the radon gas reached its

highest value in the confined room, and places with poor ventilation systems are generally ideal

for elevating radon levels (Alghamdi and Aleissa, 2014).

112

Table 5.2 Radon and thoron concentrations in different locations.

Workplace Number of Building Floor Radon Concentration (Bq m-3) Thoron concentration (Bq m-3) Category workplaces type* level Range Mean ± SD Range Mean ± SD

Office 7 A, B, C multiple 0.7 – 28.1 6.3 ± 3.2 2.2 – 25.5 6.2 ± 3.5 Classroom 3 B, C 3 1.7 – 23.0 7.3 ± 5.2 2.2 – 21.6 3.8 ± 1.1 Laboratory 6 A, B, C multiple 1.2 – 41.7 9.3 ± 7.1 2.2 – 15.2 4.4 ± 1.9 Library 3 C 2, 3, 5 1.1 – 10.2 3.8 ± 1.3 2.6 – 11.5 4.6 ± 1.3 Basement 2 B, C basement 3.8 – 54.9 26.5 ± 14.6 7.9 – 46.8 23.7 ± 3.9 Research room 2 C 2, 5 0.9 – 22.0 8.9 ± 3.4 2.3 – 14.0 6.1 ± 1.5 Storage 3 B, C 1, 4 1.7 – 35.6 11.1 ± 5.9 2.8 – 78.1 23.9 ± 23.9 Confined area 1 B ground 29.3 – 86.6 64.6 ± 6.0 0.6 – 8.4 2.2 ± 1.5 Car park 1 C ground 0.9 – 13.2 6.3 ± 2.7 3.0 – 19.6 10.4 ± 3.9 Main hall 1 B ground 20.8 – 63.1 35.8 ± 5.1 2.6 – 11.6 4.3 ± 2.2 All 29 0.7 – 86.6 10.5 ± 11.3 0.6 – 78.1 8.2 ± 1.4 * A : Metal cladded B : Brick veneers C : Prefabricated concrete

113

Workplace category

Office

Classroom

Library

Laborato

ry

Basement

Research R

oom

Stora

ges

Ra

do

n c

once

ntr

atio

n (

Bq

m-3

)

0

10

20

30

40

Workplace category

Office

Classroom

Library

Laborato

ry

Basement

Research R

oom

Stora

ges

Tho

ron c

once

ntr

atio

n (

Bq

m-3

)

0

20

40

60

Figure 5.2 A box plot showing radon and thoron concentrations across workplaces. The boxes represent 25–75% of the values with median concentrations represented by the lines inside the boxes. Whiskers represent the range of the distribution and the dots outside the range of 95% indicate values as outliers.

y = 0.12x + 7.02

R² = 0.015

Radon concentration (Bq m-3

)

0 20 40 60 80

Tho

ron c

once

ntr

atio

n (

Bq

m-3

)

0

20

40

60

80

Figure 5.3 Correlation between radon and thoron concentrations

The levels of radon and thoron in the present study were relatively low when compared

to other international reported values (Clouvas et al., 2007; Martín Sánchez et al., 2012).

Although the concentration of indoor radon is generally low in Australia (Langroo et al., 1991),

the ventilation system used in the workplaces is a possible cause of lowering radon levels, as

most examined workplaces are air-conditioned. The operation of the air conditioner can

increase the air exchange between the indoor and outdoor air through the pressure driven and

hence reduce radon levels (Lee and Yu, 2000). The location of Brisbane can be another reason,

as coastal sites have lower radon concentrations than inland sites (Arnold et al., 2009; Crawford

et al., 2011). This is because of the poor radon concentrations in the air coming from the ocean.

Additionally, the radium content, the progenitor of radon, in Australian building materials was

found to be lower than in many other places of the world (Beretka and Mathew, 1985). These

114

factors might have contributed in varying degrees to the overall radon and thoron levels in the

workplaces investigated in this study.

Among all the examined sites, the basement showed a significant increase (p < 0.05) in

both radon and thoron concentrations when compared to other locations (Figure 5.2). The

median radon and thoron concentrations in the basements were 34.5 and 22.1 Bq m-3,

respectively. Due to the close proximity of basements to the ground underneath, which

represents the main source of indoor radon, radon in the soil gas can easily enter through

cracks in the floor of the basement. Furthermore, the floors of basements are made of concrete

and generally not covered by carpets or any other materials that can disturb radon flux from the

cracks. Another possible reason is the lower air exchange in the basements, where access is

limited and the ventilation is relatively poor. Additionally, the walls in all the measured

basements were not coated with paint, which is known to reduce the exhalation rate of radon

(O'Brien et al., 1995) as well as of thoron (Wang et al., 2012). The latter is likely the main cause

of higher thoron levels in the basements, as walls and floors represent a major source of indoor

thoron.

Ln (Radon concentration)

1 2 3 4

fre

que

ncy

0

4

8

12

16

Ln (Thoron concentration)

1 2 3 4

fre

que

ncy

0

5

10

15

20

25

Figure 5.4 Histogram showing the distribution of the natural logarithms of radon and thoron concentrations in workplaces with a normal distribution fitted to the data. Geometric means of radon and thoron concentrations were 7.4 and 5.8 Bq m-3, with a geometric standard deviation of 2.2 and 2.0 Bq m-3, respectively.

The natural logarithms of the overall radon and thoron concentrations in workplaces

appeared to be normally distributed (Figure 5.4). In other words, indoor radon and thoron

concentrations followed a log-normal distribution, which perhaps indicates there are several

random variables influencing indoor radon and thoron levels in a multiplicative rather than

additive manner. The log-normal distribution of radon concentrations has been widely found in

indoor environments whether workplaces (Oikawa et al., 2006; Papachristodoulou et al., 2010)

or dwellings (Langroo et al., 1991).

115

5.3.2 Diurnal variation

Time of the day

0.00 3.00 6.00 9.00 12.00 15.00 18.00 21.00 24.00

No

rma

lise

d r

ad

on c

once

ntr

atio

n

0.0

0.5

1.0

1.5

2.0

2.5

Time of the day

0.00 3.00 6.00 9.00 12.00 15.00 18.00 21.00 24.00

No

rma

lise

d tho

ron c

once

ntr

atio

n

0.0

0.5

1.0

1.5

2.0

Figure 5.5 The variation in radon and thoron concentrations during the day.

The diurnal behaviour of radon and thoron concentrations in workplaces is illustrated

in Figure 5.5. The concentrations of radon and thoron in all workplaces were averaged on a

three-hourly basis. During normal work hours, between 09:00 am and 05:00 pm, the

concentration of radon was reduced to its lower value and remained relatively low over the

entire period. This concentration tended to rise after 05:00 pm, until it reached its maximum

value in the early morning. In fact, most workplaces are air-conditioned, which is usually

operated during the occupied hours; the air-conditioner acts as a ventilation system, and its

influence on reducing indoor radon concentrations is well-known (Marley and Phillips, 2001).

The frequent opening of doors and windows during the occupied period may also contribute in

lowering the level of indoor radon by increasing the air exchange in the room (Doi et al., 1994;

Yu et al., 2000). Reductions in the indoor radon concentration during the working hours have

been observed in a number of workplaces (Lee et al., 2013; Vaupotič et al., 2012).

Radon concentrations during and after the working hours were compared with the

average over the total time (Figure 5.6). The mean radon concentration for all workplaces

during the occupied hours was 9.6 ± 11.6, while it was 10.5 ± 11.3 Bq m-3 over the total time.

Although the concentration during the working hours was less than the average over the whole

time at most sites, no statistical significance was observed between the periods. However, a

comparison between the radon concentrations during and after working hours showed a

significant difference (p < 0.05) in the radon level between the two periods. Among workplaces,

radon concentrations in offices during the occupied time were statistically lower (p < 0.05) than

after-hours concentrations.

In some locations, such as the library and car park, the concentrations of radon during

and after work hours were equal to the average over the total time. This is possibly due to the

type of ventilation system used in those places. In the library, work hours start in the early

116

morning (7:00 am) and extend until late (around 10:00 pm); consequently, the air conditioners

are switched off only for a few hours, whereas the car park is ventilated naturally all the time.

Workplace category

Office

Classroom

Library

Laborato

ry

Research R

oom

Basement

Stora

ges

Confind a

rea

Car park

Main h

all

No

rma

lise

d r

ad

on c

once

ntr

atio

n (

Bq

m-3

)

0.0

0.5

1.0

1.5

2.0

Radon-working hours

Radon-off hours

Average radon per day

Workplace category

Office

Classroom

Library

Laborato

ry

Research R

oom

Basement

Stora

ges

Confind a

rea

Car park

Main h

all

No

rma

lise

d tho

ron c

once

ntr

atio

n (

Bq

m-3

)

0.0

0.5

1.0

1.5

2.0

Thoron-working hours

Thoron-off hours

Average thoron per day

Figure 5.6 Normalised radon and thoron concentrations during and after work hours.

The findings suggest that attention should be paid to radon concentrations during work

hours, which represent the actual exposure time, when radon levels are measured in

workplaces. As the mean radon concentration during the working hours was less than over the

total time, the average radon concentration can be estimated incorrectly. Therefore, averaging

the radon concentration in the workplace without considering the concentration during the

actual exposure time (i.e. working hours) can result in overestimating radon concentrations

and, subsequently, the effective dose.

In contrast, the diurnal variation in the thoron concentration was less observed in all

workplaces (Figure 5.6). The mean thoron concentrations during the working hours and over

the total time were 8.2 ± 10.9 and 8.2 ± 1.4 Bq m-3, respectively. The statistical analysis

indicated no difference (p = 0.984) between the thoron concentrations during and after work

hours, nor over the total time.

117

5.3.3 The influence of the measurement heights

Figure 5.7 Radon and thoron concentrations in the air at 5 cm and 120 cm above the floor for each sampling site.

Radon Thoron

Co

nce

ntr

atio

n (

Bq

m-3

)

0

10

20

30

40

50

60

70

Ground Zone

Breathin Zone

Figure 5.8 Box-whisker plot showing the overall radon and thoron concentrations at the ground and breathing zones for all location.

At each site, the radon and thoron concentrations were measured at two different

heights: 5 cm (ground zone) and 120 cm above the floor (breathing zone) (Figure 5.7). In some

sampling sites, the radon and thoron concentrations were measured at one height only due to

118

access constrains where missing concentrations in Figure 5.7. A comparison between the radon

and thoron concentrations at two different heights for all locations is shown in Figure 5.8, which

indicates the average radon concentrations near the surface of the floor and at the breathing

zone were nearly consistent at all sites. The statistical analysis revealed no significant

differences (p = 230)* between the radon concentrations of the two zones. Conversely, the level

of thoron close to the ground was significantly higher (p < 0.05) than its value at breathing

height, indicating the influence of distance on the thoron concentration. As thoron has a

relatively short half-life compared to radon, most of it decays before reaching the breathing

zone. Care, however, should be taken to distinguish between thoron and thoron progeny. Long-

lived thoron progeny, 212Pb (T1/2 = 10.6 hours), may distribute differently than thoron.

5.3.4 The influence of floor levels

Floor level

Basement Ground First Second Third and above

Ra

do

n c

once

ntr

atio

n (

Bq

m-3

)

0

10

20

30

40

50

60

70

Floor level

Basement Ground First Second Third and above

Tho

ron c

once

ntr

atio

n (

Bq

m-3

)

0

10

20

30

40

50

60

70

Figure 5.9 Box-whisker plot showing radon and thoron concentrations categorised by the floor-level.

As the measurements in this study were conducted in multi-story buildings, the

influence of the floor level on the concentration of radon and thoron was investigated. Figure

5.9 shows the radon and thoron concentrations characterised by the floor level, in that the level

of radon decreased as it moved towards the upper floors, which is not the case for thoron in our

data. The basements and ground floors had significantly (p < 0.05) higher radon concentrations

than other floor levels, with no significant difference between the basements and ground floors

(p = 0.383). In the case of thoron, only the basements were statistically different (p < 0.05) from

other floors.

* Typographic error. The value should be 0.230—Note: footnotes were added to the thesis, not present in the published paper.

119

5.3.5 The role of flooring on indoor radon and thoron concentrations

A)

Flooring type

Concrete Tile Vinyl Carpet

Ra

do

n c

once

ntr

atio

n (

Bq

m-3

)

0

10

20

30

40

50

60

70B)

Flooring type

Concrete Tile Vinyl Carpet

Tho

ron c

once

ntr

atio

n (

Bq

m-3

)

0

10

20

30

40

50

60

70

Figure 5.10. Box-whisker plot showing A) radon and B) thoron concentrations characterised by the

flooring type.

The effect of flooring on the levels of indoor radon and thoron was examined in this

study by categorising the data according to the flooring type, as can be seen in Figure 5.10.

There were four types of flooring: carpet, tile, vinyl sheet and concrete. A one-way ANOVA on

the log-transformed data showed a significant variation in radon concentration with the floor

type (p < 0.05). A post hoc test, Fisher's least significant difference (LSD), identified that the

radon concentrations in rooms with concrete floors were higher than in rooms with other

flooring types. The level of thoron was also higher in concrete rooms, indicating the effect of the

flooring type on the indoor thoron concentration. This is due to the existence of uranium and

thorium contents in concrete (Beretka and Mathew, 1985). These results suggest that covering

the floor with carpet or vinyl sheets can help mitigate indoor radon and thoron levels.

5.4 Conclusion

The concentrations of radon and thoron in various workplace categories have been

measured at Queensland University of Technology, Brisbane, Australia. The average radon and

thoron concentrations for all workplace categories were 10.5 ± 11.3 and 8.2 ± 1.4 Bq m-3,

respectively. Among workplaces, the concentration of radon in basements and confined areas

was significantly higher than in other locations, while the highest thoron level was found in

storage areas. Poor ventilation is the most likely cause of increasing the radon level in those

locations. During the working hours, radon concentration tended to be lower than at other

times of the day. This can be attributed to the ventilation systems, including air conditioners

and natural ventilation, which normally operate during working time. The diurnal variation was

less observed in the case of thoron as the change in its concentration during and after the

working hours was insignificant. The findings suggest that care should be taken when

120

measuring radon concentrations in workplaces, as its concentration during the working hours

may considerably differ from that of the entire day.

