HEALTH EFFECTS OF LOW DOSES OF RADIATION

NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
NCRP DRAFT SC 1-21 COMMENTARY
(Council and Board Draft) (3-31-15)
HEALTH EFFECTS OF LOW DOSES OF
RADIATION: INTEGRATING RADIATION
BIOLOGY AND EPIDEMIOLOGY
March 31, 2015
Note: Copyright permission is being sought for the figures and tables requiring such permission prior to
their use in the final NCRP publication.
National Council on Radiation Protection and Measurements
7910 Woodmont Avenue, Suite 400, Bethesda, Maryland 20814
1
NCRP SC 1-21
Draft of March 31, 2015 [MR]
2
NOT TO BE DISSEMINATED OR REFERENCED
Preface
3
4
In order to gain a greater understanding of the biological interactions and health effects of low
5
doses of ionizing radiation, the National Council on Radiation Protection and Measurements
6
(NCRP) has increased its activities on this subject in recent years. The NCRP 2008 Annual Meeting
7
was held on the subject of Low Dose and Low Dose-Rate Radiation Effects and Models (NCRP,
8
2009a). In December 2008 NCRP held a workshop that involved 30 participants with expertise in the
9
effects of low doses of radiation. After the workshop summary was completed,1 a decision was made
10
to convene a panel of experts to provide advice on a new NCRP Commentary related to critical
11
issues and research needs for gaining a better understanding of effects of low doses of radiation. The
12
advisory panel met in August 2010, and NCRP began the preparation of a Commentary in 2012. The
13
focus of the Commentary is on integration of results of basic science studies, including biomarkers
14
and bioindicators of cancer and other diseases, with epidemiologic studies on health effects of low
15
doses of radiation. The Committee members have expertise and experience in the following areas
16
related to low doses of ionizing radiation: epidemiology, radiation biology, radiation oncology,
17
biostatistics, health physics and dosimetry.
18
19
This Commentary was prepared by Scientific Committee 1-21 on Multiplatform National
20
Program for Providing Guidance on Integrating Basic Science and Epidemiological Studies on Low-
21
Dose Radiation Biological and Health Effects.
22
23
Serving on Scientific Committee 1-21 were:
24
25
26
27
28
29
30
31
32
Sally A. Amundson, Co-Chair
Columbia University Medical Center
New York, New York
Jonine L. Bernstein, Co-Chair
Memorial Sloan-Kettering Cancer Center
New York, New York
1
NCRP (2008). National Council on Radiation Protection and Measurements. Developing a Framework for
Incorporating New Information in the Radiation Sciences to Understanding Potential Health Risks for Low Doses of
Ionizing Radiation: Summary of NCRP Workshop held on December 1-2, 2008 (National Council on Radiation
Protection and Measurements, Bethesda, Maryland).
2
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
Members
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
John D. Boice, Jr.
Vanderbilt University
Nashville, Tennessee
Raymond A. Guilmette
Lovelace Respiratory Research
Institute
Albuquerque, New Mexico
Amy Kronenberg
Lawrence Berkeley National
Laboratory
Berkeley, California
Mark P. Little
National Cancer Institute
Bethesda, Maryland
William F. Morgan
Pacific Northwest National Laboratory
Richland, Washington
Jac A. Nickoloff
Colorado State University
Fort Collins, Colorado
Simon N. Powell
Memorial Sloan-Kettering Cancer Center
New York, New York
Daniel O. Stram
University of Southern California
Los Angeles, California
Consultant
R. Julian Preston
U.S. Environmental Protection Agency
Research Triangle Park, North Carolina
NCRP Secretariat
Marvin Rosenstein, Staff Consultant (2013 –)
Terry Pellmar, Staff Consultant (2012)
Cindy L. O’Brien, Managing Editor
Laura J. Atwell, Office Manager
David A. Smith, Executive Director (2014 –)
James R. Cassata, Executive Director (2012 – 2014)
The Council expresses appreciation to the Committee members for the time and effort
51
devoted to the preparation of this Commentary. NCRP also gratefully acknowledges the financial
52
support provided by the Centers for Disease Control and Prevention.
53
54
John D. Boice, Jr.
55
President
56
3
NCRP SC 1-21
Draft of March 31, 2015 [MR]
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
NOT TO BE DISSEMINATED OR REFERENCED
Contents
Preface ............................................................................................................................................2
Executive Summary ......................................................................................................................6
1. Introduction ................................................................................................................................9
1.1 Low Doses and Low Dose Rates ......................................................................................11
1.2 Use of Radiation Biology Data to Reduce Uncertainty in Risk Estimates ........................12
2. General Approaches to Risk Assessment ..............................................................................19
2.1 Risk Assessment Process and Associated Uncertainties ...................................................19
2.2 Modeling Uncertainties: Dosimetric, Epidemiologic and Biologic ...................................20
2.3 Uncertainties Related to Transfer of Radiation Risk between Populations and
Interactions with Carcinogens ...........................................................................................22
2.4 Extrapolating Low-Dose Effects from In Vitro to In Vivo; Animal to Human ................23
3. Studies Integrating Biology and Epidemiology ....................................................................27
3.1 Epidemiologic Studies Focused on Cancer Outcomes .....................................................27
3.1.1 Chromosome-Aberration Studies ............................................................................27
3.1.2 Atomic-Bomb Survivors ..........................................................................................28
3.1.3 Pelvic Irradiation Studies .........................................................................................30
3.1.4 Occupational Studies ...............................................................................................31
3.1.5 Environmental Background Studies ........................................................................33
3.1.6 Transgenerational Studies ........................................................................................34
3.1.7 Gene and Radiation Interaction: The Women’s Environmental Cancer and
Radiation Epidemiology Study ................................................................................35
3.1.8 Retinoblastoma ........................................................................................................36
3.2 Epidemiologic Studies Focused on Noncancer Outcomes ...............................................37
4. Bioindicators from Radiation Biology Studies .....................................................................39
4.1 Deoxyribonucleic Acid Damage and Repair ....................................................................39
4.2 Cellular Signaling at Low Doses ......................................................................................40
4.3 Chromosome Alterations ..................................................................................................41
4.4 Mutations ..........................................................................................................................42
4.5 Radiation-Induced Genomic Instability ............................................................................46
4.6 Modulators of Response ...................................................................................................49
4.6.1 Adaptive Responses .................................................................................................49
4.6.2 Bystander Effects .....................................................................................................51
4.6.3 Genetic Susceptibility and Interactions with Radiation ...........................................52
4.6.4 Interactions with Environmental and Lifestyle Factors ...........................................53
4
NCRP SC 1-21
Draft of March 31, 2015 [MR]
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
NOT TO BE DISSEMINATED OR REFERENCED
5. Recommendations for Closing the Gaps towards an Integrated Approach ......................55
5.1 Targeted Research to Identify Adverse Outcome Pathways and Bioindicators of
Response ...........................................................................................................................55
5.2 Biologically-Based Modeling ...........................................................................................57
5.2.1 Biologically-Based Dose-Response Approach ........................................................58
5.2.2 Systems Biology ......................................................................................................60
5.3 Registries of Medical-Imaging Exposures Linked to Outcomes and Biological
Materials ............................................................................................................................62
5.4 Facilities and Interdisciplinary Training ...........................................................................63
5.4.1 Facilities ....................................................................................................................63
5.4.2 Interdisciplinary Training ........................................................................................63
Glossary .......................................................................................................................................64
Symbols, Abbreviations and Acronyms ....................................................................................70
References ....................................................................................................................................71
5
NCRP SC 1-21
Draft of March 31, 2015 [MR]
116
NOT TO BE DISSEMINATED OR REFERENCED
Executive Summary
117
118
For decades, epidemiologic studies have assessed the health effects of radiation exposure from
119
multiple sources: occupational, accidental, environmental and medical. These efforts have been most
120
successful in evaluating radiation-related adverse effects (particularly cancer) at or above ~100 mGy
121
(absorbed dose). Since there are greater uncertainties inherent in epidemiologic studies of exposed
122
individuals at lower doses, it is generally agreed that the effects observed at 100 mGy and above are
123
the more reliable than those observed at <100 mGy. Despite a history of over 70 y, the field of
124
radiation risk estimation has not, or perhaps more correctly, has not been able to make much direct
125
use of radiation biology data from laboratory animal, cellular and molecular studies. Rather, such
126
data have been used extensively in a supportive role for the epidemiology-based risk estimates.
127
128
The current state of knowledge on the effects of low-dose and low dose-rate radiation (defined
129
for the purpose of this Commentary as a dose <100 mGy and a dose rate <5 mGy h–1) is incomplete
130
and uncertain with respect to understanding the shape of the dose-response relationship and the level
131
of risk at low doses. These deficiencies have long-term and profound consequences with regard to
132
radiation protection guidance, compensation programs, and environmental contamination issues.
133
There is a need to identify novel approaches to reduce these uncertainties.
134
135
The focus of this Commentary is on identifying further means to integrate results of basic
136
science studies in radiation biology, including biomarkers and bioindicators of cancer and other
137
diseases, with epidemiologic studies on health effects of low doses of radiation. A biomarker is a
138
biological phenotype (e.g., chromosome alteration, DNA adduct, gene expression change, specific
139
metabolite) that can be used to indicate a response at the cell or tissue level to an exposure. A
140
bioindicator is a cellular alteration that is on the pathway to the disease endpoint itself (a key event),
141
such as a specific mutation in a target cell that is associated with tumor formation. Towards this end,
142
the Commentary reviews general approaches to risk assessment (Section 2), existing epidemiologic
143
studies that have incorporated biological endpoints (Section 3), and areas of radiation biology with
144
potential relevance to low-dose risk (Section 4).
145
6
NCRP SC 1-21
Draft of March 31, 2015 [MR]
146
NOT TO BE DISSEMINATED OR REFERENCED
Given major advances in our understanding of the etiology of diseases, host susceptibility, and
147
the cellular processes affected by radiation, coupled with the rapid development of new
148
technologies, there now is the opportunity to integrate information from multiple disciplines in risk
149
assessment. For example, radiation biology data can be incorporated into the process of
150
extrapolating from epidemiologic data at higher doses to predict responses at low doses (<100 mGy)
151
and low dose rates (<5 mGy h–1). The key to successfully doing this is to identify and select
152
appropriate endpoints that inform the process of extrapolation of risk from the higher doses to low
153
doses. Such biological endpoints utilized for extrapolation purposes need to be predictive of the
154
apical endpoint for which risk is being estimated (i.e., a bioindicator predictive specifically of
155
radiation-related cancer or noncancer outcome). While some approaches have been suggested, a key
156
problem is the lack of bioindicators (key events) that are specific for radiogenic disease. It is not yet
157
feasible to assess the robustness of key events in the radiation risk assessment process because too
158
little reliable information has been developed.
159
160
Biologically-based dose-response models have been suggested as a way forward for low-dose
161
risk estimation. However, developing such models requires a fuller understanding of the relevant
162
biology than we currently have, as well as the collection of sufficient data for parameterization of the
163
models. Awareness of such modeling needs should be integrated with experimental design, to ensure
164
the collection of relevant data.
165
166
Radiation-induced genomic instability is one example of a specific biological response that may
167
have relevance to modeling of low-dose radiation risk. It is important to determine if specific types
168
of genomic instability are capable of transforming a cell from a normal to a cancer cell, or furthering
169
the progression of a premalignant cell towards a more aggressive cancer phenotype. Chromosome
170
instability has been shown to persist for as long as a year, but there is little information concerning
171
how cells pass the unstable phenotype through succeeding somatic cell generations.
172
173
Another gap in our understanding of radiation risk is the role of individual variation in genetic
174
susceptibility and interaction with radiation. Clinical studies over the years have identified genetic
175
disorders and a few high penetrance alleles that convey increased risk of radiation-related cancer and
176
other endpoints. Such alleles are often demonstrated in studies of second cancers, but the magnitude
7
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
177
of effect is unknown and likely to vary. As each person is exposed to a different array of lifestyle
178
and environmental factors in addition to any low-dose radiation exposure, the potential impact of
179
interactions between these factors, genetic factors, and radiation exposure must also be considered.
180
181
The nascent field of systems biology, in which global information from multiple levels (e.g.,
182
transcriptomic, proteomic, metabolomic) is integrated to develop a clearer picture of a process, for
183
example, response to radiation, may also prove informative for risk prediction. A systems level
184
understanding of radiation response could lead to a more comprehensive understanding of the
185
mechanisms involved, and the interactions between mechanisms, thus supporting more robust
186
biologically-based models of risk.
187
188
The following recommendations are made for closing the gaps towards an integrated approach of
189
basic science studies in radiation biology with epidemiologic studies on health effects of low doses
190
of radiation:
191
192

focus on key events (i.e., bioindicators) and modifying factors in adverse outcomes of
193
ionizing radiation exposure [as outlined in the parallelogram approach presented in
194
Figure 1.1 (for mechanistic studies) and Figure 1.3 (specific for bioindicators)], rather than
195
simple biomarkers of exposure;
196

197
198
develop biologically-based dose-response models to provide a path forward in low-dose
radiation risk assessment;

assemble registries of patients with extensive exposure to current and emerging medical-
199
imaging procedures (particularly those associated with higher patient doses) linked with
200
biological specimens and outcome data, in order to provide for standardized research
201
planning and data collection for prospective risk assessment;
202

203
204
maintain and develop the specialized facilities required for studies conducted at low doses
and low dose rates; and

promote and expand interdisciplinary training and integrated cross-professional research
205
programs devoted to understanding and quantifying radiation health effects at low doses in
206
order to meet the growing needs of radiation science for the next generation of radiation
207
researchers.
8
NCRP SC 1-21
Draft of March 31, 2015 [MR]
208
NOT TO BE DISSEMINATED OR REFERENCED
1. Introduction
209
210
We live in an era of increased public exposure to ionizing radiation, from the proliferation of its
211
use in diagnostic and interventional medical procedures to its nonmedical use for security screening.
212
With an increasing need for nuclear power, we also see continuing concern surrounding the public
213
exposures that may result from accidents, as dramatically illustrated by the recent Fukushima-
214
Daiichi disaster. In such an environment, meaningful risk assessment for low-dose and low dose-rate
215
ionizing radiation exposures is more critical than ever for the improvement of radiation protection,
216
environmental protection and remediation, and compensation. A fundamental issue in deriving risks
217
at low levels of dose and dose rate is the shape of the dose response for endpoints relevant to human
218
health outcomes and possible modifications by dose rate. This Commentary outlines a framework to
219
integrate data from biological studies with epidemiologic data to illuminate this issue. This
220
framework is broadly based on a parallelogram approach for mechanistic studies shown in
221
Figure 1.1 and made more specific by the use of bioindicators of response described in Figure 1.3
222
(Section 1.2). The bioindicators proposed are developed from an adverse outcome pathway/key
223
events approach initially described in Section 1.2 (Preston, 2015).
224
225
While epidemiologic studies remain the gold standard for determining risk and setting exposure
226
limits, a number of practical considerations limit the usefulness of such studies in the low-dose
227
range. In recent years, however, considerable progress has been made in low-dose radiation biology
228
at the cellular and molecular levels. Ideally, such work should contribute to our understanding of risk
229
and should help shape radiation protection guidance. In order to effectively integrate the results of
230
radiobiology studies into epidemiology and risk analysis, it is necessary to make multiple
231
comparisons, from cell and molecular studies to in vivo studies, and from animal to human on both
232
the cell and molecular level, and on the in vivo level (Figure 1.1).
233
234
Conceptually, this approach has proven useful particularly at the cellular level (e.g., Brewen
235
et al., 1973). The intercomparisons illustrated in Figure 1.1 are critical for developing a robust
236
analysis of the risk of human exposure to low-dose radiation. Studies with biological models that
237
more closely approximate tissue architecture are of particular use as a bridge between simple cellular
238
studies in vitro and organismal studies (Bissell et al., 1997). Application of extrapolation to the
9
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
239
240
241
242
Fig. 1.1. A parallelogram approach for utilizing laboratory animal and in vitro cell data to
estimate human cancer and noncancer risks for ionizing radiation.
243
10
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
244
assessment of human risk has been reviewed in NCRP Report No. 150 (NCRP, 2005), which
245
concluded that an understanding of the molecular mechanisms of radiation carcinogenesis in
246
different animal species was needed for the most robust extrapolation. Although species-specific
247
differences in critical biological processes need to be kept in mind, a focus on conservation of
248
mechanism will be useful when making intercomparisons to inform risk in humans.
249
250
The major goal of this Commentary is to develop a framework to support the useful
251
incorporation of biological endpoints into epidemiologic studies in order to enhance our
252
understanding of the risks of low-dose and low dose-rate ionizing radiation exposures.
253
254
Towards this end, this Commentary reviews general approaches to risk assessment, areas of
255
radiation biology with potential relevance to low-dose risk, and existing epidemiologic studies that
256
have incorporated biological endpoints. The Commentary suggests that focusing on key events and
257
modifying factors for adverse outcomes pathways from ionizing radiation exposure, rather than
258
simple biomarkers of exposure, in concert with the development of biologically-based dose-response
259
models will provide the best path forward in low-dose radiation risk assessment. Finally, we identify
260
the major gaps in knowledge that need to be addressed to enable the integration of basic biology and
261
epidemiologic studies for the advancement of low-dose ionizing radiation risk assessment and
262
radiation protection.
263
264
1.1 Low Doses and Low Dose Rates
265
266
Studies have been conducted over a broad range of radiation doses and dose rates referred to as
267
low, and what constitutes a low dose and a low dose rate is often due to the context in which dose
268
and dose rate are being considered (e.g., radiobiological research, epidemiologic studies, medical
269
exposure, environmental exposure). For the purposes of this Commentary, the context is the levels
270
experienced by individuals or populations to various sources of ionizing radiation, and single values
271
of low dose and low dose rate are useful for general discussion.
272
273
In assessing human health effects (in particular cancer), ICRP (1991) applied a DDREF of two to
274
obtain risk estimates for absorbed doses below 200 mGy and from higher doses when the dose rate is
11
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
275
<100 mGy h–1. The U.K. National Radiological Protection Board (NRPB) concluded that for the
276
purposes of applying a dose-rate effectiveness factor in radiation protection, dose rates of
277
<0.1 mGy min–1 (when averaged over ~1 h) and acute doses of <100 mGy may be regarded as low
278
(Muirhead et al., 1993). This dose rate is equivalent to 6 mGy h–1. Wakeford and Tawn (2010) more
279
recently agreed with the NRPB guidance and suggested that a low dose is <100 mGy delivered
280
acutely and a low dose rate is <5 mGy h–1. The rationale for selecting such dose rates appears to be
281
driven by the desire to link them to what is believed to be the linear portion of a dose-response
282
relationship. A dose of <100 mGy and a dose rate of <5 mGy h–1 are used for low doses and low
283
dose rates in this Commentary.
284
285
It is recognized that these specified values for low dose and low dose rate are most relevant to
286
exposure scenarios related to occupations in which sources of radiation and radioactive material are
287
used. In comparison, ubiquitous background dose rates tend to occur at levels of ~1 µGy h–1, which
288
is a factor of 5,000 less than the value of 5 mGy h–1 used here. Additionally, regulatory limits for
289
doses to members of the public from environmental radiation sources are typically set at levels that
290
equate to a fraction of the ubiquitous background dose rate [e.g., the NRC limit for members of the
291
public from nuclear reactor sites is 1 mSv y–1 (total effective dose equivalent) or 0.11 µSv h–1].
292
These dose rates assume chronic exposure to relatively constant radiation fields. In any case, it is
293
difficult to discern whether differences in dose rate on the order of 103 to 104 can be understood in
294
terms of mechanisms of radiation action, given our current state of knowledge.
295
296
1.2 Use of Radiation Biology Data to Reduce the Uncertainty in Risk Estimates
297
298
There is an extensive literature on the health effects of radiation exposures (occupational,
299
accidental, environmental and medical) in human populations. Many of these involve relatively high
300
acute doses, although associations at lower doses (around a few tens of milligray) and lower dose
301
rates have been reported. Given the greater uncertainties inherent in epidemiologic studies of
302
exposed individuals at low doses than at higher doses, it is generally agreed that effects observed at
303
~100 mGy and above are the more reliable than those observed at < 100 mGy. In this regard, despite
304
a history of over 70 y, the field of risk estimation has not, or perhaps more correctly, has not been
305
able to make much direct use of radiation biology data from laboratory animal, cellular and
12
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
306
molecular studies. Rather such data have been used in a more supportive role for the epidemiology-
307
based risk estimates.
308
309
Risk estimates for radiation-related cancer and noncancer effects at low doses (chronic and
310
acute) have relied almost exclusively on epidemiologic data. Exceptions are for the estimation of
311
DDREF and relative biological effectiveness of high-LET radiation that have relied quite
312
significantly on data from biological studies, although also supported by epidemiologic data. Given
313
the major advances in our understanding of the etiology of chronic diseases (especially cancer,
314
although with an increasing emphasis also on certain noncancer diseases) and of the fundamental
315
cellular processes affected by radiation (both targeted and nontargeted), it is prudent to consider how
316
this information can be incorporated into the risk assessment process. In particular, it is critical that
317
the use of data from biological studies be incorporated into the process for extrapolating from
318
epidemiologic data at medium/high doses to predict responses at low doses (<100 mGy) and low
319
dose rates (<5 mGy h–1) (Section 1.1). Given these limitations, two issues to be addressed herein are:
320
321

322
323
324
how can informative endpoints be selected that can enhance the process of extrapolation of
“risk” from high/medium to low doses; and

how can data from laboratory animal, cellular and molecular studies best be integrated with
epidemiologic studies to better predict risk.
325
326
Assistance with the task of estimating the health effects of radiation exposures at low doses and
327
low dose rates can be found in the approach being developed for estimating risks from exposures to
328
chemicals (e.g., Boobis et al., 2006; EPA, 2005; Julien et al., 2009; Preston, 2015; Seed et al., 2005;
329
Thomas et al., 2012; Vinken, 2013). The use of mechanistic and other biological data is necessary
330
for the risk assessment process for chemicals because of the lack of epidemiologic data for almost all
331
chemicals (environmental or occupational exposure). The process is built upon a framework of
332
adverse outcome pathways and key events for the development of adverse health effects following
333
exposure to a chemical or chemical mixture. An adverse outcome pathway for both cancer and
334
noncancer endpoints is defined as a sequence of key events and processes starting with interaction of
335
an agent with a cell, proceeding through operational and anatomical changes, and resulting in
336
formation of an adverse outcome (EPA, 2005). It should be noted that this adverse outcome
13
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
337
pathway/key event approach has not been used for an entire risk assessment process for any
338
chemical. A worked example has recently been developed by Adeleye et al. (2014).
339
340
For ionizing radiation, for example, cancer is considered to be produced by a DNA-reactive
341
mode of action. A key event is defined as an empirically observable precursor step that is itself a
342
necessary element of the mode of action or is a biologically based marker for such an element (EPA,
343
2005). Preston (2015) has recently presented a proposed adverse outcome pathway and associated
344
key events for low doses and low dose rates of ionizing radiation. It is more difficult to utilize such a
345
general approach for noncancer effects given the wide range of mechanisms whereby adverse
346
outcomes (e.g., neurological, reproductive, cardiovascular) can be produced. However, an initial
347
approach is described by Seed et al. (2005) for a range of chemical classes and noncancer endpoints.
348
349
Preston and Williams (2005) developed a set of key events by which a DNA-reactive chemical
350
can produce a metastatic cancer. A modified version of this scheme can be considered as being
351
applicable to ionizing radiation. As a starting point for the development of an adverse outcome
352
pathway for radiation-related cancer, an initial key event is considered to be a DNA double-strand
353
break, if in turn it can be converted into a mutation. The set of key events in Figure 1.2 is a linear
354
progression from the initial key event to a malignant tumor (i.e., an adverse outcome pathway)
355
(Preston, 2015). It is possible to expand this linear approach to include radiation effects on
356
preneoplastic lesions that may have developed prior to exposure. Another approach considers
357
radiation effects on any of the critical processes that are considered hallmarks of cancer development
358
(Boss et al., 2014; Hanahan and Weinberg, 2011).
359
360
For the purpose of maximizing the use of biological data in the dose-response assessment for
361
adverse health outcomes, it is informative to distinguish between biomarkers and bioindicators. A
362
biomarker is a biological phenotype (e.g., chromosome alteration, DNA adduct, gene expression
363
change, specific metabolite) that can be used to indicate a response to an exposure at the cell or
364
tissue level. In this regard, it is generally a measure of the potential for development of an adverse
365
outcome such as cancer (e.g., a predictor of exposure level). A bioindicator is defined as a cellular
366
alteration that is on the pathway to the disease endpoint itself, such as a specific mutation in a target
367
cell that is associated with tumor formation. Thus, a bioindicator can be perceived as informing on
14
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
368
Adverse Outcome
Pathway Steps
Interaction with Radiation
Energy Deposition
Macro-Molecular Alterations
Key Events

Exposure of Target
Tissue

Single, double and multiple DNA
breaks
Base Modifications
Protein Oxidation
Free Radical Formation
Chromosome Alterations








Gene Activation
Protein Production
Altered Signaling
Cell killing and Tissue
Disruption
Organ Responses



Altered Physiology
Disrupted Homeostasis
Altered Tissue
Development/Function
Adverse Outcome



Impaired Development
Impaired Reproduction
Cancer and Noncancer
Effects
Cellular Responses
369
370
Fig. 1.2. Schematic representation of an adverse outcome pathway for ionizing radiation-related
371
cancer and noncancer diseases showing each step along the proposed pathway and the associated key
372
events (Preston, 2015).
373
15
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
374
the shape of dose-response curve for the disease outcome or on cancer frequency itself, and
375
therefore, is equivalent to a key event. In a logical way, the more proximate a key event/bioindicator
376
is to the apical endpoint (e.g., cancer), the more predictive of this outcome it will be. Thus, for the
377
purposes of risk assessment, it is most pertinent to design research studies to assess bioindicators of
378
response rather than simply biomarkers of an exposure. In addition, it is essential to describe the
379
predicted human relevance of key events/bioindicators developed in nonhuman assay systems
380
(Boobis et al., 2006). Figure 1.1 can be modified (Figure 1.3) to show how bioindicators can be used
381
as the data input for a typical parallelogram approach (Sobels, 1993). The types of informative
382
bioindicators for cancer risk assessment include: cancer-specific (target tissue) chromosome
383
alterations and gene mutations, target tissue cell killing and regenerative cell proliferation, and target
384
tissue inflammatory responses. Genovese et al. (2014) has provided data that suggest that “a subset
385
of genes that are mutated in patients with myeloid cancers is frequently mutated in apparently
386
healthy persons; these mutations may represent characteristic early events in the development of
387
hematologic cancers.” This type of DNA sequencing approach can be extended to investigate other
388
389
390
cancer types.
391
and/or modulating factors can influence the nature of the overall response for both cancer and
392
noncancer outcomes. Consideration of such events and factors is particularly important for
393
394
395
quantitative assessments of dose response.
396
host factors that can modulate the dose-response relationship of one or more key events, thus altering
397
the probability or magnitude of the adverse outcome. As such, some modulating factors might be
398
considered as “inevitable” (fixed) because all humans are subject to their influence; examples are
399
sex, age, and ethnicity. Other modulating factors are more aptly described as “optional”
400
(controllable): these include smoking, diet, and occupation. It is relatively straightforward to
401
incorporate the inevitable modulators into a risk assessment process but much more problematic to
402
adequately account for the optional ones without considering individual risk values. Laboratory
403
animal and cellular and molecular data might be developed that can account for the magnitude of the
404
influence of modulating factors on risk but currently this is very rarely done and frequently it is not
405
possible.
Key events provide the critical input for dose-response modeling, so-called associated events
A modulating factor can be considered as a biological feature, including control mechanisms or
16
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
406
407
408
409
410
Fig. 1.3. Modified Figure 1.1, showing how bioindicators can be used as the data input for a
parallelogram approach (Sobels, 1993).
411
17
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
412
413
An associative event is a biological process that is not itself a key event but is a reliable
414
biomarker for a key event(s). Thus, an associative event can be a biomarker for exposure or closely
415
linked to the mechanism of disease formation. Recent research has identified two novel classes of
416
events, epigenetic changes and nontargeted events [e.g., bystander effects (Section 4.6.2) and
417
genomic instability (Section 4.5)] that are sometimes considered as associative events. However, at
418
present it is not clear whether epigenetic events are associative, but based on the evidence to date it
419
appears unlikely that nontargeted events are associative, since not all radiation exposures are
420
associated with instability in human subjects (Tawn et al., 2000; 2011; 2015; Whitehouse and Tawn,
421
422
423
2001).
424
health effects, it is necessary to utilize or develop some form of a biologically-based dose-response
425
model. Several approaches that have been described suggest the viability of this type of modeling
426
(Eidemuller et al., 2011; Kaiser et al., 2014; Little, 2010; Luebeck et al., 2013; Moolgavkar and
427
Knudson, 1981; Moolgavkar and Luebeck, 1992; Moolgavkar and Venzon, 1979; Shuryak et al.,
428
2010), although perhaps limited in their value to reliably predict cancer risk because they do not
429
incorporate any biological data other than the generalized mutational evidence supporting the
430
multistage model. Some recent approaches have perhaps provided a greater reliance on auxiliary
431
biological data (Heidenreich and Rosemann, 2012; Heidenreich et al., 2013). The further
432
development of truly informative models along these lines has to be a research priority given the
433
importance of accurate estimates of radiation-related health effects at low doses and low dose rates.
To fully use such key event/bioindicator data to inform dose-response assessment for adverse
434
435
The discussion herein describes an approach for integrating data from radiation biology and
436
epidemiology in the dose-response assessment phase of risk assessment for radiation-related cancer
437
and noncancer effects (Morgan and Bair, 2013). It proposes the development and application of
438
mode of action/key events/bioindicators as providing a link between epidemiology and radiation
439
biology data at the animal, cell and molecular levels.
440
18
NCRP SC 1-21
Draft of March 31, 2015 [MR]
441
442
443
NOT TO BE DISSEMINATED OR REFERENCED
2. General Approaches to Risk Assessment
2.1 Risk Assessment Process and Associated Uncertainties
444
445
Risk estimates for radiation-related cancer mortality and incidence and those for noncancer
446
diseases can play a critical role in the development of dose limits used for radiation protection
447
purposes. A recent example of the use of risk estimates for cancer is provided by the approach for
448
establishing the standards used by NASA for limits on days in space for astronauts (Cucinotta et al.,
449
2013; NA/NRC, 2012). The estimates of radiation risk currently used by international and national
450
bodies such as the International Commission on Radiological Protection (ICRP), the United Nations
451
Scientific Committee on the Effects of Atomic Radiation (UNSCEAR) , the National
452
Academies/National Research Council (NA/NRC), National Council on Radiation Protection and
453
Measurements (NCRP), and the U.S. Environmental Protection Agency (EPA) rely heavily upon
454
epidemiologic data on cancer and noncancer diseases in a variety of exposed populations
455
(particularly atomic-bomb survivors and people with estimated or known doses from occupational,
456
medical and environmental exposures). The general approach to risk assessment is to extrapolate
457
from health effects measured at medium to high doses to predict those expected at low doses.
458
Currently, for cancer this extrapolation utilizes a linear-nonthreshold model (ICRP, 2007; NA/NRC,
459
2006; NCRP, 2001). The estimates obtained in this way are corrected for a DDREF to allow for
460
calculated reductions in effect from high to low dose and from high to low dose rate. There are
461
challenges to this approach largely because of the reliance on the linear-nonthreshold model and the
462
uncertainty in a chosen value for DDREF. For noncancer diseases, the approach employed by ICRP
463
(2007; 2012) is to calculate a dose for a practical threshold level of disease response (<1 % chance of
464
occurrence) (ICRP, 2012) based on effects at medium to high doses. Because of the reliance on
465
epidemiologic data, it is extremely important to utilize the most recent and relevant data for
466
developing risk estimates and to take advantage of enhancements in computational modeling
467
approaches. However, at the same time, it is essential to appreciate that any approach used will
468
inevitably have associated uncertainties that should be considered when applying risk estimates for
469
protection purposes. NCRP Report No. 171 (NCRP, 2012) considers the types and magnitude of the
470
several uncertainties that are a component of the risk assessment process for cancer, noncancer
471
effects and heritable effects following radiation exposure. Consideration of this issue is timely
19
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
472
because new data have recently become available for cancer incidence, noncancer occurrence
473
(particularly for cataracts and cardiovascular disease) and heritable effects (reviewed in ICRP, 2007;
474
2012; UNSCEAR, 2001). NCRP (2012) builds upon the analyses in other recent NCRP reports
475
(NCRP, 2007; 2009b; 2009c) of the sources and magnitude of uncertainties in the estimation of
476
organ doses from exposure to external and internal sources of radiation. Given the heavy reliance on
477
epidemiologic studies, a significant component of uncertainty associated with these risk estimates
478
will be accounted for by comprehensive assessments of uncertainties in these epidemiological
479
studies. It is of note that little direct use has been made of the extensive data from ~70 y of
480
radiobiological study in estimating radiation-related health risks (ICRP, 2007; NA/NRC, 2006;
481
NCRP, 2005); the data have generally provided support for the epidemiology-based approach
482
currently used. The most significant uses of laboratory animal and cellular data have been for
483
developing a value for DDREF and values for quality factors and radiation weighting factors used in
484
radiation protection. However, even with the use of these experimental data, uncertainties associated
485
with these factors are large, and contribute significantly to the overall uncertainty in risk estimates
486
(Cucinotta et al., 2013). Sections 2.2 through 2.4 consider the types of uncertainty, their magnitude
487
488
489
490
491
and potential approaches for reducing these uncertainties.
492
(2012), the precision of epidemiologic risk estimates is driven primarily by the:
2.2 Modeling Uncertainties: Dosimetric, Epidemiologic and Biologic
Epidemiology is by its nature observational rather than experimental. As discussed in NCRP
493
494

range and distribution of doses;
495

sample size;
496

duration of and ages at observation;
497

baseline frequencies of the health endpoint of interest;
498

“true” strength of the radiation-disease association;
499

various types of dose uncertainties; and
500

degree of accuracy of ascertainment (or, conversely, misclassification or underascertainment)
501
of the disease of interest.
502
20
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
503
However, the degree of systematic errors (i.e., bias) of the epidemiologic risk estimates depends on a
504
different set of factors, such as:
505
506

