ArticlePDF AvailableLiterature Review

Assessing the Carcinogenic Potential of Low Dose Exposures to Chemical Mixtures in the Environment: The Challenge Ahead

Authors:

Abstract and Figures

Lifestyle factors are responsible for a considerable portion of cancer incidence worldwide, but credible estimates from the World Health Organization and the International Agency for Research on Cancer (IARC) suggest that the fraction of cancers attributable to toxic environmental exposures is between 7% and 19%. To explore the hypothesis that low-dose exposures to mixtures of chemicals in the environment may be combining to contribute to environmental carcinogenesis, we reviewed 11 hallmark phenotypes of cancer, multiple priority target sites for disruption in each area and prototypical chemical disruptors for all targets, this included dose-response characterizations, evidence of low-dose effects and cross-hallmark effects for all targets and chemicals. In total, 85 examples of chemicals were reviewed for actions on key pathways/mechanisms related to carcinogenesis. Only 15% (13/85) were found to have evidence of a dose-response threshold, whereas 59% (50/85) exerted low-dose effects. No dose-response information was found for the remaining 26% (22/85). Our analysis suggests that the cumulative effects of individual (non-carcinogenic) chemicals acting on different pathways, and a variety of related systems, organs, tissues and cells could plausibly conspire to produce carcinogenic synergies. Additional basic research on carcinogenesis and research focused on low-dose effects of chemical mixtures needs to be rigorously pursued before the merits of this hypothesis can be further advanced. However, the structure of the World Health Organization International Programme on Chemical Safety 'Mode of Action' framework should be revisited as it has inherent weaknesses that are not fully aligned with our current understanding of cancer biology. © The Author 2015. Published by Oxford University Press.
Content may be subject to copyright.
oxford
print ISSN: 0143-3334
online ISSN: 1460-2180
Scan to view this journal
on your mobile device
Volume 36 Supplement 1 June 2015
www.carcin.oxfordjournals.org
Assessing the Carcinogenic Potential of Low-Dose Exposures
to Chemical Mixtures in the Environment: The Challenge Ahead
Volume 36 Supplement 1 June 2015
Carcinogenesis Integrative Cancer Research
CARCINOGENESIS SUPPLEMENT
Assessing the Carcinogenicity of Low-Dose
Exposures to Chemical Mixtures in the
Environment: The Challenge Ahead
Editor-in-Chief
Curtis C. Harris
We gratefully acknowledge the support of the National Institute of Health (NIH)-National Institute of
Environmental Health Sciences (NIEHS) conference grant travel support [R13ES023276]
EDITORIAL
Cause and Prevention of Human Cancer
Curtis C Harris ..............................................................................S1
REVIEWS
Mechanisms of environmental chemicals that enable the cancer hallmark of evasion of growth
suppression
RitaNahta, FahdAl-Mulla, RabeahAl-Temaimi, AmedeoAmedei, RafaelaAndrade-Vieira, SarahBay,
DustinG.Brown, GloriaM.Calaf, RobertC.Castellino, KarineA.Cohen-Solal, AnnamariaColacci,
NicholaCruickshanks, PaulDent, RiccardoDiFiore, StefanoForte, GaryS.Goldberg, RoslidaA.Hamid,
HariniKrishnan, DaleW.Laird, AhmedLasfar, PaolaA.Marignani, LorenzoMemeo, ChiaraMondello,
ChristianC.Naus, RichardPonce-Cusi, JayadevRaju, DebasishRoy, RabindraRoy, ElizabethP.Ryan,
HosniK.Salem, A.IvanaScovassi, NeetuSingh, MonicaVaccari, RenzaVento, JanVondráček, MarkWade,
JordanWoodrick, and WilliamH.Bisson...........................................................S2
Disruptive chemicals, senescence and immortality
AmancioCarnero, CarmenBlanco-Aparicio, HiroshiKondoh, MatildeE.Lleonart, JuanFernando
Martinez-Leal, ChiaraMondello, A.IvanaScovassi, WilliamH.Bisson, AmedeoAmedei, RabindraRoy,
JordanWoodrick, AnnamariaColacci, MonicaVaccari, JayadevRaju, FahdAl-Mulla, RabeahAl-Temaimi,
HosniK.Salem, LorenzoMemeo, StefanoForte, NeetuSingh, RoslidaA.Hamid, ElizabethP.Ryan,
DustinG.Brown, JohnPierceWiseSr, SandraS.Wise, and HemadYasaei ...............................S19
The potential for chemical mixtures from the environment to enable the cancer hallmark of
sustained proliferative signalling
WilhelmEngström, PhilippaDarbre, StaffanEriksson, LindaGulliver, ToveHultman, MichalisV.Karamouzis,
JamesE.Klaunig, RekhaMehta, KimMoorwood, ThomasSanderson, HidekoSone, PankajVadgama,
GerardWagemaker, AndrewWard, NeetuSingh, FahdAl-Mulla, RabeahAl-Temaimi, AmedeoAmedei,
AnnaMariaColacci, MonicaVaccari, ChiaraMondello, A.IvanaScovassi, JayadevRaju, RoslidaA.Hamid,
LorenzoMemeo, StefanoForte, RabindraRoy, JordanWoodrick, HosniK.Salem, ElizabethRyan,
DustinG.Brown, and WilliamH.Bisson .........................................................S38
Causes of genome instability: the effect of low dose chemical exposures in modern society
SabineA.S.Langie, GudrunKoppen, DanielDesaulniers, FahdAl-Mulla, RabeahAl-Temaimi, AmedeoAmedei,
AmayaAzqueta, WilliamH.Bisson, DustinBrown, GunnarBrunborg, AmeliaK.Charles, TaoChen,
AnnamariaColacci, FirouzDarroudi, StefanoForte, LaetitiaGonzalez, RoslidaA.Hamid, LisbethE.Knudsen,
LucLeyns, AdelaLopezdeCerainSalsamendi, LorenzoMemeo, ChiaraMondello, CarmelMothersill,
Ann-KarinOlsen, SoaPavanello, JayadevRaju, EmilioRojas, RabindraRoy, ElizabethRyan, Patricia
Ostrosky-Wegman, HosniK.Salem, IvanaScovassi, NeetuSingh, MonicaVaccari, FrederikJ.VanSchooten,
MaharaValverde, JordanWoodrick, LuopingZhang, NikvanLarebeke, MichelineKirsch-Volders,
and AndrewR.Collins........................................................................S61
Disruptive environmental chemicals and cellular mechanisms that confer resistance to celldeath
KannanBadriNarayanan, ManafAli, BarryJ.Barclay, QiangCheng, LeandroD’Abronzo, Rita
Dornetshuber-Fleiss, ParamitaM.Ghosh, MichaelJ.GonzalezGuzman, Tae-JinLee, PoSingLeung, LinLi,
SuidjitLuanpitpong, EdwardRatovitski, YonRojanasakul, MariaFiammettaRomano, SimonaRomano,
RanjeetKumarSinha, ClementYedjou, FahdAl-Mulla, RabeahAl-Temaimi, AmedeoAmedei, DustinG.
Brown, ElizabethP.Ryan, AnnamariaColacci, RoslidaA.Hamid, ChiaraMondello, JayadevRaju, HosniK.Salem,
JordanWoodrick, IvanaScovassi, NeetuSingh, MonicaVaccari, RabindraRoy, StefanoForte, LorenzoMemeo,
SeoYunKim, WilliamH.Bisson, LeroyLowe, and HyunHoPark.......................................S89
JOURNAL OF CARCINOGENESIS
VOLUME 36, SUPPLEMENT 1
MAY 2015
Contents
Chemical compounds from anthropogenic environment and immune evasion
mechanisms: potential interactions
JuliaKravchenko, EmanuelaCorsini, MarcA.Williams, WilliamDecker, MasoudH.Manjili, TakemiOtsuki,
NeetuSingh, FahaAl-Mulla, RabeahAl-Temaimi, AmedeoAmedei, AnnaMariaColacci, MonicaVaccari,
ChiaraMondello, A.IvanaScovassi, JayadevRaju, RoslidaA.Hamid, LorenzoMemeo, StefanoForte,
RabindraRoy, JordanWoodrick, HosniK.Salem, ElizabethP.Ryan, DustinG.Brown, WilliamH.Bisson,
LeroyLowe, and H.KimLyerly ................................................................S111
The impact of low-dose carcinogens and environmental disruptors on tissue invasion and metastasis
JosiahOchieng, GladysN.Nangami, OlugbemigaOgunkua, IsabelleR.Miousse, IgorKoturbash, ValerieOdero-
Marah, LisaMcCawley, PratimaNangia-Makker, NuzhatAhmed, YunusLuqmani, ZhenbangChen,
SilvanaPapagerakis, GregoryT.Wolf, ChenfangDong, BinhuaP.Zhou, DustinG.Brown, AnnamariaColacci,
RoslidaA.Hamid, ChiaraMondello, JayadevRaju, ElizabethP.Ryan, JordanWoodrick, IvanaScovassi,
NeetuSingh, MonicaVaccari, RabindraRoy, StefanoForte, LorenzoMemeo, HosniK.Salem, AmedeoAmedei,
RabeahAl-Temaimi, FahdAl-Mulla, WilliamH.Bisson, and SakinaE.Eltom............................S128
The effect of environmental chemicals on the tumor microenvironment
StephanieC.Casey, MonicaVaccari, FahdAl-Mulla, RabeahAl-Temaimi, AmedeoAmedei, MaryHelen
Barcellos-Hoff, DustinG.Brown, MarionChapellier, JosephChristopher, ColleenCurran, StefanoForte,
RoslidaA.Hamid, PetrHeneberg, DanielC.Koch, P.K.Krishnakumar, EzioLaconi, VeroniqueMaguer-Satta,
FabioMarongiu, LorenzoMemeo, ChiaraMondello, JayadevRaju, JesseRoman, RabindraRoy, ElizabethP.Ryan,
SandraRyeom, HosniK.Salem, A.IvanaScovassi, NeetuSingh, LauraSoucek, LouisVermeulen, JonathanR.
Whiteld, JordanWoodrick, AnnamariaColacci, WilliamH.Bisson, and DeanW.Felsher...................S160
Assessing the carcinogenic potential of low-dose exposures to chemical mixtures in the
environment: focus on the cancer hallmark of tumor angiogenesis
ZhiweiHu, SamiraA.Brooks, ValérianDormoy, Chia-WenHsu, Hsue-YinHsu, Liang-TzungLin,
ThierryMassfelder, W.KimrynRathmell, MenghangXia, FahdAl-Mulla, RabeahAl-Temaimi,
AmedeoAmedei, DustinG.Brown, KalanR.Prudhomme, AnnamariaColacci, RoslidaA.Hamid,
ChiaraMondello, JayadevRaju, ElizabethP.Ryan, JordanWoodrick, A.IvanaScovassi, NeetuSingh,
MonicaVaccari, RabindraRoy, StefanoForte, LorenzoMemeo, HosniK.Salem, LeroyLowe, LasseJensen,
WilliamH.Bisson, and NicoleKleinstreuer.......................................................S184
Metabolic reprogramming and dysregulated metabolism: cause, consequence and/or enabler of
environmental carcinogenesis?
R.BrooksRobey, JudithWeisz, NancyKuemmerle, AnnaC.Salzberg, ArthurBerg, DustinBrown, LauraKubik,
RobertaPalorini, FahdAl-Mulla, RabeahAl-Temaimi, AnnamariaColacci, ChiaraMondello, JayadevRaju,
JordanWoodrick, IvanaScovassi, NeetuSingh, MonicaVaccari, RabindraRoy, StefanoForte, LorenzoMemeo,
HosniK.Salem, AmedeoAmedei, RoslidaA.Hamid, GraemeP.Williams, LeroyLowe, JoelMeyer,
FrancisL.Martin, WilliamH.Bisson, FerdinandoChiaradonna, and ElizabethP.Ryan ....................S203
Environmental immune disruptors, inammation and cancerrisk
PatriciaA.Thompson, MahinKhatami, CarolynJ.Baglole, JunSun, ShelleyHarris, Eun-YiMoon, FahdAl-Mulla,
RabeahAl-Temaimi, DustinBrown, AnnamariaColacci, ChiaraMondello, JayadevRaju, ElizabethRyan,
JordanWoodrick, IvanaScovassi, NeetuSingh, MonicaVaccari, RabindraRoy, StefanoForte, LorenzoMemeo,
HosniK.Salem, AmedeoAmedei, RoslidaA.Hamid, LeroyLowe, and WilliamH.Bisson ....................S232
Assessing the carcinogenic potential of low-dose exposures to chemical mixtures in the environment:
the challengeahead
WilliamH.GoodsonIII, LeroyLowe, DavidO.Carpenter, MichaelGilbertson, AbdulManafAli,
AdelaLopezdeCerainSalsamendi, AhmedLasfar, AmancioCarnero, AmayaAzqueta, AmedeoAmedei, AmeliaK.
Charles, AndrewR.Collins, AndrewWard, AnnaC.Salzberg, AnnamariaColacci, Ann-KarinOlsen, ArthurBerg,
BarryJ.Barclay, BinhuaP.Zhou, CarmenBlanco-Aparicio, CarolynBaglole, ChenfangDong, ChiaraMondello,
Chia-WenHsu, ChristianC.Naus, ClementYedjou, ColleenS.Curran, DaleW.Laird, DanielC.Koch, DanielleJ.
Carlin, DeanW.Felsher, DebasishRoy, DustinBrown, EdwardRatovitski, ElizabethP.Ryan, EmanuelaCorsini,
EmilioRojas, Eun-YiMoon, EzioLaconi, FabioMarongiu, FahdAl-Mulla, FerdinandoChiaradonna, FirouzDarroudi,
FrancisL.Martin, FrederikJ.VanSchooten, GaryS.Goldberg, GerardWagemaker, GladysNangami, GloriaM.
Calaf, GraemeWilliams, GregoryT.Wolf, GudrunKoppen, GunnarBrunborg, H.KimLyerly, HariniKrishnan,
HasiahAbHamid, HemadYasaei, HidekoSone, HiroshiKondoh, HosniK.Salem, Hsue-YinHsu, HyunHoPark,
IgorKoturbash, IsabelleR.Miousse, IvanaScovassi, JamesEKlaunig, JanVondráček, JayadevRaju, JesseRoman,
JohnPierceWiseSr., JonathanR.Whiteld, JordanWoodrick, JosephChristopher, JosiahOchieng,
JuanFernandoMartinez-Leal, JudithWeisz, JuliaKravchenko, JunSun, KalanR.Prudhomme,
KannanBadriNarayanan, KarineA.Cohen-Solal, KimMoorwood, LaetitiaGonzalez, LauraSoucek,
LeJian, LeandroS.D’Abronzo, Liang-TzungLin, LinLi, LindaGulliver, LisaJ.McCawley,
LorenzoMemeo, LouisVermeulen, LucLeyns, LuopingZhang, MaharaValverde, MahinKhatami,
MariaFiammettaRomano, MarionChapellier, MarcA.Williams, MarkWade, MasoudH.Manjili,
MatildeLleonart, MenghangXia, MichaelJGonzalez, MichalisV.Karamouzis, MichelineKirsch-
Volders, MonicaVaccari, NancyB.Kuemmerle, NeetuSingh, NicholaCruickshanks, NicoleKleinstreuer,
NikvanLarebeke, NuzhatAhmed, OlugbemigaOgunkua, P.K.Krishnakumar, PankajVadgama,
PaolaA.Marignani, ParamitaM.Ghosh, PatriciaOstrosky-Wegman, PatriciaThompson, PaulDent,
PetrHeneberg, PhilippaDarbre, PoSingLeung, PratimaNangia-Makker, Qiang(Shawn)Cheng,
R.BrooksRobey, RabeahAl-Temaimi, RabindraRoy, RafaelaAndrade-Vieira, RanjeetK.Sinha,
RekhaMehta, RenzaVento, RiccardoDiFiore, RichardPonce-Cusi, RitaDornetshuber-Fleiss,
RitaNahta, RobertC.Castellino, RobertaPalorini, RoslidaAbdHamid, SabineA.S.Langie,
SakinaEltom, SamiraA.Brooks, SandraRyeom, SandraS.Wise, SarahN.Bay, ShelleyHarris,
SilvanaPapagerakis, SimonaRomano, SoaPavanello, StaffanEriksson, StefanoForte,
StephanieC.Casey, SudjitLuanpitpong, Tae-JinLee, TakemiOtsuki, TaoChen, ThierryMassfelder,
ThomasSanderson, TizianaGuarnieri, ToveHultman, ValérianDormoy, ValerieOdero-Marah,
VenkataSabbisetti, VeroniqueMaguer-Satta, W.KimrynRathmell, WilhelmEngström, WilliamK.
Decker, WilliamH.Bisson, YonRojanasakul, YunusLuqmani, ZhenbangChen, and ZhiweiHu ..............S254
SUBSCRIPTIONS
A subscription to Carcinogenesis comprises 12 issues. All prices include postage,
and for subscribers outside the UK delivery is by Standard Air. Advance Access
contains papers that have recently been accepted, but have not yet been included
within an issue. Advance Access is updated weekly.
Annual subscription rate (Volume 36, 12 issues, 2015)
Institutional
Print edition and site-wide online access: £1608.00/$3216.00/€2412.00
Print edition only: £1479.00/$2959.00/€2219.00
Site-wide online access only: £1247.00/$2494.00/€1870.00
Personal
Print edition and individual online access: £564.00/$1128.00/€846.00
Please note: US$ rate applies to US & Canada, Euro (€) applies to Europe, UK£
applies to UK and Rest of World
There may by other subscription rates available, for a complete listing please
visit www.carcin.oupjournals.org/subinfo/.
Full prepayment in the correct currency is required for all orders. Payment should
be in US dollars for orders being delivered to the USA or Canada; Euros for orders
being delivered within Europe (excluding the UK); GBP Sterling for orders being
delivered elsewhere (i.e., not being delivered to USA, Canada, or Europe). All
orders should be accompanied by full payment and sent to your nearest Oxford
Journals office. Subscriptions are accepted for complete volumes only. Orders are
regarded as firm, and payments are not refundable. Our prices include dispatch by
Standard Air. Claims must be notified within four months of dispatch/order date
(whichever is later). Subscriptions in the EEC may be subject to European VAT. If
registered, please supply details to avoid unnecessary charges. For subscriptions that
include online versions, a proportion of the subscription price may be subject to UK
VAT. Subscribers in Canada, please add GST to the prices quoted. Personal rate
subscriptions are only available if payment is made by personal cheque or credit
card, delivery is to a private address, and is for personal use only.
The current year and two previous years’ issues are available from Oxford
University Press. Previous volumes can be obtained from the Periodicals Service
Company, 11 Main Street, Germantown, NY 12526, USA. Email:
psc@periodicals.com. Tel: +1 (518) 537 4700. Fax: +1 (518) 537 5899.
Contact information
Journals Customer Service Department, Oxford Journals, Great Clarendon Street,
Oxford OX2 6DP, UK. E-mail: jnls.cust.serv@oup.com. Tel: +44 (0)1865 353907.
Fax: + 44 (0)1865 353485. In the Americas, please contact: Journals Customer
Servie Department, Oxford Journals, 2001 Evans Road, Cary, NC 27513, USA.
E-mail: jnlorders@oup.com. Tel: (800) 852-7323 (toll-free in USA/Canada) or
(919) 677-0977. Fax: (919) 677-1714. In Japan, please contact: Journals Customer
Service Department, Oxford Journals, 4-5-10-8F Shiba, Minato-ku, Tokyo,
108-8386, Japan. E-mail: custserv.jp@oup.com. Tel: (81) 3 5444 5858.
Fax: (81) 3 3454 2929.
Methods of payment. Payment should be made: by cheque (to Oxford University
Press, Cashiers Office, Great Clarendon Street, Oxford, OX2 6DP, UK); by
bank transfer [to Barclays Bank Plc, Oxford Office, Oxford (bank sort code
20-65-18) (UK); overseas only Swift code BARC GB22 (GB£ Sterling
Account no. 70299332, IBAN GB89BARC20651870299332; US$ Dollars
Account no. 66014600, IBAN GB27BARC20651866014600; EU€EURO
Account no. 78923655, IBAN GB16BARC20651878923655]; or by credit card
(Mastercard, Visa, Switch or American Express).
Carcinogenesis is published monthly (ISSN 0143-3334) by Oxford University
Press, Oxford, UK.
Carcinogenesis is distributed in the USA by, Mercury Media Processing LLC,
1850 Elizabeth Ave., Suite #C, Rahway, NJ 07065. Periodicals Postage paid at
Rahway, NJ and at additional entry points.
US Postmaster: send address changes to Carcinogenesis, c/o Mercury Media
Processing 1634 E. Elizabeth Ave, Linden, NJ 07036
Environmental and ethical policies
Oxford Journals, a division of Oxford University Press, is committed to working
with the global community to bring the highest quality research to the widest
possible audience. Oxford Journals will protect the environment by implementing
environmentally friendly policies and practices wherever possible. Please
see http://www.oxfordjournals.org/ethicalpolicies.html for further information on
environmental and ethical policies.
Articles in this journal are now assigned a Digital Object Identifier (DOI).
A DOI is a unique identifier of an article and can be used either to link to the
online version of the article, or as additional information in a citation. Unlike a
standard URL, the DOI of an article never changes, and so will work in
perpetuity. To use a DOI as a link to the online version of an article, prefix the
DOI with http://dx.doi.org/ or go to the dx.doi.org website and enter the DOI.
To include the DOI in a citation, insert the DOI at the end of the reference,
prefixed with ‘‘doi:’’.
Permissions
For information on how to request permissions to reproduce articles/
information from this journal, please visit www.oxfordjournals.org/permissions.
Advertising
Advertising, inserts and artwork enquiries should be addressed to Advertising and
Special Sales, Oxford Journals, Oxford University Press, Great Clarendon Street,
Oxford, OX2 6DP, UK. Tel: +44 (0)1865 354767; Fax +44 (0)1865 353774; E-mail:
jnlsadvertising@oup.com.
Disclaimer
Statements of fact and opinion in the articles in Carcinogenesis are those of the
respective authors and contributors and not of Carcinogenesis or Oxford
University Press. Neither Oxford University Press nor Carcinogenesis make any
representation, express or implied, in respect of the accuracy of the material
in this journal and cannot accept any legal responsibility or liability for any
errors or omissions that may be made. The reader should make his/her own
evaluation as to the appropriateness or otherwise of any experimental technique
described.
© Oxford University Press 2015.
All rights reserved; no part of this publication may be reproduced, stored in a
retrieval system, or transmitted in any form or by any means, electronic,
mechanical, photocopying, recording, or otherwise without
prior written permission of the Publishers, or a licence permitting restricted
copying issued in the UK by the Copyright Licensing Agency Ltd,
90 Tottenham Court Road, London W1P 9HE, or in the USA by the Copyright
Clearance Center, 222 Rosewood Drive, Danvers, MA 01923.
Typeset by Newgen Knowledge Works Pvt. Ltd, Chennai, India.
Printed in China by CTPS Digiprints.
Theo Colborn (March 28, 1927 - December 14, 2014)
We dedicate this special issue on the challenges associated with assessing the carcinogenic potential of low dose exposures to
chemical mixtures in the environment, to the memory of Dr Theodora (Theo) Colborn. Theo was a pioneer in the science of the
effects of low dose exposures to environmental chemicals and for the past twenty-ve years, was instrumental in the develop-
ment and integration of the eld of endocrine disruption. Theo introduced us to one another about four-years-ago which led to the
founding of Getting to Know Cancer, and ultimately the launch of the Halifax Project (which has been a tremendously productive
collaboration for the integration of cancer biology and environmental toxicology). So we want to thank her for her legacy of work in
this area, her inuence on our research, and her encouragement.
Theo was well known internationally for her tireless commitment to the protection of public health, but not everyone knew that
she was also a tremendously generous and insightful scientist who assembled researchers from a variety of specialties and allowed
them to discover for themselves what she had understood about the inuences of low-dose exposures to certain environmental
chemicals on embryonic and fetal development. Indeed, she nurtured cross-disciplinary collaboration and it was that collegiality
and spirit of sharing that produced seminal insights that opened up the entire eld of endocrine disruption. So we have attempted
to use a similar approach to help us understand the importance of ongoing low dose exposures to mixtures of chemicals in the
environment and their relevance for cancer and carcinogenesis. In other words, this is truly an extension of her work, and we want
to pay tribute and offer thanks for her wisdom, her generosity and her legacy.
Leroy Lowe and Michael Gilbertson,
Cofounders, Getting To Know Cancer
June, 2015
© The Author 2015. Published by Oxford University Press. All rights reserved. For Permissions, please email: journals.permissions@oup.com
Carcinogenesis, 2015, Vol. 36, Supplement 1, S1
doi: 10.1093/carcin/bgv047
Editorial
S1
Copyedited by: <CE_Initials>
Editorial
Cause and Prevention of Human Cancer
The relative contribution of the environment, genetic suscepti-
bility and DNA replication errors to cancer causation has been
a longstanding area of investigation in the elds of molecular
epidemiology of cancer and carcinogenesis. A recent report by
Tomasetti etal. (1) attributing DNA replication errors within stem
cells and ‘bad luck’ as a major cause of a select group of cancers
has stirred debate within community of cancer researchers espe-
cially those in cancer prevention (2–7). Tomasetti et al. (8) have
also written a balanced response to many of these concerns.
Carcinogenesis is joining this debate by publishing in this
issue a series of reviews on the carcinogenic potential of expo-
sure to low doses and mixtures of chemicals. The reviews utilize
a framework of the Hallmarks of Cancer (9) and are the prod-
uct of the Halifax Project Task Force initiated by Leroy Lowe
and Michael Gilbertson. They engaged international teams with
input of nearly 200 cancer biologists and toxicologist to review
the literature in each of the 11 Hallmarks of Cancer. The reviews
are multiauthored, condensed by a peer review and extensively
referenced. The primary recommendation is a research and
regulatory strategy using the Hallmarks of Cancer framework to
identify priority mixtures of chemicals, i.e. ‘….those with sub-
stantial carcinogenic relevance’, for future investigations ‘…. to
inform risk assessment practices worldwide’ (10).
Carcinogenesis will also publish a review of cancer prevention
this summer, which will be written by Christopher Wild, Director
of the International Agency on Research of Cancer.
References
1. Tomasetti, C. etal. (2015) Cancer etiology. Variation in cancer
risk among tissues can be explained by the number of stem
cell divisions. Science, 347, 78–81.
2. Crossan, G.P. etal. (2015) Do mutational dynamics in stem cells
explain the origin of common cancers? Cell Stem Cell, 16, 111–112.
3. O’Callaghan, M. (2015) Cancer risk: accuracy of literature. Sci-
ence, 347, 729.
4. Ashford, N.A. et al. (2015) Cancer risk: role of environment.
Science, 347, 727.
5. Song, M. etal. (2015) Cancer risk: many factors contribute. Sci-
ence, 347, 728–729.
6. Potter, J.D. etal. (2015) Cancer risk: tumors excluded. Science,
347, 727.
7. Wild, C. etal. (2015) Cancer risk: role of chance overstated. Sci-
ence, 347, 728.
8. Tomasetti, C. etal. (2015) Cancer risk: role of environment—
response. Science, 347, 729–731.
9. Hanahan, D. etal. (2000) The hallmarks of cancer. Cell, 100,
57–70.
10. Goodson, W.H. III, etal. (2015) Assessing the carcinogenesis
potential of low dose exposures to chemical mixtures in the
environment: the challenge ahead. Carcinogenesis.
Curtis CHarris,
Editor-in-Chief
Received: August 7, 2014; Revised: January 23, 2015; Accepted: January 31, 2015
© The Author 2015. Published by Oxford University Press.
Carcinogenesis, 2015, Vol. 36, Supplement 1, S254–S296
doi:10.1093/carcin/bgv039
Review
S254
This is an Open Access article distributed under the terms of the Creative Commons Attribution Non-Commercial License
(http://creativecommons.org/licenses/by-nc/4.0/), which permits non-commercial re-use, distribution, and reproduction in any
medium, provided the original work is properly cited. For commercial re-use, please contact journals.permissions@oup.com

Assessing the carcinogenic potential of low-dose
exposures to chemical mixtures in the environment: the
challengeahead
William H.Goodson III*, LeroyLowe1,2, David O.Carpenter3,
MichaelGilbertson4, AbdulManaf Ali5,
AdelaLopez de Cerain Salsamendi6, AhmedLasfar7,
AmancioCarnero8, AmayaAzqueta6, AmedeoAmedei9,
Amelia K.Charles10, Andrew R.Collins11, AndrewWard12,
Anna C.Salzberg13, AnnamariaColacci14, Ann-KarinOlsen15,
ArthurBerg13, Barry J.Barclay16, Binhua P.Zhou17,
CarmenBlanco-Aparicio18, Carolyn J.Baglole19, ChenfangDong17,
ChiaraMondello20, Chia-WenHsu21, Christian C.Naus22,
ClementYedjou23, Colleen S.Curran24, Dale W.Laird25, Daniel C.Koch26,
Danielle J.Carlin27, Dean W.Felsher28, DebasishRoy29,
Dustin G.Brown30, EdwardRatovitski31, Elizabeth P.Ryan30,
EmanuelaCorsini32, EmilioRojas33, Eun-YiMoon34, EzioLaconi35,
FabioMarongiu35, FahdAl-Mulla36,
FerdinandoChiaradonna37,38, FirouzDarroudi39, Francis L.Martin2,
Frederik J.Van Schooten40, Gary S.Goldberg41,
GerardWagemaker42, GladysNangami43, Gloria M.Calaf44,45,
GraemeWilliams46, Gregory T.Wolf47, GudrunKoppen48,
GunnarBrunborg15, H.Kim Lyerly49, HariniKrishnan41,
HasiahAb Hamid50, HemadYasaei51, HidekoSone52,
HiroshiKondoh53, Hosni K.Salem54, Hsue-YinHsu55,
Hyun HoPark56, IgorKoturbash57, Isabelle R.Miousse57, A.IvanaScovassi20,
James E.Klaunig58, JanVondráček59, JayadevRaju60, JesseRoman61,62,
John PierceWise Sr.63, Jonathan R.Whiteld64,
JordanWoodrick65, Joseph A.Christopher66, JosiahOchieng43,
Juan FernandoMartinez-Leal67, JudithWeisz68, JuliaKravchenko49,
JunSun69, Kalan R.Prudhomme70, Kannan BadriNarayanan56,
Karine A.Cohen-Solal71, KimMoorwood12, LaetitiaGonzalez72,
LauraSoucek64,73, LeJian74,75, Leandro S.D’Abronzo76, Liang-TzungLin77,
W.H.Goodson et al. | S255
Carcinogenesis, 2015, Vol. 36, Supplement 1, S254–S296
doi:10.1093/carcin/bgv039
Review
LinLi78, LindaGulliver79, Lisa J.McCawley80, LorenzoMemeo81,
LouisVermeulen82, LucLeyns72, LuopingZhang83,
MaharaValverde33, MahinKhatami84,
Maria FiammettaRomano85, MarionChapellier86, Marc A.Williams87,
MarkWade88, Masoud H.Manjili89, MatildeLleonart90, MenghangXia21,
Michael J.Gonzalez91, Michalis V.Karamouzis92,
MichelineKirsch-Volders72, MonicaVaccari14, Nancy B.Kuemmerle93,94,
NeetuSingh95, NicholaCruickshanks96, NicoleKleinstreuer97,
Nikvan Larebeke98, NuzhatAhmed99, OlugbemigaOgunkua43,
P.K.Krishnakumar100, PankajVadgama101, Paola A.Marignani102,
Paramita M.Ghosh76, PatriciaOstrosky-Wegman33, PatriciaThompson103,
PaulDent96, PetrHeneberg104, PhilippaDarbre105, PoSing Leung78,
PratimaNangia-Makker106, Qiang (Shawn)Cheng107, R.BrooksRobey93,94,
RabeahAl-Temaimi108, RabindraRoy65, RafaelaAndrade-Vieira102,
Ranjeet K.Sinha109, RekhaMehta60, RenzaVento110,111, RiccardoDi Fiore110,
RichardPonce-Cusi45, RitaDornetshuber-Fleiss112,113, RitaNahta114,
Robert C.Castellino115,116, RobertaPalorini37,38, RoslidaAbd Hamid50,
Sabine A.S.Langie48, SakinaEltom43, Samira A.Brooks117, SandraRyeom118,
Sandra S.Wise63, Sarah N.Bay119, Shelley A.Harris120,121,
SilvanaPapagerakis47, SimonaRomano85, SoaPavanello122,
StaffanEriksson123, StefanoForte81, Stephanie C.Casey26,
SudjitLuanpitpong124, Tae-JinLee125, TakemiOtsuki126, TaoChen127,
ThierryMassfelder128, ThomasSanderson129, TizianaGuarnieri130,131,132,
Tove Hultman133, ValérianDormoy128,134, ValerieOdero-Marah135,
VenkataSabbisetti136, VeroniqueMaguer-Satta87,
W.KimrynRathmell117, WilhelmEngström137, William K.Decker138,
William H.Bisson70, YonRojanasakul139, YunusLuqmani140,
ZhenbangChen43 and ZhiweiHu141
California Pacic Medical Center Research Institute, 2100 Webster Street #401, San Francisco, CA 94115, USA, 1Getting to
Know Cancer, Room 229A, 36 Arthur Street, Truro, Nova Scotia B2N 1X5, Canada, 2Lancaster Environment Centre, Lancaster
University, Bailrigg, Lancaster LA1 4AP, UK, 3Institute for Health and the Environment, University at Albany, 5 University
Pl., Rensselaer, NY 12144, USA, 4Getting to Know Cancer, Guelph N1G 1E4, Canada, 5School of Biotechnology, Faculty of
Agriculture Biotechnology and Food Sciences, Sultan Zainal Abidin University, Tembila Campus, 22200 Besut, Terengganu,
Malaysia, 6Department of Pharmacology and Toxicology, Faculty of Pharmacy, University of Navarra, Pamplona 31008,
Spain, 7Department of Pharmacology and Toxicology, Ernest Mario School of Pharmacy, Rutgers, State University of New
Jersey, Piscataway, NJ 08854, USA, 8Instituto de Biomedicina de Sevilla, Consejo Superior de Investigaciones Cienticas.