The influence of other parameters on indoor radon and thoron concentrations was

investigated from different point of view. At each site, the concentrations of radon and thoron

were measured at two heights, 5 cm and 120 cm above the floor. The level of radon showed no

difference between the two heights, while the change in thoron concentration was significant,

indicating the impact of the distance from the source on the thoron concentration. The floor

level and flooring type affected the level of indoor radon and thoron concentrations which are

elevated in basements and on the ground floor level, as well as in rooms with concrete flooring.

Acknowledgements

The authors are thankful for the support provided by Colin Phipps for arranging access to

the sampling sites.

121

References

Alghamdi, Abdulrahman S and Khalid A Aleissa. 2014. "Influences on indoor radon concentrations in Riyadh, Saudi Arabia." Radiation Measurements 62: 35-40.

Alharbi, Sami H and Riaz A Akber. 2014. "Radon-222 activity flux measurement using activated charcoal canisters: Revisiting the methodology." Journal of Environmental Radioactivity 129: 94-99.

Arnold, D, A Vargas and X Ortega. 2009. "Analysis of outdoor radon progeny concentration measured at the Spanish radioactive aerosol automatic monitoring network." Applied Radiation and Isotopes 67 (5): 833-838.

Baysson, H, S Billon, D Laurier, A Rogel and M Tirmarche. 2003. "Seasonal correction factors for estimating radon exposure in dwellings in France." Radiation Protection Dosimetry 104 (3): 245-252.

Beretka, J and PJ Mathew. 1985. "Natural radioactivity of Australian building materials, industrial wastes and by-products." Health Physics 48 (1): 87-95.

Bochicchio, F, G Campos-Venuti, S Piermattei, C Nuccetelli, S Risica, L Tommasino, G Torri, M Magnoni, G Agnesod and G Sgorbati. 2005. "Annual average and seasonal variations of residential radon concentration for all the Italian Regions." Radiation Measurements 40 (2): 686-694.

Burghele, BD and C Cosma. 2012. "Thoron and radon measurements in Romanian schools." Radiation Protection Dosimetry 152 (1-3): 38-41.

Clouvas, A, S Xanthos and M Antonopoulos-Domis. 2007. "Pilot study of indoor radon in Greek workplaces." Radiation Protection Dosimetry 124 (2): 68-74.

Crawford, Jagoda, David D Cohen, Wlodek Zahorowski, Scott Chambers and Eduard Stelcer. 2011. "A new method to combine IBA of fine aerosols with Radon-222 to determine source characteristics." Nuclear Instruments and Methods in Physics Research Section B: Beam Interactions with Materials and Atoms 269 (19): 2041-2051.

Doi, Masahiro, Kenzo Fujimoto, Sadayoshi Kobayashi and Hidenori Yonehara. 1994. "Spatial distribution of thoron and radon concentrations in the indoor air of a traditional Japanese wooden house." Health Physics 66 (1): 43-49.

Faheem, Munazza and N Mati. 2007. "Seasonal variation in indoor radon concentrations in dwellings in six districts of the Punjab province, Pakistan." Journal of Radiological Protection 27 (4): 493.

Fujimoto, K, S Kobayashi and H Yonehara. 1994. "Spatial distribution of thoron and radon concentrations in the indoor air of a traditional Japanese wooden house." Health Physics 66 (1): 43-49.

Gillmore, Gavin K, Malcom Sperrin, Paul Phillips and Antony Denman. 2000. "Radon hazards, geology, and exposure of cave users: A case study and some theoretical perspectives." Ecotoxicology and Environmental Safety 46 (3): 279-288.

Gunby, JA, SC Darby, JC Miles, BM Green and DR Cox. 1993. "Factors affecting indoor radon concentrations in the United Kingdom." Health Physics 64 (1): 2-12.

IAEA. 2003. Radiation protection against radon in workplaces other than mines, Safety Reports Series 33 Vienna: International Atomic Energy Agency.

IAEA. 2011. Radiation protection and safety of radiation sources: international basic safety standards - Interim edition. Vienna: International Atomic Energy Agency.

IAEA. 2013. National and regional surveys of radon concentration in dwellings. Vienna: International Atomic Energy Agency.

ICRP. 1993. "Protection Against Radon-222 at Home and at Work." ICRP Publication 65. Ann. ICRP 23 (2).

ICRP. 2014. "Radiological Protection against Radon Exposure. ICRP Publication 126." Ann. ICRP 43 (3) (3).

Inoue, Kazumasa, Masahiro Hosoda, Shinji Tokonami, Tetsuo Ishikawa and Masahiro Fukushi. 2013. "Investigation of radon and thoron concentrations in a landmark skyscraper in Tokyo." Journal of Radioanalytical and Nuclear Chemistry 298 (3): 2009-2015.

122

Laiolo, M, C Cigolini, D Coppola and D Piscopo. 2012. "Developments in real-time radon monitoring at Stromboli volcano." Journal of Environmental Radioactivity 105: 21-29.

Langroo, MK, KN Wise, JC Duggleby and LH Kotler. 1991. "A nationwide survey of 222Rn and gamma radiation levels in Australian homes." Health Physics 61 (6): 753-761.

Lee, Jung Min, Chan Hee Park, Shin Jae Kim and Joo Hyun Moon. 2013. "Investigation of the areas of high radon concentration in Gyeongju." Applied Radiation and Isotopes 81: 248-254.

Lee, Thomas KC and KN Yu. 2000. "Effects of air conditioning, dehumidification and natural ventilation on indoor concentrations of 222Rn and 220Rn." Journal of Environmental Radioactivity 47 (2): 189-199.

Marley, Frederick and Paul S Phillips. 2001. "Investigation of the potential for radon mitigation by operation of mechanical systems affecting indoor air." Journal of Environmental Radioactivity 54 (2): 205-219.

Martín Sánchez, A, J De la Torre Pérez, AB Ruano Sánchez and FL Naranjo Correa. 2012. "Radon in workplaces in Extremadura (Spain)." Journal of Environmental Radioactivity 107: 86-91.

Miles, J. 2001. "Temporal variation of radon levels in houses and implications for radon measurement strategies." Radiation Protection Dosimetry 93 (4): 369-375.

Moreno, V, C Baixeras, Ll Font and J Bach. 2008. "Indoor radon levels and their dynamics in relation with the geological characteristics of La Garrotxa, Spain." Radiation Measurements 43 (9): 1532-1540.

NHMRC. 1995. National Health and Medical Research Council., Recommendations for limiting exposure to ionising radiation. Guidance note: NOHSC:3022. . Canberra: Australian Government Publishing Service.

NORDIC. 2000. Naturally occurring radioactivity in the Nordic countries - recommendations: The Radiation Protection Authorities in Denmark, Finland, Iceland, Norway and Sweden.

O'Brien, RS, JR Peggie and IS Leith. 1995. "Estimates of inhalation doses resulting from the possible use of phospho-gypsum plaster-board in Australian homes." Health Physics 68 (4): 561-570.

Oikawa, Shinji, Nobuyuki Kanno, Tetsuya Sanada, Joji Abukawa and Hideo Higuchi. 2006. "A survey of indoor workplace radon concentration in Japan." Journal of Environmental Radioactivity 87 (3): 239-245.

Papachristodoulou, CA, DL Patiris and KG Ioannides. 2010. "Exposure to indoor radon and natural gamma radiation in public workplaces in north-western Greece." Radiation Measurements 45 (7): 865-871.

Pinel, J, T Fearn, SC Darby and JCH Miles. 1995. "Seasonal correction factors for indoor radon measurements in the United Kingdom." Radiation Protection Dosimetry 58 (2): 127-132.

Porstendorfer, J, A Wicke and A Schraub. 1978. "The influence of exhalation, ventilation and deposition processes upon the concentration of radon (222Rn), thoron (220Rn) and their decay products in room air." Health Physics 34 (5): 465-473.

Prasad, G., T. Ishikawa, M. Hosoda, A. Sorimachi, SK Sahoo, N. Kavasi, S. Tokonami, M. Sugino and S. Uchida. 2012. "Seasonal and diurnal variations of radon/thoron exhalation rate in Kanto-loam area in Japan." Journal of Radioanalytical and Nuclear Chemistry 292 (3): 1385-1390.

Ramola, RC, GS Gusain, BS Rautela, DV Sagar, G Prasad, SK Shahoo, T Ishikawa, Y Omori, M Janik and A Sorimachi. 2012. "Levels of thoron and progeny in high background radiation area of southeastern coast of Odisha, India." Radiation Protection Dosimetry 152 (1-3): 62-65.

Shaikh, AN, TV Ramachandran and A Vinod Kumar. 2003. "Monitoring and modelling of indoor radon concentrations in a multi-storey building at Mumbai, India." Journal of Environmental Radioactivity 67 (1): 15-26.

Shang, B., B. Chen, Y. Gao, Y. Wang, H. Cui and Z. Li. 2005. "Thoron levels in traditional Chinese residential dwellings." Radiation and environmental biophysics 44 (3): 193-199.

Singh, Surinder, Rohit Mehra and Kulwant Singh. 2005. "Seasonal variation of indoor radon in dwellings of Malwa region, Punjab." Atmospheric Environment 39 (40): 7761-7767.

123

Solomon, SB, MB Cooper, RS O'Brien and L Wilkinson. 1992. "Radon exposure in a limestone cave." Radiation Protection Dosimetry 45 (1-4): 171-174.

Solomon, SB, R Langroo, RG Lyons and JM James. 1996. "Radon exposure to tour guides in Australian show caves." Environment International 22: 409-413.

Song, Gang, Boyou Zhang, Xinming Wang, Jingping Gong, Daniel Chan, John Bernett and SC Lee. 2005. "Indoor radon levels in selected hot spring hotels in Guangdong, China." Science of the Total Environment 339 (1): 63-70.

Sundal, AV, H Henriksen, O Soldal and T Strand. 2004. "The influence of geological factors on indoor radon concentrations in Norway." Science of the Total Environment 328 (1): 41-53.

Tokonami, S, J Pan, M Matsumoto, M Furukawa, K Fujimoto, K Fujitaka and R Kurosawa. 1996. "Radon measurements in indoor workplaces." Radiation Protection Dosimetry 67 (2): 143-146.

UNSCEAR. 2000. Sources and effects of ionizing radiation. UNSCEAR 2000 report to the General Assembly with scientific annexes. New York: United Nations.

UNSCEAR. 2006. Sources-to-effects assessment for radon in homes and workplaces: United Nations Scientific Committee on the Effects of Atomic Radiation

Urosevic, V., D. Nikezic and S. Vulovic. 2008. "A theoretical approach to indoor radon and thoron distribution." Journal of Environmental Radioactivity 99 (12): 1829-1833.

Vaupotič, J, M Bezek, Norbert Kavasi, Tetsuo Ishikawa, Hidenori Yonehara and Shinji Tokonami. 2012. "Radon and thoron doses in kindergartens and elementary schools." Radiation Protection Dosimetry 152 (1-3): 247-252.

Wang, J, O Meisenberg, YH Chen, L Bi and J Tschiersch. 2012. "Mitigation of thoron exposure by application of wallpaper as a diffusion barrier." Radiation Protection Dosimetry 152 (1-3): 94-97.

WHO. 2009. Handbook on indoor radon. Geneva: World Health Organization. Willmott, Warwick, Neville Stevens, David Trezise and Jhon Simon. 2012. Rocks and Landscapes

of Brisbane and Ipswich: Geological Society of Australia Incorporated, Queensland Division.

Yamada, Y., Q. Sun, S. Tokonami, S. Akiba, W. Zhuo, C. Hou, S. Zhang, T. Ishikawa, M. Furukawa and K. Fukutsu. 2006. "Radon–Thoron discriminative measurements in Gansu Province, China, and their implication for dose estimates." Journal of Toxicology and Environmental Health, Part A 69 (7-8): 723-734.

Yu, KN, T Cheung, ZJ Guan, BWN Mui and YT Ng. 2000. "222Rn, 220Rn and their progeny concentrations in offices in Hong Kong." Journal of Environmental Radioactivity 48 (2): 211-221.

Zahorowski, W, S Whittlestone and JM James. 1998. "Continuous measurements of radon and radon progeny as a basis for management of radon as a hazard in a tourist cave." Journal of Radioanalytical and Nuclear Chemistry 236 (1): 219-225.

124

Chapter 6 A broad energy germanium detector for

routine and rapid analysis of naturally occurring

radioactive materials

Sami H. Alharbi*, Riaz A. Akber

*School of Physical and Chemical Sciences, Queensland University of Technology, 2 George

Street, Brisbane, Qld 4000, Australia

Journal :

Journal of Radioanalytical and Nuclear Chemistry (In Press)

125

Statement of Contribution of Co-Authors for Thesis by Published Paper The authors listed below have certified that:

1. they meet the criteria for authorship in that they have participated in the conception,

execution, or interpretation, of at least that part of the publication in their field of

expertise;

2. they take public responsibility for their part of the publication, except for the

responsible author who accepts overall responsibility for the publication;

3. there are no other authors of the publication according to these criteria;

4. potential conflicts of interest have been disclosed to (a) granting bodies, (b) the editor

or publisher of journals or other publications, and (c) the head of the responsible

academic unit, and

5. they agree to the use of the publication in the student’s thesis and its publication on the

Australasian Research Online database consistent with any limitations set by publisher

requirements.

In the case of this chapter:

A broad energy germanium detector for routine and rapid analysis of naturally occurring

radioactive materials

Contributor Statement of contribution

Sami Alharbi

Signature:

Date: 31/03/2016

Identified concept for the research paper, prepared samples of

analysis, conducted experimental work, analysed and interpreted the

data, wrote the manuscript

Riaz Akber Assisted with editorial process, aided in concept development

Helped in experimental design, reviewed the manuscript.

Principal Supervisor Confirmation

I have sighted email or other correspondence from all Co-authors confirming their certifying

authorship.

Name:

Signature:

Date:

126

Abstract

The analysis of naturally occurring radioactive materials using a broad energy germanium

detector is reported, along with a description of the rapid measurement of 226Ra and the

determination of 227Ac in environmental samples. The detector calibration procedure, sample

preparation, and the necessary correction factors have been described. The detector was found

useful for the routine analysis of natural radionuclides, with the ability to detect low photon

emission peaks at lower gamma energies. The results showed that the direct determination of

226Ra via the 186 keV peak is feasible and the determination of 227Ac using gamma spectrometry

is reliable.