507
systematic personal reporting errors (with respect to dose-related items, disease experience,
or risk-factor covariates);
508

insufficient statistical adjustment for other risk factors;
509

dose-related inequalities in disease ascertainment;
510

errors in assigning average values for shared dosimetry factors (e.g., an incorrect estimate of
511
the magnitude of a radioactive release);
512

failure to correct for individual measurement errors; or
513

failure to adjust for the effects of disease-related covariates.
514
515
The importance of these sources of uncertainty varies from study to study.
516
517
Uncertainties and biases resulting from design and methodologic issues in epidemiologic studies
518
have been considered by a number of scientific committees (NCRP, 2012; UNSCEAR, 2008a). Of
519
special importance are biases, that is, any process at any stage in the conduct of the study that tends
520
to produce results or conclusions that differ systematically from the truth (Sackett, 1979), and
521
include:
522
523

follow-up bias:
524

ascertainment bias:
525

recall bias; and
526

confounding.
527
528
Statistical methods to deal with the complexities of measurement and ascertainment error corrections
529
are still evolving (Carroll et al., 2006). A new generation of epidemiologic studies that incorporate
530
biologic measures for estimating radiation risk will help sharpen assessments and correct for some
531
uncertainties and biases.
532
21
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
533
2.3 Uncertainties Related to Transfer of Radiation Risk between Populations
534
535
536
and Interactions with Carcinogens
Despite the relatively large number of data on radiation risk, the question of how to transfer risk
537
estimates derived from one population to a different population remains unanswered. It should not
538
be surprising that the relationship between radiation-related and underlying risk in different
539
populations is not consistent for different cancer sites. The available data suggest that there is no
540
541
542
simple solution to the problem (Muirhead et al., 1993; UNSCEAR, 1994),
543
transfer of risk between populations, or indeed the use of some sort of hybrid approach, such as that
544
employed by Little et al. (1999a) and Muirhead and Darby (1987), may be appropriate. The principal
545
scientific committees implicitly assume that risk transfer is intermediate between additive and
546
multiplicative (ICRP, 2007; NA/NRC, 2006). The subject is discussed in NCRP (2012), highlighting
547
the complexity of the evidence, so that for example for breast cancer a more nearly additive transfer
548
549
550
was suggested, whereas for liver cancer relative risk transfer was more likely to be valid.
551
exposed group to another, it is necessary to also consider the influence of factors other than radiation
552
on the cancer rates. Thus, much of environmental, nutritional and occupational cancer epidemiology
553
is concerned with identifying risk factors that might account for some part of the variation of site-
554
specific underlying cancer rates among populations. While there has been much progress, the
555
problem is complex and there is only limited information on the interaction between radiation dose
556
and lifestyle, or constitutional factors, in terms of cancer risk. Thus it is likely that, for the
557
foreseeable future, the most useful information relevant to transferring radiation-related risk
558
coefficients from one population to another will come from multinational comparisons of site-
559
specific radiation-related risk, rather than from investigations of underlying cancer risk factors and
560
their interactions with radiation dose. There has been some work comparing aggregate radiation-
561
associated cancer mortality risks in humans and in various animals (Carnes et al., 2003), suggesting
562
that a simple scaling of radiation-related mortality may be appropriate. However, without more
563
detailed investigation taking account of the different spectrum of tumors in humans and in these
564
experimental systems, and their distinct etiology, this approach may not be very productive.
The available information suggests that, depending on circumstances, relative or absolute
When considering the issue of how to transfer site-specific cancer rates from one population or
22
NCRP SC 1-21
Draft of March 31, 2015 [MR]
565
566
NOT TO BE DISSEMINATED OR REFERENCED
Sometimes a meta-analysis, based on published findings from several studies, may be performed.
567
Where feasible, as noted below, it is preferable to combine the original data and analyse them using
568
a common format, in other words to perform a pooled analysis. Pooled analyses have been
569
conducted of various cohorts of radiation workers (Cardis et al., 2007) to assess:
570
571

572
effects of radon daughter exposure in relation to lung cancer risk in underground miners
(UNSCEAR, 2000);
573

thyroid cancer risk in various (mainly medically exposed) cohorts (Ron et al., 1995); and
574

breast cancer risk in various populations (Howe and McLaughlin, 1996; Little and Boice,
575
1999; Preston et al., 2002).
576
577
Less commonly, analyses combining cohort and case-control data, for example in relation to
578
579
580
581
582
leukaemia risk (Little et al., 1999b), have been conducted.
583
events that occur in response to acute moderate to high doses of radiation, including activation of
584
DNA damage checkpoints, DNA repair pathways, mutagenesis, genomic instability, cell
585
transformation, and the regulation of cell death pathways (Hall and Giaccia, 2011). The radiation-
586
induced genetic changes observed in cells in vitro are similar to those observed in radiation-induced
587
tumors in animal models. As a result, studies of model genetic loci are frequently extrapolated to
588
cancer relevant genetic changes seen in animals and humans. Such genetic events include oncogene
589
activation via chromosome translocation or gene amplification and inactivation of tumor suppressor
590
genes via point mutations or larger scale changes such as deletions, insertions and inversions.
2.4 Extrapolating Low-Dose Effects from In Vitro to In Vivo; Animal to Human
Several decades of laboratory research have given us a detailed view of molecular and cellular
591
592
The complexities of extrapolation between animal models and humans has been reviewed in
593
depth in NCRP Report No. 150 (NCRP, 2005), which concluded that studies of life shortening in
594
animals may provide reliable estimates of radiation risks to humans. The strength of life shortening
595
data is that it integrates more completely the detrimental effects of radiation in the entire animal,
23
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
596
rather than focusing on just one endpoint such as cancer induction. NCRP (2005) further concluded
597
that risk extrapolation from animal models to humans will:
598
599

600
require robust, well annotated databases that must be widely disseminated to investigators
interested in radiation risk estimation;
601

include experimental data from animal studies and data from human epidemiologic studies;
602

further develop our understanding of molecular mechanisms of cancer and other disease
603
604
605
induction; and