Hospital Universitario Virgen del Rocio, Univ. de Sevilla., Avda Manuel Siurot sn. 41013 Sevilla, Spain, 9Department of
Experimental and Clinical Medicine, University of Firenze, Florence 50134, Italy, 10School of Biological Sciences, University
of Reading, Hopkins Building, Reading, Berkshire RG6 6UB, UK, 11Department of Nutrition, University of Oslo, Oslo, Norway,
12Department of Biochemistry and Biology, University of Bath, Claverton Down, Bath BA2 7AY, UK, 13Department of Public
Health Sciences, College of Medicine, Pennsylvania State University, Hershey, PA 17033, USA, 14Center for Environmental
Carcinogenesis and Risk Assessment, Environmental Protection and Health Prevention Agency, 40126 Bologna, Italy,
15Department of Chemicals and Radiation, Division of Environmental Medicine, Norwegian Institute of Public Health, Oslo
N-0403, Norway, 16Planet Biotechnologies Inc., St Albert, Alberta T8N 5K4, Canada, 17Department of Molecular and Cellular
Biochemistry, University of Kentucky, Lexington, KY 40508, USA, 18Spanish National Cancer Research Centre, CNIO, Melchor
Fernandez Almagro, 3, 28029 Madrid, Spain, 19Department of Medicine, McGill University, Montreal, Quebec H4A 3J1,
Canada, 20Istituto di Genetica Molecolare, CNR, Via Abbiategrasso 207, 27100 Pavia, Italy, 21Division of Preclinical Innovation,
National Center for Advancing Translational Sciences, National Institutes of Health, 9800 Medical Center Drive, Bethesda,
MD 20892–3375, USA, 22Department of Cellular and Physiological Sciences, Life Sciences Institute, Faculty of Medicine,
The University of British Columbia, Vancouver, British Columbia V5Z 1M9, Canada, 23Department of Biology, Jackson State
University, Jackson, MS 39217, USA, 24Department of Molecular and Environmental Toxicology, University of Wisconsin-
Madison, Madison, WI 53706, USA, 25Department of Anatomy and Cell Biology, University of Western Ontario, London,
S256 | Carcinogenesis, 2015, Vol. 36, Supplement 1
Ontario N6A 3K7, Canada, 26Stanford University Department of Medicine, Division of Oncology, Stanford, CA 94305, USA,
27Superfund Research Program, National Institute of Environmental Health Sciences, Research Triangle Park, NC 27560,
USA, 28Department of Medicine, Oncology and Pathology, Stanford University, Stanford, CA 94305, USA, 29Department of
Natural Science, The City University of New York at Hostos Campus, Bronx, NY 10451, USA, 30Department of Environmental
and Radiological Health Sciences, Colorado State University, Fort Collins, CO 80523–1680, USA, 31Department of Head and
Neck Surgery/Head and Neck Cancer Research, Johns Hopkins University School of Medicine, Baltimore, MD 21205, USA,
32Department of Pharmacological and Biomolecular Sciences, Università degli Studi di Milano, 20133 Milan, Italy,
33Department of Genomic Medicine and Environmental Toxicology, Institute for Biomedical Research, National Autonomous
University of Mexico, Mexico City 04510, México, 34Department of Bioscience and Biotechnology, Sejong University, Seoul
143–747, Korea, 35Department of Biomedical Sciences, University of Cagliari, 09124 Cagliari, Italy, 36Department of Pathology,
Kuwait University, Safat 13110, Kuwait, 37Department of Biotechnology and Biosciences, University of Milano-Bicocca, 20126
Milan, Italy, 38SYSBIO Centre of Systems Biology, Department of Biotechnology and Biosciences, University of Milano-Bicocca,
20126 Milan, Italy, 39Human Safety and Environmental Research, Department of Health Sciences, College of North Atlantic,
Doha 24449, State of Qatar, 40Department of Toxicology, NUTRIM School for Nutrition, Toxicology and Metabolism, Maastricht
University, Maastricht 6200, The Netherlands, 41Department of Molecular Biology, School of Osteopathic Medicine, Rowan
University, Stratford, NJ 08084, USA, 42Hacettepe University, Center for Stem Cell Research and Development, Ankara 06640,
Turkey, 43Department of Biochemistry and Cancer Biology, Meharry Medical College, Nashville, TN 37208, USA, 44Center for
Radiological Research, Columbia University Medical Center, New York, NY 10032, USA, 45Instituto de Alta Investigacion,
Universidad de Tarapaca, Arica, Chile, 46School of Biological Sciences, University of Reading, Reading, RG6 6UB, UK,
47Department of Otolaryngology - Head and Neck Surgery, University of Michigan Medical School, Ann Arbor, MI 48109, USA,
48Environmental Risk and Health Unit, Flemish Institute for Technological Research, 2400 Mol, Belgium, 49Department of
Surgery, Pathology, Immunology, Duke University Medical Center, Durham, NC 27710, USA, 50Department of Biomedical
Sciences, Faculty of Medicine and Health Sciences, 43400 Universiti Putra Malaysia, Serdang, Selangor, Malaysia,
51Department of Life Sciences, College of Health and Life Sciences and the Health and Environment Theme, Institute of
Environment, Health and Societies, Brunel University Kingston Lane, Uxbridge, Middlesex UB8 3PH, UK, 52National Institute
for Environmental Studies, 16-2 Onogawa, Tsukuba, Ibraki 3058506, Japan, 53Department of Geriatric Medicine, Kyoto
University Hospital 54 Kawaharacho, Shogoin, Sakyo-ku Kyoto, 606–8507, Japan, 54Department of Urology, Kasr Al-Ainy School
of Medicine, Cairo University, El Manial, Cairo 11559, Egypt, 55Department of Life Sciences, Tzu-Chi University, Hualien 970,
Taiwan, 56School of Biotechnology, Yeungnam University, Gyeongbuk 712-749, South Korea, 57Department of Environmental
and Occupational Health, University of Arkansas for Medical Sciences, Little Rock, AR 72205, USA, 58Department of
Environmental Health, Indiana University, School of Public Health, Bloomington, IN 47405, USA, 59Department of Cytokinetics,
Institute of Biophysics Academy of Sciences of the Czech Republic, Brno, CZ-61265, Czech Republic, 60Regulatory Toxicology
Research Division, Bureau of Chemical Safety, Food Directorate, Health Canada, Ottawa, Ontario K1A 0K9, Canada,
61Department of Medicine, University of Louisville, Louisville, KY 40202, USA, 62Robley Rex VA Medical Center, Louisville, KY
40202, USA, 63Department of Applied Medical Sciences, University of Southern Maine, 96 Falmouth St., Portland, ME 04104,
USA, 64Mouse Models of Cancer Therapies Group, Vall d’Hebron Institute of Oncology (VHIO), 08035 Barcelona, Spain,
65Lombardi Comprehensive Cancer Center, Georgetown University Medical Center, Washington DC 20057, USA, 66Cancer
Research UK. Cambridge Institute, University of Cambridge, Robinson Way, Cambridge CB2 0RE, UK, 67Department of Cell
Biology, Pharmamar-SAU, Avda. De los Reyes, 1.28770-Colmenar Viejo, Madrid, Spain, 68Departments of Obstetrics and
Gynecology and Pathology, Pennsylvania State University College of Medicine, Hershey PA 17033, USA, 69Department of
Biochemistry, Rush University, Chicago, IL 60612, USA, 70Environmental and Molecular Toxicology, Environmental Health
Science Center, Oregon State University, Corvallis, OR 97331, USA, 71Department of Medicine/Medical Oncology, Rutgers
Cancer Institute of New Jersey, New Brunswick, NJ 08903, USA, 72Laboratory for Cell Genetics, Vrije Universiteit Brussel, 1050
Brussels, Belgium, 73Catalan Institution for Research and Advanced Studies (ICREA), Barcelona 08010, Spain, 74School of Public
Health, Curtin University, Bentley, WA 6102, Australia, 75Public Health and Clinical Services Division, Department of Health,
Government of Western Australia, WA 6004, Australia, 76Department of Urology, University of California Davis, Sacramento,
CA 95817, USA, 77Department of Microbiology and Immunology, School of Medicine, College of Medicine, Taipei Medical
University, Taipei 11031, Taiwan, 78School of Biomedical Sciences, The Chinese University of Hong Kong, Shatin, NT, Hong
Kong SAR, The People’s Republic of China, 79Faculty of Medicine, University of Otago, Dunedin 9054, New Zealand,
80Department of Biomedical Engineering and Cancer Biology, Vanderbilt University, Nashville, TN 37235, USA, 81Department of
Experimental Oncology, Mediterranean Institute of Oncology, Via Penninazzo 7, Viagrande (CT) 95029, Italy, 82Center for
Experimental Molecular Medicine, Academic Medical Center, Meibergdreef 9, Amsterdam 1105 AZ, The Netherlands,
83Division of Environmental Health Sciences, School of Public Health, University of California, Berkeley, CA 94720-7360, USA,
84Inammation and Cancer Research, National Cancer Institute (NCI) (Retired), National Institutes of Health, Bethesda, MD
20892, USA, 85Department of Molecular Medicine and Medical Biotechnology, Federico II University of Naples, 80131 Naples,
Italy, 86Centre De Recherche En Cancerologie, De Lyon, Lyon, U1052-UMR5286, France, 87United States Army Institute of Public
Health, Toxicology Portfolio-Health Effects Research Program, Aberdeen Proving Ground, Edgewood, MD 21010-5403, USA,
88Center for Genomic Science of IIT@SEMM, Fondazione Istituto Italiano di Tecnologia, Via Adamello 16, 20139 Milano, Italy,
89Department of Microbiology and Immunology, Virginia Commonwealth University, Massey Cancer Center, Richmond, VA
23298, USA, 90Institut De Recerca Hospital Vall D’Hebron, Passeig Vall d’Hebron, 119–129, 08035 Barcelona, Spain, 91University
of Puerto Rico, Medical Sciences Campus, School of Public Health, Nutrition Program, San Juan 00921, Puerto Rico,
92Department of Biological Chemistry, Medical School, University of Athens, Institute of Molecular Medicine and Biomedical
Research, 10676 Athens, Greece, 93White River Junction Veterans Affairs Medical Center, White River Junction, VT 05009, USA,
W.H.Goodson et al. | S257
94Geisel School of Medicine at Dartmouth, Hanover, NH 03755, USA, 95Advanced Molecular Science Research Centre (Centre
for Advanced Research), King George’s Medical University, Lucknow, Uttar Pradesh 226 003, India, 96Departments of
Neurosurgery and Biochemistry and Massey Cancer Center, Virginia Commonwealth University, Richmond, VA 23298, USA,
97Integrated Laboratory Systems Inc., in support of the National Toxicology Program Interagency Center for the Evaluation of
Alternative Toxicological Methods, RTP, NC 27709, USA, 98Analytische, Milieu en Geochemie, Vrije Universiteit Brussel, Brussel
B1050, Belgium, 99Department of Obstetrics and Gynecology, University of Melbourne, Victoria 3052, Australia, 100Center for
Environment and Water, Research Institute, King Fahd University of Petroleum and Minerals, Dhahran 3126, Saudi Arabia,
101School of Engineering and Materials Science, Queen Mary University of London, Mile End Road, London, E1 4NS, UK,
102Department of Biochemistry and Molecular Biology, Dalhousie University, Halifax, Nova Scotia B3H 4R2, Canada,
103Department of Pathology, Stony Brook School of Medicine, Stony Brook University, The State University of New York, Stony
Brook, NY 11794-8691, USA, 104Charles University in Prague, Third Faculty of Medicine, CZ-100 00 Prague 10, Czech Republic,
105School of Biological Sciences, The University of Reading, Whiteknights, Reading RG6 6UB, England, 106Department of
Pathology, Wayne State University, Detroit, MI 48201, USA, 107Computer Science Department, Southern Illinois University,
Carbondale, IL 62901, USA, 108Human Genetics Unit, Department of Pathology, Faculty of Medicine, Kuwait University, Jabriya
13110, Kuwait, 109Department of Molecular and Experimental Medicine, The Scripps Research Institute, La Jolla, CA 92037,
USA, 110Department of Biological, Chemical, and Pharmaceutical Sciences and Technologies, Polyclinic Plexus, University of
Palermo, Palermo 90127, Italy, 111Sbarro Institute for Cancer Research and Molecular Medicine, Temple University,
Philadelphia, PA 19122, USA, 112Department of Pharmacology and Toxicology, University of Vienna, Vienna A-1090, Austria,
113Institute of Cancer Research, Department of Medicine, Medical University of Vienna, Wien 1090, Austria, 114Departments of
Pharmacology and Hematology and Medical Oncology, Emory University School of Medicine and Winship Cancer Institute,
Atlanta, GA 30322, USA, 115Division of Hematology and Oncology, Department of Pediatrics, Children’s Healthcare of Atlanta,
GA 30322, USA, 116Department of Pediatrics, Emory University School of Medicine, Emory University, Atlanta, GA 30322, USA,
117Lineberger Comprehensive Cancer Center, University of North Carolina at Chapel Hill, NC 27599, USA, 118Department of
Cancer Biology, Perelman School of Medicine at the University of Pennsylvania, Philadelphia, PA 19104, USA, 119Program in
Genetics and Molecular Biology, Graduate Division of Biological and Biomedical Sciences, Emory University, Atlanta, GA
30322, USA, 120Population Health and Prevention, Research, Prevention and Cancer Control, Cancer Care Ontario, Toronto,
Ontario, M5G 2L7, Canada, 121Departments of Epidemiology and Occupational and Environmental Health, Dalla Lana School
of Public Health, University of Toronto, Toronto, Ontario, M5T 3M7, Canada, 122Department of Cardiac, Thoracic and Vascular
Sciences, Unit of Occupational Medicine, University of Padova, Padova 35128, Italy, 123Department of Anatomy, Physiology and
Biochemistry, The Swedish University of Agricultural Sciences, PO Box 7011, VHC, Almas Allé 4, SE-756 51, Uppsala, Sweden,
124Siriraj Center of Excellence for Stem Cell Research, Faculty of Medicine Siriraj Hospital, Mahidol University, Bangkok 10700,
Thailand, 125Department of Anatomy, College of Medicine, Yeungnam University, Daegu 705–717, South Korea,126Department
of Hygiene, Kawasaki Medical School, Matsushima Kurashiki, Okayama 701-0192, Japan, 127Division of Genetic and Molecular
Toxicology, National Center for Toxicological Research, United States Food and Drug Administration, Jefferson, AR 72079, USA,
128INSERM U1113, team 3‘Cell Signalling and Communication in Kidney and Prostate Cancer’, University of Strasbourg,
Faculté de Médecine, 67085 Strasbourg, France, 129INRS-Institut Armand-Frappier, 531 Boulevard des Prairies, Laval, QC H7V
1B7, Canada, 130Department of Biology, Geology and Environmental Sciences, Alma Mater Studiorum Università di Bologna,
Via Francesco Selmi, 3, 40126 Bologna, Italy, 131Center for Applied Biomedical Research, S.Orsola-Malpighi University Hospital,
Via Massarenti, 9, 40126 Bologna, Italy, 132National Institute of Biostructures and Biosystems, Viale Medaglie d’ Oro, 305, 00136
Roma, Italy, 133Department of Biosciences and Veterinary Public Health, Faculty of Veterinary Medicine, Swedish University of
Agricultural Sciences, PO Box 7028, 75007 Uppsala, Sweden, 134Department of Cell and Developmental Biology, University of
California, Irvine, CA 92697, USA, 135Department of Biology/Center for Cancer Research and Therapeutic Development, Clark
Atlanta University, Atlanta, GA 30314, USA, 136Harvard Medical School/Brigham and Women’s Hospital, Boston, MA 02115,
USA, 137Department of Biosciences and Veterinary Public Health, Faculty of Veterinary Medicine, Swedish University of
Agricultural Sciences, PO Box 7028, 75007 Uppsala, Sweden, 138Baylor College of Medicine, Houston, TX 77030, USA,
139Department of Pharmaceutical Sciences, West Virginia University, Morgantown, WV, 26506, USA 140Department of
Pharmaceutical Chemistry, Faculty of Pharmacy, Kuwait University, PO Box 24923, Safat 13110, Kuwait and 141Department of
Surgery, The Ohio State University College of Medicine, The James Comprehensive Cancer Center, Columbus, OH 43210, USA
*To whom correspondence should be addressed. William H.Goodson III, California Pacic Medical Center Research Institute, 2100 Webster Street #401,
San Francisco, CA 94115, USA. Tel:+41 59 233925; Fax: +41 57 761977; Email: whg3md@att.net
Correspondence may also be addressed to Leroy Lowe. Tel:+90 28 935362; Fax: +90 28 935610; Email: leroy.lowe@gettingtoknowcancer.org
Part of the special issue on ‘Assessing the Carcinogenic Potential of Low-Dose Exposures to Chemical Mixtures in the Environment: The Challenge Ahead’
Abstract
Lifestyle factors are responsible for a considerable portion of cancer incidence worldwide, but credible estimates from the
World Health Organization and the International Agency for Research on Cancer (IARC) suggest that the fraction of cancers
attributable to toxic environmental exposures is between 7% and 19%. To explore the hypothesis that low-dose exposures to
mixtures of chemicals in the environment may be combining to contribute to environmental carcinogenesis, we reviewed
11 hallmark phenotypes of cancer, multiple priority target sites for disruption in each area and prototypical chemical
S258 | Carcinogenesis, 2015, Vol. 36, Supplement 1
disruptors for all targets, this included dose-response characterizations, evidence of low-dose effects and cross-hallmark
effects for all targets and chemicals. In total, 85 examples of chemicals were reviewed for actions on key pathways/
mechanisms related to carcinogenesis. Only 15% (13/85) were found to have evidence of a dose-response threshold, whereas
59% (50/85) exerted low-dose effects. No dose-response information was found for the remaining 26% (22/85). Our analysis
suggests that the cumulative effects of individual (non-carcinogenic) chemicals acting on different pathways, and a variety
of related systems, organs, tissues and cells could plausibly conspire to produce carcinogenic synergies. Additional basic
research on carcinogenesis and research focused on low-dose effects of chemical mixtures needs to be rigorously pursued
before the merits of this hypothesis can be further advanced. However, the structure of the World Health Organization
International Programme on Chemical Safety ‘Mode of Action’ framework should be revisited as it has inherent weaknesses
that are not fully aligned with our current understanding of cancer biology.
Introduction
Cancer is a burden on humanity and among the leading causes
of morbidity and mortality worldwide, with ~14 million new
cases and 8.2 million cancer-related deaths in 2012 (1). In gen-
eral, both genetic and environmental factors play a role in an
individual’s cancer susceptibility (2,3), so there has been a long-
standing emphasis on avoidable ‘lifestyle’ factors (i.e. those
that can be modied to reduce the incidence of the disease) and
a parallel focus on exogenous chemical exposures (e.g. agricul-
tural, occupational and so on) (4). But advances in our under-
standing of the complexity of cancer biology have resulted in
serious critiques of current risk assessment practices related
to exogenous exposures (5) along with calls for an expanded
focus on research that will allow us to evaluate the (potentially
carcinogenic) effects of in-utero exposures and low-level expo-
sures to combinations of chemicals that occur throughout our
lifetime (6,7).
The 2008–09 President’s Cancer Panel Annual Report in the
USA (8) opined that the ‘true burden of environmentally induced
cancer has been grossly underestimated’ (7), whereas Parkin
etal. (9) estimated in a British study that the fraction of cancer
that can now be attributed to both lifestyle and environmen-
tal factors is only 43% (i.e. the underlying cause of 57% of all
cancers is still unexplained). However, an expanded focus on
research that will allow us to evaluate the (potentially carcino-
genic) contribution of low-level exposures to combinations of
chemicals that occur in utero and throughout our lifetime is not
a trivial undertaking.
First of all, the number of chemicals to which we are exposed
is substantial, and many have not been adequately tested.
Christiani (6) cited increased and persistently high incidence
rates of various cancers and called on the National Institutes of
Health to expand their investigation of environmental causes
of cancer noting that ‘Massive gaps exist in toxicologic data,
even in the case of widely used synthetic chemicals. Only about
50% of chemicals classied by the Environmental Protection
Agency (EPA) as “high production volume” have undergone
even minimal testing for carcinogenicity’. But even though the
incidence of cancer attributable to environmental exposures
has not been denitively established (3,6), it remains an impor-
tant focus of our prevention efforts [with credible estimates
from the World Health Organization [WHO] and the IARC sug-
gesting that the fraction of cancers attributable to toxic envi-
ronmental exposures is between 7% and 19%] (10,11).
The possibility that unanticipated low-dose effects (LDE) are
also a factor in environmental carcinogenesis further compli-
cates matters. Vandenberg etal. (12) recently reviewed the accu-
mulating evidence that points to LDE that occur at levels that are
well below those used for traditional toxicological studies. This
review identied several hundred examples of non-monotonic
dose-response relationships (i.e. examples where the relation-
ship between dose and effect is complex and the slope of the
curve changes sign—from positive to negative or vice versa
somewhere within the range of doses examined). Drawing on
the known actions of natural hormones and selected environ-
mental chemicals examined in cell cultures, animals and epi-
demiology, the authors emphasized that when non-monotonic
dose-response curves occur, the effects of low doses cannot be
predicted by the effects observed at high doses. However, endo-
crine disruption research to this point has been aimed primarily
at chemicals that disrupt developmental processes through a rel-
atively small subset of hormones (e.g. estrogen, androgen, thyroid
and so on), and thus, many commonly encountered chemicals
have not been tested at all for these effects (at environmentally
relevant dose levels) and, to date, mechanisms that relate to car-
cinogenesis have typically not been the focus of these studies.
Generally for chemical risk assessments, toxicity stud-
ies are conducted with individual chemicals in animal mod-
els based on regulatory test guidelines [e.g. Organization for
Economic Co-operation and Development (OECD) test guide-
lines (13)] with a key objective of providing a dose-response
assessment that estimates a point of departure [traditionally
the no-observed-adverse-effect level or the lowest-observed-
adverse-effect level (LOAEL)], which is then used to extrapo-
late the quantity of substance above which adverse effects can
be expected in humans. The no-observed-adverse-effect level,
combined with uncertainty factors (which acknowledge gaps
in the available data), is then used to establish safety criteria
Abbreviations
AhR aryl hydrocarbon receptor
BPA bisphenol A
EMT epithelial-mesenchymal transition
EPA environmental protection agency
HTS high-throughput screening
IARC International Agency for Research on Cancer
IL interleukin
LDE low-dose effects
LOAEL lowest-observed-adverse-effect level
LOEL lowest observed effect level
miRNA microRNAs
4-NP nonylphenol
MXC methoxychlor
NF-κB nuclear factor-κB
PBDE polybrominated diphenyl ethers
PPAR peroxisome proliferator-activated receptor
ROS reactive oxygen species
W.H.Goodson et al. | S259
for human exposure. However, in order to be able to detect
adverse effects utilizing classical toxicological endpoints, dose
selection has historically involved the use of high dose levels
and appropriate dose level spacing to obtain the LOAEL or no-
observed-adverse-effect level thresholds. Techniques such as
linear extrapolation or benchmark dose modeling (14) are then
employed to predict safety margins for low-dose exposures.
This approach to risk assessment depends on the use of appro-
priate and sensitive endpoints, and on valid assumptions for
extrapolation estimates (e.g. dose-response linearity) and cal-
culations, and on the existence of thresholds of effects (15–17).
So when the potential for non-linear dose-response relation-
ships is combined with the possibility of synergism between
and amongst low doses of mixtures of individual chemicals in
the environment, it appears plausible that chemicals that are
not individually carcinogenic may be capable of producing car-
cinogenic synergies that would be missed using current risk
assessment practices.
The complex nature of the biology of cancer adds another
layer of complexity for risk assessment. In a landmark paper
in 1979, Ames (18) noted that damage to DNA appeared to
be a major cause of most cancers and suggested that natu-
ral chemicals in the human diet and the tens of thousands
of man-made chemicals that had been introduced into the
environment in the preceding decades be tested for their abil-
ity to damage DNA. In doing so, he sketched out the difculty
of dealing with complex chemical mixtures and he proposed
the use of rapid mutagenicity assays to identify environ-
mental mutagens and carcinogens. The strategy was sound
at the time, but it led to a scientic and regulatory emphasis
on ‘mutagens as carcinogens’, whereas the issue of complex
environmental mixtures, or carcinogens that are not muta-
gens, was never vigorously pursued. Instead, what followed
was an international quest to nd individual chemicals and
a few well-dened mixtures (e.g. diesel exhaust) that could be
shown to be ‘complete’ carcinogens (i.e. substances that could
cause cancer on theirown).
However, advances in cancer biology have revealed the
limitations of this approach. Armitage and Doll rst laid out
a multistage theory of carcinogenesis in 1954 (19), and by
1990, initiation and promotion were well established as dis-
crete steps in the evolution towards malignancy, along with
the inuence of ‘free radicals’, proto-oncogenes, oncogenes,
epigenetic mechanisms and other synergistic or antagonistic
factors (20). In 2000, Hanahan etal. (21) gave structure to this
rapidly growing eld of research with the proposal that ‘the
vast catalog of cancer cell genotypes [could be organized into]
a manifestation of six essential alterations in cell physiology
that collectively dictate malignant growth’. They called these
alterations the Hallmarks of Cancer, dened as ‘… acquired
capabilities’ common to most cancers that ‘… incipient cancer
cells … [must acquire to] enable them to become tumorigenic
and ultimately malignant.’ The hallmarks delineated at the
time were as follows:
• Self-sufciency in growth signals (later renamed proliferative
signaling)—cancer cells grow at a seemingly unlimited rate.
• Insensitivity to antigrowth signals (evading growth suppres-
sors)—cancer cells are not subject to antigrowth signals or
withdrawal of normal growth signals.
• Evading apoptosis (resisting cell death)—cancer cells avoid the
usual process whereby abnormal or redundant cells trigger
internal self-destroying (as opposed to cell death) mecha-
nisms.
• Limitless replicative potential (enabling replicative immortal-
ity)—cancer cells do not senesce (or age) and die after a lim-
ited number of cell divisions.
• Sustained angiogenesis (inducing angiogenesis)—cancer cells
elicit new blood vessels to sustain growth.
• Tissue invasion and metastasis (activating invasion and
metastasis)—in situ or non-invasive cancers, e.g. ductal carci-
noma in situ in the breast or carcinoma in situ in colon polyps,
grow into pre-existing spaces but invasive tumors must cre-
ate a space to expand into normal tissue.
From this perspective risk assessments based on limited ‘mode
of action’ information, assumptions of linear dose-response
relationships and a focus on individual chemicals (as complete
carcinogens) appeared to be inadequate to estimate human can-
cer risks. So in 2005, a scientist at the United States EPA called
for a shift in risk assessment practices that would move the eld
towards the development of biomarkers directly related to the
pathways found within the Hallmarks of Cancer framework (22).
The Hallmarks of Cancer framework was subsequently revis-
ited by Hanahan et al. (21) and expanded to encompass addi-
tional areas suggested by subsequent cancer research (23). This
expansion included the following:
• Two enabling characteristics:
• Genome instability and mutation, which allows changes in one
cell to pass to daughter cells through mutation or epigenetic
changes in the parent cell DNA.
• Tumor-promoting inammation, which helps cancer cells grow
via the same growth signals normal cells provide to each
other during wound healing and embryonic growth; inam-
mation further contributes to the survival of malignant cells,
angiogenesis, metastasis and the subversion of adaptive
immunity (24).
• Two ‘emerging’ hallmarks:
• Avoiding immune destruction whereby tumor cells avoid
immune surveillance that would otherwise mark them for
destruction.
• Dysregulated metabolism, one of the most recognizable fea-
tures of cancer; its exclusion from the original list of hall-
marks (21) probably represented a signicant oversight, as it
constitutes one of the earliest described hallmarks of cancer
(25,26). It is needed to support the increased anabolic and cat-
abolic demands of rapid proliferation and is likely an enabler
of cancer development and its other associated hallmarks.
Unfortunately, risk assessment practices that are currently used
to assess the carcinogenic potential of chemicals have changed
very little (despite the vast literature that now underpins the
main tenants of the Hallmarks of Cancer framework). For exam-
ple, a chemical that disrupts DNA repair capacity might prove
to be non-carcinogenic at any level of exposure (when tested on
its own), but that very same chemical may have the potential to
be an important contributor to carcinogenesis (e.g. in the pres-
ence of mutagens that cause DNA damage). Similarly, a chemical
that has immuno-suppressive qualities may not be carcinogenic
on its own, but if it acts to suppress the immune response, it
may contribute to carcinogenesis (by dismantling an important
layer of defense) in the presence of other disruptive chemicals.
Considering the multistep nature of cancer and the acquired
capabilities implied by each of these hallmarks, it is therefore
a very small step to envision how a series of complementary
exposures acting in concert might prove to be far more carci-
nogenic than predictions related to any single exposure might
suggest (see Figure 1). Interacting contributors need not act
S260 | Carcinogenesis, 2015, Vol. 36, Supplement 1
simultaneously or continuously, they might act sequentially or
discontinuously. So a sustained focus on the carcinogenicity of
individual chemicals may miss the sorts of synergies that might
reasonably be anticipated to occur when combinations of disrup-
tive chemicals (i.e. those that can act in concert on the key mech-
anisms/pathways related to these hallmarks) are encountered.
To address the biological complexity issue associated with
chronic diseases, the EPA and other agencies have begun to pursue
risk assessment models that incorporate biological information.
This is the basis of the Adverse Outcome Pathway concept, a con-
struct that is gaining momentum because it ties existing knowl-
edge of disease pathology (i.e. concerning the linkage between
a direct molecular initiating event and an adverse outcome at a
biological level of organization) to risk assessment (27,28). This
line of thinking inspired a recent initiative by the EPA, where
the agency tested a proposal for characterizing the carcinogenic
potential of chemicals in humans, using in-vitro high-through-
put screening (HTS) assays. The selected HTS assays speci-
cally matched key targets and pathways within the Hallmarks
of Cancer framework. The authors tested 292 chemicals in 672
assays and were successfully able to correlate the most disrup-
tive chemicals (i.e. those that were most active across the vari-
ous hallmarks) with known levels of carcinogenicity. Chemicals
were classied as ‘possible’/‘probable’/‘likely’ carcinogens or des-
ignated as ‘not likely’ or with ‘evidence of non-carcinogenicity’
and then compared with in-vivo rodent carcinogenicity data in
the Toxicity Reference Database to evaluate their predictions. The
model proved to be a good predictive tool, but it was developed
only as a means to help the EPA prioritize many untested indi-
vidual chemicals for their carcinogenic potential (i.e. in order to
establish priorities for individual chemical testing (29)).
What is still needed, is an approach employing the Hallmarks
of Cancer framework that can be used to identify priority
mixtures (i.e. those with substantive carcinogenic potential).
Without a way to anticipate the carcinogenicity of complex
mixtures, an important gap in capability exists and it creates
a signicant weakness in current risk assessment practices.
Countries around the globe have made a signicant investment
in the regulatory infrastructure and risk assessment practices
that protect us from unwanted exposures to harmful chemicals
and carcinogens, so we wanted to review the biology of cancer
to map out the challenges associated with the development of
an approach that would help us assess the carcinogenic poten-
tial of low-dose exposures to chemical mixtures in the envi-
ronment. Such an approach was seen as a reasonable step to
provide impetus for progress in this area of research and ulti-
mately to inform risk assessment practices worldwide.
Materials and methods
In 2012, the non-prot organization ‘Getting to Know Cancer’ instigated
an initiative called ‘The Halifax Project’ to develop such an approach using
the ‘Hallmarks of Cancer’ framework as a starting point. The aim of the
project was to produce a series of overarching reviews of the cancer hall-
marks that would collectively assess biologically disruptive chemicals (i.e.
chemicals that are known to have the ability to act in an adverse manner
on important cancer-related mechanisms, but not deemed to be carcino-
genic to humans) that might be acting in concert with other seemingly
innocuous chemicals and contributing to various aspects of carcinogen-
esis (i.e. at levels of exposure that have been deemed to be safe via the
traditional risk assessment process). The reviews were to be written by 12
writingteams.
The writing teams were recruited by Getting to Know Cancer circu-
lating an email in July 2012 to a large number of cancer researchers, ask-
ing about their interest in the project. Respondents were asked to submit
personal details through a dedicated webpage that provided additional
project information. Atotal of 703 scientists responded to the email,
and from that group, 11 team leaders were selected to lead reviews of
each hallmark (10 Hallmarks plus an 11th team to consider the tumor
microenvironment as a whole) and one leader for the cross-validation
Figure1. Disruptive potential of environmental exposures to mixtures of chemicals. Note that some of the acquired hallmark phenotypes are known to be involved in
many stages of disease development, but the precise sequencing of the acquisition of these hallmarks and the degree of involvement that each has in carcinogenesis
are factors that have not yet been fully elucidated/dened. This depiction is therefore only intended to illustrate the ways in which exogenous actions might contribute
to the enablement of these phenotypes.
W.H.Goodson et al. | S261
team (see below). Writing group leaders were asked to form individual
teams drawn from the pool of researchers who expressed interest in
the project and from their own circles of collaborators. Leaders were
encouraged to engage junior researchers as well. Team leaders received
project participation guidelines and ongoing communication from the
project leaders, L.Lowe and M.Gilbertson. Each team included: a lead
author with a published expertise in the hallmark area; domain experts
who assisted in the production of the descriptive review of the biology;
environmental health specialists (e.g. specialists in toxicology, endo-
crine disruption, or other related disciplines) and support researchers.
Each writing team was charged to describe the hallmark, its systemic
and cellular dysfunctions and its relationships to other hallmarks. Apri-
ority list of relevant (i.e. prototypical) target sites for disruption was to
be developed by the team and a list of corresponding chemicals in the
environment that have been shown to have the potential to act on those
targets was requested, along with a discussion of related issues and future
research needed (in the context of project goals).
Selection of target sites for disruption
A ‘target’ was broadly dened as a procarcinogenic disruption at the sys-
tem level (e.g. the hypothalamic–pituitary–gonadal axis), organ level, tis-
sue level or cellular level. It was assumed from the outset that a project
intended to develop an approach for the assessment of the carcinogenic
potential of low-dose exposures to chemical mixtures in the environment
would encounter a practical upper limit to the number of potential targets
that any given team could realistically review. Therefore, each team was
asked to identify up to 10 relevant targets for their domain (bearing in
mind that each target would also serve as a starting point for the identi-
cation of a disruptive environmental chemical that had already shown
a demonstrated ability to act on that target). In theory, it was understood
that this could lead to as many as 110 targets for the entire project, and
as the teams were also asked to select one disruptive chemical for each
target, a maximum of 110 chemicals.
In this phase, teams were asked to focus on specic gene changes
common to many cancers as identied by The Cancer Genome Project
(30) in order to estimate how the function of specic genes might be
altered, not by specic gene mutations, but rather either by direct
action or by epigenetic changes that might lead to the same functional
ends. Most of these pathways and processes are found within both
the hallmarks of cancer and the genomic frameworks, so teams were
asked to evaluate both models and consider non-mutagenic/epigenetic
pathways of interference as well (given that epigenetic changes such as
DNA methylation and histone acetylation are relevant for cancer and
often inducible by chemicals and may be transmitted to daughter cells).
Selection of disruptive chemicals
Teams were then asked to identify ‘prototypical’ chemicals in the envi-
ronment that had demonstrated an ability to act on the selected targets.
During workshops in Halifax, the teams settled on the following criteria
to guide their choices:
• Chemicals should be ubiquitous in the environment because we
wanted the broadest possible relevance for the general popula-
tion.
• Chemicals should selectively disrupt individual targets such as
specic receptors, specic pathways or specic mechanisms. Hypo-
thetically speaking, a chemical could affect more than one pathway,
receptor and so on; indeed, we expected that most chemicals would
likely exert a multitude of actions. However, we used the term ‘selec-
tively disruptive’ to encourage teams to avoid choosing mutagens
that are randomly destructive in their action (i.e. unpredictable and
capable of producing varying types of damage across a wide range
of pathways).
• Chemicals should not be ‘lifestyle’ related, such as those encountered
from tobacco, poor diet choices (e.g. red meats, French fries, lack of
fruit and vegetables and so on), alcohol consumption, obesity, infec-
tions (e.g. human papillomavirus) and so on.
• Chemicals should not be known as ‘carcinogenic to humans’ (i.e. not
IARC Group1, carcinogens).
The choice to focus on environmental pollutants in this project was
intentionally restrictive. Countries around the globe have made sig-
nicant investments in regulatory infrastructure and risk assessment
practices to protect us from unwanted exposures to harmful chemicals
and carcinogens. Therefore, we focused on chemicals that are com-
monly encountered in the environment. Primarily, we wanted to gen-
erate insights that would be valuable for cancer researchers who are
specically interested in environmental chemical exposures to chemical
mixtures and/or those who are focused on risk assessment practices in
general.
Dose-response characterizations andLDE
Given that much of the evidence in the toxicological literature that docu-
ments the disruptive actions of various chemicals has been produced
under a wide range of differing experimental circumstances, we wanted
to assess the quality and relevance of data that were gathered for expo-
sures discussed in this review. Specically, for each chemical selected
and each mechanism identied, teams were additionally tasked to iden-
tify any dose-response characterization results and/or relevant low-dose
research evidence that might exist. The term ‘low dose’ was dened using
the European Food Safety Authority denition (i.e. responses that occur at
doses well below the traditional lowest dose of 1 mg/kg that are used in
toxicology tests) and the denition for ‘LDE’ was based on the EPA deni-
tion (31)—as follows:
Any biological changes occurring
(a) in the range of typical human exposuresor
(b) at doses lower than those typically used in standard testing proto-
cols, i.e. doses below those tested in traditional toxicology assess-
ments (32),or
(c) at a dose below the lowest dose for a specic chemical that has
been measured in the past, i.e. any dose below the lowest observed
effect level (LOEL) or LOAEL (33)
(d) occurring at a dose administered to an animal that produces blood
concentrations of that chemical in the range of what has been
measured in the general human population (i.e. not exposed oc-
cupationally, and often referred to as an environmentally relevant
dose because it creates an internal dose relevant to concentrations
of the chemical measured in humans) (34,35).
Each team was then asked to categorize each chemical by using one of
ve possible categories (to determine the relevance and relative strength
of the underlying evidence for each of the chemicals being considered).
The categories were as follows: (i) LDE (i.e. levels that are deemed relevant
given the background levels of exposure that exist in the environment);
(ii) linear dose-response with LDE; (iii) non-linear dose-response with LDE;
(iv) threshold (i.e. this action on this mechanism/pathway does not occur
at low-dose levels) and (v) unknown. Additional details of the descriptions
for each of these categories are shown in Table1.
Cross-hallmark relationships
In recognition of the network of signaling pathways involved and the
degree of overlap/interconnection between the acquired capabilities
described in each hallmark area, the project included a cross-validation
step to create a more complete mapping of the actions that might be
anticipated as the result of an action on the target sites identied or by
the disruptive effects of the chemicals selected. Given the diversity of the
targets involved in the 11 hallmark areas, it was anticipated that inhibit-
ing or stimulating a target relevant to one hallmark may have an impact
on other targets that are relevant, especially if both are linked via signal-
ing pathways.
Accordingly, the cross-validation team conducted additional back-
ground literature review of submitted targets and chemicals from each
writing team, searching for evidence to identify cross-hallmark activity.
Each potential target-hallmark or approach-hallmark interaction was
assessed to determine whether the inhibition or activation of each tar-
get and the corresponding biological activity of each chemical might
reasonably be expected to have either a procarcinogenic or anticarcinogenic
effect on key pathways/processes in the various hallmark areas.