Keywords: BEGe; NORM; 226Ra; 227Ac; Rapid measurement

6.1 Introduction

The information obtained from individual radionuclides of naturally occurring

radioactive material (NORM) has become a powerful tool for many scientific disciplines, such as,

radiometric dating (Ugur et al., 2003, Aba et al., 2014), atmospheric studies (Baskaran, 2011,

Doering and Akber, 2008), soil erosion and sediment deposition rates (Matisoff, 2014, Zapata,

2003). Selections of measurement methods and sample preparations are dictated by a project’s

specific needs, the required accuracy and precision of results, and by the availability of

resources. Analytical methods for NORM radioisotopes are generally time-and resource-

consuming—and, therefore, expensive. Gamma-ray spectrometry is one of the most effective

measurement techniques for identifying radionuclides and quantifying the activity levels of

materials in a non-destructive manner (Engelbrecht and Schwaiger, 2008). For example,

Vesterbacka et al. (2009) compared various methods of detecting radioisotopes by alpha, beta

and gamma spectrometry and found gamma spectrometry to be superior in terms of the speed

of analysis.

Over the past few decades, several types of gamma detectors have been developed for

various applications. The broad energy germanium (BEGe) detection system (manufactured by

Canberra Industries Inc.) represents the latest technology that is commercially available for

routine gamma analysis. According to the manufacturer, the BEGe detector has a resolution

equivalent to the conventional planar detector at lower gamma energies and comparable to the

coaxial detector at higher energies. Han and Choi (2010) tested different types of germanium

detectors. In their study, the BEGe detector had the highest efficiency at lower gamma energies

than other detectors. This may be beneficial for analysing environmental samples in which a

number of key radionuclides emit gamma-rays in low-energy regions.

127

The BEGe detector has been extensively used in nuclear physics (e.g., in neutrino-less

double beta decay experiments (Aalseth et al., 2011, Budjáš et al., 2013)) and in medical science

(e.g., in lung counting measurements (Webb and Kramer, 2001)). However, the use of BEGe

detectors in environmental studies is still limited. This paper examines the potential

implications of BEGe detectors for the routine measurement of NORM. The detector was also

employed to investigate two other topics in the context of NORM: the rapid measurements of

226Ra and the determination of 227Ac.

6.1.1 Routine analysis of NORM

Routine measurement of NORM targets radionuclides in the uranium, actinium and

thorium decay series as well as potassium-40. Some of radionuclides in the natural series

cannot be measured directly with gamma spectrometry as they do not emit gamma-ray or their

emitted gamma-ray is very weak. In such cases, these radionuclides can be determined through

their daughter radionuclides assuming a state of equilibrium between the daughter and parent

radionuclides. For example, 238U emits a very weak gamma line (0.064%) at 49.55 keV, which

makes the direct measure of 238U through that line impracticable. Therefore, the activity of 238U

is typically determined through the gamma-rays of its daughter nuclides, such as, 234Th, 234mPa,

214Pb or 214Bi.

6.1.2 Rapid measurements of 226Ra

226Ra is an important natural radionuclide that has many applications in environmental

and dating studies (Voltaggio et al., 1995, Shuktomova and Rachkova, 2011, Condomines and

Rihs, 2006, Paytan et al., 1996). It has a long half-life of 1600 years, and it decays to 222Rn (half-

life of 3.8 days) by emitting alpha particles. This decay is accompanied with gamma-rays at

186.1 keV, with a low emission probability (3.6%) (Ekström and Firestone, 2004).

226Ra can be measured in environmental samples using gamma spectrometry (Murray et

al., 1987, Dowdall et al., 2004, Landsberger et al., 2013). In the gamma-ray spectrometry

technique, 226Ra is commonly determined through its progenies: 214Pb (at 295.2 and 351.9 keV)

and 214Bi (at 609.3, 1120.3 and 1764.5 keV) (Al-Ghamdi et al., 2016, Khandaker et al., 2012).

This is because the 186.1 keV peak of 226Ra interferes with the 185.7 keV (57.2%) peak of 235U,

the 186.1 keV (0.009%) peak of 230Th, and the 186.2 keV (1.76%) peak of 234Pa. The latter two

peaks are usually ignored due to the extremely low intensity of 230Th and the low branching

ratio of 234Pa. However, this procedure is time consuming because it requires a delay period of

at least three to four weeks to achieve secular equilibrium between 226Ra and its daughters,

214Pb and 214Bi.

In order to overcome the delay time, a number of methods for the rapid determination

of 226Ra have been attempted. Some researchers have adopted the method of deconvolution of

the overlapping peak at 186 keV (Saïdou et al., 2008, Zhang et al., 2009). However, this

128

approach is unlikely to provide accurate results due to the small distance between the

overlapped peaks as demonstrated by Dowdall et al. (2004). Another approach of determining

226Ra is by direct measurement of the186.1 keV peak after subtracting the contribution of 235U

(e.g. (De Corte et al., 2005, Justo et al., 2006, Dowdall et al., 2004, Johnston and Martin, 1997,

Völgyesi et al., 2014)). The contribution of 235U can be deduced through the natural activity ratio

of 235U/238U using 238U daughter, 234Th (Dowdall et al., 2004) or using an independent method

for measuring 235U concentration (Johnston and Martin, 1997), which would add extra cost to

the analysis. The main issues with the former approach are the long waiting time for secular

equilibrium of 234Th, which can be guaranteed after 240 days. Whereas the problem with the

latter approach is the alterations that could occur to the 235U/238U ratio as a result of certain

nuclear activities (Danesi et al., 2003, Yoshida et al., 2000, Sansone et al., 2001). A different

technique was developed by Johnston and Martin (1997) utilising the 222Rn ingrowth for the

rapid analysis of 226Ra in water samples. Nevertheless, their method leads to an overestimate of

the 226Ra concentration as a result of a systematic error.

Alternatively, the contribution of 235U can be obtained from its gamma lines at 143.8,

163.3 and 205.3 keV (De Corte et al., 2005, Justo et al., 2006). For example, Justo et al. (2006)

used 235U peak at 143.8 keV to directly measure 226Ra from 186.1 keV peak. However, no

comparison has been made to assess the accuracy of their results. Moreover, in another study

conducted by De Corte et al (2005), the 143.8 and 205.3 keV peaks were examined and

demonstrated to have unsatisfactory results due to large uncertainties associated with it. The

present study investigates the feasibility of using 235U peaks (143.8, 163.3 and 205.3 keV) for

the rapid determination of 226Ra.

6.1.3 Determination of 227Ac using gamma-ray spectrometry

The actinium isotope 227Ac is a member of the 235U series derived from the decay of

231Pa. It is a highly toxic radionuclide that can pose radiological health concerns to both humans

and the environment. The effective dose coefficients for inhalation and ingestion intakes of 227Ac

are more restrictive than those for 226Ra (ICRP, 2012). Although the activity of 227Ac in

environmental samples is expected to be extremely low, as the natural abundance of the whole

235U series is only 0.0072 of the 238U, it may be elevated to a higher level through certain human

activities. In an oil and gas industry, a level of 2.5 Bq g-1 of 227Ac was reported in scale samples

(Kolb and Wojcik, 1985). This level is 25 times higher than the exemption level recommended

by the International Atomic Energy Agency (IAEA, 2014).

The measurement of 227Ac is associated with difficulties due mostly to its low

abundance. The current analytical techniques available for the determination of 227Ac are ion

exchange with alpha spectrometry (Martin et al., 1995, Bojanowski et al., 1987), delayed

coincidence counting (Shaw and Moore, 2002), liquid scintillation counting (Kossert et al.,

129

2015), and gamma-ray spectrometry (Van Beek et al., 2010). All of these techniques, with the

exception of gamma-ray spectrometry, require complicated chemical separation and extraction

processes. The present work uses a BEGe detector to establish analytical techniques for

measuring 227Ac using gamma-ray spectrometry in soil samples.

6.2 Material and methods

6.2.1 Routine analysis of NORM

In the analysis of NORM with gamma spectrometry, a number of key radionuclides are

routinely measured. These radionuclides are listed in Table 6.1. This table has been generated

using the nuclear data given in the Table of Isotopes (Ekström and Firestone, 2004). In the

literature, several authors have commented on the photon peaks behaviour during

measurements. These comments are discussed in the paragraphs below and also summarised in

Table 6.1.

In the 238U series, seven nuclides, namely, 234Th, 234mPa, 230Th, 226Ra, 214Pb, 214Bi, and

210Pb, emit detectable gamma-rays. In practice, the direct daughters 234Th and 234mPa are most

preferable for such measurements (Yücel et al., 2009, Huy and Luyen, 2004, Lenka et al., 2009,

Yücel et al., 1998) due to their short half-lives of 24.1 d and 1.17 min, respectively. Moreover,

these radioisotopes are the early members of the 238U series, which means that they require less

time to achieve secular equilibrium with their parent than other radionuclides in the series.

234Th emits several gamma-lines, at 63.3, 92.4 and 92.8 keV. The latter peaks are doublets and

interfere with fluorescence x-rays, such as Kα1 of thorium series radionuclides (228Ac) at 93.3

keV. This makes it difficult to deconvolute. Therefore, this peak was avoided, although some

authors found this peak provided acceptable results (Kaste et al., 2006, El-Daoushy and

Hernández, 2002).

The peaks at 63.3 (4.80 %) and 1001 keV (0.84 %) are more practical for estimating the

activity concentration of 238U. The gamma line at 63.3 keV suffers from self-absorption and from

interference with 232Th at 63.8 keV (0.26%) and 231Th at 63.9 keV (0.02%). However, this self-

absorption can be corrected (Huy et al., 2013), and the contributions of 232Th and 231Th to the

peak 63.3 keV in environmental samples is normally negligible, unless the sample has higher

232Th or 235U content (Yücel et al., 2009). In this work, the interference is corrected by

subtracting the contributions of 232Th, such that the activity of 234Th (A234) is given as:

𝐴234 = [𝐶63

𝜀63 𝑚− ( 𝐼232 𝐴232)] .

𝑓𝑠

𝐼234 (6.1)

where C63 is the count rate of the 63 keV peak, A232 is the activity of 232Th (as determined

through its daughter, 228Ac), 63 is the calibration efficiency of the 63 keV peak, m is the mass of

the sample, fs is the correction factor of the self-absorption, and I232 and I234 are the photon

130

emission intensities of 63.8 and 63.3 keV peaks of 232Th and 234Th, respectively. Equation. 1

assumes a secular equilibrium between 238U and 234Th. This assumption may not be valid under

some special cases where 238U has been separated from the remainder of the series during the

past four months.

This research also employed the 1001 keV (0.84%) peak of 234mPa to estimate the

activity concentration of 238U. Despite having low gamma emissions, this peak has been

successfully used to determine 238U in environmental samples (Yücel et al., 1998). It also

exhibits potential interference with a very weak 228Ac gamma line at 1000.7 keV (0.005%). This

interference was ignored, since the contribution of 228Ac to the 1001 keV peak is negligible.

However, overlooking this contribution in materials rich in thorium may cause an error of 0.05

to 2.98% (Yücel et al., 2009), which is likely to be well within the range of experimental

uncertainties.

Neither the weighted averages of 214Pb and 214Bi nor the natural isotopic activity ratios

of 235U/238U were used to estimate the activity concentration of 238U. In the first case, the

equilibrium between these isotopes and 238U cannot be guaranteed in samples, because of the

differences in the chemical properties and environmental behaviours of the various long-lived

isotopes in between. The second approach is not always viable, because, in some parts of the

world, the ratio 238U/235U has been altered due to nuclear contamination (Danesi et al., 2003,

Yoshida et al., 2000, Sansone et al., 2001). In addition, the natural abundance of 235U is very low

(0.72%), which could increase the measurement’s uncertainty.

Other important radionuclides in the 238U series that have many applications in

environmental studies are 230Th, 226Ra and 210Pb. 230Th emits a number of weak gamma

radiations, such as 67.7 keV (0.38%) and 143.9 keV (0.049%). Only the 67.7 keV peak is

employed for gamma spectrometry (Lozano et al., 2001, Yücel et al., 2010). The higher efficiency

of the BEGe detector at lower gamma-rays makes the determination of 230Th through the 67.7

keV peak possible. The measurement of 226Ra is discussed in further detail in the next section.

The 210Pb nuclide has a single gamma-line at 46.5 keV (4.25%). It should be noted that both the

46.5 and 67.7 keV peaks are affected by self-absorption, which was corrected for each sample.

The actinium series, which was named after the first element discovered in the series,

227Ac, is headed by 235U. Within this series, four nuclides (235U, 227Th, 223Ra, and 219Rn) are

applicable to gamma spectrometry. The measurement of this series is often problematic due to

its low abundance in nature. The major peak of 235U at 185.7 keV (57.2%) interferes with the

186.2 keV (3.5%) peak of 226Ra. This interference can be corrected by subtracting 226Ra

contribution at the peak. The portion of 226Ra can be obtained from its daughters, 214Pb and

214Bi, in equilibrium. Care must be taken, so, 222Rn, the gaseous daughter of 226Ra and the direct

parent of 214Pb and 214Bi, loss not occur in these samples. Other peaks of 235U at 143.8 (10.96%),

131

163.3 (5.08%) and 205 keV (5.01%) can be used as alternative measures of 235U (Pöllänen et al.,

2003, De Corte et al., 2005), despite the probability for true summing with the 19.6 keV gamma

line (Garcıa-Talavera et al., 2001).

In the 232Th series, several radioisotopes, particularly 228Ac, 224Ra, 212Pb, 212Bi and 208Tl,

have strong gamma emissions. 232Th itself has a very weak gamma line at 63.8 kev (0.27 %),

which interferes with 234Th and 231Th. It is more easily determined through its short-lived

daughter, 228Ac, which has strong gamma lines at 911.2 (25.8%), 968.9 (15.8%), and 338.3

(11.3%) keV assuming secular equilibrium. 228Ac is also used to determine the activity

concentration of 228Ra, which emits gamma-rays that difficult to be detected due to their low

intensities.

Another important radionuclide in the 232Th series is 228Th. Unfortunately, this isotope

has a very weak gamma line at 215.98 keV (0.26%), which may be useful in thorium-rich

samples. In this study, the activity concentration of 228Th was determined through its 215.98

keV peak, as well as through its daughters, 224Ra, 212Pb, 212Bi and 208Tl. The gamma line of 224Ra

at 241.0 keV (4.1 %) interferes with the 242.0 keV (7.0%) line of 214Pb, which makes the use of

its daughters, 212Pb, 212Bi, and 208Tl, more practical. 212Pb is typically measured using its

strongest gamma line of 238.6 keV (44%). The other peak at 300.1 keV (3.3%) interferes with

other peaks from the actinium series (231Pa and 227Th) and thus is not recommended.