account for effects of age and genetic heterogeneity (or lack thereof) of animal models versus
humans.
606
607
While the benefits of life shortening studies are clear, the most recent study that validated risk
608
estimates for humans using life shortening measures in mice and dogs demonstrated that there was
609
little effect on life span with acute radiation doses <1 Gy (Carnes et al., 2003). Earlier studies of life
610
shortening in mice indicated that acute radiation doses reduced life span by ~28 d Gy–1 (Grahn,
611
1960). However, with protracted, low dose-rate exposures (<0.1 Gy d–1), life shortening was greatly
612
reduced to ~4 d Gy–1 (Grahn and Sacher, 1968). Therefore, despite the advantages of life shortening
613
as an integrated endpoint of the detrimental effects of radiation, this approach does not appear to
614
have sufficient sensitivity to detect low-dose or low dose-rate effects on animal life span, and even
615
with strong tools to extrapolate risk from animal models to humans (Carnes et al., 2003) this lack of
616
sensitivity poses problems when attempting to use these animal data to extrapolate human risk to
617
618
619
low-dose or low dose-rate radiation exposures.
620
many challenges due to significant environmental and genetic differences among cells in culture
621
versus cells in animals or humans. For example, long-term studies of cells in culture have
622
historically focused on cancer cell lines, which are not only different from the normal cells from
623
which they were derived, but are often genetically unstable. Studies of normal (primary) cells in
624
culture are typically limited to ~50 cell divisions, and immortalization by expression of telomerase
625
reverse transcriptase (TERT, hTERT in humans) to maintain telomeres, introduces another departure
626
from normal tissues. Cells in culture are generally studied as monolayers, which is far different from
Extrapolations of in vitro cell culture results to animals and to human cancer risks also pose
24
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
627
their in vivo three dimensional (3D) organization. The development of 3D artificial tissue models
628
has provided an important experimental bridge between traditional monolayer models and animal
629
models, and there is clear evidence of differential signaling and other cellular responses to radiation
630
in two dimensional (2D) monolayer versus 3D tissue models (Barcellos-Hoff and Costes, 2006). For
631
example, 3D tissue models have been used to investigate transcriptome and secretome changes in
632
response to low- versus high-dose radiation (Mezentsev and Amundson, 2011; Zhang et al., 2014),
633
and there is evidence of reduced high-dose radiation cytotoxicity in long-term 3D cultures versus 2D
634
cultures, reflecting differential apoptotic responses in 2D versus 3D cultures despite similar effects
635
on checkpoint delay (Sowa et al., 2010). In addition, cell signaling in vitro may also differ markedly
636
from the behavior of the parent cell in its normal tissue environment within an animal, where
637
638
639
communication among cells of different types may contribute to the response to radiation.
640
understanding low-dose radiation effects in humans, but inbred mice may not accurately reflect
641
effects in outbred human populations with diverse histories of radiation (and other genotoxin)
642
exposures, diet and lifestyles. There is clear evidence of variation in radiation responses in human
643
cells [e.g., cells that are heterozygous for ataxia-telangiectasia mutated (ATM), or the retinoblastoma
644
susceptibility gene (Fernet et al., 2004; Kato et al., 2009; Wilson et al., 2010a)], but at least in mice,
645
ATM heterozyosity does not enhance radiation carcinogenesis (Mao et al., 2008). The idea that low-
646
dose ionizing radiation may have significantly different effects in outbred populations is further
647
supported by a study of metabolic markers in outbred mice exposed to 100 mGy ionizing radiation
648
(Lee Do et al., 2012), and a study of mammary tissue responses to four weekly doses of 75 mGy
649
(Snijders et al., 2012). Similarly, Okayasu et al. (2000) and Yu et al. (2001) demonstrated clear
650
differences in (high-dose) radiation carcinogenesis in different strains of mice, tracing the effects to
651
hypomorphic mutations in the DNA-PKcs gene, a key factor in nonhomologous end joining repair of
652
ionizing radiation-induced double-strand breaks (DSBs). As rodents express DNA-PKcs at ~20-fold
653
lower levels than humans, this highlights yet another point to be considered when extrapolating
654
655
656
experimental results in mice to assess risk in humans.
657
mechanism. For instance, transgenerational effects have been seen in progeny of whole-body
658
irradiated male mice (Barber et al., 2002; Carls and Schiestl, 1999) but there is little evidence of
Animal studies represent a critical translational bridge between in vitro studies and
Extrapolation of risk from animal models to humans depends on conservation of the underlying
25
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
659
transgenerational effects in exposed human populations (Dauer et al., 2010; Little et al., 2013),
660
suggesting that at least for this endpoint, there are fundamental mechanistic differences between
661
animal models and humans. Particularly pertinent to extrapolation of risk from animals to humans
662
with regard to effects of low doses and low dose rates are inverse dose-rate effects, which have been
663
described in several species (mouse, rat, human cells) and several endpoints (cytotoxicity,
664
mutagenesis, tumor induction) (Brenner, 1994; Lubin et al., 1995; Spiess and Mays, 1973; Ullrich,
665
1984), characterized by greater biological effects of equal doses delivered at low versus high dose
666
rates. While data on inverse dose-rate effects are limited, these effects have been explained in terms
667
of differential checkpoint responses at low- versus high-dose rates. Thus, low dose rates do not
668
activate checkpoint arrest, and low doses absorbed by cycling cells result in greater biological effects
669
per unit dose than higher dose rates which cause arrest and thereby reduce the number of cells in
670
sensitive S and G2 phases of the cell cycle (Hall and Giaccia, 2011). The generality of inverse dose-
671
rate effects across species and endpoints argues for conserved mechanisms, and such effects should
672
be further investigated in animal models to refine risk estimates of low-dose and low dose-rate
673
radiation exposures in humans.
674
26
NCRP SC 1-21
Draft of March 31, 2015 [MR]
675
NOT TO BE DISSEMINATED OR REFERENCED
3. Studies Integrating Biology and Epidemiology
676
677
There are numerous epidemiologic studies documenting radiation-associated cancer
678
(UNSCEAR, 2008a) and certain noncancer effects (UNSCEAR, 2008b). Section 3.1 provides
679
examples of a few studies that successfully incorporate biological measures, sound epidemiologic
680
methods and radiation measures, which is the main focus of this Commentary. Section 3.2 discusses
681
studies of various noncancer diseases and effects, concentrating in particular on circulatory disease
682
and cataracts. Emerging epidemiologic studies focused on chronic low-dose radiation, such as the
683
Million Worker Study (Boice, 2012; 2014a; Bouville et al., 2015), incorporate radiation dosimetry
684
with biological and epidemiologic data, and are being designed to examine both cancer and
685
noncancer outcomes. Although currently in preliminary phases, the integrative approach combining
686
multiple measures, including individual dosimetry measures, and the planned analyses of multiple
687
outcomes promise improvements over prior studies designed in singular contexts.
688
689
3.1 Epidemiologic Studies Focused on Cancer Outcomes
690
691
3.1.1 Chromosome-Aberration Studies
692
693
A small number of radiation epidemiologic studies have also incorporated measures of
694
chromosome aberrations on a subset of the study population, in particular for support of dose
695
estimation.
696
697
Chromosome aberrations in blood lymphocytes are the most thoroughly studied biological
698
sentinel of radiation-induced cellular injury. Unstable aberrations such as rings and dicentrics are
699
used to estimate exposures in the short term, and stable aberrations such as symmetrical
700
translocations are used to estimate exposures many years after they occurred (Ainsbury et al., 2011;
701
Kleinerman et al., 2006; Tucker, 2008; Tucker and Luckinbill, 2011). Epidemiologic observations
702
that have incorporated chromosome-aberration evaluations include atomic-bomb survivors (Kodama
703
et al., 2001), patients treated with radiation for ankylosing spondylitis (Buckton, 1983), cervical
704
cancer (Kleinerman et al., 1989), benign gynecological disorders (Kleinerman et al., 1994), enlarged
705
thymus gland or tonsils (Kleinerman et al., 1990); patients receiving diagnostic radiation for
27
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
706
tuberculosis monitoring (Kleinerman et al., 1990; Littlefield et al., 1991) or given Thorostrast for
707
cerebral angiography (Littlefield et al., 1997); workers in radiation occupations (Bhatti et al., 2008;
708
Little et al., 2014a; Tucker et al., 1997a); and populations living in areas of high natural background
709
radiation (Wang et al., 1990).
710
711
While chromosome aberrations in lymphocytes are indirect indicators of radiation dose and thus
712
correlated with cancer risk, there is evidence that chromosome aberrations may predict human
713
cancer risk independently of such exposure and thus be relevant to the mechanistic and biological
714
understanding of the carcinogenic process (Bonassi et al., 2008; Hagmar et al., 2004; Tucker et al.,
715
1997b). Examples are given below for several epidemiologic investigations where concurrent studies
716
of chromosome aberrations have provided insights into the possible mechanisms of carcinogenesis
717
and other human health detriments.
718
719
3.1.2 Atomic-Bomb Survivors
720
721
The principal atomic-bomb survivor cohorts consist of the LSS, which provides follow-up for
722
mortality and cancer incidence of about 90,000 survivors with dose estimates, and the Adult Health
723
Study (AHS), which provides biennial medical follow-up for a subcohort of initially about 20,000
724
LSS members. These two studies are primary sources for investigating the relationship between
725
radiation exposure and cancer and noncancer outcomes. It is from the LSS, for example, that we
726
know that most types of cancer can be associated with radiation exposure, and that increased cancer
727
risk from an acute radiation exposure continues for more than 60 y (through 2009) from initial
728
exposure (Grant et al., 2014; Hsu et al., 2013; Ozasa et al., 2012; Preston et al., 2007).
729
730
There are few known biological effects or lifestyle variables apart from cancer, mortality from
731
various causes, and the standard socio-demographic variables (e.g., age, sex and city) determined in
732
the full LSS cohort. Much more comprehensive biological and lifestyle information was collected in
733
the AHS. The AHS includes a large fraction of LSS members with moderate to high radiation
734
exposures, as well as, for control purposes, survivors who were essentially unexposed. Within this
735
study group a large number of effects and biomarkers have been assessed, many of which have
736
focused on noncancer outcomes. The following presents an overview of those atomic-bomb survivor
28
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
737
studies that have incorporated chromosome aberrations as a biological endpoint and/or a biomarker
738
of exposure. A significant fraction of the AHS members (~4,000 total) have been assessed for stable
739
chromosome aberrations, evaluated in lymphocytes either by traditional Giemsa methods or more
740
recently by fluorescent in situ hybridization. Stable chromosome aberrations (primarily
741
translocations) are clearly related to radiation dose but there is large inter-individual variability in
742
response to dose (Kodama et al., 2001).
743
744
Significant excesses of leukemia have been observed among children and adults exposed to the
745
Hiroshima and Nagasaki atomic bombs (Hsu et al., 2013), but not among survivors who were
746
exposed in utero (Jablon and Kato, 1970). Studies of chromosome aberrations mirror these
747
epidemiologic observations. There was little evidence of dose response for chromosome aberrations
748
among atomic-bomb survivors exposed in utero, although there was a small but significant increase
749
in chromosome-aberration rate at <100 mGy (weighted fetal dose)2, whereas dose-response
750
relationships of the expected magnitude were seen for translocation frequencies in lymphocytes of
751
the mothers of those exposed in utero (Ohtaki et al., 2004). Murine experimental data (using
752
B6C3F1 mice) confirmed the human observational data, indicating that chromosome aberrations in
753
lymphocytes do not persist after an in utero dose of 1 to 2 Gy (x rays) (Nakano et al., 2007). A
754
slightly different, although essentially compatible result was observed in a Sprague-Dawley rat
755
model, where mammary cells from animals irradiated in utero with a dose of 2 Gy (gamma rays)
756
appeared to show the normal chromosome-aberration dose response, but hemato-lymphoid spleen
757
cells did not (Nakano et al., 2014).
758
759
The lack of a dose response at high doses for chromosome aberrations among those exposed in
760
utero may provide insights into the differences seen in excess risks among those exposed in utero
761
and during early childhood for leukemia. However, it should be noted that for solid cancers there is
762
an appreciable risk following in utero exposure, manifest in adulthood, which is similar to that
763
among those survivors exposed in early childhood (Preston et al., 2008). As one possible explanation
764
for these human and experimental observations, Nakano et al., (2007) suggested that the abnormal
765
cells in the fetus were replaced by normal fetal stem cells during the postnatal growth period.
766
Another possibility, suggested by Nakano et al. (2014) is that following relatively high doses, the
2
The estimated fetal dose from gamma rays plus 10 times the estimated fetal dose from neutrons, presented in milligray.
29
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
767
high degree of induced cell sterilization will leave too small a pool of hematopoietic stem cells for
768
any malignancy (i.e., leukemia) to be observed, given the relatively low expected mutation rate (10–5
769
to 10–4 per gene per gray). If this phenomenon is true in general, it might suggest that the fetus is less
770
vulnerable to the oncogenic effect of high-dose radiation exposure than the child, but that this may
771
not be the case at lower doses. However, the apparent contrast between the null findings in relation
772
to leukemia following in utero exposure in the atomic-bomb survivors and in the various studies of
773
leukemia following diagnostic obstetric exposure (Bithell and Stewart, 1975; Harvey et al., 1985;
774
Monson and MacMahon, 1984; Stewart et al., 1956) should not be overstated, since the results are in
775
fact statistically compatible (Wakeford and Little, 2003).
776
777
In addition to the atomic-bomb survivors’ cohort, other studies (Sections 3.1.3 through 3.1.8)
778
have incorporated chromosome aberration measures to evaluate further whether these markers are
779
predictive of dose and possibly cancer risk.
780
781
3.1.3 Pelvic Irradiation Studies
782
783
Significant excesses of leukemia have been observed in women treated with pelvic radiation
784
therapy for cervical cancer (Boice et al., 1987) and for nonmalignant gynecological bleeding
785
disorders (Sakata et al., 2012). Following radiation treatments for cervical cancer, the leukemia risk
786
increased up to doses of ~4 Gy, and then decreased at higher levels. A similar pattern was seen for
787
radiation-induced chromosomal aberrations suggesting that cell killing might predominate over
788
transformation at very high doses (Kleinerman et al., 1989; Littlefield et al., 1991). Excess leukemia
789
was also reported following lower dose partial-body radiation therapy for benign gynecological
790
disease. However, the level of risk was nearly the same as seen among the cervical cancer patients,
791
about twofold, despite a tenfold difference in average dose to active bone marrow (8 versus 0.7 Gy,
792
respectively) (Kleinerman et al., 1994). High-dose cell killing was again postulated as one
793
explanation for this apparent inconsistency. Remarkably, the rate of stable aberrations, which
794
reflects nonlethal damage in surviving stem cells, was also similar among the cancer patients and
795
benign gynecological disease patients. Apparently, the lower-dose treatments for benign disorders
796
resulted in much higher chromosomal-aberration yields per unit dose than seen following radiation
797
therapy for cervical cancer. Assuming that cytogenetically aberrant stem cells that survive radiation
30
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
798
therapy contribute to the leukemogenic process, the chromosome-aberration data in these two
799
irradiated populations are consistent with the leukemia outcomes in the two populations. In these
800
circumstances of partial-body radiation therapy, stable aberrations many years after exposure appear
801
to be indicators of effective risk rather than biomarkers of radiation dose.
802
803
3.1.4 Occupational Studies
804
805
Studies of chromosome aberrations among working populations with accurate measures of
806
chronic radiation exposures can provide insights into the ability of human cellular systems to repair
807
radiation damage incurred at a low dose rate. Understanding human biological responses to radiation
808
exposure has implications concerning radiation protection guidelines where a DDREF is applied to
809
scale the high dose-rate exposure risks among atomic-bomb survivors to low dose-rate exposure
810
circumstances (ICRP, 2007). Studies of workers have indicated that stable chromosome aberrations
811
can be detected and related to the amount of radiation experienced over time (Evans et al., 1979;
812
Little et al., 2014a; Tucker et al., 1997a).
813
814
The ability to detect a biological effect at the chromosome level does not mean that there will be
815
a biological consequence in the population exposed, and it may be informative to contrast
816
chromosome findings in studies for which effects have and have not been demonstrated. In studies of
817
Chernobyl cleanup workers from Estonia, increases in chromosome aberrations (Littlefield et al.,
818
1998) and in leukemia have not been observed (Rahu et al., 2013), suggesting that the population
819
size may have been too small and the doses too low to demonstrate effects following protracted
820
exposure had there been any. Other studies of low-dose occupational exposure and chromosome
821
aberrations suggested a consistency with the cytogenetic data among atomic-bomb survivors (Little
822
et al., 2014a) and there is also evidence for an increase in leukemia risk in the radiologic
823
technologist population, albeit based only on number of years worked before 1950, when doses were
824
highest (Liu et al., 2014).
825
826
Studies of workers at Sellafield nuclear facilities have revealed increased chromosome
827
aberrations that are dose related (Tucker et al., 1997a; Tawn et al., 2004; 2006). A separate study
828
found increased leukemia occurrences (Muirhead et al., 2009) that are also dose related. A similar
31
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
829
dose-related excess of chromosome aberrations has been observed in the Mayak nuclear workers
830
(Burak et al., 2001), at a rate lower than, albeit statistically compatible with, those in the Japanese
831
atomic-bomb survivors (Kodama et al., 2001). This parallels a dose-related excess risk of leukemia
832
(Shilnikova et al., 2003) in that cohort, also compatible with that in the LSS (Hsu et al., 2013). The
833
cytogenetic data, however, indicated that DDREF for chromosome aberrations was about 6,
834
suggesting to the authors that “radiation-induced double-strand breaks may be repaired more
835
efficiently at low dose rates, either because there are fewer breaks competing for a limited number of
836
repair enzymes or because chronic exposure results in a higher steady state of induction of these
837
same enzymes.” However, the parallels between the independent studies of dose-related
838
chromosome aberrations and of dose-related leukemia in these studies are only suggestive, as there
839
is no linkage between the two; there is no evidence that the leukemia cases themselves had elevated
840
chromosome aberration rates before disease development. Inferences on DDREF for a disease
841
endpoint in humans should not be drawn from these data.
842
843
Studies of Chernobyl clean-up workers provide limited evidence for chromosome damage
844
among workers whose mean whole-body doses were on the order of 100 mGy (Jones et al., 2002;
845
Littlefield et al., 1997; Moore and Tucker, 1999). This may simply reflect the fact that excess stable
846
translocations cannot be detected at much lower doses than 100 mGy (Tucker and Luckinbill, 2011).
847
Related to this, among liquidators exposed at higher doses there is unambiguous evidence of excess
848
translocations (Sevankaev et al., 2005). The evidence for a leukemia excess in large cohorts of
849
Chernobyl recovery workers also remains equivocal. While recent reports suggest an increase in the
850
incidence of leukemia among the recovery operation workers (UNSCEAR, 2011; Zablotska et al.,
851
2013), the limitations of these studies are severe and include low participation rates, proxy
852
interviews for dose reconstruction, uncertainties in dose reconstruction and ascertainment of cases,
853
and internal inconsistencies. An example is a high radiation risk for chronic lymphocytic leukemia, a
854
malignancy that is not generally found to be increased in other exposed populations (UNSCEAR,
855
2008a), although there is significant (p < 0.05) excess in the latest LSS incidence data (Hsu et al.,
856
2013), a possibly fragile finding based on 12 cases. This suggests potential biases or confounding
857
factors (UNSCEAR, 2011). The chromosome-aberration data, although limited because of the low
858
doses involved, add additional caution to the interpretation of the reports of leukemia among
32
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
859
Chernobyl recovery workers, and confirm that the worker exposure was generally relatively low and
860
the risk of exposure-related cancer accordingly low (Jones et al., 2002).
861
862
3.1.5 Environmental Background Studies
863
864
There are excess chromosome translocations in populations exposed to discharges from the
865
Mayak nuclear plant, living along the Techa River (Vozilova et al., 2012), at a rate lower than, albeit
866
statistically compatible with those in the Japanese atomic-bomb survivors (Kodama et al. 2001).
867
This group also exhibits excess risk of leukemia (Krestinina et al., 2013), consistent (per unit dose)
868
with those in the LSS (Hsu et al. 2013) and as observed also in various chronically exposed groups
869
(Ivanov et al., 2012, Zablotska et al., 2013).
870
871
Excess prevalence of micronuclei has been observed in a group of people exposed to 137Cs
872
contamination at near background dose rates from a research reactor in Taiwan (Jen et al., 2002).
873
Another population in Taiwan was exposed to prolonged, near background dose-rate 60Co
874
contamination, and again excess prevalence of chromosome translocations and deletions were
875
observed (Hsieh et al., 2002). This population has also been followed up for cancer, and significant
876
excess incidence of leukemia, and a borderline significant excess incidence of breast cancer
877
observed (Hwang et al., 2008), in both cases consistent with, although slightly less than, risks that
878
can be derived by linear extrapolation from the Japanese atomic-bomb survivors. Reduced fertility
879
has also been observed in this cohort, with fertility decreasing according to dose rate, especially of
880
mothers, but not in relation to cumulative dose (Lin et al., 2010); fertility recovered to normal among
881
those who moved out of the 60Co contaminated buildings.
882
883
In contrast, studies of populations living in areas of high natural background radiation have
884
failed to reveal convincing evidence for increases in leukemia, thyroid cancer or other malignancies
885
(Nair et al., 2009; Tao et al., 2012; Wang et al., 1990), yet increased chromosome aberrations are
886
observed in the resident populations (Hayata et al., 2004; Wang et al., 1990). Such data, limited by
887
both sample size and low doses, indicate caution when interpreting the biological significance of
888
chromosome aberrations in circulating lymphocytes as well as the absence of an effect in exposed
889
populations. Nonetheless, the absence of a leukemia response and a lower chromosome-aberration
33
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
890
yield than anticipated from studies of acute exposures (Tucker et al., 1997a) conceivably might be
891
attributable to a decrease in the tumorigenic effectiveness of low-LET radiation with decreasing dose
892
rate, as has been observed for tumorigenic effects on many tissues in laboratory animals (NCRP,
893
1980; Upton, 1990).
894
895
3.1.6 Transgenerational Studies
896
897
To date, no radiation-related genetic effects (i.e., hereditary diseases) have been
898
demonstrated in a human population exposed to ionizing radiation. A number of long-term studies of
899
the children of survivors of the atomic bombings in Hiroshima and Nagasaki have not detected any
900
transmitted genetic effects (transgenerational effects) of radiation exposure (Neel et al., 1980; Satoh
901
et al., 1996).
902
903
Children of cancer survivors have also been studied to learn of possible adverse genetic effects
904
associated with curative radiation treatments of the parents (Boice et al., 2003; Green et al., 2009;
905
Winther and Olsen, 2012). Blood samples taken from survivors of cancer, their spouses or partners
906
and their children have been analyzed for genomic instability, inherited mutations in minisatellite
907
DNA and in mitochondrial DNA, chromosome radiosensitivity and DNA polymorphic variation, and
908
the occurrence of cytogenetic abnormalities. Molecular studies found no evidence for inheritance of
909
minisatellite DNA mutations or mitochondrial DNA changes related to gonadal radiation (Guo et al.,
910
2012; Tawn et al., 2011). There was no evidence for genomic instability (Tawn et al., 2005) or
911
consistent evidence for G2 chromosome radiosensitivity (Wilding et al., 2007), although inheritance
912
of chromosome radiosensitivity was evident (Curwen et al., 2010). The molecular analyses
913
supported the epidemiologic findings of no detectable heritable genetic changes in the risk of
914
cytogenetic abnormalities, single gene disorders, birth defects, stillbirths, neonatal deaths and cancer
915
in the children of men exposed to testicular irradiation or women exposed to ovarian irradiation
916
(Signorello et al., 2012; Winther et al., 2012). These epidemiologic and cellular studies of the
917
children of cancer survivors are consistent with the studies of the children of atomic-bomb survivors
918
in Japan (Schull, 2003).
919
34
NCRP SC 1-21
Draft of March 31, 2015 [MR]
920
NOT TO BE DISSEMINATED OR REFERENCED
In contrast however, Dubrova and colleagues have described an elevated minisatellite mutation
921
rate in the offspring of workers involved in radiation cleanup after the Chernobyl accident (Dubrova
922
et al., 1996; 1997), as well as populations living along the Techa River (Dubrova et al., 2006).
923
However, these studies have been somewhat controversial (Livshits et al., 2001; Slebos et al., 2004).
924
Possible reasons for the discrepancy between these observations have been considered by Bouffler
925
et al. (2006), Little (2014) and Tawn et al. (2015).
926
927
928
3.1.7 Gene and Radiation Interaction: The Women’s Environmental Cancer and Radiation
Epidemiology Study
929
930
The Women’s Environmental Cancer and Radiation Epidemiology (WECARE) Study is an
931
international multicenter, population-based case-control study of breast cancer designed specifically
932
to examine the joint roles of interaction of gene carrier status and radiation exposure in the etiology
933
of breast cancer (Stovall et al., 2008). The underlying hypothesis is that a woman who carries certain
934
genetic mutations will be more susceptible to radiation-associated breast cancer than a woman who
935
is not a carrier. The study is nested within seven population-based cancer registries. 1,508 women
936
with asynchronous contralateral breast cancer were individually matched to 2,200 controls with
937
unilateral breast cancer on date and age at diagnosis of the first breast cancer, race, and registry
938
region.
939
940
Only young women with breast cancer were included because they were more likely to be
941
genetically predisposed to breast cancer and to have had a greater susceptibility to radiation-related
942
cancer. All study subjects were identified, recruited, and interviewed through seven population-
943
based cancer registries. The average age of cases and controls was 45 y and the average time
944
between primaries for cases was 5 y. All women were interviewed using a questionnaire that
945
included known and suspected risk factors for breast cancer and detailed treatment. Tumor
946
characteristic information and biospecimens were also collected. For each patient who received
947
radiation therapy, absorbed dose was estimated for each of the four quadrants and nipple area for the
948
contralateral breast. The average dose to the contralateral breast was 1 Gy (Stovall et al., 2008).
949
35
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
950
To date, the WECARE Study participants have been screened for candidate breast cancer
951
susceptibility genes, genes that lie in the ATM-CHEK2 pathway and across the genome using an
952
agnostic approach in order to determine genetic variation associated with contralateral breast cancer
953
and radiation-associated contralateral breast cancer in particular. Screening 2,100 WECARE Study
954
participants for the ATM gene yielded an elevated risk among carriers who were treated with
955
radiation [RR for carrying missense mutations and receiving >1 Gy (3.3, 95 % CI 1.4 to 8.0)]. The
956
risk was stronger among those who were treated at young ages and who carried mutations predicted
957
to be deleterious [RR for those under age 45 y at exposure (10.4, 95 % CI 2.3 to 47.2) (Stovall et al.,
958
2008)]. There was no elevated risk in the absence of radiation. The WECARE Study incorporates
959
epidemiologic methods, genetic data, and individual dosimetry for each participant, making the
960
study a paradigm for future interdisciplinary integrated genetic radiation research.
961
962
3.1.8 Retinoblastoma
963
964
Retinoblastoma is an embryonic childhood cancer of the eye that originates in the cells of the
965
retina (Gallie et al., 1999). It has both a heritable and a nonheritable etiology (Knudson, 1971).
966
Similar to other familial susceptibility syndromes, hereditary retinoblastoma is an autosomal-
967
dominant cancer resulting from deleterious mutations of a single gene of high penetrance, the RB1
968
tumor-suppressor gene (Taylor et al., 2007). Most patients with hereditary retinoblastoma develop
969
cancer in both eyes, but carriers of low-penetrance alleles may develop cancer in one eye only or not
970
at all. Nonhereditary retinoblastoma rises sporadically as a result of biallelic somatic mutation and
971
usually affects only one eye (Little et al., 2012a; Lohmann, 2010). Unlike patients with sporadic
972
(nonhereditary) disease, children with hereditary retinoblastoma are at a notable increased risk for
973
developing a second cancer later in life (Eng et al., 1993; Kleinerman et al., 2007). Radiation therapy
974
further increases the risk of sarcoma and second cancer development (Kleinerman et al., 2005;
975
NCRP, 2011; Wong et al., 1997; Yu et al., 2009). Radiation-related sarcomas occur most frequently
976
around the orbit of the eye that received curative treatments, whereas the nonradiation associated
977
sarcomas occur in other parts of the body. For germline (hereditary) retinoblastoma, every cell in the
978
body contains a deleterious mutation of the retinoblastoma tumor-suppressor gene and studies have
979
attempted to evaluate the possible interaction between high-dose radiation and underlying genetic
980
susceptibilities in causing second primary cancers (Kleinerman, 2009). While there is compelling
36
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
981
evidence for an enhancement of risk of certain types of second cancer following radiation therapy
982
among children with heritable retinoblastoma contrasted with children with the nonheritable type of
983
retinoblastoma (i.e., for a gene and radiation interaction), the data are limited by small numbers and
984
statistical uncertainties (Draper et al., 1986; Wong et al., 1997; Yu et al., 2009). For some cancer
985
sites (e.g., breast) there appears to be little or no excess radiation risk among those with heritable
986
retinoblastoma, in contrast to the much larger risk in those without the heritable disease (Little et al.,
987
2014b). Nonetheless, it is interesting that studies of cells from retinoblastoma families involving the
988
-H2AX assay (Kato et al., 2007) and the G2 chromosome-radiosensitivity assay (Wilson et al.,
989
2010) indicate elevated radiation sensitivities with respect to cell killing and DNA repair or DNA
990
damage processing functions. Conceivably these cellular studies point to a defect in genome
991
maintenance, including the capacity to process radiation-induced DNA damage, that results in
992
increased germline mutation rates, predisposition to cancer, and increased susceptibility to radiation-
993
related cancer. Identification of such genes or factors involved could provide insights into
994
995
996
mechanisms leading to retinoblastoma, cancer predisposition and radiation carcinogenesis.
997
radiation therapy-related cancers (Travis et al., 2006). Germline mutations in the RB1 tumor-
998
suppressor gene predispose children to a high risk of osteosarcomas, soft-tissue sarcomas, and a
999
number of other cancers. Radiation therapy further enhances the risk of tumors arising in the
Retinoblastoma provides an example of how genetic mutations may influence the risk of
1000
radiation field. Mutations in the RB1 gene and perhaps other genes appear important in the cellular
1001
response to DNA damage and may confer an increased risk of radiation-related cancer. It is unclear,
1002
however, the extent to which radiation therapy for hereditary retinoblastoma, a genetic disorder due
1003
to mutations in a gene of high penetrance, can be generalized to genetic conditions of lower
1004
penetrance (or to polygenic circumstances) following much lower absorbed doses at or below
1005
1006
1007
1008
100 mGy.
1009
3.2. Epidemiologic Studies Focused on Noncancer Outcomes
There is emerging evidence of radiation-associated excess risk of circulatory disease in a number
1010
of groups exposed at moderate doses (e.g., 500 mGy) (Little et al., 2012b). In particular, radiation-
1011
associated excess risk of mortality and morbidity for a number of circulatory disease endpoints has
1012
been observed in the Japanese atomic-bomb survivor data (Shimizu et al., 2010); Yamada et al. (2004).
37
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
1013
There is very rich covariate information available within the AHS subcohort of atomic-bomb
1014
survivors, and the radiation associations for a number of these have been analyzed in a number of
1015
studies. For example, elevated levels of the pro-inflammatory cytokines IL6, CRP, TNF-α and INF-γ,
1016
but also increased levels of the (generally) anti-inflammatory cytokine IL10, have been observed in
1017
the Japanese atomic-bomb survivors (Hayashi et al., 2003; 2005). There was also a dose-related
1018
elevation in the erythrocyte sedimentation rate and in the levels of IgG, IgA and total
1019
immunoglobulins, all markers of systemic inflammation, in this cohort (Hayashi et al., 2005).
1020
Certain T-cell and B-cell population numbers are known to vary with radiation dose among the
1021
Japanese atomic-bomb survivors (Kusunoki et al., 1998), suggesting a role for the immune system.
1022
The atomic-bomb survivors also demonstrate dose-dependent decreases in levels of CD4+ helper T
1023
cells (Hayashi et al., 2003); decreased levels of helper T cells have also been found in blood samples
1024
from Japanese atomic-bomb survivors with myocardial infarction (Kusunoki et al., 1999). However,
1025
none of this covariate data has been analyzed together with circulatory disease outcome data. There
1026
is known to be similarly rich covariate information in the Mayak nuclear worker data, in which there
1027
is also significant radiation-associated excess risk of circulatory disease (Azizova et al., 2010a;
1028
2010b; 2011; 2014; Moseeva et al., 2014). However, this covariate data has not been separately
1029
1030
analyzed, nor is there any analysis in which this has been combined with health outcome data.
1031
The literature on radiation and cataract and other lenticular opacities has been the subject of a
1032
number of recent reviews (Ainsbury et al., 2009; ICRP, 2012; Little, 2013; Shore et al., 2010),
1033
including a recent systematic review (Hammer et al., 2013), all suggesting that there may be
1034
radiation-associated risk at moderate and low doses. However, there appear to be no studies
1035
combining analysis of cataract outcome data with intermediate bioindicator or biomarker covariates.
1036
For all other noncancer disease endpoints, the evidence for radiation etiology is much weaker,
1037
whether in directly exposed populations (Little, 2013) or in the offspring of exposed individuals
1038
(Little et al., 2013).
1039
38
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
1040
1041
1042
4. Bioindicators from Radiation Biology Studies
1043
outcomes (and to a lesser extent noncancer outcomes) was introduced in Section 1.2. This included a
1044
description of how bioindicators of effect could be used as the key events for such an approach. In
1045
Section 4, a discussion is presented of the types of bioindicators that can be measured in cellular,
1046
laboratory animal and human systems, with a parallelogram approach for linking these
1047
measurements to adverse health outcomes in humans. In this way, it is feasible to both describe the
1048
shape of the dose response at low doses and low dose rates as well as the risk estimates themselves
1049
for these exposure scenarios. The rapid development of new technologies that permit whole-genome
1050
sequencing in a very short time has and will continue to enhance the ability to identify and measure
1051
many of these types of informative bioindicators. To fully utilize this type of approach to develop a
1052
bioindicator of response, it is likely that a systems biology type approach (Section 5.2.2) will be
1053
1054
1055
1056
1057
needed.
1058
indirectly through production of reactive oxygen species, producing a plethora of lesions, including
1059
many types of base damage such as broken rings, oxidized bases, small and large adducts including
1060
DNA-DNA and DNA-protein crosslinks, single-strand breaks (SSBs), and DSBs, including
1061
relatively reparable simple DSBs, and more difficult to repair complex DNA lesions. Under typical
1062
conditions, ~40 DSBs are created per cell per gray, and many more base lesions and SSBs are
1063
produced than DSBs (Ward, 1998). However, DSBs have the greatest biologic significance since
1064
DSBs correlate with cytotoxicity, chromosome aberration formation, and cellular transformation
1065
1066
1067
(Frankenberg-Schwager, 1990).
1068
2011). Specific lesions are typically repaired by a specific pathway or subpathway, but if the primary
1069
pathway fails, repair may be attempted by another pathway due to redundancies in the process. DNA
1070
damage and repair do not occur in a static environment, but rather in the context of other dynamic
1071
processes including DNA replication and transcription, and in a structurally complex and dynamic
The concept of using a key event/adverse outcome pathway approach for estimating cancer
4.1 Deoxyribonucleic Acid Damage and Repair
Ionizing radiation causes damage to DNA by direct energy absorption and to a larger extent,
There are at least five types of DNA repair pathways, each with subpathways (Shaheen et al.,
39
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
1072
chromatin environment. DNA repair systems maintain genome stability and confer resistance to
1073
ionizing radiation and other DNA damaging agents, and defects in DNA repair systems typically
1074
predispose to cancer and other diseases (Friedberg et al., 2005). Three sets of pathways act on
1075
single-strand damage, including base-excision repair (nonbulky lesions) (Krokan and Bjoras, 2013),
1076
nucleotide excision repair [bulky, helix distorting lesions (Kamileri et al., 2012)], and mismatch
1077
repair (Hsieh and Yamane, 2008). Classical and alternative nonhomologous end joining, and
1078
homologous recombination, are redundant DSB repair pathways that can restore the chemical
1079
1080
1081
integrity of DNA, but with lesser or greater capacity to accurately restore genetic sequences.
1082
when forks encounter other types of lesions (single-strand breaks, many types of base damage, most
1083
DNA adducts, pyrimidine dimers, and intra- and inter-strand crosslinks) (Allen et al., 2011;
1084
Budzowska and Kanaar, 2009). DSBs are marked by phosphorylated histone H2AX (-H2AX)
1085
(Rogakou et al., 1998), which plays important roles in DNA damage checkpoint signaling and DSB
1086
repair (Chanoux et al., 2008; Downey and Durocher, 2006). -H2AX is a fairly reliable biomarker of
1087
DSBs, detected by fluorescence microscopy using phospho-specific antibodies, hence -H2AX is
1088
often used in kinetic studies to monitor both DSB induction and repair (Lobrich et al., 2010), and as
1089
a radiation biodosimeter (Rothkamm et al., 2013). Proteins involved in nonhomologous end joining
1090
and homologous recombination migrate to DSBs, such as the early DNA damage response protein
1091
53BP1, and the late acting homologous recombination protein RAD51. The resulting protein
1092
aggregates (“repair factories”) can be detected with fluorescence microscopy using cognate
1093
antibodies as subnuclear foci. These foci thus complement -H2AX as additional biomarkers of
1094
DSBs and their repair (Allen et al., 2011). -H2AX foci represent a robust biomarker of DNA
1095
damage by radiation, and induced foci or residual (unrepaired) -H2AX foci show promise as
1096
potential biomarkers of risk for various cancers (Fernandez et al., 2013; He et al., 2013; Matthaios
1097
1098
1099
et al., 2013).
DSBs are produced immediately by radiation, or they can arise later during DNA replication
4.2 Cellular Signaling at Low Doses
1100
1101
1102
Cells respond to DNA damage in part through the protein kinase mediated activation of signaling
pathways that can culminate in cell-cycle arrest, stimulation of DNA repair, induction of cell death,
40
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
1103
or terminal differentiation and senescence. The cellular DNA-damage response includes two major
1104
checkpoint signaling pathways; one responding to DSBs through ATM activation that leads to
1105
CHEK2 activation and p53 stabilization (Postiglione et al., 2010), and one centered on ATM and
1106
Rad3-related activation, which activates CHEK1 in response to single-strand breaks and gaps (i.e., at
1107
replication forks) (Fokas et al., 2014). Damage to cellular components in addition to DNA can also
1108
trigger signal transduction pathways, such as the ceramide pathway that responds to damage to the
1109
cell membrane (Zeidan and Hannun, 2010), redox signaling from mitochondria (Hei et al., 2008), as
1110
well as signaling between directly hit and non-hit bystander cells (Section 4.6.2).
1111
1112
Checkpoint pathways are not “on or off” but show different responses depending on the level of
1113
damage. This is reflected in different gene expression responses at different levels of radiation
1114
exposure and in different physiological outcomes (Ding et al., 2005; Gridley et al., 2009; Lu et al.,
1115
2010; Mezentsev and Amundson, 2011). DNA damage response thresholds are genetically regulated
1116
(Peng et al., 2010; Putnam et al., 2009), and thresholds may vary for each pathway (Fernet et al.,
1117
2010). With minimal DNA damage, as with low-dose ionizing radiation, some cells may not activate
1118
the DNA repair process (Rothkamm and Lobrich, 2003) while others may activate repair pathways
1119
but not cell-cycle arrest or death pathways. Activation of repair or other damage signaling pathways
1120
by low-dose ionizing radiation probably underlies adaptive responses that enhance survival, prevent
1121
genomic instability, or limit deleterious effects when cells are subsequently exposed to large ionizing
1122
radiation doses (Section 4.6.1). DNA repair efficiency may vary with the level of damage, with more
1123
efficient repair at low ionizing radiation doses than at high doses (Neumaier et al., 2012). At higher
1124
levels of damage, cells arrest at G1, S, and/or G2/M phases to prevent cell-cycle progression with
1125
DNA damage to reduce the chance of mutations, aneuploidy, and mitotic catastrophe, and above
1126
threshold levels of damage (which vary with cell type), cell death and nonproliferation pathways are
1127
activated, including apoptosis, autophagy, necrosis and senescence (Branzei and Foiani, 2009;
1128
Budzowska and Kanaar, 2009; Eriksson and Stigbrand, 2010; Galluzzi and Kroemer, 2008;
1129
1130
1131
1132
1133
Vakifahmetoglu et al., 2008; Vandenabeele et al., 2010).
1134
of radiation dose, dose rate and radiation quality, and are the best-documented biomarker of
4.3 Chromosome Alterations
Following exposure to ionizing radiation, chromosomal rearrangements are induced as a function
41
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
1135
radiation exposure, even to doses as low as 200 mGy (Lloyd et al., 1992; Pohl-Ruling et al., 1983;
1136
Tucker and Luckinbill, 2011). Chromosomal rearrangements arise as a consequence of the
1137
processing of directly and indirectly induced DNA DSBs and at high dose rates increase in a linear-
1138
quadratic fashion for sparsely ionizing radiation, suggesting that the observed rearrangements are the
1139
result of two or more interacting molecular events. There is a long and well-documented literature on
1140
the dose-response relationships for the induction of chromosomal aberrations evaluated after Giemsa
1141
staining, or by fluorescence in situ hybridization using chromosome-specific molecular probes, and
1142
more recently by combinatorial hybridization techniques (reviewed in Cornforth, 2001).
1143
1144
Using these more modern technologies it appears that the shape of the dose-response profile for
1145
chromosomal rearrangements is influenced by the production of complex rearrangements involving
1146
three or more chromosomes (Loucas et al., 2013), especially following exposure to low dose-rate
1147
radiation exposures (Loucas et al., 2004).
1148
1149
While chromosomal alterations are a well-documented biomarker of exposure and dose, and
1150
remain the gold standard in this area following potential radiation exposure (Tucker, 2008), it should
1151
be noted that rearrangement frequency can vary significantly between tissues (Brooks et al., 2003).
1152
Ongoing studies suggest that chromosomal aberrations might predict human cancer independently of
1153
exposure to carcinogens (Bonassi et al., 2000; Hagmar et al., 1994). However, it remains to be seen
1154
if a molecular, biochemical, or cellular signature of radiation exposure can be correlated with
1155
chromosomal aberration formation (a biomarker of radiation exposure) and be used as a predictor of
1156
health risk(s) associated with irradiation and predict adverse effects, particularly at low radiation
1157
doses.
1158
1159
4.4 Mutations
1160
1161
Mutations are essential features of cancer (Hanahan and Weinberg, 2011), and ionizing
1162
radiation was the first environmental mutagen identified (Muller, 1927). Mutations can serve as
1163
bioindicators of radiation effects as a function of dose, dose rate and radiation quality, and most
1164
mutation studies to date have made use of model systems that assess changes at reporter loci that are
1165
not directly linked to tumor formation.
42
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
1166
1167
Most mutation studies to date have utilized model reporter gene systems in which changes in the
1168
DNA sequence of a gene result in the failure to produce a functional protein product, allowing
1169
chemical selection of the mutant cells. Such mutations can reflect small changes in the protein
1170
product that render it unable to perform its normal function, or the failure to produce any of the
1171
protein, but the proteins involved generally are not related to cancer development.. However,
1172
information learned from these studies can be useful in helping to guide our understanding of
1173
mutational mechanisms and the types of mutational events likely to occur at loci that are directly
1174
linked to tumor formation following exposure to ionizing radiation. All of the studies presented in
1175
Section 4.4 focus on the identification of mutations that are thought to be directly induced by the
1176
radiation at short times after exposure.
1177
1178
Tumor suppressor genes generally require mutation of two alleles, often involving a point
1179
mutation and subsequent loss of the second allele to enable tumor formation. Model systems have
1180
been engineered to allow quantification of mutations at such heterozygous autosomal loci, and
1181
analysis of the molecular changes in DNA. In these models, one copy of a target gene, such as
1182
thymidine kinase (TK1, or Tk) or adenine phosphoribosyltransferase (APRT, or Aprt), has been
1183
inactivated and a mutagenic event that results in the loss of production of any functional protein
1184
from the second copy of the gene can be selected using a chemical that will kill only the normal
1185
cells. Mutations can also occur in dominantly acting oncogenes. In this case, mutation of a single
1186
copy of the gene can lead to tumorigenesis, but the mutational mechanisms are the same.
1187
1188
Radiation has been shown to induce mutations by many mechanisms, including all types of
1189
single basepair substitutions (Grosovsky et al., 1988), small insertions and deletions within the gene
1190
itself (intragenic events), deletion of the whole gene, multi-locus deletions, recombination-mediated
1191
events leading to loss of heterozygosity, and loss of whole chromosomes (a form of aneuploidy)
1192
(reviewed in Turker et al., 2009; Wiese et al., 2001).
1193
1194
Studies using single copy genes, such as the X-linked hypoxanthine phosphoribosyltransferase
1195
(HPRT, or Hprt) gene, can detect a subset of the radiation-induced mutational mechanisms that can
1196
occur at an autosomal locus (Liber et al., 1989). Similarly, the S1/MIC1/CD59 locus in artificially
43
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
1197
constructed human x hamster hybrid cells has been used to demonstrate that ionizing radiation often
1198
produces chromosomal scale events, including whole chromosome loss, although such events are
1199
often lethal in normal human cells (Kraemer et al., 2001).
1200
1201
There are also data on radiation mutagenesis at low doses and/or low dose rates. In one study,
1202
induction of HPRT mutations in a human lymphoblast cell line by protracted exposure could not be
1203
distinguished from the mutant frequencies measured in samples exposed to the same total doses in a
1204
single acute exposure, suggesting that there was no dose-rate effect (Grosovsky and Little, 1985).
1205
Other studies using the same cell line have found an inverse dose rate effect for TK1 mutation that
1206
was dependent on dose rate (Amundson and Chen, 1996; Brenner et al., 1996).
1207
1208
The Aprt heterozygous mouse model provides a good example of translation of mutagenesis
1209
studies from in vitro to in vivo. Mutations have been characterized in kidney cells irradiated in vitro,
1210
or in cells retrieved from irradiated kidneys in mice that had total body exposures. All of the types of
1211
mutational events that occurred in kidney cells exposed in vitro also occurred in vivo, from small
1212
intragenic events to whole chromosome loss (Ponomareva et al., 2002; Turker et al., 2009; 2013).
1213
The chromosomal milieu of the Aprt locus in the mouse is similar to that for the TK1 locus in human
1214
cells. This is important for cross-species extrapolations of mutational risk, although one class of
1215
radiation-induced mutations (whole chromosome loss) that can be observed in the engineered mouse
1216
model is not detected in the human cells, most likely due to functional hemizygosity of linked loci in
1217
the “outbred” human genome that would result in cell death from such an event in the human cells.
1218
1219
Recent studies on radiation-related acute myeloid leukemia (AML) in mice have examined
1220
mutations at the PU.1 locus on mouse chromosome 2. Deletion of one allele of PU.1 was proposed
1221
as a candidate bioindicator for radiogenic AML in a susceptible mouse strain (Peng et al. 2009).
1222
Further work indicates that point mutation of the second PU.1 allele is not radiation related, and that
1223
mutation of other genes located near PU.1 may be particularly important in radiogenic AML in
1224
different mouse strains (Genik et al., 2014).
1225
1226
Radiation-induced autosomal mutations have also been studied in Dlb-1 heterozygous mice.
1227
Mutations were measured in intestinal stem cells and their progeny (cells which are of particular
44
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
1228
interest since mutations in stem cells are thought to be important in the carcinogenic process). Fewer
1229
gamma ray-induced mutations were observed at the Dlb-1 locus in the mouse small Muintestine
1230
when the dose rate was lowered to 0.01 Gy min–1 as compared with acute exposures at 1.8 Gy min–1
1231
(Winton et al., 1989). However, this study was limited to high total doses.
1232
1233
In the early 1950s and 1960s hybrid male mice were irradiated and then mated to unirradiated
1234
females that harbored a series of seven recessive loci and mutations were assessed in these marker
1235
loci in the offspring. Mutant frequency results indicated that low-LET radiation delivered to stem
1236
cell spermatogonia at exposure rates of 0.8 R min–1 and below induced only about one-third as many
1237
mutations in the offspring as those delivered as single acute exposures (Russell and Kelly, 1982;
1238
Russell et al., 1958). While the total exposures in these experiments were not low (86 to 1,000 R),
1239
the exposure rates were as low as 1 mR min–1 and lower in some cases. The data gathered in these
1240
very large-scale studies were re-examined in light of the availability of fine-structure genetic or
1241
molecular analyses of the seven marker loci and their surrounding chromosomal regions (Russell
1242
and Hunsicker, 2012), and indicated that the reduction in mutant frequency at the lower exposure-
1243
rates is associated with a decline in “large lesion” mutations. Since these data reflect radiation effects
1244
on the irradiated male germline, it is more challenging to extrapolate these findings to somatic cells
1245
in adult humans, where most human cancers develop.
1246
1247
New approaches to the analysis of mutations have become available with the advent of next-
1248
generation DNA sequencing, in addition to molecular cytogenetic approaches that can detect copy
1249
number variations with increasing precision. These methods are being considered now for use in the
1250
analysis of irradiated cells in culture as well as in irradiated tissues. Associating radiation-induced
1251
changes with particular loss of heterozygosity or other events that may be causal to tumor formation
1252
will be challenging, given the wide variety of molecular events that lead to radiation-induced
1253
mutations and the large number of driver and passenger mutations found in human tumors (Bozic
1254
et al., 2010; Vogelstein et al., 2013). An example of the use of such an approach comes from
1255
experimental studies of genetic and chemically-induced lung cancers in mice (Westcott et al., 2014).
1256
In this study, the carcinogen-induced tumors display signatures of the initiating chemical
1257
carcinogens. The situation may be more challenging for ionizing radiation-induced events, where
1258
specific radiation-signature events have been more difficult to identify.
45
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
1259
1260
4.5 Radiation-Induced Genomic Instability
1261
1262
Genomic instability is characterized by an increased rate of genetic change in the progeny of an
1263
irradiated cell. Instability manifests as alterations in chromosomal rearrangements, micronuclei,
1264
increased gene copy number, minisatellite and microsatellite instabilities in DNA repeat sequences,
1265
transformation, and cell killing and can be induced through both targeted and nontargeted bystander-
1266
like processes (Little, 2000; Morgan et al., 2002). Genome instability is a key feature of cancer cells
1267
and it is now well established that defects in DNA repair and DNA damage response systems
1268
predispose to cancer. Defects in DNA repair or DNA damage response systems can allow more rapid
1269
accumulation of instability that underlies cancer development. However it is still controversial
1270
whether there is a causal association between instability and cancer (Little, 2010). Genomic
1271
instability can be produced from endogenous or induced DNA damage, and as for nontargeted
1272
bystander effects, the mechanisms of instability share features with inflammatory responses
1273
characterized by intercellular signaling, production of chemokines, cytokines, and reactive oxygen
1274
and nitrogen species.
1275
1276
Ionizing radiation causes DNA damage that directly leads to large- or small-scale genetic
1277
changes, ranging from single-base (point) mutations to gene rearrangements, translocations, and
1278
whole chromosome gain or loss (Section 4.4). Studies over the past decade have shown that ionizing
1279
radiation (low- and high-LET) and nonionizing radiation (e.g., ultraviolet light) also induce genome
1280
instability many cell generations after the exposure (Durant et al., 2006; Huang et al., 2004; 2007;
1281
Morgan, 2003a; 2003b). Some of these instability phenotypes are correlated and therefore likely
1282
share a common means of induction or are otherwise related mechanistically (Limoli et al., 1997).
1283
For example, cells that display ionizing radiation-induced chromosome instability typically show
1284
delayed death, probably reflecting the low viability of cells with large-scale karyotypic changes and
1285
aneuploidy as a result of massive gene expression imbalances. In contrast, ionizing radiation-
1286
induced delayed hyper-recombination is not associated with delayed death; therefore chromosome
1287
instability and delayed hyper-recombination are mechanistically distinct processes (Huang et al.,
1288
2004).
1289
46
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
1290
One commonality among the many instabilities investigated to date is the fact that they can be
1291
induced by both high (1 to 10 Gy) and low (0.01 to 0.1 Gy) ionizing radiation doses. However, in
1292
some instances a threshold dose has been observed and there are also examples of the effect
1293
saturating at a specific radiation dose (Limoli et al., 1999). Like many immediate effects of
1294
radiation, the induction of genome instability is also subject to adaptive responses. As an example,
1295
exposing cells to low doses of ionizing radiation offered some protection for induction of
1296
instabilities by a subsequent higher dose (Huang et al., 2007). However, the adaptive response was
1297
obviously not a complete or even a strong protective mechanism, since even low doses were
1298
sufficient to induce instabilities.
1299
1300
Two other features common to ionizing radiation-induced instabilities is that they occur at high
1301
frequencies (often 10 % or more of cells that survive low to moderate doses) and they typically do
1302
not show a dose response (Little, 2000; Morgan, 2003a). These features suggest that the target for
1303
induced instability is large. It is difficult to imagine instabilities arising from gene inactivation (even
1304
if only one of a large family of genes such as those in the DNA damage response pathway needed to
1305
be inactivated). Instead, the target is likely to be as large as the nucleus or even the whole cell.
1306
1307
There is considerable variation in the manifestation of induced instability between cell lines in
1308
vitro, and organisms in vivo. This variation in response has a genetic component (Ponnaiya et al.,
1309
1997) and may reflect the age of the organism at the time of irradiation as well as the cells/organisms
1310
ability to detoxify cellular reactive oxygen/nitrogen species (Spitz et al., 2004). Once initiated
1311
however, there is evidence that instability can be perpetuated over time by dicentric chromosome
1312
formation followed by bridge breakage fusion cycles (Marder and Morgan, 1993) as well as
1313
recombinational events involving interstitial telomere like repeat sequences (Day et al., 1998). There
1314
is also increasing evidence that epigenetic factors (Aypar et al., 2011; Koturbash et al., 2006),
1315
dysfunctional mitochondria (Kim et al., 2006a; 2006b), and inflammatory type reactions (Lorimore
1316
and Wright, 2003; Mukherjee et al., 2012), presumably involving reactive oxygen and nitrogen
1317
species as well as cytokines and chemokines might be involved in driving the unstable phenotype
1318
(Laiakis et al., 2007; Hei et al., 2008). To this end there is convincing evidence for such reactions
1319
being involved in another nontargeted effect associated with ionizing radiation, the bystander effect
1320
(Morgan et al., 2002). Clearly the link between induced instability and bystander effects suggests
47
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
1321
common processes and inflammatory type reactions will likely be the subject of future investigation,
1322
as will the clinical implications of induced instability in terms of a marker for increased cancer
1323
risk/early detection, and as a consequence of radiation therapy that may result in the induction of a
1324
therapy-related second malignancy (Goldberg, 2003).
1325
1326
While the precise molecular, cellular, organ, and organismal mechanisms contributing to
1327
radiation-induced genomic instability are not known, the phenotype is well documented in a number
1328
of in vivo model systems (Morgan, 2003b; Watson et al., 2001). This raises a number of questions
1329
regarding this intriguing phenotype.
1330
1331

What, if any, specific types of radiation-induced genomic instability are capable of
1332
transforming a cell from a normal to a cancer cell, or furthering the progression of a
1333
premalignant cell towards a more aggressive cancer phenotype?
1334

1335
1336
What, if any, is the role of induced instability in secondary cancers observed in radiation
therapy patients?

Could analysis of radiation-induced genomic instability in peripheral blood lymphocytes be
1337
used as a biomarker of an individual’s radiation sensitivity and potential radiation-related
1338
cancer risk?
1339
1340
Recently the integration of radiobiological effects in a two-step clonal expansion model of
1341
carcinogenesis and applications to radioepidemiologic data have been proposed (Jacob et al., 2010).
1342
First, a model version with radiation-induced genomic instability was shown to be a possible
1343
explanation for the age dependence of the radiation-related cancer mortality in the Techa River
1344
Cohort. Second, it was demonstrated that inclusion of a bystander effect with a dose threshold allows
1345
an improved description of the lung cancer mortality risk for the Mayak workers cohort due to
1346
incorporation of plutonium. It is envisaged that a further combination of epidemiologic, biological
1347
and modeling interactions will address these questions in the future (Eidemuller et al., 2011; Kaiser
1348
1349
et al., 2014).
48
NCRP SC 1-21
Draft of March 31, 2015 [MR]
1350
NOT TO BE DISSEMINATED OR REFERENCED
4.6 Modulators of Response
1351
1352
Classical radiation biology was built on the concept of target theory wherein the deleterious
1353
effects of exposure including chromosome aberrations, mutations and carcinogenesis are assumed to
1354
arise from damage to a cellular target (e.g., nuclear DNA) via energy deposition in the cell at risk
1355
(Hall and Giaccia, 2011; Lea, 1946; UNSCEAR, 1993). The deposition of energy after irradiation is
1356
well understood, but the biological processes that modify response to that damage may vary between
1357
individuals and in response to low versus high dose and dose rate.
1358
1359
In recent years, much research has focused on responses to doses of 100 mGy and below.
1360
While some responses appear to follow a continuous dose-response relationship from high to low
1361
doses, some other responses may only occur at either low or high doses, thus complicating the
1362
extrapolation from high dose to low dose. Current radiation risk extrapolations to low doses assume
1363
that:
1364
1365