S262 | Carcinogenesis, 2015, Vol. 36, Supplement 1
Table1. Dose-response characterization
Review team Chemical name Disruptive action on key mechanism/pathway Low-dose effect (LDE, LLDE, NLDE, threshold, unknown)
Angiogenesis Diniconazole Vascular cell adhesion molecule and cytokine signaling Threshold (H-PC) (36)
Ziram Vascular cell adhesion molecule and cytokine signaling Threshold (H-PC) (36,37)
Chlorothalonil Thrombomodulin, vascular proliferation and cytokine signaling Unknown (H-PC) (36), NLDE (A-in vivo) (38)
Biphenyl Angiogenic cytokine signaling Unknown (H-PC) (36)
Tributyltin chloride Vascular cell proliferation and adhesion molecule signaling Unknown (H-PC) (36)
Methylene bis(thiocyanate) Plasminogen activating system and cytokine signaling Unknown (H-PC) (36)
HPTE Vascular cell adhesion molecule and cytokine signaling Unknown (H-PC) (36), threshold (A-Ia) (39), LDE (A-Ia) (40)
PFOS Angiogenic cytokine signaling Threshold (H-PC) (36), LDE (H-CL) (41)
Bisphenol AF Matrix metalloproteinase expression and estrogen receptor signal-
ing
Unknown (H-PC) (36)
C.I.solvent yellow 14 AhR and hypoxic signaling Unknown (H-PC) (36)
Dysregulated metabolism Cypermethrin AR and ER expression, reduction of ATP and mitochondrial en-
zymes, mitochondrial membrane potential
LLDE (A-I) (42), NLDE (A-I) (42), NLDE (H-CL) (36,43,44)
Acrolein p53 activation, DNA repair inhibition, PERK phosphorylation, mito-
chondrial dysfunction, cell survival
LLDE (A-I, A-CL, H-PC, H-CL) (45–50), NLDE (49), threshold (46)
Rotenone Cell cycle, DNA damage response, proliferation, differentiation,
mitochondria
LLDE (H-CL) (51–53), NLDE (H-CL) (51,53), unknown (H-CL,H-
PC) (36)
Copper p53 activation, p21 up-regulation, cell viability LLDE (H-CL) (54–56)
Nickel Neutrophil apoptosis, E-cadherin regulation, matrix metallopepti-
dase (MMP) production
LLDE (H-CL) (57), NLDE (H-CL) (58), Threshold (H-CL) (58)
Cadmium p53-dependent apoptosis, cell proliferation LLDE (H-CL) (59), threshold (H-CL) (60)
Diazinon AChE activity, neuronal cytotoxicity Unknown (A-PC) (61), LLDE (H-CL) (62), threshold (H-CL) (36)
Iron KRAS mutations LLDE (A-I) (63)
Malathion Lymphocyte Mutations, Cytotoxicity Unknown (H-PC, H-E) (36,64)
Tissue invasion and
metastasis
BPA MMP-2 and MMP-9 expression, increased migration, invasion, EMT,
oxidative stress, ER signaling
LDE (H-CL) (65,66), threshold (H-CL, H-PC) (36)
Hexacholorobenzene Activation of c-Src, HER1, STAT5b and ERK1/2 signaling LLDE (H-CL, A-I) (67)
Sulfur dioxide MMP-9 expression Unknown (A-PC) (68)
Phthalates MMP-2 and MMP-9 expression LDE (H-CL) (66),Unknown (H-CL, H-PC) (36)
Iron ROI generation, NF-κB activation, uPA expression Unknown (H-CL) (69)
Biorhythms/melatonin GSK3β activation, EMT regulation Unknown (H-CL, H-E) (70,71)
Resistance to cell death BPA Inhibition of GJIC, activation of mTOR pathway, down-regulation of
p53, p21 and BAX, binding to ER-α, weakly binds to TH receptor
and AR, activation of ERK1/2 and p38
NLDE(H-CL, A-CL) (72–74)Threshold (H-CL, H-PC) (36)
Dibutyl phthalate Activation of PPAR-α, inhibition of GJIC, expression of cyclin D and
cdk-4, activation of AhR/HDAC6/c-Myc pathway
NLDE (H-CL) (75), unknown (H-CL, H-PC) (36)
Chlorothalonil Up-regulation of ErbB-2 tyrosine kinase and MAP kinase, aromatase
inhibitor
Threshold-based (i.e. non-linear) (A-I) (76), unknown (H-PC)
(36), threshold (H-CL) (36)
Lindane Induction of MAPK/ERK pathways Threshold-based (i.e. non-linear) (A-I) (77), threshold (H-CL)
(36)
Dichlorvos Expression of p16, Bcl-2 and c-myc LLDE (A-I) (78), threshold (H-CL) (36)
MXC Binding to ER-α receptor, up-regulation of cyclin D1, down-regula-
tion of p21
LLDE (H-CL, A-CL) (75,79), unknown (H-PC) (36), threshold
(H-CL) (36)
Oxyuorfen Expression of Cyp2b10 and Cyp4a10 transcripts (markers of PPAR-α
activation)
Threshold (A-I) (80), unknown (H-CL, H-PC) (36)
DEHP Activation of PPAR-α, inhibition of GJIC Threshold-based (i.e. non-linear) (A-I) (81)
Linuron Hypersecretion of LH, inhibition of GJIC Unknown (H-CL) (82)
W.H.Goodson et al. | S263
Review team Chemical name Disruptive action on key mechanism/pathway Low-dose effect (LDE, LLDE, NLDE, threshold, unknown)
Replicative immortality Nickel-derived compounds, (e.g.
nickel chloride)
Epigenetic silencing of p16 LLDE (H-CL, A-PC) (83)
Diethylstilbestrol Allelic loss and point mutation in ETRG-1 gene LLDE (A-I) (84)
Reserpine Epigenetic modications Unknown (A-PC) (85), threshold (H-CL) (36)
Phenobarbital Reduces expression of the CDKN1A product p21, CAR activation LLDE (A-I) (86,87)
Acetaminophen Cellular energy loss, mitochondrial damage, telomerase activation LDE (H-CL, A-I, A-CL) (88–92)
Cotinine Telomerase activation LLDE (H-PC) (93)
Nitric oxide p53 inactivation LLDE (H-PC, H-CL, A-CL, A-I) (94)
Na-selenite p53 promoter methylation LLDE (A-CL, A-I) (95,96)
Lead p53 inactivation LLDE (H-PC, H-CL, A-CL, A-I) (94)
Sustained proliferative
signaling
BPA Estrogen receptor activation, cell cycle/senescence LLDE (A-I, H-CL, H-E) (12,97), NLDE (A-I) (98,99), threshold (H-
CL) (36)
Cyprodinil Increased proliferation signaling, AhR activation Unknown (H-PC, H-CL) (36,100,101), threshold (H-CL) (36)
Imazalil AR signaling NLDE (A-I) (102,103), threshold (H-CL, H-PC) (36)
Maneb Nitric oxide signaling Unknown (A-CL, H-CL, H-PC) (36,104,105)
Methoxyclor ER signaling Threshold (H-CL) (36), LDE (A-I) (106,107), NLDE (A-I) (108)
PFOS Nuclear hormone receptors Threshold (H-CL) (36), LLDE (A-I) (109,110)
Phthalates CAR, ER signaling Unknown (H-CL) (36), LDE (A-I) (111–113)
Phosalone Increased proliferation, PXR signaling Unknown (H-PC, H-CL) (36,114,115)
PBDEs ER signaling LDE (A-I) (116,117)
Prochloraz ER signaling LDE (A-I) (118,119)
Trenbolone acetate Insulin-like growth hormone-1 and AR signaling Unknown, LDE (A-I, H-CL, H-E) (120,121)
Tumor-promoting inamma-
tion
BPA Immune cell proliferation, proinammatory cytokine induction Threshold (H-PC) (36), LDE (A-I, H-CL, H-E) (122–126)
Phthalates Immunomodulation of macrophages, lymphocytes, eosinophils and
neutrophils
Unknown (H-PC, H-CL, H-E) (36,127)
PBDEs Induction of pro-inammatory cytokines (IL-6, IL8 and CRP), inhibi-
tion of anti-inammatory cytokines (IL-10)
Threshold (H-PC, H-CL) (128–131)
Atrazine Immunomodulation of T cell and B cells, proinammatory cy-
tokines
Unknown (H-PC, A-I) (36,132,133)
Vinclozolin Proinammatory cytokine induction, NF-κB activation Unknown (H-PC, A-I) (36,134–136)
4-NP Proinammatory cytokine induction, NF-κB activation, iNOS induc-
tion
Unknown (A-CL, H-CL, H-PC) (36,137,138)
Immune system evasion Pyridaben Chemokine signaling, TGF-β, FAK, HIF-1a, IL-1a pathways Unknown (H-CL, H-PC, A-CL) (36,139,140), threshold (A-I) (141)
Triclosan Chemokine signaling, TGF-β, FAK, IL-1a pathways Threshold (H-CL, H-PC, A-I) (36,142–144), LDE (A-I, H-CL)
(145,146)
Pyraclostrobin Chemokine signaling, TGF-β, IL-1a pathways Unknown (H-CL, H-PC) (36)
Fluoxastrobin Chemokine signaling, EGR, HIF-1a, IL-1a pathways Unknown (H-CL, H-PC) (36)
BPA Chemokine signaling, TGF-β pathway Threshold (H-PC) (36), LDE (A-I) (12), NLDE (H-CL) (147), NLDE
(A-CL) (148–151), NLDE (A-I) (152–155)
Maneb PI3K/Akt signaling, chemokine signaling, TGF-β, FAK, IGF-1, IL-6,
IL-1a pathways
Unknown (H-CL, H-PC) (36,139,156–158), LDE (A-I) (159),
threshold (A-I) (139,160), threshold (A-CL, A-I) (161)
Table1. Continued
S264 | Carcinogenesis, 2015, Vol. 36, Supplement 1
Review team Chemical name Disruptive action on key mechanism/pathway Low-dose effect (LDE, LLDE, NLDE, threshold, unknown)
Evasion of antigrowth
signaling
DDT Induces MDM2, cyclin D1, E2F1 expression, disrupts gap junctions NLDE (A-I, H-CL, A-CL) (162–164)
Chlorpyrifos Increases proliferation LDE (H-CL, H-PC) (165,166)
Folpet Disrupts G1–S checkpoint kinases, down-regulates p53, promotes
proliferation
LDE(A-C) (167)
Atrazine Induces estrogen production and proliferation LDE(H-CL, A-I) (168–170)
BPA Reduced p53, reduced connexin 43 expression, increased prolifera-
tion
NLDE (H-CL, A-I) (171–174)
Tumor microenvironment Nickel ROS and cellular stress NLDE (A-I) (175)
BPA IL-6 expression, improper DC maturation and polarization, ROS
production
LLDE (A-I) (176), NLDE (A-I) (176)
Butyltins (such as tributyltin) NK cell inhibition LDE (A-I) (177)
MeHg Chronic oxidative stress LDE (H-PC, H-CL) (178,179)
Paraquat Chronic ROS production, cellular stress Unknown (A-I) (180)
Genome instability Lead Dysfunctional DNA repair, defect in telomere maintenance Unknown (A-CL) (181–183), threshold (H-CL, H-E) (184,185)
Acrylamide Inactivation of DNA repair proteins/enzymes Unknown (A-CL, A-I, H-CL) (186,187)
Quinones Affect free cysteine residues in catalytic center of DNA methyl-
transferases (DNMT)
Unknown (A-CL) (188)
Nickel Affect enzymes that modulate post-translational histone modica-
tion
LDE (H-E) (189,190), LDE (A-CL, H-CL) (191)
BPA Epigenetic changes via interactions with miRNA Threshold (H-PC) (192)
Alloy particles (tungsten/nickel/
cobalt)
Disruption of DNA damage/redox signaling involving Nrf, NF-κB,
Egr, and so on
LDE (A-I) (193)
Titanium dioxide NPs Decreased NADH levels and impaired mitochondrial membrane
potential and mitochondrial respiration, ROS generation
Unknown (A-PC) (194)
Benomyl Spindle defects leading to formation of micronuclei Threshold (H-CL) (195), Threshold (A-CL) (196)
Carbon nanotubes Spindle defects leading to formation of micronuclei LLDE (A-CL) (197,198), unknown (A-I) (198)
Each chemical in the table was categorized by using one of ve possible categories (to determine the relevance and relative strength of the underlying evidence for each of the chemicals being considered)—as follows: (1) LDE
(low-dose effect)—the ability of this chemical to exert this particular effect is not well characterized at a range of dose levels, but the evidence suggests that this chemical can exert this effect at low-dose levels (i.e. levels that
are deemed relevant given the background levels of exposure that exist in the environment and as further dened below). (2) LLDE (linear dose-response with low-dose effects)—the ability of this chemical to exert this particular
effect is well characterized at a range of dose levels and the evidence suggests that a linear dose-response relationship exists with effects at low-dose levels being evident (i.e. levels that are lower than the LOEL/LOAEL or thresh-
old and deemed relevant given the background levels of exposure that exist in the environment). Note: a linear dose-response model implies no threshold. Effects at low doses are the same as at higher doses even if at a lesser
extent. The effect is directly proportional to the dose. (3) NLDE (non-linear dose-response with low-dose effects)—the ability of this chemical to exert this particular effect is well characterized at a range of dose levels and the
evidence suggests that a non-linear dose-response relationship exists with exaggerated effects at low-dose levels being evident (i.e. levels that are lower than the LOEL/LOAEL or threshold and deemed relevant given the back-
ground levels of exposure that exist in the environment). Note: a non-linear dose-response with low-dose effect implies that the effect does not vary according to the dose of the agent. The effect at low doses may be the same
as at the higher doses or different. The non-linear dose-response may have or not have a threshold. It is represented by a sigmoid curve. The non-linear dose-response at low doses may be a non-monotonic dose-response. (4)
Threshold—the ability of this chemical to exert this particular effect is well characterized at a range of dose levels, and a threshold has been established for this chemical that suggests that this action on this mechanism/path-
way does not occur at low-dose levels (i.e. levels that are lower than the threshold and deemed relevant given the background levels of exposure that exist in the environment). (5) Unknown—although the ability of this chemi-
cal to exert this particular effect has been shown at higher dose levels, this effect is not well characterized at a range of dose levels, so a LOEL /LOAEL or a threshold has not been determined for this chemical and there is no
evidence showing that this chemical exerts this action at low-dose levels (i.e. levels that are lower than the LOEL/LOAEL or threshold and deemed relevant given the background levels of exposure that exist in the environment).
A-I, in-vivo animal models; A-CL, animal cell lines; A-PC, animal primary cells; H-PC, human primary cells; H-CL, human cell lines; H-E, human epidemiological studies. With respect to the human primary cell (H-PC) data from
ToxCast (36): unknown signies that the compound was tested across a range of doses and showed statistically signicant activity against the specied targets at the lowest test concentrations (~0.01µM); therefore, a threshold
could not be established. Threshold in this data set signies that there was no activity against the targets at one or more of the lowest concentrations tested.
aExtrapolated from in-vivo data on the parent compound, MXC.
Table1. Continued
W.H.Goodson et al. | S265
The cross-validation team also sought out controversial interactions
(i.e. mixed indications of hallmark-like effects) and instances where no
known relationship existed. It was our belief that target sites or chemicals
that demonstrated a substantial number of ‘anticarcinogenic’ effects in
other hallmark areas would be less suitable to serve as instigating con-
stituents in the design of carcinogenic mixtures (where procarcinogenic
synergy was being sought).
It is important to note that the cross-validation team was not
given any restrictions for literature selection for this effort, and con-
tributing authors were restricted neither to results from low-dose
testing, nor to that of cancer-related research. This approach was
taken because it was realized at the outset that this sort of breadth
and homogeneity (of low-dose evidence) does not yet exist in the lit-
erature. As a result, the types and sources of data gathered in this
effort varied considerably, resulting in an admixture of reviews and
original studies. Moreover, many studies that were cited in this effort
only considered a chemical’s ability to instigate or promote an action
that mimics a hallmark phenotype in a manner directionally consist-
ent with changes that have been associated with cancer. So, although
we have referred to these actions as procarcinogenic and anticarci-
nogenic, as these changes are frequently neither fixed nor specific
for cancer, the specificity of these changes and implications for car-
cinogenesis cannot and should not be immediately inferred from this
data set. Short-term toxicity and toxic responses—particularly in data
from in-vitro HTS platforms—must be distinguished from truly ‘carci-
nogenic’ long-term changes. In other words, the tabularized results
from this particular aspect of the project were only compiled to serve
as a starting point for future research. Where cross-hallmark effects
were reported (at any dose level and in any tissue type), we wanted
samples of that evidence to share with researchers who might be try-
ing to anticipate the types of effects that might be encountered in
future research on mixtures of chemicals (in a wide range of possible
research contexts).
Results
The results are presented roughly sequenced in a manner that
captures the acquired capabilities found in many/most cancers.
The section begins with two enabling characteristics found in
most cancers Genetic instability and Tumor-promoting inam-
mation, followed by Sustained proliferative signaling and
Insensitivity to antigrowth signals, the two related hallmarks
that ensure that proliferation is unabated in immortalized cells.
These sections are followed by Resistance to cell death and
Replicative immortality, two critical layers of defense that are
believed to be bypassed in all cancers and then by dysregulated
metabolism. Sections on Angiogenesis and Tissue invasion and
metastasis follow and speak to the progression of the disease,
and nally, the Tumor microenvironment and Avoiding immune
destruction sections offer summaries related to the very last
lines of defense that are defeated in most cancers. Additionally,
dose-response characterizations and evidence of LDE are then
presented for all of these areas and the results from the cross-
validation activity are summarized and reviewed.
Genetic instability
The phenotypic variations underlying cancer result from interac-
tions among many different environmental and genetic factors,
occurring over long time periods (199). One of the most important
effects of these interactions is genome instability—loosely dened
as an increased likelihood of the occurrence of potentially muta-
genic and carcinogenic changes in the genome. The term is used
to describe both the presence of markers of genetic change (such
as DNA damage and aneuploidy) and intrinsic factors that per-
mit or induce such change (such as specic gene polymorphisms,
defective DNA repair or changes in epigenetic regulation).
DNA damage—which can be caused by exposure to external
chemicals or radiation, or by endogenous agents such as reactive
oxygen or faulty replication—is an event that can initiate the
multistep process of carcinogenesis (200). Protection is afforded
at different levels; removal of damaging agents before they
reach the DNA, by antioxidant defenses and the phase I/phase
II xenobiotic metabolizing enzymes; a second line of defense,
DNA repair, operating on the damage that occurs despite the
primary protection; and as a last resort, apoptosis (programmed
cell death), disposing of heavily damagedcells.
A clear sign of genome instability is aneuploidy—a deviation
from the normal number of chromosomes (201). Aneuploidy is
a very common feature of human cancers. Another hallmark of
cancer is loss of the normal mechanism of telomere shortening,
which allows abnormal cells to escape senescence, by avoid-
ing the body’s ‘editing’ processes that normally eliminate aging
cells with their accumulated genome aberrations (202,203).
The genes of most signicance for cancer are the (proto)-
oncogenes which, if defective, or abnormally expressed, lead
to uncontrolled cell proliferation; tumor suppressor genes, the
normal products of which tend to switch off replication to allow
repair, and promote cell death if damage is excessive; and genes
such as those involved in DNA repair that can—if faulty—lead
to a ‘mutator phenotype’. Mutated proto-oncogenes and tumor
suppressor genes are found in most if not all cancers and
play key roles in cancer etiology (204–207). Rare mutations in
DNA repair genes greatly increase the risk of cancer (208,209).
However, the evidence for links between common variants of
repair genes and cancer is generally inconclusive (210).
The term ‘epigenetics’ refers to covalent modications of the
DNA (methylation of cytosine in ‘CpG islands’ within regula-
tory regions of genes) or of the histones. These modications
can control gene expression and the pattern of modications
is altered in many cancers (211,212). For instance, hypometh-
ylation of proto-oncogenes can lead to overexpression, which
is undesirable. MicroRNAs (miRNAs) are responsible for specic
down-regulation of gene expression at a post-transcriptional
level, by preventing translation from messenger RNAs. miRNAs
participate in DNA damage responses and some miRNAs are
deregulated in many cancers (213–215).
Mutations in germ and stem cells are potentially more seri-
ous than those in other cells as they are passed to the cells’
progeny within the developing embryo or regenerating tissue
(216,217). There is a presumed survival benet when stem cells
tend to show a particularly stringent maintenance of genome
integrity through cell cycle regulation and enhanced responses
to DNA damage (218).
The selected ‘chemical disruptors’ that induce genome
instability include chemicals that not only directly damage DNA
or cause mutations, but act indirectly, via pathways such as DNA
damage signaling, DNA repair, epigenetic regulation or mito-
chondrial function. They include the following:
Metals such as lead, nickel, cobalt and mercury (common
water pollutants) are known to disrupt DNA repair (181,219),
whereas nickel also affects epigenetic histone modication
(189,191) and lead causes defective telomere maintenance
(184,220). Alloy particles, containing tungsten, nickel and cobalt,
can be inhaled and disrupt redox signaling (193,221). Titanium
dioxide nanoparticles are also common in many consumer prod-
ucts and foods and have been reported to disrupt mitochondrial
function and increase oxidative stress, as well as inhibit DNA
repair and disrupt mitosis (194,222,223).
Acrylamide occurs in many fried and baked food products,
and (apart from the well-known DNA adduct formation) can
S266 | Carcinogenesis, 2015, Vol. 36, Supplement 1
inactivate many critical proteins by binding sulfhydryl groups
(186).
Bisphenol A (BPA) is a plasticizer used for manufactur-
ing polycarbonate plastics and epoxy resins, and it can leach
from plastics into food and water. It is implicated in disruption
of DNA methylation, histone acetylation and disturbance of
miRNA binding (192,224,225), redox signaling (226) and induc-
tion of micronuclei through spindle defects in mitosis (227).
The fungicide benomyl is metabolized to carbendazim;
both are classied as possible human carcinogens at present.
The route of exposure is most likely ingestion via residues in
crops. Benomyl disrupts the microtubules involved in the func-
tion of the spindle apparatus during cell division, leading to
production of micronuclei (Frame,S.R. etal., unpublished report,
Schneider,P.W. etal., unpublished report, (228)).
Halobenzoquinones are disinfection by-products in chlo-
rinated drinking water (229). Quinones are electrophilic com-
pounds, known to react with proteins and DNA to form adducts.
These electrophylic chemicals can interact with functional thiol
groups via Michaelis–Menton type addition, causing modica-
tion of enzymes involved in methylation and demethylation
(188). This mechanism might be shared by other xenobiotics
that increase reactive oxygen species(ROS).
Human exposure to nano-sized materials used in cosmetics,
biomedical compounds, textiles, food, plastics and paints has
increased not only in a conscious way but also passively by the
leakage of nanomaterials from different objects. Nanoparticles
can induce genome instability via mitochondrial-related apop-
tosis (230), decreased DNA repair (222,230,231), hypoacetyla-
tion of histones (232), disruption of DNA methylation (231),
up-regulation of miRNA (233), reducing telomerase activity
(220) and—more specically for carbon nanotubes—interact-
ing with components of the mitotic spindle during cell division
or interacting with proteins directly or indirectly involved in
chromosome segregation (197,234). Nano-sized materials can
also produce inammation and alteration of the antioxidant
defenses that can lead to genome instability.
Tumor-promoting inammation
One of the earliest hypothesized causes of tumors subsequently
supported experimentally was the irritation hypothesis pro-
posed by Virchow. Although it was recognized initially that injury
alone was insufcient for carcinogenesis, it was also recognized
that irritation may have an accessory or predisposing inuence
in tumor formation, and that it may be enough nally to upset
the balance of a group of cells which for some other reason were
already hovering on the brink of abnormal growth’ (235). Indeed,
it is now recognized that inammatory responses, similar to
those associated with wound healing or infection, support the
development of invasive carcinomas by altering the microen-
vironment in favor of proliferation, cell survival, angiogenesis
and tumor cell dissemination while also disrupting antitumor
immune surveillance mechanisms. In other words, inamma-
tion plays a critical role in tumorigenesis (23,24).
Inammation is an immediate and necessary host defense
mechanism in response to infection or tissue injury by noxious
stimuli. In tumor-associated inammation, both the epithelium
and the immune cells express receptors that signal the activa-
tion and production of a wide array of biologically active proteins
most analogous to an unhealed wound. The sustained or uncon-
trolled release of potent and reactive molecules such as prosta-
glandins, cytokines, ROS and chemokines from both the tumor
cell and the microenvironment constituents lead to progressive
genomic instability, alterations in the integrity and function of
the microenvironment including alterations in the vasculature
(e.g. vascular hyperpermeability, neovascularization and angio-
genesis), as well as alterations in local immune dynamics. The
cellular and molecular mechanisms include a diverse array of
immune- and tumor-cell-derived effector molecules such as the
proinammatory reactive oxygen and nitrogen species, a num-
ber or cytokines, chemokines as well as cyclooxygenase-2 and
its product, prostaglandin E2.
In general, there is a paucity of experimentation, and when
present, inconsistent ndings for the role of environmental
chemicals as proinammatory molecules and more so for a pro-
inammatory action as a co-factors in carcinogenesis. However,
some recent studies provide a credible mechanistic basis, par-
ticularly early life exposures that might act by disrupting the
immune cell balance toward inammation, and that manifest in
adulthood. One example is BPA, one of the most abundant and
best studied environmental endocrine disruptors, and its con-
troversial role as an immune disruptor. Specically, studies in
male rats found that early life BPA exposure leads to the devel-
opment of prostate intraepithelial neoplasia (a prostate cancer
precursor lesion) through a pathological process that includes
BPA-dependent epigenetic reprogramming of genes involved in
the development of lateral prostate inammation in adulthood
(236,237).
This work in prostate is complemented by a much more
extensive study of BPA effects on immune cell components,
particularly the T-cell compartment, demonstrating that BPA
acts as an immune disruptor by promoting ‘immune’ cell pro-
liferation though the exact nature of the effect on specic cells
of the immune system is poorly delineated. Most interesting is
the work by Yan etal. (122), who reported ndings suggesting
that the timing of BPA exposure during development (prenatally,
early life or adult) alters the effect of BPA on regulatory T cells.
BPA actions also map over to the effects on the immune sys-
tem including the promiscuity of BPA for a number of nuclear
receptors relevant to immune cells such as the estrogen recep-
tor and the aryl hydrocarbon receptor (AhR). As well, bulky BPA
analogs may act as antagonists of members of the peroxisome
proliferator-activated receptor (PPAR) family, an important fam-
ily of nuclear receptors with potent anti-inammatory function
(238,239). Effects on the PPAR nuclear receptors may also explain
inammation-associated phenotypes observed with exposures
to certain phthalates and nonylphenol (4-NP).
A second example is the reported immunotoxic effects of
atrazine (6-chloro-N-ethyl-N-(1-methylethyl)-1,3,5-triazine-2,4-
diamine) (240), a chemical that is the most commonly detected
triazine herbicide in USA soil and water. Atrazine is banned
by the European Union and drinking water exposures are sup-
posed to be limited in the USA to <3µg/l (although exposures
exceed this limit regularly), but the use of this chemical is high
and increasing in Asia and other countries. Thus, atrazine is an
important pesticide to which humans are exposed. Atrazine
exhibits weak mutagenicity and low oncogenic properties, but
research by a number of authors is emerging that suggests that
immune system disruption might be a concern (132,240,241).
Although the majority of work on atrazine has been focused
on its endocrine disrupting properties, there is also evidence
to support immunotoxicity including effects on T-lymphocytes
composition with oral dosing (242,243), modulation of nitric
oxide production (244) and potential generation of ROS (245,246).
The local production of reactive nitrogen species and ROS
by mast cells and macrophages are among the better stud-
ied immune modulatory molecules for which recent evidence
W.H.Goodson et al. | S267
supports important roles both in the tumor microenvironment
and in the tumor progression (247–249). Notably, these reactive
species trigger oxidative/nitrosative modications, which can
initiate redox signaling that tightly modulates the inammatory
response in a manner that is highly relevant for carcinogenesis
(250,251).
We also looked at polybrominated diphenyl ethers (PBDEs)
and their effects on inammatory cytokines. Peltier etal. (128)
recently found that placental explants treated with a mix-
ture of the cogeners BDE-47, -99 and -100 and then exposed
to Escherichia coli were ‘reprogrammed’ toward a proinam-
matory response (increased IL-1β and tumor necrosis factor
α) and away from the expected anti-inammatory response
(decreased IL-10) compared with untreated placenta. Although
these studies are preliminary, chronic PBDE exposure may
lower the threshold for bacteria to stimulate a proinammatory
response, which has potential relevance given the established
link between bacteria and certain cancers (e.g. Helicobacter pylori
and gastric cancer), where tumor development is dependent on
inammation.
Vinclozolin was also of particular interest as an environmen-
tal chemical because transient early life exposures in utero have
been linked to both adult-onset disease and transgenerational
disease that involves inammation (134,135). For example, tran-
sient vinclozolin exposure in utero has been shown to promote
inammation in the prostate (prostatitis) of postpubertal rats
coupled with a down-regulation of the androgen receptor and
increase in nuclear factor-κB (NF-κB). The late or delayed effect
of exposure is hypothesized to reect a mechanism whereby
vinclozolin exposure during a critical development window
imprints an irreversible alteration in DNA methyltransferase
activity, leading to reprogramming of the androgen receptor (AR)
gene(s), which manifest as inammation in early adult life with
adverse effects on spermatid number.
Similarly, 4-NP has been shown to increase progenitor white
adipose levels, body weight and overall body size in rodents
exposed prenatally. Like vinclozolin, 4-NP effects on adipogen-
esis in the perinatal period confer transgenerational inheritance
of the obesogenic effects observable in F2 offspring, consistent
with genome reprogramming through an epigenetic process
(252) and others have reported immune and inammation-
related effects (137,138) making it relevant to carcinogenesis a
deserving further investigation.
Sustained proliferative signaling
Sustained proliferative potential is an essential component of
cancerous growth. Progressive conversion of normal cells into
cancer cells requires a series of genetic alterations, where each
alteration confers one or more types of growth advantage. One
such alteration that affords the transformed cell a distinct
growth advantage over its normal counterparts is the acquired
capacity of the cancer cell to proliferate in a sustained manner,
so as to crowd out and outnumber the normal cell population
(23). One of the fundamental differences between a normal and
a transformed cell is that normal cells halt proliferation when
subjected to growth inhibitory signals or in the absence of
growth stimulatory signals (253). But tumor cells act to sustain
proliferative signaling in several different ways. They can acti-
vate specic genes to produce relevant growth factors, which in
turn bind to signaling receptors giving rise to an autocrine loop
(254). Growth factors produced by tumor cells can also stimulate
the proliferation of stromal cells that in turn produce growth
factors to sustain tumor cell proliferation (255). Sustained pro-
liferation can additionally be maintained at the receptor level
by truncation of signaling receptor proteins whereby the ligand-
activated switch is missing (256). Alternatively, the number of
high-afnity receptor proteins may be increased to levels that
will sustain proliferative signaling in otherwise normal growth
factor levels. Finally, sustained proliferative signaling may well
be the result of perpetual activation of the intracellular sign-
aling chain independent of growth factors or receptors (e.g.
mutated ras (257) or truncated src (258) are intermediaries of a
normal proliferation signaling chain responsible for sustained
proliferation).
We hypothesized that disruptive environmental chemicals
acting in a procarcinogenic manner by inducing what is referred
to as ‘sustained cell proliferation’ likely exerted their action by
interfering with some basic control mechanisms (23,253). For
instance, they could achieve this by positively regulating tar-
gets within and outside the cell known to promote cell prolif-
eration or negatively regulating targets within and outside the
cell known to halt cell proliferation. In this way, such chemicals
could confer proliferative advantage to a distinct cell population
and contribute to that population’s capability to successfully
breach innate anticancer defense mechanisms and to become
progressively autonomous.
Specically, we identied a total of 15 ubiquitous chemical
disruptors capable of producing sustained cell proliferation. The
majority of these chemicals interacted with multiple targets,
and we have tabled this information in our review. In summary,
we identied several commonly used insecticides and fungi-
cides capable of causing sustained proliferation. These included
cyprodinil, etoxazole, imazalil, lactofen, maneb, methoxychlor
(MXC), phosalone, prochloraz and pyridaben, all of which tar-
geted estrogen receptor α and frequently other steroid hormone
receptors such as androgen receptor (102,259–275). Most of
these chemicals also targeted growth factors and their recep-
tors (264,267) and induced cytokines and cytokine receptors
(identied by ToxCast high throughput assay). Top disrupting
chemical fungicides and insecticides were MXC and cyprodinil,
which each interacted with a total of six individual targets that
further included the AhR (100), B-lymphocyte markers (ToxCast
2009 high-throughput assay, both chemicals), AP-1 proteins/
transcription/translation regulators, downstream signaling
molecules and cell cycle regulators (276,277). Other strong per-
formers for sustained proliferation were BPA (activated all tar-
gets activated by the insecticides and fungicides above except
growth factors and their receptors, B lymphocyte markers and
PPAR, but included cell cycle regulators alongside AP-1 proteins/
transcription/translation regulators and downstream signaling)
(272,276,278,279) (also identied in ToxCast high-throughput
assay, 2009), polyuorinated octinoid sulfate and polybromi-
nated diphenylethers (ame retardants) that either activated
AhR (280,281) or up to ve other targets that included steroid
receptors, growth factors, cytokines and cell cycle regulators
(109) (ToxCast high-throughput assay 2009). Three other con-
tenders were phthalates (plasticizers that acted via three tar-
gets that included AhR, steroid hormone receptors and PPAR)
(282–285), trenbolone acetate (a synthetic anabolic steroid
that unsurprisingly acted through steroid hormone receptors)
(120,286–290) and nally, edible oil adulterants (food contami-
nants produced during food processing that acted via down-
stream signaling) (291,292).
We have shown estrogen and androgen receptors to be
important targets in relation to sustained proliferative signaling
(293), and note that environmental estrogens and androgens are
frequently recognized as prototypical disruptor(s) of this hall-
mark. Although this is a small sample, there are a great number
S268 | Carcinogenesis, 2015, Vol. 36, Supplement 1
of chemicals in the environment, both naturally occurring and
man-made, are estrogenic, interact with estrogen receptor and
produce estrogen metabolites (just as naturally derived ovarian
estrogen does during metabolic breakdown). Catechol estrogens
(hydroxyl derivatives of estrogens), which are formed during
estradiol metabolism, are also potentially important mediators
of endogenous estradiol levels, and therefore of sustained pro-
liferative signaling and oncogenesis (294).
Insensitivity to antigrowth signals
Cell cycle arrest is important for maintaining genomic integrity
and for preventing genetic errors from being propagated. The
normal cell cycle contains multiple checkpoints to safeguard
against DNA-damaging agents. Specic proteins at these check-
points are activated in response to harmful stimuli, ensuring
that cellular proliferation, growth and/or division of cells with
damaged DNA are blocked.
There are multiple key mediators of growth inhibition that
may become compromised during carcinogenesis. Some, such
as p53, RB1, and checkpoint kinases cause cells to arrest at the
G1–S phase transition when they are activated by DNA damage.
Mutations in the p53 gene occur in ~50% of all cancers, although
certain tumor types, such as lung and colon, show a higher than
average incidence (295). Similarly, pRb hyperphosphorylation
(296), direct mutations (297), loss of heterozygosity (298) and dis-
ruption of the INK4–pRb pathway (INK4–CDK4/6–pRb–E2Fs) (299)
are common events in the development of most types of can-
cer. Cancer cells may also evade the growth inhibitory signals
of transforming growth factor-β (TGF-β) (300) and modulate the
action of downstream effectors as well as crosstalk with other
pathways.
Cells also receive growth inhibitory signals through intercel-
lular communication via gap junctions. Gap junctions disperse
and dilute growth-inhibiting signals, thereby suppressing cell
proliferation. In contrast, loss of gap junctions increases intra-
cellular signaling, leading to enhanced proliferation and tumor
formation. The molecular components of gap junctions are the
connexin proteins (301). Connexins are recognized as tumor sup-
pressors and have been documented to reduce tumor cell growth.
Numerous environmental stimuli have been reported to directly
affect gap junction intercellular communication. Adherens junc-
tion machinery mediates contact inhibition of growth, and loss
of contact inhibition is a mediator of tumor cell growth.
Chemicals that may contribute to insensitivity to antigrowth
signals through multiple targets of this hallmark are BPA, a
common constituent of everyday plastics, and pesticides such
as DDT, folpet and atrazine. BPA promotes proliferation by dis-
rupting the growth inhibitory signals of p53 and gap junction
communication (171,302). DDT has also been shown to enhance
proliferation by increasing the expression of Ccnd1 (cyclin D1)/
E2f, inducing phosphorylation of pRb, increasing the expression
of p53-degrading protein Mdm2 (a negative regulator of p53)
(162) and disrupting gap-junctional intercellular communica-
tion (163,164). Folpet down-regulates the functions of p53 and
ATM/ATR checkpoint kinases (167) and promotes proliferation,
whereas atrazine shows genotoxic effects at subacute dose on
Wistar rats. Genotoxicity was also associated with increased
transcription of connexin accompanied with increased oxida-
tive stress (303).
Resistance to celldeath
Cell death is an actively controlled and genetically regulated
program of cell suicide that is essential for maintaining tis-
sue homeostasis and for eliminating cells in the body that are
irreparably damaged. Cell death programs include: apoptosis,
necrosis, autophagy senescence and mitotic catastrophe (21).
Defects in these pathways are associated with initiation and
progression of tumorigenesis. Normally, cells accumulate from
an imbalance of cell proliferation and cell death, permissive cell
survival amidst antigrowth signals such as hypoxia and con-
tact inhibition, resistance to the killing mechanisms of immune
cell attack and anoikis resistance (304). Increased resistance to
apoptotic cell death involves inhibition of both intrinsic and
extrinsic apoptotic pathways.
The link between malignancy and apoptosis is exemplied
by the ability of oncogenes, such as MYC and RAS, and tumor
suppressor genes, such as TP53 and RB, to engage both apop-
tosis and the aberrant alterations of apoptosis regulatory pro-
teins such as BCL-2 and c-FLIP in various solid tumors (305). This
variety of signals driving tumor evolution provides the selective
pressure to alter apoptotic programs during tumor development.
Some chemical carcinogens and sources of radiation cause DNA
damage and increase genetic and/or epigenetic alterations of
oncogenes and tumor suppressor genes leading to loss of cel-
lular homeostasis (306). Other signals include growth/survival
factor depletion, hypoxia, oxidative stress, DNA damage, cell
cycle checkpoint defects, telomere malfunction and oncogenic
mutations, and exposure to chemotherapeutic agents and heavy
metals (307,308).
Cancer cells resist apoptotic cell death by up-regulation of
antiapoptotic molecules and the down-regulation, inactivation
or alteration of pro-apoptotic molecules. Activation of p53 usu-
ally induces expression of pro-apoptotic proteins (Noxa and
PUMA) and facilitates apoptotic cell death (309). Antiapoptotic
Bcl-2 family proteins suppress pro-apoptotic Bax/Bak [which
would otherwise inhibit mitochondrial outer membrane perme-
abilization]. Mitochondrial outer membrane permeabilization
releases cytochrome c and triggers apoptosis through an intrin-
sic pathway (310). Thus, regulation of apoptosis can be achieved
by inhibiting the antiapoptotic Bcl-2 family proteins and Bcl-XL
proteins as this restores a cell’s ability to undergo apoptosis. In
the process, mitochondrial outer membrane permeabilization,
mitochondrial proteins (Smac/DIABLO and Omi/HtrA2), which
inhibit the X-linked inhibitor of the apoptosis protein, are leaked
to trigger caspase activity in apoptosis (311,312).