The gamma lines of 212Bi are not frequently used, as its strongest peak at 727.3 keV

(6.6%) interferes with the 726.9 keV (0.64 %) peak of 228Ac. Other 212Bi gamma lines at 1620.7

(1.5 %) and 39.9 keV (1.5 %) are weak, and the latter suffers from serious self-absorption.

These peaks can be used after correcting the results for interference and self-absorption.

However, the 208Tl peaks at 2614.53 (99 %), 583.2 (85 %), and 510.8 (25.8 %) were employed

instead. The accuracy of these peaks may be affected by potential interference at the latter peak,

with the annihilation peak occurring at 511 keV and a high probability of true coincidence

summing at the other two peaks (Garcıa-Talavera et al., 2001, Şahiner and Meriç, 2014).

The naturally occurring radionuclide 40K can also be quantified using gamma

spectrometry. The gamma line at 1460.8 keV (11%) was used to determine the activity

concentration of 40K. This gamma line interferes with the 1459.14 keV (0.80%) peak of 228Ac,

and the spectral interference can cause serious errors of up to 100% or more in the measured

activity of 40K, depending on the 232Th/40K ratio (Lavi et al., 2004). However, the contribution of

228Ac at the 1460.8 keV peak was corrected as described by Lavi et al. (2004), and the 40K

analysis was performed after that correction. The error in the activity concentration of 40K

caused by 228Ac can be calculated by multiplying the intensity ratio (I228Ac / I40K) at the peak by

the activity concentration ratio (232Th/40K). The uncertainty in the measured activity

132

concentration of 40K due to peak interference should be corrected accordingly. An example is

shown in Section 2.8.

Table 6.1 Gamma energy lines of key natural radionuclides (Ekström and Firestone, 2004)

Isotope Half-life Energy

(keV)

Emission

(%)

Note

238U series

234Th 24.1 d 63.29 4.80 Self-absorption is pronounced - Interferes with 232Th at

63.83keV(0.267%) 234mPa 1.17 min 1001.03 0.84 Potential interference with 1000.7 keV (0.005%) peak of 228Ac 230Th 7.54 ×104 a 67.67 0.38 Self-absorption is pronounced 226Ra 1600 a 186.10 3.50 Overlapping with 185.7 keV (57.2 %) peak of 235U 214Pb 26.8 min 351.92

295.21

35.80

18.50

214Bi 19.9 min 609.31

1120.29

1764.49

1238.11

44.80

14.80

15.36

5.86

A probability of true coincidence summing effect

A probability of true coincidence summing effect

A probability of true coincidence summing effect 210Pb 22.3 a 46.54 4.25 Self-absorption is more pronounced

232Th series

228Ac 6.15 h 911.20

968.97

338.32

964.77

463.01

26.60

16.20

11.27

5.11

4.44

A probability of true coincidence summing effect

Interfering with 223Ra (338keV, 2.79%)

228Th 1.912 a 215.98 0.26 Can be useful in thorium rich samples 224Ra 3.66 d 240.99 3.97 Interferes with 241 keV (7%) peak of 214Pb 212Pb 10.64 h 238.63 43.30 212Bi 60.55 min 39.86

727.33

1.09

6.58

Self-absorption is more pronounced

Interferes with 726.8 keV (0.64%) peak of 228Ac 208Tl 3.053 min 2614.53

583.19

510.77

860.56

99.00

84.50

22.60

12.42

All peaks are affected by the true coincidence summing

potential interference with 214Pb at 580.15 keV (0.35%)

interference with the annihilation peak at 511 keV

235U series

235U 7.038×108 a 185.71

143.76

163.36

205.31

57.2

10.96

5.08

5.01

Interferes with 226Ra.

Interferes with 223Ra(143.1keV,3.22%) and 230Th( 143.87keV,

0.049%) and a probability of true coincidence summing effect.

Potential interference with 228Ac at 204.03 keV (0.12%) and 228Th at 205.93 keV (0.0204%) in thorium rich samples

227Th 18.72 d 235.97

50.13

256.25

12.3

8.0

7.0

Self-absorption is more pronounced

223Ra 11.44 d 269.46

154.21

144.23

13.7

5.62

3.22

Interferes with 271.23 keV (10.8%) peak of 219Rn and 270.24

keV (3.43%) peak of 228Ac

Interferes with 153.98 keV (0.74%) of 228Ac

219Rn 3.96 s 271.23

401.81

10.80

6.24

Interferes with 223Ra and 228Ac

40K 1.28×109 a 1460.83 11.00 Interferes with 1459.14 keV (0.80 %) peak of 228Ac

133

6.2.2 Rapid measurements of 226Ra

The activity concentration of 226Ra was determined directly from the 186 keV peak by

subtracting the contribution of 235U from the peak. The 235U portion was estimated from its

gamma lines at 143.8, 163.3, and 205.3 keV. Although these gamma lines may be affected by

true coincidence summing with the 19.6 keV line (Garcıa-Talavera et al., 2001), the use of

standard calibration sources identical to those of the sample can minimise this effect (Gilmore,

2008). The activity concentration of 226Ra (ARa) is given by the following formula:

𝐴𝑅𝑎 = [𝐶186

𝜀186 𝑚 − ( 𝐼𝑈 𝐴𝑈)] .

𝑓𝑠

𝐼𝑅𝑎 (6.2)

where C186 is the count rate of the peak at 186 keV, AU is the activity of 235U, 186 is the

calibration efficiency of the 186 keV peak, and IU and IRa are the gamma emission intensities of

the 186 keV peak of 235U and 226Ra, respectively.

The activity concentration of 235U obtained from the peak at 143 keV was corrected for

the interference caused by 223Ra and 230Th, as follows:

𝐴𝑈(143 𝑘𝑒𝑉) = [𝐶143

𝜀143 𝑚 − ( 𝐼223𝑅𝑎 𝐴223𝑅𝑎) − ( 𝐼230𝑇ℎ 𝐴230𝑇ℎ)] .

𝑓𝑠

𝐼𝑈 (6.3)

where C143 is the count rate of the peak at 143 keV, A223Ra and A230Th are the activity

concentrations of 223Ra and 230Th, 143 is the calibration efficiency of the 143 keV peak, and I223Ra

and I230Th are the gamma emission intensities of the 143 keV peak of 223Ra and 230Th,

respectively. The 223Ra activity was measured using its daughter 219Rn at the 401.8 keV (as the

269.5 keV peak has more interferences), while the activity of 230Th was determined at the 67.7

keV peak.

The activity of 226Ra was also determined indirectly through its daughters, 214Pb and

214Bi, after three weeks of waiting time. The results were then compared with those of the direct

method. For the reference materials IAEA-434 and RGU-1, the measurements were repeated

twice (after 1 day and 30 days of preparing the sample) to examine the influence of the elapsed

time. The results were then compared with the IAEA values. The quoted uncertainties in this

method were two standard deviations and arise from the counting statistic, efficiency

calibration and peak interference corrections.

6.2.3 Determination of 227Ac using gamma ray spectrometry

227Ac decays through the emission of low-energy beta particles (98.62%) to 227Th and

alpha particles (1.38%) to 223Fr. This decay is associated with extremely low gamma intensity,

which makes the direct measurement of 227Ac via gamma spectrometry very challenging.

Instead, in this research, 227Ac was determined by measuring its daughters, 227Th, 223Ra, and

219Rn, after a delay period of at least four months. The delay period was dictated by the half-life

134

of 227Th, which is 18.7 days. Some samples may have already reached equilibrium state between

227Ac and 227Th; in that case, the waiting time is not required. The direct daughter, 227Th, emits a

number of detectable gamma-rays (Table. 6.1). The strongest line is 12.3% at 235.97 keV. This

line is useful to use, but is also likely to interfere with the 212Pb gamma line at 238.63 keV

(43.3%), especially in samples rich in thorium. The next gamma line of 227Th occurs at 50.13 keV

(8.0%), where the self-absorption is more pronounced. Another peak that can be used is the

256.25 keV (7%) peak, which is in close proximity to the 214Pb peak at 258.79 keV (0.55%).

223Ra and its short-lived daughter, 219Rn, have very useful gamma lines that can be used

to determine 227Ac. The peak of 223Ra at 269.46 keV (13.7 %) has been used after deconvoluting

the spectral interference due to the 270.24 keV (3.43 %) peak of 228Ac and the 271.23 keV

(10.8 %) peak of 219Rn (Van Beek et al., 2010). Desideri et al. (2008) used the gamma peak of

219Rn at 401.81 keV (6.6 %) to determine 227Ac in waste samples. 219Rn has another gamma line

at 271.23 keV (10.8 %); however, it interferes with the two gamma lines emitted from 223Ra and

228Ac.

227Ac was determined through its daughters 227Th (at 50.13 and 256.25 keV), 223Ra (at

269.46 keV), and 219Rn (at 401.81 keV). The 235.97 keV peak of 227Th was deliberately not used

due to its complex interference and the 269.46 keV peak was corrected for the contributions of

other radionuclides (i.e., 219Rn and 228Ac). The results were compared with the expected values

for 227Ac, as well as with the activity of 235U.

6.2.4 Sample preparation

Thirteen samples of different types were prepared in a simple and reproducible way. Six

of these samples are reference materials with known activity concentrations and seven are

representative samples. It should be noted that the activities given for the IAEA-434 sample

were corrected for decay and ingrowth, since the equilibrium is disturbed in the 238U series. The

representative samples were selected to cover a range of activity concentrations (from a few

tens of Becquerel to a few mega Becquerel per kilogram) and to represent different radioactive

equilibrium conditions. A description of all samples is provided in Table 6.2.

The samples were dried overnight at 110oC, crushed and sieved through a 2 mm sieve to

obtain homogeneity. Samples TI and TII were initially in clay boulder form. They were coarsely

ground in mortar pestle to generate a sample that was sieved. Each sample was then weighed,

compressed in a polyester petri dish (6 cm in diameter and 1.2 cm in height) using a 10 kpa

hydraulic press and sealed airtight with a silicone adhesive to retain radon. It has been found

that placing the reference material RGU-1 in a sealed cylindrical plastic container with a

diameter of 6 cm can reduce radon loss to 0.8% (Scholten et al., 2013).

135

Table 6.2 Sample descriptions

Sample code Description Bulk Density (g cm-3)

Reference materials

IAEA-434 Certified reference material-Phosphogypsum 0.91

RGU-1 Certified reference material- uranium ore 1.31

RGTh-1 Certified reference material- thorium ore 1.30

RGK-1 Certified reference material- potassium sulphate 1.35

SS Spiked SiO2 1.39

SG Spiked grass 0.25

Representative samples

CHI Soil sample 1.12

CHII Soil sample 1.10

TI Uranium tailing 1.59

TII Uranium tailing 1.59

BC Black concentrate 2.09

CS Contaminated soil* 2.36

MO Monazite rich with thorium 3.51 * Contaminated due to a past mineral sand processing operation

6.2.5 Analytical technique

The BEGe detector (BE3820) used in this study is manufactured by Canberra and is

made of a high-purity germanium (HPGe) crystal housed in a 100 mm thick cylindrical graded

lead shield (Model 747, Canberra). The detector has good resolution, with a Full Width at Half

Maximum (FWHM) of 0.59 keV at 122 keV and 1.77 keV at 1332 keV and an optimum efficiency

across a wide range of gamma energies between 3 keV and 3 MeV.

The counting time was set at 48 hrs to achieve better statistics. The net count at the peak

of interest was obtained using Genie 2000 software supplied by Canberra. The activity

concentrations of the natural radionuclides in each sample were calculated via direct

comparison with a calibration standard with the same geometry. The calibration standard was

prepared by diluting certified reference materials supplied by New Brunswick Laboratory

(NBL) in silica powder. The diluted materials contain uranium and thorium ores. Reagent grade

potassium chloride supplied by Ajax Chemicals (Australia) was also added to the diluted

materials. A blank sample with the same matrix and composition as the standard was prepared,

using identical geometry to make background corrections.

The comparative method is highly recommended for environmental analysis (Gilmore,

2008). In this method, the effects of geometry, self-absorption, and coincidence summing are

cancelled out in the standard and the sample. However, in environmental samples, some

variations may exist mostly in the density, mineralogy and activity strength. In such situations,

correction factors are required for both self-absorption and random summing. Corrections to

the true summing may not be required, since the standard and the sample have identical

geometries and are analysed for the same nuclides.

136

The calibration data were obtained empirically using the prepared calibration standard.

The counts per photon emission of the calibration standard as a function of photon energy are

plotted in Figure 6.1(A). The results of the reference materials are plotted in the same figure as

well to check the performance of the detector. These values were obtained using the same

geometry, petri dishes. The measured counts per emission of the calibration standard and

reference materials were consistent. The results indicate a higher efficiency of the detector at

low gamma energy. The overall uncertainty in the calibration data is estimated to be 4% and it

arises from the uncertainty in the source activity and counting statistics.

The empirical calibration of the BEGe detector was compared with an HPGe coaxial

detector after preparing the calibration standard material in Marinelli Beaker geometry. Figure

6.1 (B) shows a superior efficiency of the BEGe detector at lower gamma energy and

comparable values at higher gamma energy. Furthermore, 39.86 and 67.67 keV peaks of 212Bi

and 234Th were only detected with the BEGe detector.

The analytical performance of the laboratory was tested using two statistical

parameters: relative bias and u-score. The relative bias identified the potential difference

between the analyst’s results and the IAEA assigned values using the following relation:

𝑅𝑒𝑙𝑎𝑡𝑖𝑣𝑒 𝑏𝑖𝑎𝑠 =(𝐴𝑛𝑎𝑙𝑦𝑠𝑡 𝑣𝑎𝑙𝑢𝑒−𝐴𝑠𝑠𝑖𝑔𝑛𝑒𝑑 𝑣𝑎𝑙𝑢𝑒)

𝐴𝑠𝑠𝑖𝑔𝑛𝑒𝑑 𝑣𝑎𝑙𝑢𝑒× 100 (6.4)

The u-scores examined whether the analyst’s results differed significantly from the

certified values, taking into account their uncertainties. The u-scores were calculated using the

following equation (Brookes et al., 1979):

𝑢𝑠𝑐𝑜𝑟𝑒 =(𝐴𝑛𝑎𝑙𝑦𝑠𝑡 𝑣𝑎𝑙𝑢𝑒−𝐴𝑠𝑠𝑖𝑔𝑛𝑒𝑑 𝑣𝑎𝑙𝑢𝑒)

√[(𝐴𝑛𝑎𝑙𝑦𝑠𝑡 𝑢𝑛𝑐𝑒𝑟𝑡𝑎𝑖𝑛𝑡𝑦)2+(𝐴𝑠𝑠𝑖𝑔𝑛𝑒𝑑 𝑢𝑛𝑐𝑒𝑟𝑡𝑎𝑖𝑛𝑡𝑦)2] (6.5)

The analysed and assigned uncertainties are quoted as two standard deviations. In order

to determine the significance of the comparison of the measured and expected values, the u-

scores were compared with the critical values listed in the t-test table. The u-test parameter of

1.96 for a confidence interval level of 95% was applied. Thus, the results were considered to

pass the test when u-score 1.96.