the primary mode of action is linearly related to dose; and
1366

cancer is clonal in origin (i.e., the individual cell is the unit of risk), although cellular
1367
responses are likely modulated by the tissue microenvironment.
1368
1369
Although the targeted effects of radiation (i.e., responses in hit cells) may be proportional to dose,
1370
other processes, including effects in non-hit cells (e.g., induced genomic instability and bystander
1371
effects) may show nonlinear responses. Some of these low-dose responses are highlighted in
1372
Section 4.6.
1373
1374
4.6.1 Adaptive Responses
1375
1376
Two different adaptive responses have been described. The classical adaptive response was first
1377
described by Olivieri et al. (1984). They demonstrated that when human lymphocytes were exposed
1378
to a very low dose of radiation, a priming dose, followed a short time later with a larger higher dose,
1379
the challenge dose, the frequency of chromosomal aberrations induced by the challenge dose was
1380
less than that from the challenge dose given alone. A second adaptive response suggests that when
49
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
1381
cells with a high background frequency of phenotypic alterations (e.g., mutations or transformation)
1382
were exposed to very low doses of ionizing radiation, these doses can actually reduce that frequency
1383
to below background (Azzam et al., 1996; Redpath et al., 2001).
1384
1385
While adaptive responses are well characterized in certain situations in vitro, there is also some
1386
evidence that they can occur in humans occupationally (Barquinero et al., 1995) or clinically
1387
(Monsieurs et al., 2000) exposed to ionizing radiation. As always in a biological system there is
1388
considerable variation both in vitro (Bosi and Olivieri, 1989) and in vivo (Padovani et al., 1995;
1389
Tedeschi et al., 1995). In fact, not only are reduced levels of DNA alterations observed in some
1390
systems, in other individuals no response has been observed, while in others an additive increase in
1391
the manifestations of irradiation has been observed (Tapio and Jacob, 2007; UNSCEAR, 1994;
1392
2012).
1393
1394
Adaptive response(s), if reproducibly induced by low doses of ionizing radiation, have potential
1395
clinical and radiation protection implications. For example, that radiation signals danger to the
1396
immune system through the Toll-like receptor family can cause cells to release damage-associated
1397
molecular patterns to activate canonical inflammatory pathways and/or initiate immunity (McBride
1398
et al., 2004). Through induction of superoxide dismutase and anti-inflammatory responses, low-dose
1399
ionizing radiation could also be used prior to surgery or other therapies that cause inflammation or
1400
ischemia. Howell et al. (2013) indicated that exposure of mice in utero to a dose rate of 10 to
1401
13 mGy d–1 [Chernobyl soils (primary radionuclides being 137Cs and 90Sr) embedded into bioplastic]
1402
for 10 d had no deleterious effects on litter size, bone marrow stem cells, and white blood cell
1403
counts, yet this exposure had significant protective effects when pups were later exposed to an acute
1404
dose of 2.4 Gy (calibrated 137Cs source). Alternatively, if low dose radiation exposures can induce an
1405
adaptive response then a CT scan for treatment planning purposes may result in the subsequent
1406
fractionated radiation dose being less effective in tumor cell kill, thus limiting the impact of radiation
1407
therapy. These outcomes are likely to vary between individuals and a future challenge is to
1408
determine which individuals will manifest one or other of these alternatives.
1409
50
NCRP SC 1-21
Draft of March 31, 2015 [MR]
1410
NOT TO BE DISSEMINATED OR REFERENCED
4.6.2 Bystander Effects
1411
1412
Bystander effects occur in nonirradiated cells that receive a signal from an irradiated cell. Such
1413
signals may be secreted and/or shed by irradiated cells or communicated via molecules transported
1414
between cells by cell-to-cell gap junction communication. Not all cells are able to produce bystander
1415
signals or respond to them. The bystander effect has been the subject of a number of recent reviews
1416
(Blyth and Sykes, 2011; Butterworth et al., 2013; Ilnytskyy and Kovalchuk, 2011; Mancuso et al.,
1417
2012).
1418
1419
The key question in the context of this Commentary is whether bystander effects can be
1420
beneficial or detrimental to the nonirradiated cell, tissue, or organism. On the cellular level there is
1421
evidence of both potentially protective effects, such as apoptosis (Belyakov et al., 2005) and
1422
differentiation (Belyakov et al., 2006), and detrimental effects, such as DNA damage (Dickey et al.,
1423
2009; Nagasawa and Little, 1992), chromosomal instability (Dickey et al., 2009), mutation (Zhang
1424
et al., 2009; Zhou et al., 2000), and transformation (Sawant et al., 2001). More recently, bystander
1425
responses including mutagenesis (Chai et al., 2013a; 2013b) and carcinogenesis (Mancuso et al.,
1426
2008; 2011; 2013) have been convincingly demonstrated in animals in vivo. Neither the net
1427
protective or detrimental effect of the different processes affected, nor their relevance to human risk
1428
are yet fully understood.
1429
1430
It is sometimes thought that the implications of bystander effects for radiation protection and
1431
ultimately for interpreting epidemiologic studies are not likely to be significant – in that they are
1432
likely already “built into” organ risks and consequent tissue weighting factors. However, at low
1433
doses and/or dose rates this is probably incorrect. The bystander effect would not be observed (e.g.,
1434
in the LSS data) because most cells are traversed by at least 100 electron tracks (= 0.1 Gy). But this
1435
is not the case extrapolating to an occupational setting, for example, where, with low-LET doses of
1436
1 mGy y–1, most cells would receive only one electron track per year. By definition, bystander
1437
effects imply that the target for radiation effects in low-dose exposures will be greater than the target
1438
volume actually irradiated.
1439
51
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
1440
A number of bystander effect models have been fitted to experimental (Brenner et al., 2001;
1441
Little et al., 2005; Nikjoo and Khvostunov, 2003) and epidemiologic (Brenner and Sachs, 2003;
1442
Little, 2004; Little and Wakeford, 2001) data. The available data suggest that despite compelling
1443
experimental findings there is not strong evidence of the involvement of the bystander effect as a
1444
modifier of radiation-related cancer risk in humans (Little, 2004; Little and Wakeford, 2001; Little
1445
et al., 2009). However, as bystander responses could alter the shape of the dose-response curve at
1446
low doses, an understanding of this area could add confidence to biological models and
1447
extrapolations.
1448
1449
4.6.3 Genetic Susceptibility and Interactions with Radiation
1450
1451
Over the past decade or so, there has been an extensive increase in knowledge of the genetic
1452
basis of diseases, especially cancers of many different types (Weinberg, 2013). For example, for so-
1453
called high penetrance genes, excess spontaneous cancers are expressed in a large proportion of
1454
carriers. In addition, there are gene-gene and gene-environment interactions that can lead to
1455
variations in the likelihood of cancer across the population. How such information might convert
1456
into inter-individual differences in susceptibility to radiation-related cancer has been quite
1457
extensively discussed by ICRP (1998), NA/NRC (2006), and UNSCEAR (2001; 2008a). However, it
1458
remains unclear as to the magnitude of the effect of any such sensitivity even at the individual level,
1459
let alone a population level.
1460
1461
Sankaranarayanan and Chakraborty (2001) conducted an extensive analysis of potential impacts
1462
at the individual and population levels. They used a Mendelian one-locus, two allele autosomal
1463
dominant model for predicting the impact of cancer predisposition and increased radiosensitivity on
1464
the risk of radiation-related cancers in the population and in relatives of affected individuals using
1465
breast cancer due to BRCA-1 mutations. The general conclusions from their study were that, when
1466
the proportion of cancers due to the susceptible genotype is small (<10 %), the attributable risks are
1467
small. These risks only become large when cancer susceptibility is very large (>10-fold). On the
1468
other hand, when the proportion of cancers resulting from the susceptible genotypes is large (10 %),
1469
there can be significant increases in attributable risk for relatively small increases in cancer
1470
susceptibility (>10-fold) and radiosensitivity (>100-fold) in the susceptible populations. This means
52
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
1471
that the increase in cancer risk to a heterogeneous population of susceptible and nonsusceptible
1472
genotypes will generally be quite small if the proportion of susceptible individuals is small and the
1473
radiation sensitivity relatively small. On the other hand, cancer risks assessed on the basis of an
1474
individual (in radiation therapy situations, for example) can be greatly influenced by genotype. For
1475
the purposes of radiation protection, dose limits for the public will not be significantly influenced by
1476
genetic radiosensitivity whereas they certainly could be in some occupational settings and for
1477
individual assessments for defined medical exposures.
1478
1479
There is evidence to suggest that efficiency in DNA DSB-repair modulates radiation toxicity risk
1480
because patients with genetic syndromes involving DNA DSB-repair factors, such as ataxia
1481
telangiectasia, have well recognized acute and late sensitivities to radiation. However, apart from
1482
these rare genetic syndromes, the genetic determinants of radiation toxicity are largely unknown.
1483
However, progress is being made in demonstrating experimentally that there are complex
1484
interactions that underlie the expression of cancer-predisposing genes of lower penetrance
1485
(NA/NRC, 2006). This knowledge highlights the difficulty that will be encountered in trying to
1486
incorporate facets of genetic radiosensitivity into the radiation risk assessment process, especially at
1487
low doses and doses rates, and determinants of clinical endpoints.
1488
1489
4.6.4 Interactions with Environmental and Lifestyle Factors
1490
1491
Lifetime risk estimates are reported for radiation exposures essentially considering the radiation
1492
exposure as the single stressor. Modulating factors have been identified to some extent and a few
1493
have been incorporated into the risk paradigm (e.g., age at exposure, sex, smoking for radon
1494
exposures) (Section 2.2). A major issue is to be able to identify those cancers associated with
1495
radiation exposure among the high frequency of background cancers, so that the impact of any
1496
modulating factor on the radiation component can be ascertained. This is currently an intractable
1497
problem since radiation-specific disease signatures are not known. The issue has been addressed by
1498
UNSCEAR (2000) and by Wakeford (2012). However, there is no simple approach to the problem.
1499
As a starting point it is necessary to be able to describe the nature of exposures to which individuals
1500
or populations are subject. Because the U.S. Environmental Protection Agency’s guidelines for
1501
estimation of risks to environmental stressors (chemicals and others) requires a consideration of
53
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
1502
aggregate or cumulative risk, methods have been proposed and are being developed for this purpose
1503
(EPA, 2003). A novel approach that takes advantage of current technical advances in exposure
1504
science, including biomonitoring, is the concept of a so-called exposome [see an overview at NIOSH
1505
(2014) and reviews by Rappaport (2011) and Vrijheid (2014)]. The exposome is broadly defined as a
1506
measure of the total exposure of an individual in their lifetime and the relationship of this exposure
1507
to health outcomes. In this context, exposure begins before birth and includes both environmental
1508
and occupational exposures. As noted by NIOSH (2014), “Understanding how exposures from our
1509
environment, diet, lifestyle etc. interact with our own unique characteristics such as genetics,
1510
physiology, and epigenetics impact our health” is how the exposome will be articulated. This issue is
1511
very pertinent to this Commentary for which ionizing radiation is just one of the environmental/
1512
occupational exposures and considerations of risk from radiation will need to be placed in the
1513
purview of total exposure and total risk. Exposomics is the name given to the study of the exposome
1514
and utilizes internal and external exposure methods.
1515
1516
As discussed by Rappaport (2011), it is appropriate for the application of an exposome-based
1517
approach for exposure assessment, to regard the “environment” as the body’s internal chemical (and
1518
physical) environment and to define “exposures” as levels of biologically active chemical and
1519
physical agents in and impacting this environment. He further proposes that to assess the exposome,
1520
the more applicable approach might be a top down one based on biomonitoring, rather than a
1521
bottom-up one that relies on analysis of exposures from external sources (e.g., air water, food). This
1522
general approach is extended by Wild et al. (2013) to provide a view of how an understanding of the
1523
complete exposure profiles of individuals in large populations can be linked to a detailed analysis of
1524
their omics profiles. As noted such approaches are in their early stages of development but already
1525
indicate the reasonable likelihood of linking total exposure measures to biological effects and
1526
subsequently adverse health outcomes. In the context of this Commentary, the “omics” component is
1527
intended to indicate that molecular methods based on omics techniques (e.g., genomics,
1528
transcriptomics, proteomics, and metabolomics) can be used in the selection of informative
1529
biomarkers and bioindicators of external exposure and internal dose. The more that is known about
1530
the etiology of diseases such as cancer, the more readily such informative biological surrogates can
1531
be identified.
1532
54
NCRP SC 1-21
Draft of March 31, 2015 [MR]
1533
NOT TO BE DISSEMINATED OR REFERENCED
5. Recommendations for Closing the Gaps towards an Integrated Approach
1534
1535
The current state of knowledge on the effects of low-dose radiation is incomplete and uncertain
1536
with respect to understanding the shape of the dose-response relationship and the level of risk at low
1537
doses. These deficiencies have long-term and profound consequences with regard to radiation
1538
protection guidance, compensation programs, and environmental contamination issues. There is a
1539
need to identify novel approaches to reduce these uncertainties. Research is needed to identify,
1540
evaluate, and utilize informative bioindicators of the health effects of interest (i.e., cancer and
1541
noncancer effects). Ideally, a bioindicator would be a quantitative predictor of the adverse health
1542
effect and measureable over the dose range of interest, particularly including low doses (i.e.,
1543
<100 mGy). Such information would extend the knowledge of the dose-response relationship at
1544
doses below those that can be observed through epidemiologic studies.
1545
1546
To be able to utilize an approach for incorporating radiation biology data into the risk assessment
1547
process for low-dose and low dose-rate exposures, it is necessary to establish a framework for
1548
research planning and data collection. Such a framework has been presented in this Commentary and
1549
is basically one that involves the use of adverse outcome pathways and key events defining such
1550
pathways. Research should address this need in a prospective manner rather than in a retrospective
1551
one. Lastly, to ensure progress in multi-disciplinary radiation research and continuity of low-dose
1552
radiation research in general, it is essential that low-dose facilities are maintained and broadened and
1553
educational training programs are expanded to meet the growing needs of cross-disciplinary
1554
radiation science in the next generation of radiation researchers.
1555
1556
5.1 Targeted Research to Identify Adverse Outcome Pathways and Bioindicators of Response
1557
1558
To utilize such an approach outlined above, it is essential to identify adverse outcome pathways
1559
for radiation-related adverse health outcomes and the key events that characterize these. The types of
1560
bioindicators that will be informative in this regard are discussed above in Section 4. It remains very
1561
difficult to develop the required information for human tumors because of the lack of ability to
1562
accurately identify radiation-related tumors. Thus, it is necessary, in part, to use animal models of
1563
human tumors together with cellular systems to support the evaluation of the potential predictive
55
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
1564
value of any prospective bioindicator. This general approach is depicted in the parallelogram
1565
presented in Figure 1.3. It has been proposed that adverse outcome pathways are likely to be
1566
conserved across species in support of integrated approaches to testing for chemical risks (Tollefsen
1567
et al., 2014)
1568
1569
It is initially necessary to establish the robustness of specific bioindicators and other intermediate
1570
endpoints for predicting the adverse outcome under study. Thus, it is required that any biological
1571
endpoint utilized for extrapolation purposes be predictive of the apical endpoint for which risk is
1572
being estimated. Such predictions can describe the shape of the dose-response curve or an estimate
1573
of the risk itself, depending on the nature of the specific biological endpoint. Of course, it is
1574
pertinent to establish, whenever possible, if there are additional dose-dependent modifications
1575
beyond the position of a particular bioindicator along the AOP. This can, in part, be overcome by
1576
using a suite of key-event bioindicators for steps along the AOP. As discussed in Sections 1 and 4,
1577
biomarkers are considered to be surrogates for key events in the formation of the apical endpoint,
1578
whereas bioindicators are defined as being key events themselves in an adverse outcome pathway for
1579
a particular adverse health outcome. For a quantitative assessment of risk, it is necessary to make use
1580
of bioindicators at dose levels below those at which the apical endpoint itself can be assessed with
1581
any degree of accuracy. In the case of cancer, it is also feasible to assess pre-neoplastic lesions as
1582
predictors of cancer outcome in some human studies. These are likely to be the most accurate
1583
predictors of a cancer outcome, given their proximity to the malignant cancer itself.
1584
1585
There are basically two classes of predictive biomarkers or bioindicators: those that can be
1586
predictive of the adverse health outcome (cancer or noncancer) and those that are predictive
1587
specifically of radiation-related cancer or noncancer outcomes. Knowledge of the cancer process has
1588
increased dramatically leading to the description of the Hallmarks of Cancer by Hanahan and
1589
Weinberg (2000; 2011) and the concept of gene alterations in cancer cells being classified as
1590
“drivers” and “passengers” (Pleasance et al., 2010). The use of this type of information forms the
1591
basis for the U.S. Environmental Protection Agency’s Guidelines for Carcinogen Risk Assessment
1592
(EPA, 2005). The approach is predicated on the fact that for the majority of chemicals (and exposure
1593
scenarios) there are very few epidemiologic data sets available other than those for occupational
1594
exposures. Thus, data from other sources (laboratory animal and in vitro cell studies) are used to
56
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
1595
support the epidemiology. The specific approach relies upon the assessment of key events in the
1596
cancer process to be used as bioindicators in a qualitative manner today, but is anticipated to be more
1597
quantitative in the near future. A similar use of key events has been proposed for noncancer
1598
endpoints using toxicity data (Julien et al., 2009; Seed et al., 2005). It would seem feasible for the
1599
general principles developed for these diverse set of exposures and disease endpoints to be applied
1600
for the enhancement of the radiation risk assessment process. The aim, in this case, would be to
1601
design research that specifically measured key events for a particular disease outcome — in the first
1602
instance, likely to be cancer given the wealth of knowledge on the underlying molecular mechanisms
1603
of cancer development.
1604
1605
At this time, it is not readily feasible to assess the robustness of surrogate markers or key events
1606
in the radiation (or chemical) risk assessment process because too little reliable information has been
1607
developed. There is some developing information that will prove useful with enhancement. For
1608
example, Jarabek et al. (2009) describe the use of specific DNA adducts as bioindicators for use in
1609
the cancer risk assessment process, and chromosome alterations have been used to predict cancer
1610
outcomes at the group (not the individual) level (Bonassi et al., 2008). The real problem is the lack
1611
of bioindicators that are specific for radiation exposure. The advent of ultra-high-throughput DNA
1612
sequencing and the identification of a limited set of genes that can be mutated in cancer cells [about
1613
120 according to Greenman et al. (2007)] could lead to the identification of radiation-specific
1614
signature alterations.
1615
1616
5.2 Biologically-Based Modeling
1617
1618
The limitations of the direct use of epidemiologic data for estimating cancer and noncancer risks
1619
at low radiation doses were discussed in (Sections 2.1, 2.2 and 2.3). In cases where such direct use
1620
has been presented (e.g., Kendall et al. 2013; Mathews et al., 2013; Pearce et al., 2012), possible
1621
confounders [in particular in relation to the brain cancer findings in the study of Pearce et al. (2012)]
1622
and modest power [in the case of the study of Kendall et al. (2013)] make interpretation somewhat
1623
problematic. The present Commentary is predicated on the potential use of biological data to support
1624
the current risk assessment approach. The specific application would be to enhance the extrapolation
1625
from available high/medium dose epidemiologic data to dose levels and dose rates that are necessary
57
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
1626
for radiation protection purposes (e.g., <100 mGy). In this regard, it has been proposed that some
1627
form of biologically-based dose-response modeling would meet this need (e.g., Conolly et al., 2004;
1628
Curtis et al., 2002; Little et al., 2008a; NCRP, 2012; Shuryak et al., 2010). However, here also,
1629
modeling on such a biological basis is uncertain outside the range of the available data that are used
1630
as input parameters to a biologically-based model.
1631
1632
5.2.1 Biologically-Based Dose-Response Approach
1633
1634
For risk assessment for environmental chemicals, because of the paucity of epidemiologic data, it
1635
is recommended that a biologically-based dose-response (BBDR) approach be used for predicting
1636
adverse health outcomes under low-dose chronic exposure scenarios (EPA, 2005). However, the
1637
approach has been used very sparingly in the regulatory process itself, largely because of the lack of
1638
availability of the appropriate data sets to provide input parameters to a BBDR model. The need for
1639
expanded use of BBDR models has been concisely expressed in Section 5.7 of NCRP (2012):
1640
1641
“The challenge of developing a biologically-based computational model to minimize
1642
uncertainty in dose-response modeling can be summarized as: understanding a
1643
sufficient amount of the relevant biology; acquiring enough data to parameterize the
1644
1645
1646
model; and developing the computational model.”
Of course, this simple statement hides a degree of complexity that must be addressed. For
1647
example, how much of the detailed biology do we need to know and when are there sufficient data to
1648
sustain model development? This situation can be significantly alleviated if research is directed
1649
towards the needs identified by BBDR models. Research targeted to the risk assessment process is
1650
needed.
1651
1652
The initial development of BBDR as applied to carcinogenesis is described by Moolgavkar and
1653
Knudson (1981) and Moolgavkar and Venzon (1979) in their two-stage cancer model [usually
1654
referred to as the Moolgavkar-Venzon-Knudson model]. This simplified model has been expanded
1655
upon in subsequent years, as reviewed in Little (2010). The basic principle is that carcinogenesis can
1656
be considered as a two-step process. These two critical steps in carcinogenesis are considered to be
58
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
1657
specific, irreversible, and inherited (at the cell level). The first step is assumed to give a small growth
1658
advantage through partial alteration of growth control, and the second is assumed to lead to total
1659
abrogation of growth control. Other assumptions include: that cancer arises from a single cell;
1660
transformation of stem cells are independent events; and the time from transformation of a single
1661
cell to a tumor is constant. Thus, what is developed is essentially a “working” model rather than one
1662
that is truly representative of the cancer process as described, for example, by Hanahan and
1663
Weinberg (2000; 2011).
1664
1665
The two-stage carcinogenesis model has been applied to a number of situations, for example, by
1666
Heidenreich et al. (2004) and Krewski et al. (2003) for radiation and by Conolly et al. (2004) for
1667
formaldehyde exposures. There remain some concerns that the use of this two-stage model does not
1668
make use of the considerable body of data that has been generated over the past 20 y on the
1669
understanding of cancer development and, in the present context, of radiation-related effects and
1670
radiation carcinogenesis. Multi-stage generalizations of the two-stage model have been developed
1671
(Little, 1995), and applied to model ionizing radiation exposure (Little, 1996; Little et al., 2002).
1672
Further generalizations of the model have been developed to incorporate genomic instability, and
1673
used for modeling population colon cancer data (Little and Li, 2007; Little and Wright, 2003; Little
1674
et al., 2008b). However, it is equally fair to note that no viable alternative BBDR has been proposed
1675
and applied to radiation risk assessment, with the possible exception of the multistage model
1676
developed by Shuryak et al. (2010) for the analysis of cancer risk patterns as a function of age-at-
1677
exposure in Japanese atomic-bomb survivors. The uncertainties inevitably associated with BBDR
1678
approaches are discussed in detail in NCRP (2012).
1679
1680
In addition to a simplified model approach (such as with the two-stage carcinogenesis model), it
1681
is feasible to use more realistic models of carcinogenesis, that is, models that establish if specific
1682
processes such as genomic instability or adaptive responses are predictive of the cancer response in
1683
specific studies (e.g., Eidemuller et al., 2011). Third, and perhaps the most informative type of
1684
BBDR model, are ones that are developed to examine the role of specific measured biological
1685
variables in the genesis of radiogenic cancer in a particular study. Examples of this informative
1686
approach have been reported by Heidenreich and Rosemann (2012) and Heidenreich et al. (2013).
59
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
1687
Further development of this latter type of model using key-event bioindicators as input parameters is
1688
encouraged.
1689
1690
5.2.2 Systems Biology
1691
1692
Systems biology, defined here as the quantitative study of biological networks and pathways as
1693
integrated systems rather than as a set of isolated parts, is a young but rapidly evolving field that
1694
takes a data-intensive approach to biologically-based modeling. Systems radiation biology
1695
emphasizes the interactions of cell signaling pathways and networks, the distribution of effects and
1696
responses throughout these networks, as well as the redundancy between and within such networks
1697
for maintenance of cellular homeostasis and subsequent tissue function. This approach uses ‘omics-
1698
level data (e.g., genomics, transcriptomics, proteomics and metabolomics) to reverse engineer the
1699
biological circuitry of a cell, tissue, or organism with the goal of being able to predict the response to
1700
a perturbation, such as the mutation of a gene that effectively removes a node from the network, or
1701
an extrinsic stress, such as radiation exposure. Ultimately, systems biology also attempts to consider
1702
tissue responses in the context of the whole irradiated system rather than as the response of
1703
individual cells.
1704
1705
While much is known about the DNA damage induced by the deposition of energy from
1706
exposure to ionizing radiation and the subsequent cellular responses, there is a significant gap in our
1707
understanding of how these might lead to detrimental health effects years after exposure. Current
1708
radiation modeling and risk evaluation tend to be limited to simplistic models that do not reflect the
1709
inherent biological diversity of humans or the mechanisms that underlie radiation risks (Barcellos-
1710
Hoff, 2008). For instance, the interactions between irradiated cells within a tissue and the
1711
modulating effects of the tissue microenvironment (Barcellos-Hoff, 2005) are not accounted for in
1712
many models. Linking mechanisms of cellular and molecular responses to radiation exposure to
1713
macroscopic processes at the tissue, organ, and whole organism levels would improve our ability to
1714
predict cancer risk in response to irradiation. This is our multiscale systems challenge: to understand
1715
the sequence of events occurring in tissues following irradiation and why in some instances exposure
1716
can lead to cancer or other phenotypic alterations, and in other instances no long-term phenotypic
1717
changes are observed.
60
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
1718
1719
Recent advances in high-throughput ‘omics technologies enable interrogation of induced
1720
radiation effects at the genomic [reviewed in Amundson (2008)], proteomic (Lin et al., 2009;
1721
Menard, et al. 2006; Park et al., 2006), and metabolomic (Lanz et al., 2009; Tyburski et al., 2008;
1722
2009; Zhang, et al., 2014) levels. Radiation systems biology attempts to integrate the resulting
1723
‘omics data to enable modeling of radiation dose and dose-rate effects in a complex tissue as a
1724
function of time, and relate these to observable functional physiological and /or physical changes in
1725
that tissue after irradiation.
1726
1727
It is clear that vast amounts of data can, and will, be generated. The usefulness of these data will
1728
depend on the development of computational modeling approaches to organize an abundance of
1729
complex biological data in a predictive, multicellular framework. The challenge is to understand the
1730
flow of information between cell types in a given tissue and why different tissues vary in their
1731
response to radiation. This demands approaches that can elucidate the connectivity among and
1732
between signaling networks rather than focusing on reducing analysis to individual signaling
1733
components.
1734
1735
Network construction of radiation responses will require quality, dose-dependent measurements
1736
over a protracted time frame (Eschrich et al., 2009). Multiple doses, dose rates and time points will
1737
need to be interrogated at the global level in order to address the question at hand. The impact of
1738
post-transcriptional and post-translational modifications as well as potential epigenetic modifications
1739
that influence cell signaling will have to be evaluated and incorporated into the model. In this way
1740
the combined behavior of interdependent processes that dictate cell responses and the fundamental
1741
properties of the networks involved, such as redundancy, modularity, robustness, and feedback
1742
1743
1744
1745
control can be assessed.
The systems biology approach for radiation risk assessment will require:
1746

multi-level ‘omics data from well-designed experiments;
1747

knowledge of the pathways involved;
1748

developing the computational modeling approaches to organize complex biological data in a
1749
predictive, multi-cellular framework that can be integrated with radiation risk models;
61
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
1750