Normal cellular metabolism is important for the sur-
vival of cells, whereas dysregulated metabolism in cells (see
Dysregulated metabolism) can induce either apoptosis or resist-
ance to apoptotic stimuli (313). In the liver, nearly every enzyme
in glycolysis, in the tricarboxylic acid cycle, in the urea cycle, in
gluconeogenesis and in fatty acid and glycogen metabolism is
found to be acetylated, and this N-α-acetylation confers sensi-
tivity to apoptotic stimuli (314). The antiapoptotic protein, Bcl-xL
reduces the efux of acetyl-CoA from the mitochondria to the
cytosol in the form of citrate and decreases N-α-acetylation of
apoptotic proteins, which enables cells less sensitive toward
apoptotic stimuli to mediate cell proliferation, growth and sur-
vival. Thus, N-α-acetylation might be a major factor in overcom-
ing apoptotic resistance in cancer cells (315,316).
Death receptor ligands such as TRAIL—which is bound to
DR4/DR5—induce receptor oligomerization and recruitment of
FADD and caspase-8 to form death-inducing signaling com-
plex, which leads to subsequent cell death via apoptosis. Thus,
expression of death receptors and their decoy receptors (Dcr1/2)
mediates apoptosis in tumor cells (317). When normal cells lose
contact with their extracellular matrix or neighboring cells, they
undergo an apoptotic cell death pathway known as ‘anoikis’
(304). During the metastatic process, cancerous cells acquire
W.H.Goodson et al. | S269
anoikis resistance and dissociate from primary sites, travel
through the vascular system and proliferate in distant target
organs.
A blockage of gap junction intracellular communication
(GJIC) between normal and preneoplastic cells also creates an
intra-tissue microenvironment in which tumor-initiated prene-
oplastic cells are isolated from growth controlling factors of nor-
mal surrounding cells resulting in clonal expansion (318). Gap
junction channels and Cxs control cell apoptosis by facilitating
the inux and ux of apoptotic signals between adjacent cells
and hemi-channels between the intracellular and extracellular
environments, and Cx proteins in conjunction with their intra-
cytoplasmic localization, may act as signaling effectors that are
able to activate the canonical mitochondrial apoptotic pathway
(319).
Several anthropogenic chemicals can affect resistance to cell
death. For example, BPA has been shown to strikingly impair
TP53 activity and its downstream targets, cell cycle regulators,
p21WAF1 and RB, or pro-apoptotic BAX, thereby enhancing the
threshold for apoptosis (172).
Chlorothalonil, a broad-spectrum fungicide that is used on
vegetables, fruit trees and agricultural crops, is considered to be
non-genotoxic but classied as ‘likely’ to be a human carcino-
gen by all routes of exposure (29). In a eukaryotic system, chlo-
rothalonil reacted with proteins and decreased cell viability by
formation of substituted chlorothalonil-reduced glutathione
derivatives and inhibition of specic nicotinamide adenine dinu-
cleotide thiol-dependent glycolytic and respiratory enzymes (320).
Caspases (cysteine-dependent proteases) and transglutaminase
are some of the thiol-dependent enzymes involved in apoptosis,
so inhibition of these thiol-dependent enzymes in tumor-initiated
cells may disrupt apoptotic cell death and aid in tumor survival.
Dibutyl phthalate and diethylhexyl phthalate (DEHP) are
diesters of phthalic acid and commonly referred to as phtha-
lates. In general, mimic the function or activity of the endoge-
nous estrogen 17β-estradiol (E2) and bind to estrogen receptors.
Interestingly, phthalates can mimic estrogen in the inhibition
of TAM-induced apoptosis in human breast cancer cell lines by
increasing intracellular Bcl-2/Bax ratio in breast cancer (321).
Lindane, an organochlorine pesticide, bioaccumulates in
wildlife and humans. Exposure to lindane induces tumor for-
mation in the mouse 42GPA9 Sertoli cell line by disrupting the
autophagic pathway and sustained activation of the mitogen-
activated protein kinase (MAPK)/extracellular signal-regulated
kinase (ERK) pathway (322).
MXC (1,1,1-trichloro-2,2-bis(4-methoxyphenyl)ethane) is a
DDT derivative that was developed after the ban of DDT and
it exhibits antiandrogenic and estrogenic activity. MXC stimu-
lates proliferation and human breast cancer cell growth by the
up-regulation of genes that involve cell cycle (cyclin D1), and the
down-regulation of genes p21 and Bax affecting G1/S transition
and apoptosis, respectively, through ERα signaling (323).
Replicative immortality
Cellular senescence is a state of irreversible arrest of cellular
proliferation characterized by changes in transcription, chro-
matin conformation, cytoplasmic and nuclear morphology,
DNA damage signaling and a strong increase in the secretion of
proinammatory cytokines (324) Senescence is the rst line of
defense against potentially transformed cells (325). Progression
to malignancy correlates with a bypass of cellular senescence.
Thus, senescence inhibits the activation of the tumorigenic
process (325). Senescence has been observed in vitro and in vivo
in response to various stimuli, including telomere shortening
(replicative senescence), oncogenic stress, oxidative stress and
chemotherapeutic agents (326).
Cellular senescence exhibits several layers of redundant
regulatory pathways. These pathways converge to arrest the cell
cycle through the inhibition of CDKs. The best-known effector
pathways are the p16INK4a/pRB, the p19ARF/p53/p21CIP1 and
the PI3K/mammalian target of rapamycin (mTOR)/FOXO path-
ways (327–330), which show a high degree of interconnection.
Additionally, the pRb and the mTOR pathways are two routes
that have been proposed to be responsible for permanent arrest
of the cell cycle (331). More pathways and genes are being dis-
covered, increasing the complexity of our knowledge of this
physiological process (329). Most, if not all of these genes have
been related to human tumorigenesis.
Despite the relevance of senescence as a gatekeeper in the
process of tumorigenesis, there is not a large body of infor-
mation exploring the effect of chemicals on this safeguard.
Little research has been undertaken on chemicals that alter
gene expression regulating senescence and few genes have
been identied (e.g. telomerase, p53, pRb, INK4a) (83,332,333).
Traditional protocols for the assessment of the carcinogenic
risk rely on the detection of tumors induced by agents that
alter many different pathways at the same time (includ-
ing senescence). These agents are mainly unspecic muta-
gens or epigenetic modiers. The effect of some compounds
is being explored including nickel-derived compounds (e.g.
nickel chloride), diethylstilbestrol, reserpine or phenobarbital
(83,334–337).
There may be environmental chemicals that are not muta-
gens or epigenetic modiers, but that target specic proteins on
the senescence pathways and may affect the initiation of tumo-
rigenesis by other compounds allowing senescence bypass. The
contribution of these compounds to the carcinogenesis process
is largely unknown. A few compounds bypass senescence in
this specic manner—acetaminophen, cotinine, nitric oxide,
Na-selenite and lead. Other chemicals known to alter senes-
cence only are mostly unknown (86,88–91,338–341).
Senescence has strong fail-safe mechanisms, and experi-
mental attempts to bypass senescence are usually recognized
as unwanted signals and trigger a senescence response anyway.
However, these conclusions are based on the interpretations of
experimental designs in which acute molecular or cellular alter-
ations are produced. There are few experiments regarding the
effects of chronic, low-dose alterations and even fewer studies
that consider the different cellular and molecular contexts that
can arise over the course of a lifetime.
Dysregulated metabolism
The highly glycolytic cancer phenotype described by Warburg
etal. (25) in the early 20th century determined much of the initial
direction in cancer research (26). Other characteristic metabolic
abnormalities have also been described (25,26,342,343) and have
recently garnered increased attention (344–348). These changes
are neither xed nor specic for cancer (349–351), but the uni-
versality of metabolic dysregulation suggests major roles in can-
cer genesis, maintenance and progression. Precise denitions of
what constitutes cancer metabolism, and when such changes
rst occur during the course of cancer development, are lacking.
From a teleological perspective, alterations in both intermediary
metabolism and its control are not surprising insofar as highly
proliferative cancer cells exhibit increased energy demands
and expanded requirements for macromolecular precursors to
S270 | Carcinogenesis, 2015, Vol. 36, Supplement 1
support nucleic acid and protein biosynthesis, as well as mem-
brane biogenesis, for increased biomass. Metabolic reprogram-
ming ostensibly equips cancer cells to cope with these demands,
as well as accompanying cellular stresses. Although much of
the attention on cancer metabolism has focused on enhanced
glucose utilization via glycolytic and pentose phosphate path-
ways, cancer cells are also capable of the oxidative utilization of
carbohydrates, lipids and peptides, and the metabolism of these
individual substrate classes remain intimately intertwined as in
normal cells (26,345,352).
Major control of glycolysis is traditionally ascribed to glu-
cose transport, hexokinase, phosphofructokinase and pyru-
vate kinase (352). Glyceraldehyde-3-phosphate dehydrogenase
also normally couples glycolytic ux to mitochondrial metabo-
lism in the presence of oxygen and to lactate generation in its
absence, but this relationship is fundamentally altered in can-
cer (26,345,353,354). Given the central importance of the pen-
tose phosphate pathway to anabolic metabolism and redox
homeostasis, glucose-6-phosphate dehydrogenase and its redox
coupling partners represent similarly attractive carcinogenic
targets (355). In addition, the enzymes of the tricarboxylic acid
cycle, such as fumarate hydratase, succinate dehydrogenase
and isocitrate dehydrogenase, play crucial roles in oxidative
energy metabolism and the interconversion of metabolic inter-
mediates, making them appealing candidates for study as well
(356,357).
The central importance of the mitochondrial electron trans-
port chain to oxidative energy metabolism and its established
role in toxic responses and dysregulated mitochondrial func-
tion in cancer makes its assembly and function attractive topics
for study (358–360). Despite well-established roles for lipid and
amino acid metabolism in cancer development and progression,
they have historically received less attention than carbohydrate
metabolism (26). Lipogenic, lipolytic and lipophagic pheno-
types are now widely recognized (344,361–363), so targets such
as acetyl-CoA carboxylase, fatty acid synthase, cellular lipases
and lipid transporters represent additional attractive targets
for study. Amino acid metabolism—particularly glutamine and
serine metabolism—also has well-established roles in cancer
(364–366), providing additional potential targets for study that
include 3-phosphoglycerate dehydrogenase (346,365,367,368)
and cellular transaminase coupling mechanisms. Study of
both lipid and protein metabolism must accommodate the fact
that cancer cells exhibit substrate preferences, including well-
described endogenous lipid- and protein-sparing effects of
exogenous glucose availability in cancercells.
The metabolic capacity of both normal cells and cancer cells
generally exceed their catabolic and anabolic requirements
(364,369,370), and only a fraction of the available potential energy is
ultimately required for cell survival (371,372). Moreover, very small
changes in metabolic ux can have profound phenotypic conse-
quences, and metabolic control analysis has suggested that the
importance of increased cancer-associated glycolytic and glutami-
nolytic uxes may lie not in their magnitudes, but in the mainte-
nance and control of smaller branched pathway uxes (364). For
these reasons, rigorous functional validation is needed for all can-
cer-associated changes in gene expression or metabolite accumu-
lation. Well-described moonlighting functions for many metabolic
enzymes (373–375), including the novel antiapoptotic roles of mito-
chondrial hexokinases (376), cannot be simply extrapolated from
our knowledge of classical roles in cellular metabolism.
These enzymes and their pathways constitute broad cat-
egories of potential targets for disruption that could serve to
enable the observed metabolic phenotypes of cancer cells (377).
Although metabolic control is broadly distributed over all indi-
vidual steps for a given pathway (352,378), the most obvious
targets for conceptual and experimental scrutiny involve major
rate-controlling elements of pathways capable of supporting the
anabolic and catabolic needs of rapidly proliferating cancercells.
Numerous studies have demonstrated cancer-associated
changes in metabolism or related gene expression (26). We
looked at acrolein, copper, cypermethrin, diazinon, hexythi-
azox, iron, malathion and rotenone as chemicals that had been
reported to show relevant disruptive potential (51,379–383);
however, the toxicological data that are available for many
suspected or known environmental disruptors, generally lacks
mechanistic information regarding their potential roles as
determinants of the observed metabolic hallmarks of cancer.
Even prior metabolic screening platforms, including tetrazolium
reduction assays, have limited specicity and can be profoundly
inuenced by experimental screening conditions. Unfortunately,
standardized chemical screening has typically not been con-
ducted under controlled or limiting substrate conditions that
would directly inform our understanding of the functional rel-
evance of observed changes. None have established unambigu-
ous causal relationships between specic chemical exposures
and the parallel or sequential development of dysregulated
metabolism of cancer in the same model, and most observed
changes in gene expression with potential relevance to cancer
metabolism have not been accompanied by validating func-
tional studies.
Angiogenesis
Angiogenesis, the process of formation of new blood vessels
from existing blood vessels, is a critical process for normal organ
function, tissue growth and regeneration (e.g. wound healing,
female menstruation, ovulation and pregnancy) as well as for
pathological conditions (e.g. cancer and numerous non-cancer-
ous diseases, such as age-related macular degeneration, dia-
betic retinopathy, rheumatoid arthritis, endometriosis, diabetes
and psoriasis) (384,385).
Tumor angiogenesis is an early critical event for tumor
development: Atumor cannot grow beyond 1 mm3 (by estimate)
without angiogenesis (386). Tumor growth, invasion and metas-
tasis depend on blood vessels and neovascular development
to provide nutrients, oxygen and removal of metabolic waste
as tumors grow in primary sites, invade adjacent tissues and
metastasize to distant organs (387,388). Inhibition or eradication
of tumor angiogenesis by antiangiogenic inhibitors (389,390) or
by antineovascular agents (such as vascular-disrupting agents
(391–393) and fVII/IgG Fc (394), the latter also called ICON (395–
397)) can treat pathological angiogenesis-dependent diseases,
including cancer and many non-cancerous diseases.
Under physiological conditions, angiogenesis is well bal-
anced and controlled by endogenous proangiogenic factors
and antiangiogenic factors. Factors produced by cancer cells
can shift the balance to favor tumor angiogenesis. Such factors
include vascular endothelial growth factor (VEGF) and tissue
factor (TF). VEGF, one of the most potent proangiogenic factors
produced by cancer stem cells and cancer cells, binds to vascular
endothelial cells via its receptor VEGFR, initiating VEGF/VEGFR
intracellular signal transduction pathways and activating many
gene transcriptions and translations toward angiogenesis. TF
is a transmembrane receptor (398) not expressed on quiescent
endothelial cells (399,400). Upon stimulation of VEGF, TF is selec-
tively expressed by angiogenic endothelial cells, the inner layer
of the tumor neovasculature. Thus, TF is a specic biomarker
for tumor angiogenesis (408–410). Both of the membrane-bound
W.H.Goodson et al. | S271
receptors VEGFR and TF can mediate separate intracellular sign-
aling pathways that contribute to tumor angiogenesis.
Environmental exposures can promote tumor development,
but the role of chemicals in tumor angiogenesis, particularly the
role of low-dose non-carcinogens, is largely unknown. Some food-
use pesticides that are non-genotoxic act as tumor promoters,
and other chemicals affect various hallmarks such as apoptosis,
proliferative signaling, evading growth suppression, enabling
replicative immortality, metastasis, avoiding immune destruc-
tion, tumor-promoting inammation and deregulating cellular
energetics—in addition to tumor angiogenesis.
Chemical disruptors that may promote tumor angiogenesis
included diniconazole, 2,2-bis-(p-hydroxyphenyl)-1,1,1-trichlo-
roethane (HPTE), methylene bis(thiocyanate), peruorooctane
sulfonate (PFOS), Ziram, biphenyl, chlorothalonil, tributyltin
chloride and bisphenol AF. Diniconazole (pesticide), for exam-
ple, targets certain angiogenic molecules (CXCL9, CXCL10,
MMP1, uPAR, VCAM1 and THBD) in vitro (29). MXC (the parent
compound to HPTE) induces histological expression of angio-
genic factors such as VEGF, VEGFR2 and ANG1 in rat pituitary
and uterus (39), and exposure to PFOS induces actin lament
remodeling, endothelial permeability changes and ROS produc-
tion in human microvascular endothelial cells (41). Ziram can
induce angiogenesis through activation of MAPK and decreases
cytolytic protein levels in human natural killer (NK) cells
(404,405).
Tissue invasion and metastasis
Tissue invasion and metastasis are also key processes of tumor
progression. In normal cells, E-cadherin holds the epithelial
cells together as a society of cells that are well differentiated
and otherwise quiescent (406). Carcinomas constitute almost
90% of cancers and upon oncogenic transformation, the process
of tissue invasion and metastasis begins with the down-regu-
lation of E-cadherin. Concomitant with this down-regulation of
E-cadherin is the conversion of epithelial to mesenchymal cells
(EMT) (407). The transcription factors that control EMT, such as
snail, slug, Twist and Zeb1/2, are some of the best-character-
ized signaling molecules in biology (408,409). During the pro-
cess of EMT, a number of inammatory cells are attracted to
the growing tumor mass (410). Upon attaining mesenchymal
characteristics, tumor cells are able to move out of their natu-
ral environment, aided by cross talk between them and stromal
cells, resulting in the secretion of matrix degrading enzymes
such as matrix metalloproteinases (411). This process is accel-
erated by chronic inammation mediated by NF-κB (410). Other
invasion mediating molecules include hepatocyte growth fac-
tor, secreted mainly by tumor-associated broblasts to signal
metastatic cells to move upon their interactions with their cell
surface receptor cMet (412).
Attracted by chemokines, metastatic cells move to the near-
est blood vessel or lymphatic vessel, where they complete the
process of intravasation, entering the capillaries and are then
transported to the capillary bed in their colonized site or new
environment (413). In this new location, tumor cells undergo
extravasation where they come out of the capillaries or lym-
phatic vessels, most likely again following the cues emanat-
ing from the chemokines in their new microenvironments. To
survive in their new home, they may have to revert back and
assume the cuboidal morphology of epithelial cells-undergo-
ing the reversal of EMT otherwise known as mesenchymal to
epithelial transition (414). At this point, they may remain dor-
mant for a very long time until conditions for their division and
growth become favorable.
Mounting evidence supports the involvement of exosomes
(nano-vesicles secreted by tumor or cancer-associated bro-
blasts) in adhesion and motility of metastatic cells. The secretion
of exosomes is accelerated by increases in intracellular cal-
cium ions, and low-dose environmental mixtures that increase
intracellular calcium may promote the secretion of exosomes
and the subsequent invasion and metastasis processes of the
tumorcells.
Environmental chemicals, such as tetrabromobisphenol
Aand its metabolites, BPA and tetrabromobisphenol Adimethyl
ether, which mediate the activation of EMT enzymes or drive
their synthesis, may also contribute to the process of tissue
invasion (415). Low-dose exposure to hexavalent chromium may
accelerate the EMT transition (416). Other contributing factors
may also be low-dose environmental contaminants, such as for-
maldehyde, or bacteria, e.g. H.pylori, that drive the transcription
of NF-κB and exacerbate the process (417,418).
Tumor microenvironment
The tumor microenvironment is a complex mix of cells in addi-
tion to tumor cells themselves; it is constructed of a complex
balance of blood vessels that feed the tumor, the extracellu-
lar matrix that provides structural and biochemical support,
signaling molecules that send messages, soluble factors such
as cytokines and many other cell types. Tumors can inuence
the microenvironment and vice versa. The micro-environmental
reaction to early tumor cells begins with the recruitment and
activation of multipotent stromal cells/mesenchymal stem cells,
broblasts, endothelial cell precursors, antigen-presenting cells
such as dendritic cells (DCs) and other white blood cells. All of
these tumor stromal cells secrete a variety of growth factors and
chemokines that, together with the tumor cells and secreted
factors, culminate in the generation of the tumor microenviron-
ment (419–422).
The tumor microenvironment is important because any cell
within this process has the potential to be affected by carcino-
gens, either alone or in mixtures, or by the inammation that
results from the carcinogenic insult (423). Although often asso-
ciated with infection, chronic inammation can be caused by
exposure to carcinogens such as irradiation or environmental
chemicals. Carcinogenesis can also be fostered via effects on the
tissue context surrounding preneoplastic lesions. For example,
transplantation experiments of preneoplastic cells have clearly
documented that a growth-constrained tissue microenviron-
ment can promote the growth and progression of preneoplastic
cell populations (424).
Several compounds appear to inuence the complex het-
erogeneity that forms the support network for cancer growth.
The exposure to nickel chloride has been associated with the
generation of ROS and inammation (425). ROS are impor-
tant because they can stimulate the induction of angiogenesis
growth factors, such as VEGF, and can promote cell proliferation
and immune evasion and play a role in cell survival (57,426–428).
Prenatal exposure to BPA in experimental animals disrupts ERα
and triggers angiogenesis, and other BPA exposure studies have
demonstrated that BPA interplays with cell proliferation (226),
genomic instability (429), inammation (430) and cell immor-
talization (431). Butyltins, and specically tributyltin, which is
suspected to act as an endocrine disruptor, have been found
to inhibit the cytotoxic activity of NK cells (432), affect inam-
mation (432) and disrupt membrane metalloproteinases (432).
Cooperatively, disruption of these processes can lead to prolif-
eration, migration and angiogenesis. Methylmercury (MeHg) is a
S272 | Carcinogenesis, 2015, Vol. 36, Supplement 1
neurotoxic compound deriving from metallic mercury through
bacteria-supported metabolism in an aquatic environment.
Bio-concentration in sh and shellsh poses a risk for sensi-
tive population categories such as pregnant women and infants.
MeHg-induced ROS production may be involved in inamma-
tion and apoptosis (433) as well as endothelial cell cytotoxicity
(434). We also looked at paraquat, which may also have rele-
vance for the tumor microenvironment via its role in oxidative
stress (435,436).
Avoiding immune destruction
The concept of immune surveillance suggests that the host
immune system could identify tumor cells and destroy them.
If this is true, tumor cells need to be poor stimulators of or
challenging targets for the host immune system. To provide
an effective immune response, multiple types of the cells are
involved within innate and adaptive immune ‘arm’ with some
cells (e.g. DCs and the NK cells) ‘bridging’ these two types
of immunity (437). To avoid a strong immune response of the
host, the expression of tumor antigens may be down-regulated
or altered (resulting in decreased or impossible recognition of
malignant cells) (438) and various soluble factors and cytokines
may be released resulting in subverted effectiveness of antitu-
mor immune response (439–441). Tumor cells can also escape
host immune response by inducing apoptosis in activated T
cells (442).
Multiple genes involved in immune evasion mechanisms
and, therefore, can interfere with chemical exposures from
anthropogenic environment: ADORA1 (adenosine A1 receptor),
AKT1 (v-akt murine thymoma viral oncogene homolog 1), CCL2
(chemokine C-C motif ligand 2), CCL26 (chemokine C-C motif
ligand 26), CD40, CD69, COL3A1 (type III collagen of extracellu-
lar matrix), CXCL10 (also called interferon-inducible protein-10),
CXCL9 (monokine induced by interferon-γ), EGR1 (early growth
response protein 1), HIF-1α (hypoxia-inducible factor), IGF1R
(insulin-like growth factor 1 receptor) and interleukins (IL) such
as IL-1α and IL-6. Based on available studies, several candi-
date signaling pathways that are related to the host immune
response can be identied for further study; e.g. the pathways
involving PI3K/Akt, chemokines, TGF-β, FAK, IGF-1, HIF-1α, IL-6,
IL-1α, CTLA-4 and PD-1/PDL-1.
Biologically disruptive environmental chemicals can affect
the host immune responses as follows: (i) if a certain chemical
is immunotoxic, and, in particular, if it affects activity of DCs, T
cells or NK cells, it is also likely to affect tumor immuno-surveil-
lance and enable malignant growth to proceed; (ii) if a chemi-
cal targets the immune system, it can increase the cancer risk
related to other factors/exposures; (iii) exposures to certain tox-
ins or toxicants can dramatically increase the number of can-
cerous cells and impact immuno-regulatory signals suppressing
the mechanisms of immune control. Collectively, these sorts of
actions suppress the immune system, so it cannot be effectively
stimulated and cannot eliminate tumor cells, thus allowing
some tumor cells to escape and metastasize.
We looked at several groups of environmentally ubiqui-
tous chemicals such as pesticides and personal care products
that might potentially interrelate with mechanisms of tumor
immuno-surveillance. Although none of them are recognized as
human carcinogens (443–445), the research on these chemicals
and their interactions with the immune response may be valu-
able. For example, the fungicide maneb is a cortisol disruptor
(446) that has shown a wide spectrum of potential effects on
multiple pathways, including some that are relevant to immune
evasion (139,156–158,447). By comparison, pyraclostrobin and
uoxastrobin (448) interfere with a narrower spectrum of can-
cer hallmarks (36,449–452). Atrazine has also shown potential
to impact immune system evasion by directly targeting matura-
tion of DCs and decreasing the levels of major histocompatibility
complex classImolecules (243,453). The insecticides pyridaben
and azamethiphos can also both be disruptive to immuno-sur-
veillance (139,140,454,455).
Commonly used in personal care products, triclosan and BPA
(456), are endocrine disruptors (457–459) that are often detected
in waters downstream in urban areas (460,461). In addition to
immune evasion mechanisms (36,142,145), they interfere with
wide spectrum of cancer-related mechanisms (36,173,429,462–
464). DEHP (472) is also an endocrine disruptor (466,467) that can
impact multiple hallmarks such as immune evasion, resistance
to cell death, evasion of antiproliferative signaling, sustained
proliferative signaling and tumor-promoting inammation
(36,288,468,469).
Knowing whether or not cumulative low-dose exposures to
these chemicals interfere with the host immune response can
help to stimulate further studies (e.g. on screening of lesions at
the pre-malignant stage of tumor development) to determine
the inuence of such exposures on host immunity and to evalu-
ate their potential to increase the risk of tumor cell survival.
Dose-response characterizations andLDE
For all the chemicals selected and target sites for disruption that
were identied, dose-response characterization results and/or
relevant low-dose research evidence were reviewed and catego-
rized using the criteria mentioned in the Materials and meth-
ods. Table1 sets out these results and the supporting references.
In total, 85 examples of environmental chemicals were
reviewed (for specic actions on key pathways/mechanisms that
are important for carcinogenesis) and 59% of them (i.e. 50/85)
were found to exert LDE (at levels that are deemed relevant given
the background levels of exposure that exist in the environment)
with 15 of the 50 demonstrating their LDE in a non-linear dose-
response pattern. Indeed, all of the teams selected at least one or
more disruptive chemicals that exerted their effects on the tar-
get sites at low-dose levels. In contrast, only 15% of the chemicals
reviewed (i.e. 13/85) showed evidence of a threshold.
The remaining 26% of the chemicals reviewed (i.e. 22/85)
were categorized as ‘unknown’. Some of these chemicals (5 of
the 22) had been tested using human primary cell data from
ToxCast and had showed statistically signicant activity across
a full range of doses against the specied targets (i.e. they
were active even at the lowest test concentrations of ~0.01µM).
However, even though no threshold could be discerned for these
chemicals, we did not characterize them as having LDE (because
it was not clear that the lowest test concentrations were low
enough to be equated to levels of exposure that are normally
seen in the environment).
Evidence of cross-hallmark relationships
Teams then evaluated the chemicals selected and target sites for
disruption for known effects on the other cancer hallmark path-
ways. Evidence in the literature that showed procarcinogenic
actions or anticarcinogenic actions in other hallmark areas were
reported, and in instances where no literature support was found,
this was documented as well. The same approach was used for
the chemicals that were reviewed. Asample of these cross-hall-
mark results is provided in Table 2—Sample of cross-hallmark
relationships of target pathways/mechanisms and in Table 3
Cross-hallmark relationships of selected chemical disruptors.
W.H.Goodson et al. | S273
Table2. Sample of cross-hallmark relationships of target pathways/mechanisms
Insensitivity to antigrowth signals (targets) Antigrowth Dysreg metab Gen instab Angio Cell death Immun Immort Prolif Metas Inamm Tumor micro PRO ANTI MIX
p53 n/a +/− - +/− − +/− + + 2 5 3
pRB n/a +/− 0 + + 2 6 1
TGF-βn/a + + + + + + + 7 3 0
LKB1 n/a + + +/− 0 0 + + + 5 2 1
Connexins n/a 0 0 0 0 +/− + + 2 3 1
Contact inhibition n/a +/− 0 0 + 0 + 2 4 1
One set of results (from the insensitivity to antigrowth signals review) is shown here without references to support a discussion on the range of effects that have been reported for the selected targets in each article. Specic
references supporting these effects for any given hallmark area can be found in the individual reviews within this special issue. Cross-hallmark relationships are reported in the rst 11 columns of the table—table
heading abbreviations are as follows:gen instab, genetic instability; dysreg metab, dysregulated metabolism; antigrowth, insensitivity to antigrowth signals; angio, angiogenesis; cell death, resistance to cell death; immun, avoid-
ing immune destruction; immort, replicative immortality; prolif, sustained proliferative signaling; metas, tissue invasion and metastasis; inamm, tumor-promoting inammation; tumor micro, tumor microenvironment. The
number of procarcinogenic (PRO), anticarcinogenic (ANTI) and mixed (MIX) (i.e. procarciniogenic and anticarcinogenic reports) cross-hallmark relationships for each target have been summed and are reported in the last three
columns of the table. Target pathways/mechanisms for each hallmark area were evaluated by each team for known effects in other cancer hallmark pathways. Targets that were found to have anticarcinogenic actions in another
hallmark area were indicated with ‘−’, whereas targets that were found to have procarcinogenic actions in another hallmark area were indicated with ‘+’. In instances where reports on relevant actions in other hallmark areas
were mixed (i.e. reports showing both procarcinogenic potential and anticarcinogenic potential), the symbol ‘+/−’ was used. Finally, in instances where no literature support was found to document the relevance of a target in a
particular aspect of cancer’s biology, we documented this as ‘0’.
Table3. Cross-hallmark relationships of selected chemical disruptors
Insensitivity to antigrowth signals (disruptors) Antigrowth Dereg metab Gen instab Angio Cell death Immun Immort Prolif Metas Inamm Tumor micro PRO ANTI MIX
BPA n/a + + + +/− 0 + + + + 0 7 0 1
DDT n/a 0 + + + + + + 0 + 0 7 0 0
Folpet n/a 0 + 0 + 0 0 + 0 + 0 4 0 0
Atrazine n/a 0 + 0 0 0 0 + 0 + 0 3 0 0
One set of results (from the insensitivity to antigrowth signals review) is shown here without references to support a discussion on the range of effects that have been reported for the selected disruptors in each review. Specic
references supporting these effects for any given hallmark area can be found in the individual reviews within this special issue. Cross-hallmark relationships are reported in the rst 11 columns of the table—
table heading abbreviations are as follows:gen instab, genetic instability; dereg metab, dysregulated metabolism; antigrowth, insensitivity to antigrowth signals; angio, angiogenesis; cell death, resistance to cell death; immun,
avoiding immune destruction; immort, replicative immortality; prolif, sustained proliferative signaling; metas, tissue invasion and metastasis; inamm, tumor-promoting inammation; tumor micro, tumor microenvironment.
The number of procarcinogenic (PRO), anticarcinogenic (ANTI) and mixed (MIX) (i.e. procarciniogenic and anticarcinogenic reports) cross-hallmark relationships for each target have been summed and are reported in the last
three columns of the table. Prototypical chemical disruptors selected by each team were evaluated for reported actions in other cancer hallmark pathways. Disruptors that were found to have anticarcinogenic actions in a
particular hallmark area were indicated with ‘−’, whereas disruptors that were found to have procarcinogenic actions in a particular hallmark area were indicated with ‘+’. In instances where reports on relevant actions in other
hallmarks were mixed (i.e. reports showing both procarcinogenic potential and anticarcinogenic potential), the symbol ‘+/−’ was used. Finally, in instances where no literature support was found to document the relevance of a
chemical in a particular aspect of cancer’s biology, we documented this as ‘0’. Specic references supporting these effects for any given area can be found in the individual reviews in this special issue.
S274 | Carcinogenesis, 2015, Vol. 36, Supplement 1
Table4. Aggregated evidence of cross-hallmark effects for selected pathways/mechanisms
Key targets Originating review Procarcinogenic Anticarcinogenic Mixed
Adenosine A1 receptor (ADORA1) ISE 3 1 0
AhR Ang 7 0 2
SPS 7 0 2
Bcl-2/p53 RCD 6 2 1
Cell cycle/cell division: spindle defect GI 6 0 1
Checkpoint kinase 1 and checkpoint kinase 2 (Chk1/2) EAS 4 3 1
Chemokine (C-C motif) ligand 2 (CCL2) Ang 8 1 0
Chemokine (C--XC motif) ligand 10 (CXCL10) Ang 4 1 1
Chemokine (C--XC motif) ligand 9 (CXCL9) Ang 3 2 0
Chemokine signaling pathway (CCL2, CCL26, CXCL9, CXCL10) ISE 6 1 2
Chronic oxidative stress TM 6 1 1
Clock-genes-mediated metastasis TIM 5 1 0
Collagen type III (COLIII) Ang 3 0 0
Contact inhibition EAS 4 3 0
cSrc/Her1/STAT5B/ERK1/2 TIM 3 1 1
Cyclin D, IL8, CXCL SPS 4 0 2
Cyclooxygenase expression and stimulation calcium signaling in migration. TIM 8 1 0
Cyclooxygenase-2 TPI 8 1 0
DNA damage signaling: disturbed by Redox signaling (NF-κB, Nrf, EGR) GI 8 1 0
DNA repair pathways GI 6 2 1
Eck fatty acid metabolism DM 6 1 2
Electron transport chain complexes II and IV DM 3 2 0
Epidermal growth factor receptor SPS 6 0 1
Epigenetic pathways
Disturbed miRNA binding GI 6 0 2
DNA methylation GI 7 0 1
Histone acetylation GI 6 1 1
EMT TIM 5 0 1
EMT, catenin-Wnt pathway TIM 6 1 1
ErbB-2/HER-2 tyrosine kinase RCD 6 1 0
ERK/MAPK RCD 8 2 0
Estrogen receptor TPI 5 3 1
Estrogen receptor α (binding to) RCD 5 1 1
Gap junction connexins EAS 2 2 2
GJIC RCD 2 1 1
Gluconeogenesis DM 5 3 0
Glycolysis DM 8 1 0
Hexokinase 2 DM 6 1 0
H-Ras SPS 6 1 2
Hypersecretion of luteinizing hormone by gonadotroph cells in pituitary gland RCD 2 1 0
HIF-1-α pathway ISE 8 0 2
Inducible nitric oxide synthase TPI 6 1 0
IGF-1 signaling pathway ISE 6 2 1
Intercellular adhesion molecule 1 (ICAM1) Ang 6 3 0
W.H.Goodson et al. | S275
Key targets Originating review Procarcinogenic Anticarcinogenic Mixed
IL-6 TPI 7 0 0
IL-6 expression, improper DC maturation and polarization TM 5 2 0
Jun/Fos/AP1 SPS 4 1 3
Lipid metabolism/cholesterol metabolism DM 4 2 1
Liver kinase B1 (Lkb1) EAS 4 2 2
MMP 1 Ang 6 1 0
MMP-9 activation TIM 5 1 1
Mitochondrial function GI 5 2 2
MAPK RCD 9 0 1
mTOR activation DM 7 1 1
mTOR inactivation RI 3 6 1
NK cell inhibition TM 4 3 0
NF-κB TPI 4 2 0
Oxidative stress and IL-6 production TM 3 1 1
P16/p53 RCD 4 4 0
P53 inactivation EAS 10 0 0
RCD 10 0 0
RI 10 0 0
PPAR SPS 5 2 0
PPAR-αRCD 3 3 1
PI3K/Akt signaling pathway ISE 9 0 1
Pyruvate dehydrogenase (PDH) DM 1 5 0
ROS (increase) DM 6 0 4
ROS and cellular stress TM 5 0 4
Retinoblastoma protein (pRb) inactivation EAS 9 0 0
RI 9 0 0
Steroid hormone receptors SPS 5 0 1
Telomerase activation RI 9 1 0
Telomere loss GI 4 4 0
The tricarboxylic acid cycle DM 5 4 0
Thrombomodulin Ang 2 3 0
Transforming growth factor βEAS 6 3 1
Tumor necrosis factor αTPI 8 0 1
Urokinase receptor (uPAR) Ang 6 2 0
Vascular cell adhesion molecule 1 (VCAM1) Ang 6 0 0
Aggregated number of procarcinogenic actions, anticarcinogenic actions and instances where mixed actions (i.e. procarciniogenic and anticarcinogenic) where cross-hallmark effects have been reported (for each pathway/
mechanism across the full range of hallmark domains—i.e. from all of the areas covered by the reviews in this special issue)—see samples of this data in Table2. Note: fully referenced data for these cross-hallmark effects can
be found in each of the reviews in this special issue. ANG, angiogenesis; DM, dysregulated metabolism; EAS, evasion of antigrowth signaling; GI, genetic instability; ISE, immune system evasion; RCD, resistance to cell death; RI,
replicative immortality; SPS, sustained proliferative signaling; TIM, tissue invasion and metastasis; TM, tumor microenvironment; TPI, tumor-promoting inammation.