137

(A)

Energy (keV)

100 1000

cp

s/p

ho

ton e

mis

sio

n

0.00

0.02

0.04

0.06

0.08

0.10

0.12

0.14

RG-set

IAEA-434

SS

SG

Calibration standard

(B)

Energy (keV)

100 1000

cp

s/p

ho

ton e

mis

sio

n

0.00

0.02

0.04

0.06

0.08

0.10

0.12

0.14

BEGe_Petri dish

HPGe_Marinelli beaker

Figure 6.1 (A) Count per emission versus the gamma energy for the reference materials using the BEGe detector. (B) A comparison between the detection efficiency of a BEGe detector and an HPGe coaxial detector. The results are for a calibration standard in petri dish geometry (50 g) in BEGe and Marinelli Beaker geometry (1000 g) in HPGe-coaxial.

6.2.6 Gamma self-attenuation

The phenomenon of self-attenuation occurs when gamma-rays pass through different

materials, and it depends on both the gamma energy of the incident photon and the density and

composition of the material. The effect is more pronounced at low gamma energies, which are

associated with some natural radionuclides, such as 210Pb, 234Th, and 230Th (Pilleyre et al., 2006).

In the comparison method used to determine the activity concentration, the measured sample is

ideally required to have a geometry, density, and composition identical to those of the

calibration standard. This approach ensures that the self-attenuation effect and the other effects

in the standard are similar to those of the sample and, thus, cancel each other out. However,

variations in density and composition do exist. In order to cover a wide range of environmental

samples, one option may be to prepare a large number of calibration standards with different

densities and compositions, which is costly and time consuming. Alternatively, a correction

factor can be worked out and applied to the results.

There are several methods to estimate the correction factor of the self-attenuation

effect. In this study, a practical approach using the direct transmission of gamma-rays was used.

This method was proposed by Cushal (1983) and it is based on measuring the transmission of

gamma-rays through a sample by positioning a point source above the sample. The self-

attenuation correction factor (fs) for a sample relative to the calibration standard is given as

(Cutshall et al., 1983):

𝑓𝑠 =ln(𝐼 𝐼0)⁄

𝑠𝑎𝑚𝑝𝑙𝑒

ln(𝐼 𝐼0)⁄𝑠𝑡𝑎𝑛𝑑𝑎𝑟𝑑

×(𝐼0−𝐼𝑠𝑡𝑎𝑛𝑑𝑎𝑟𝑑)

(𝐼0−𝐼𝑠𝑎𝑚𝑝𝑙𝑒) (6.6)

where I and Io are the intensities with and without attenuation, respectively, expressed in count

rates. The correction factors are limited to gamma energies emitted from the available sources;

241Am, 137Cs, and 60Co (their activities ranging between 8 to 40 kBq). For other gamma energies,

138

the correction factors were interpolated from a self-absorption curve thus produced. The values

of fs for different samples used in this study are listed in Table 6.3†.

6.2.7 Coincidence summing

Coincidence summing occurs when two or more photons are recorded almost

simultaneously by a detector. If the incident photons are emitted by the same nuclide, the

summing is true, and if they originate from different nuclides, the summing is random. True

coincidence summing depends on the decay scheme of the radionuclide, as well as on the

geometry of the sample. Coincidence summing can introduce substantial error into the

measurement of activity concentrations, which requires correction. In the present study, the

summing errors in the standard and the sample are cancelled out, since the standard contains

reference materials with the same nuclides and geometry as the sample.

On the other hand, random summing is count-rate-dependant, and as long as the

activities of the sample and the calibration standard are comparable, it will be cancelled out. For

environmental samples, the count rate is typically low, which means negligible random

summing. However, if the activity of the sample is significantly higher than that of the standard,

random summing should be taken into account. This study has adopted an approach, which is

based on a method described by Wyttenbach (1971). The random summing correction factor

(fR) is calculated as follows (Wyttenbach, 1971):

𝑓𝑅 =1

𝑒−𝑊𝑓 (

𝑇𝑡

−1) (6.7)

where T is the real time, t is the live time, and Wf is the Wyttenbach factor = (2τ/θ)—where τ

represents the resolution time and θ is the sum of the fixed conversion time and the mean

channel-dependent conversion time of the memory cycle of the multi-channel analyser. The Wf

factor was determined experimentally using an approach given in (Wyttenbach, 1971) by

placing a 137Cs point source at a fixed position (100 mm) above the detector to yield an

uncertainty of less than 1% in the net area of the 661.6 keV peak. Another source, 241Am, was

introduced at a position above the 137Cs source and was moved closer to the detector through a

stepwise approach to vary the dead time. The ratio of the count rates at the peak (661.6 keV)

was plotted against the ratio of the real time to the live time, and the gradient was Wf. The

correction factors presented in Table 6.3 were given relative to the calibration standard. The fR

values for reference materials are equal to unity, as they have the same matrices and

† Other methods have been proposed to estimate the self-attenuation correction factor. For example,

theoretical and simulation approaches have been developed by some researchers (Dababneh et al., 2014;

McMahon et al., 2004; Pilleyre et al., 2006). However, this technique requires precise knowledge of the

sample geometry and composition, as well as of the detector characteristics. (Note: footnotes were added

to the thesis, not present in the published paper)

139

comparable activities with the calibration standards. This was confirmed by the experiment as

well. The fR values of CHII and TII samples were similar to that of CHI and TII, respectively.

Table 6.3 Correction factors for self-absorption (fs) and random summing (fR); material densities (g cm-3) are given between the brackets. A brief description of samples is given in Table. 2.

Correction Energy (keV) Nuclide SG CHI TI BC CS MO factor (0.25) (1.10) (1.59) (2.09) (2.36) (3.51) fs 46.54 210Pb 0.69 0.98 1.41 3.39 5.51 26.50 50.13 227Th 0.70 0.98 1.39 3.05 4.88 23.92 63.29 234Th 0.71 0.99 1.34 2.19 3.34 17.33 67.67 230Th 0.72 0.99 1.33 2.00 3.00 15.80 143.76 235U 0.77 0.99 1.19 1.34 1.89 2.02 186.00 226Ra+235U 0.79 0.99 1.14 1.25 1.56 1.85 205.31 228Th 0.80 0.99 1.12 1.22 1.44 1.78 215.99 228Th 0.81 0.99 1.12 1.20 1.39 1.75 238.63 212Pb 0.81 0.99 1.10 1.17 1.29 1.69 295.21 214Pb 0.83 0.99 1.07 1.11 1.10 1.57 351.92 214Pb 0.85 0.99 1.04 1.06 1.07 1.48 401.81 219Rn 0.86 0.99 1.02 1.06 1.07 1.41 609.31 214Bi 0.89 0.99 1.00 1.04 1.07 1.22 911.21 228Ac 0.93 1.00 1.00 1.00 1.06 1.18 1001.03 234mPa 0.94 1.00 1.00 1.00 1.06 1.17 1460.83 40K 0.97 1.00 1.00 1.00 1.06 1.14 2614.53 208Tl 1.00 1.00 1.00 1.00 1.05 1.09 fR 1.000 1.001 1.004 1.001 1.035 1.205

6.2.8 Uncertainties

All uncertainties in this paper have been computed to two standard deviations

stemming from the statistics of counting only. Other sources of uncertainties including detection

efficiency, peak interference correction factors, and the uncertainties in the certified activities

were also taken into account. The latter uncertainty was specified by the supplier of the

standard materials. The individual uncertainties were combined to give the overall

uncertainties according to ISO (1995). There are a number of uncertainties assumed to be

negligible due to their fairly low values. These include uncertainties in the acquisition time, the

half-life, the sample weight, and in the sample preparation procedure. Uncertainties of

interference correction that occur at some peaks are also considered. For example, the

associated uncertainty with the activity concentration of 226Ra (u(ARa)) determined through the

rapid analysis, i.e. eq.6.2, is calculated as follows:

If: 𝐴𝑅𝑎 = [𝐶186

𝜀186 𝑚 − ( 𝐼𝑈 𝐴𝑈)] .

𝑓𝑠

𝐼𝑅𝑎 (equation. 6.2), then:

u(𝐴𝑅𝑎) = √(𝐶186 𝑓𝑠

𝜀186 𝑚 𝐼𝑅𝑎 )

2[(

𝑢(𝐶186)

𝐶186 )

2+ (

𝑢(𝜀186)

𝜀186 )

2] + (

𝐴𝑈 𝐼𝑈𝑓𝑠

𝐼𝑅𝑎 )

2[(

𝑢(𝐴𝑈)

𝐴𝑈 )

2] (6.8)

where u(AU), u(C186), and u(186) are the standard uncertainties for the activity concentrations of

235U, the count rate, and the efficiency at the 186 keV peak. Note that u(ARa) represents

140

uncertainty to one standard deviation. In this paper, the reported values are to two standard

deviations, i.e. 2 u(ARa).

6.3 Results and discussion

6.3.1 Routine analysis of NORM

The activity concentrations of the key radioisotopes were first determined for the

reference materials. The measured activity concentrations of radionuclides in the reference

materials were compared with certified values as shown in Figure 6.2. In general, the measured

values were in excellent agreement with the certified values (Table 6.4). The results of the u-

score statistic revealed no significant difference between the analyst’s and assigned values

(Table 6.4), such that the calculated u-scores were lower than 1.96 at a 95% confidence interval.

The relative biases of the reference materials were within a deviation range of -5.50% to 4.39%.

The activity concentrations of 235U and its daughters were only reported for the RGU-1

sample, as the expected values were given by IAEA. 235U was measured using its gamma lines at

the 185.7 keV peak. 227Th was measured using its gamma lines at 50.13 and 256.25 keV, while

its short-lived progeny, 223Ra and 219Rn, were measured through their gamma lines at 269.46

and 401.81 keV, respectively. The 269.46 keV peak was corrected for the interference resulting

from 228Ac and 219Rn. The results of 235U and its daughters, 227Th, 223Ra, and 219Rn, agreed with

the recommended values within the range of statistical accuracy.

The RGTh-1 standard material was used to check the activity measurements of several

radionuclides in the 232Th series (Figure 6.2(A)). The activity of 228Ac is often used to estimate

the activity concentrations of 232Th and 228Ra as well. The measured and expected values of

228Ac indicate excellent agreement with a u-score of 0.28 and a relative bias of 1.6%.

Interestingly, the 228Th activity could also be measured directly through its weak gamma line

(0.263 %) at 215.98 keV. Although the measured activity agreed with the expected value, the

uncertainty at this peak was relatively high (Figure 6.2(A)). The other short-lived daughters in

the series (namely, 212Pb, 212Bi and 208Tl), had activities comparable to the recommended values.

The activity concentrations of 40K in the reference materials, RGK-1, SS and SG, were

measured and compared with the given values (Figure 6.2(C)). The measured activities of 40K

were found to be in excellent agreement with the assigned values (the relative bias for the RGK-

1 was -1.93 % with u-score value of 0.23).

In Figure 6.2(B), ratios between the measured and expected activities of radionuclides

in the 238U decay chain in reference materials fell between 1.01 ± 0.08% and 1.08 ± 0.20% for

SS, and between 0.87 ± 0.12% and 1.05 ± 0.15% for SG. The corresponding values for the 232Th

series radionuclides, the measured activities of the SS and SG materials were between 0.91 ±

0.10% and 1.17 ± 0.11% of the expected values, respectively.

141

RGU-1

234Th

234mPa

230Th

226Ra

210Pb

235U

227Th

223Ra

Activi

ty c

once

ntr

atio

n (

Bq

kg

-1)

200

5000

Measured

Expected

± 2SD

SS and SG - 238

U series

Radionuclide

234Th

234mPa

230Th

226Ra

210Pb

Activi

ty r

atio

0.0

0.5

1.0

1.5

2.0

SS

SG

Expected

RGTh-1

228Ac

228Th

212Pb

212Bi

208Tl

Activi

ty c

once

ntr

atio

n (

Bq

kg

-1)

2500

3000

3500

4000

Measured

Expected

± 2SD

SS and SG - 232

Th series

Radionuclide

228Ac

228Th

212Pb

212Bi

208Tl

Activi

ty r

atio

0.0

0.5

1.0

1.5

2.0

SS

SG

Expected

(B) Activity ratios of several radionuclides in the uranium and thorium series for the spiked samples (SS and SG) prepared from the NBL standards. The continuous line represents a ratio of 1:00 and the error bars are the uncertainty to 2.

IAEA-434

234Th

234mPa

230Th

226Ra

210Pb

Activi

ty c

once

ntr

atio

n (

Bq

kg

-1)

0

200

600

800

1000

Measured

Expected

± 2SD

40K

Sample

SS SG RGK-1

40K

activi

ty c

once

ntr

atio

n (

Bq

kg

-1)

0

500

15000 Measured

Expected

± 2SD

142

Figure 6.2 (A) The measured activity concentration (dots) compared to the certified values (continuous line) for radioisotopes in the reference materials used in the present study, RGU-1, RGTh-1, and IAEA-434. The comparison is for uranium, thorium, and actinium radioisotopes. The uncertainty (2) in the certified values is given as dashed lines and in the measured values as error bars.

The results of the representative samples are presented in Table 6.5. In all samples, the

activity of 234Th at the 63.29 keV peak agreed with that of 234mPa obtained from the 1001.03 keV

peak. Both peaks were able to provide reliable estimations for 238U in most of the analysed

samples. However, the 63.29 keV peaks in the two highest density samples, CS and MO, did not

provide results consistent with those obtained from 234mPa. This discrepancy could be related to

the larger self-absorption effect, because of the large densities of these samples, occurring at the

63.29 keV peak than at the 1001.03 keV peak.