a long-term commitment from federal funding agencies; and
1751

dedicated cross-disciplinary teams of investigators to link studies at all levels to relevant
1752
1753
1754
1755
1756
health outcomes.
5.3 Registries of Medical-Imaging Exposures Linked to Outcomes and Biological Materials
The use of ionizing radiation for medical imaging has grown significantly in recent years. In the
1757
United States, exposure to ionizing radiation from medical-imaging procedures3 in 2006 was
1758
approximately six times greater than in 1980 (NCRP, 2009d), with an average effective dose to a
1759
member of the U.S. population of ~3 mSv, comparable to the annual effective dose from natural
1760
background radiation in 2006; the dose from natural background was the same as in 1980. This
1761
increased dose from medical-imaging procedures was attributed to the greatly increased usage in
1762
2006 of higher-dose procedures particularly computed tomography (CT) and nuclear medicine
1763
compared to 1980 (NCRP, 2009d). Brenner and Hall (2007) reported that almost 82 million CT
1764
procedures were performed annually in the United States in 2011, up from 46 million in 2000 and
1765
13 million in 1990. Mettler et al. (2009), reported that cardiac diagnostic nuclear procedures
1766
increased from 1 % of the total number of diagnostic nuclear medicine examinations performed in
1767
1973 to 57 % in 2005). Further, Brenner and Hricak (2010) state that an increase in emerging
1768
imaging modalities such as positron-emission tomography CT, single-photon emission CT, and
1769
potentially CT for screening of high-risk asymptomatic patients (e.g., smokers screened for early
1770
1771
1772
lung cancer detection) are likely.
1773
patient demographics, as available in other countries (e.g., Nordic countries) could be useful for
1774
tracking long-term sequelae of these procedures, including the potential for carcinogenic risk
1775
associated with extensive medical imaging using these technologies. An ideal situation would also
1776
include the collection and storage of biospecimens. In addition, there may be other potentially
1777
informative radiation-exposed groups, such as astronauts, where follow-up and collection of samples
1778
would be valuable. However, such a comprehensive endeavor may not be economically feasible as
Therefore registries of medical-imaging procedures, including indication for examination and
3
Medical-imaging procedures included computed tomography, converntional radiography and fluoroscopy,
interventional fluoroscopy, and nuclear medicine.
62
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
1779
the cost of establishing and maintaining the registry, harmonizing the input and follow-up
1780
1781
1782
1783
1784
1785
1786
procedures and the complex issues of sample collection and storage may be extensive.
1787
information and strong quantitative data to inform risk assessments for human exposure. This
1788
research requires the availability of facilities that can provide accurate dose delivery at the relevant
1789
dose-rates. The necessary facilities have variable requirements. Some must be able to accommodate
1790
cell culture systems, including tissue-mimetic models, and these require ancillary equipment to
1791
provide detailed dosimetry and to maintain defined oxygen tensions. Some animal work will be
1792
possible on a small scale using these facilities, but larger-scale, dedicated facilities with the
1793
capability to conduct long-term exposures at low dose rates are critical for some aspects of this
1794
work. Specialized facilities are also needed for studies with low-energy sources, charged-particle
1795
microbeams, and inhaled sources, and to investigate long-term low dose-rate effects. Support for the
1796
maintenance and development of this diverse group of facilities is essential to the conduct of a robust
1797
1798
1799
1800
1801
research program at low doses and low dose rates.
5.4.2 Interdisciplinary Training
1802
critical elements that are required to support this highly interdisciplinary field (Boice, 2014b; NCRP,
1803
2015). Individuals trained across disciplines are essential to meet the health needs of the nation as
1804
we evaluate radiation risk. Current training programs are often narrowly focused and rarely include
1805
the breadth of coursework and practical training needed to develop robust research programs.
1806
Integrated training programs with representation of the disciplines of radiation biology, molecular
1807
biology, pathology, radiation dosimetry, health physics, epidemiology, and statistics should be
1808
developed and supported to ensure the availability of a cadre of scientists who can advance our
1809
understanding of radiation health effects at low doses.
5.4 Facilities and Interdisciplinary Training
5.4.1 Facilities
Basic and applied research studies at low doses and low dose rates can provide mechanistic
The science of radiation protection will continue to develop only if individuals are trained in
1810
63
NCRP SC 1-21
Draft of March 31, 2015 [MR]
1811
NOT TO BE DISSEMINATED OR REFERENCED
Glossary
1812
1813
absorbed dose (D): The quotient of dε by dm, where dε is the mean energy imparted to matter of mass dm
1814
(i.e., D = dε /dm). The unit for D is joule per kilogram (J kg–1) with the special name gray (Gy) (see also
1815
organ dose).
1816
1817
1818
1819
1820
1821
ascertainment bias: Arises when there is variation in ascertainment of disease status, perhaps correlated with
exposure variables.
adaptive response: In radiobiology, the theory that a low dose of ionizing radiation can reduce the effect of a
higher dose of ionizing radiation when the higher dose is administered after a short time delay.
Adult Health Study (AHS): A subcohort of the Life Span Study, which provides biennial medical follow-up
for a subcohort of about 20,000 members of the Life Span Study (see Life Span Study).
1822
as low as reasonably achievable (ALARA): A principle of radiation protection philosophy that requires that
1823
exposures to ionizing radiation should be kept as low as reasonably achievable, economic and social
1824
factors being taken into consideration. The protection from radiation exposure is ALARA when the
1825
expenditure of further resources would be unwarranted by the reduction in exposure that would be
1826
achieved.
1827
background radiation: The amount of radiation to which a member of the population is exposed from
1828
natural sources, such as terrestrial radiation from naturally-occurring radionuclides in the soil, cosmic
1829
radiation originating in outer space, and naturally-occurring radionuclides deposited in the human body.
1830
The natural background radiation received by an individual depends on geographic location and living
1831
habits. In the United States, the average effective dose from background radiation is ~1 mSv y–1,
1832
excluding indoor radon which amounts to ~2 mSv y–1 on average.
1833
Berkson error: The Berkson error model assumes that true dose is equal to the observed dose plus an error
1834
introduced by individual peculiarity, which is defined as a random variable that has a mean of zero and is
1835
independent of observed dose.
1836
bias (epidemiology): Any process at any stage in the conduct of the study that tends to produce results or
1837
conclusions that differ systematically from the truth (see follow-up bias, ascertainment bias, recall bias,
1838
and confounding factor).
1839
bioindicator: A cellular alteration that is on the pathway to the disease endpoint itself, such as a specific
1840
mutation in a target cell that is associated with tumor formation. Thus, a bioindicator can be perceived as
1841
informing on the shape of dose-response curve for the disease outcome or on cancer frequency itself, and
1842
therefore, is equivalent to a key event.
1843
1844
biomarker: A biological phenotype (e.g., chromosome alteration, DNA adduct, gene expression change,
specific metabolite) that can be used to indicate a response at the cell or tissue level to an exposure. In this
64
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
1845
regard, it is generally a measure of the potential for development of an adverse outcome such as cancer
1846
(e.g., a predictor of exposure level).
1847
bystander effect: In radiobiology, the term is used to describe an effect on cells in which the energy had not
1848
been directly deposited. In most instances, the cells so affected were neighbors of the cells directly
1849
impacted by the radiation.
1850
1851
cancer: A general term for more than 100 diseases characterized by abnormal and uncontrolled growth of
cells.
1852
carcinogen: An agent that is associated with an increased risk of cancer.
1853
carcinogenesis: Induction of cancer by radiation or any other agent (a somatic effect).
1854
case-control study: An epidemiologic study in which people with disease and a similarly composed control
1855
1856
group are compared in terms of exposures to a putative causative agent.
Chernobyl recovery workers: The civil and military personnel cleanup workers who were called upon to
1857
deal with consequences of the 1986 Chernobyl nuclear disaster on the site of the event in the
1858
former Soviet Union.
1859
Classical measurement error: The Classical measurement error model assumes that observed dose equals
1860
true dose plus measurement error, where measurement error is a random variable that is independent of
1861
the true dose and has a mean of zero.
1862
cohort study: An epidemiologic study in which groups of people (the cohort) are identified with respect to
1863
the presence or absence of exposure to a disease-causing agent, and in which the outcomes of disease
1864
rates are compared; also called a follow-up study.
1865
confounding factor: In statistics, an extraneous variable in a statistical model that correlates (directly or
1866
inversely) with both the dependent variable and the independent variable (e.g., a factor that is correlated
1867
both with the disease under study and with an exposure of interest).
1868
1869
1870
1871
1872
chromosome aberration: A missing, extra, or irregular portion of chromosomal DNA. It can be from an
atypical number of chromosomes or a structural abnormality in one or more chromosomes.
cytogenetic: Relating to the branch of genetics that is concerned with the study of the structure and function
of the cell, especially the chromosomes.
deoxyribonucleic acid (DNA): Genetic material of cells; a complex molecule of high molecular weight
1873
consisting of deoxyribose, phosphoric acid, and four bases which are arranged as two long chains that
1874
twist around each other to form a double helix joined by hydrogen bonds between the complementary
1875
components.
1876
1877
DNA damage: An alteration in the chemical structure of DNA, such as a break in a strand of DNA, a base
missing from the backbone of DNA, or a chemically changed base.
65
NCRP SC 1-21
Draft of March 31, 2015 [MR]
1878
1879
1880
NOT TO BE DISSEMINATED OR REFERENCED
dose: A general term used when the context is not specific to a particular dose quantity. When the context is
specific, the name for the quantity is used (e.g., absorbed dose).
dose and dose-rate effectiveness factor (DDREF): A judged factor that generalizes the usually lower
1881
biological effectiveness (per unit of dose) of radiation exposures at low doses and low dose rates as
1882
compared with exposures at high doses and high dose rates.
1883
dose limit: A limit on radiation dose that is applied for exposure to individuals or groups of individuals in
1884
order to prevent the occurrence of radiation-induced tissue reactions or to limit the probability of
1885
radiation-related stochastic effects to an acceptable level (see tissue reaction and stochastic effect).
1886
dose rate: Dose delivered per unit time. Can refer to any dose quantity (e.g., organ dose).
1887
dose reconstruction: Retrospective assessment of dose to identifiable or representative individuals or
1888
populations by any means, especially absorbed dose to specific organs or tissues.
1889
dose response: The way in which an effect, or response, depends on dose.
1890
effective dose: The sum over specified tissues of the products of the equivalent dose in a tissue or organ and
1891
the tissue weighting factor for that tissue or organ. Equivalent dose is the product of the radiation
1892
weighting factor and the mean absorbed dose in a tissue or organ (organ dose).The SI unit for effective
1893
dose is joule per kilogram (J kg–1), with the special name sievert (Sv). A similar quantity used by the
1894
Nuclear Regulatory Commission is named total effective dose equivalent.
1895
1896
electron: Subatomic charged particle. Negatively charged particles are parts of stable atoms. Both negatively
and positively charged electrons may be expelled from the radioactive atom when it disintegrates.
1897
exposome: (see exposomics).
1898
exposomics: The name given to the study of the exposome. The exposome is broadly defined as a measure of
1899
the total exposure of an individual in a lifetime and the relationship of this exposure to health outcomes.
1900
The “omics” component is intended to indicate that molecular methods based on omics techniques (e.g.,
1901
genomics, transcriptomics, proteomics, and metabolomics) can be used in the selection of informative
1902
biomarkers and bioindicators of the exposure.
1903
excess absolute risk (EAR): The difference in absolute risk between exposed and control populations.
1904
excess relative risk (ERR): The ratio of the excess risk of a specified disease to the probability of the same
1905
effect in the unexposed population (i.e., the relative risk minus one). The relative is ratio of the incidence
1906
rate of a given disease in those exposed to the incidence rate of that disease in those not exposed.
1907
exposure: In this Commentary, exposure is used in a general sense meaning to be irradiated.
1908
follow-up bias: Arises when there is a lack of follow-up information, for example if persons have, unknown
1909
to the investigator, migrated outside of the study area, so that their health status cannot be reported.
1910
gamma rays: Short-wavelength electromagnetic radiation of nuclear origin (approximate range of energy:
1911
10 keV to 9 MeV).
66
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
1912
genomic instability (delayed genetic effects): An increased rate of acquiring genetic change.
1913
genotoxin: Any substance capable of causing damage to cellular DNA and thus
1914
producing mutations or cancer.
1915
gray (Gy): The International System (SI) unit of absorbed dose of radiation, 1 Gy = 1 J kg–1.
1916
germline mutation: Any detectable and heritable variation in the lineage of germ cells. Mutations in these
1917
cells are transmitted to offspring,
1918
incidence: The rate of occurrence of a disease, usually expressed in number of cases per million.
1919
ionizing radiation: Any electromagnetic or particulate radiation capable of producing ions, directly or
1920
1921
1922
1923
1924
indirectly, in its passage through matter.
key event: An empirically observable precursor step that is itself a necessary element of the mode of action or
is a biologically-based marker for such an element (see also bioindicator).
Life Span Study (LSS): Study of the Japanese atomic-bomb survivors; the sample consists of 93,741 persons
(in city) of whom 86,572 survivors have a defined dose assigned to them.
1925
lifetime risk: The probability during one’s lifetime of expressing a given health outcome.
1926
linear energy transfer (LET): Average amount of energy lost per unit of particle track length and expressed
1927
in keV μm–1.
1928
low-LET: Radiation having a low linear energy transfer; for example, electrons, x rays, and gamma rays.
1929
high-LET: Radiation having a high linear energy transfer; for example, alpha particles and interaction
1930
products of fast neutrons.
1931
linear-nonthreshold model: In radiation protection, a model that assumes that the long term, biological
1932
damage caused by ionizing radiation (essentially the cancer risk) is directly proportional to the dose.
1933
linkage study: A study that looks for patterns of inheritance of genetic markers in large families in which a
1934
1935
particular condition is common.
mode of action: For both cancer and noncancer endpoints, a sequence of key events and processes starting
1936
with interaction of an agent with a cell, proceeding through operational and anatomical changes, and
1937
resulting in formation of an adverse outcome (see key event).
1938
Monte Carlo simulation: Computation of a probability distribution of an output of a model on the basis of
1939
repeated calculations using random sampling of input variables from specified probability distributions.
1940
natural background radiation (see background radiation):
1941
neutrons: Particles with a mass similar to that of a proton, but with no electrical charge. Because they are
1942
1943
1944
electrically neutral, they cannot be accelerated in an electrical field.
noncancer: Health effects other than cancer (e.g., cataracts, cardiovascular disease) that occur in the exposed
individual.
67
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
1945
nonionizing radiation: Electromagnetic radiation that includes the ultraviolet, visible, infrared, microwave,
1946
radiofrequency, and extremely-low-frequency portions of the electromagnetic spectrum. Unlike ionizing
1947
radiation, nonionizing radiation is unable to ionize atoms in its interactions with matter.
1948
1949
1950
1951
1952
1953
1954
occupational exposure: The exposure received by an individual in a restricted area, or in the course of
employment in which the individual’s duties necessarily involve exposure to radiation.
’omics: Refers to molecular methods for measuring all of a certain molecular specie in a cell (e.g., genomics,
transcriptomics, proteomics, and metabolomics).
organ dose: The mean absorbed dose in an organ or tissue, obtained by integrating or averaging absorbed
doses at points in the organ or tissue.
parallelogram approach: As initially applied to the estimation of heritable risk: the principle that genetic
1955
damage in human germ cells can be estimated by measuring a common end point in humans and animals,
1956
such as mutations or chromosomal aberrations in lymphocytes, and a corresponding genetic end point in
1957
germ cells of animals, the desired target tissue generally inaccessible in man.
1958
partial body: Exposure to only part of a body, as opposed to exposure of the whole body.
1959
polygene: A group of nonallelic genes that together influence a phenotypic trait.
1960
quality factor: The factor by which absorbed dose at a point is modified in order to express the effectiveness
1961
of a given type of ionizing radiation in inducing stochastic effects.
1962
radiation: (1) The energy propagated as waves; radiation or radiant energy, when unqualified, usually refers
1963
to electromagnetic radiation; commonly classified by frequency (e.g., infrared, visible, ultraviolet, x rays,
1964
and gamma rays), and (2) corpuscular emission, such as alpha and beta particles.
1965
radiation weighting factor: A factor used to allow for differences in the biological effectiveness between
1966
different radiations when calculating equivalent dose. These factors are independent of the tissue or organ
1967
irradiated. Equivalent dose is the product of the radiation weighting factor and the mean absorbed dose in
1968
a tissue or organ (organ dose).
1969
1970
radiosensitivity: The relative susceptibility of cells, tissues, organs or organisms to the harmful effect
of ionizing radiation.
1971
recall bias: Arises when information, for example on exposure, is collected retrospectively, and patients, or
1972
their relatives, are subject to differential recall of this information, depending on their disease status.
1973
relative biological effectiveness: A factor used to compare the biological effectiveness of absorbed doses
1974
from different types of ionizing radiation, determined experimentally. Relative biological effectiveness is
1975
the ratio of the absorbed dose of a reference radiation (often taken as 250 kVp x rays) to the absorbed
1976
dose of the radiation in question required to produce an identical biological effect in a particular
1977
experimental organism or tissue.
1978
risk: The probability of a specified effect or response occurring.
68
NCRP SC 1-21
Draft of March 31, 2015 [MR]
1979
1980
NOT TO BE DISSEMINATED OR REFERENCED
risk estimate: The number of cases (or deaths) that are projected to occur in a specified exposed population
per unit dose for a defined exposure regime and expression period.
1981
sievert (Sv): The special name for the unit of effective dose. 1 Sv = 1 J kg–1.
1982
somatic mutation: A mutation occurring in a somatic cell as opposed to a germ cell.
1983
stochastic effect: An effect for which the probability of occurrence, rather than its severity, is a function of
1984
1985
1986
1987
1988
1989
radiation dose without threshold (e.g., cancer).
systems biology: The quantitative study of biological networks and pathways as integrated systems rather
than as a set of isolated parts.
tissue reaction: Injury in population of cells, characterized by a threshold dose and an increase in the severity
of the reaction as the dose is increased further.
tissue weighting factor: A factor representing the ratio of risk of stochastic effects attributable to
1990
irradiation of a given organ or tissue to the total risk when the whole body is irradiated
1991
uniformly. The factor is assumed to be independent of the type of radiation or energy of the
1992
radiation.
1993
total effective dose equivalent: (see effective dose).
1994
uncertainty: Lack of sureness or confidence in predictions of models or results of measurements.
1995
Uncertainties may be categorized as those due to stochastic variation, or as those due to lack of
1996
knowledge founded on an incomplete characterization, understanding or measurement of a system.
1997
69
NCRP SC 1-21
Draft of March 31, 2015 [MR]
1998
NOT TO BE DISSEMINATED OR REFERENCED
Symbols, Abbreviations and Acronyms
1999
2000
AHS
Adult Health Study (Japanese atomic-bomb survivors)
2001
ALARA
as low as reasonably achievable
2002
AML
acute myeloid leukemia
2003
ATM
ataxia-telangiectasia mutated
2004
BBDR
biologically-based dose-response
2005
BEIR
Committee on the Biological Effects of Ionizing Radiation
2006
CT
computed tomography
2007
DDREF
dose and dose-rate effectiveness factor
2008
DNA
deoxyribonucleic acid
2009
DSBs
double-strand breaks (DNA)
2010
EAR
excess absolute risk
2011
ERR
excess relative risk
2012
GWAS
genome-wide association study(ies)
2013
LET
linear energy transfer
2014
LSS
Life Span Study (Japanese atomic-bomb survivors)
2015
SNP
single nucleotide polymorphism
2016
SSBs
single-strand breaks (DNA)
2017
TERT, hTERT
telomerase reverse transcriptase (hTERT in humans)
2018
WARP
Where are the Radiation Professionals?
2019
WECARE
Women’s Environmental Cancer and Radiation Epidemiology (Study)
2020
70
NCRP SC 1-21
Draft of March 31, 2015 [MR]
2021
NOT TO BE DISSEMINATED OR REFERENCED
References
2022
2023
ADELEYE, Y., ANDERSEN, M., CLEWELL, R., DAVIES, M., DENT, M., EDWARDS, S., FOWLER, P.,
2024
MALCOMBER, S., NICOL, B., SCOTT, A., SCOTT, S., SUN, B., WESTMORELAND, C., WHITE, A.,
2025
ZHANG, O. and CARMICHAEL, P.L. (2014). “Implementing Toxicity Testing in the 21st Century
2026
(TT21C): Making safety decisions using toxicity pathways and progress in a prototype risk assessment,”
2027
Toxicology pii: S0300-483X(14)00032-8. doi: 10.1016/j.tox.2014.02.007 (Epub ahead of print).
2028
2029
2030
AINSBURY, E.A., BOUFFLER, S.D., DORR, W., GRAW, J., MUIRHEAD, C.R., EDWARDS, A.A. and
COOPER, J. (2009). “Radiation cataractogenesis: A review of recent studies,” Radiat. Res. 172(1), 1–9.
AINSBURY, E.A., BAKHANOVA, E., BARQUINERO, J.F., BRAI, M., CHUMAK, V., CORRECHER, V.,
2031
DARROUDI, F., FATTIBENE, P., GRUEL, G., GUCLU, I., HORN, S., JAWORSKA, A., KULKA, U.,
2032
LINDHOLM, C., LLOYD, D., LONGO, A., MARRALE, M., MONTEIRO GIL, O., OESTREICHER,
2033
U., PAJIC, J., RAKIC, B., ROMM, H., TROMPIER, F., VERONESE, I., VOISIN, P., VRAL, A.,
2034
WHITEHOUSE, C.A., WIESER, A., WODA, C., WOJCIK, A. and ROTHKAMM, K. (2011). “Review
2035
of retrospective dosimetry techniques for external ionising radiation exposures,” Radiat. Prot. Dosim.
2036
147(4), 573–592.
2037
2038
2039
2040
2041
2042
2043
2044
ALLEN, C., ASHLEY, A.K., HROMAS, R. and NICKOLOFF, J.A. (2011). “More forks on the road to
replication stress recovery,” J. Mol. Cell. Biol. 3(1), 4–12.
AMUNDSON, S.A. (2008). “Functional genomics in radiation biology: A gateway to cellular systems-level
studies,” Radiat. Environ. Biophys. 47(1), 25–31.
AMUNDSON, S.A. and CHEN, D.J. (1996). “Inverse dose-rate effect for mutation induction by gamma-rays
in human lymphoblasts,” Int. J. Radiat. Biol. 69(5), 555–563.
AYPAR, U., MORGAN, W.F. and BAULCH, J.E. (2011). “Radiation-induced genomic instability: Are
epigenetic mechanisms the missing link?,” Int. J. Radiat. Biol. 87(2), 179–191.
2045
AZIZOVA, T.V., MUIRHEAD, C.R., DRUZHININA, M.B., GRIGORYEVA, E.S., VLASENKO, E.V.,
2046
SUMINA, M.V., O’HAGAN, J.A., ZHANG, W., HAYLOCK, R.G. and HUNTER, N. (2010a).
2047
“Cardiovascular diseases in the cohort of workers first employed at Mayak PA in 1948–1958,” Radiat.
2048
Res. 174(2), 155–168.
2049
AZIZOVA, T.V., MUIRHEAD, C.R., DRUZHININA, M.B., GRIGORYEVA, E.S., VLASENKO, E.V.,
2050
SUMINA, M.V., O’HAGAN, J.A., ZHANG, W., HAYLOCK, R.G. and HUNTER, N. (2010b).
2051
“Cerebrovascular diseases in the cohort of workers first employed at Mayak PA in 1948–1958,” Radiat.
2052
Res. 174(6), 851–864.
2053
2054
AZIZOVA, T.V., MUIRHEAD, C.R., MOSEEVA, M.B., GRIGORYEVA, E.S., SUMINA, M.V.,
O'HAGAN, J., ZHANG, W., HAYLOCK, R.J. and HUNTER, N. (2011). “Cerebrovascular diseases in
71
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
2055
nuclear workers first employed at the Mayak PA in 1948–1972,” Radiat. Environ. Biophys. 50(4), 539–
2056
552.
2057
AZIZOVA, T.V., HAYLOCK, R.G., MOSEEVA, M.B., BANNIKOVA, M.V. and GRIGORYEVA, E.S.
2058
(2014). “Cerebrovascular diseases incidence and mortality in an extended Mayak Worker Cohort
2059
1948–1982,” Radiat. Res. 182(5), 529–544.
2060
AZZAM, E.I., DE TOLEDO, S.M., RAAPHORST, G.P. and MITCHEL, R.E. (1996). “Low-dose ionizing
2061
radiation decreases the frequency of neoplastic transformation to a level below the spontaneous rate in
2062
C3H 10T1/2 cells,” Radiat. Res. 146(4), 369–373.
2063
BARBER, R., PLUMB, M.A., BOULTON, E., ROUX, I. and DUBROVA, Y.E. (2002). “Elevated mutation
2064
rates in the germ line of first- and second-generation offspring of irradiated male mice,” Proc. Natl. Acad.
2065
Sci. USA 99(10), 6877–6882.
2066
2067
2068
2069
2070
2071
2072
BARCELLOS-HOFF, M.H. (2005). “Integrative radiation carcinogenesis: Interactions between cell and
tissue responses to DNA damage,” Semin. Cancer Biol. 15(2), 138–148.
BARCELLOS-HOFF, M.H. and COSTES, S.V. (2006). “A systems biology approach to multicellular and
multi-generational radiation responses,” Mutat. Res. 597(1–2), 32–38.
BARCELLOS-HOFF, M.H. (2008). “Cancer as an emergent phenomenon in systems radiation biology,”
Radiat. Environ. Biophys 47(1), 33–38.
BARQUINERO, J.F., BARRIOS, L., CABALLIN, M.R., MIRO, R., RIBAS, M., SUBIAS, A. and
2073
EGOZCUE, J. (1995). “Occupational exposure to radiation induces an adaptive response in human
2074
lymphocytes,” Int. J. Radiat. Biol. 67(2), 187–191.
2075
BELYAKOV, O.V., MITCHELL, S.A., PARIKH, D., RANDERS-PEHRSON, G., MARINO, S.A.,
2076
AMUNDSON, S.A., GEARD, C.R. and BRENNER, D.J. (2005) “Biological effects in unirradiated
2077
human tissue induced by radiation damage up to 1 mm away,” Proc. Natl. Acad. Sci. USA, 102(40),
2078
14203–14208.
2079
BELYAKOV, O.V., FOLKARD, M., MOTHERSILL, C., PRISE, K.M. and MICHAEL, B.D. (2006).
2080
“Bystander-induced differentiation: A major response to targeted irradiation of a urothelial explant
2081
model,” Mutat. Res. 597(1–2), 43–49.
2082
BHATTI, P., DOODY, M.M., PRESTON, D.L., KAMPA, D., RON, E., WEINSTOCK, R.W., SIMON, S.,
2083
EDWARDS, A.A. and SIGURDSON, A.J. (2008). “Increased frequency of chromosome translocations
2084
associated with diagnostic x-ray examinations,” Radiat. Res. 170(2), 149–155.
2085
BISSELL, M.J., WARNER, H.R., BERGET, S.M., FRY, R.J.M., HANAWALT, P.C., KASTAN, M.,
2086
KORNBERG, A., LUTZE-MANN, L., SOUZA, K.A., ULLRICH, R., VIJG, J. and CHATTERJEE, A.
2087
(1997). Modeling Human Risk: Cell and Molecular Biology in Context, DE-97007347, LBNL-40278
2088
(National Technical Information Service, Springfield, Virginia).
72
NCRP SC 1-21
Draft of March 31, 2015 [MR]
2089
2090
2091
2092
2093
2094
2095
2096
2097
2098
2099
NOT TO BE DISSEMINATED OR REFERENCED
BITHELL, J.F. and STEWART, A.M. (1975). “Pre-natal irradiation and childhood malignancy: A review of
British data from the Oxford Survey,” Br. J. Cancer 31(3), 271–287.
BLYTH, B.J. and SYKES, P.J. (2011). “Radiation-induced bystander effects: What are they, and how
relevant are they to human radiation exposures?” Radiat. Res. 176(2), 139–157.
BOICE, J.D., JR. (2012). “A study of one million U.S. radiation workers and veterans: A new NCRP
initiative (DOE Grant Awarded September 2012),” Health Phys. News XL(11), 7–10.
BOICE, J.D., JR. (2014a). “Invited editorial: The importance of radiation worker studies,” J. Radiol. Prot.
34(2014), E7–E12.