Table4. Continued
S276 | Carcinogenesis, 2015, Vol. 36, Supplement 1
Note that Tables 2 and 3 contain just a single set of unref-
erenced results from the review on the hallmark insensitivity to
antigrowth signals. This is intended only to illustrate the catego-
ries of cross-hallmark effects that were reviewed and to show
how they were presented. Fully referenced results for each hall-
mark area can be found in each of the individual reviews within
this specialissue.
The decision to review target sites for disruption and proto-
typical disruptors for cross-hallmark effects was driven by the
fact that many individual studies and reviews of chemical expo-
sures fail to account systematically for the spectrum of inci-
dental actions that can result from exposures to a single given
chemical. It was our belief that this approach constitutes a bet-
ter way to ensure that we had assembled a reasonably complete
view of the literature (i.e. where any sort of evidence of cross-
hallmark activity had been reported). Future research will likely
involve empirical testing of mixtures, so we wanted to create a
heuristic that could serve as a starting point for other research-
ers who might be considering such research.
For researchers focused on low-dose exposure research
intended to produce carcinogenesis, we anticipated that there
would be interest in chemicals that had been reported to exhibit
a large number of procarcinogenic actions across a number of
hallmarks and we anticipated that a lack of anticarcinogenic
potential would be important to identify (as targets or approaches
that exert anticarcinogenic actions would potentially represent
a confounding inuence/factor in empirical research aimed at
the identication of carcinogenic synergies). To that end, Table4
provides a summary of the aggregated number of procarcino-
genic actions, anticarcinogenic actions and instances where
mixed actions (i.e. procarciniogenic and anticarcinogenic) have
been found for each pathway/mechanism (across the full range
of hallmark domains—i.e. from all of the areas covered by the
reviews in this special issue). Similarly, Table5 provides a sum-
mary of the aggregated number of procarcinogenic actions,
anticarcinogenic actions and mixed actions (i.e. procarcinogenic
and anticarcinogenic), where cross-hallmark effects have been
reported for each chemical (across the full range of hallmark
domains—i.e. from all of the areas covered by the reviews in this
special issue).
Note that, in some instances, the underlying evidence used
to support the indication of cross-hallmark relationships was
robust, consisting of multiple studies involving detailed in-vitro
and in-vivo ndings. In other instances, the underlying evidence
that was used to report the existence of a cross-hallmark rela-
tionship was quite weak (e.g. consisting of only a single in-vitro
study involving a single cell-type). The selected prototypical dis-
ruptors are likely biased towards agents that have been exten-
sively studied, and not necessarily those that will prove to be
the most important biologically. Finally, there are examples of
chemicals that are known to exert different effects at different
dose levels, but dose levels were not used to discriminate when
gathering evidence of cross-hallmark effects. So, the referenced
cross-validation results in the individual tables (reported in the
many reviews within this special issue) should be seen only as
a starting point for those who are pursuing mixtures research
(e.g. references would need to be further scrutinized to deter-
mine whether or not the dose levels noted for specic results
are suitable points of reference for the type of research that is
being undertaken).
Particular attention should also be given to results related to
the endocrine system due to mechanistic complexity. For exam-
ple, xeno-estrogen compounds are typically compared with
estradiol based on binding afnity strength. However, many
xeno-estrogens that are ‘weak’ by this measure can alter the
steroidogenic cascade (e.g. signicantly up-regulate the activ-
ity of P450 aromatase, the enzyme that increases intracellular
estradiol synthesis within estrogen-sensitive cells (470–473) or
alter levels of ERα or the ratio of ERα:ERβ (260)). In other words, a
weak xeno-estrogen can stimulate the production of estradiol, a
potent endogenous carcinogen (474) or alter the receptors with
which a cell will respond to estrogen.
Nonetheless, given that the overarching goal in this project
was to create a foundation that would allow researchers to look
systematically across the literature in each of these areas, the
tables should serve as a useful starting point as long as they
are approached with these caveats in mind. We believe that
this heuristic will be useful to consider synergies that might be
anticipated in testing that involves certain target sites for dis-
ruption and/or mixtures of chemical constituents that are being
considered for procarcinogenic effects. Future research efforts
to improve this approach could involve a large-scale collabora-
tive effort to generate high-quality in-vitro data and low-dose in-
vivo data in a range of predened tissues.
Discussion
Getting to Know Cancer hosted the initial project meeting in
Halifax, Nova Scotia giving participants an opportunity to have
presentations, break-out sessions, and chances for conversation
and debate among experts who came from a range of different
disciplines. Cancer biologists with specialized expertise in areas
related to individual hallmarks met with specialists from other
areas such as environmental health, toxicology and endocrinol-
ogy. Although some researchers in the eld of environmental
health are cancer scientists in their own right, many conference
participants commented on the novelty of having an oppor-
tunity to work so closely with cancer biology specialists. As a
result, many interdisciplinary barriers were removed and the
discussions that ensued were challenging but productive.
At the outset, participants overwhelmingly agreed that the
Hallmarks of Cancer provides a useful organizing heuristic for
systematic review of ways that biologically disruptive chemicals
might exert procarcinogenic and anticarcinogenic inuences in
biological systems. Most of the individual writing teams were
then readily able to identify ubiquitous environmental contami-
nants with disruptive potential in their respective areas of study.
The only teams that had signicant challenges in this regard
were the ones that focused on the bypassing of senescence (i.e.
replicative immortality) and dysregulated metabolism, both being
areas of cancer research that have not yet received a lot of atten-
tion from researchers in the eld of toxicology.
Considerable discussion was devoted to the criteria that
were used to select prototypical disruptors from the long list of
known potential contaminants. Indeed, it seems that much of
the population is now exposed to a wide variety of exogenous
chemicals that have some disruptive potential, but we did not
have any intention of implicating any of the selected chemicals
as being carcinogenic per se. It was simply agreed that chemi-
cals would be chosen that met the basic criteria and that then
could be used as ‘prototypical’ disruptors. In other words, the
chemicals that were selected for this review were not deemed to
be the most important, and they were not selected to somehow
imply (based on current information) that they are endanger-
ing us. Rather, we simply wanted to illustrate that many non-
carcinogenic chemicals (that are ubiquitous in the environment)
have also been shown to exert effects at low doses, which are
highly relevant to the process of carcinogenesis. We also wanted
W.H.Goodson et al. | S277
Table5. Aggregated evidence of cross-hallmark effects for selected chemical disruptors
Chemicals Originating review Procarcinogenic Anticarcinogenic Mixed
12-O-Tetradecanoylphorbol-13-acetate SPS 5 1 0
HPTE ANG 4 0 0
Acetaminophen RI 0 4 2
Acrolein DM 3 3 3
Acrylamide GI 3 1 1
Atrazine ISE 3 0 1
EAS 4 0 1
TPI 3 0 1
Azamethiphos ISE 1 0 0
Benomyl GI 0 3 1
Benzo(a)pyrene SPS 8 1 0
Biorhythms TIM 3 2 0
Biphenyl ANG 2 2 1
BPA EAS 6 0 1
GI 6 0 1
ISE 7 0 1
RCD 7 0 0
SPS 6 0 1
TIM 7 0 1
TM 7 0 1
TPI 6 0 1
Bisphenol AF ANG 5 1 0
Butyltins (such as tributyltin) TM 4 2 0
C.I.solvent yellow 14 ANG 4 0 0
Carbendazim GI 0 2 1
Carbon black GI 5 1 0
Chlorothalonil ANG 5 1 0
RCD 5 0 0
Cobalt GI 5 2 0
Copper DM 6 0 3
Cotinine RI 4 1 0
Cypermethrin DM 5 0 0
DDT EAS 6 0 0
Diazinon DM 2 3 0
Dibutyl phthalate RCD 4 0 0
Dichlorvos RCD 4 0 0
DEHP ISE 4 0 1
RCD 4 0 0
Diniconazole ANG 2 0 0
Fluoxastrobin ISE 2 1 0
Folpet EAS 2 1 0
Hexachlorobenzene TIM 5 2 0
Hexythiazox DM 0 0 0
Imazalil SPS 3 1 0
Iron DM 5 1 3
TIM 5 1 2
Lactofen SPS 2 0 0
Lead GI 3 1 0
RI 3 1 0
Lindane RCD 5 0 0
Linuron RCD 2 0 0
Malathion DM 5 0 0
Maneb ISE 4 2 0
Mercury GI 3 2 1
MXC RCD 3 0 0
Methylene bis(thiocyanate) ANG 2 1 0
MeHg TM 5 2 0
Na-selenite RI 0 4 2
Nickel GI 6 1 1
TM 6 1 1
Nickel chloride RI 6 0 2
Nitric oxide RI 5 2 2
4-NP TPI 2 1 0
Oxyuorfen RCD 4 0 0
S278 | Carcinogenesis, 2015, Vol. 36, Supplement 1
to lay out a heuristic framework that would be helpful for other
researchers who are interested in considering these and other
chemicals as potential constituents for low-dose mixtures
research.
LDE, chemical mixtures and carcinogenicity
Although we did not specically ask the teams to focus on
disruptive chemicals that were known to exert LDE, the sum-
mary of dose-response characterizations for the chemicals
that were selected by these teams is dominated by chemi-
cals (i.e. 50/85) that have been shown to produce LDE, and
15 of the 50 showed evidence of a non-linear dose-response.
Surprisingly, only 15% of the chemicals reviewed (i.e. 13/85)
showed evidence of a threshold. We believe that this helps to
validate the idea that chemicals can act disruptively on key
cancer-related mechanisms at environmentally relevant lev-
els of exposure.
Historically, the axiom ‘the dose makes the poison’ has had
some merit, so many people remain skeptical about the idea
that adverse outcomes can result from minute exposures to
commonly encountered chemicals. But we are now at a point in
time where our knowledge of the biology of cancer has advanced
considerably, and we know that carcinogenesis can begin when
key events have occurred in a single cell, between cells or in
the surrounding microenvironment. So the idea that LDE from
many environmental chemicals (acting together) might serve
to instigate, support or fully enable carcinogenesis, no longer
appears to be an unreasonable assertion.
At this stage, we are not making any assumptions about
whether or not future empirical research will nd support for
this hypothesis, nor are we assuming that this a signicant
problem. We are simply impressed by the fact that we are now
starting to see evidence of a wide range of LDE (that are directly
related to carcinogenesis) that can be exerted by chemicals that
are ubiquitous and unavoidable in the environment. As a result,
we are compelled to explore and consider this possibility.
In-utero exposures and transgenerational effects
Additionally, a number of the teams cited in-utero exposure
studies in their reviews and presented evidence on transgen-
erational effects. Although this detail is not fully captured in
the team summaries offered in this capstone paper (please see
the individual reviews in this special issue for complete details),
these effects are important to acknowledge. For example, the
inammation team noted that transient early life exposures in
utero to vinclozolin have been linked to both adult-onset dis-
ease and transgenerational disease that involves inamma-
tion. Similarly, the immune system evasion team reported that
there is increasing evidence from animal studies that in-utero
or neonatal exposures to BPA are associated with higher risk of
immune system dysregulation that may develop later inlife.
Taken together, these and other similar types of examples
raise intriguing possibilities about vulnerabilities at the popu-
lation level, and the contributions that in utero and early life
exposures to mixtures of those chemicals might make towards
cancer susceptibility. Single-generation experimental models
are inadequate to detect this sort of disruptive activity (for expo-
sures to a given chemical or to mixtures of chemicals), but these
sorts of effects may increase cancer risks by promoting and/or
enabling tumorigenesis.
The interplay between genetic factors and
environmental factors
Given the number of key cancer-related mechanisms that can
apparently be disrupted by chemicals that are commonly found
in the environment, and the possibility that in-utero and/or early
life exposures may also contribute to population vulnerabil-
ity, the interplay between genetic factors and environmental
Chemicals Originating review Procarcinogenic Anticarcinogenic Mixed
Paraquat GI 4 2 0
TM 4 2 0
PFOS ANG 4 1 0
SPS 4 1 0
Phosalone SPS 1 1 0
Phthalates TIM 6 0 1
TPI 6 0 1
PBDEs TPI 2 0 2
Pyraclostrobin ISE 2 1 0
Pyridaben ISE 1 3 1
Quinones GI 1 6 1
Rotenone DM 2 5 1
Sulfur dioxide TIM 5 1 0
Titanium dioxide NPs GI 3 1 1
Tributyltin chloride ANG 3 1 0
Triclosan GI 2 2 1
ISE 3 2 1
Tungsten GI 2 1 1
Vinclozolin TPI 2 1 0
Ziram ANG 3 1 1
Aggregated number of procarcinogenic actions, anticarcinogenic actions and mixed actions (i.e. procarciniogenic and anticarcinogenic) where cross-hallmark effects
have been reported (for each chemical across the full range of hallmark domains—i.e. from all of the areas covered by the reviews in this special issue)—see samples
of this how this data were reported in Table3. Note: fully referenced data for these cross-hallmark effects can be found in each of the reviews in this special issue.
ANG, angiogenesis; DM, dysregulated metabolism; EAS, evasion of antigrowth signaling; GI, genetic instability; ISE, immune system evasion; RCD, resistance to cell
death; RI, replicative immortality; SPS, sustained proliferative signaling; TIM, tissue invasion and metastasis; TM, tumor microenvironment; TPI, tumor-promoting
inammation.
Table5. Continued
W.H.Goodson et al. | S279
factors should also be mentioned. For example, a hereditary
genetic vulnerability (such as mutations to BRCA1/2 genes
which greatly increase the lifetime risk of breast and ovarian
cancer (482)) can predispose someone to a higher risk of cancer.
But many hereditary genetic mutations and somatic mutations
do not result in cancer, presumably because additional actions
(e.g. sustained proliferative signaling) are needed or additional
biological safeguards still need to be suppressed or defeated
(e.g. apoptosis, senescence, immuno-surveillance and so on)
before a fully immortalized cellular phenotype can emerge. In
these instances, cancer may not be assured, but it is easy to
see how the disruptive effects of low-dose exposures to certain
chemicals might act on key pathways/mechanisms and play a
supporting role in the steps involved in carcinogenesis and/or
increase the overall risk of getting cancer.
This same issue applies to other sensitive subpopulations
who might be predisposed to higher levels of cancer risk. In some
instances, vulnerabilities that exist are genetic in nature (e.g.
cancer patients in remission), due to endogenous factors (e.g.
due to obesity) or due to external inuences (i.e. smoking). But
in all cases, the enhanced risks in these subpopulations leave
the affected individuals vulnerable to carcinogenesis. Although
a detailed investigation of this type of interaction is beyond the
scope of this project, it is important to consider that low dose,
disruptive chemical effects on key pathways and mechanisms
in these subpopulations may serve to further enhance cancer
susceptibility, or even fully enable carcinogenesis.
The low-dose carcinogenesis hypothesis
It is important to reiterate that this group has no interest in
implicating any of the chemicals that were reviewed in this
project as individual carcinogens per se. We fully realized
at the outset that much of the evidence in the toxicological
literature that documented the disruptive actions of these
chemicals had been produced under a wide range of differing
experimental circumstances. So it was agreed at the beginning
that we would not make leaps between different lines of evi-
dence nor draw any specic conclusions about chemical mix-
tures that might prove to be carcinogenic. Nonetheless, we are
intrigued by the number of chemicals that we reviewed that
were found to be capable of disruptive LDE on key pathways/
mechanisms across all of the areas that were reviewed. Many
of the environmental chemicals that we chose are well known
as environmental contaminants, but they represent only a
small fraction of the thousands of chemicals that are now
ubiquitous and unavoidable in the environment. So although
we cannot draw any rm conclusions at this stage, we emerge
from this effort with a better understanding of the evidence
that is available to support the merits of our initial hypothesis
(i.e. that low-dose exposures to disruptive chemicals that are
not individually carcinogenic may be capable of instigating
and/or enabling carcinogenesis).
Although the breadth and scope of this review effort was
daunting, we now believe that we have enough supporting evi-
dence to offer a holistic overview of this issue. At a minimum,
we hope that the studies cited in this review, the gaps that we
have identied and the framework that we have proposed for
future research will be useful to researchers who are encour-
aged to explore this hypothesis in greater detail.
The implications for risk assessment
Thirty-ve years ago, the work of Ames and others who fol-
lowed set in motion a quest for individual chemicals as (com-
plete) ‘carcinogens’ that became a dominant paradigm that has
shaped our thinking for decades (226). So dominant has the
focus been on single chemicals, that combinations of chemi-
cals are rarely tested or even considered. For example, although
IARC has focused on extensive monographs of the carcinogenic
nature of individual chemicals, little has been done to evaluate
the possibility of carcinogenic effects attributable to chemical
mixtures except in a few instances where mixtures of concern
are encountered during occupational exposures (e.g. polychlo-
rinated dibenzo-p-dioxins and polychlorinated dibenzofurans)
or as a result of personal and cultural habits (e.g. cigarette
smoke, diesel and gasoline engine exhausts).
But the search for mutagenic carcinogens was never
matched with a corresponding search for chemicals that might
contribute to the promotion of carcinogenesis along with other
chemicals. We now know that individual chemicals can produce
unique disruptions of cellular biology and specic combinations
of non-carcinogenic chemicals have been able to demonstrate
potent carcinogenic effects. Yet, we have only scratched the sur-
face of the biology of mixtures, and we need to look carefully at
the synergistic effects.
In risk assessments, the risks associated with exposures to
mixtures of chemicals are often estimated using relatively sim-
ple, component-based approaches (476). Risk analysts evalu-
ate information regarding the mode of action associated with
individual mixture components and then use either ‘dose addi-
tion’ or ‘response addition’ to predict effects. Dose addition is
an appropriate approach to assess mixtures risks, when the
chemicals of interest act through a common mode of action.
Although response addition assumes that constituent agents
act independently of each other (cause the same outcome via
different modes of action). In general, a dose addition approach
would be appropriate for mixtures risk assessment if we wanted
to consider a series of chemicals that were carcinogenic in their
own right, and if they all produced the cancer by the same mode
of action. The Hallmarks of Cancer framework suggests that
we should be equally, if not more, concerned about mixtures of
chemicals that are not individually carcinogenic but disruptive
in a manner that is collectively procarcinogenic (i.e. potentially
capable of producing carcinogenic synergies when combined
with other chemicals that are acting on the diverse series of
mechanisms involved in carcinogenesis).
With this in mind, there should be concern that the World
Health Organization International Programme on Chemical
Safety (WHO IPCS) has spent the past decade developing a
risk analysis agenda predicated mainly on a ‘Mode of Action’
framework (477–480), where ‘mode of action’ is dened as a
sequence of key events and processes, starting with interaction
of an agent with a cell, proceeding through operational and ana-
tomical changes and resulting in an adverse outcome, in this
case, cancer formation. The OECD guidance on the conduct and
design of chronic toxicity and carcinogenicity (which is followed
by many nations) now also reects this approach (480). This
analysis of risks from cumulative effects of chemical exposures
is restrictive because it suggests that regulators should only
focus on groupings of individual chemicals that are as follows:
(a) known to act via a common sequence of key events and
processes;
(b) known to act on a common target/tissue and
(c) known to produce a common adverse outcome (e.g.
cancer).
So, for example, in the USA, the Food Quality Protection Act pro-
vides legislated guidance on testing for cumulative effects by
using the term ‘common mechanism of toxicity’ (481), which is
interpreted to mean ‘mode of action’ or ‘the major steps leading
to an adverse health effect following interaction of a pesticide
S280 | Carcinogenesis, 2015, Vol. 36, Supplement 1
with biological targets’. Similarly, in Canada, the Pest Control
Products Act requires the government to assess the cumulative
effects of pest control products that have a ‘common mecha-
nism of toxicity’. In the USA, there has also been a tradition of
employing an additional restriction requiring chemical struc-
tural similarity when selecting groups of chemicals to be sub-
jected to mixtures risk assessment (other than a few instances
where whole mixtures have been assessed, e.g. diesel exhaust,
combinations of chemicals that are not similar structurally have
been largely ignored (489)). In light of current knowledge of can-
cer biology, these criteria appear to be inappropriately restric-
tive, and thus demand a number of considerations—as follows:
Cumulative risk assessment should anticipate synergies of chemicals
acting via dissimilar sequences/processes. From the Hallmarks of
Cancer framework, it becomes evident that chemicals that act
via dissimilar pathways/targets or that produce different sorts
of key events and/or employ different processes could very
well produce synergies within carcinogenesis that would be
relevant for cumulative risk assessment purposes. For example,
ethylenediaminetetraacetic acid (EDTA) is a ubiquitous,
presumably non-carcinogenic chemical that disrupts DNA
repair (483,484), and it is well established that it inuences
chromosome breakage by mutagenic agents. In particular,
when applied in combination with chemical mutagens, EDTA
enhances mutagen-induced aberration frequencies and
contributes to genetic instability (485). But within the mode of
action framework, a chemical that is a mutagenic carcinogen,
would not be assessed for the cumulative risks associated with
an additional exposure to a chemical that disrupts DNA repair (a
key layer of cancer defense) because it is not known to produce
a common sequence of key events and processes.
A 2008 report on phthalates and cumulative risk assess-
ment emphasized that the chemicals considered for cumula-
tive risk assessment should be ones that cause the same health
outcomes or the same types of health outcomes, not ones that
cause the health outcomes only by a specic pathway (486).
Similarly, The European Food Safety Authority Panel on Plant
Protection Products and their Residues (PPR Panel) produced a
scientic opinion on the relevance of dissimilar modes of action
and their relevance for cumulative risk assessment of pesticides
residues in food (482). The PPR Panel found good evidence that
combination effects can arise from co-exposure to chemicals
that produce common (adverse) outcomes through entirely
different modes of action and recommended cumulative risk
assessment methods to evaluate mixtures of pesticides in foods
that have dissimilar modes of action (396).
Cumulative risk assessment should anticipate synergies of chemicals
acting on different targets/tissues. The Hallmarks of Cancer
framework suggest that spatiotemporal aspects of chemical
exposures are likely important as well. For example, the many
constituent parts of the immune system and its distributed
nature (e.g. lymph vessels, thymus, bone marrow and so on), the
hypothalamic–pituitary–adrenal axis and cortisol in circulation,
which are used to suppress macrophage migration inhibitory
factor and control inammation (487–489) and the surrounding
tissues of the tumor microenvironment, are all relevant targets
that could be chemically disrupted to produce procarcinogenic
contributions to carcinogenesis.
For example, as noted previously, maneb is a fungicide with
a potentially disrupting effect on cortisol (446), which could
impact the body’s response to inammation suppression,
whereas atrazine affects the host immune response by directly
targeting maturation of DCs and decreasing the levels of major
histocompatibility complex class I molecules (243,453). Both
are highly relevant forms of disruption for carcinogenesis, but
within the mode of action framework, the cumulative effects
of these chemicals (and other chemicals acting on these and
similarly distributed targets) would never be assessed together
because they do not act on a common biological target.
The PPR Panel recently pointed out that there is no empiri-
cal evidence for the validity of independent action as a predic-
tive concept for multicomponent mixtures in the mammalian
toxicological literature. Further, they argued that although over-
lapping toxic effects in different organs/systems may exist, it is
difcult to identify a combination effect. Thus, the panel speci-
cally restricted their focus to chemicals that ultimately produce
a common adverse outcome (e.g. cancer) in the same target
organ/system (482). Although it may be difcult to identify this
sort of an effect, that does not mean, however, that we should
ignore this possibility (i.e. now that our understanding of the
biology of cancer has improved).
Cumulative risk assessment should anticipate synergies of non-
carcinogens. The WHO IPCS mode of action framework accepts
the notion of a common toxic endpoint and therefore that
chemicals need to rst be carcinogens themselves before they
can be considered as possible constituents of carcinogenic
mixtures. However, it is now evident that not every
procarcinogenic action resulting from a chemical exposure
must be the result of a chemical that is a carcinogen itself.
Continued focus on individual carcinogens reects a lingering
paradigm that overlooks the examples of synergies such as
those highlighted in this project. Low-dose mechanistic
effects may be very important so approaches are needed
that take this into account. In chronic and complex diseases,
establishing dose thresholds using the whole disease as the
endpoint (e.g. cancer) may be inappropriate, especially when
exposures to individual chemicals can produce relevant (but
not disease causing) mechanistic effects at much lower dose
levels.
Cumulative risk assessment should anticipate synergies of
structurally dissimilar chemicals. The EPA’s emphasis on
structurally similar classes of chemicals for mixtures risk
assessments is unnecessarily restrictive. The dissimilar
chemicals reviewed within this special issue are testament to
the fact that similar disruptive effects can be produced by a wide
range of chemical structures and failure to adapt testing to this
fact is no longer acceptable (486).
In sum, it is concerning that the WHO IPCS approach is so
highly restrictive when it comes to the assessment of cumu-
lative effects. The OECD guidelines acknowledge that cancers
originating from at least some cell types may arise by a vari-
ety of independent pathways, but the guidance is fundamen-
tally focused on the identication of individual carcinogens and
cumulative effects of carcinogens, specically noting that the
approach is intended to ‘avoid misidentication of non-tumorigenic
compounds as possible human carcinogens’ (480). But in practice, as
in-vitro and in-vivo evidence for many chemicals is frequently
not available (i.e. to prove that they individually act via a com-
mon sequence of key events or process a common target/tissue
to produce cancer), it means that risk assessments of the cumu-
lative effects of exposures to mixtures of chemicals on carcino-
genesis are rarely conducted.
W.H.Goodson et al. | S281
The International Life Sciences Institute, which is a non-
prot organization with members comprised largely of major
corporate interests from the food and beverage, agricultural,
chemical and pharmaceutical industries, has worked closely
with the WHO IPCS to support this approach. But while it may
serve to ensure the avoidance of the misidentication of (non-
tumorigenic) chemicals/compounds as possible human car-
cinogens, it simultaneously discourages regulatory agencies
from exploring the sorts of synergies that might plausibly be
expected to occur. Indeed, the biology of cancer suggests that
the cumulative effects of non-carcinogenic chemicals acting on
different pathways that are relevant to cancer, and on a variety
of cancer-relevant systems, organs, tissues and cells may very
well conspire to produce carcinogenic synergies that will be
overlooked entirely as long as the mode of action framework
(and the restrictions that it imposes) remains inuse.
As mentioned briey previously, a considerable effort has
been made by toxicologists to advance a new approach called
the Adverse Outcome Pathway framework. This is an extension
of the Mode of Action framework and is primarily being devel-
oped as an alternative solution to in-vivo toxicity testing. The
framework is based on the idea that any adverse human health
effect caused by exposure to an exogenous substance can be
described by a series of causally linked biochemical or biologi-
cal key events with measurable parameters (28,490). Although
the Adverse Outcome Pathway framework anticipates the pos-
sibility that multiple pathways may need to be dened (i.e. dif-
ferent pathways that can produce the same adverse human
health effect), the concept is currently aligned with the mode
of action approach and focuses mainly on individual chemi-
cal effects that follow a well-described pathway to produce an
adverse health outcome. So as it is currently conceived, it has
some of the same limitations that apply to the mode of action
framework.
Nonetheless, this focus at a mechanistic level is progressive
in nature and some researchers in this area are starting to call
for the adoption of practices within the framework that can
account for epigenetic effects, transgenerational effects and
chronic toxicity (detrimental effects arising in individual or at
the population level following long-term continuous or uctu-
ating exposure to chemicals at sublethal concentrations—i.e.
concentrations not high enough to cause mortality or directly
observable impairment following acute, short-term exposure,
but able to induce specic effects potentially leading to adverse
outcomes occurring at a later point in time) (28).
So this framework may be suitable for research that is
focused on mixtures of chemicals and the pathways involved
in carcinogenesis, so long as the adherents to this approach
are open to the possibility that all relevant pathways need not
have adverse health outcomes as endpoints, and that synergies
between pathways may need to be anticipated. In other words,
a series of seemingly benign actions on different pathways may
be needed to conspire to produce the adverse health outcome
that is of interest. This is the case in cancer. There are so many
layers of redundancy and safeguards in place that individual
disruptions of certain pathways may never cause disease on
their own. Yet, when a number of these pathways are enabled,
they can produce a discernable adverse health outcome (i.e.
cancer). If this approach is robust enough to anticipate this type
of complexity, it may be a model that will allow us to move past
the limitations imposed by the mode of actionmodel.
Many regulatory agencies that conduct chemical risk assess-
ments also have a mandate to ensure that adequate safety
margins are in place to protect sensitive subpopulations. So
they will need to place an increasing emphasis on the inter-
play between environmental factors and genetic factors and
also consider in-utero exposures and the potential for transgen-
erational effects. Some progress has been made in tackling the
gene-environment interaction problem using pathway analysis
to demonstrate the role of genetic variants in exposure-related
cancer susceptibility (c.f. Malhotra et al. (498)), but very little
research has been done on in-utero exposures to mixtures of
chemicals that act on cancer-related mechanisms. An approach
that focuses on dening mixtures of constituents that act dis-
ruptively on key mechanisms that are related to individual hall-
marks may serve as a useful starting point to nd evidence of
relevant transgenerational effects (c.f. Singh et al. (499)). This
is denitely an area where additional research and regulatory
input is needed.
Research needs: cancer versus carcinogenesis
One of the main challenges in this project has been the need
to better understand carcinogenesis as a process characterized
by a long latency—and the corollary possibility of both direct
and indirect effects—rather than cancer as a disease endpoint
that must occur rapidly and in the majority of exposed persons
to be relevant. This was also accompanied the recognition that
the Hallmarks of Cancer are frequently neither xed nor spe-
cic for cancer (349–351). Numerous experimental models have
been used in cancer research over the years, and Vineis etal.
(493) summarized them into at least ve separate classes of
models—seebelow:
(a) Mutational models
(b) Genome instability
(c) Models based on non-genotoxic mechanisms, clonal
expansion and epigenetics
(d) ‘Darwinian’ or ‘somatic cellular selection’, and
(e) ‘Tissue organization’.
All of these models have had signicant support in the scientic
literature (based upon empirical evidence) and there is consid-
erable overlap between them. But our collective understanding
of carcinogenesis is still largely constrained by a historically
monolithic toxicology-based approach that has been focused on
the effects of mutagens and the disease itself. So although the
Hallmarks of Cancer framework helps us to better conceptualize
the many acquired capabilities of the disease, it leaves much to
the imagination when it comes to advancing our understanding
of carcinogenesis per se. This lacuna was recently highlighted by
Brash etal. (494,495) in an article on what they called ‘the mys-
terious steps in carcinogenesis’.
Carcinogenesis appears to be an evolution of factors that
ultimately conspire towards various acquired capabilities (i.e.
those delineated within the Hallmarks of Cancer framework),
but how much does the sequencing of these acquired capabili-
ties matter and in what order are these capabilities acquired?
Figure 1 implies a rough sequencing of these capabilities, but
do we know for certain that all hallmarks for established cancer
are important for carcinogenesis as well (i.e. which hallmarks
are necessary for all tumors, and of those, which are sufcient
or perhaps distinct for certain cancers?). Other important ques-
tions to ask relate to whether or not the individual hallmarks
are a cause or a consequence of cancer development? Do the
individual hallmarks need to be expressed simultaneously or
sequentially along the continuum of carcinogenesis (from expo-
sure to unambiguous cancer phenotype development)? More
S282 | Carcinogenesis, 2015, Vol. 36, Supplement 1
importantly, how does our understanding of this framework
inform our general approach to the study of carcinogenesis?
We have partial answers to some of these questions, but
some of these questions remain unanswered, and given the
prolonged latency of many cancers, these are important ques-
tions. Our lack of knowledge in this regard makes it difcult to
draw immediate conclusions about the effects that exposures to
mixtures of disruptive chemicals might cause and the synergies
they might produce. Public health protection is challenged by
the combinatorial complexity posed, not only by multiple expo-
sures to chemicals at environmentally relevant doses (either
simultaneously or sequentially) but also through the different
mechanisms played out in temporospatial manners (includ-
ing life stages of development, which are different from those
applied in traditional toxicologic and carcinogenic screening).
We, therefore, need to consider an expanded research agenda
to include the origins, determinants and temporospatial evolu-
tion of the various cancer hallmarks and their interrelatedness.
The key questions of reversibility and of cause versus conse-
quence must also be rigorously addressed at every step from ini-
tiating carcinogenic exposure to established cancer, recognizing
that not all hallmarks are either xed or specic for any given
cancer type.
Research needs: the Hallmarks ofCancer
Current approaches to the study of chemical exposures and car-
cinogenesis have not been designed to address effects at low
concentrations or in complex mixtures. Procarcinogenic agents
may be directly genotoxic, indirectly genotoxic or non-genotoxic.
In principle, not every disruptive effect resulting in a change
that mimics a cancer hallmark is necessarily carcinogenic. Such
associations, when observed, still require rigorous validation to
ensure that exposures are unequivocally linked to the develop-
ment of both cancer and accompanying phenotypic hallmarks.
These complex interactional possibilities, coupled with the fact
that low-dose combinatorial effects on cancer development
and progression have not been rigorously or comprehensively
addressed, speak to major gaps in our understanding of envi-
ronmental cancer risk and the specic role that mixtures of
environmental chemical exposures might play in the incidence
of cancer at the populationlevel.
Unfortunately, the known effects for chemicals examined in
isolation and at higher concentrations cannot be readily extrap-
olated to effects at lower concentrations. Interactions within
complex mixtures will also occur against the backdrop of com-
plex interactions with other environmental, genetic and epige-
netic factors, so there is a need for expanded or complementary
conceptual and experimental frameworks to better understand
the determinants and specic functional contributions of envi-
ronmental exposures in cancer.
A considerable amount of energy is now being placed on the
development of research and technologies that can support the
‘exposome’ (496), an emerging concept aimed at representing
the totality of chemical exposures received by a person during
a lifetime. This approach encompasses all sources of toxicants
and is intended to help researchers discern some of the contrib-
uting factors that are driving chronic diseases such as cancer.
Related projects are expected to involve extensive biomoni-
toring (e.g. blood and urine sampling) and other techniques to
assess biomarkers that might be relevant, and this information
should be extremely helpful. Longitudinal studies should also
be carried out in animal models to assess the tissue distribution
of mixtures of chemical metabolites. To truly make good use
of this information, we are going to need a better mechanistic
understanding of the process of carcinogenesis itself and better
early markers of cancer development.
It therefore makes sense to pursue empirical research based
on our current understandings of the disease to test the effects
of real-world environmental mixtures at relevant dose levels.
Basic studies should be designed to test joint toxic action (of
carefully designed combinations of chemicals) to assess both
dose additivity (via common mode of action) and response addi-
tivity (via disparate modes of action). Research designs should
anticipate the many layers of inherent defense and incorporate
chemical constituents specically intended to demonstrate pre-
dictable synergies and mechanistic relevance. It would also be
useful to know whether or not the chemical induction of certain
numbers/combinations of hallmarks is sufcient to consistently
produce in-vivo carcinogenesis.
Mixtures research that focuses on the carcinogenic synergies
of non-carcinogenic constituents would be particularly useful.
In addition, compounds or classes of chemicals already consid-
ered to be (complete) carcinogens in the classical sense may also
contribute to carcinogenesis in complex mixtures at concentra-
tions not traditionally deemed carcinogenic. For this reason
and for completeness, ‘classic’ carcinogens with an established
environmental presence at levels that are presumed to be incon-
sequential may still have pathogenic relevance and should be
routinely included in the analysis.
Target sites that are being manipulated and disruptive chem-
icals that are being selected to produce carcinogenic effects
should be scrutinized for confounding effects. Table4 contains
aggregated evidence of cross-hallmark effects for selected path-
ways/mechanisms, and although some target sites for disrup-
tion may be compelling starting points for researchers focused
on a given phenotype (e.g. genetic instability), cross-hallmark
relationships should be explored. So, for example, telomere loss
is seen as a disruptive (procarcinogenic) effect from the per-
spective of the the genetic instability team (i.e. the group in this
project who selected this target) and it has also been shown to
exert procarcinogenic effects in four other hallmark areas. But
evidence also exists that suggests that telomere loss can have
anticarcinogenic effects in four other hallmark areas. The exact
circumstances of the various studies that support these cross-
hallmark relationships would need to be reviewed to better
understand the implications/relevance of these reported effects.
But checking planned disruptions of each target across all of
the other hallmark areas is a way to ensure that confounding
(i.e. anticarcinogenic) effects are not inadvertantly introduced
into experiments that are aimed at producing carcinogenesis,
or phenotypes that can support/contribute to carcinogenesis.
Similarly, Table 5 contains aggregated evidence of cross-hall-
mark effects for the chemical disruptors in this review, so this
table can be used for the same purpose.
It may also be productive to identify ‘reference compounds’
(ideal and prototypical disruptors) for each hallmark path-
way as a guide to predict different combinations of chemicals
that might act in a procarcinogenic manner on any one of
the hallmarks. This may involve different systems and organs
that have relevance to cancer and this sort of research could
also be combined with similar sorts of research on other refer-
ence compounds or mixtures that are shown to enable other
hallmarks. In doing so, researchers should evaluate epigenetic
changes in multiple samples/organs/tissues from exposed ani-
mals/other experimental models using gene array technology,
‘omics’ approaches, real-time imaging of tumors in 3D both in-
vitro (primary cells) and in-vivo models combined with molecu-
lar biomarkers of disease progression, and cellular immune
W.H.Goodson et al. | S283
parameters. The combination of use of computational chemical
genomics virtual screening (497), system biology/pharmacology
and high-quality imaging techniques should help us nd quan-
titative-structure-activity-relationship correlations between the
chemical structure of dissimilar disruptors and experimental
data on biological activity, physiological changes, in-vivo toxicity
and 3D cellular protein dynamics.