The 230Th activity was measured via its gamma line at 67.67 keV. Despite its lower

probability of gamma emission and higher self-absorption at this peak, the activities obtained

appeared to align well with those of other radionuclides in the series. This was true for samples

for which the equilibrium in the series was guaranteed (i.e. when the activities of 238U, 230Th,

226Ra, and 210Pb were equivalent).

The activity concentration of 226Ra was calculated as the weighted average of its short-

lived progeny 214Pb and 214Bi, after allowing four weeks for secular equilibrium. The activity of

226Ra was consistent with its daughter isotope 210Pb in all samples, indicating equilibrium in the

lower part of the series with no excess 210Pb activity.

Table 6.4 U-scores and relative bias for reference materials

Reference material Nuclide Activity concentration (Bq kg-1) Bias U-Score

IAEA Measured (%)

IAEA-434 238U 120 ±22 119 ± 8 -0.59 0.03

230Th 211 ± 18 209 ± 28 -0.73 0.05

226Ra 778 ± 124 773 ± 16 -0.67 0.04

210Pb 701 ± 120 732 ± 17 4.39 0.37

RGU-1 238U 4940 ± 60 4832 ± 150 -2.17 0.67

230Th 4967 ± 286 0.54 0.09

226Ra 4990 ± 82 1.02 0.49

210Pb 5073 ± 182 2.70 0.69

235U 228 ± 4 221 ± 10 -2.94 0.62

227Th 232 ± 12 1.58 0.29

223Ra 226 ± 4 -0.95 0.38

219Rn 215 ± 28 -5.50 0.44

RGTh-1 228Ac 3250 ± 180 3302 ± 36 1.60 0.28

228Th 3229 ± 254 -0.63 0.07

212Pb 3295 ± 22 1.38 0.25

208Tl 3317 ± 46 2.05 0.33

RGK-1 40K 14000 ± 800 13730 ± 440 -1.93 0.23

(c) A comparison of 40K measured activity concentrations and referenced values for the reference samples SS, SG, and RGK-1.

143

For the upper part of the series, an equilibrium state between 226Ra and its parent 238U

was observed in most samples (Table 6.5). However, for the TI, TII and BC samples, the activity

concentrations for the lower part of the series (i.e., 226Ra and its progeny) were higher than

those for the upper part. This implies that the equilibrium states of these samples were

disturbed suggesting that during past industrial processing uranium has been extracted from

bulk materials.

The reported activities of radionuclides in the 235U series followed the same equilibrium

pattern as those in the 238U series (Table 6.5). The 235U/238U ratio remained close to the natural

value of 0.046 in all samples. The 235U/226Ra ratio was altered in disequilibrium samples. This is

evidence that 235U atoms were also removed with the removal of 238U atoms and left the series.

In the 232Th series, equilibrium was present in all analysed samples (Table 6.5). Moreover, the

direct measurement of 228Th from its weak gamma line at 215.98 keV (0.26%) appeared to be

possible with the BEGe detector. In most samples, the measured activity of 228Th through the

215.98 keV peak was comparable to that obtained from its short-lived daughters, 212Pb, 212Bi,

and 208Tl. However, in samples with low thorium concentrations (100 Bq kg-1), such as CHI and

CHII, the 215.98 keV peak was unreliable. This is most likely related to the low probability of

gamma emissions within that line, along with other factors (e.g., self-absorption).

In the spectrum of samples with high count rates, additional interferences have been

observed. For example, the 238.63 keV peak of 212Pb overlapped with the 235.97 keV peak of

227Th and the 241 keV doublet peaks of 214Pb and 224Ra. This is because, at high count rates, the

resolution is degraded, which may broaden the peaks (Debertin and Helmer, 1988) and cause

adjacent peaks to overlap. However, the 238.63 keV peak was successfully corrected for this

interference. Furthermore, the 510.77 and 583.19 keV peaks of 208Tl were avoided due to the

complex interference among the annihilation peak at 511 keV, the 214Pb peak at 580.15 keV

(0.35%) and the 228Ac at 583.4 keV (0.11%). Instead, the 2614.5 (99 %) and 860.6 keV (12.4 %)

peaks were used. Likewise, the 235U peak at 143.8 keV overlapped with two 228Ac peaks at

145.85 (0.16%) and 141.02 keV (0.05 %). This occurred in addition to the contributions of 230Th

and 223Ra at the same peak. The correction for such interference can be risky due to the large

number of interferences and will probably introduce additional errors to the final results.

The spectral interferences that occurred at the 63, 143, and 270 keV peaks of 234Th, 235U and

223Ra, respectively, were successfully corrected in most samples. The contributions at the

interfered peaks were due mainly to radionuclides in the 232Th series, including 232Th at 63 keV

and 228Ac at the other two peaks. In the CS and MO samples, however, the results obtained from

the overlapped peaks were unreliable. These samples exhibited higher thorium contents, such

that the 232Th/238U ratio was greater than five in the CS and seven in the MO sample. However,

144

the unreliability of the correction may not be related to the 232Th/238U ratio. For example, in the

case of RGTh-1 sample, correction could be successfully applied despite the 232Th/238U ratio

exceeding 40. Other factors, such as, random summing and self-absorption would have been

responsible for this effect. These factors are discussed below.

The activity strengths of MO and CS samples are higher than the activity of the calibration

standard by a factor of 30 or more. This has an impact on the comparative method commonly

used for NORM samples analysis. The probability of random summing, which is count-rate

dependant, in these samples will be much greater than it is in the calibration standard. As a

result, there will be a gain or loss in the number of events at the peaks of interest. Another

observation in the CS and MO samples is the elevated spectral background, which originated

from Compton scattering. This can reduce the signal-to-noise ratio. Most of the overlapped

peaks lay in the region dominated by Compton scattering. Moreover, the self-absorptions of the

CS and MO samples were relatively large. Although, the random summing and self-absorption in

these samples have been corrected, the large values of such factors may add complexity in a way

that is not understandable.

The activity concentration of 40K was reported only for TI and TII samples. The level of

40K in these samples was within the global population weighted average (420 Bq kg-1)

(UNSCEAR, 2008). In the CHI and CHII samples, the activity of 40K was below the detection limit.

This is due to the natural formation of the underlying geology of the area from which the

samples were collected, which consists of basaltic rocks known to have low potassium content

and consequently low 40K. In other samples, in which the 232Th/238U ratio is high (i.e., MO and

CS), the 1460 keV peak was due mostly to 228Ac interfering with 40K. The error in 40K activity

caused by this interference was greater than 100%.

145

Table 6.5 The measured activity concentration of several radionuclides of NORM (Bq kg-1) in various samples used in the present study.

Series Nuclide CHI CHII TI TII BC CS MO Soil Soil U-tailing U-tailing Concentrate Contaminated soil Monazite 238U 234Th 1032 ± 56 666 ± 38 2006 ± 102 1976 ± 100 1295 ± 68 3460 ± 164 N.A 234mPa 946 ± 146 551 ± 102 1929 ± 228 1866 ± 222 1223 ± 150 6407 ± 566 34887 ± 3120 230Th 927 ± 74 567 ± 50 8557 ± 468 8599 ± 470 2346 ± 156 6639 ± 392 34652 ± 3420 226Ra 956 ± 50 555 ± 32 6347 ± 106 6337 ± 106 13833 ± 204 6242 ± 106 34194 ± 602 210Pb 984 ± 42 552 ± 26 5319 ± 194 5380 ± 196 15221 ± 540 6427 ± 250 33080 ± 1838 232Th 228Ac 83 ± 4 95 ± 4 167 ± 6 171 ± 6 83 ± 6 33763 ± 1208 248796 ± 5096 228Th BDL BDL 155 ± 32 55 ± 20 79 ± 20 35264 ± 2320 255584 ± 17658 212Pb 78 ± 4 96 ± 4 165 ± 6 169 ± 6 63 ± 2 35930 ± 2384 233136 ± 9324 212Bi 82 ± 10 95 ± 8 166 ± 14 156 ± 12 83 ± 8 33334 ± 1272 234724 ± 9192 208Tl 85 ± 4 96 ± 6 164 ± 10 163 ± 10 79 ± 6 33449 ± 704 235187 ± 6510 235U 235U 45 ± 4 31 ± 4 115 ± 8 119 ± 8 63 ± 6 312 ± 30 N.A 227Th 55 ± 6 29 ± 4 285 ± 18 239 ± 18 601 ± 12 240 ± 20 N.A 223Ra 44 ± 10 27 ± 10 338 ± 68 299 ± 60 646 ± 130 N.A N.A 219Rn 43 ± 10 34 ± 8 288 ± 34 267 ± 32 634 ± 66 337 ± 36 1902 ± 178 40K BDL BDL 575 ± 44 598 ± 44 BDL BDL BDL BDL = Below Detection Limit N.A = Not Applicable

146

6.3.2 Rapid measurements of 226Ra

The results of the rapid determination of 226Ra for the reference materials RGU-1 and

IAEA-434 after 1 day and 30 days of preparing the samples are illustrated in Figure 6.3(A). The

measured 226Ra activities obtained from 186 keV peak corrected through various gamma lines

of 235U and through the natural 235U/238U ratio fell within the statistical uncertainty of IAEA

values at all times. The relative bias between the measured 226Ra obtained through the rapid

approach and the IAEA values ranged from – 6.2% to 3.6% and the u-score values were found

satisfactory (u<1.96).

On the other hand, the results of 226Ra obtained from its progeny, 214Pb and 214Bi,

appeared to be dependent on the delay time required for the secular equilibrium. The activity

concentration of 226Ra, measured after a day of preparing the sample, was lower than that after

30 days. The discrepancy would depend on the loss of any 222Rn that has emanated from the

grains and released from the interstitial space during sample preparation. The results

demonstrate that measuring 226Ra soon after preparing the sample is possible using the 186 keV

peak, with results equivalent to those obtained from 214Pb and 214Bi after a waiting time of 30

days. Without this delay period, the latter method may underestimate the results. Of the

alternative suggested for rapid 226Ra measurement, the method based on 235U/238U ratio

assumes the natural activity concentration ratio of 0.046 between the two uranium isotopes.

This value may not may not be true for depleted or enriched uranium samples. The methods

based on the direct measurement of 235U peaks are not influenced either by the 235U/238U ratio

or the state of equilibrium between the uranium or actinium series.

The results of the measurements of 226Ra from the reference materials SS and SG are

plotted in Figure 6.3(B). Most of the observed data points are scattered fairly close to the line of

equality, indicating good agreement between the measured and expected values of 226Ra, with

comparable results across the examined methods. The results of the gamma line at 163.3 keV in

the SG sample were higher than the expected values by more than 30%; however, the result was

consistent with the statistical error.

147

RGU-1

Method

(a) Progeny (b) 143 keV (c) 163 keV (d) 205 keV (e) 235U(f) 238U/235U ratio

Activi

ty c

once

ntr

atio

n (

Bq

kg

-1)

3500

4000

4500

5000

5500

6000

1-2 days

After 30 days

Expected value

± 2SD

IAEA-434

Method

(a) Progeny (b) 143 keV (c) 163 keV (d) 205 keV (e) 235U(f) 238U/235U ratio

Activi

ty c

once

ntr

atio

n (

Bq

kg

-1)

400

600

800

1000

1200 1-2 days

After 30 days

Expected value

± 2SD

Figure 6.3 (A) The 226Ra activity concentration measured in the IAEA standards, RGU-1 and IAEA-434, using (a) 222Rn progeny gamma lines and the 186 keV peak after correcting it for the 235U contribution, calculated from 235U individual gamma lines at (b) 143, (c) 163, (d) 205 keV, (e) the combined average from these lines, and (f) from the 238U/235U ratio compared to the certified values.

148

Method

(a) Progeny (b) 143 keV (c) 163 keV (d) 205 keV (e) 235U(f) 238U/235U ratio

Re

lative

to

the

exp

ecte

d v

alu

e

0.0

0.5

1.0

1.5

2.0

SS

SG

Expected

Figure 6.3 (B) The 226Ra activity measured using different approaches compared to the expected values in the spiked samples.

The direct measure of 226Ra activity using the 186 keV peak was also applied to the

representative samples. The measured activity of 226Ra from each method was compared to the

activities obtained from 226Ra progeny, 214Pb and 214Bi, at equilibrium (Table 6.6). The ratios of

the activity of 226Ra, as measured through several gamma energy lines of 235U, to those of 214Pb

and 214Bi were close to unity for most samples.

The results for the 226Ra obtained via the 143 keV peak in the MO and CS samples were

unsatisfactory. Those samples were rich in 232Th; hence, two peaks of 228Ac at 145.85 (0.16%)

and 141.02 keV (0.05%) became pronounced and overlapped with the triplet peak at 143 keV

due to 235U, 223Ra, and 230Th. The complexity of the interference, along with the high Compton

continuum background and the large self-absorption factor, would lead to enormous errors at

this peak. Recall the low natural abundance of 235U and the weak gamma intensity at the 143

keV peak; the same is true for the 205.3 keV peak in the MO sample where it interferes with the

204.0 keV (0.12%) 228Ac peak. These results are consistent with those reported by De Corte et

al. (2005), where the use of 143 and 205.3 keV peaks yielded systematically high results.

The rapid determination of 226Ra appeared to be unreliable in the case of the MO sample.

The difference in the activity concentrations measured through the 226Ra progeny and measured

149

directly via 235U gamma-lines was significant, with the exception of the measurements taken at

the 163.3 keV peak. This peak resulted in an activity ratio of 1.17 ± 0.12% and was free of

interference. This is most likely due to the random summing and Compton continuum

backgrounds that result from the high count rates of this sample.

The uncertainty of the rapid 226Ra measurement is estimated to be 25% or lower. This

uncertainty arises mainly from counting statistics and calibration efficiencies. It is directly

related to the low count rates of 235U peaks, which are associated with low natural abundance,

as well as the low probability of emission of the 235U gamma lines employed in the direct

method. In addition, the interference that occurred at the targeted peaks can introduce another

source of error. This can be seen clearly in the high uncertainty associated with the results of

the 143 keV peak that interferes with 223Ra, 230Th, and in some cases, 228Ac.