BOICE, J.D., JR. (2014b). “WARP—where are the radiation professionals?” Health Phys. News XLLII(4),
22.
BOICE, J.D., JR., BLETTNER, M., KLEINERMAN, R.A., STOVALL, M., MOLONEY, W.C.,
2100
ENGHOLM, G., AUSTIN, D.F., BOSCH, A., COOKFAIR, D.L. and KREMENTZ, E.T. (1987).
2101
“Radiation dose and leukemia risk in patients treated for cancer of the cervix,” J. Natl. Cancer Inst. 79(6),
2102
1295–1311.
2103
BOICE, J.D., JR., TAWN, E.J., WINTHER, J.F., DONALDSON, S.S., GREEN, D.M., MERTENS, A.C.,
2104
MULVIHILL, J.J., OLSEN, J.H., ROBISON, L.L. and STOVALL, M. (2003). “Genetic effects of
2105
radiation therapy for childhood cancer,” Health Phys. 85(1), 65–80.
2106
BONASSI, S., HAGMAR, L., STROMBERG, U., MONTAGUD, A.H., TINNERBERG, H., FORNI, A.,
2107
HEIKKILA, P., WANDERS, S., WILHARDT, P., HANSTEEN, I.L., KNUDSEN, L.E. and NORPPA,
2108
H. (2000). “Chromosomal aberrations in lymphocytes predict human cancer independently of exposure to
2109
carcinogens,” Cancer Res. 60(6), 1619–1625.
2110
BONASSI, S., NORPPA, H., CEPPI, M., STROMBERG, U., VERMEULEN, R., ZNAOR, A., CEBULSKA-
2111
WASILEWSKA, A., FABIANOVA, E., FUCIC, A., GUNDY, S., HANSTEEN, I.L., KNUDSEN, L.E.,
2112
LAZUTKA, J., ROSSNER, P., SRAM, R.J. and BOFFETTA, P. (2008). “Chromosomal aberration
2113
frequency in lymphocytes predicts the risk of cancer: Results from a pooled cohort study of 22 358
2114
subjects in 11 countries,” Carcinogenesis 29(6), 1178–1183.
2115
BOOBIS, A.R., COHEN, S.M., DELLARCO, V., MCGREGOR, D., MEEK, M.E., VICKERS, C.,
2116
WILLCOCKS, D. and FARLAND, W. (2006). “IPCS framework for analyzing the relevance of a cancer
2117
mode of action for humans,” Crit. Rev. Toxicol. 36(10), 781–792.
2118
2119
2120
2121
BOSI, A. and OLIVIERI, G. (1989). “Variability of the adaptive response to ionizing radiations in humans,”
Mutat. Res. 211(1), 13–17.
BOSS, M., BRISTOW, R. and DEWHIRST, M. (2014). “Linking the history of radiation biology to the
hallmarks of cancer,” Radiat. Res. 181(6), 561–577.
73
NCRP SC 1-21
Draft of March 31, 2015 [MR]
2122
NOT TO BE DISSEMINATED OR REFERENCED
BOUFFLER, S.D., BRIDGES, B.A., COOPER, D.N., DUBROVA, Y., MCMILLAN, T.J., THACKER, J.,
2123
WRIGHT, E.G. and WATERS, R. (2006). “Assessing radiation-associated mutational risk to the
2124
germline: Repetitive DNA sequences as mutational targets and biomarkers,” Radiat. Res. 165(3), 249–
2125
268.
2126
BOUVILLE, A., TOOHEY, R.E., BOICE, J.D., JR., BECK, H.L., DAUER, L.T., ECKERMAN, K.F.,
2127
HAGEMEYER, D., LEGGETT, R.W., MUMMA, M., NAPIER, B., PRYOR, K.H., ROSENSTEIN, M.,
2128
SCHAUER, D.A., SHERBINI, S., STRAM, D.O., THOMPSON, J.L., TILL, J.E., YODER, C. and
2129
ZEITLIN, C. (2015). “Dose reconstruction for the million worker study: Status and guidelines,” Health
2130
Phys. 108(2), 206–220.
2131
BOZIC, I., ANTAL, T., OHTSUKI, H., CARTER, H., KIM, D., CHEN, S., KARCHIN, R., KINZLER,
2132
K.W., VOGELSTEIN, B. and NOWAK, M.A. (2010). “Accumulation of driver and passenger mutations
2133
during tumor progression,” Proc. Natl. Acad. Sci. (USA) 107(43), 18545–18550.
2134
2135
2136
2137
2138
2139
2140
2141
2142
2143
2144
BRANZEI, D. and FOIANI, M. (2009). “The checkpoint response to replication stress,” DNA Repair (Amst.)
8(9), 1038–1046.
BRENNER, D.J. (1994). “The significance of dose rate in assessing the hazards of domestic radon exposure,”
Health Phys. 67(1), 76–79.
BRENNER, D.J. and HALL, E.J. (2007). “Computed tomography—an increasing source of radiation
exposure,” N. Engl. J. Med. 357(22), 2277–2284.
BRENNER, D.J. and HRICAK, H. (2010). “Commentary: Radiation exposure from medical imaging: Time
to regulate?” JAMA 304(2), 208–209.
BRENNER, D.J. and SACHS, R.K. (2003). “Domestic radon risks may be dominated by bystander effects—
but the risks are unlikely to be greater than we thought,” Health Phys. 85(1), 103–108.
BRENNER, D.J., HAHNFELDT, P., AMUNDSON, S.A. and SACHS, R.K. (1996). “Interpretation of
2145
inverse dose-rate effects for mutagenesis by sparsely ionizing radiation,” Int. J. Radiat. Biol. 70(4), 447–
2146
458.
2147
2148
2149
2150
2151
2152
BRENNER, D.J., LITTLE, J.B. and SACHS, R.K. (2001). “The bystander effect in radiation oncogenesis: II.
A quantitative model,” Radiat. Res. 155(3), 402–408.
BREWEN, J.G., PRESTON, R.J., JONES, K.P. and GOSSLEE, D.G. (1973). “Genetic hazards of ionizing
radiations: cytogenetic extrapolations from mouse to man,” Mutat. Res. 17(2), 245–254.
BROOKS, A.L., LEI, X.C. and RITHIDECH, K. (2003). “Changes in biomarkers from space radiation may
reflect dose not risk,” Adv. Space Res. 31(6), 1505–1512.
2153
BUCKTON, K.E. (1983). “Chromosome aberrations in patients treated with x-irradiation for ankylosing
2154
spondylitis,” pages 491 to 511 in Radiation-Induced Chromosome Damage in Man, Ishihara, T. and
2155
Sasaki, M.S. Eds. (AR Liss, New York).
74
NCRP SC 1-21
Draft of March 31, 2015 [MR]
2156
2157
2158
NOT TO BE DISSEMINATED OR REFERENCED
BUDZOWSKA, M. and KANAAR, R. (2009). “Mechanisms of dealing with DNA damage-induced
replication problems,” Cell Biochem. Biophys. 53(1), 17–31.
BURAK, L.E., KODAMA, Y., NAKANO, M., OHTAKI, K., ITOH, M., OKLADNIKOVA, N.D.,
2159
VASILENKO, E.K., COLOGNE, J.B. and NAKAMURA, N. (2001). “FISH examination of lymphocytes
2160
from Mayak workers for assessment of translocation induction rate under chronic radiation exposures,”
2161
Int. J. Radiat. Biol. 77(8), 901–908.
2162
BUTTERWORTH, K.T., MCMAHON, S.J., HOUNSELL, A.R., O'SULLIVAN, J.M. and PRISE K.M.
2163
(2013) “Bystander signalling: Exploring clinical relevance through new approaches and new models,”
2164
Clin. Oncol. (R. Coll. Radiol.), 25(10), 586–592.
2165
CARDIS, E., VRIJHEID, M., BLETTNER, M., GILBERT, E., HAKAMA, M., HILL, C., HOWE, G.,
2166
KALDOR, J., MUIRHEAD, C.R., SCHUBAUER-BERIGAN, M., YOSHIMURA, T., BERMANN, F.,
2167
COWPER, G., FIX, J., HACKER, C., HEINMILLER, B., MARSHALL, M., THIERRY-CHEF, I.,
2168
UTTERBACK, D., AHN, Y.O., AMOROS, E., ASHMORE, P., AUVINEN, A., BAE, J.M., BERNAR,
2169
J., BIAU, A., COMBALOT, E., DEBOODT, P., DIEZ SACRISTAN, A., EKLOF, M., ENGELS, H.,
2170
ENGHOLM, G., GULIS, G., HABIB, R.R., HOLAN, K., HYVONEN, H., KEREKES, A.,
2171
KURTINAITIS, J., MALKER, H., MARTUZZI, M., MASTAUSKAS, A., MONNET, A., MOSER, M.,
2172
PEARCE, M.S., RICHARDSON, D.B., RODRIGUEZ-ARTALEJO, F., ROGEL, A., TARDY, H.,
2173
TELLE-LAMBERTON, M., TURAI, I., USEL, M. and VERESS, K. (2007). “The 15-Country
2174
Collaborative Study of Cancer Risk among Radiation Workers in the Nuclear Industry: Estimates of
2175
radiation-related cancer risks,” Radiat. Res. 167(4), 396–416.
2176
2177
2178
2179
CARLS, N. and SCHIESTL, R.H. (1999). “Effect of ionizing radiation on transgenerational appearance of
p(un) reversions in mice,” Carcinogenesis 20(12), 2351–2354.
CARNES, B.A., GRAHN, D. and HOEL, D. (2003). “Mortality of atomic bomb survivors predicted from
laboratory animals,” Radiat. Res. 160, 159–167.
2180
CARROLL, R.J., RUPPERT, D., STEFANSKI, L.A. and CRAINICEANU, C.M. (2006). Measurement Error
2181
in Nonlinear Models. A Modern Perspective, 2nd ed. (Chapman and Hall/CRC, Boca Raton, Florida).
2182
CHAI, Y., CALAF, G.M., ZHOU, H., GHANDHI, S.A., ELLISTON, C.D., WEN, G., NOHMI, T.,
2183
AMUNDSON, S.A. and HEI, T.K. (2013a). “Radiation induced COX-2 expression and mutagenesis at
2184
non-targeted lung tissues of gpt delta transgenic mice,” Br. J. Cancer, 108(1), 91–98.
2185
CHAI, Y., LAM, R.K., CALAF, G.M., ZHOU, H., AMUNDSON, S. and HEI, T.K. (2013b). “Radiation-
2186
induced non-targeted response in vivo: Role of the TGFβ-TGFBR1-COX-2 signalling pathway,” Br. J.
2187
Cancer, 108(5), 1106–1112.
75
NCRP SC 1-21
Draft of March 31, 2015 [MR]
2188
NOT TO BE DISSEMINATED OR REFERENCED
CHANOUX, R.A., YIN, B., URTISHAK, K.A., ASARE, A., BASSING, C.H. and BROWN, E.J. (2008).
2189
“ATR and H2AX cooperate in maintaining genome stability under replication stress,” J. Biol. Chem.
2190
284(9), 5994–6003.
2191
CONOLLY, R.B., KIMBELL, J.S., JANSZEN, D., SCHLOSSER, P.M., KALISAK, D., PRESTON, J. and
2192
MILLER, F.J. (2004). “Human respiratory tract cancer risks of inhaled formaldehyde: Dose-response
2193
predictions derived from biologically-motivated computational modeling of a combined rodent and
2194
human dataset,” Toxicol. Sci. 82(1), 279–296.
2195
2196
2197
CORNFORTH, M.N. (2001). “Analyzing radiation-induced complex chromosome rearrangements by
combinatorial painting,” Radiat Res. 155(5), 643–659.
CUCINOTTA, F.A., KIM, M.H.Y. and CHAPPELL, L.J. (2013a). Space Radiation Cancer Risk Projections
2198
and Uncertainties – 2012, NASA/TP-2013-217375, http://www.csc.caltech.edu/references/
2199
SpaceRadiationCancerRiskProjectionsandUncertainties.pdf (accessed November 8, 2013) (National
2200
Aeronautics and Space Administration, Washington).
2201
CURTIS, S.B., LUEBECK, E.G., HAZELTON, W.D. and MOOLGAVKAR, S.H. (2002). “A new
2202
perspective of carcinogenesis from protracted high-LET radiation arises from the two-stage clonal
2203
expansion model,” Adv. Space. Res. 30(4), 937–944.
2204
CURWEN, G.B., CADWELL, K.K., WINTHER, J.F., TAWN, E.J., REES, G.S., OLSEN, J.H.,
2205
RECHNITZER, C., SCHROEDER, H., GULDBERG, P., CORDELL, H.J. and BOICE, J.D., JR. (2010).
2206
“The heritability of G2 chromosomal radiosensitivity and its association with cancer in Danish cancer
2207
survivors and their offspring,” Int. J. Radiat. Biol. 86(11), 986–995.
2208
DAUER, L.T., BROOKS, A.L., HOEL, D.G., MORGAN, W.F., STRAM, D. and TRAN, P. (2010). “Review
2209
and evaluation of updated research on the health effects associated with low-dose ionising radiation,”
2210
Radiat. Prot. Dosim. 140(2), 103–136.
2211
DAY, J.P., LIMOLI, C.L. and MORGAN, W.F. (1998). “Recombination involving interstitial telomere
2212
repeat-like sequences promotes chromosomal instability in Chinese hamster cells,” Carcinogenesis 19(2),
2213
259–265.
2214
DICKEY, J.S., BAIRD, B.J., REDON, C.E., SOKOLOV, M.V., SEDELNIKOVA, O.A. and BONNER,
2215
W.M. (2009). “Intercellular communication of cellular stress monitored by γ-H2AX induction,”
2216
Carcinogenesis 30(10), 1686–1695.
2217
DING, L.H., SHINGYOJI, M., CHEN, F., HWANG, J.J., BURMA, S., LEE, C., CHENG, J.F. and CHEN,
2218
D.J. (2005). “Gene expression profiles of normal human fibroblasts after exposure to ionizing radiation:
2219
A comparative study of low and high doses,” Radiat. Res. 164(1), 17–26.
2220
2221
DOWNEY, M. and DUROCHER, D. (2006). “H2AX as a checkpoint maintenance signal,” Cell Cycle 5(13),
1376–1381.
76
NCRP SC 1-21
Draft of March 31, 2015 [MR]
2222
2223
2224
NOT TO BE DISSEMINATED OR REFERENCED
DRAPER, G.J., SANDERS, B.M. and KINGSTON, J.E. (1986). “Second primary neoplasms in patients with
retinoblastoma,” Br. J. Cancer 53(5), 661–671.
DUBROVA, Y.E., GRANT, G., CHUMAK, A.A., STEZHKA, V.A. and KARAKASIAN, A.N. (1996).
2225
“Elevated minisatellite mutation rate in the post-Chernobyl families from Ukraine,” Am. J. Hum. Genet.
2226
71(4), 801–809.
2227
DUBROVA, Y.E., NESTEROV, V.N., KROUCHINSKY, N.G., OSTAPENKO, V.A., VERGNAUD, G.,
2228
GIRAUDEAU, F., BUARD, J. and JEFFREYS, A.J. (1997). “Further evidence for elevated human
2229
minisatellite mutation rate in Belarus eight years after the Chernobyl accident,” Mutat. Res. 381(2), 267–
2230
278.
2231
2232
2233
DUBROVA, Y.E., PLOSHCHANSKAYA, O.G., KOZIONOVA, O.S. and AKLEYEV, A.V. (2006).
“Minisatellite germline mutation rate in the Techa River population,” Mutat. Res. 602(1–2), 74-82.
DURANT, S.T., PAFFETT, K.S., SHRIVASTAV, M., TIMMINS, G.S., MORGAN, W.F. and NICKOLOFF,
2234
J.A. (2006). “UV radiation induces delayed hyperrecombination associated with hypermutation in human
2235
cells,” Mol. Cell. Biol. 26(16), 6047–6055.
2236
EIDEMULLER, M., HOLMBERG, E., JACOB, P., LUNDELL, M. and KARLSSON, P. (2011). “Breast
2237
cancer risk after radiation treatment at infancy: Potential consequences of radiation-induced genomic
2238
instability,” Radiat. Prot. Dosim. 143(2–4), 375–379.
2239
ENG, C., LI, F.P., ABRAMSON, D.H., ELLSWORTH, R.M., WONG, F.L., GOLDMAN, M.B., SEDDON,
2240
J., TARBELL, N. and BOICE, J.D., JR. (1993). “Mortality from second tumors among long-term
2241
survivors of retinoblastoma,” J. Natl. Cancer Inst. 85(14), 1121–1128.
2242
EPA (2003). U.S. Environmental Protection Agency. Framework for Cumulative Risk Assessment,
2243
EPA/630/P-02/001F, http://www.epa.gov/raf/publications/pdfs/frmwrk_cum_risk_assmnt.pdf (accessed
2244
January 16, 2015) (U.S. Environmental Protection Agency, Washington).
2245
EPA (2005). U.S. Environmental Protection Agency. Guidelines for Carcinogen Risk Assessment,
2246
EPA/630/P-03/001F, http://www.epa.gov/raf/publications/pdfs/CANCER_GUIDELINES_FINAL_3-25-
2247
05.PDF (accessed January 14, 2015) (U.S. Environmental Protection Agency, Washington).
2248
2249
ERIKSSON, D. and STIGBRAND, T. (2010). “Radiation-induced cell death mechanisms,” Tumour Biol
31(4), 363–372.
2250
ESCHRICH, S., ZHANG, H., ZHAO, H., BOULWARE, D., LEE, J.H., BLOOM, G. and TORRES-ROCA,
2251
J.F. (2009). “Systems biology modeling of the radiation sensitivity network: A biomarker discovery
2252
platform,” Int. J. Radiat. Oncol. Biol. Phys. 75(2), 497–505.
2253
2254
EVANS, H.J., BUCKTON, K.E., HAMILTON, G.E. and CAROTHERS, A. (1979). “Radiation-induced
chromosome aberrations in nuclear-dockyard workers,” Nature 277(5697), 531–534.
77
NCRP SC 1-21
Draft of March 31, 2015 [MR]
2255
NOT TO BE DISSEMINATED OR REFERENCED
FERNANDEZ, M.I., GONG, Y., YE, Y., LIN, J., CHANG, D.W., KAMAT, A.M. and WU, X. (2013).
2256
“γ-H2AX level in peripheral blood lymphocytes as a risk predictor for bladder cancer,” Carcinogenesis
2257
34(11), 2543–2547.
2258
FERNET, M., MOULLAN, N., LAUGE, A., STOPPA-LYONNET, D. and HALL, J. (2004). “Cellular
2259
responses to ionising radiation of AT heterozygotes: Differences between missense and truncating
2260
mutation carriers,” Br. J. Cancer 90(4), 866–873.
2261
FERNET, M., MEGNIN-CHANET, F., HALL, J. and FAVAUDON, V. (2010). “Control of the G2/M
2262
checkpoints after exposure to low doses of ionising radiation: Implications for hyper-radiosensitivity,”
2263
DNA Repair (Amst.) 9(1), 48–57.
2264
FOKAS, E., PREVO, R., HAMMOND, E.M., BRUNNER, T.B., MCKENNA, W.G. and MUSCHEL, R.J.
2265
(2014). “Targeting ATR in DNA damage response and cancer therapeutics,” Cancer Treat. Rev. 40(1),
2266
109–117.
2267
2268
2269
2270
2271
2272
2273
2274
FRANKENBERG-SCHWAGER, M. (1990). “Induction, repair and biological relevance of radiation-induced
DNA lesions in eukaryotic cells,” Radiat. Environ. Biophys. 29(4), 273–292.
FRIEDBERG, E.C., WALKER, G.C., SIEDE, W., WOOD, R.D., SCHULTZ, R.A. and ELLENBERGER, T.
(2005). DNA Repair and Mutagenesis, 2nd ed. (ASM Press. Washington).
GALLIE, B.L., CAMPBELL, C., DEVLIN, H., DUCKETT, A. and SQUIRE, J.A. (1999). “Developmental
basis of retinal-specific induction of cancer by RB mutation,” Cancer Res. 59(7 Suppl.), 1731s–1735s.
GALLUZZI, L. and KROEMER, G. (2008). “Necroptosis: A specialized pathway of programmed necrosis,”
Cell 135(7), 1161–1163.
2275
GENIK, P.C., VYAZUNOVA, I., STEFFEN, L.S., BACHER, J.W., BIELEFELDT-OHMANN, H.,
2276
MCKERCHER, S., ULLRICH, R.L., FALLGREN, C.M., WEIL, M.M. and RAY, F.A. (2014).
2277
“Leukemogenesis in heterozygous PU.1 knockout mice,” Radiat. Res. 182(3), 310–315.
2278
GENOVESE, G., KAHLER, A.K., HANDSAKER, R.E., LINDBERG, J., ROSE, S.A., BAKHOUM, S.F.,
2279
CHAMBERT, K., MICK, E., NEALE, B.M., FROMER, M., PURCELL, S.M., SVANTESSON, O.,
2280
LANDEN, M., HOGLUND, M., LEHMANN, S., GABRIEL, S.B., MORAN, J.L., LANDER, E.S.,
2281
SULLIVAN, P.F., SKLAR, P., GRONBERG, H., HULTMAN, C.M. and MCCARROLL, S.A. (2014).
2282
“Clonal hematopoiesis and blood-cancer risk inferred from blood DNA sequence,” N. Engl. J. Med.
2283
371(26), 2477–2487.
2284
2285
2286
GOLDBERG, Z. (2003). “Clinical implications of radiation-induced genomic instability,” Oncogene 22(45),
7011–7017.
GRAHN, D. (1960). “Genetic control of physiological processes: The genetics of radiation toxicity in
2287
animals,” pages 181 to 190 in A Symposium on Radioisotopes in the Biosphere, Caldecott, R.S. and
2288
Snyder, L.A., Eds. (University of Minnesota, Minneapolis).
78
NCRP SC 1-21
Draft of March 31, 2015 [MR]
2289
NOT TO BE DISSEMINATED OR REFERENCED
GRAHN, D, and SACHER, GA. (1968). “Fractionation and protraction factors and the late effects of
2290
radiation in small mammals,” pages 2-1 to 2-27 in The Proceedings of a Symposium on Dose Rate in
2291
Mammalian Radiation Biology, Brown, D.G., Cragle, R.G. and Noonan, T.R., Eds., U.S. Atomic Energy
2292
Commission CONF-680410 (National Technical Information Service, Springfield, Virginia).
2293
GRANT, E.J., SUGIYAMA, H., OZASA, A.V.K. and RERF-NCI Cancer Risk Analysis Group (2014).
2294
“Abstract: Solid cancer incidence among atomic bomb survivors: Overview of dosimetry, follow-up and
2295
risk updates,” Conference on Radiation and Health, http://www.ameetingbydesign.com/ambddrop/
2296
files/basic-html/page39.html (accessed January 16, 2015) (Radiation Effects Research Foundation,
2297
Hiroshima).
2298
GREEN, D.M., SKLAR, C.A., BOICE, J.D., JR., MULVIHILL, J.J., WHITTON, J.A., STOVALL, M. and
2299
YASUI, Y. (2009). “Ovarian failure and reproductive outcomes after childhood cancer treatment: Results
2300
from the Childhood Cancer Survivor Study,” J. Clin. Oncol. 27(14), 2374–2381.
2301
GREENMAN, C., STEPHENS, P., SMITH, R., DALGLIESH, G.L., HUNTER, C., BIGNELL, G., DAVIES,
2302
H., TEAGUE, J., BUTLER, A., STEVENS, C., EDKINS, S., O’MEARA, S., VASTRIK, I., SCHMIDT,
2303
E.E., AVIS, T., BARTHORPE, S., BHAMRA, G., BUCK, G., CHOUDHURY, B., CLEMENTS, J.,
2304
COLE, J., DICKS, E., FORBES, S., GRAY, K., HALLIDAY, K., HARRISON, R., HILLS, K.,
2305
HINTON, J., JENKINSON, A., JONES, D., MENZIES, A., MIRONENKO, T., PERRY, J., RAINE, K.,
2306
RICHARDSON, D., SHEPHERD, R., SMALL, A., TOFTS, C., VARIAN, J., WEBB, T., WEST, S.,
2307
WIDAA, S., YATES, A., CAHILL, D.P., LOUIS, D.N., GOLDSTRAW, P., NICHOLSON, A.G.,
2308
BRASSEUR, F., LOOIJENGA, L., WEBER, B.L., CHIEW, Y.E., DEFAZIO, A., GREAVES, M.F.,
2309
GREEN, A.R., CAMPBELL, P., BIRNEY, E., EASTON, D.F., CHENEVIX-TRENCH, G., TAN, M.H.,
2310
KHOO, S.K., TEH, B.T., YUEN, S.T., LEUNG, S.Y., WOOSTER, R., FUTREAL, P.A. and
2311
STRATTON, M.R. (2007). “Patterns of somatic mutation in human cancer genomes,” Nature 446(7132),
2312
153–158.
2313
GRIDLEY, D.S., RIZVI, A., LUO-OWEN, X., MAKINDE, A.Y. and PECAUT, M.J. (2009). “Low dose,
2314
low dose rate photon radiation modifies leukocyte distribution and gene expression in CD4(+) T cells,” J.
2315
Radiat. Res. (Tokyo) 50(2), 139–150.
2316
2317
2318
GROSOVSKY, A.J. and LITTLE, J.B. (1985). “Evidence for linear response for the induction of mutations in
human cells by x-ray exposures below 10 rads,” Proc. Natl. Acad. Sci. (USA) 82(7), 2092–2095.
GROSOVSKY, A.J., DE BOER, J.G., DE JONG, P.J., DROBETSKY, E.A. and GLICKMAN, B.W. (1988).
2319
“Base substitutions, frameshifts, and small deletions constitute ionizing radiation-induced point mutations
2320
in mammalian cells,” Proc. Natl. Acad. Sci USA 85(1), 185–188.
2321
2322
GUO, Y., CAI, Q., SAMUELS, D.C., YE, F., LONG, J., LI, C.I., WINTHER, J.F., TAWN, E.J., STOVALL,
M., LAHTEENMAKI, P., MALILA, N., LEVY, S., SHAFFER, C., SHYR, Y., SHU, X.O. and BOICE,
79
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
2323
J.D., JR. (2012). “The use of next generation sequencing technology to study the effect of radiation
2324
therapy on mitochondrial DNA mutation,” Mutat. Res. 744(2), 154–160.
2325
HAGMAR, L., BROGGER, A., HANSTEEN, I.L., HEIM, S., HOGSTEDT, B., KNUDSEN, L., LAMBERT,
2326
B., LINNAINMAA, K., MITELMAN, F., NORDENSON, I., REUTERWALL, C., SALOMAN, S.,
2327
SKERFVING, S. and SORSA, M. (1994). “Cancer risk in humans predicted by increased levels of
2328
chromosomal aberrations in lymphocytes: Nordic study group on the health risk of chromosome damage,”
2329
Cancer Res. 54(11), 2919–2922.
2330
HAGMAR, L., STROMBERG, U., TINNERBERG, H. and MIKOCZY, Z. (2004). “Epidemiological
2331
evaluation of cytogenetic biomarkers as potential surrogate end-points for cancer,” IARC Sci. Publ. (157),
2332
207–215.
2333
2334
2335
HALL, E.J. and GIACCIA, A.J. (2011). Radiobiology for the Radiologist, 7th ed. (Lippincott Williams and
Wilkins, Philadelphia).
HAMMER, G.P., SCHEIDEMANN-WESP, U., SAMKANGE-ZEEB, F., WICKE, H., NERIISHI, K. and
2336
BLETTNER, M. (2013). “Occupational exposure to low doses of ionizing radiation and cataract
2337
development: A systematic literature review and perspectives on future studies,” Radiat. Environ.
2338
Biophys. 52(3), 303–319.
2339
HANAHAN, D. and WEINBERG, R.A. (2000). “The hallmarks of cancer,” Cell 100(1), 57–70.
2340
HANAHAN, D. and WEINBERG, R.A. (2011). “Hallmarks of cancer: The next generation,” Cell 144(5),
2341
2342
2343
2344
646–674.
HARVEY, E.B., BOICE, J.D., JR., HONEYMAN, M. and FLANNERY, J.T. (1985). “Prenatal x-ray
exposure and childhood cancer in twins,” N. Engl. J. Med. 312(9), 541–545.
HAYASHI, T., KUSUNOKI, Y., HAKODA, M., MORISHITA, Y., KUBO, Y., MAKI, M., KASAGI, F.,
2345
KODAMA, K., MACPHEE, D.G. and KYOIZUMI, S. (2003). “Radiation dose-dependent increases in
2346
inflammatory response markers in A-bomb survivors,” Int. J. Radiat. Biol. 79(2), 129–136.
2347
HAYASHI, T., MORISHITA, Y., KUBO, Y., KUSUNOKI, Y., HAYASHI, I., KASAGI, F., HAKODA, M.,
2348
KYOIZUMI, S. and NAKACHI, K. (2005). “Long-term effects of radiation dose on inflammatory
2349
markers in atomic bomb survivors,” Am. J. Med. 118(1), 83–86.
2350
HAYATA, I., WANG, C., ZHANG, W., CHEN, D., MINAMIHISAMATSU, M., MORISHIMA, H., WEI, L.
2351
and SUGAHARA, T. (2004). “Effect of high-level natural radiation on chromosomes of residents in
2352
southern China,” Cytogenet. Genome Res. 104(1–4), 237–239.
2353
HE, Y., GONG, Y., LIN, J., CHANG, D.W., GU, J., ROTH, J.A. and WU, X. (2013). “Ionizing radiation-
2354
induced γ-H2AX activity in whole blood culture and the risk of lung cancer,” Cancer Epidemiol.
2355
Biomarkers Prev. 22(3), 443–451.
80
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
2356
HEI, T.K., ZHOU, H., IVANOV, V.N., HONG, M., LIEBERMAN, H.B., BRENNER, D.J., AMUNDSON,
2357
S.A. and GEARD, C.R. (2008). “Mechanism of radiation-induced bystander effects: A unifying model,”
2358
J. Pharm. Pharmacol. 60(8), 943–950.
2359
HEIDENREICH, W.F. and ROSEMANN, M. (2012). “Genetic background and 227Thorium as risk factors in
2360
biologically based models for induction of bone cancer in mice,” Radiat. Environ. Biophys. 51(2),179–
2361
185.
2362
HEIDENREICH, W.F., TOMASEK, L., ROGEL, A., LAURIER, D. and TIRMARCHE, M. (2004). “Studies
2363
of radon-exposed miner cohorts using a biologically based model: Comparison of current Czech and
2364
French data with historic data from China and Colorado,” Radiat. Environ. Biophys. 43(4), 247–256.
2365
2366
HEIDENREICH, W.F., SARAN, A., ATKINSON, M. and PAZZAGLIA, S. (2013). “A mechanistic model
for medulloblastoma induction in mice,” Radiat. Res. 179(5), 610–614.
2367
HOWE, G.R. and MCLAUGHLIN, J. (1996). “Breast cancer mortality between 1950 and 1987 after exposure
2368
to fractionated moderate-dose-rate ionizing radiation in the Canadian Fluoroscopy Cohort Study and a
2369
comparison with breast cancer mortality in the Atomic Bomb Survivors Study,” Radiat. Res. 145(6), 694–
2370
707.
2371
2372
2373
2374
HOWELL, E.K., GASCHAK, S.P., GRIFFITH, K.D. and RODGERS, B.E. (2013). “Radioadaptive response
following in utero low-dose irradiation,” Radiat. Res. 179(1), 29–37.
HSIEH, P. and YAMANE, K. (2008). “DNA mismatch repair: Molecular mechanism, cancer, and ageing,”
Mech. Ageing Dev. 129(7–8), 391–407.
2375
HSIEH, W.A., NI, C., HWANG, J.J., FANG, J.S., LIN, S.P., LIN, Y.A., HUANG, T.W. and CHANG, W.P.
2376
(2002). “Evaluation of the frequencies of chromosomal aberrations in a population exposed to prolonged
2377
low dose-rate 60Co gamma-irradiation,” Int. J. Radiat. Biol. 78(7), 625–633.
2378
HSU, W.L., PRESTON, D.L., SODA M, SUGIYAMA, H., FUNAMOTO, S., KODAMA, K., KIMURA, A.,
2379
KAMADA, N., DOHY, H., TOMONAGA, M., IWANAGA, M., MIYAZAKI, Y., CULLINGS, H.M.,
2380
SUYAMA, A., OZASA, K., SHORE, R.E. and MABUCHI, K. (2013). “The incidence of leukemia,
2381
lymphoma and multiple myeloma among atomic bomb survivors: 1950–2001,” Radiat. Res. 179(3), 361–
2382
382.
2383
HUANG, L., GRIMM, S., SMITH, L.E., KIM, P.M., NICKOLOFF, J.A., GOLOUBEVA, O.G. and
2384
MORGAN, W.F. (2004). “Ionizing radiation induces delayed hyperrecombination in mammalian cells,”
2385
Mol. Cell. Biol. 24(11), 5060–5068.
2386
HUANG, L., KIM, P.M., NICKOLOFF, J.A. and MORGAN, W.F. (2007). “Targeted and non-targeted
2387
effects of low-dose ionizing radiation on delayed genomic instability in human cells,” Cancer Res. 67(3),
2388
1099–1104.
81
NCRP SC 1-21
Draft of March 31, 2015 [MR]
2389
NOT TO BE DISSEMINATED OR REFERENCED
HWANG, S.L., HWANG, J.S., YANG, Y.T., HSIEH, W.A., CHANG, T.C., GUO, H.R., TSAI, M.H.,
2390
TANG, J.L., LIN, I.F. and CHANG, W.P. (2008). “Estimates of relative risks for cancers in a population
2391
after prolonged low-dose-rate radiation exposure: A follow-up assessment from 1983 to 2005,” Radiat.
2392
Res. 170(2), 143–148.
2393
ICRP (1991). International Commission on Radiological Protection. 1990 Recommendations of the
2394
International Commission on Radiological Protection. ICRP Publication 60, Ann. ICRP 21(1–3) (Sage
2395
Publications, Thousand Oaks, California).
2396
2397
2398
ICRP (1998). International Commission on Radiological Protection. Genetic Susceptibility to Cancer, ICRP
Publication 79, Ann. ICRP 28(1–2) (Sage Publications, Thousand Oaks, California).
ICRP (2007). International Commission on Radiological Protection. The 2007 Recommendations of the
2399
International Commission on Radiological Protection, ICRP Publication 103, Ann. ICRP 37(2–4) (Sage
2400
Publications, Thousand Oaks, California).
2401
ICRP (2012). International Commission on Radiological Protection. ICRP Statement on Tissue Reactions and
2402
Early and Late Effects of Radiation in Normal Tissues and Organs—Threshold Doses for Tissue
2403
Reactions in a Radiation Protection Context, ICRP Publication 118, Ann. ICRP 41(1/2) (Sage
2404
Publications, Thousand Oaks, California).
2405
2406
2407
ILNYTSKYY, Y. and KOVALCHUK, O. (2011). “Non-targeted radiation effects-an epigenetic connection,”
Mutat. Res. 714(1–2), 113–125.
IVANOV, V.K., TSYB, A.F., KHAIT, S.E., KASHCHEEV, V.V., CHEKIN, S.Y., MAKSIOUTOV, M.A.
2408
and TUMANOV, K.A. (2012). “Leukemia incidence in the Russian cohort of Chernobyl emergency
2409
workers,” Radiat. Environ. Biophys. 51(2), 143–149.
2410
JACOB, P., MECKBACH, R., KAISER, J.C. and SOKOLNIKOV, M. (2010). “Possible expressions of
2411
radiation-induced genomic instability, bystander effects or low-dose hypersensitivity in cancer
2412
epidemiology,” Mutat. Res. 687(1–2), 34–39.
2413
2414
2415
JABLON, S. and KATO, H. (1970). “Childhood cancer in relation to prenatal exposure to atomic-bomb
radiation,” Lancet 2(7681), 1000–1003.
JARABEK, A.M., POTTENGER, L.H., ANDREWS, L.S., CASCIANO, D., EMBRY, M.R., KIM, J.H.,
2416
PRESTON, R.J., REDDY, M.V., SCHOENY, R., SHUKER, D., SKARE, J., SWENBERG, J.,
2417
WILLIAMS, G.M. and ZEIGER, E. (2009). “Creating context for the use of DNA adduct data in cancer
2418
risk assessment: I. Data organization,” Crit. Rev. Toxicol. 39(8), 659–678.
2419
JEN, M.H., HWANG, J.J., YANG, J.Y., NABYVANETS, Y.B., HSIEH, W.A., TSAI, M.H., GUO, S.D. and
2420
CHANG, W.P. (2002). “Micronuclei and nuclear anomalies in urinary exfoliated cells of subjects in
2421
radionuclide-contaminated regions,” Mutat. Res. 520(1–2), 39–46.
82
NCRP SC 1-21
Draft of March 31, 2015 [MR]
2422
NOT TO BE DISSEMINATED OR REFERENCED
JONES, I.M., GALICK, H., KATO, P., LANGLOIS, R.G., MENDELSOHN, M.L., MURPHY, G.A.,
2423
PLESHANOV, P., RAMSEY, M.J., THOMAS, C.B., TUCKER, J.D., TUREVA, L., VOROBTSOVA, I.
2424
and NELSON, D.O. (2002). “Three somatic genetic biomarkers and covariates in radiation-exposed
2425
Russian cleanup workers of the Chernobyl nuclear reactor 6-13 years after exposure,” Radiat. Res.
2426
158(4), 424–442.
2427
JULIEN, E., BOOBIS, A.R. and OLIN, S.S. (2009). “The key events dose-response framework: A cross-
2428
disciplinary mode-of-action based approach to examining dose-response and thresholds,” Crit. Rev. Food
2429
Sci. Nutr. 49(8), 682–689.
2430
2431
2432
2433
2434
KAISER, J.C., MECKBACH, R. and JACOB, P. (2014). “Genomic instability and radiation risk in molecular
pathways to colon cancer,” PLoS One 9(10):e111024.
KAMILERI, I., KARAKASILIOTI, I. and GARINIS, G.A. (2012). “Nucleotide excision repair: New tricks
with old bricks,” Trends Genet. 28(11), 566–573.
KATO, T.A., WILSON, P.F., NAGASAWA, H., FITZEK, M.M., WEIL, M.M., LITTLE, J.B. and
2435
BEDFORD, J.S. (2007). “A defect in DNA double strand break processing in cells from unaffected
2436
parents of retinoblastoma patients and other apparently normal humans,” DNA Repair (Amst). 6(6), 818–
2437
829.
2438
KATO, T.A., WILSON, P.F., NAGASAW, H., PENG, Y., WEIL, M.M., LITTLE, J.B. and BEDFORD, J.S.
2439
(2009). “Variations in radiosensitivity among individuals: A potential impact on risk assessment?” Health
2440
Phys. 97(5), 470–480.
2441
KENDALL, G.M, LITTLE, M.P., WAKEFORD, R., BUNCH, K.J., MILES, J.C., VINCENT, T.J., MEARA,
2442
J.R. and MURPHY, M.F. (2013). “A record-based case-control study of natural background radiation and
2443
the incidence of childhood leukaemia and other cancers in Great Britain during 1980–2006,” Leukemia
2444
27(1), 3–9.
2445
KIM, G.J., CHANDRASEKARAN, K. and MORGAN, W.F. (2006a). “Mitochondrial dysfunction,
2446
persistently elevated levels of reactive oxygen species and radiation-induced genomic instability: A