It is also conceivable that the combined effects of hundreds of
chemicals in the environment may be involved in the process of
enabling carcinogenesis at the population level, so basic empiri-
cal research that can demonstrate carcinogenic effects with min-
imalistic combinations may initially be needed to reveal the more
granular aspects of carcinogenesis. For example, initial research
might test our assumptions of the step-wise progression of car-
cinogenesis using targeted mixtures of chemicals that exert LDE
to test combinations of 2, 3, 4 chemicals etc. against specic
hallmarks and then adding additional targets to move through
the various steps that are believed to be needed to fully enable
the process. Experiments of this nature may reveal increases
as well as decreases in cancer risk when different mechanisms
are disrupted and corresponding hallmark phenotypes are ena-
bled (depending on the timing of various disruptive exposures).
Batteries of tests may ultimately be needed to evaluate whole
mixtures and key components individually and in various com-
binations. HTS approaches will be particularly helpful here, and
a tiered approach may make sense to look for disruptive combi-
nations, which can then be applied in vivo. Exposure sequenc-
ing and dosage may also be important and should be evaluated
based on our current understandings of the biology of cancer.
In terms of setting research priorities, tissue fate is also a
matter for consideration. It has been known for many years that
certain chemicals have afnities for certain tissues, and radi-
otracer labeling studies that have been conducted on chemicals
for regulatory purposes illustrate how certain chemicals tend
to accumulate in certain tissues. Additionally, it is well known
that some tissue types give rise to human cancers millions of
times more often than other tissue types (498). So, researchers
may want to focus their work on mixtures of disruptive chemi-
cals that prove to be complementary at a mechanistic level and
individually known to accumulate in the same types of tissues,
while at the same time choosing tissue types that are known to
produce cancers more rapidly.
The work that has been done by the WHO IPCS on mode of
action has been very useful. Understanding when chemicals
operate through the same mode of action is denitely good
information for analytical purposes, but given that we now rec-
ognize that non-carcinogens acting at very low-dose levels on
different targets and mechanisms can still activate carcinogen-
esis-related pathways, the combined (carcinogenic) potential of
the many commonly encountered chemicals within the envi-
ronment still needs to be evaluated.
Increasingly, our information is improving and there are
several tools that researchers can use to improve their research
designs. For example, ToxCast™ is an approach launched by
the EPA in 2007 to develop ways to predict potential toxicity of
chemicals and to develop a cost-effective approach for prior-
itizing the thousands of chemicals that need toxicity testing.
The ToxCast™ database was used in this project by a number
of the teams and an incredible amount of data are available on
in-vitro tests (produced using HTS) for a wide range of chemi-
cals. For example, there are many results that are direct meas-
ures of actions related to important mechanisms found within
the Hallmarks of Cancer framework, which would be useful for
research focused along theselines.
Although the hallmark phenotypes in this project repre-
sent areas of cancer research for which there is considerable
agreement, one critique of this framework is that it ignores the
‘missing hallmark’ of dedifferentiation (351). As well, the com-
plexity encompassed by each of these areas of research is hum-
bling. Moreover, cancer is not a singular or xed entity, which
frequently limits the ability to generalize about cancer biology
(349–351). In a recent reection on his career, Weinberg et al.
(499) noted not only widespread acceptance of the ‘Hallmarks of
Cancer’ heuristic but also that this attempt to simplify the dis-
ease is rapidly being eclipsed by calls from the next generation
of researchers who are now focused on assembling and analyz-
ing enormous data sets to gain an increasingly sophisticated
understanding of cancer (e.g. genomes, transcriptomes, pro-
teomes—including isoforms, post-translational modications
and proteoforms, epigenomes, kinomes, methylomes, glycomes
and matrisomes—each one of which encompasses staggering
amounts of accumulated information) (499).
Many researchers have called for an analytical use of sys-
tems biology to transcend the study of individual genes/proteins
and to integrate this complexity into higher order phenotypes
(500,501). Systems biology enables researchers to identify prop-
erties that emerge from complex chemical–biological systems
by probing how changes in one part affect the others and the
behavior of the whole system. The combined effects of tens,
if not hundreds, of simultaneous exposures may need to be
accounted for. The fundamental challenge is that such models
require parameters that are driven by data, but there are very
few good examples of research on mixtures at environmentally
relevant dose levels (502) (c.f. Porter etal. (510)), and there are
fewer still that are focused on cancer.
Nonetheless, in the near term, this basic framework should
serve as a useful starting point for foundational research and
government funding agencies should consider new ways to
support large-scale, team-based holistic approaches to this
problem.
Regulatory priorities (in the face of combinatorial
complexity)
It will take time before we fully understand the carcinogenic
potential of low-dose exposures to chemical mixtures in the
environment. Nonetheless, we cannot afford to lose sight of
the fact that the incidence of cancer remains unacceptably
high, and that the unavoidable (i.e. not lifestyle related) causa-
tive factors that are, in part, underpinning this trend are still
not fully understood (9–11,504,505). Populations worldwide are
continually exposed to a wide range of chemicals, so keeping
the precautionary principle in mind (506), there is a need to take
the risks related to the cumulative effects of these chemicals
seriously (422). Of primary concern is the fact that WHO IPCS
mode of action framework (477) and the OECD guidelines for
risk assessment (480) are restrictive to the point that regulators
could be underestimating the risks posed by exposures to low
doses of mixtures of chemicals.
National regulatory agencies and cancer research founda-
tions must proactively pursue empirical research programs to
assess any basic relationships that can be discerned between
exposures to mixtures of commonly encountered chemicals and
carcinogenicity. For example, systematic exploratory research in
appropriate rodent models exposed to ‘whole-mixtures’ that
consist of multiple chemical constituents at environmentally
relevant dose levels could demonstrate the carcinogenic poten-
tial of complex mixtures that are relevant to the population.
S284 | Carcinogenesis, 2015, Vol. 36, Supplement 1
There is also a compelling need for complementary basic
research to address specic causal relationships between envi-
ronmental exposures and the associated development of cancer
and its characteristic hallmarks.
Hypothetically speaking, such a ‘whole mixture’ should be
composed of non-carcinogens and potential carcinogens given
that individual chemicals that are not carcinogenic could act on a
range of different systems, tissues and/or cells and act synergisti-
cally with other chemicals to instigate carcinogenesis. The goal of
such investigations would not be to single out any given chemical
as a carcinogen, but rather to determine whether or not unantici-
pated (procarcinogenic) synergies of many commonly encoun-
tered chemicals when combined are endangering public health.
In line with the 3Rs (Reduction, Replacement and Renement)
guiding principles for more ethical use of animals in scientic
experiments, there has been a signicant push for research-
ers and regulatory agencies to move away from in-vivo testing
(e.g. European Union REACH legislation and in the USA, the NRC
Toxicology for the 21st Century vision (507)) to take advantage
of HTS and other new technologies. The EPA’s effort to search
for environmental chemicals that are most active in relevant
assays across the various cancer hallmarks, and then to com-
pare those results with in-vivo rodent carcinogenicity data for
the same chemicals, was a denite step in this direction (29).
However, HTS models of carcinogenicity will require validation,
and signicant hurdles remain before this sort of testing will
be ready to replace in-vivo research (508). Therefore, in the near
term, in-vivo testing still remains an important avenue for devel-
oping data sets to address cancer risks of complex mixtures.
Summary/Conclusions
For several decades, there has been a concerted effort to iden-
tify individual chemicals and other agents that are carcinogenic.
At the same time, however, little has been done to determine
whether or not chronic lifetime exposures to mixtures of non-
carcinogenic chemicals in the environment (at low-dose lev-
els) have carcinogenic potential. Many chemicals are known to
accumulate in bodily tissues over time, but little is known about
their combined effects at a mechanistic level and their impact
on cancer-related mechanisms and carcinogenesis. In this pro-
ject, teams of cancer biologists worked with researchers in the
eld of environmental health for the very rst time to explore
this possibility.
Teams that reviewed these cancer-related phenotypes (i.e.
genetic instability, tumor-promoting inammation, sustained
proliferative signaling, insensitivity to antigrowth signals, resist-
ance to cell death, angiogenesis, tissue invasion and metastasis,
the tumor microenvironment and avoiding immune destruc-
tion) readily identied individual (non-carcinogenic) chemicals
that are ubiquitous in the environment that have some poten-
tial to act on key/priority functional targets in each of these
domains. In contrast, the teams focused on replicative immortality
and dysregulated metabolism found examples of chemicals to con-
sider but noted a signicant lack of useful toxicological research
in theseareas.
In total, 85 examples of environmental chemicals were
reviewed as prototypical disruptors (for specic actions on key
pathways/mechanisms that are important for carcinogenesis)
and 59% of them (i.e. 50/85) were found to exert LDE (at levels that
are deemed relevant given the background levels of exposure that
exist in the environment) with 15 of the 50 demonstrating their
LDE in a non-linear dose-response pattern. Only 15% of the chem-
icals reviewed (i.e. 13/85) were found to have a dose-response
threshold and the remaining 26% (i.e. 22/85) were categorized as
‘unknown’ due to a lack of dose-response information.
Cross-hallmark effects for all target sites for disruption and
for all chemicals were found, but the evidence supporting these
results varied considerably in strength and in context.
A number of the teams also cited relevant in-utero expo-
sure studies in their reviews and presented data on transgen-
erational effects related to different aspects of the disease (e.g.
inammation, immune evasion and so on). These examples
raise intriguing possibilities about vulnerabilities at the popu-
lation level, and the contributions that in-utero and early life
exposures to mixtures of those chemicals might make towards
cancer susceptibility.
Therefore, current regulations in many countries (that con-
sider only the cumulative effects of exposures to individual
carcinogens that act via a common sequence of key events and
processes on a common target/tissue to produce cancer) should
be revisited. Our current understanding of the biology of can-
cer suggests that the cumulative effects of (non-carcinogenic)
chemicals acting on different pathways that are relevant to
cancer, and on a variety of cancer-relevant systems, organs, tis-
sues and cells could conspire to produce carcinogenic synergies
that will be overlooked using current risk assessment methods.
Cumulative risk assessment methods that are based on ‘com-
mon mechanisms of toxicity’ or common ‘modes of action’ may
therefore be underestimating cancer-related risks. In-utero and
early life exposures, transgenerational effects and the interplay
between the low-dose mechanistic effects of chemical mixtures
in the environment and the vulnerabilities of subpopulations
who are predisposed to cancer (i.e. via genetics or other inu-
ences) must also be considered. Current policies and practices
do not adequately address these issues and should therefore be
revisited if regulatory agencies hope to better understand and
assess theserisks.
Finally, given the long latency period in most cancers, early
detection to cancer is key so an improved understanding of the
biology within originating tissues (during the latency period)
would be very helpful. If we can use the heuristic presented in
this review to better assess the combined effects of common
exposures to chemical mixtures in the environment, it will help
us improve our understanding of carcinogenesis and identify
exogenous triggers and enabling factors (in utero and during
this important latency period), all of which will be key for the
development of effective strategies for prevention and early
detection.
Contributions
The rst draft of this manuscript (prepared by W.H.G.) was distrib-
uted to all of the contributors within the task force for feedback
and additional inputs. The many responses that followed were
managed by W.H.G. (with the assistance of L.Lo., M.G.and D.O.C.).
Then, multiple rounds of inputs were solicited from the entire task
force with several subsequent rounds of revisions and renements
prior to submission. In addition to the contributions mentioned
previously, The Halifax Project also beneted from the involve-
ment of D.J.C. (who provided details, at the workshop in Halifax,
Nova Scotia, Canada, of National Institute of Environmental
Health Sciences priorities and the agency’s interest in unravel-
ling the health effects of environmental mixtures) and from Glenn
Rice (who gave a Halifax workshop presentation on the chemical
mixtures as a consideration in cancer risk assessment). Both of
these presenters were included in the iterative rounds of manu-
script revisions mentioned previously, and both offered inputs that
W.H.Goodson et al. | S285
resulted in renements to the manuscript. Finally, the journal’s
peer-review process was important, and resulted in the collection
of additional evidence from the teams that related to thresholds,
LDE and of non-monotonic dose-response relationships. The
reviewer’s critical analysis on these topics resulted in a substantial
improvement to the data presented in this capstone document,
which ultimately served to highlight the extent to which low-dose
exposures to individual chemical constituents (within mixtures of
environmental chemicals) might have relevance for the process of
carcinogenesis. Dose-response characterization data and inputs
were then submitted by all teams and subsequently reviewed and
compiled by N.K., A.Co. and R.M.
The Halifax Project Task Force that worked on this manuscript
involved nearly 200 people, many of whom contributed to, and
signed on to this capstone article. The design of the Halifax Project
was conceived by L.Lo. with scientic advice from M.G. Funding
provided by the National Institute for Environmental Health
Sciences was arranged by D.O.C., and this manuscript was rst
drafted by W.H.G. Starting with the Hallmarks of Cancer frame-
work (Hanahan etal. (21)), 11 teams of international cancer biolo-
gists and toxicologists were established to review the literature
on key cancer-related mechanisms/pathways in their respective
domains and to also look at the disruptive potential of low-dose
exposures to chemicals commonly encountered in the environ-
ment (i.e. as it relates to those same mechanisms/pathways).
Each team had a leader and each team was responsible for con-
tributing a section of related content within the capstone manu-
script. The contributing authors from these teams are as follows:
(1) Angiogenesis (Z.H., C-W.H., H-Y.H., L-T.L., M.X., N.K., S.A.B.,
T.M., V.D., W.K.R.); (2) Dysregulated metabolism (R.B.R., A.C.S.,
A.B., E.Ry., D.B., F.C., F.L.M., G.Wi., J.We., N.B.K., R.P.); (3) Evasion of
antigrowth signaling (R.N., A.L., C.C.N., D.W.L., D.R., G.S.G., G.M.C.,
H.Kr., J.V., K.A.C-S., M.W., N.C., P.A.M., P.De., R.A-V., R.V., R.D.F., R.P-
C., R.C.C., S.N.B.), (4) Genetic instability (S.A.S.L., A.L.d.C.S., A.Az.,
A.K.C., A.R.C., A-K.O., E.Ro., F.D., F.J.V.S., G.K., G.B., L.Go., L.Le., L.Z.,
M.Val., M.K-V., N.v L., P.O-W., S.Pav., T.C.); (5) Immune system eva-
sion (H.K.L., E.C., J.K., M.A.W., M.H.M., T.O., W.K.D.), (6) Replicative
immortality (A.Ca., C.B-A., H.Y., H.Ko., J.P.W., J.F.M-L., M.L., S.S.W.);
(7) Resistance to cell death (H.H.P., A.M.A., B.J.B., C.Y., E.R., K.B.N.,
L.S.D’A., L.Li., M.F.R., M.J.G., P.M.G., P.S.L., Q.(S.) C., R.K.S., R.D., S.Ro.,
S.L., T-J.L., Y.R.); (8) Sustained proliferative signaling (W.E., A.W.,
G.Wa., H.S., J.E.K., J.R., K.M., L.Gu., M.V.K., P.V., P.Da., R.M., S.Er.,
T.S., T.H.); (9) Tissue invasion and metastasis (J.O., B.P.Z., C.D., G.N.,
G.T.W., I.K., I.R.M., L.J.M., N.A., O.O., P.N-M., S.El., S.Pap., V.O-M., Y.L.,
Z.C.); (10) Tumor microenvironment (D.W.F., C.S.C., D.C.K., E.L.,
F.M., J.Ro., J.C., J.R.W., L.S., L.V., M.C., P.K.K., P.H., S.Ry., S.C.C., V.M-
S.) and (11) Tumor-promoting inammation (P.T., C.B., E-Y. M., J.S.,
L.J., M.K., S.H., T.G., V.S.).** Additionally, a special cross-functional
team was established to investigate whether or not the chemi-
cals that were identied by the teams as having disruptive poten-
tial for key mechanisms/pathways in a particular domain might
also have been shown in other research to exert relevant effects
on mechanisms/pathways in other domains. The results of the
efforts from this team have been compiled and summarized in
this article and can be found within Table4. This team was com-
prised as follows: W.H.B., A.Am., I.S., A.Co., C.M., D.B., E.Ry., F.A-
M., H.A.H., H.K.S., J.Ra., J.Wo., K.R.P., L.M., M.Vac., N.S., R.A-T., R.R.,
R.A.H.and S.F.** **Note that team leaders are denoted by the rst
set of initials in each team list.
Funding/Acknowledgements
We gratefully acknowledge the support of the National Institute
of Health-National Institute of Environmental Health Sciences
(NIEHS) conference grant travel support (R13ES023276); Glenn
Rice, Ofce of Research and Development, United States
Environmental Protection Agency, Cincinnati, OH, USA also
deserves thanks for his thoughtful feedback and inputs on the
manuscript; William H.Goodson III was supported by the
California Breast Cancer Research Program, Clarence Heller
Foundation and California Pacic Medical Center Foundation;
Abdul M.Ali would like to acknowledge the nancial support of
the University of Sultan Zainal Abidin, Malaysia; Ahmed Lasfar
was supported by an award from the Rutgers Cancer Institute of
New Jersey; Ann-Karin Olsen and Gunnar Brunborg were sup-
ported by the Research Council of Norway (RCN) through its
Centres of Excellence funding scheme (223268/F50), Amancio
Carnero’s lab was supported by grants from the Spanish Ministry
of Economy and Competitivity, ISCIII (Fis: PI12/00137, RTICC:
RD12/0036/0028) co-funded by FEDER from Regional Development
European Funds (European Union), Consejeria de Ciencia e
Innovacion (CTS-1848) and Consejeria de Salud of the Junta de
Andalucia (PI-0306-2012); Matilde E. Lleonart was supported by a
trienal project grant PI12/01104 and by project CP03/00101 for per-
sonal support. Amaya Azqueta would like to thank the Ministerio
de Educacion y Ciencia (‘Juande la Cierva’ programme, 2009) of
the Spanish Government for personal support; Amedeo Amedei
was supported by the Italian Ministry of University and Research
(2009FZZ4XM_002), and the University of Florence (ex60%2012);
Andrew R.Collins was supported by the University of Oslo;
Annamaria Colacci was supported by the Emilia-Romagna Region
- Project ‘Supersite’ in Italy; Carolyn Baglole was supported by a
salary award from the Fonds de recherche du Quebec-Sante (FRQ-
S); Chiara Mondello’s laboratory is supported by Fondazione
Cariplo in Milan, Italy (grant n.2011-0370); Christian C.Naus holds
a Canada Research Chair; Clement Yedjou was supported by a
grant from the National Institutes of Health (NIH-NIMHD grant
no. G12MD007581); Daniel C.Koch is supported by the Burroughs
Wellcome Fund Postdoctoral Enrichment Award and the Tumor
Biology Training grant: NIH T32CA09151; Dean W.Felsher would
like to acknowledge the support of United States Department of
Health and Human Services, NIH grants (R01 CA170378 PQ22, R01
CA184384, U54 CA149145, U54 CA151459, P50 CA114747 and
R21 CA169964); Emilio Rojas would like to thank CONACyT sup-
port 152473; Ezio Laconi was supported by AIRC (Italian
Association for Cancer Research, grant no. IG 14640) and by the
Sardinian Regional Government (RAS); Eun-Yi Moon was sup-
ported by grants from the Public Problem-Solving Program
(NRF-015M3C8A6A06014500) and Nuclear R&D Program
(#2013M2B2A9A03051296 and 2010-0018545) through the National
Research Foundation of Korea (NRF) and funded by the Ministry
of Education, Science and Technology (MEST) in Korea;
Fahd Al-Mulla was supported by the Kuwait Foundation for
the Advancement of Sciences (2011-1302-06); Ferdinando
Chiaradonna is supported by SysBioNet, a grant for the Italian
Roadmap of European Strategy Forum on Research Infrastructures
(ESFRI) and by AIRC (Associazione Italiana Ricerca sul Cancro; IG
15364); Francis L.Martin acknowledges funding from Rosemere
Cancer Foundation; he also thanks Lancashire Teaching Hospitals
NHS trust and the patients who have facilitated the studies he
has undertaken over the course of the last 10 years; Gary
S.Goldberg would like to acknowledge the support of the New
Jersey Health Foundation; Gloria M.Calaf was supported by Fondo
Nacional de Ciencia y Tecnología (FONDECYT), Ministerio de
Educación de Chile (MINEDUC), Universidad de Tarapacá (UTA);
Gudrun Koppen was supported by the Flemish Institute for
Technological Research (VITO), Belgium; Hemad Yasaei was sup-
ported from a triennial project grant (Strategic Award) from the
S286 | Carcinogenesis, 2015, Vol. 36, Supplement 1
National Centre for the Replacement, Renement and Reduction
(NC3Rs) of animals in research (NC.K500045.1 and G0800697);
Hiroshi Kondoh was supported in part by grants from the Ministry
of Education, Culture, Sports, Science, and Technology of Japan,
Japan Science and Technology Agency and by JST, CREST; Hsue-
Yin Hsu was supported by the Ministry of Science and Technology
of Taiwan (NSC93-2314-B-320-006 and NSC94-2314-B-320-002);
Hyun Ho Park was supported by the Basic Science Research
Program through the National Research Foundation of Korea
(NRF) of the Ministry of Education, Science and Technology
(2012R1A2A2A01010870) and a grant from the Korea Healthcare
Technology R&D project, Ministry of Health and Welfare, Republic
of Korea (HI13C1449); Igor Koturbash is supported by the UAMS/
NIH Clinical and Translational Science Award (UL1TR000039 and
KL2TR000063) and the Arkansas Biosciences Institute, the major
research component of the Arkansas Tobacco Settlement
Proceeds Act of 2000; Jan Vondráček acknowledges funding from
the Czech Science Foundation (13-07711S); Jesse Roman thanks
the NIH for their support (CA116812); John Pierce Wise Sr. and
Sandra S.Wise were supported by National Institute of
Environmental Health Sciences (ES016893 to J.P.W.) and the Maine
Center for Toxicology and Environmental Health; Jonathan
Whiteld acknowledges support from the FERO Foundation in
Barcelona, Spain; Joseph Christopher is funded by Cancer
Research UK and the International Journal of Experimental
Pathology; Julia Kravchenko is supported by a philanthropic
donation by Fred and Alice Stanback; Jun Sun is supported by a
Swim Across America Cancer Research Award; Karine A.Cohen-
Solal is supported by a research scholar grant from the American
Cancer Society (116683-RSG-09-087-01-TBE); Laetitia Gonzalez
received a postdoctoral fellowship from the Fund for Scientic
Research–Flanders (FWO-Vlaanderen) and support by an
InterUniversity Attraction Pole grant (IAP-P7-07); Laura Soucek is
supported by grant #CP10/00656 from the Miguel Servet Research
Contract Program and acknowledges support from the FERO
Foundation in Barcelona, Spain; Liang-Tzung Lin was supported
by funding from the Taipei Medical University (TMU101-AE3-Y19);
Linda Gulliver is supported by a Genesis Oncology Trust (NZ)
Professional Development Grant, and the Faculty of Medicine,
University of Otago, Dunedin, New Zealand; Louis Vermeulen is
supported by a Fellowship of the Dutch Cancer Society (KWF,
UVA2011-4969) and a grant from the AICR (14–1164); Mahara
Valverde would like to thank CONACyT support 153781; Masoud
H. Manjili was supported by the ofce of the Assistant Secretary
of Defense for Health Affairs (USA) through the Breast Cancer
Research Program under Award No. W81XWH-14-1-0087 Neetu
Singh was supported by grant #SR/FT/LS-063/2008 from the
Department of Science and Technology, Government of India;
Nicole Kleinstreuer is supported by NIEHS contracts (N01-ES
35504 and HHSN27320140003C); P.K. Krishnakumar is supported
by the Funding (No. T.K. 11-0629) of King Abdulaziz City for
Science and Technology, Riyadh, Saudi Arabia; Paola A.Marignani
is supported by the Dalhousie Medical Research Foundation, The
Beatrice Hunter Cancer Institute and CIHR and the Nova Scotia
Lung Association; Paul Dent is the holder of the Universal Inc.
Chair in Signal Transduction Research and is supported with
funds from PHS grants from the NIH (R01-CA141704,
R01-CA150214, R01-DK52825 and R01-CA61774); Petr Heneberg
was supported by the Charles University in Prague projects UNCE
204015 and PRVOUK P31/2012, and by the Czech Science
Foundation projects P301/12/1686 and 15-03834Y; Po Sing Leung
was supported by the Health and Medical Research Fund of Food
and Health Bureau, Hong Kong Special Administrative Region,
Ref. No: 10110021; Qiang Cheng was supported in part by grant
NSF IIS-1218712; R. Brooks Robey is supported by the United
States Department of Veterans Affairs; Rabindra Roy was sup-
ported by United States Public Health Service Grants (RO1
CA92306, RO1 CA92306-S1 and RO1 CA113447); Rafaela Andrade-
Vieira is supported by the Beatrice Hunter Cancer Research
Institute and the Nova Scotia Health Research Foundation; Renza
Vento was partially funded by European Regional Development
Fund, European Territorial Cooperation 2007–13 (CCI 2007 CB 163
PO 037, OP Italia-Malta 2007–13) and grants from the Italian
Ministry of Education, University and Research (MIUR) ex-60%,
2007; Riccardo Di Fiore was a recipient of fellowship granted by
European Regional Development Fund, European Territorial
Cooperation 2007–2013 (CCI 2007 CB 163 PO 037, OP Italia-Malta
2007–2013); Rita Dornetshuber-Fleiss was supported by the
Austrian Science Fund (FWF, project number T 451-B18) and the
Johanna Mahlke, geb.-Obermann-Stiftung; Roberta Palorini is
supported by a SysBioNet fellowship; Roslida Abd Hamid is sup-
ported by the Ministry of Education, Malaysia-Exploratory
Research Grant Scheme-Project no: ERGS/1-2013/5527165; Sabine
A.S.Langie is the beneciary of a postdoctoral grant from the AXA
Research Fund and the Cec-LRI Innovative Science Award 2013;
Sakina Eltom is supported by NIH grant SC1CA153326; Samira
A.Brooks was supported by National Research Service Award (T32
ES007126) from the National Institute of Environmental Health
Sciences and the HHMI Translational Medicine Fellowship;
Sandra Ryeom was supported by The Garrett B.Smith Foundation
and the TedDriven Foundation; Thierry Massfelder was supported
by the Institut National de la Santé et de la Recherche Médicale
INSERM and Université de Strasbourg; Thomas Sanderson is sup-
ported by the Canadian Institutes of Health Research (CIHR; MOP-
115019), the Natural Sciences and Engineering Council of Canada
(NSERC; 313313) and the California Breast Cancer Research
Program (CBCRP; 17UB-8703); Tiziana Guarnieri is supported by a
grant from Fundamental Oriented Research (RFO) to the Alma
Mater Studiorum University of Bologna, Bologna, Italy and thanks
the Fondazione Cassa di Risparmio di Bologna and the Fondazione
Banca del Monte di Bologna e Ravenna for supporting the Center
for Applied Biomedical Research, S.Orsola-Malpighi University
Hospital, Bologna, Italy; W.Kimryn Rathmell is supported by the V
Foundation for Cancer Research and the American Cancer
Society; William K.Decker was supported in part by grant
RP110545 from the Cancer Prevention Research Institute of Texas;
William H.Bisson was supported with funding from the NIH P30
ES000210; Yon Rojanasakul was supported with NIH grant
R01-ES022968; Zhenbang Chen is supported by NIH grants
(MD004038, CA163069 and MD007593); Zhiwei Hu is grateful for
the grant support from an institutional start-up fund from The
Ohio State University College of Medicine and The OSU James
Comprehensive Cancer Center (OSUCCC) and a Seed Award from
the OSUCCC Translational Therapeutics Program.
Disclaimer
This manuscript is the work product of a large group of research-
ers collectively representing a diverse array of institutions and a
signicant number of sponsors; however, the statements, opin-
ions and conclusions expressed herein do not necessarily repre-
sent the views or positions of these institutions and sponsors.
Conict of interest statement: None declared
References
1. (2014) World cancer report 2014. In Wild, C.P. and Stewart B.W. (eds).
World Health Organization.
W.H.Goodson et al. | S287
2. Malhotra, J. (2014) Molecular and genetic epidemiology of cancer in
low- and medium-income countries. Ann. Glob. Health, 80, 418–425.
3. McGuinn, L.A. etal. (2012) Cancer and environment: denitions and
misconceptions. Environ. Res., 112, 230–234.
4. Sankpal, U.T. etal. (2012) Environmental factors in causing human can-
cers: emphasis on tumorigenesis. Tumour Biol., 33, 1265–1274.
5. Trosko, J.E. etal. (2005) The emperor wears no clothes in the eld of car-
cinogen risk assessment: ignored concepts in cancer risk assessment.
Mutagenesis, 20, 81–92.
6. Christiani, D.C. (2011) Combating environmental causes of cancer. N.
Engl. J.Med., 364, 791–793.
7. Clapp, R. (2011) Chemicals policy in the 2008-2009 President's Cancer
Panel Report. New Solut., 21, 447–455.
8. Reuben, S.H. (2008–2009) Reducing environmental cancer risk: what we
can do now. In Panel, T.P.s.C. (ed.), Bethesda, Maryland.
9. Parkin, D.M. etal. (2011) The fraction of cancer attributable to lifestyle and
environmental factors in the UK in 2010. Br. J.Cancer, 105(Suppl 2), S77–S81.
10. (2009) Global Health Risks: Mortality and Burden of Disease Attribut-
able to Selected Major Risks. World Health Organization, Geneva.
11. Straif, K. (2008) The burden of occupational cancer. Occup. Environ.
Med., 65, 787–788.
12. Vandenberg, L.N. et al. (2012) Hormones and endocrine-disrupting
chemicals: low-dose effects and nonmonotonic dose responses.
Endocr. Rev., 33, 378–455.
13. (2009) OECD Guidelines for the Testing of Chemicals, Section 4 Health
Effects, Test No. 451: Carcinogenicity Studies. OECD.
14. Wignall, J.A. etal. (2014) Standardizing benchmark dose calculations to
improve science-based decisions in human health assessments. Envi-
ron. Health Perspect., 122, 499–505.
15. Myers, J.P. etal. (2009) A clash of old and new scientic concepts in
toxicity, with important implications for public health. Environ. Health
Perspect., 117, 1652–1655.
16. Vandenberg, L.N. etal. (2013) Regulatory decisions on endocrine dis-
rupting chemicals should be based on the principles of endocrinology.
Reprod. Toxicol., 38, 1–15.
17. Bergman, A. etal. (2013) Science and policy on endocrine disrupters
must not be mixed: a reply to a “common sense” intervention by toxi-
cology journal editors. Environ. Health, 12, 69.
18. Ames, B.N. (1979) Identifying environmental chemicals causing muta-
tions and cancer. Science, 204, 587–593.
19. Armitage, P. et al. (1954) The age distribution of cancer and a multi-
stage theory of carcinogenesis. Br. J.Cancer, 8, 1–12.
20. Truhaut, R. (1990) [Recent progress in the evaluation of the dangers of
chemical carcinogens]. J. Pharm. Belg., 45, 131–140.
21. Hanahan, D. etal. (2000) The hallmarks of cancer. Cell, 100, 57–70.
22. Preston, R.J. (2005) Extrapolations are the Achilles heel of risk assess-
ment. Mutat. Res., 589, 153–157.
23. Hanahan, D. etal. (2011) Hallmarks of cancer: the next generation. Cell,
144, 646–674.
24. Colotta, F. et al. (2009) Cancer-related inammation, the seventh hall-
mark of cancer: links to genetic instability. Carcinogenesis, 30, 1073–1081.
25. Warburg, O. (ed.) (1930) The Metabolism of Tumours: Investigations
from the Kaiser Wilhelm Institute for Biology, Berlin-Dahlen. Constable
& Company Limited, London.
26. Aisenberg, A.C. (1961) The Glycolysis and Respiration of Tumors. Aca-
demic Press, New York, NY.
27. Ankley, G.T. et al. (2010) Adverse outcome pathways: a conceptual
framework to support ecotoxicology research and risk assessment.
Environ. Toxicol. Chem., 29, 730–741.
28. Groh, K.J. etal. (2015) Development and application of the adverse out-
come pathway framework for understanding and predicting chronic
toxicity: I.Challenges and research needs in ecotoxicology. Chemos-
phere, 120, 764–777.
29. Kleinstreuer, N.C. etal. (2013) In vitro perturbations of targets in cancer
hallmark processes predict rodent chemical carcinogenesis. Toxicol.
Sci., 131, 40–55.
30. Yates, L.R. etal. (2012) Evolution of the cancer genome. Nat. Rev. Genet.,
13, 795–806.
31. (2001) National Toxicology Program’s report of the endocrine disrup-
tors low dose peer review. National Institute of Environmental Health
Sciences, National Toxicology Program, Research Triangle Park, NC.
32. Melnick, R. etal. (2002) Summary of the National Toxicology Program's
report of the endocrine disruptors low-dose peer review. Environ.
Health Perspect., 110, 427–431.
33. Welshons, W.V. et al. (2006) Large effects from small exposures. III.
Endocrine mechanisms mediating effects of bisphenol Aat levels of
human exposure. Endocrinology, 147(Suppl 6), S56–S69.
34. Vandenberg, L.N. et al. (2007) Human exposure to bisphenol A(BPA).
Reprod. Toxicol., 24, 139–177.
35. Brucker-Davis, F. etal. (2001) Signicant effects of mild endogenous
hormonal changes in humans: considerations for low-dose testing.
Environ. Health Perspect., 109(Suppl 1), 21–26.
36. EPA, U.S. The U.S. Environmental Protection Agency ToxCast Phase I/II
data. http://www.epa.gov/ncct/toxcast/data.html.
37. Taylor, T.R. et al. (2011) Ziram activates mitogen-activated protein
kinases and decreases cytolytic protein levels in human natural killer
cells. Toxicol. Mech. Methods, 21, 577–584.
38. McMahon, T.A. etal. (2011) The fungicide chlorothalonil is nonlinearly
associated with corticosterone levels, immunity, and mortality in
amphibians. Environ. Health Perspect., 119, 1098–1103.
39. Goldman, J.M. et al. (2004) Methoxychlor-induced alterations in the
histological expression of angiogenic factors in pituitary and uterus. J.
Mol. Histol., 35, 363–375.
40. Chapin, R.E. etal. (1997) The effects of perinatal/juvenile methoxychlor
exposure on adult rat nervous, immune, and reproductive system
function. Fundam. Appl. Toxicol., 40, 138–157.
41. Qian, Y. etal. (2010) Peruorooctane sulfonate (PFOS) induces reactive
oxygen species (ROS) production in human microvascular endothelial
cells: role in endothelial permeability. J. Toxicol. Environ. Health A, 73,
819–836.
42. Hu, J.X. etal. (2013) Toxic effects of cypermethrin on the male repro-
ductive system: with emphasis on the androgen receptor. J. Appl. Toxi-
col., 33, 576–585.
43. Jin, M. etal. (2010) Estrogenic activities of two synthetic pyrethroids
and their metabolites. J. Environ. Sci. (China), 22, 290–296.
44. Kakko, I. etal. (2004) Oestradiol potentiates the effects of certain pyre-
throid compounds in the MCF7 human breast carcinoma cell line.
Altern. Lab. Anim., 32, 383–390.
45. Feng, Z. etal. (2006) Acrolein is a major cigarette-related lung can-
cer agent: preferential binding at p53 mutational hotspots and
inhibition of DNA repair. Proc. Natl Acad. Sci. USA, 103, 15404–
15409.
46. Günther, M. etal. (2008) Acrolein: unwanted side product or contribu-
tion to antiangiogenic properties of metronomic cyclophosphamide
therapy? J. Cell. Mol. Med., 12(6B), 2704–2716.
47. Luo, C. etal. (2013) A cigarette component acrolein induces accelerated
senescence in human diploid broblast IMR-90 cells. Biogerontology,
14, 503–511.
48. Roy, J. etal. (2010) Acrolein induces apoptosis through the death recep-
tor pathway in A549 lung cells: role of p53. Can. J.Physiol. Pharmacol.,
88, 353–368.
49. Tanel, A. et al. (2014) Acrolein activates cell survival and apoptotic
death responses involving the endoplasmic reticulum in A549 lung
cells. Biochim. Biophys. Acta, 1843, 827–835.
50. Tang, M.S. et al. (2011) Acrolein induced DNA damage, mutagenicity
and effect on DNA repair. Mol. Nutr. Food Res., 55, 1291–1300.
51. Cabeza-Arvelaiz, Y. etal. (2012) Transcriptome analysis of a rotenone
model of parkinsonism reveals complex I-tied and -untied toxicity
mechanisms common to neurodegenerative diseases. PLoS One, 7,
e44700.
52. Deng, Y.T. et al. (2010) Rotenone induces apoptosis in MCF-7 human
breast cancer cell-mediated ROS through JNK and p38 signaling. Mol.
Carcinog., 49, 141–151.
53. Gonçalves, A.P. etal. (2011) Involvement of p53 in cell death following
cell cycle arrest and mitotic catastrophe induced by rotenone. Biochim.
Biophys. Acta, 1813, 492–499.
54. Li, Y. et al. (2013) Copper induces cellular senescence in human glio-
blastoma multiforme cells through downregulation of Bmi-1. Oncol.
Rep., 29, 1805–1810.
55. Ostrakhovitch, E.A. etal. (2005) Role of p53 and reactive oxygen species
in apoptotic response to copper and zinc in epithelial breast cancer
cells. Apoptosis, 10, 111–121.
S288 | Carcinogenesis, 2015, Vol. 36, Supplement 1
56. Parr-Sturgess, C.A. etal. (2012) Copper modulates zinc metalloprotein-
ase-dependent ectodomain shedding of key signaling and adhesion
proteins and promotes the invasion of prostate cancer epithelial cells.
Mol. Cancer Res., 10, 1282–1293.
57. Freitas, M. etal. (2013) Nickel induces apoptosis in human neutrophils.
Biometals, 26, 13–21.