The findings confirm the potential for using the gamma lines of 235U at 143.8, 163.3 and

205.3 keV for the rapid determination of 226Ra. This approach was found to be equivalent to the

conventional method of measuring 226Ra using gamma spectrometry (i.e., through 226Ra progeny

in equilibrium). However, the usefulness of the rapid approach appeared to be limited by the

systematic errors arising from the large differences in activity strengths between the sample

and the calibration standard. Due to these errors, the Compton scattering backgrounds, the

random summing in samples with high count rate spectra are likely to be elevated. High count

rates allow greater spectral interference to occur and cause complexity in correcting

interference that occurs at the peak of interest.

Table 6.6 Ratios of 226Ra activity as measured using different approaches to the ratio obtained via its progeny

Sample 143 163 205 235U 235U/238U

(keV) (keV) (keV) (weighted mean) Ratio

CHI 0.92 ± 0.10 1.00 ± 0.10 0.96 ± 0.10 0.97 ± 0.10 0.97 ± 0.10

CHII 0.93 ± 0.10 0.95 ± 0.10 1.32 ± 0.14 1.14 ± 0.12 1.05 ± 0.10

TI 1.13 ± 0.11 1.18 ± 0.08 1.09 ± 0.08 1.14 ± 0.10 1.17 ± 0.11

TII 1.11 ± 0.10 1.18 ± 0.08 1.07 ± 0.08 1.13 ± 0.10 1.19 ± 0.11

BC 1.22 ± 0.08 1.27 ± 0.08 1.28 ± 0.08 1.27 ± 0.08 1.27 ± 0.08

CS 4.26 ± 0.42 0.96 ± 0.12 1.08 ± 0.08 1.28 ± 0.11 1.17 ± 0.24

MO 7.03 ± 0.32 1.17 ± 0.12 5.44 ± 0.20 2.71 ± 0.14 0.27 ± 0.04

6.3.3 Determination of 227Ac using gamma ray spectrometry

The activity concentration of 227Ac was obtained as a weighted average of its daughter

radionuclides, 227Th, 223Ra, and 219Rn, after a waiting time sufficient for equilibrium. The

measured 227Ac activity concentration was compared with those of 235U, 238U, and 226Ra, as

presented in Table 6.7. The 227Ac/235U activity ratio was close to unity in the samples exhibiting

the equilibrium condition. In these samples, the 227Ac/238U and 227Ac/226Ra activity ratios

corresponded to the natural value of 0.046. On the other hand, the 227Ac /235U ratio was greater

150

than 2 in the IAEA-434, TI, TII, and BC samples. This confirms the disequilibrium state of the

235U series. The activity ratio of 227Ac/238U in these samples was also changed from the natural

value of 0.046, while the 227Ac/226Ra ratio remained the same. In case of samples TI, TII, and BC,

the removal of element uranium would have resulted in this behaviour. These samples have

been taken from the product and waste streams of mineral processing plants. However, the

exact reason of the IAEA-434 sample behaviour is not clear.

Figure 6.4 shows a plot of the mean activity concentration of 227Ac against the measured

activity of 235U in all samples. An excellent correlation was found between the two results in

samples that maintained equilibrium, as can be seen in the data points distributed closely to the

line of equality in Figure 6.4. When the equilibrium state was lost, the activity of 227Ac differed

significantly from that of 235U, as indicated by the outliers in Figure 6.4.

It is worth noting that under certain conditions, the use of 227Th and its daughters to

determine the activity concentration of 227Ac may be misleading due to the differences in their

chemical properties. In an acidic pH condition, for example, the activity concentration of 227Ac,

as measured through 227Th, was found to be incorrectly estimated due to the mobility of 227Th in

that condition (Patra et al., 2013).

Table 6.7 Activity ratios of 227Ac to 235U, 238U and 226Ra

Sample 227Ac/235U 227Ac/238U 227Ac/226Ra IAEA434 4.514 ± 1.108 0.208 ± 0.050 0.032 ± 0.008 RGU-1 1.062 ± 0.226 0.046 ± 0.008 0.046 ± 0.008 SS 0.859 ± 0.178 0.040 ± 0.008 0.039 ± 0.008 SG 1.068 ± 0.136 0.049 ± 0.014 0.050 ± 0.006 CHI 1.137 ± 0.150 0.052 ± 0.012 0.055 ± 0.006 CHII 0.961 ± 0.182 0.047 ± 0.012 0.053 ± 0.008 TI 2.855 ± 0.274 0.145 ± 0.024 0.045 ± 0.002 TII 2.332 ± 0.226 0.126 ± 0.022 0.038 ± 0.002 BC 10.876 ± 1.338 0.536 ± 0.088 0.049 ± 0.002 CS 0.918 ± 0.052 0.060 ± 0.012 0.040 ± 0.002 MO n.a 0.055 ± 0.010 0.056 ± 0.006

151

235U Activity (Bq kg

-1)

10 100 1000

22

7A

Activi

ty (

Bq

kg

- 1)

10

100

1000

IAEA-434

RGU-1

SS

SG

CHI

CHII

TI

TII

BC

CS

MO

Line of equality

Figure 6.4 The activity concentrations of 227Ac versus 235U in various samples examined during the present study. For most samples, the values lie along the line of equality. Samples such as TI, TII, and BC show the 227Ac:235U ratio 1. This behaviour is most likely due to extraction of uranium from these samples.

6.4 Conclusion

The BEGe detector was successfully employed for routine analyses of NORM samples.

The detector was found to be efficient for measuring radionuclides with low gamma energy and

intensity that have many applications in the field of environmental radioactivity, such as 46.5,

63.3, and 67.7 keV, of 210Pb, 234Th, and 230Th, respectively. Analysis of the same standard using a

BEGe and an HPGe coaxial detector demonstrated the superiority of the BEGe detection system

in the 20 – 200 keV range. A comparison of the activity concentrations of the analysed

radionuclides in reference materials measured by BEGe detector and assigned by the supplier

showed a good agreement with a bias range between -5.5% and 4.4% and the u-score values

were less than 1.96. Care should be exercised when using the comparative method to calibrate

the detector as the large difference in the densities and activity strengths between the sample

and the calibration standards can lead to unreliable results.

The findings demonstrated the feasibility of the rapid determination of 226Ra through

the direct measure of the 186 keV peak after subtracting the contribution of 235U at the peak.

The gamma lines at 143.8, 163.3, and 205.3 keV were succeeded, for most samples, in

152

estimating the activity concentration of 235U. The validity of this approach has been confirmed

by a statistical comparison with reference materials. It revealed that the activity concentrations

of 226Ra measured using the rapid approach agreed with the certified values to within two

standard deviations. The results suggested the use of this approach to save time in determining

the activity concentration of 226Ra in environmental samples and accomplish the analysis within

two days. However, the reliability of this approach may be limited by the large variations in the

source strength and density between the sample and the comparator, which may resulting in

systematic errors.

The activity concentration of 227Ac was successfully determined in the environmental

samples through its daughters using gamma-ray spectrometry. The results were consistent with

the expected values, as well as with the values for the parent nuclide, 235U, in samples where

secular equilibrium was attained. The drawback of this approach is, perhaps, the delay period

required for the secular equilibrium state (227Th, T1/2 = 18.7 d; 223Ra, T1/2 = 11.4 d). However,

this equilibrium may already exist in some samples.

Acknowledgements

The authors would like to thank Dr Jamie Trapp for preparing the calibration standard.

153

References

Aalseth, Craig E, M Amman, Frank T Avignone, Henning O Back, Alexander S Barabash, PS Barbeau, M Bergevin, FE Bertrand, M Boswell and V Brudanin. 2011. "Astroparticle physics with a customized low-background broad energy germanium detector." Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment 652 (1): 692-695.

Aba, A, S Uddin, M Bahbahani and A Al-Ghadban. 2014. "Radiometric dating of sediment records in Kuwait’s marine area." Journal of Radioanalytical and Nuclear Chemistry 301 (1): 247-255.

Al-Ghamdi, H, A Al-Muqrin and A El-Sharkawy. 2016. "Assessment of natural radioactivity and 137 Cs in some coastal areas of the Saudi Arabian gulf." Marine Pollution Bulletin 104 (1): 29-33.

Baskaran, M. 2011. "Po-210 and Pb-210 as atmospheric tracers and global atmospheric Pb-210 fallout: a review." Journal of Environmental Radioactivity 102 (5): 500-513.

Bojanowski, R, E Holm and NE Whitehead. 1987. "Determination of 227 Ac in environmental samples by ion-exchange and alpha spectrometry." Journal of Radioanalytical and Nuclear Chemistry 115 (1): 23-37.

Brookes, Charles Joseph, IG Betteley and SM Loxston. 1979. Fundamentals of mathematics and statistics New York: Wiley.

Budjáš, Dusan, M Agostini, L Baudis, E Bellotti, L Bezrukov, R Brugnera, C Cattadori, A di Vacri, R Falkenstein and A Garfagnini. 2013. "Isotopically modified Ge detectors for GERDA: from production to operation." Journal of Instrumentation 8 (4): P04018.

Condomines, M and S Rihs. 2006. "First 226 Ra–210 Pb dating of a young speleothem." Earth and Planetary Science Letters 250 (1): 4-10.

Cutshall, Norman H, Ingvar L Larsen and Curtis R Olsen. 1983. "Direct analysis of 210Pb in sediment samples: Self-absorption corrections." Nuclear Instruments and Methods in Physics Research 206 (1): 309-312.

Dababneh, Saed, Ektimal Al-Nemri and Jamal Sharaf. 2014. "Application of Geant4 in routine close geometry gamma spectroscopy for environmental samples." Journal of Environmental Radioactivity 134: 27-34.

Danesi, PR, A Bleise, W Burkart, T Cabianca, MJ Campbell, M Makarewicz, J Moreno, C Tuniz and M Hotchkis. 2003. "Isotopic composition and origin of uranium and plutonium in selected soil samples collected in Kosovo." Journal of Environmental Radioactivity 64 (2): 121-131.

De Corte, Frans, Herman Umans, Dimitri Vandenberghe, Antoine De Wispelaere and P Van den Haute. 2005. "Direct gamma-spectrometric measurement of the 226Ra 186.2 keV line for detecting 238U/226Ra disequilibrium in determining the environmental dose rate for the luminescence dating of sediments." Applied Radiation and Isotopes 63 (5-6): 589.

Debertin, Klaus and Richard G Helmer. 1988. Gamma-and X-ray spectrometry with semiconductor detectors. Amsterdam: North-Holland.

Desideri, D, C Roselli, MA Meli and L Feduzi. 2008. "Analytical methods for the characterization and the leachability evaluation of a solid waste generated in a phosphoric acid production plant." Microchemical Journal 88 (1): 67-73.

Doering, Che and Riaz Akber. 2008. "Beryllium-7 in near-surface air and deposition at Brisbane, Australia." Journal of Environmental Radioactivity 99 (3): 461-467.

Dowdall, M, ØG Selnæs, JP Gwynn and C Davids. 2004. "Simultaneous determination of 226Ra and 238U in soil and environmental materials by gamma-spectrometry in the absence of radium progeny equilibrium." Journal of Radioanalytical and Nuclear Chemistry 261 (3): 513-521.

Ekström, LP and RB Firestone. 2004. "WWW Table of Radioactive Isotopes: database version 2/28/99." Accessed 10/01/ 2016. http://nucleardata.nuclear.lu.se/toi/.

El-Daoushy, Farid and Francisco Hernández. 2002. "Gamma spectrometry of 234 Th (238 U) in environmental samples." Analyst 127 (7): 981-989.

154

Engelbrecht, Rudolf and Martina Schwaiger. 2008. "State of the art of standard methods used for environmental radioactivity monitoring." Applied Radiation and Isotopes 66 (11): 1604-1610.

Garcıa-Talavera, M, JP Laedermann, M Decombaz, MJ Daza and B Quintana. 2001. "Coincidence summing corrections for the natural decay series in γ-ray spectrometry." Applied Radiation and Isotopes 54 (5): 769-776.

Gilmore, Gordon. 2008. Practical gamma-ray spectroscopy. 2nd ed. Chichester: Wiley Han, J.H. and J.H. Choi. 2010. "Broad Energy HPGe Gamma Spectrometry for Dose Rate

Estimation for Trapped Charge Dating." Journal of Analytical Science & Technology 1 (2): 98-108.

Huy, Ngo Quang, Vo Xuan An, Truong Thi Hong Loan and Nguyen Thanh Can. 2013. "Self-absorption correction in determining the 238 U activity of soil samples via 63.3 keV gamma ray using MCNP5 code." Applied Radiation and Isotopes 71 (1): 11-20.

Huy, NQ and TV Luyen. 2004. "A method to determine 238U activity in environmental soil samples by using 63.3-keV-photopeak-gamma HPGe spectrometer." Applied Radiation and Isotopes 61 (6): 1419-1424.

IAEA. 2014. Radiation protection and safety of radiation sources : international basic safety standards. Vienna: International Atomic Energy Agency.

ICRP. 2012. "Compendium of dose coefficients based on ICRP Publication 60. ICRP Publication 119." Ann. ICRP 41(Suppl.) (4).

ISO. 1995. Guide to the Expression of Uncertainty in Measurement. Geneva: International Organization for Standardization.

Johnston, A and P Martin. 1997. "Rapid analysis of 226 Ra in waters by γ-ray spectrometry." Applied Radiation and Isotopes 48 (5): 631-638.

Justo, J, H Evangelista and AS Paschoa. 2006. "Direct determination of 226 Ra in NORM/TENORM matrices by gamma-spectrometry." Journal of Radioanalytical and Nuclear Chemistry 269 (3): 733-737.

Kaste, James M, Benjamin C Bostick and Arjun M Heimsath. 2006. "Determining 234 Th and 238 U in rocks, soils, and sediments via the doublet gamma at 92.5 keV." Analyst 131 (6): 757-763.

Khandaker, MU, PJ Jojo, HA Kassim and YM Amin. 2012. "Radiometric analysis of construction materials using HPGe gamma-ray spectrometry." Radiation Protection Dosimetry: ncs145.

Kolb, WA and M Wojcik. 1985. "Enhanced radioactivity due to natural oil and gas production and related radiological problems." Science of the Total Environment 45: 77-84.

Kossert, Karsten, Karen Bokeloh, Rainer Dersch and Ole Nähle. 2015. "Activity determination of 227 Ac and 223 Ra by means of liquid scintillation counting and determination of nuclear decay data." Applied Radiation and Isotopes 95: 143-152.