2447
review,” Mutagenesis 21(6), 361–367.
2448
2449
2450
2451
2452
KIM, G.J., FISKUM, G.M. and MORGAN, W.F. (2006b). “A role for mitochondrial dysfunction in
perpetuating radiation-induced genomic instability,” Cancer Res. 66(21), 10377–10383.
KLEINERMAN, R.A. (2009). “Radiation-sensitive genetically susceptible pediatric sub-populations,”
Pediatr. Radiol. 39(Suppl. 1), S27–S31.
KLEINERMAN, R.A., LITTLEFIELD, L.G., TARONE, R.E., MACHADO, S.G., BLETTNER, M.,
2453
PETERS, L.J. and BOICE, J.D., JR. (1989). “Chromosome aberrations in peripheral lymphocytes and
2454
radiation dose to active bone marrow in patients treated for cancer of the cervix,” Radiat. Res. 119(1),
2455
176–190.
83
NCRP SC 1-21
Draft of March 31, 2015 [MR]
2456
NOT TO BE DISSEMINATED OR REFERENCED
KLEINERMAN, R.A., LITTLEFIELD, L.G., TARONE, R.E., SAYER, A.M., HILDRETH, N.G.,
2457
POTTERN, L.M., MACHADO, S.G. and BOICE, J.D., JR. (1990). “Chromosome aberrations in relation
2458
to radiation dose following partial-body exposures in three populations,” Radiat. Res. 123(1), 93–101.
2459
KLEINERMAN, R.A., LITTLEFIELD, L.G., TARONE, R.E., SAYER, A.M., COOKFAIR, D.L.,
2460
WACTAWSKI-WENDE, J., INSKIP, P.D., BLOCK, A., RAMESH, K.H. and BOICE, J.D., JR. (1994).
2461
“Chromosome aberrations in lymphocytes from women irradiated for benign and malignant
2462
gynecological disease,” Radiat. Res. 139(1), 40–46.
2463
KLEINERMAN, R.A., TUCKER, M.A., TARONE, R.E., ABRAMSON, D.H., SEDDON, J.M., STOVALL,
2464
M., LI, F.P. and FRAUMENI, J.F., JR. (2005). “Risk of new cancers after radiation therapy in long-term
2465
survivors of retinoblastoma: An extended follow-up,” J. Clin. Oncol. 23(10), 2272–2279.
2466
KLEINERMAN, R.A, ROMANYUKHA, A.A., SCHAUER, D.A. and TUCKER, J.D. (2006). “Retrospective
2467
assessment of radiation exposure using biological dosimetry: Chromosome painting, electron
2468
paramagnetic resonance and the glycophorin a mutation assay,” Radiat. Res. 166(1 Pt 2), 287–302.
2469
KLEINERMAN, R.A., TUCKER, M.A., ABRAMSON, D.H., SEDDON, J.M., TARONE, R.E. and
2470
FRAUMENI, J.F., JR. (2007). “Risk of soft tissue sarcomas by individual subtype in survivors of
2471
hereditary retinoblastoma,” J. Natl. Cancer Inst. 99(1), 24–31.
2472
2473
2474
KNUDSON, A.G., JR. (1971). “Mutation and cancer: Statistical study of retinoblastoma,” Proc. Natl. Acad.
Sci. USA 68(4), 820–823.
KODAMA, Y., PAWEL, D., NAKAMURA, N., PRESTON, D., HONDA, T., ITOH, M., NAKANO, M.,
2475
OHTAKI, K., FUNAMOTO, S. and AWA, A.A. (2001). “Stable chromosome aberrations in atomic
2476
bomb survivors: Results from 25 years of investigation,” Radiat. Res. 156(4), 337–346.
2477
KOTURBASH, I., BAKER, M., LOREE, J., KUTANZI, K., HUDSON, D., POGRIBNY, I.,
2478
SEDELNIKOVA, O., BONNER, W. and KOVALCHUK, O. (2006). “Epigenetic dysregulation underlies
2479
radiation-induced transgenerational genome instability in vivo,” Int. J. Radiat. Oncol. Biol. Phys. 66(2),
2480
327–330.
2481
KRAEMER, S.M., VANNAIS, D.B., KRONENBERG, A., UENO, A. and WALDREN, C.A. (2001).
2482
“Gamma-ray mutagenesis studies in a new human-hamster hybrid, A(L)CD59(+/-), which has two human
2483
chromosomes 11 but is hemizygous for the CD59 gene,” Radiat. Res. 156(1), 10–19.
2484
KRESTININA, L.Y., DAVIS, F.G., SCHONFELD, S., PRESTON, D.L., DEGTEVA, M., EPIFANOVA, S.
2485
and AKLEYEV, A.V. (2013). “Leukaemia incidence in the Techa River Cohort: 1953–2007,” Br. J.
2486
Cancer 109(11), 2886–2893.
2487
KREWSKI, D., ZIELINSKI, J.M., HAZELTON, W.D., GARNER, M.J. and MOOLGAVKAR, S.H. (2003).
2488
“The use of biologically based cancer risk models in radiation epidemiology,” Radiat. Prot. Dosim.
2489
104(4), 367–376.
84
NCRP SC 1-21
Draft of March 31, 2015 [MR]
2490
2491
2492
NOT TO BE DISSEMINATED OR REFERENCED
KROKAN, H.E. and BJORAS, M. (2013). “Base excision repair,” Cold Spring Harb. Perspect. Biol. 5(4),
a012583.
KUSUNOKI, Y., KYOIZUMI, S., HIRAI, Y., SUZUKI, T., NAKASHIMA, E., KODAMA, K. and
2493
SEYAMA, T. (1998). “Flow cytometry measurements of subsets of T, B and NK cells in peripheral blood
2494
lymphocytes of atomic bomb survivors,” Radiat. Res. 150(2), 227–236.
2495
KUSUNOKI, Y., KYOIZUMI, S., YAMAOKA, M., KASAGI, F., KODAMA, K. and SEYAMA, T. (1999).
2496
“Decreased proportion of CD4 T cells in the blood of atomic bomb survivors with myocardial infarction,”
2497
Radiat. Res. 152(5), 539–543.
2498
LANZ, C., PATTERSON, A.D., SLAVIK, J., KRAUSZ, K.W., LEDERMANN, M., GONZALEZ, F.J. and
2499
IDLE, J.R. (2009). “Radiation metabolomics. 3. Biomarker discovery in the urine of gamma-irradiated
2500
rats using a simplified metabolomics protocol of gas chromatography-mass spectrometry combined with
2501
random forests machine learning algorithm,” Radiat. Res. 172(2), 198–212.
2502
LAIAKIS, E.C., BAULCH, J.E. and MORGAN, W.F. (2007). “Cytokine and chemokine responses after
2503
exposure to ionizing radiation: Implications for the astronauts,” Adv. Space Res. 39(6), 1019–1025.
2504
LEA, D.E. (1946). Actions of Radiations on Living Cells (Cambridge University Press, Cambridge).
2505
LEE DO, Y., BOWEN, B.P., NGUYEN, D.H., PARSA, S., HUANG, Y., MAO, J.H. and NORTHEN, T.R.
2506
(2012). “Low-dose ionizing radiation-induced blood plasma metabolic response in a diverse genetic
2507
mouse population,” Radiat. Res. 178(6), 551–555.
2508
LIBER, H., YANDELL, D.W. and LITTLE, J.B. (1989). “A comparison of mutation induction at the tk and
2509
hprt loci in human lymphoblastoid cells; quantitative differences are due to an additional class of
2510
mutations at the autosomal tk locus,” Mutat. Res. 216(1), 9–17.
2511
LIMOLI, C.L., KAPLAN, M.I., CORCORAN, J., MEYERS, M., BOOTHMAN, D.A. and MORGAN, W.F.
2512
(1997). “Chromosomal instability and its relationship to other end points of genomic instability,” Cancer
2513
Res. 57(24), 5557–5563.
2514
LIMOLI, C.L., CORCORAN, J.J., MILLIGAN, J.R., WARD, J.F. and MORGAN, W.F. (1999). “Critical
2515
target and dose and dose-rate responses for the induction of chromosomal instability by ionizing
2516
radiation,” Radiat. Res. 151(6), 677–685.
2517
LIN, R.X., ZHAO, H.B., LI, C.R., SUN, Y.N., QIAN, X.H. and WANG, S.Q. (2009). “Proteomic analysis of
2518
ionizing radiation-induced proteins at the subcellular level,” J. Proteome Res. 8(1), 390–399.
2519
LIN, C.M., CHANG, W.P., DOYLE, P., WANG, J.D., LEE, L.T., LEE, C.L. and CHEN, P.C. (2010).
2520
“Prolonged time to pregnancy in residents exposed to ionising radiation in cobalt-60-contaminated
2521
buildings,” Occup. Environ. Med. 67(3), 187–195.
85
NCRP SC 1-21
Draft of March 31, 2015 [MR]
2522
NOT TO BE DISSEMINATED OR REFERENCED
LITTLE, M.P. (1995). “Are two mutations sufficient to cause cancer? Some generalizations of the two-
2523
mutation model of carcinogenesis of Moolgavkar, Venzon, and Knudson, and of the multistage model of
2524
Armitage and Doll,” Biometrics 51(4), 1278–1291.
2525
2526
LITTLE, M.P. (1996). “Generalisations of the two-mutation and classical multi-stage models of
carcinogenesis fitted to the Japanese atomic bomb survivor data,” J. Radiol. Prot. 16(1), 7–24.
2527
LITTLE, J.B. (2000). “Radiation carcinogenesis,” Carcinogenesis 21(3), 397–404.
2528
LITTLE, M.P. (2004). “The bystander effect model of Brenner and Sachs fitted to lung cancer data in 11
2529
cohorts of underground miners, and equivalence of fit of a linear relative risk model with adjustment for
2530
attained age and age at exposure,” J. Radiol. Prot. 24(3), 243–255.
2531
2532
2533
2534
2535
2536
2537
LITTLE, M.P. (2010). “Cancer models, genomic instability and somatic cellular Darwinian evolution,” Biol.
Direct 5, 19.
LITTLE, M.P. (2013). “A review of non-cancer effects, especially circulatory and ocular diseases,” Radiat.
Environ. Biophys. 52(4), 435–449.
LITTLE, M.P. (2014). “Germline minisatellite mutations in the offspring of irradiated parents,” J. Radiat.
Prot. 35(1), E1–E4.
LITTLE, M.P. and BOICE, J.D., JR. (1999). “Comparison of breast cancer incidence in the Massachusetts
2538
Tuberculosis Fluoroscopy Cohort and in the Japanese atomic bomb survivors,” Radiat. Res. 151(2), 218–
2539
224.
2540
2541
2542
2543
2544
2545
LITTLE, M.P. and LI, G. (2007). “Stochastic modelling of colon cancer: Is there a role for genomic
instability?” Carcinogenesis 28(2), 479–487.
LITTLE, M.P. and WAKEFORD, R. (2001). “The bystander effect in C3H 10T1/2 cells and radon-induced
lung cancer,” Radiat. Res. 156(6), 695–699.
LITTLE, M.P. and WRIGHT, E.G. (2003). “A stochastic carcinogenesis model incorporating genomic
instability fitted to colon cancer data,” Math. Biosci. 183(2), 111–134.
2546
LITTLE, M.P., MUIRHEAD, C.R. and CHARLES, M.W. (1999a). “Describing time and age variations in the
2547
risk of radiation-induced solid tumour incidence in the Japanese atomic bomb survivors using generalized
2548
relative and absolute risk models,” Stat. Med. 18(1), 17–33.
2549
LITTLE, M.P., WEISS, H.A., BOICE, J.D. JR., DARBY, S.C., DAY, N.E. and MUIRHEAD, C.R. (1999b).
2550
“Risks of leukemia in Japanese atomic bomb survivors, in women treated for cervical cancer, and in
2551
patients treated for ankylosing spondylitis,” Radiat. Res. 152(3), 280–292.
2552
LITTLE, M.P., HAYLOCK, R.G.E. and MUIRHEAD, C.R. (2002). “Modelling lung tumour risk in radon-
2553
exposed uranium miners using generalizations of the two-mutation model of Moolgavkar, Venzon and
2554
Knudson,” Int. J. Radiat. Biol. 78(1), 49–68.
86
NCRP SC 1-21
Draft of March 31, 2015 [MR]
2555
NOT TO BE DISSEMINATED OR REFERENCED
LITTLE, M.P., FILIPE, J.A., PRISE, K.M., FOLKARD, M. and BELYAKOV, O.V. (2005). “A model for
2556
radiation-induced bystander effects, with allowance for spatial position and the effects of cell turnover,”
2557
J. Theor. Biol. 232(3), 329–338.
2558
LITTLE, M.P., HOEL, D.G., MOLITOR, J., BOICE, J.D., WAKEFORD, R. and MUIRHEAD, C.R. (2008a).
2559
“New models for evaluation of radiation-induced lifetime cancer risk and its uncertainty employed in the
2560
UNSCEAR 2006 report,” Radiat. Res. 169(6), 660–676.
2561
LITTLE, M.P., VINEIS, P. and LI, G. (2008b). “A stochastic carcinogenesis model incorporating multiple
2562
types of genomic instability fitted to colon cancer data,” J. Theoret. Biol. 254(2), 229-238 [Erratum
2563
255(2), 268].
2564
LITTLE, M.P., WAKEFORD, R., TAWN, E.J., BOUFFLER, S.D. and BERRINGTON DE GONZALEZ, A.
2565
(2009). “Risks associated with low doses and low dose rates of ionizing radiation: Why linearity may be
2566
(almost) the best we can do,” Radiology 251(1), 6–12.
2567
LITTLE, M.P., KLEINERMAN, R.A., STILLER, C.A., LI, G., KROLL, M.E. and MURPHY, M.F. (2012a).
2568
“Analysis of retinoblastoma age incidence data using a fully stochastic cancer model,” Int. J. Cancer
2569
130(3), 631–640.
2570
LITTLE, M.P., AZIZOVA, T.V., BAZYKA, D., BOUFFLER, S.D., CARDIS, E., CHEKIN, S., CHUMAK,
2571
V.V., CUCINOTTA, F.A., DE VATHAIRE, F., HALL, P., HARRISON, J.D., HILDEBRANDT, G.,
2572
IVANOV, V., KASHCHEEV, V.V., KLYMENKO, S.V., KREUZER, M., LAURENT, O., OZASA, K.,
2573
SCHNEIDER, T., TAPIO, S., TAYLOR, A.M., TZOULAKI, I., VANDOOLAEGHE, W.L.,
2574
WAKEFORD, R., ZABLOTSKA, L.B., ZHANG, W. and LIPSHULTZ, S.E. (2012b). “Systematic
2575
review and meta-analysis of circulatory disease from exposure to low-level ionizing radiation and
2576
estimates of potential population mortality risks,” Environ. Health Perspect. 120(11), 1503–1511.
2577
LITTLE, M.P., GOODHEAD, D.T., BRIDGES, B.A. and BOUFFLER, S.D. (2013). “Evidence relevant to
2578
untargeted and transgenerational effects in the offspring of irradiated parents,” Mutat. Res. Rev. 753(1),
2579
50–67.
2580
LITTLE, M.P., KWON, D., DOI, K., SIMON, S.L., PRESTON, D.L., DOODY, M.M., LEE, T., MILLER,
2581
J.S., KAMPA, D.M., BHATTI, P., TUCKER, J.D., LINET, M.S. and SIGURDSON, A.J. (2014a).
2582
“Association of chromosome translocation rate with low dose occupational radiation exposures in U.S.
2583
radiologic technologists,” Radiat. Res. 182(1), 1–17.
2584
LITTLE, M.P., SCHAEFFER, M.L., REULEN, R.C., ABRAMSON, D.H., STOVALL, M., WEATHERS,
2585
R., DE VATHAIRE, F., DIALLO, I., SEDDON, J.M., HAWKINS, M.M., TUCKER, M.A. and
2586
KLEINERMAN, R.A. (2014b). “Breast cancer risk after radiation therapy for heritable and non-heritable
2587
retinoblastoma: A US–UK study,” Br. J. Cancer 110(10), 2623–2632.
87
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
2588
LITTLEFIELD, L.G., KLEINERMAN, R.A., SAYER, A.M., TARONE, R. and BOICE, J.D., JR. (1991).
2589
“Chromosome aberrations in lymphocytes—biomonitors of radiation exposure,” Prog. Clin. Biol. Res.
2590
372, 387–397.
2591
LITTLEFIELD, L.G., TRAVIS, L.B., SAYER, A.M., VOELZ, G.L., JENSEN, R.H. and BOICE, J.D., JR.
2592
(1997). “Cumulative genetic damage in hematopoietic stem cells in a patient with a 40-year exposure to
2593
alpha particles emitted by thorium dioxide,” Radiat. Res. 148(2), 135–144.
2594
LITTLEFIELD, L.G., MCFEE, A.F., SALOMAA, S.I., TUCKER, J.D., INSKIP, P.D., SAYER, A.M.,
2595
LINDHOLM, C., MAKINEN, S., MUSTONEN, R., SORENSEN, K., TEKKEL, M., VEIDEBAUM, T.,
2596
AUVINEN, A. and BOICE, J.D., JR. (1998). “Do recorded doses overestimate true doses received by
2597
Chernobyl cleanup workers? Results of cytogenetic analyses of Estonian workers by fluorescence in situ
2598
hybridization,” Radiat. Res. 150(2), 237–249.
2599
LIU, J.J., FREEDMAN, D.M., LITTLE, M.P., DOODY, M.M., ALEXANDER, B.H., KITAHARA, C.M.,
2600
LEE, T., RAJARAMAN, P., MILLER, J.S., KAMPA, D.M., SIMON S.L., PRESTON, D.L. and LINET,
2601
M.S. (2014). “Work history and mortality risks in 90,268 US radiological technologists,” Occup. Environ.
2602
Med. 71(12), 819–835.
2603
LIVSHITS, L.A., MALYARCHUK, S.G., KRAVCHENKO, S.A., MATSUKA, G.H., LUKYANOVA, E.M.,
2604
ANTIPKIN, Y.G., ARABSKAYA, L.P., PETIT, E., GIRAUDEAU, F., GOURMELON, P.,
2605
VERGNAUD, G. and LE GUEN, B. (2001). “Children of Chernobyl cleanup workers do not show
2606
elevated rates of mutations in minisatellite alleles,” Radiat. Res. 155(1), 74–80.
2607
LLOYD, D.C., EDWARDS, A.A., LEONARD, A., DEKNUDT, G.L., VERSCHAEVE, L., NATARAJAN,
2608
A.T., DARROUDI, F., OBE, G., PALITTI, F. and TANZARELLA, C. (1992). “Chromosomal
2609
aberrations in human lymphocytes induced in vitro by very low doses of x-rays,” Int. J. Radiat. Biol.
2610
61(3), 335–343.
2611
LOBRICH, M., SHIBATA, A., BEUCHER, A., FISHER, A., ENSMINGER, M., GOODARZI, A.A.,
2612
BARTON, O. and JEGGO, P.A. (2010). “H2AX foci analysis for monitoring DNA double-strand break
2613
repair: Strengths, limitations and optimization,” Cell Cycle 9(4), 662–669.
2614
LOHMANN, D. (2010). “Retinoblastoma,” Adv. Exp. Med. Biol. 685, 220–227.
2615
LORIMORE, S.A. and WRIGHT, E.G. (2003). “Radiation-induced genomic instability and bystander effects:
2616
related inflammatory-type responses to radiation-induced stress and injury? A review,” Int. J. Radiat.
2617
Biol. 79(1), 15–25.
2618
LOUCAS, B.D., EBERLE, R., BAILEY, S.M. and CORNFORTH, M.N. (2004). “Influence of dose rate on
2619
the induction of simple and complex chromosome exchanges by gamma rays,” Radiat. Res. 162(4), 339–
2620
349.
88
NCRP SC 1-21
Draft of March 31, 2015 [MR]
2621
NOT TO BE DISSEMINATED OR REFERENCED
LOUCAS, B.D., DURANTE, M., BAILEY, S.M. and CORNFORTH, M.N. (2013). “Chromosome damage in
2622
human cells by  rays,  particles and heavy ions: Track interactions in basic dose-response
2623
relationships,” Radiat. Res. 179(1), 9–20.
2624
LU, T.P., LAI, L.C., LIN, B.I., CHEN, L.H., HSIAO, T.H., LIBER, H.L., COOK, J.A., MITCHELL, J.B.,
2625
TSAI, M.H. and CHUANG, E.Y. (2010). “Distinct signaling pathways after higher or lower doses of
2626
radiation in three closely related human lymphoblast cell lines,” Int. J. Radiat. Oncol. Biol. Phys 76(1),
2627
212–219.
2628
LUBIN, J.H., BOICE, J.D., JR., EDLING, C., HORNUNG, R.W., HOWE, G., KUNZ, E., KUSIAK, R.A.,
2629
MORRISON, H.I., RADFORD, E.P., SAMET, J.M., TIRMARCHE, M., WOODARD, A. and YAO,
2630
S.X. (1995). “Radon-exposed underground miners and inverse dose-rate (protraction enhancement)
2631
effects,” Health Phys. 69(4), 494–500.
2632
2633
2634
LUEBECK, E.G, CURTIUS, K., JEON, J. and HAZELTON, W.D. (2013). “Impact of tumor progression on
cancer incidence curves,” Cancer Res. 73(3), 1086–1096.
MANCUSO, M., PASQUALI, E., LEONARDI, S., TANORI, M., REBESSI, S., DI MAJO, V.,
2635
PAZZAGLIA, S., PIA TONI, M., PIMPINELLA, M., COVELLI, V. and SARAN, A. (2008).
2636
“Oncogenic bystander radiation effects in Patched heterozygous mouse cerebellum,” Proc. Natl. Acad.
2637
Sci. USA 105(34), 12445–12450.
2638
MANCUSO, M., PASQUALI, E., LEONARDI, S., REBESSI, S., TANORI, M., GIARDULLO, P., BORRA,
2639
F., PAZZAGLIA, S., NAUS, C.C., DI MAJO, V. and SARAN, A. (2011). “Role of connexin43 and ATP
2640
in long-range bystander radiation damage and oncogenesis in vivo,” Oncogene 30(45), 4601–4608.
2641
MANCUSO, M., PASQUALI, E., GIARDULLO, P., LEONARDI, S., TANORI, M., DI MAJO, V.,
2642
PAZZAGLIA, S. and SARAN, A. (2012). “The radiation bystander effect and its potential implications
2643
for human health,” Curr. Mol. Med. 12(5), 613–624.
2644
MANCUSO, M., GIARDULLO, P., LEONARDI, S., PASQUALI, E., CASCIATI, A., DE STEFANO, I.,
2645
TANORI, M., PAZZAGLIA, S. and SARAN, A. (2013). “Dose and spatial effects in long-distance
2646
radiation signaling in vivo: Implications for abscopal tumorigenesis,” Int. J. Radiat. Oncol. Biol. Phys.
2647
85(3), 813–819.
2648
MAO, J.H., WU, D., DELROSARIO, R., CASTELLANOS, A., BALMAIN, A. and PEREZ-LOSADA, J.
2649
(2008). “Atm heterozygosity does not increase tumor susceptibility to ionizing radiation alone or in a p53
2650
heterozygous background,” Oncogene 27(51), 6596–6600.
2651
2652
MARDER, B.A. and MORGAN, W.F. (1993). “Delayed chromosomal instability induced by DNA damage,”
Mol. Cell Biol. 13(11), 6667–6677.
2653
MATHEWS, J.D., FORSYTHE, A.V., BRADY, Z., BUTLER, M.W., GOERGEN, S.K., BYRNES, G.B.,
2654
GILES, G.G., WALLACE, A.B., ANDERSON, P.R., GUIVER, T.A., MCGALE, P., CAIN, T.M.,
89
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
2655
DOWTY, J.G., BICKERSTAFFE, A.C. and DARBY, S.C. (2013). “Cancer risk in 680 000 people
2656
exposed to computed tomography scans in childhood or adolescence: Data linkage study of 11 million
2657
Australians,” BMJ 346, f2360.
2658
2659
MATTHAIOS, D., HOUNTIS, P., KARAKITSOS, P., BOUROS, D. and KAKOLYRIS, S. (2013). “H2AX a
promising biomarker for lung cancer: A review,” Cancer Invest. 31(9), 582–599.
2660
MCBRIDE, W.H., CHIANG, C.S., OLSON, J.L., WANG, C.C., HONG, J.H., PAJONK, F., DOUGHERTY,
2661
G.J., IWAMOTO, K.S., PERVAN, M. and LIA, Y.P. (2004). “A sense of danger from radiation,” Radiat.
2662
Res. 162(1), 1–19.
2663
MENARD, C., JOHANN, D., LOWENTHAL, M., MUANZA, T., SPROULL, M., ROSS, S., GULLEY, J.,
2664
PETRICOIN, E., COLEMAN, C.N., WHITELEY, G., LIOTTA, L. and CAMPHAUSEN, K. (2006).
2665
“Discovering clinical biomarkers of ionizing radiation exposure with serum proteomic analysis,” Cancer
2666
Res. 66(3), 1844–1850.
2667
METTLER, F.A. JR., BHARGAVAN, M., FAULKNER, K., GILLEY, D.B., GRAY, J.E., IBBOTT, G.S.,
2668
LIPOTI, J.A., MAHESH, M., MCCROHAN, J.L., STABIN, M.G., THOMADSEN, B.R. and
2669
YOSHIZUMI, T.T. (2009). “Radiologic and nuclear medicine studies in the United States and worldwide:
2670
Frequency, radiation dose, and comparison with other radiation sources—1950–2007. Radiology 253(2),
2671
520–531.
2672
2673
2674
MEZENTSEV, A. and AMUNDSON, S.A. (2011). “Global gene expression responses to low- or high-dose
radiation in a human three-dimensional tissue model,” Radiat. Res. 175(6), 677–688.
MONSIEURS, M.A., THIERENS, H.M., VRAL, A.M., VAN DE WIELE, C., DE RIDDER, L.I. and
2675
DIERCKX, R.A. (2000). “Adaptive response in patients treated with 131I,” J. Nucl. Med. 41(1), 17–22.
2676
MONSON, R.R. and MACMAHON, B. (1984). “Prenatal x-ray exposure and cancer in children,” pages 97 to
2677
105 in Radiation Carcinogenesis: Epidemiology and Biological Significance, Boice, J.D. Jr. and
2678
Fraumeni, J.F., Jr., Eds. (Raven Press, New York).
2679
2680
2681
2682
2683
2684
2685
MOOLGAVKAR, S.H. and KNUDSON, A.G., JR. (1981). “Mutation and cancer: A model for human
carcinogenesis,” J. Natl. Cancer Inst. 66(6), 1037–1052.
MOOLGAVKAR, S.H. and LUEBECK, E.G. (1992). “Multistage carcinogenesis: Population-based model
for colon cancer,” J. Natl. Cancer Inst. 84(8), 610–618.
MOOLGAVKAR, S.H. and VENZON, D.J. (1979). “Two-event models for carcinogenesis: Incidence curves
for childhood and adult tumors,” Math. Biosci. 47(1–2), 55–77.
MOORE, D.H., II and TUCKER, J.D. (1999). “Biological dosimetry of Chernobyl cleanup workers:
2686
Inclusion of data on age and smoking provides improved radiation dose estimates,” Radiat. Res. 152(6),
2687
655–664.
90
NCRP SC 1-21
Draft of March 31, 2015 [MR]
2688
2689
2690
NOT TO BE DISSEMINATED OR REFERENCED
MORGAN, W.F. (2003a). “Non-targeted and delayed effects of exposure to ionizing radiation: I. Radiationinduced genomic instability and bystander effects in vitro,” Radiat. Res. 159(5), 567–580.
MORGAN, W.F. (2003b). “Non-targeted and delayed effects of exposure to ionizing radiation: II. Radiation-
2691
induced genomic instability and bystander effects in vivo, clastogenic factors and transgenerational
2692
effects,” Radiat. Res. 159(5), 581–596.
2693
2694
2695
2696
MORGAN, W.F. and BAIR, W.J. (2013). “Issues in low dose radiation biology: The controversy continues.
A perspective,” Radiat. Res. 179(5), 501–510.
MORGAN, W.F., HARTMANN, A., LIMOLI, C.L., NAGAR, S. and PONNAIYA, B. (2002). “Bystander
effects in radiation-induced genomic instability,” Mutat. Res. 504(1–2), 91–100.
2697
MOSEEVA, M.B., AZIZOVA, T.V., GRIGORYEVA, E.S. and HAYLOCK, R. (2014). “Risks of circulatory
2698
diseases among Mayak PA workers with radiation doses estimated using the improved Mayak Worker
2699
Dosimetry System 2008,” Radiat. Environ. Biophys. 53(2), 469–477.
2700
2701
2702
MUIRHEAD, C.R. and DARBY, S.C. (1987). “Modelling the relative and absolute risks of radiation-induced
cancers,” J. R. Statist. Soc. A 150(2), 83–118.
MUIRHEAD, C.R., COX, R., STATHER, J.W., MACGIBBON, B.H., EDWARDS, A.A. and HAYLOCK,
2703
R.G.E. (1993). “Estimates of late radiation risks to the UK population,” pages 15 to 157 in Documents of
2704
the NRPB 4(4) (Public Health England, London).
2705
MUIRHEAD, C.R., O’HAGAN, J.A., HAYLOCK, R.G., PHILLIPSON, M.A., WILLCOCK, T.,
2706
BERRIDGE, G.L. and ZHANG, W. (2009). “Mortality and cancer incidence following occupational
2707
radiation exposure: Third analysis of the National Registry for Radiation Workers,” Br. J. Cancer 100(1),
2708
206–212.
2709
MUKHERJEE, D., COATES, P.J., LORIMORE, S.A. and WRIGHT, E.G. (2012). “The in vivo expression of
2710
radiation-induced chromosomal instability has an inflammatory mechanism,” Radiat. Res. 177(1), 18–24.
2711
MULLER, H.J. (1927). “Artificial transmutation of the gene,” Science 66(1699), 84–87.
2712
NAIR, R.R., RAJAN, B., AKIBA, S., JAYALEKSHMI, P., NAIR, M.K., GANGADHARAN, P., KOGA, T.,
2713
MORISHIMA, H., NAKAMURA, S. and SUGAHARA, T. (2009). “Background radiation and cancer
2714
incidence in Kerala, India-Karanagappally cohort study,” Health Phys. 96(1), 55–66.
2715
NAKANO, M., KODAMA, Y., OHTAKI, K., NAKASHIMA, E., NIWA, O., TOYOSHIMA, M. and
2716
NAKAMURA, N. (2007). “Chromosome aberrations do not persist in the lymphocytes or bone marrow
2717
cells of mice irradiated in utero or soon after birth,” Radiat. Res. 167(6), 693–702.
2718
NAKANO, M. NISHIMURA, M., HAMASAKI, K., MISHIMA, S., YOSHIDA, M., NAKATA, A.,
2719
SHIMADA, Y., NODA, A., NAKAMURA, N. and KODAMA, Y. (2014). “Fetal irradiation of rats
2720
induces persistent translocations in mammary epithelial cells similar to the level after adult irradiation,
2721
but not in hematolymphoid cells,” Radiat. Res. 181(2) 172–176.
91
NCRP SC 1-21
Draft of March 31, 2015 [MR]
2722
2723
2724
NOT TO BE DISSEMINATED OR REFERENCED
NA/NRC (2006). National Academies/National Research Council. Health Risks from Exposure to Low
Levels of Ionizing Radiation, BEIR VII, Phase 2 (National Academies Press, Washington).
NA/NRC (2012). National Academies/National Research Council. Technical Evaluation of the NASA Model
2725
for Cancer Risk to Astronauts Due to Space Radiation (National Academies Press, Washington).
2726
NAGASAWA, H. and LITTLE, J.B. (1992). “Induction of sister chromatid exchanges by extremely low
2727
2728
doses of alpha-particles,” Cancer Res. 52(22), 6394–6396.
NCRP (1980). National Council on Radiation Protection and Measurements. Influence of Dose and Its
2729
Distribution in Time on Dose-Response Relationships for Low-LET Radiations, NCRP Report No. 64
2730
(National Council on Radiation Protection and Measurements. Bethesda, Maryland).
2731
NCRP (2001). National Council on Radiation Protection and Measurements. Evaluation of the Linear-
2732
Nonthreshold Dose-Response Model for Ionizing Radiation, NCRP Report No. 136 (National Council on
2733
Radiation Protection and Measurements. Bethesda, Maryland).
2734
NCRP (2005). National Council on Radiation Protection and Measurements. Extrapolation of Radiation-
2735
Induced Cancer Risks from Nonhuman Experimental Systems to Humans, NCRP Report No. 150
2736
(National Council on Radiation Protection and Measurements. Bethesda, Maryland).
2737
NCRP (2007). National Council on Radiation Protection and Measurements. Uncertainties in the
2738
Measurement and Dosimetry of External Radiation, NCRP Report No. 158 (National Council on
2739
Radiation Protection and Measurements. Bethesda, Maryland).
2740
NCRP (2009a). National Council on Radiation Protection and Measurements. Low Dose and Low Dose-Rate
2741
Radiation Effects and Models, Proceedings of the Forty-Fourth Annual Meeting, Health Phys. 97(5),
2742
373–541.
2743
NCRP (2009b). National Council on Radiation Protection and Measurements. Uncertainties in Internal
2744
Radiation Dose Assessment, NCRP Report No. 164 (National Council on Radiation Protection and
2745
Measurements. Bethesda, Maryland).
2746
NCRP (2009c). National Council on Radiation Protection and Measurements). Radiation Dose
2747
Reconstruction: Principles and Practices, NCRP Report No. 163 (National Council on Radiation
2748
Protection and Measurements. Bethesda, Maryland).
2749
NCRP (2009d). National Council on Radiation Protection and Measurements. Ionizing Radiation Exposure of
2750
the Populations of the United States, NCRP Report No.160 (National Council on Radiation Protection
2751
and Measurements. Bethesda, Maryland).
2752
NCRP (2011). National Council on Radiation Protection and Measurements. Second Primary Cancers and
2753
Cardiovascular Disease After Radiation Therapy, NCRP Report No. 170 (National Council of Radiation
2754
Protection and Measurements, Bethesda, Maryland).
92
NCRP SC 1-21
Draft of March 31, 2015 [MR]
2755
NOT TO BE DISSEMINATED OR REFERENCED
NCRP (2012). National Council on Radiation Protection and Measurements. Uncertainties in the Estimation
2756
of Radiation Risks and Probability of Disease Causation, NCRP Report No. 171 (National Council on
2757
Radiation Protection and Measurements, Bethesda, Maryland).
2758
NCRP (2015). National Council on Radiation Protection and Measurements. Where Are the Radiation
2759
Professionals (WARP)?, Synopsis of NCRP Statement No. 12, http://ncrponline.org/Publications/
2760
Statements/WARP_Synopsis_Statement12.pdf (accessed March 22, 2015) (National Council on
2761
Radiation Protection and Measurements, Bethesda, Maryland).
2762
NEEL, J.V., SATOH, C., HAMILTON, H.B., OTAKE, M., GORIKI, K., KAGEOKA, T., FUJITA, M.,
2763
NERIISHI, S. and ASAKAWA, J. (1980). “Search for mutations affecting protein structure in children of
2764
atomic bomb survivors: Preliminary report,” Proc. Natl. Acad. Sci. USA 77(7), 4221–4225.
2765
NEUMAIER, T., SWENSON, J., PHAM, C., POLYZOS, A., LO, A.T., YANG, P., DYBALL, J.,
2766
ASAITHAMBY, A., CHEN, D.J., BISSELL, M.J., THALHAMMER, S. and COSTES, S.V. (2012).
2767
“Evidence for formation of DNA repair centers and dose-response nonlinearity in human cells,” Proc.
2768
Natl. Acad. Sci. USA 109(2), 443–448.
2769
2770
NIKJOO, H. and KHVOSTUNOV, I.K. (2003). “Biophysical model of the radiation-induced bystander
effect,” Int. J. Radiat. Biol. 79(1), 43–52.
2771
NIOSH (2014). National Institute for Occupational Safety and Health. Exposome and Exposomics: Overview,
2772
http://www.cdc.gov/niosh/topics/exposome/ (accessed 2014) (Centers for Disease Control and
2773
Prevention, Atlanta, Georgia).
2774
OHTAKI, K., KODAMA, Y., NAKANO, M., ITOH, M., AWA, A.A., COLOGNE, J. and NAKAMURA, N.
2775
(2004). “Human fetuses do not register chromosome damage inflicted by radiation exposure in lymphoid
2776
precursor cells except for a small but significant effect at low doses,” Radiat. Res. 161(4), 373–379.
2777
OKAYASU, R., SUETOMI, K., YU, Y., SILVER, A., BEDFORD, J.S., COX, R. and ULLRICH, R.L.
2778
(2000). “A deficiency in DNA repair and DNA-PKcs expression in the radiosensitive BALB/c mouse,”
2779
Cancer Res. 60(16), 4342–4345.
2780
2781
2782
OLIVIERI, G., BODYCOTE, J. and WOLFF, S. (1984). “Adaptive response of human lymphocytes to low
concentrations of radioactive thymidine,” Science 223(4636), 594–597.
OZASA, K., SHIMIZU, Y., SUYAMA, A., KASAGI, F., SODA, M., GRANT, E.J., SAKATA, R.,
2783
SUGIYAMA, H. and KODAMA, K. (2012). “Studies of the mortality of atomic bomb survivors, Report
2784
14, 1950–2003: An overview of cancer and noncancer diseases,” Radiat. Res. 177(3), 229–243.
2785
PADOVANI, L., APPOLLONI, M., ANZIDEI, P., TEDESCHI, B., CAPOROSSI, D., VERNOLE, P. and
2786
MAURO, F. (1995). “Do human lymphocytes exposed to the fallout of the Chernobyl accident exhibit an