58. Wu, C.H. etal. (2012) Nickel-induced epithelial-mesenchymal transi-
tion by reactive oxygen species generation and E-cadherin promoter
hypermethylation. J. Biol. Chem., 287, 25292–25302.
59. Aimola, P. etal. (2012) Cadmium induces p53-dependent apoptosis in
human prostate epithelial cells. PLoS One, 7, e33647.
60. Yuan, D. et al. (2013) Long-term cadmium exposure leads to the
enhancement of lymphocyte proliferation via down-regulating p16 by
DNA hypermethylation. Mutat. Res., 757, 125–131.
61. Aluigi, M.G. etal. (2010) Apoptosis as a specic biomarker of diazinon
toxicity in NTera2-D1 cells. Chem. Biol. Interact., 187, 299–303.
62. Giordano, G. etal. (2007) Organophosphorus insecticides chlorpyrifos
and diazinon and oxidative stress in neuronal cells in a genetic model
of glutathione deciency. Toxicol. Appl. Pharmacol., 219, 181–189.
63. Gilsing, A.M. etal. (2013) Dietary heme iron and the risk of colorectal
cancer with specic mutations in KRAS and APC. Carcinogenesis, 34,
2757–2766.
64. Pluth, J.M. etal. (1996) Increased frequency of specic genomic deletions
resulting from in vitro malathion exposure. Cancer Res., 56, 2393–2399.
65. Chen, Z.J. etal. (2014) Bisphenol Amodulates colorectal cancer protein
prole and promotes the metastasis via induction of epithelial to mes-
enchymal transitions. Arch Toxicol.
66. Zhu, H. etal. (2010) Environmental endocrine disruptors promote inva-
sion and metastasis of SK-N-SH human neuroblastoma cells. Oncol.
Rep., 23, 129–139.
67. Pontillo, C.A. et al. (2013) Action of hexachlorobenzene on tumor
growth and metastasis in different experimental models. Toxicol. Appl.
Pharmacol., 268, 331–342.
68. O'Brien, D.W. etal. (2004) A mechanism of airway injury in an epithelial
model of mucociliary clearance. Respir. Res., 5, 10.
69. Ornstein, D.L. etal. (2007) Iron stimulates urokinase plasminogen acti-
vator expression and activates NF-kappa B in human prostate cancer
cells. Nutr. Cancer, 58, 115–126.
70. Mao, L. et al. (2012) Circadian gating of epithelial-to-mesenchymal
transition in breast cancer cells via melatonin-regulation of GSK3β.
Mol. Endocrinol., 26, 1808–1820.
71. Papagerakis, S. etal. (2014) The circadian clock in oral health and dis-
eases. J. Dent. Res., 93, 27–35.
72. Bouskine, A. etal. (2009) Low doses of bisphenol Apromote human
seminoma cell proliferation by activating PKA and PKG via a mem-
brane G-protein-coupled estrogen receptor. Environ. Health Perspect.,
117, 1053–1058.
73. Hernández, L.G. etal. (2013) A mode-of-action approach for the identi-
cation of genotoxic carcinogens. PLoS One, 8, e64532.
74. Wetherill, Y.B. etal. (2002) The xenoestrogen bisphenol Ainduces inap-
propriate androgen receptor activation and mitogenesis in prostatic
adenocarcinoma cells. Mol. Cancer Ther., 1, 515–524.
75. Park, S.H. etal. (2009) Cell growth of ovarian cancer cells is stimulated
by xenoestrogens through an estrogen-dependent pathway, but their
stimulation of cell growth appears not to be involved in the activation
of the mitogen-activated protein kinases ERK-1 and p38. J. Reprod. Dev.,
55, 23–29.
76. Wilkinson, C.F. etal. (1996) A mechanistic interpretation of the onco-
genicity of chlorothalonil in rodents and an assessment of human rel-
evance. Regul. Toxicol. Pharmacol., 24(1 Pt 1), 69–84.
77. Vesselinovitch, S.D. etal. (1983) Lindane bioassay studies and human
cancer risk assessment. Toxicol. Pathol., 11, 12–22.
78. Wang, Q.L. etal. (2013) Risk assessment of mouse gastric tissue can-
cer induced by dichlorvos and dimethoate. Oncol. Lett., 5, 1385–1389.
79. Lee, H.R. etal. (2012) Treatment with bisphenol Aand methoxychlor
results in the growth of human breast cancer cells and alteration of
the expression of cell cycle-related genes, cyclin D1 and p21, via an
estrogen receptor-dependent signaling pathway. Int. J.Mol. Med., 29,
883–890.
80. Stagg, N.J. etal. (2012) Assessment of possible carcinogenicity of oxy-
uorfen to humans using mode of action analysis of rodent liver
effects. Toxicol. Sci., 128, 334–345.
81. Doull, J. et al. (1999) A cancer risk assessment of di(2-ethylhexyl)
phthalate: application of the new U.S. EPA Risk Assessment Guide-
lines. Regul. Toxicol. Pharmacol., 29, 327–357.
82. Mazzoleni, G. et al. (1994) Inuence of the herbicide Linuron on
growth rate and gap-junctional intercellular communication of cul-
tured endothelial cells. J. Environ. Pathol. Toxicol. Oncol., 13, 1–10.
83. Yasaei, H. etal. (2013) Carcinogen-specic mutational and epigenetic
alterations in INK4A, INK4B and p53 tumour-suppressor genes drive
induced senescence bypass in normal diploid mammalian cells.
Oncogene, 32, 171–179.
84. Singh, K.P. et al. (2008) Allelic loss and mutations in a new ETRG-1
gene are early events in diethylstilbestrol-induced renal carcinogen-
esis in Syrian hamsters. Gene, 408, 18–26.
85. Tsutsui, T. etal. (1994) Reserpine-induced cell transformation without
detectable genetic effects in Syrian hamster embryo cells in culture.
Carcinogenesis, 15, 11–14.
86. Martens, U. etal. (1996) Low expression of the WAF1/CIP1 gene prod-
uct, p21, in enzyme-altered foci induced in rat liver by diethylnitrosa-
mine or phenobarbital. Cancer Lett., 104, 21–26.
87. Geter, D.R. etal. (2014) Dose-response modeling of early molecular
and cellular key events in the CAR-mediated hepatocarcinogenesis
pathway. Toxicol. Sci., 138, 425–445.
88. Bader, A. et al. (2011) Paracetamol treatment increases telomerase
activity in rat embryonic liver cells. Pharmacol. Rep., 63, 1435–1441.
89. Tsuruga, Y. etal. (2008) Establishment of immortalized human hepat-
ocytes by introduction of HPV16 E6/E7 and hTERT as cell sources for
liver cell-based therapy. Cell Transplant., 17, 1083–1094.
90. Nguyen, T.H. etal. (2005) Treatment of acetaminophen-induced acute
liver failure in the mouse with conditionally immortalized human
hepatocytes. J. Hepatol., 43, 1031–1037.
91. Bode-Böger, S.M. etal. (2005) Aspirin reduces endothelial cell senes-
cence. Biochem. Biophys. Res. Commun., 334, 1226–1232.
92. Heinloth, A.N. etal. (2004) Gene expression proling of rat livers reveals
indicators of potential adverse effects. Toxicol. Sci., 80, 193–202.
93. Jacob, T. etal. (2009) The effect of cotinine on telomerase activity in human
vascular smooth muscle cells. J. Cardiovasc. Surg. (Torino), 50, 345–349.
94. Brüne, B. etal. (2001) Transcription factors p53 and HIF-1alpha as tar-
gets of nitric oxide. Cell. Signal., 13, 525–533.
95. Davis, C.D. et al. (2000) Dietary selenium and arsenic affect DNA
methylation in vitro in Caco-2 cells and in vivo in rat liver and colon. J.
Nutr., 130, 2903–2909.
96. Arnér, E.S. etal. (2006) The thioredoxin system in cancer. Semin. Can-
cer Biol., 16, 420–426.
97. vom Saal, F.S. etal. (2007) Chapel Hill bisphenol Aexpert panel con-
sensus statement: integration of mechanisms, effects in animals
and potential to impact human health at current levels of exposure.
Reprod. Toxicol., 24, 131–138.
98. Qin, X.Y. etal. (2012) Effects of bisphenol A exposure on the prolif-
eration and senescence of normal human mammary epithelial cells.
Cancer Biol. Ther., 13, 296–306.
99. Peluso, M.E. etal. (2014) Bisphenol-A exposures and behavioural aber-
rations: median and linear spline and meta-regression analyses of 12
toxicity studies in rodents. Toxicology, 325, 200–208.
100. Fang, C.C. etal. (2013) Cyprodinil as an activator of aryl hydrocarbon
receptor. Toxicology, 304, 32–40.
101. Bharadwaj, R. etal. (2004) The spindle checkpoint, aneuploidy, and
cancer. Oncogene, 23, 2016–2027.
102. Orton, F. etal. (2011) Widely used pesticides with previously unknown
endocrine activity revealed as in vitro antiandrogens. Environ. Health
Perspect., 119, 794–800.
103. Tanaka, T. etal. (2013) Effects of maternal exposure to imazalil on
behavioral development in F-generation mice. Birth Defects Res. B
Dev. Reprod. Toxicol., 98, 334–342.
104. Ahmad, I. etal. (2008) The involvement of nitric oxide in maneb- and
paraquat-induced oxidative stress in rat polymorphonuclear leuko-
cytes. Free Radic. Res., 42, 849–862.
105. US Environmental Protection Agency (1988) US Integrated Risk Infor-
mation System—Maneb (CASRN 12427-38-2). http://www.epa.gov/
iris/subst/0249.htm.
106. Miller, K.P. etal. (2006) Methoxychlor metabolites may cause ovarian
toxicity through estrogen-regulated pathways. Toxicol. Sci., 93, 180–
188.
W.H.Goodson et al. | S289
107. Du, X. et al. (2014) Perinatal exposure to low-dose methoxychlor
impairs testicular development in C57BL/6 mice. PLoS One, 9, e103016.
108. Palanza, P. etal. (2001) Effects of prenatal exposure to low doses of
diethylstilbestrol, o,p'DDT, and methoxychlor on postnatal growth
and neurobehavioral development in male and female mice. Horm.
Behav., 40, 252–265.
109. Du, G. etal. (2013) Peruorooctane sulfonate (PFOS) affects hormone
receptor activity, steroidogenesis, and expression of endocrine-
related genes in vitro and in vivo. Environ. Toxicol. Chem., 32, 353–360.
110. Kim, H.S. etal. (2011) Induction of apoptosis and CYP4A1 expression
in Sprague-Dawley rats exposed to low doses of peruorooctane sul-
fonate. J. Toxicol. Sci., 36, 201–210.
111. Eveillard, A. etal. (2009) Di-(2-ethylhexyl)-phthalate (DEHP) activates
the constitutive androstane receptor (CAR): a novel signalling path-
way sensitive to phthalates. Biochem. Pharmacol., 77, 1735–1746.
112. Nakai, M. etal. (1999) Binding characteristics of dialkyl phthalates for
the estrogen receptor. Biochem. Biophys. Res. Commun., 254, 311–314.
113. Grande, S.W. etal. (2006) A dose-response study following in utero and
lactational exposure to di(2-ethylhexyl)phthalate: effects on female
rat reproductive development. Toxicol. Sci., 91, 247–254.
114. Kojima, H. et al. (2011) Comparative study of human and mouse
pregnane X receptor agonistic activity in 200 pesticides using in vitro
reporter gene assays. Toxicology, 280, 77–87.
115. (2006) Phosalone Reregistration Eligibility Decision (RED). The United
States Environmental Protection Agency Ofce of Pesticide Programs.
116. Li, X. etal. (2013) Structure-dependent activities of hydroxylated poly-
brominated diphenyl ethers on human estrogen receptor. Toxicology,
309, 15–22.
117. Berger, R.G. et al. (2014) Exposure to an environmentally relevant
mixture of brominated ame retardants affects fetal development in
Sprague-Dawley rats. Toxicology, 320, 56–66.
118. Hofmeister, M.V. etal. (2004) Effects of the pesticides prochloraz and
methiocarb on human estrogen receptor alpha and beta mRNA levels
analyzed by on-line RT-PCR. Toxicol. In Vitro, 18, 427–433.
119. Jacobsen, P.R. etal. (2012) Persistent developmental toxicity in rat off-
spring after low dose exposure to a mixture of endocrine disrupting
pesticides. Reprod. Toxicol., 34, 237–250.
120. Kamanga-Sollo, E. etal. (2008) Roles of IGF-I and the estrogen, andro-
gen and IGF-I receptors in estradiol-17beta- and trenbolone acetate-
stimulated proliferation of cultured bovine satellite cells. Domest.
Anim. Endocrinol., 35, 88–97.
121. Yarrow, J.F. etal. (2010) Tissue selectivity and potential clinical appli-
cations of trenbolone (17beta-hydroxyestra-4,9,11-trien-3-one): a
potent anabolic steroid with reduced androgenic and estrogenic
activity. Steroids, 75, 377–389.
122. Yan, H. etal. (2008) Exposure to bisphenol Aprenatally or in adult-
hood promotes T(H)2 cytokine production associated with reduction
of CD4CD25 regulatory T cells. Environ. Health Perspect., 116, 514–519.
123. Erden, E.S. etal. (2014) Investigation of Bisphenol Aas an endocrine
disruptor, total thiol, malondialdehyde, and C-reactive protein levels
in chronic obstructive pulmonary disease. Eur. Rev. Med. Pharmacol.
Sci., 18, 3477–3483.
124. Kharrazian, D. (2014) The potential roles of bisphenol A(BPA) patho-
genesis in autoimmunity. Autoimmune Dis., 2014, 743616.
125. Liu, Y. etal. (2014) Modulation of cytokine expression in human mac-
rophages by endocrine-disrupting chemical Bisphenol-A. Biochem.
Biophys. Res. Commun., 451, 592–598.
126. Rogers, J.A. etal. (2013) Review: endocrine disrupting chemicals and
immune responses: a focus on bisphenol-A and its potential mecha-
nisms. Mol. Immunol., 53, 421–430.
127. Deutschle, T. etal. (2008) A controlled challenge study on di(2-ethyl-
hexyl) phthalate (DEHP) in house dust and the immune response in
human nasal mucosa of allergic subjects. Environ. Health Perspect.,
116, 1487–1493.
128. Peltier, M.R. etal. (2012) Polybrominated diphenyl ethers enhance the
production of proinammatory cytokines by the placenta. Placenta,
33, 745–749.
129. Park, H.R. etal. (2014) Involvement of reactive oxygen species in bro-
minated diphenyl ether-47-induced inammatory cytokine release
from human extravillous trophoblasts in vitro. Toxicol. Appl. Pharma-
col., 274, 283–292.
130. Park, H.R. et al. (2014) Protective effect of nuclear factor E2-related
factor 2 on inammatory cytokine response to brominated diphe-
nyl ether-47 in the HTR-8/SVneo human rst trimester extravillous
trophoblast cell line. Toxicol. Appl. Pharmacol., 281, 67–77.
131. Koike, E. etal. (2014) Penta- and octa-bromodiphenyl ethers promote
proinammatory protein expression in human bronchial epithelial
cells in vitro. Toxicol. In Vitro, 28, 327–333.
132. Zhao, S. etal. (2013) Sub-acute exposure to the herbicide atrazine sup-
presses cell immune functions in adolescent mice. Biosci. Trends, 7,
193–201.
133. Rowe, A.M. etal. (2006) Immunomodulatory effects of maternal atra-
zine exposure on male Balb/c mice. Toxicol. Appl. Pharmacol., 214,
69–77.
134. Skinner, M.K. etal. (2007) Epigenetic transgenerational actions of vin-
clozolin on the development of disease and cancer. Crit. Rev. Oncog.,
13, 75–82.
135. Anway, M.D. etal. (2008) Transgenerational effects of the endocrine
disruptor vinclozolin on the prostate transcriptome and adult onset
disease. Prostate, 68, 517–529.
136. Cowin, P.A. etal. (2008) Early-onset endocrine disruptor-induced pros-
tatitis in the rat. Environ. Health Perspect., 116, 923–929.
137. Zhou, H.R. etal. (2003) Rapid, sequential activation of mitogen-acti-
vated protein kinases and transcription factors precedes proinam-
matory cytokine mRNA expression in spleens of mice exposed to the
trichothecene vomitoxin. Toxicol. Sci., 72, 130–142.
138. Shin, S.G. etal. (2005) Suppression of inducible nitric oxide synthase
and cyclooxygenase-2 expression in RAW 264.7 macrophages by ses-
quiterpene lactones. J. Toxicol. Environ. Health A, 68, 2119–2131.
139. Gollamudi, S. etal. (2012) Concordant signaling pathways produced
by pesticide exposure in mice correspond to pathways identied in
human Parkinson's disease. PLoS One, 7, e36191.
140. Morgan, J.B. etal. (2010) The marine sponge metabolite mycothiazole:
a novel prototype mitochondrial complex I inhibitor. Bioorg. Med.
Chem., 18, 5988–5994.
141. (1997) BASF Corporation Pyridaben (Sanmite) Pesticide Tolerance
Petition 3/97, US EPA [PF-721; FRL-5592 -7], http//pmep.cce.cornell.
edu/profiles/insect-mite/mevinphos-propargite/pyridaben/
pyridab_pet_0397.html (accessed 7 May 2015)
142. Barros, S.P. et al. (2010) Triclosan inhibition of acute and chronic
inammatory gene pathways. J. Clin. Periodontol., 37, 412–418.
143. Wallet, M.A. etal. (2013) Triclosan alters antimicrobial and inamma-
tory responses of epithelial cells. Oral Dis., 19, 296–302.
144. Bhargava, H.N. et al. (1996) Triclosan: applications and safety. Am.
J.Infect. Control, 24, 209–218.
145. Winitthana, T. etal. (2014) Triclosan potentiates epithelial-to-mesen-
chymal transition in anoikis-resistant human lung cancer cells. PLoS
One, 9, e110851.
146. Stoker, T.E. et al. (2010) Triclosan exposure modulates estrogen-
dependent responses in the female wistar rat. Toxicol. Sci., 117, 45–53.
147. Shioda, T. etal. (2006) Importance of dosage standardization for inter-
preting transcriptomal signature proles: evidence from studies of
xenoestrogens. Proc. Natl Acad. Sci. USA, 103, 12033–12038.
148. Welshons, W.V. etal. (1999) Low-dose bioactivity of xenoestrogens in ani-
mals: fetal exposure to low doses of methoxychlor and other xenoestro-
gens increases adult prostate size in mice. Toxicol. Ind. Health, 15, 12–25.
149. Alyea, R.A. etal. (2009) Differential regulation of dopamine transporter
function and location by low concentrations of environmental estro-
gens and 17beta-estradiol. Environ. Health Perspect., 117, 778–783.
150. Wozniak, A.L. etal. (2005) Xenoestrogens at picomolar to nanomolar
concentrations trigger membrane estrogen receptor-alpha-mediated
Ca2+ uxes and prolactin release in GH3/B6 pituitary tumor cells.
Environ. Health Perspect., 113, 431–439.
151. Jeng, Y.J. et al. (2011) Combinations of physiologic estrogens with
xenoestrogens alter ERK phosphorylation proles in rat pituitary
cells. Environ. Health Perspect., 119, 104–112.
152. Cabaton, N.J. etal. (2011) Perinatal exposure to environmentally rel-
evant levels of bisphenol Adecreases fertility and fecundity in CD-1
mice. Environ. Health Perspect., 119, 547–552.
153. Jones, B.A. et al. (2011) Pre- and postnatal bisphenol Atreatment
results in persistent decits in the sexual behavior of male rats, but
not female rats, in adulthood. Horm. Behav., 59, 246–251.
S290 | Carcinogenesis, 2015, Vol. 36, Supplement 1
154. Lemos, M.F. et al. (2010) Protein differential expression induced by
endocrine disrupting compounds in a terrestrial isopod. Chemos-
phere, 79, 570–576.
155. Markey, C.M. etal. (2001) The mouse uterotrophic assay: a reevalua-
tion of its validity in assessing the estrogenicity of bisphenol A. Envi-
ron. Health Perspect., 109, 55–60.
156. Filipov, N.M. etal. (2005) Manganese potentiates in vitro production of
proinammatory cytokines and nitric oxide by microglia through a
nuclear factor kappa B–dependent mechanism. Toxicol. Sci., 84, 139–
148.
157. Knudsen, T.B. etal. (2011) Disruption of embryonic vascular develop-
ment in predictive toxicology. Birth Defects Res. Part C, 93, 312–323.
158. Qin, R. et al. (2011) Protection by tetrahydroxystilbene glucoside
against neurotoxicity induced by MPP+: the involvement of PI3K/Akt
pathway activation. Toxicol. Lett., 202, 1–7.
159. Manfo, F.P. etal. (2011) Effects of maneb on testosterone release in
male rats. Drug Chem. Toxicol., 34, 120–128.
160. Matsushita, T. et al. (1976) Experimental study on contact dermati-
tis caused by dithiocarbamates maneb, mancozeb, zineb, and their
related compounds. Int. Arch. Occup. Environ. Health, 37, 169–178.
161. Barlow, B. et al. (2005) Modulation of antioxidant defense systems by
the environmental pesticide Maneb in dopaminergic cells. Neurotoxi-
col., 26, 63–75.
162. Kazantseva, Y.A. etal. (2013) Dichlorodiphenyltrichloroethane tech-
nical mixture regulates cell cycle and apoptosis genes through the
activation of CAR and ERα in mouse livers. Toxicol. Appl. Pharmacol.,
271, 137–143.
163. Lin, Z.X. etal. (1986) Inhibition of gap junctional intercellular com-
munication in human teratocarcinoma cells by organochlorine pesti-
cides. Toxicol. Appl. Pharmacol., 83, 10–19.
164. Ruch, R.J. et al. (1987) Inhibition of intercellular communication
between mouse hepatocytes by tumor promoters. Toxicol. Appl. Phar-
macol., 87, 111–120.
165. Ventura, C. etal. (2012) Differential mechanisms of action are involved
in chlorpyrifos effects in estrogen-dependent or -independent breast
cancer cells exposed to low or high concentrations of the pesticide.
Toxicol. Lett., 213, 184–193.
166. Mense, S.M. etal. (2006) The common insecticides cyfluthrin and
chlorpyrifos alter the expression of a subset of genes with diverse
functions in primary human astrocytes. Toxicol. Sci., 93, 125–135.
167. Santucci, M.A. etal. (2003) Cell-cycle deregulation in BALB/c 3T3 cells
transformed by 1,2-dibromoethane and folpet pesticides. Environ.
Mol. Mutagen., 41, 315–321.
168. Albanito, L. et al. (2008) G-protein-coupled receptor 30 and estrogen
receptor-alpha are involved in the proliferative effects induced by atra-
zine in ovarian cancer cells. Environ. Health Perspect., 116, 1648–1655.
169. Tsuda, H. etal. (2005) High susceptibility of human c-Ha-ras proto-
oncogene transgenic rats to carcinogenesis: a cancer-prone animal
model. Cancer Sci., 96, 309–316.
170. Wetzel, L.T. etal. (1994) Chronic effects of atrazine on estrus and
mammary tumor formation in female Sprague-Dawley and Fischer
344 rats. J. Toxicol. Environ. Health, 43, 169–182.
171. Andersson, H. etal. (2012) Proangiogenic effects of environmentally
relevant levels of bisphenol A in human primary endothelial cells.
Arch. Toxicol., 86, 465–474.
172. Dairkee, S.H. etal. (2013) Bisphenol-A-induced inactivation of the p53 axis
underlying deregulation of proliferation kinetics, and cell death in non-
malignant human breast epithelial cells. Carcinogenesis, 34, 703–712.
173. Betancourt, A.M. etal. (2012) Altered carcinogenesis and proteome in
mammary glands of rats after prepubertal exposures to the hormo-
nally active chemicals bisphenol Aand genistein. J. Nutr, 142, 1382S–
1388S.
174. Andrysík, Z. etal. (2013) Aryl hydrocarbon receptor-mediated disrup-
tion of contact inhibition is associated with connexin43 downregu-
lation and inhibition of gap junctional intercellular communication.
Arch. Toxicol., 87, 491–503.
175. Haber, L.T. et al. (2000) Hazard identication and dose response of
inhaled nickel-soluble salts. Regul. Toxicol. Pharmacol., 31, 210–230.
176. LN, V. et al. (2013) Low dose effects of bisphenol A: an integrated
review of in vitro, laboratory animal and human studies. Endocrine
Disruptors, 1, e1.1–e1.20.
177. Tryphonas, H. etal. (2004) Oral (gavage), in utero and post-natal expo-
sure of Sprague-Dawley rats to low doses of tributyltin chloride. Part
II: effects on the immune system. Food Chem. Toxicol., 42, 221–235.
178. Watanabe, J. etal. (2013) Low dose of methylmercury (MeHg) exposure
induces caspase mediated-apoptosis in cultured neural progenitor
cells. J. Toxicol. Sci., 38, 931–935.
179. Petroni, D. etal. (2012) Low-dose methylmercury-induced oxidative
stress, cytotoxicity, and tau-hyperphosphorylation in human neuro-
blastoma (SH-SY5Y) cells. Environ. Toxicol., 27, 549–555.
180. McCormack, A.L. et al. (2005) Role of oxidative stress in paraquat-
induced dopaminergic cell degeneration. J. Neurochem., 93, 1030–
1037.
181. Hartwig, A. etal. (2002) Interference by toxic metal ions with DNA
repair processes and cell cycle control: molecular mechanisms. Envi-
ron. Health Perspect., 110(Suppl 5), 797–799.
182. Asmuss, M. etal. (2000) Differential effects of toxic metal compounds
on the activities of Fpg and XPA, two zinc nger proteins involved in
DNA repair. Carcinogenesis, 21, 2097–2104.
183. McNeill, D.R. etal. (2004) Inhibition of Ape1 nuclease activity by lead,
iron, and cadmium. Environ. Health Perspect., 112, 799–804.
184. Pottier, G. etal. (2013) Lead exposure induces telomere instability in
human cells. PLoS One, 8, e67501.
185. Zhang, X. etal. (2013) Environmental and occupational exposure to
chemicals and telomere length in human studies. Occup. Environ.
Med., 70, 743–749.
186. Exon, J.H. (2006) A review of the toxicology of acrylamide. J. Toxicol.
Environ. Health B.Crit. Rev., 9, 397–412.
187. Sickles, D.W. etal. (2007) Acrylamide effects on kinesin-related pro-
teins of the mitotic/meiotic spindle. Toxicol. Appl. Pharmacol., 222,
111–121.
188. Wang, X. et al. (2013) Epigenotoxicity of environmental pollutants
evaluated by a combination of DNA methylation inhibition and capil-
lary electrophoresis-laser-induced uorescence immunoassay. Anal.
Bioanal. Chem., 405, 2435–2442.
189. Arita, A. etal. (2012) Global levels of histone modications in periph-
eral blood mononuclear cells of subjects with exposure to nickel.
Environ. Health Perspect., 120, 198–203.
190. Cantone, L. etal. (2011) Inhalable metal-rich air particles and histone
H3K4 dimethylation and H3K9 acetylation in a cross-sectional study
of steel workers. Environ. Health Perspect., 119, 964–969.
191. Chervona, Y. etal. (2012) Carcinogenic metals and the epigenome:
understanding the effect of nickel, arsenic, and chromium. Metal-
lomics, 4, 619–627.
192. Avissar-Whiting, M. etal. (2010) Bisphenol Aexposure leads to spe-
cic microRNA alterations in placental cells. Reprod. Toxicol., 29,
401–406.
193. Roedel, E.Q. etal. (2012) Pulmonary toxicity after exposure to military-
relevant heavy metal tungsten alloy particles. Toxicol. Appl. Pharma-
col., 259, 74–86.
194. Freyre-Fonseca, V. etal. (2011) Titanium dioxide nanoparticles impair
lung mitochondrial function. Toxicol. Lett., 202, 111–119.
195. Elhajouji, A. etal. (2011) Potential thresholds for genotoxic effects by
micronucleus scoring. Mutagenesis, 26, 199–204.
196. Ermler, S. etal. (2013) Seven benzimidazole pesticides combined at
sub-threshold levels induce micronuclei in vitro. Mutagenesis, 28,
417–426.
197. Sargent, L.M. et al. (2012) Single-walled carbon nanotube-induced
mitotic disruption. Mutat. Res., 745, 28–37.
198. Muller, J. etal. (2008) Clastogenic and aneugenic effects of multi-wall
carbon nanotubes in epithelial cells. Carcinogenesis, 29, 427–433.
199. Thomas, D. (2010) Gene–environment-wide association studies:
emerging approaches. Nat. Rev. Genet., 11, 259–272.
200. Santella, R.M. etal. (2005) DNA adducts, DNA repair genotype/pheno-
type and cancer risk. Mutat. Res., 592, 29–35.
201. (2011) Cytogenetic Dosimetry: Applications in Preparedness for and
Response to Radiation Emergencies. International Atomic Energy
Agency, Vienna.
202. De Lange, T. (2005) Telomere-related genome instability in cancer.
Cold Spring Harb. Symp. Quant. Biol., 70, 197–204.
203. Frias, C. etal. (2012) Telomere dysfunction and genome instability.
Front. Biosci. (Landmark Ed), 17, 2181–2196.
W.H.Goodson et al. | S291
204. Hollstein, M. etal. (1991) p53 mutations in human cancers. Science,
253, 49–53.
205. Jang, J.W. et al. (2006) Isoform-specic ras activation and oncogene
dependence during MYC- and Wnt-induced mammary tumorigen-
esis. Mol. Cell. Biol., 26, 8109–8121.
206. Muñoz, D.M. etal. (2013) Loss of p53 cooperates with K-ras activation
to induce glioma formation in a region-independent manner. Glia, 61,
1862–1872.
207. Pierotti, M.A. et al. Mechanisms of oncogene activation. Kufe DW, Pol-
lock RE, Weichselbaum RR, and et al. Holland-Frei Cancer Medicine.
6th. 2003. Hamilton (ON), BC Decker.
208. Mazzei, F. etal. (2013) Role of MUTYH in human cancer. Mutat. Res.,
743–744, 33–43.
209. Sancar, A. (1995) Excision repair in mammalian cells. J. Biol. Chem.,
270, 15915–15918.
210. Vineis, P. etal. (2009) A eld synopsis on low-penetrance variants in
DNA repair genes and cancer susceptibility. J. Natl Cancer Inst., 101,
24–36.
211. Bohacek, J. etal. (2013) Epigenetic inheritance of disease and disease
risk. Neuropsychopharmacology, 38, 220–236.
212. Esteller, M. (2007) Cancer epigenomics: DNA methylomes and his-
tone-modication maps. Nat. Rev. Genet., 8, 286–298.
213. Caffarelli, E. etal. (2011) Epigenetic regulation in cancer development.
Front. Biosci. (Landmark Ed), 16, 2682–2694.
214. Croce, C.M. (2009) Causes and consequences of microRNA dysregula-
tion in cancer. Nat. Rev. Genet., 10, 704–714.
215. Wang, Y. etal. (2013) MicroRNAs and DNA damage response: implica-
tions for cancer therapy. Cell Cycle, 12, 32–42.
216. Devalle, S. etal. (2012) Implications of aneuploidy for stem cell biology
and brain therapeutics. Front. Cell. Neurosci., 6, 36.
217. Linschooten, J.O. etal. (2013) Paternal lifestyle as a potential source of
germline mutations transmitted to offspring. FASEB J., 27, 2873–2879.
218. Leyns, L. etal. (2012) Genomic integrity of mouse embryonic stem
cells. In Embryogenesis. Intech. pp. 333–358.
219. Blessing, H. etal. (2004) Interaction of selenium compounds with zinc
nger proteins involved in DNA repair. Eur. J. Biochem., 271, 3190–
3199.
220. Zhang, X. etal. (2013) Environmental and occupational exposure to
chemicals and telomere length in human studies. Postgrad. Med. J.,
89, 722–728.
221. Lombaert, N. etal. (2013) Hard-metal (WC-Co) particles trigger a sign-
aling cascade involving p38 MAPK, HIF-1α, HMOX1, and p53 activation
in human PBMC. Arch. Toxicol., 87, 259–268.
222. Jugan, M.L. etal. (2012) Titanium dioxide nanoparticles exhibit geno-
toxicity and impair DNA repair activity in A549 cells. Nanotoxicology,
6, 501–513.
223. Song, M.F. etal. (2012) Metal nanoparticle-induced micronuclei and
oxidative DNA damage in mice. J. Clin. Biochem. Nutr., 50, 211–216.
224. Doshi, T. etal. (2011) Hypermethylation of estrogen receptor promoter
region in adult testis of rats exposed neonatally to bisphenol A. Toxi-
cology, 289, 74–82.
225. Kundakovic, M. et al. (2013) Sex-specic epigenetic disruption and
behavioral changes following low-dose in utero bisphenol Aexposure.
Proc. Natl Acad. Sci. USA, 110, 9956–9961.
226. Pupo, M. etal. (2012) Bisphenol Ainduces gene expression changes
and proliferative effects through GPER in breast cancer cells and can-
cer-associated broblasts. Environ. Health Perspect., 120, 1177–1182.
227. Ribeiro-Varandas, E. etal. (2013) Bisphenol Aat concentrations found
in human serum induces aneugenic effects in endothelial cells.
Mutat. Res., 751, 27–33.
228. Marshall, H. (2002) Fact sheet: carbendazim. Pesticides News, 57, 20–
21.
229. Zhao, Y. et al. (2010) Characterization and determination of chloro-
and bromo-benzoquinones as new chlorination disinfection byprod-
ucts in drinking water. Anal. Chem., 82, 4599–4605.
230. Piao, M.J. et al. (2011) Silver nanoparticles induce oxidative cell dam-
age in human liver cells through inhibition of reduced glutathione and
induction of mitochondria-involved apoptosis. Toxicol. Lett., 201, 92–100.
231. Gong, C. etal. (2012) Methylation of PARP-1 promoter involved in the
regulation of nano-SiO2-induced decrease of PARP-1 mRNA expres-
sion. Toxicol. Lett., 209, 264–269.
232. Choi, A.O. etal. (2008) Quantum dot-induced epigenetic and geno-
toxic changes in human breast cancer cells. J. Mol. Med. (Berl), 86,
291–302.
233. Balansky, R. et al. (2013) Transplacental clastogenic and epigenetic
effects of gold nanoparticles in mice. Mutat. Res., 751–752, 42–48.
234. Liu, Y. etal. (2013) Understanding the toxicity of carbon nanotubes.
Acc. Chem. Res., 46, 702–713.
235. Chisholm, H. (1910–1911) 11th Edition of Encyclopedia Britannica.
Cambridge University Press, Cambridge, UK.
236. Stoker, T.E. et al. (1999) Prepubertal exposure to compounds that
increase prolactin secretion in the male rat: effects on the adult pros-
tate. Biol. Reprod., 61, 1636–1643.
237. Ho, S.M. etal. (2006) Developmental exposure to estradiol and bisphe-
nol Aincreases susceptibility to prostate carcinogenesis and epige-
netically regulates phosphodiesterase type 4 variant 4. Cancer Res.,
66, 5624–5632.
238. Riu, A. et al. (2011) Characterization of novel ligands of ERalpha,
Erbeta, and PPARgamma: the case of halogenated bisphenol A and
their conjugated metabolites. Toxicol. Sci, 122, 372–82.
239. Riu, A. et al. (2011) Peroxisome proliferator-activated receptor γ is a
target for halogenated analogs of bisphenol A. Environ. Health Per-
spect., 119, 1227–1232.
240. Thueson, L.E. etal. (2015) In vitro exposure to the herbicide atrazine
inhibits T cell activation, proliferation, and cytokine production and
signicantly increases the frequency of Foxp3+ regulatory T cells.
Toxicol. Sci., 143, 418–429.
241. Hooghe, R.J. etal. (2000) Effects of selected herbicides on cytokine pro-
duction in vitro. Life Sci., 66, 2519–2525.
242. Filipov, N.M. etal. (2005) Immunotoxic effects of short-term atrazine
exposure in young male C57BL/6 mice. Toxicol. Sci., 86, 324–332.
243. Karrow, N.A. etal. (2005) Oral exposure to atrazine modulates
cell-mediated immune function and decreases host resistance to
the B16F10 tumor model in female B6C3F1 mice. Toxicology, 209,
15–28.
244. Basini, G. etal. (2012) Atrazine disrupts steroidogenesis, VEGF and
NO production in swine granulosa cells. Ecotoxicol. Environ. Saf., 85,
59–63.
245. Chen, J.Y. etal. (2013) Immunotoxicity of atrazine in Balb/c mice. J.
Environ. Sci. Health B., 48, 637–645.
246. Chen, J. etal. (2015) Effects of atrazine on the proliferation and cyto-
toxicity of murine lymphocytes with the use of carboxyuorescein
succinimidyl ester-based ow cytometric approaches. Food Chem.
Toxicol., 76, 61–69.
247. Danelli, L. etal. (2015) Mast cells boost myeloid-derived suppressor
cell activity and contribute to the development of tumor-favoring
microenvironment. Cancer Immunol. Res., 3, 85–95.
248. Grimm, E.A. et al. (2013) Molecular pathways: inammation-asso-
ciated nitric-oxide production as a cancer-supporting redox mech-
anism and a potential therapeutic target. Clin. Cancer Res., 19,
5557–5563.
249. Costa, A. etal. (2014) The role of reactive oxygen species and metabo-
lism on cancer cells and their microenvironment. Semin. Cancer
Biol., 25, 23–32.
250. Lei, Y. etal. (2015) Redox regulation of inammation: old elements, a
new story. Med. Res. Rev., 35, 306–340.
251. Wu, Y. etal. (2014) Molecular mechanisms underlying chronic inam-
mation-associated cancers. Cancer Lett., 345, 164–173.