Landsberger, S, C Brabec, B Canion, J Hashem, C Lu, D Millsap and G George. 2013. "Determination of 226Ra, 228Ra and 210Pb in NORM products from oil and gas exploration: Problems in activity underestimation due to the presence of metals and self-absorption of photons." Journal of Environmental Radioactivity 125: 23-26.

Lavi, N, F Groppi and ZB Alfassi. 2004. "On the measurement of 40K in natural and synthetic materials by the method of high-resolution gamma-ray spectrometry." Radiation Measurements 38 (2): 139-143.

Lenka, P, SK Jha, S Gothankar, RM Tripathi and VD Puranik. 2009. "Suitable gamma energy for gamma-spectrometric determination of 238U in surface soil samples of a high rainfall area in India." Journal of Environmental Radioactivity 100 (6): 509-514.

Lozano, JC, P Blanco Rodriguez and F Vera Tome. 2001. "A simple screening method for the determination of 230Th in soil and sediment samples by gamma-spectrometry." Journal of Radioanalytical and Nuclear Chemistry 247 (1): 101-105.

Martin, P, GJ Hancock, S Paulka and RA Akber. 1995. "Determination of 227 Ac by α-particle spectrometry." Applied Radiation and Isotopes 46 (10): 1065-1070.

155

Matisoff, Gerald. 2014. "210Pb as a tracer of soil erosion, sediment source area identification and particle transport in the terrestrial environment." Journal of Environmental Radioactivity (In press).

McMahon, CA, MF Fegan, J Wong, SC Long, TP Ryan and PA Colgan. 2004. "Determination of self-absorption corrections for gamma analysis of environmental samples: comparing gamma-absorption curves and spiked matrix-matched samples." Applied Radiation and Isotopes 60 (2): 571-577.

Murray, AS, R Marten, A Johnston and P Martin. 1987. "Analysis for naturally occuring radionuclides at environmental concentrations by gamma spectrometry." Journal of Radioanalytical and Nuclear Chemistry 115 (2): 263-288.

Patra, A Chakrabarty, S Mohapatra, SK Sahoo, RM Tripathi and VD Puranik. 2013. "Radioactive series disequilibria in underground uranium deposits of the Singhbhum Shear Zone, Eastern India." Journal of Radioanalytical and Nuclear Chemistry 295 (1): 675-683.

Paytan, A, WS Moore and M Kastner. 1996. "Sedimentation rate as determined by 226 Ra activity in marine barite." Geochimica et cosmochimica acta 60 (22): 4313-4319.

Pilleyre, T, S Sanzelle, D Miallier, J Faïn and F Courtine. 2006. "Theoretical and experimental estimation of self-attenuation corrections in determination of 210Pb by γ-spectrometry with well Ge detector." Radiation Measurements 41 (3): 323-329.

Pöllänen, R, TK Ikäheimonen, S Klemola, V-P Vartti, K Vesterbacka, S Ristonmaa, T Honkamaa, P Sipilä, I Jokelainen and A Kosunen. 2003. "Characterisation of projectiles composed of depleted uranium." Journal of Environmental Radioactivity 64 (2): 133-142.

Şahiner, Eren and Niyazi Meriç. 2014. "A trapezoid approach for the experimental total-to-peak efficiency curve used in the determination of true coincidence summing correction factors in a HPGe detector." Radiation Physics and Chemistry 96: 50-55.

Saïdou, François Bochud, Jean-Pascal Laedermann, MG Kwato Njock and Pascal Froidevaux. 2008. "A comparison of alpha and gamma spectrometry for environmental natural radioactivity surveys." Applied Radiation and Isotopes 66 (2): 215-222.

Sansone, Umberto, Pier Roberto Danesi, Sabrina Barbizzi, Maria Belli, Michael Campbell, Stefania Gaudino, Guogang Jia, Rita Ocone, Alessandra Pati and Silvia Rosamilia. 2001. "Radioecological survey at selected sites hit by depleted uranium ammunitions during the 1999 Kosovo conflict." Science of the Total Environment 281 (1): 23-35.

Scholten, Jan C, Iolanda Osvath and Mai Khanh Pham. 2013. "226Ra measurements through gamma spectrometric counting of radon progenies: How significant is the loss of radon?" Marine Chemistry 156: 146-152.

Shaw, Timothy J and Willard S Moore. 2002. "Analysis of 227 Ac in seawater by delayed coincidence counting." Marine Chemistry 78 (4): 197-203.

Shuktomova, I. I. and N. G. Rachkova. 2011. "Determination of 226Ra and 228Ra in slightly mineralised natural waters." Journal of Environmental Radioactivity 102 (2): 84-87.

Ugur, Aysun, Juan-Carlos Miquel, Scott W Fowler and Peter Appleby. 2003. "Radiometric dating of sediment cores from a hydrothermal vent zone off Milos Island in the Aegean Sea." Science of the Total Environment 307 (1): 203-214.

UNSCEAR. 2008. Exposures of the public and workers from various sources of radiation, UNSCEAR 2000 report to the General Assembly with scientific annexes New York: United Nations.

Van Beek, P, M Souhaut and J-L Reyss. 2010. "Measuring the radium quartet (228 Ra, 226 Ra, 224 Ra, 223 Ra) in seawater samples using gamma spectrometry." Journal of Environmental Radioactivity 101 (7): 521-529.

Vesterbacka, P, S Klemola, K Salahel-Din and M Saman. 2009. "Comparison of analytical methods used to determine 235U, 238U and 210Pb from sediment samples by alpha, beta and gamma spectrometry." Journal of Radioanalytical and Nuclear Chemistry 281 (3): 441-448.

Völgyesi, Péter, Zoltán Kis, Zsuzsanna Szabó and Csaba Szabó. 2014. "Using the 186-keV peak for 226Ra activity concentration determination in Hungarian coal-slag samples by gamma-ray spectroscopy." Journal of Radioanalytical and Nuclear Chemistry 302 (1): 375-383.

156

Voltaggio, Mario, Marilì Branca, Paola Tuccimei and Francesca Tecce. 1995. "Leaching procedure used in dating young potassic volcanic rocks by the 226 Ra/230 Th method." Earth and Planetary Science Letters 136 (3): 123-131.

Webb, Joel L and Gary H Kramer. 2001. "An evaluation of germanium detectors employed for the measurement of radionuclides deposited in lungs using an experimental and Monte Carlo approach." Health Physics 81 (6): 711-719.

Wyttenbach, A. 1971. "Coincidence losses in activation analysis." Journal of Radioanalytical and Nuclear Chemistry 8 (2): 335-343.

Yoshida, Satoshi, Yasuyuki Muramatsu, Keiko Tagami, Shigeo Uchida, Tadaaki Ban-nai, Hidenori Yonehara and Sarat Sahoo. 2000. "Concentrations of uranium and 235U/238U ratios in soil and plant samples collected around the uranium conversion building in the JCO campus." Journal of Environmental Radioactivity 50 (1): 161-172.

Yücel, H, MA Cetiner and H Demirel. 1998. "Use of the 1001keV peak of 234m Pa daughter of 238 U in measurement of uranium concentration by HPGe gamma-ray spectrometry." Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated Equipment 413 (1): 74-82.

Yücel, H, AN Solmaz, E Köse and D Bor. 2009. "Spectral interference corrections for the measurement of 238U in materials rich in thorium by a high resolution γ-ray spectrometry." Applied Radiation and Isotopes 67 (11): 2049-2056.

Yücel, H, AN Solmaz, E Köse and D Bor. 2010. "Methods for spectral interference corrections for direct measurements of 234U and 230Th in materials by gamma-ray spectrometry." Radiation Protection Dosimetry 138 (3): 264-277.

Zapata, Felipe. 2003. "The use of environmental radionuclides as tracers in soil erosion and sedimentation investigations: recent advances and future developments." Soil and Tillage Research 69 (1): 3-13.

Zhang, W., K. Ungar, J. Chen, N. St-Amant and BL Tracy. 2009. "An accurate method for the determination of 226Ra activity concentrations in soil." Journal of Radioanalytical and Nuclear Chemistry 280 (3): 561-567.

157

Chapter 7 Summary and Conclusion

7.1 Conclusion

Despite the marked improvement in the radiation protection system, NORM is still an

evolving subject, and many challenges have to be overcome. This thesis critically reviews the

current regulation of NORM, identifies emerging challenges and provides a better

understanding of how the measurements and monitoring of NORM should be carried out to

adequately demonstrate compliance with regulatory regimes. In particular, the measurements

used in quantifying the activity concentration of radionuclide members of NORM in solid

materials, as well as the release of radon from material surfaces and its concentration in indoor

air, were investigated.

The current radiological protection system of NORM was comprehensively reviewed. A

number of regulatory issues were highlighted and discussed, including: the absence of a

complete framework; the interference between planned and existing exposure situations; the

lack of a homogeneous approach; the ubiquity of NORM; the source of low doses; societal

attitudes; measurement challenges; waste and residues, and the ethical approach. The absence

of a complete framework, along with the interference that might occur between planned and

existing exposure situations, has created inconsistency in the regulation of NORM. This has led

to confusion among users of NORM guidelines. More clarity may be required to eliminate such

confusion. In addition, societal attitudes toward natural sources of radiation have influenced the

decision in terms of radiation protection against NORM, resulting in less attention being paid to

it. The impact of the measurement (which is necessary to implement and comply with the

regulation) on the management of NORM is obvious. Improving NORM measurement techniques

is still needed to address the challenges associated with it.

The review of the regulation of NORM has also shown that each industry involving

NORM has its own merits. It is vital from a regulation point of view, therefore, to characterise

exposure to NORM and its potential pathways in each industry on a case-by-case basis.

Implications of the implementation of the new radon and thoron policies were also discussed

from economic and practical perspectives. Further steps need to be taken by national

authorities to develop and implement a national radon protection strategy to control indoor

radon exposure. Characterising such exposure cannot be accomplished without quantifying

radon levels in the air, and its exhalation flux from surfaces.

The measurement of radon flux density from surfaces using the activated charcoal

canister technique was examined. The distribution of the adsorbed 222Rn in the charcoal bed of

the measured canister was found to be slightly different from that in the standard canister, and

158

became more uniform with longer exposure times. This is because calibration standards were

prepared by a uniform mixing of 226Ra in the charcoal matrix, while in the field 222Rn was

adsorbed more readily in the charcoal layer closer to the exposed surface. A correction factor

was worked out by differences in the counting geometries between the uniformly distributed

standards and the measured canisters with non-uniform distribution. Its value for one day of

exposure, which recorded the lowest degree of uniformity, was at 14% or less. The influence of

exposure time on the measured 222Rn activity flux was also investigated. The measured value of

the activity flux was found to decrease as exposure time increased. The reduction was

attributed to the back diffusion of 222Rn from the canister through a canister-soil interface, and

possibly the absorption of water molecules by the activated charcoal. These findings introduced

limitations to the methodology, which need to be taken into account when interpreting the

measurements.

The levels of indoor radon (222Rn) and thoron (220Rn) were measured in various

buildings representing workplaces in Brisbane CBD, Australia. The average 222Rn and 220Rn

concentrations for all workplaces were 10.5 ± 11.3 Bq m−3 and 8.2 ± 1.4 Bq m−3 respectively. The

highest 222Rn concentration was found in basements and confined areas, while the maximum

220Rn level was detected in storage rooms. Poor ventilation in these locations is most likely the

main cause of the increased concentration levels. Diurnal variation in 222Rn level was observed

in the workplaces-concentrations were lower during working hours due to the use of air

conditioners and natural ventilation during this time. Changes in 220Rn concentration were

insignificant during and after working hours. The study extended to investigating the

parameters that could influence indoor 222Rn and 220Rn concentrations. 222Rn was more

influenced by floor level and flooring type, while 220Rn was most affected by the distance from

the source of origin. The findings suggest that care should be exercised when measuring 222Rn

concentrations in workplaces, as its concentration during and after working hours differ

considerably. The study demonstrated that indoor 222Rn and 220Rn levels are influenced by the

parameters examined in this study.

The suitability of the BEGe detector for routine analysis of NORM was investigated. The

detector was found to be useful for the activity concentration measurement of natural

radionuclides and was capable of detecting peaks with low probabilities of gamma emissions at

lower energy regions, such as 46.5, 50.1, 63.3 and 67.7 keV, of 210Pb, 227Th, 234Th and 230Th,

respectively. The detector was used for the rapid determination of 226Ra through the direct

measure of the 186 keV peak. Although this approach succeeded in most cases, its reliability is

limited by the systematic errors arising from the large variations in activity strength and

density between the sample and the calibration standard. The outcome of this study is

159

encouraging, and the detector has a promising future in a variety of applications in the field of

environmental radioactivity.

7.2 Future Direction

This thesis has shown the importance of understanding the measurement and

regulation of NORM, and has opened new horizons for future research projects on the topic. It

has been shown that the regulation of NORM within the relevant industries has to be treated on

a case-by-case basis. Information about NORM in non-nuclear industries is scarce, so

investigation of different types of industries and industrial processes would be beneficial in

assessing the radiological impact of NORM on a wide scale. The regulatory challenges associated

with NORM that have been identified in this thesis should be addressed and taken into account

in future recommendations and policies. Additionally, the influence of measurement issues (e.g.,

the cost of the analysis, uncertainties of the measurement, the background radiation of

detectors, and the state of radioactive equilibrium) on adequate compliance with NORM

regulations needs further investigation. A particularly important example is the development of

a cost-effective method for measuring certain radionuclides of NORM in specific environmental

media.

The findings of each measurement technique investigated in this thesis highlight a

number of further works that are yet to be explored:

The study of measuring radon activity flux density using an activated charcoal canister

outlined the limitations of the method owing to exposure time and the distribution of

radon in the charcoal bed. A study of the effects of the charcoal type and charcoal grain

size on radon adsorption would provide a better understanding of the adsorption

efficiency of the charcoal.

The measurements of radon and thoron in workplaces provide the average

concentrations of these gases but do not determine radon and thoron progeny

distribution in indoor air. A comprehensive approach that includes a time series of

radon and thoron, as well as their progeny, with suitable modelling and analysis

techniques would allow for a better radiological assessment of exposure to radon and

thoron in indoor environments.

The BEGe detector was used only for the measurement of solid NORM samples.

Investigation of the use of the detector for the analysis of filters and water samples

would be beneficial for a comprehensive analysis of NORM encountered in industrial

situations.

The results of these research projects will significantly improve the measurement and

regulation of NORM.