2787
adaptive response? 1. Challenge with ionizing radiation,” Mutat. Res. 332(1–2), 33–38.
93
NCRP SC 1-21
Draft of March 31, 2015 [MR]
2788
NOT TO BE DISSEMINATED OR REFERENCED
PARK, E.C., YOON, J.B., SEONG, J.S., CHOI, K.S., KONG, E.S., KIM, Y.J., PARK, Y.M. and PARK,
2789
E.M. (2006). “Effect of ionizing radiation on rat tissue: Proteomic and biochemical analysis,” Prep.
2790
Biochem. Biotechnol. 36(1), 19–35.
2791
PEARCE, M.S., SALOTTI, J.A., LITTLE, M.P., MCHUGH, K., LEE, C., KIM, K.P., HOWE, N.L.,
2792
RONCKERS, C.M., RAJARAMAN, P., CRAFT, A.W., PARKER, L. and BERRINGTON DE
2793
GONZALEZ, A. (2012). “Radiation exposure from CT scans in childhood and subsequent risk of
2794
leukaemia and brain tumours: A retrospective cohort study,” Lancet 380(9840), 499–505.
2795
PENG, Y., BROWN, N., FINNON, R., WARNER, C.L., LIU, X., GENIK, P.C., CALLAN, M.A., RAY,
2796
F.A., BORAK, T.B., BADIE, C., BOUFFLER, S.D., ULLRICH, R.L., BEDFORD, J.S. and WEIL, M.M.
2797
(2009). “Radiation leukemogenesis in mice: Loss of PU.1 on chromosome 2 in CBA and C57BL/6 mice
2798
after irradiation with 1 GeV/nucleon 56Fe ions, x rays or gamma rays. Part I. Experimental observations,”
2799
Radiat. Res. 171(4), 474–483.
2800
PENG, A., LEWELLYN, A.L., SCHIEMANN, W.P. and MALLER, J.L. (2010). “Repo-man controls a
2801
protein phosphatase 1-dependent threshold for DNA damage checkpoint activation,” Curr. Biol. 20(5),
2802
387–396.
2803
PLEASANCE, E.D., STEPHENS, P.J., O’MEARA, S., MCBRIDE, D.J., MEYNERT, A., JONES, D., LIN,
2804
M.L., BEARE, D., LAU, K.W., GREENMAN, C., VARELA, I., NIK-ZAINAL, S., DAVIES, H.R.,
2805
ORDONEZ, G.R., MUDIE, L.J., LATIMER, C., EDKINS, S., STEBBINGS, L., CHEN, L., JIA, M.,
2806
LEROY, C., MARSHALL, J., MENZIES, A., BUTLER, A., TEAGUE, J.W., MANGION, J., SUN,
2807
Y.A., MCLAUGHLIN, S.F., PECKHAM, H.E., TSUNG, E.F., COSTA, G.L., LEE, C.C., MINNA, J.D.,
2808
GAZDAR, A., BIRNEY, E., RHODES, M.D., MCKERNAN, K.J., STRATTON, M.R., FUTREAL, P.A.
2809
and CAMPBELL, P.J. (2010). “A small-cell lung cancer genome with complex signatures of tobacco
2810
exposure,” Nature 463(7278), 184–190.
2811
POHL-RULING, J., FISCHER, P., HAAS, O., OBE, G., NATARAJAN, A.T., VAN BUUL, P.P.W.,
2812
BUCKTON, K.E., BIANCHI, N.O., LARRAMENDY, M., KUCEROVA, M., POLIKOVA, Z.,
2813
LEONARD, A., FABRY, L., PALITTI, F., SHARMA, T., BINDER, W., MUKHERJEE, R.N. and
2814
MUKHERJEE, U. (1983). “Effect of low-dose acute x-irradiation on the frequencies of chromosomal
2815
aberrationsin human peripheral lymphocytes in vitro,” Mutat. Res. 110(1), 71–82.
2816
PONNAIYA, B., CORNFORTH, M.N. and ULLRICH, R.L. (1997). “Radiation-induced chromosomal
2817
instability in BALB/c and C57BL/6 mice: The difference is as clear as black and white,” Radiat Res.
2818
147(2), 121–125.
94
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
2819
PONOMAREVA, O.N., ROSE, J.A., LASAREV, M., RASEY, J. and TURKER, M.S. (2002). “Tissue-
2820
specific deletion and discontinuous loss of heterozygosity are signatures for the mutagenic effects of
2821
ionizing radiation in solid tissues,” Cancer Res. 62(5), 1518–1523.
2822
POSTIGLIONE, I., CHIAVIELLO, A. and PALUMBO, G. (2010). “Twilight effects of low doses of ionizing
2823
radiation on cellular systems: A bird's eye view on current concepts and research,” Med. Oncol. 27(2),
2824
495–509.
2825
2826
2827
2828
PRESTON, R.J. (2015). “Integrating basic radiobiological science and epidemiological studies: Why and
how,” Health Phys. 108(2), 125–130.
PRESTON, R.J. and WILLIAMS, G.M. (2005). “DNA-reactive carcinogens: Mode of action and human
cancer hazard,” Crit. Rev. Toxicol. 35(8–9), 673–683.
2829
PRESTON, D.L., MATTSSON, A., HOLMBERG, E., SHORE, R., HILDRETH, N.G. and BOICE, J.D., JR.
2830
(2002). “Radiation effects on breast cancer risk: A pooled analysis of eight cohorts,” Radiat. Res. 158(2),
2831
220–235.
2832
PRESTON, D., RON, E., TOKUOKA, S., FUNAMOTO, S., NISHI, N., SODA, M., MABUCHI, K. and
2833
KODAMA, K. (2007). “Solid cancer incidence in atomic bomb survivors: 1958–1998,” Radiat. Res.
2834
168(1), 1–64.
2835
PRESTON, D.L., CULLINGS, H., SUYAMA, A., FUNAMOTO, S., NISHI, N., SODA, M., MABUCHI, K.,
2836
KODAMA, K., KASAGI, F. and SHORE, R.E. (2008). “Solid cancer incidence in atomic bomb survivors
2837
exposed in utero or as young children,” J. Natl. Cancer Inst. 100(6), 428–436.
2838
PUTNAM, C.D., JAEHNIG, E.J. and KOLODNER, R.D. (2009). “Perspectives on the DNA damage and
2839
replication checkpoint responses in Saccharomyces cerevisiae,” DNA Repair (Amst.) 8(9), 974–982.
2840
RAHU, K., AUVINEN, A., HAKULINEN, T., TEKKEL, M., INSKIP, P.D., BROMET, E.J., BOICE, J.D.,
2841
JR. and RAHU, M. (2013). “Chernobyl cleanup workers from Estonia: Follow-up for cancer incidence
2842
and mortality,” J. Radiol. Protect. 33(2), 395–411.
2843
2844
2845
RAPPAPORT, S.M. (2011). “Implications of the exposome for exposure science,” J. Expo. Sci. Environ.
Epidemiol. 21(1), 5–9.
REDPATH, J. L., LIANG, D., TAYLOR, T. H., CHRISTIE, C. and ELMORE, E. (2001). “The shape of the
2846
dose-response curve for radiation-induced neoplastic transformation in vitro: Evidence for an adaptive
2847
response against neoplastic transformation at low doses of low-LET radiation,” Radiat. Res. 156(6), 700–
2848
707.
2849
ROGAKOU, E.P., PILCH, D.R., ORR, A.H., IVANOVA, V.S. and BONNER, W.M. (1998). “DNA double-
2850
stranded breaks induce histone H2AX phosphorylation on serine 139,” J. Biol. Chem. 1273(10), 5858–
2851
5868.
95
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
2852
RON, E., LUBIN, J.H., SHORE, R.E., MABUCHI, K., MODAN, B., POTTERN, L.M., SCHNEIDER, A.B.,
2853
TUCKER, M.A. and BOICE, J.D., JR. (1995). “Thyroid-cancer after exposure to external radiation: A
2854
pooled analysis of seven studies,” Radiat. Res. 141(3), 259–277.
2855
2856
2857
ROTHKAMM, K. and LOBRICH, M. (2003). “Evidence for a lack of DNA double-strand break repair in
human cells exposed to very low x-ray doses,” Proc. Natl. Acad. Sci. USA, 100(9), 5057–5062.
ROTHKAMM, K., BARNARD, S., AINSBURY, E.A., AL-HAFIDH, J., BARQUINERO, J.F.,
2858
LINDHOLM, C., MOQUET, J., PERALA, M., ROCH-LEFEVRE, S., SCHERTHAN, H., THIERENS,
2859
H., VRAL, A. and VANDERSICKEL, V. (2013). “Manual versus automated γ-H2AX foci analysis
2860
across five European laboratories: Can this assay be used for rapid biodosimetry in a large scale radiation
2861
accident?,” Mutat. Res. 756(1–2), 170–173.
2862
RUSSELL, L.B. and HUNSICKER, P.R. (2012). “The effect of dose rate on the frequency of specific-locus
2863
mutations induced in mouse spermatogonia is restricted to larger lesions; a retrospective analysis of
2864
historical data,” Radiat. Res. 177(5), 555–564.
2865
2866
2867
2868
RUSSELL, W.L. and KELLY, E.M. (1982). “Mutation frequencies in male mice and the estimation of
genetic hazards of radiation in men,” Proc. Natl. Acad. Sci. USA 79(2), 542–544.
RUSSELL, W.L., RUSSELL, L.B. and KELLY, E.M. (1958). “Radiation dose rate and mutation frequency,”
Science 128(3338), 1546–15450.
2869
SACKETT, D.L. (1979). “Bias in analytic research,” J. Chronic Dis. 32(1–2), 51–63.
2870
SAKATA, R., KLEINERMAN, R.A., MABUCHI, K., STOVALL, M., SMITH, S.A., WEATHERS, R.,
2871
WACTAWSKI-WENDE, J., COOKFAIR, D.L., BOICE, J.D., JR. and INSKIP, P.D. (2012). “Cancer
2872
mortality following radiation therapy for benign gynecologic disorders,” Radiat. Res. 178(4), 266–279.
2873
SANKARANARAYANAN, K. and CHAKRABORTY, R. (2001). “Impact of cancer predisposition and
2874
radiosensitivity on the population risk of radiation-induced cancers,” Radiat. Res. 156(5 Pt 2), 648–656.
2875
SATOH, C., TAKAHASHI, N., ASAKAWA, J., KODAIRA, M., KUICK, R., HANASH, S.M. and NEEL,
2876
J.V. (1996). “Genetic analysis of children of atomic bomb survivors,” Environ. Health Persp. 104(Suppl.
2877
3), 511–519.
2878
SAWANT, S.G., RANDERS-PEHRSON, G., GEARD, C.R., BRENNER, D.J. and HALL, E.J. (2001). “The
2879
bystander effect in radiation oncogenesis: I. Transformation in C3H 10T1/2 cells in vitro can be initiated
2880
in the unirradiated neighbors of irradiated cells,” Radiat. Res. 155(3), 397–401.
2881
2882
2883
SCHULL, W.J. (2003). “The children of atomic bomb survivors: A synopsis,” J. Radiol. Prot. 23(4), 369–
384.
SEED, J., CARNEY, E.W., CORLEY, R.A., CROFTON, K.M., DESESSO, J.M., FOSTER, P.M.,
2884
KAVLOCK, R., KIMMEL, G., KLAUNIG, J., MEEK, M.E., PRESTON, R.J., SLIKKER, W., JR.,
2885
TABACOVA, S., WILLIAMS, G.M., WILTSE, J., ZOELLER, R.T., FENNER-CRISP, P. and PATTON,
96
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
2886
D.E. (2005). “Overview: Using mode of action and life stage information to evaluate the human relevance
2887
of animal toxicity data,” Crit. Rev. Toxicol. 35(8–9), 664–672.
2888
SELMANSBERGER, M., FEUCHTINGER, A., ZURNADZHY, L., MICHNA, A., KAISER, J.C., ABEND,
2889
M., BRENNER, A., BOGDANOVA, T., WALCH, A., UNGER, K., ZITZELSBERGER, H. and HESS, J.
2890
(2014). “CLIP2 as radiation biomarker in papillary thyroid carcinoma,” Oncogene. doi:
2891
10.1038/onc.2014.311.
2892
SEVANKAEV, A.V., LLOYD, D.C., EDWARDS, A.A., KHVOSTUNOV, I.K., MIKHAILOVA, G.F.,
2893
GOLUB, E.V., SHEPEL, N.N., NADEJINA, N.M., GALSTIAN, I.A., NUGIS, V.YU., BARRIOS L.,
2894
CABALLIN, M.R. and BARQUINERO, J.F. (2005). “A cytogenetic follow-up of some highly irradiated
2895
victims of the Chernobyl accident,” Radiat. Prot. Dosim. 113(2), 152–161.
2896
2897
2898
SHAHEEN, M., ALLEN, C., NICKOLOFF, J.A. and HROMAS, R. (2011). “Synthetic lethality: Exploiting
the addiction of cancer to DNA repair,” Blood 117(23), 6074–6082.
SHILNIKOVA, N.S., PRESTON, D.L., RON, E., GILBERT, E.S., VASSILENKO, E.K., ROMANOV, S.A.,
2899
KUZNETSOVA, I.S., SOKOLNIKOV, M.E., OKATENKO, P.V., KRESLOV, V.V. and
2900
KOSHURNIKOVA, N.A. (2003). “Cancer mortality risk among workers at the Mayak nuclear complex,”
2901
Radiat. Res. 159(6), 787–798.
2902
SHIMIZU, Y., KODAMA, K., NISHI, N., KASAGI, F., SUYAMA, A., SODA, M., GRANT, E.J.,
2903
SUGIYAMA, H., SAKATA, R., MORIWAKI, H., HAYASHI, M., KONDA, M. and SHORE, R.E.
2904
(2010). “Radiation exposure and circulatory disease risk: Hiroshima and Nagasaki atomic bomb survivor
2905
data, 1950–2003,” BMJ 340, b5349.
2906
2907
2908
2909
SHORE, R.E., NERIISHI, K. and NAKASHIMA, E. (2010). “Epidemiologic studies of cataract risk at lowto-moderate radiation doses: (Not) seeing is believing,” Radiat. Res. 174(6), 889–894.
SHURYAK, I., SACHS, R.K. and BRENNER, D.J. (2010). “Cancer risks after radiation exposure in middle
age,” J. Natl. Cancer Inst. 102(21), 1628–1636.
2910
SIGNORELLO, L.B., MULVIHILL, J.J., GREEN, D.M., MUNRO, H.M., STOVALL, M., WEATHERS,
2911
R.E., MERTENS, A.C., WHITTON, J.A., ROBISON, L.L. and BOICE, J.D., JR. (2012). “Congenital
2912
anomalies in the children of cancer survivors: A report from the Childhood Cancer Survivor Study,”
2913
J. Clin. Oncol. 30(3), 239–245.
2914
SLEBOS, R.J.C., LITTLE, R.E., UMBACH, D.M., ANTIPKIN, Y., ZADAOROZHNAJA, T.D., MENDEL,
2915
N.A., SOMMER, C.A., CONWAY, K., PARRISH, E., GULINO, S. and TAYLOR, J.A. (2004).
2916
“Mutation research/genetic toxicology and environmental mutagenesis,” Mutat. Res. 559(1–2), 143–151.
2917
SNIJDERS, A.M., MARCHETTI, F., BHATNAGAR, S., DURU, N., HAN, J., HU, Z., MAO, J.H., GRAY,
2918
J.W. and WYROBEK, A.J. (2012). “Genetic differences in transcript responses to low-dose ionizing
2919
radiation identify tissue functions associated with breast cancer susceptibility,” PLoS One 7(10), e45394.
97
NCRP SC 1-21
Draft of March 31, 2015 [MR]
2920
2921
2922
NOT TO BE DISSEMINATED OR REFERENCED
SOBELS, F.H. (1993). “Approaches to assessing genetic risks from exposure to chemicals,” Environ. Health
Perspect. 101(Supp 3), 327–332.
SOWA, M.B., CHRISLER, W.B., ZENS, K.D., ASHJIAN, E.J. and OPRESKO, L.K. (2010). “Three-
2923
dimensional culture conditions lead to decreased radiation induced cytotoxicity in human mammary
2924
epithelial cells,” Mutat. Res. 687(1–2), 78–83.
2925
SPIESS, H. and MAYS, C.W. (1973). “Protraction effect on bone sarcoma-induction of 224Ra in children and
2926
adults,” pages 437–450 in Radionuclide Carcinogenesis, AEC Symposium Series 29, CONF-720505,
2927
Sanders, C. L., Busch, R. H., Ballou, J. E. and Mahlum, D.D., Eds., (National Technical Information
2928
Service, Springfield, Virginia).
2929
SPITZ, D.R., AZZAM, E I., LI, J.J. and GIUS, D. (2004). “Metabolic oxidation/reduction reactions and
2930
cellularresponses to ionizing radiation: A unifying concept in stress response biology,” Cancer Metastasis
2931
Rev. 23(3–4), 311–322.
2932
2933
STEWART, A., WEBB, J., GILES, D. and HEWITT, D. (1956). “Malignant disease in childhood and
diagnostic irradiation in utero,” Lancet 268(6940), 447.
2934
STOVALL, M. SMITH, S.A., LANGHOLZ, B.M., BOICE, J.D., JR., SHORE, R.E., ANDERSSON, M.,
2935
BUCHHOLZ, T.A., CAPANU, M., BERNSTEIN, L., LYNCH, C.F., MALONE, K.E., ANTON-
2936
CULVER, H., HAILE, R.W., ROSENSTEIN, B.S., REINER, A.S., THOMAS, D.C. and BERNSTEIN,
2937
J.L. (2008). “Dose to the contralateral breast from radiation therapy and risk of second primary breast
2938
cancer in the WECARE Study,” Int. J. Radiat. Oncol. Biol. Phys. 72(4), 1021–1030.
2939
TAO, Z., AKIBA, S., ZHA, Y., SUN, Q., ZOU, J., LI, J., LIU, Y., YUAN, Y., TOKONAMI, S.,
2940
MORISHOMA, H., KOGA, T., NAKAMURA, S., SUGAHARA, T. and WEI, L. (2012). “Cancer and
2941
non-cancer mortality among inhabitants in the high background radiation area of Yangjiang, China
2942
(1979–1998),” Health Phys. 102(2), 173–181.
2943
2944
TAPIO, S. and JACOB, V. (2007). “Radioadaptive response revisited,” Radiat. Environ. Biophys. 46(1), 1–
12.
2945
TAWN, E.J., WHITEHOUSE, C.A. and MARTIN, F.A. (2000). “Sequential chromosome aberration analysis
2946
following radiotherapy - no evidence for enhanced genomic instability,” Mutat. Res. 465(1–2), 45–51.
2947
2948
2949
TAWN, E.J., WHITEHOUSE, C.A. and TARONE, R.E. (2004). “FISH chromosome aberration analysis on
retired radiation workers from the Sellafield nuclear facility,” Radiat. Res. 162(3), 249–256.
TAWN, E.J., WHITEHOUSE, C.A., WINTHER, J.F., CURWEN, G.B., REES, G.S., STOVALL, M.,
2950
OLSEN, J.H., GULDBERG, P., RECHNITZER, C., SCHRODER, H. and BOICE, J.D., JR. (2005).
2951
“Chromosome analysis in childhood cancer survivors and their offspring—no evidence for radiation
2952
therapy-induced persistent genomic instability,” Mutat. Res. 583(2), 198–206.
98
NCRP SC 1-21
Draft of March 31, 2015 [MR]
2953
2954
NOT TO BE DISSEMINATED OR REFERENCED
TAWN, E.J., WHITEHOUSE, C.A. and RIDDELL, A.E. (2006). “FISH chromosome analysis of plutonium
workers from the Sellafield nuclear facility,” Radiat. Res. 165(5), 592–597.
2955
TAWN, E.J., REES, G.S., LEITH, C., WINTHER, J.F., CURWEN, G.B., STOVALL, M., OLSEN, J.H.,
2956
RECHNITZER, C., SCHROEDER, H., GULDBERG, P. and BOICE, J.D., JR. (2011). “Germline
2957
minisatellite mutations in survivors of childhood and young adult cancer treated with radiation,” Int. J.
2958
Radiat. Biol. 87(3), 330–340.
2959
TAWN, E.J., CURWEN, G.B., REES, G.S. and JONAS, P. (2015). “Germline minisatellite mutations in
2960
workers occupationally exposed to radiation at the Sellafield nuclear facility,” J. Radiat. Protect. 35(1),
2961
21–36.
2962
TAYLOR, M., DEHAINAULT, C., DESJARDINS, L., DOZ, F., LEVY, C., SASTRE, X., COUTURIER, J.,
2963
STOPPA-LYONNET, D., HOUDAYER, C. and GAUTHIER-VILLARS, M. (2007). “Genotype-
2964
phenotype correlations in hereditary familial retinoblastoma,” Hum. Mutat. 28(3), 284–293.
2965
TEDESCHI, B., CAPOROSSI, D.,VERNOLE, P., PADOVANI, L., APPOLLONI, M., ANZIDEI, P. and
2966
MAURO, F. (1995). “Do human lymphocytes exposed to the fallout of the Chernobyl accident exhibit an
2967
adaptive response? 2. Challenge with bleomycin,” Mutat. Res. 332(1–2), 39–44).
2968
THOMAS, R.S., CLEWELL, H.J., III, ALLEN, B.C., YANG, L., HEALY, E. and ANDERSEN, M.E.
2969
(2012). “Integrating pathway-based transcriptomic data into quantitative chemical risk assessment: A five
2970
chemical case study,” Mutat. Res. 746(2), 135–143.
2971
TOLLEFSEN, K.E., SCHOLZ, S., CRONIN, M.T., EDWARDS, S.W., DE KNECHT, J., CROFTON, K.,
2972
GARCIA-REYERO, N., HARTUNG, T., WORTH, A. and PATLEWICZ, G.(2014). “Applying adverse
2973
outcome pathways (AOPs) to support integrated approaches to testing and assessment (IATA),” Regul.
2974
Toxicol. Pharmacol. 70(3), 629–640.
2975
TRAVIS, L.B., RABKIN, C.S., BROWN, L.M., ALLAN, J.M., ALTER, B.P., AMBROSONE, C.B., BEGG,
2976
C.B., CAPORASO, N., CHANOCK, S., DEMICHELE, A., FIGG, W.D., GOSPODAROWICZ, M.K.,
2977
HALL, E.J., HISADA, M., INSKIP, P., KLEINERMAN, R., LITTLE, J.B., MALKIN, D., NG, A.K.,
2978
OFFIT, K., PUI, C.H., ROBISON, L.L., ROTHMAN, N., SHIELDS, P.G., STRONG, L., TANIGUCHI,
2979
T., TUCKER, M.A. and GREENE, M.H. (2006). “Cancer survivorship— genetic susceptibility and
2980
second primary cancers: Research strategies and recommendations,” J. Natl. Cancer Inst. 98(1), 15–25.
2981
2982
2983
2984
TUCKER, J.D. (2008). “Low-dose ionizing radiation and chromosome translocations: A review of the major
considerations for human biological dosimetry,” Mutat. Res. 659(3), 211–220.
TUCKER, J.D. and LUCKINBILL, L.S. (2011). “Estimating the lowest detectable dose of ionizing radiation
by FISH whole-chromosome painting,” Radiat. Res. 175(5), 631–637.
99
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
2985
TUCKER, J.D., TAWN, E.J., HOLDSWORTH, D., MORRIS, S., LANGLOIS, R., RAMSEY, M.J., KATO,
2986
P., BOICE, J.D., JR., TARONE, R.E. and JENSEN, R.H. (1997a). “Biological dosimetry of radiation
2987
workers at the Sellafield nuclear facility,” Radiat. Res. 148(3), 216–226.
2988
2989
TUCKER, J.D., EASTMOND, D.A. and LITTLEFIELD, L.G. (1997b). “Cytogenetic end-points as biological
dosimeters and predictors of risk in epidemiological studies,” IARC Sci. Publ. (142), 185–200.
2990
TURKER, M.S. CONNOLLY, L, DAN, C., LASAREV, M., GAUNY, S., KWOH, E. and KRONENBERG,
2991
A. (2009). “Comparison of autosomal mutations in mouse kidney epithelial cells exposed to iron ions in
2992
situ or in culture,” Radiat. Res. 172(5), 558–566.
2993
TURKER, M.S. GRYGORYEV, D., DAN, C., ECKELMANN, B., LASAREV, M., GAUNY, S., KWOH, E.
2994
KRONENBERG, A. (2013). “Autosomal mutations in mouse kidney epithelial cells exposed to high-
2995
energy protons in vivo or in culture,” Radiat. Res. 179(5), 521–529.
2996
TYBURSKI, J.B., PATTERSON, A.D., KRAUSZ, K.W., SLAVIK, J., FORNACE, A.J., JR., GONZALEZ,
2997
F.J. and IDLE, J.R. (2008). “Radiation metabolomics. 1. Identification of minimally invasive urine
2998
biomarkers for gamma-radiation exposure in mice,” Radiat. Res. 170(1), 1–14.
2999
TYBURSKI, J.B., PATTERSON, A.D., KRAUSZ, K.W., SLAVIK, J., FORNACE, A.J., JR., GONZALEZ,
3000
F.J. and IDLE, J.R. (2009). “Radiation metabolomics. 2. Dose- and time-dependent urinary excretion of
3001
deaminated purines and pyrimidines after sublethal gamma-radiation exposure in mice,” Radiat. Res.
3002
172(1), 42–57.
3003
3004
3005
ULLRICH, R.L. (1984). “Tumor induction in BALB/c mice after fractionated or protracted exposures to
fission-spectrum neutrons,” Radiat. Res. 97(3), 587–597.
UNSCEAR (1993). United Nations Scientific Committee on the Effects of Atomic Radiation. Sources and
3006
Effects of Ionizing Radiation. UNSCEAR 1993 Report to the General Assembly, with Scientific
3007
Annexes, E.94.IX.2 (United Nations Publications, New York).
3008
UNSCEAR (1994). United Nations Scientific Committee on the Effects of Atomic Radiation. Sources and
3009
Effects of Ionizing Radiation. UNSCEAR 1994 Report to the General Assembly, with Scientific
3010
Annexes, E.94.IX.11 (United Nations Publications, New York).
3011
UNSCEAR (2000). United Nations Scientific Committee on the Effects of Atomic Radiation. Sources and
3012
Effects of Ionizing Radiation. UNSCEAR 2000 Report to the General Assembly, with Scientific
3013
Annexes. Volume II: Effects (United Nations Publications, New York).
3014
UNSCEAR (2001). United Nations Scientific Committee on the Effects of Atomic Radiation. Hereditary
3015
Effects of Radiation. UNSCEAR 2001 Report to the General Assembly, with Scientific Annex,
3016
http://www.unscear.org/unscear/en/publications/2001.html (accessed 2014), E.01.IX.2 (United Nations
3017
Publications, New York).
100
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
3018
UNSCEAR (2008a). United Nations Scientific Committee on the Effects of Atomic Radiation. “Annex A.
3019
Epidemiological studies of radiation and cancer,” pages 13 to 322 in Effects of Ionizing Radiation,
3020
UNSCEAR 2006 Report to the General Assembly, with Scientific Annexes. Volume I: Report to the
3021
General Assembly, Scientific Annexes A and B (United Nations Publications, New York).
3022
UNSCEAR (2008b). United Nations Scientific Committee on the Effects of Atomic Radiation. “Annex B.
3023
Epidemiological evaluation of cardiovascular disease and other non-cancer diseases following radiation
3024
exposure,” pages 325 to 383 in Effects of Ionizing Radiation, UNSCEAR 2006 Report to the General
3025
Assembly, with Scientific Annexes. Volume I: Report to the General Assembly, Scientific Annexes A
3026
and B (United Nations Publications, New York).
3027
UNSCEAR (2011). United Nations Scientific Committee on the Effects of Atomic Radiation. “Health effects
3028
due to radiation from the Chernobyl accident,”Annex D in Sources and Effects of Ionizing Radiation.
3029
UNSCEAR 2008 Report to the General Assemblywith Scientific Annexes, Volume II,
3030
http://www.unscear.org/docs/reports/2008/11-80076_Report_2008_Annex_D.pdf. (accessed 2014),
3031
E.11.IX.3 (United Nations Publications, New York).
3032
UNSCEAR (2012). United Nations Scientific Committee on the Effects of Atomic Radiation Biological
3033
Mechanisms of Radiation Actions at Low Doses. A White Paper to Guide the Scientific Committee's
3034
Future Programme of Work, Report V.12-57831 (United Nations Publications, New York).
3035
3036
3037
3038
3039
UPTON, A.C. (1990). “Carcinogenic effects of low-level ionizing radiation,” J. Natl. Cancer Inst. 82(6), 448–
449.
VAKIFAHMETOGLU, H., OLSSON, M. and ZHIVOTOVSKY, B. (2008). “Death through a tragedy:
Mitotic catastrophe,” Cell Death Differ. 15(7), 1153–1162.
VANDENABEELE, P., GALLUZZI, L., VANDEN BERGHE, T. and KROEMER, G. (2010). “Molecular
3040
mechanisms of necroptosis: An ordered cellular explosion,” Nat. Rev. Mol. Cell Biol. 11(10), 700–714.
3041
VINKEN, M. (2013). “The adverse outcome pathway concept: A pragmatic tool in toxicology,” Toxicology
3042
3043
3044
3045
312, 158–165.
VOGELSTEIN, B., PAPADOPOULOS, N., VELCULESCU, V.E., ZHOU, S., DIAZ, L.A., JR. and
KINZLER, K.W. (2013). “Cancer genome landscapes,” Science 339(6127), 1546–1558.
VOZILOVA, A.V., SHAGINA, N.B., DEGTEVA, M.O., EDWARDS, A.A., AINSBURY, E.A., MOQUET,
3046
J.E., HONE, P., LLOYD, D.C., FOMINA, J.N. and DARROUDI, F. (2012). “Preliminary FISH-based
3047
assessment of external dose for residents exposed on the Techa River,” Radiat. Res. 177(1), 84–91.
3048
3049
3050
3051
VRIJHEID, M. (2014). “The exposome: A new paradigm to study the impact of environment on health,”
Thorax 69(9), 876–878.
WAKEFORD, R. (2012). “Radiation effects: Modulating factors and risk assessment—an overview,” Ann.
ICRP 41(3–4), 98-107.
101
NCRP SC 1-21
Draft of March 31, 2015 [MR]
3052
3053
3054
3055
NOT TO BE DISSEMINATED OR REFERENCED
WAKEFORD, R. and LITTLE, M.P. (2003). “Risk coefficients for childhood cancer after intrauterine
irradiation: A review,” Int. J. Radiat. Biol. 79(5), 293–309.
WAKEFORD, R. and TAWN, E.J. (2010). “Editorial: The meaning of low dose and low dose-rate,” J.
Radiol. Protect. 30(1), 1–3.
3056
WANG, Z.Y., BOICE, J.D., JR., WEI, L.X., BEEBE, G.W., ZHA, Y.R., KAPLAN, M.M., TAO, Z.F.,
3057
MAXON, H.R., III, ZHANG, S.Z., SCHNEIDER, A.B., TAN, B., WESSELER, T.A., CHEN, D.,
3058
ERSHOW, A.G., KLEINERMAN, R.A., LITTLEFIELD, L.G. and PRESTON, D. (1990). “Thyroid
3059
nodularity and chromosome aberrations among women in areas of high background radiation in China,”
3060
J. Natl. Cancer Inst. 82(6), 478–485.
3061
WARD, J.F. (1998). “Nature of lesions formed by ionizing radiation,” pages 65 to 84 in DNA Damage and
3062
Repair: Volume II: DNA Repair in Higher Eukaryotes. Nickoloff, J.A. and Hoekstra, M.F., Eds. (Humana
3063
Press, Totowa, New Jersey).
3064
WATSON, G.E., LORIMORE, S.A. and WRIGHT, E.G. (1996). “Long-term in vivo transmission of alpha-
3065
particle-induced chromosomal instability in murine haemopoietic cells,” Int. J. Radiat. Biol. 69(2), 175–
3066
182.
3067
WATSON, G.E., POCOCK, D.A., PAPWORTH, D., LORIMORE, S.A. and WRIGHT, E.G. (2001). “In vivo
3068
chromosomal instability and transmissible aberrations in the progeny of haemopoietic stem cells induced
3069
by high- and low-LET radiations,” Int. J. Radiat. Biol. 77(4), 409–417.
3070
WEINBERG, R.A. (2013). The Biology of Cancer, 2nd ed. (Garland Science, New York).
3071
WESTCOTT, P.M., HALLIWILL, K.D., TO, M.D., RASHID, M., RUST, A.G., KEANE, T.M.,
3072
DELROSARIO, R., JEN, K.Y., GURLEY, K.E., KEMP, C.J., FREDLUND, E., QUIGLEY, D.A.,
3073
ADAMS, D.J. and BALMAIN, A. (2014). “The mutational landscapes of genetic and chemical models of
3074
Kras-driven lung cancer,” Nature, doi: 10.1038/nature13898
3075
3076
WHITEHOUSE, C.A. and TAWN, E.J. (2001). “No evidence for chromosomal instability in radiation
workers with in vivo exposure to plutonium,” Radiat. Res. 156(5), 467–475.
3077
WIESE, C., GAUNY, S.S., LIU, W.C., CHERBONNEL-LASSERRE, C.L. and KRONENBERG, A. (2001).
3078
“Different mechanisms of radiation-induced loss of heterozygosity in two human lymphoid cell lines
3079
from a single donor,” Cancer Res. 61(3), 1129–1137.
3080
3081
WILD, C.P., SCALBERT, A. and HERCEQ, Z. (2013). “Measuring the exposome: A powerful basis for
evaluating environmental exposures and cancer risk,” Environ. Mol. Mutagen 54(7), 480–499.
3082
WILDING, C.S., CURWEN, G.B., TAWN, E.J., SHENG, X., WINTHER, J.F., CHAKRABORTY, R. and
3083
BOICE, J.D., JR. (2007). “Influence of polymorphisms at loci encoding DNA repair proteins on G2
3084
chromosomal radiosensitivity,” Environ. Mol. Mutagen 48(1), 48–57.
102
NCRP SC 1-21
Draft of March 31, 2015 [MR]
3085
NOT TO BE DISSEMINATED OR REFERENCED
WILSON, P.F., NAGASAWA, H., FITZEK, M.M., LITTLE, J.B. and BEDFORD, J.S. (2010). “G2-phase
3086
chromosomal radiosensitivity of primary fibroblasts from hereditary retinoblastoma family members and
3087
some apparently normal controls,” Radiat. Res. 173(1), 62–70.
3088
3089
3090
WINTHER, J.F. and OLSEN, J.H. (2012). “Does cancer treatment in childhood induce transgenerational
genetic damage?,” J. Clin. Oncol. 30(3), 225–226.
WINTHER, J.F., OLSEN, J.H., WU, H., SHYR, Y., MULVIHILL, J.J., STOVALL, M., NIELSE, A.,
3091
SCHMIEGELOW, M. and BOICE, J.D., JR. (2012). “Genetic disease in the children of Danish survivors
3092
of childhood and adolescent cancer,” J. Clin. Oncol. 30(1), 27–33.
3093
3094
WINTON, D.J., PEACOCK, J.H. and PONDER, B.A. (1989). “Effect of gamma radiation at high- and lowdose rate on a novel in vivo mutation assay in mouse intestine,” Mutagenesis 4(5), 404–406.
3095
WONG, F.L., BOICE, J.D., JR., ABRAMSON, D.H., TARONE, R.E., KLEINERMAN, R.A., STOVALL,
3096
M., GOLDMAN, M.B., SEDDON, J.M., TARBELL, N., FRAUMENI, J.F., JR. and LI, F.P. (1997).
3097
“Cancer incidence after retinoblastoma. Radiation dose and sarcoma risk,” JAMA 278(15), 1262–1267.
3098
3099
3100
YAMADA, M., WONG, F.L., FUJIWARA, S., AKAHOSHI, M. and SUZUKI, G. (2004). “Noncancer
disease incidence in atomic bomb survivors, 1958–1998,” Radiat. Res. 161(6), 622–632.
YU, Y., OKAYASU, R., WEIL, M.M., SILVER, A., MCCARTHY, M., ZABRISKIE, R., LONG, S., COX,
3101
R. and ULLRICH, R.L. (2001). “Elevated breast cancer risk in irradiated BALB/c mice associates with
3102
unique functional polymorphism of the Prkdc (DNA-dependent protein kinase catalytic subunit) gene,”
3103
Cancer Res. 61(5), 1820–1824.
3104
YU, C.L., TUCKER, M.A., ABRAMSON, D.H., FURUKAWA, K., SEDDON, J.M., STOVALL, M.,
3105
FRAUMENI, J.F., JR. and KLEINERMAN, R.A. (2009). “Cause-specific mortality in long-term
3106
survivors of retinoblastoma,” J. Natl. Cancer Inst. 101(8), 581–591.
3107
ZABLOTSKA, L.B., BAZYKA, D., LUBIN, J.H., GUDZENKO, N., LITTLE, M.P., HATCH, M., FINCH,
3108
S., DYAGIL, I., REISS, R.F., CHUMAK, V.V., BOUVILLE, A., DROZDOVITCH, V., KRYUCHKOV,
3109
V.P., GOLOVANOV, I., BAKHANOVA, E., BABKINA, N., LUBARETS, T., BEBESHKO, V.,
3110
ROMANENKO, A. and MABUCHI, K. (2013). “Radiation and the risk of chronic lymphocytic and other
3111
leukemias among Chornobyl cleanup workers,” Environ. Health Perspect. 121(1), 59–65.
3112
3113
3114
ZEIDAN, Y.H. and HANNUN, Y.A. (2010). “The acid sphingomyelinase/ceramide pathway: biomedical
significance and mechanisms of regulation,” Curr. Mol. Med. 10(5), 454–466.
ZHANG, Y., ZHOU, J., BALDWIN, J., HELD, K.D., PRISE, K.M., REDMOND, R.W. and LIBER, H.L.
3115
(2009). “Ionizing radiation-induced bystander mutagenesis and adaptation: quantitative and temporal
3116
aspects,” Mutat. Res. 671(1–2), 20–25.
103
NCRP SC 1-21
Draft of March 31, 2015 [MR]
NOT TO BE DISSEMINATED OR REFERENCED
3117
ZHANG, Q., MATZKE, M., SCHEPMOES, A.A., MOORE, R.J., WEBB-ROBERTSON, B.J., HU, Z.,
3118
MONROE, M.E., QIAN, W.J., SMITH, R.D. and MORGAN, W.F. (2014). “High and low doses of
3119
ionizing radiation induce different secretome profiles in a human skin model,” PLoS One 9(3), 92332.
3120
ZHOU, H., RANDERS-PEHRSON, G., WALDREN, C.A., VANNAIS, D., HALL, E.J. and HEI, T.K.
3121
(2000). “Induction of a bystander mutagenic effect of alpha particles in mammalian cells,” Proc. Natl.
3122
Acad. Sci USA 97(5), 2099–2104.
104