252. Zhang, H.Y. etal. (2014) Perinatal exposure to 4-nonylphenol affects
adipogenesis in rst and second generation rats offspring. Toxicol.
Lett., 225, 325–332.
253. Mitchison, J. (1971) The Biology of the Cell Cycle. Cambridge Univer-
sity Press.
254. Keating, M.T. etal. (1988) Autocrine stimulation of intracellular PDGF
receptors in v-sis-transformed cells. Science, 239, 914–916.
255. Skobe, M. et al. (1998) Tumorigenic conversion of immortal human
keratinocytes through stromal cell activation. Proc. Natl Acad. Sci.
USA, 95, 1050–1055.
256. Lemmon, M.A. (2009) Ligand-induced ErbB receptor dimerization.
Exp. Cell Res., 315, 638–648.
257. Kerkhoff, E. etal. (1998) Cell cycle targets of Ras/Raf signalling. Onco-
gene, 17, 1457–1462.
S292 | Carcinogenesis, 2015, Vol. 36, Supplement 1
258. Mezquita, B. etal. (2014) Unlocking doors without keys: activation of
Src by truncated C-terminal intracellular receptor tyrosine kinases
lacking tyrosine kinase activity. Cells, 3, 92–111.
259. Grünfeld, H.T. etal. (2004) Effect of in vitro estrogenic pesticides on
human oestrogen receptor alpha and beta mRNA levels. Toxicol. Lett.,
151, 467–480.
260. Symonds, D.A. etal. (2005) Methoxychlor induces proliferation of the
mouse ovarian surface epithelium. Toxicol. Sci., 83, 355–362.
261. Murono, E.P. etal. (2004) The effects of the reported active metabo-
lite of methoxychlor, 2,2-bis(p-hydroxyphenyl)-1,1,1-trichloroethane,
on testosterone formation by cultured Leydig cells from young adult
rats. Reprod. Toxicol., 19, 135–146.
262. Gaido, K.W. etal. (2000) Interaction of methoxychlor and related com-
pounds with estrogen receptor alpha and beta, and androgen recep-
tor: structure-activity studies. Mol. Pharmacol., 58, 852–858.
263. Paulose, T. et al. (2011) Increased sensitivity of estrogen receptor
alpha overexpressing antral follicles to methoxychlor and its metabo-
lites. Toxicol. Sci., 120, 447–459.
264. Wilard, S. et al. (2009) Growth factors differentially augment the
effects of HPTE on estrogen response element-mediated gene tran-
scription in a dose- and time-dependent manner among human
breast cancer cell lines. Res. J.Med. Med. Sci. 4, 171–180.
265. Kojima, H. etal. (2010) Endocrine-disrupting potential of pesticides via
nuclear receptors and aryl hydrocarbon receptor. J. Health Sci., 56, 374–386.
266. Noriega, N.C. etal. (2005) Late gestational exposure to the fungicide
prochloraz delays the onset of parturition and causes reproductive
malformations in male but not female rat offspring. Biol. Reprod., 72,
1324–1335.
267. Cocco, P. (2002) On the rumors about the silent spring. Review of the scien-
tic evidence linking occupational and environmental pesticide exposure
to endocrine disruption health effects. Cad. Saude. Publica., 18, 379–402.
268. Kleinstreuer, N.C. et al. (2011) Identifying developmental toxicity
pathways for a subset of ToxCast chemicals using human embryonic
stem cells and metabolomics. Toxicol. Appl. Pharmacol., 257, 111–121.
269. Cummings, A.M. et al. (1989) Antifertility effect of methoxychlor in
female rats: dose- and time-dependent blockade of pregnancy. Toxi-
col. Appl. Pharmacol., 97, 454–462.
270. Gray, L.E. Jr etal. (1989) A dose-response analysis of methoxychlor-
induced alterations of reproductive development and function in the
rat. Fundam. Appl. Toxicol., 12, 92–108.
271. Metcalf, J.L. etal. (1996) Methoxychlor mimics the action of 17 beta-
estradiol on induction of uterine epidermal growth factor receptors in
immature female rats. Reprod. Toxicol., 10, 393–399.
272. Kuiper, G.G. et al. (1998) Interaction of estrogenic chemicals and phy-
toestrogens with estrogen receptor beta. Endocrinology, 139, 4252–4263.
273. Andersen, H.R. etal. (2002) Effects of currently used pesticides in
assays for estrogenicity, androgenicity, and aromatase activity in vitro.
Toxicol. Appl. Pharmacol., 179, 1–12.
274. Vinggaard, A.M. etal. (2006) Prochloraz: an imidazole fungicide with
multiple mechanisms of action. Int. J.Androl., 29, 186–192.
275. Liu, C. et al. (2011) Effects of prochloraz or propylthiouracil on the
cross-talk between the HPG, HPA, and HPT axes in zebrash. Environ.
Sci. Technol., 45, 769–775.
276. Zhang, W. etal. (2013) Known and emerging factors modulating estro-
genic effects of estrogen-disrupting chemicals. Environ. Rev., 21, 1–12.
277. Medjakovic, S. etal. (2014) Effect of nonpersistent pesticides on estro-
gen receptor, androgen receptor, and aryl hydrocarbon receptor. Envi-
ron. Toxicol., 29, 1201–1216.
278. Jenkins, S. etal. (2009) Oral exposure to bisphenol a increases dimeth-
ylbenzanthracene-induced mammary cancer in rats. Environ. Health
Perspect., 117, 910–915.
279. Goodson, W.H. III et al. (2011) Activation of the mTOR pathway by
low levels of xenoestrogens in breast epithelial cells from high-risk
women. Carcinogenesis, 32, 1724–1733.
280. Meironyté, D. etal. (1999) Analysis of polybrominated diphenyl ethers
in Swedish human milk. Atime-related trend study, 1972–1997. J. Tox-
icol. Environ. Health A, 58, 329–341.
281. Brown, D.J. etal. (2004) Analysis of Ah receptor pathway activation by
brominated ame retardants. Chemosphere, 55, 1509–1518.
282. Hsieh, T.H. etal. (2012) Phthalates induce proliferation and invasive-
ness of estrogen receptor-negative breast cancer through the AhR/
HDAC6/c-Myc signaling pathway. FASEB J., 26, 778–787.
283. Janjua, N.R. etal. (2007) Systemic uptake of diethyl phthalate, dibutyl
phthalate, and butyl paraben following whole-body topical applica-
tion and reproductive and thyroid hormone levels in humans. Envi-
ron. Sci. Technol., 41, 5564–5570.
284. Liu, W.L. et al. (2010) [Distribution characteristics of phthalic acid
esters in soils and plants at e-waste recycling sites in Taizhou of Zhe-
jiang, China]. Ying Yong Sheng Tai Xue Bao, 21, 489–494.
285. Wormuth, M. etal. (2006) What are the sources of exposure to eight
frequently used phthalic acid esters in Europeans? Risk Anal., 26,
803–824.
286. Galbraith, H. (2002) Hormones in international meat production: bio-
logical, sociological and consumer issues. Nutr. Res. Rev., 15, 293–314.
287. Boettcher, M. etal. (2011) Low-dose effects and biphasic effect pro-
les: is trenbolone a genotoxicant? Mutat. Res., 723, 152–157.
288. Kongsuwan, K. etal. (2012) The effect of combination treatment with
trenbolone acetate and estradiol-17β on skeletal muscle expression
and plasma concentrations of oxytocin in sheep. Domest. Anim.
Endocrinol., 43, 67–73.
289. Hotchkiss, A.K. et al. (2007) An environmental androgen, 17beta-
trenbolone, affects delayed-type hypersensitivity and reproductive
tissues in male mice. J. Toxicol. Environ. Health A, 70, 138–140.
290. Zhao, J.X. etal. (2011) Trenbolone enhances myogenic differentiation
by enhancing β-catenin signaling in muscle-derived stem cells of cat-
tle. Domest. Anim. Endocrinol., 40, 222–229.
291. Ansari, K.M. et al. (2010) Skin tumor promotion by argemone oil/
alkaloid in mice: evidence for enhanced cell proliferation, ornithine
decarboxylase, cyclooxygenase-2 and activation of MAPK/NF-kappaB
pathway. Food Chem. Toxicol., 48, 132–138.
292. Mishra, V. etal. (2012) Role of ErbB2 mediated AKT and MAPK pathway
in gall bladder cell proliferation induced by argemone oil and butter
yellow. Argemone oil and butter yellow induced gall bladder cell pro-
liferation. Cell Biol. Toxicol., 28, 149–159.
293. Parker, M. (1991) Nuclear Hormone Receptors: Molecular Mechanisms,
Cellular Functions, Clinical Abnormalities. Academic Press, London.
294. Gulliver, L.S.M. (2013) Estradiol synthesis and metabolism and risk
of ovarian cancer in older women taking prescribed or plant-derived
estrogen supplementation. J. Steroids Horm. Sci., S12:003.
295. Leroy, B. etal. (2014) TP53 mutations in human cancer: database reas-
sessment and prospects for the next decade. Hum. Mutat., 35, 672–688.
296. De Blasio, A. etal. (2005) Differentiative pathway activated by 3-amin-
obenzamide, an inhibitor of PARP, in human osteosarcoma MG-63
cells. FEBS Lett., 579, 615–620.
297. Nielsen, G.P. etal. (1998) CDKN2A gene deletions and loss of p16
expression occur in osteosarcomas that lack RB alterations. Am.
J.Pathol., 153, 159–163.
298. Pietruszewska, W. etal. (2008) Loss of heterozygosity for Rb locus and pRb
immunostaining in laryngeal cancer: a clinicopathologic, molecular and
immunohistochemical study. Folia Histochem. Cytobiol., 46, 479–485.
299. Myong, N.H. (2008) Cyclin D1 overexpression, p16 loss, and pRb inac-
tivation play a key role in pulmonary carcinogenesis and have a prog-
nostic implication for the long-term survival in non-small cell lung
carcinoma patients. Cancer Res. Treat., 40, 45–52.
300. Ikushima, H. etal. (2010) TGFbeta signalling: a complex web in cancer
progression. Nat. Rev. Cancer, 10, 415–424.
301. Su, V. etal. (2014) Connexins: mechanisms regulating protein levels
and intercellular communication. FEBS Lett., 588, 1212–1220.
302. Li, M.W. etal. (2010) Connexin 43 is critical to maintain the homeo-
stasis of the blood-testis barrier via its effects on tight junction reas-
sembly. Proc. Natl Acad. Sci. USA, 107, 17998–18003.
303. Campos-Pereira, F.D. etal. (2012) Early cytotoxic and genotoxic effects
of atrazine on Wistar rat liver: a morphological, immunohistochemi-
cal, biochemical, and molecular study. Ecotoxicol. Environ. Saf., 78,
170–177.
304. Stenner-Liewen, F. et al. (2003) Apoptosis and cancer: basic mecha-
nisms and therapeutic opportunities in the postgenomic era. Cancer
Res., 63, 263–268.
W.H.Goodson et al. | S293
305. Thompson, C.B. (1995) Apoptosis in the pathogenesis and treatment
of disease. Science, 267, 1456–1462.
306. Shortt, J. etal. (2012) Oncogenes in cell survival and cell death. Cold
Spring Harb. Perspect. Biol., 4, a009829.
307. Alberts, B. et al. (2002) Extracellular control of cell division, cell
growth, and apoptosis. In Molecular Biology of the Cell. Garland Sci-
ence, New York, NY.
308. Roos, W.P. etal. (2006) DNA damage-induced cell death by apoptosis.
Trends Mol. Med., 12, 440–450.
309. Fridman, J.S. et al. (2003) Control of apoptosis by p53. Oncogene, 22,
9030–9040.
310. Adams, J.M. (2003) Ways of dying: multiple pathways to apoptosis.
Genes Dev., 17, 2481–2495.
311. Deveraux, Q.L. etal. (1999) IAP family proteins–suppressors of apopto-
sis. Genes Dev., 13, 239–252.
312. Yang, Y.L. etal. (2000) The IAP family: endogenous caspase inhibitors
with multiple biological activities. Cell Res., 10, 169–177.
313. Wu, W. etal. (2013) Metabolic changes in cancer: beyond the Warburg
effect. Acta Biochim. Biophys. Sin. (Shanghai), 45, 18–26.
314. Gonzalez, M.J. etal. (2012) The bio-energetic theory of carcinogenesis.
Med. Hypotheses, 79, 433–439.
315. Ferreira, L.M. et al. (2012) Metabolic reprogramming of the tumor.
Oncogene, 31, 3999–4011.
316. Yi, C.H. etal. (2011) Metabolic regulation of protein N-alpha-acetyla-
tion by Bcl-xL promotes cell survival. Cell, 146, 607–620.
317. Wu, G.S. (2009) TRAIL as a target in anti-cancer therapy. Cancer Lett.,
285, 1–5.
318. Klaunig, J.E. etal. (1990) Gap-junctional intercellular communication
and murine hepatic carcinogenesis. Prog. Clin. Biol. Res., 331, 277–291.
319. Carette, D. etal. (2014) Connexin a check-point component of cell
apoptosis in normal and physiopathological conditions. Biochimie,
101, 1–9.
320. Leung-Toung, R. etal. (2006) Thiol proteases: inhibitors and potential
therapeutic targets. Curr. Med. Chem., 13, 547–581.
321. Kim, I.Y. etal. (2004) Phthalates inhibit tamoxifen-induced apoptosis
in MCF-7 human breast cancer cells. J. Toxicol. Environ. Health A, 67,
2025–2035.
322. Corcelle, E. etal. (2006) Disruption of autophagy at the maturation
step by the carcinogen lindane is associated with the sustained mito-
gen-activated protein kinase/extracellular signal-regulated kinase
activity. Cancer Res., 66, 6861–6870.
323. Kim, J.Y. etal. (2014) Methoxychlor and triclosan stimulates ovarian
cancer growth by regulating cell cycle- and apoptosis-related genes
via an estrogen receptor-dependent pathway. Environ. Toxicol. Phar-
macol., 37, 1264–1274.
324. Carnero, A. (2013) Markers of cellular senescence. Methods Mol. Biol.,
965, 63–81.
325. Serrano, M. etal. (2001) Putting the stress on senescence. Curr. Opin.
Cell Biol., 13, 748–753.
326. Shay, J.W. etal. (2004) Hallmarks of senescence in carcinogenesis and
cancer therapy. Oncogene, 23, 2919–2933.
327. Ohtani, N. et al. (2004) The p16INK4a-RB pathway: molecular link
between cellular senescence and tumor suppression. J. Med. Invest.,
51, 146–153.
328. Sherr, C.J. etal. (2002) The RB and p53 pathways in cancer. Cancer Cell,
2, 103–112.
329. Vergel, M. et al. (2010) Bypassing cellular senescence by genetic
screening tools. Clin. Transl. Oncol., 12, 410–417.
330. Zanella, F. etal. (2010) Understanding FOXO, new views on old tran-
scription factors. Curr. Cancer Drug Targets, 10, 135–146.
331. Ruiz, L. etal. (2008) Characterization of the p53 response to oncogene-
induced senescence. PLoS One, 3, e3230.
332. Newbold, R.F. etal. (1982) Induction of immortality is an early event
in malignant transformation of mammalian cells by carcinogens.
Nature, 299, 633–635.
333. Russo, I. etal. (1998) A telomere-independent senescence mechanism
is the sole barrier to Syrian hamster cell immortalization. Oncogene,
17, 3417–3426.
334. Newbold, R.F. etal. (1980) Mutagenicity of carcinogenic methylating
agents is associated with a specic DNA modication. Nature, 283,
596–599.
335. Lehman, T.A. etal. (1993) p53 mutations in human immortalized epi-
thelial cell lines. Carcinogenesis, 14, 833–839.
336. Lafarge-Frayssinet, C. etal. (1989) Over expression of proto-oncogenes:
ki-ras, fos and myc in rat liver cells treated in vitro by two liver tumor
promoters: phenobarbital and biliverdin. Cancer Lett., 44, 191–198.
337. Arita, A. etal. (2009) Epigenetics in metal carcinogenesis: nickel, arse-
nic, chromium and cadmium. Metallomics, 1, 222–228.
338. Trott, D.A. etal. (1995) Mechanisms involved in the immortalization
of mammalian cells by ionizing radiation and chemical carcinogens.
Carcinogenesis, 16, 193–204.
339. Rivedal, E. et al. (2000) Morphological transformation and effect on
gap junction intercellular communication in Syrian hamster embryo
cells as screening tests for carcinogens devoid of mutagenic activity.
Toxicol. In Vitro, 14, 185–192.
340. Zhou, X. etal. (2009) Effects of nickel, chromate, and arsenite on his-
tone 3 lysine methylation. Toxicol. Appl. Pharmacol., 236, 78–84.
341. Creton, S. etal. (2010) Acute toxicity testing of chemicals-Opportuni-
ties to avoid redundant testing and use alternative approaches. Crit.
Rev. Toxicol., 40, 50–83.
342. Dickens, F. etal. (1933) The metabolism of normal and tumour tis-
sue: The effects of lactate, pyruvate and deprivation of substrate. Bio-
chem. J., 27, 1134–1140.
343. MEDES, G. etal. (1953) Metabolism of neoplastic tissue. IV. Astudy
of lipid synthesis in neoplastic tissue slices in vitro. Cancer Res., 13,
27–29.
344. Menendez, J.A. etal. (2007) Fatty acid synthase and the lipogenic phe-
notype in cancer pathogenesis. Nat. Rev. Cancer, 7, 763–777.
345. Deberardinis, R.J. etal. (2008) Brick by brick: metabolism and tumor
cell growth. Curr. Opin. Genet. Dev., 18, 54–61.
346. Vander Heiden, M.G. etal. (2011) Metabolic pathway alterations that
support cell proliferation. Cold Spring Harb. Symp. Quant. Biol., 76,
325–334.
347. Currie, E. etal. (2013) Cellular fatty acid metabolism and cancer. Cell
Metab., 18, 153–161.
348. Kamphorst, J.J. et al. (2013) Hypoxic and Ras-transformed cells sup-
port growth by scavenging unsaturated fatty acids from lysophos-
pholipids. Proc. Natl Acad. Sci. USA, 110, 8882–8887.
349. Lazebnik, Y. (2010) What are the hallmarks of cancer? Nat. Rev. Can-
cer, 10, 232–233.
350. Berridge, M.V. etal. (2010) Metabolic exibility and cell hierarchy in
metastatic cancer. Mitochondrion, 10, 584–588.
351. Floor, S.L. etal. (2012) Hallmarks of cancer: of all cancer cells, all the
time? Trends Mol. Med., 18, 509–515.
352. Newsholme, E.A. et al. (1973) Regulation in Metabolism. Wiley, Lon-
don, New York.
353. Shim, H. etal. (1997) c-Myc transactivation of LDH-A: implications for
tumor metabolism and growth. Proc. Natl Acad. Sci. USA, 94, 6658–6663.
354. Fantin, V.R. etal. (2006) Attenuation of LDH-A expression uncovers a
link between glycolysis, mitochondrial physiology, and tumor main-
tenance. Cancer Cell, 9, 425–434.
355. Stanton, R.C. (2012) Glucose-6-phosphate dehydrogenase, NADPH,
and cell survival. IUBMB Life, 64, 362–369.
356. Raimundo, N. etal. (2011) Revisiting the TCA cycle: signaling to tumor
formation. Trends Mol. Med., 17, 641–649.
357. Mullen, A.R. etal. (2012) Genetically-dened metabolic reprogram-
ming in cancer. Trends Endocrinol. Metab., 23, 552–559.
358. Wallace, D.C. (2005) A mitochondrial paradigm of metabolic and
degenerative diseases, aging, and cancer: a dawn for evolutionary
medicine. Annu. Rev. Genet., 39, 359–407.
359. Santidrian, A.F. et al. (2013) Mitochondrial complex I activity and
NAD+/NADH balance regulate breast cancer progression. J. Clin.
Invest., 123, 1068–1081.
360. Lapuente-Brun, E. et al. (2013) Supercomplex assembly determines
electron ux in the mitochondrial electron transport chain. Science,
340, 1567–1570.
361. Zaidi, N. etal. (2013) Lipogenesis and lipolysis: the pathways exploited
by the cancer cells to acquire fatty acids. Prog. Lipid Res, 52, 585–589.
362. Weidberg, H. etal. (2009) Lipophagy: selective catabolism designed for
lipids. Dev. Cell, 16, 628–630.
363. Liu, K. et al. (2013) Regulation of lipid stores and metabolism by
lipophagy. Cell Death Differ., 20, 3–11.
S294 | Carcinogenesis, 2015, Vol. 36, Supplement 1
364. Newsholme, E.A. et al. (1991) Application of metabolic-control logic
to fuel utilization and its signicance in tumor cells. Adv. Enzyme
Regul., 31, 225–246.
365. Kalhan, S.C. etal. (2012) Resurgence of serine: an often neglected but
indispensable amino Acid. J. Biol. Chem., 287, 19786–19791.
366. Locasale, J.W. (2013) Serine, glycine and one-carbon units: cancer
metabolism in full circle. Nat. Rev. Cancer, 13, 572–583.
367. Cho, H.M. etal. (2000) Nucleotide sequence and differential expres-
sion of the human 3-phosphoglycerate dehydrogenase gene. Gene,
245, 193–201.
368. Owen, O.E. etal. (2002) The key role of anaplerosis and cataplerosis for
citric acid cycle function. J. Biol. Chem., 277, 30409–30412.
369. Meyerhof, O. (1951) Mechanisms of glycolysis and fermentation. Can.
J.Med. Sci., 29, 63–77.
370. Lowry, O.H. etal. (1964) The relationships between substrates and
enzymes of glycolysis in brain. J. Biol. Chem., 239, 31–42.
371. Warburg, O. etal. (1958) [Partial anaerobiosis and radiation-sensitivity
of cancer cells]. Arch. Biochem. Biophys., 78, 573–586.
372. Robey, R.B. (2011) On dogma and the metabolic gestalt of tumor
cells, Science, E-Letters. https://www.sciencemag.org/content/330/
6009/1338/reply (7 May 2015, date last accessed).
373. Copley, S.D. (2003) Enzymes with extra talents: moonlighting func-
tions and catalytic promiscuity. Curr. Opin. Chem. Biol., 7, 265–272.
374. Luo, W. etal. (2012) Emerging roles of PKM2 in cell metabolism and
cancer progression. Trends Endocrinol. Metab., 23, 560–566.
375. Gao, X. etal. (2012) Pyruvate kinase M2 regulates gene transcription
by acting as a protein kinase. Mol. Cell, 45, 598–609.
376. Robey, R.B. etal. (2006) Mitochondrial hexokinases, novel mediators
of the antiapoptotic effects of growth factors and Akt. Oncogene, 25,
4683–4696.
377. Hu, J. etal. (2013) Heterogeneity of tumor-induced gene expression
changes in the human metabolic network. Nat. Biotechnol., 31, 522–529.
378. Kacser, H. et al. (1973) The control of ux. Symp. Soc. Exp. Biol., 27,
65–104.
379. Agarwal, A.R. etal. (2013) Metabolic shift in lung alveolar cell mito-
chondria following acrolein exposure. Am. J.Physiol. Lung Cell. Mol.
Physiol., 305, L764–L773.
380. Ishida, S. etal. (2013) Bioavailable copper modulates oxidative phos-
phorylation and growth of tumors. Proc. Natl Acad. Sci. USA, 110,
19507–19512.
381. George, J. etal. (2011) Cypermethrin exposure leads to regulation of
proteins expression involved in neoplastic transformation in mouse
skin. Proteomics, 11, 4411–4421.
382. Tsitsimpikou, C. etal. (2013) Histopathological lesions, oxidative stress
and genotoxic effects in liver and kidneys following long term expo-
sure of rabbits to diazinon and propoxur. Toxicology, 307, 109–114.
383. Abdollahi, M. et al. (2004) Pesticides and oxidative stress: a review.
Med. Sci. Monit., 10, RA141–RA147.
384. Folkman, J. (1995) Angiogenesis in cancer, vascular, rheumatoid and
other disease. Nat. Med., 1, 27–31.
385. Carmeliet, P. (2005) Angiogenesis in life, disease and medicine. Nature,
438, 932–936.
386. Folkman, J. (1971) Tumor angiogenesis: therapeutic implications. N.
Engl. J.Med., 285, 1182–1186.
387. Carmeliet, P. etal. (2011) Molecular mechanisms and clinical applica-
tions of angiogenesis. Nature, 473, 298–307.
388. Ferrara, N. etal. (2005) Angiogenesis as a therapeutic target. Nature,
438, 967–974.
389. Folkman, J. (2003) Fundamental concepts of the angiogenic process.
Curr. Mol. Med., 3, 643–651.
390. Folkman, J. (2003) Angiogenesis inhibitors: a new class of drugs. Can-
cer Biol. Ther., 2, S127–S133.
391. Siemann, D.W. etal. (2005) Differentiation and denition of vascular-
targeted therapies. Clin. Cancer Res., 11(2 Pt 1), 416–420.
392. Thorpe, P.E. (2004) Vascular targeting agents as cancer therapeutics.
Clin. Cancer Res., 10, 415–427.
393. Patterson, D.M. etal. (2007) Vascular damaging agents. Clin. Oncol. (R.
Coll. Radiol.), 19, 443–456.
394. Hu, Z. etal. (1999) Targeting tumor vasculature endothelial cells and
tumor cells for immunotherapy of human melanoma in a mouse
xenograft model. Proc. Natl Acad. Sci. USA, 96, 8161–8166.
395. Hu, Z. etal. (2000) Intratumoral injection of adenoviral vectors encod-
ing tumor-targeted immunoconjugates for cancer immunotherapy.
Proc. Natl Acad. Sci. USA, 97, 9221–9225.
396. Hu, Z. etal. (2001) Targeting tissue factor on tumor vascular endothe-
lial cells and tumor cells for immunotherapy in mouse models of pro-
static cancer. Proc. Natl Acad. Sci. USA, 98, 12180–12185.
397. Hu, Z. etal. (2010) Natural killer cells are crucial for the efcacy of Icon
(factor VII/human IgG1 Fc) immunotherapy in human tongue cancer.
BMC Immunol., 11, 49.
398. Konigsberg, W.H. et al. (1988) Molecular cloning of the cDNA for
human tissue factor. Cell, 52, 639–640.
399. Contrino, J. etal. (1996) In situ detection of tissue factor in vascu-
lar endothelial cells: correlation with the malignant phenotype of
human breast disease. Nat. Med., 2, 209–215.
400. Hu, Z. (2011) Factor VII-Targeted Photodynamic Therapy for Breast
Cancer and Its Therapeutic Potential for Other Solid Cancers and
Leukemia, Breast Cancer—Current and Alternative Therapeutic
Modalities. In Esra Gunduz and Mehmet Gunduz (eds), InTech, Rijeka,
Croatia. http://www.intechopen.com/articles/show/title/factor-vii-
targeted-photodynamic-therapy-for-breast-cancer-and-its-therapeu-
tic-potential-for-other-s. ISBN: 978-953-307-776-5.
401. Hu, Z. etal. (2010) Targeting tissue factor on tumour cells and angio-
genic vascular endothelial cells by factor VII-targeted verteporn
photodynamic therapy for breast cancer in vitro and in vivo in mice.
BMC Cancer, 10, 235.
402. Duanmu, J. etal. (2011) Effective treatment of chemoresistant breast
cancer in vitro and in vivo by a factor VII-targeted photodynamic ther-
apy. Br. J.Cancer, 104, 1401–1409.
403. Hu, Z. etal. (2011) Selective and effective killing of angiogenic vascu-
lar endothelial cells and cancer cells by targeting tissue factor using
a factor VII-targeted photodynamic therapy for breast cancer. Breast
Cancer Res. Treat., 126, 589–600.
404. Jessen-Eller, K.K.J. etal. (2002) A new invertebrate member of the p53
gene family is developmentally expressed and responds to polychlo-
rinated biphenyls. Environ. Health Perspect., 110, 9.
405. Taylor, T.R. et al. (2011) Ziram activates mitogen-activated protein
kinases and decreases cytolytic protein levels in human natural killer
cells. Toxicol. Mech. Methods, 21, 8.
406. Berx, G. etal. (2009) Involvement of members of the cadherin super-
family in cancer. Cold Spring Harb. Perspect. Biol., 1, a003129.
407. Klymkowsky, M.W. etal. (2009) Epithelial-mesenchymal transition: a can-
cer researcher's conceptual friend and foe. Am. J.Pathol., 174, 1588–1593.
408. Micalizzi, D.S. etal. (2010) Epithelial-mesenchymal transition in can-
cer: parallels between normal development and tumor progression. J.
Mammary Gland Biol. Neoplasia, 15, 117–134.
409. Yang, J. etal. (2008) Epithelial-mesenchymal transition: at the cross-
roads of development and tumor metastasis. Dev. Cell, 14, 818–829.
410. Kawata, M. et al. (2012) TGF-β-induced epithelial-mesenchymal
transition of A549 lung adenocarcinoma cells is enhanced by pro-
inammatory cytokines derived from RAW 264.7 macrophage cells. J.
Biochem., 151, 205–216.
411. Nagase, H. etal. (2006) Structure and function of matrix metallopro-
teinases and TIMPs. Cardiovasc. Res., 69, 562–573.
412. Chen, H.C. et al. (1998) Tyrosine phosphorylation of focal adhesion
kinase stimulated by hepatocyte growth factor leads to mitogen-acti-
vated protein kinase activation. J. Biol. Chem., 273, 25777–25782.
413. Al-Mehdi, A.B. etal. (2000) Intravascular origin of metastasis from the
proliferation of endothelium-attached tumor cells: a new model for
metastasis. Nat. Med., 6, 100–102.
414. Chang, C.C. etal. (2013) Connective tissue growth factor activates
pluripotency genes and mesenchymal-epithelial transition in head
and neck cancer cells. Cancer Res., 73, 4147–4157.
415. McCormick, J.M. et al. (2010) Embryonic exposure to tetrabromobis-
phenol Aand its metabolites, bisphenol Aand tetrabromobisphenol
A dimethyl ether disrupts normal zebrash (Danio rerio) develop-
ment and matrix metalloproteinase expression. Aquat. Toxicol., 100,
255–262.
416. Ding, S.Z. et al. (2013) Epithelial-mesenchymal transition during
oncogenic transformation induced by hexavalent chromium involves
reactive oxygen species-dependent mechanism in lung epithelial
cells. Toxicol. Appl. Pharmacol., 269, 61–71.
W.H.Goodson et al. | S295
417. Hsiang, C.Y. etal. (2007) Acetaldehyde induces matrix metalloprotein-
ase-9 gene expression via nuclear factor-kappaB and activator pro-
tein 1 signaling pathways in human hepatocellular carcinoma cells:
association with the invasive potential. Toxicol. Lett., 171, 78–86.
418. Seo, J.H. etal. (2004) Helicobacter pylori in a Korean isolate activates
mitogen-activated protein kinases, AP-1, and NF-kappaB and induces
chemokine expression in gastric epithelial AGS cells. Lab. Invest., 84,
49–62.
419. Olumi, A.F. etal. (1999) Carcinoma-associated broblasts direct tumor
progression of initiated human prostatic epithelium. Cancer Res., 59,
5002–5011.
420. Le Bitoux, M.A. et al. (2008) Tumor-host interactions: the role of
inammation. Histochem. Cell Biol., 130, 1079–1090.
421. Mafni, M.V. etal. (2005) Stromal regulation of neoplastic develop-
ment: age-dependent normalization of neoplastic mammary cells by
mammary stroma. Am. J.Pathol., 167, 1405–1410.
422. Weaver, V.M. etal. (2004) Watch thy neighbor: cancer is a communal
affair. J. Cell Sci., 117, 1287–1290.
423. Maguer-Satta, V. (2011) The stem cell niche: the black master of can-
cer. Cancer Stem Cells Theories and Practices, ISBN 978-953-307-225-
8.
424. Laconi, S. et al. (2001) A growth-constrained environment drives
tumor progression invivo. Proc. Natl Acad. Sci. USA, 98, 7806–7811.
425. Ding, J. etal. (2009) TNF-alpha induction by nickel compounds is spe-
cic through ERKs/AP-1-dependent pathway in human bronchial epi-
thelial cells. Curr. Cancer Drug Targets, 9, 81–90.
426. Li, J. etal. (2004) Nickel compounds act through phosphatidylinositol-
3-kinase/Akt-dependent, p70(S6k)-independent pathway to induce
hypoxia inducible factor transactivation and Cap43 expression in
mouse epidermal Cl41 cells. Cancer Res., 64, 94–101.
427. Ding, J. etal. (2006) Nickel compounds render anti-apoptotic effect to
human bronchial epithelial Beas-2B cells by induction of cyclooxy-
genase-2 through an IKKbeta/p65-dependent and IKKalpha- and
p50-independent pathway. J. Biol. Chem., 281, 39022–39032.
428. Zhang, J. etal. (2013) The alteration of miR-222 and its target genes in
nickel-induced tumor. Biol. Trace Elem. Res., 152, 267–274.
429. Allard, P. et al. (2010) Bisphenol A impairs the double-strand break
repair machinery in the germline and causes chromosome abnor-
malities. Proc. Natl Acad. Sci. USA, 107, 20405–20410.
430. Hassan, Z.K. etal. (2012) Bisphenol Ainduces hepatotoxicity through
oxidative stress in rat model. Oxid. Med. Cell. Longev., 2012, 194829.
431. Takahashi, A. et al. (2004) Bisphenol A from dental polycarbonate
crown upregulates the expression of hTERT. J. Biomed. Mater. Res.
B.Appl. Biomater., 71, 214–221.
432. Hurt, K. etal. (2013) Tributyltin and dibutyltin alter secretion of tumor
necrosis factor alpha from human natural killer cells and a mixture
of T cells and natural killer cells. J. Appl. Toxicol., 33, 503–510.
433. Patel, E. etal. (2013) Methylmercury impairs motor function in early
development and induces oxidative stress in cerebellar granule cells.
Toxicol. Lett., 222, 265–272.
434. Sherwani, S.I. etal. (2013) Eicosanoid signaling and vascular dysfunc-
tion: methylmercury-induced phospholipase D activation in vascular
endothelial cells. Cell Biochem. Biophys., 67, 317–329.
435. Black, A.T. etal. (2008) Increased oxidative stress and antioxidant
expression in mouse keratinocytes following exposure to paraquat.
Toxicol. Appl. Pharmacol., 231, 384–392.
436. Chang, X. etal. (2013) Paraquat inhibits cell viability via enhanced oxi-
dative stress and apoptosis in human neural progenitor cells. Chem.
Biol. Interact., 206, 248–255.
437. Khatami, M. (2014) Chronic inammation: synergistic interactions of
recruiting macrophages (TAMs) and eosinophils (Eos) with host mast
cells (MCs) and tumorigenesis in CALTs. M-CSF, suitable biomarker
for cancer diagnosis! Cancers (Basel), 6, 297–322.
438. Whiteside, T.L. (2006) Immune suppression in cancer: effects on
immune cells, mechanisms and future therapeutic intervention.
Semin Cancer Biol., 16, 3–15.
439. Whiteside, T.L. (2002) Tumor-induced death of immune cells: its
mechanisms and consequences. Semin Cancer Biol., 12, 43–50.
440. Yang, L. etal. (2010) TGF-β and immune cells: an important regulatory
axis in the tumor microenvironment and progression. Trends Immu-
nol., 31, 220–227.
441. Mocellin, S. etal. (2004) The multifaceted relationship between IL-10
and adaptive immunity: putting together the pieces of a puzzle.
Cytokine Growth Factor Rev., 15, 61–76.
442. Zou, W. (2005) Immunosuppressive networks in the tumour environ-
ment and their therapeutic relevance. Nat. Rev. Cancer, 5, 263–274.
443. IARC (2013) IARC Monographs on the Evaluation of Carcinogenic
Risks to Humans. Agents classied by the IARC Monographs. Interna-
tional Agency for Research and Cancer, Lyon.
444. IARC (1982) Monograph on the Evaluation of Carcinogenic Risk to
Humans. Some Industrial Chemicals and Dyestuffs. Di(2-ethylhexyl)
Phthalate. International Agency for Research and Cancer, Lyon, pp.
257–280.
445. IARC (2000) Monograph on the Evaluation of Carcinogenic Risk to
Humans. Some Industrial Chemicals. Di(2-ethylhexyl) Phthalate.
International Agency for Research and Cancer, Lyon, pp. 41–148.
446. Odermatt, A. etal. (2008) Disruption of glucocorticoid and mineralo-
corticoid receptor-mediated responses by environmental chemicals.
CHIMIA Int. J.Chem., 62, 335–339.
447. Santos, P.M. etal. (2009) Insights into yeast adaptive response to the
agricultural fungicide mancozeb: a toxicoproteomics approach. Prot-
eomics, 9, 657–670.
448. Kuster, M. etal. (2009) Liquid chromatography–tandem mass spectro-
metric analysis and regulatory issues of polar pesticides in natural
and treated waters. J. Chromatogr. A, 1216, 520–529.
449. Kim, J.H. etal. (2013) Synergism of antifungal activity between mito-
chondrial respiration inhibitors and kojic acid. Molecules, 18, 1564–
1581.
450. Çayır, A. etal. (2014) Micronuclei, nucleoplasmic bridges, and nuclear
buds induced in human lymphocytes by the fungicide signum and its
active ingredients (boscalid and pyraclostrobin). Environ. Toxicol, 29,
723–732.
451. Judson, R.S. et al. (2010) In vitro screening of environmental chemi-
cals for targeted testing prioritization: the ToxCast project. Environ.
Health Persp. (Online), 118, 485.
452. Pereboeva, L. etal. (2013) DNA damage responses and oxidative stress
in dyskeratosis congenita. PLoS One, 8, e76473.
453. Pinchuk, L.M. etal. (2007) In vitro atrazine exposure affects the phe-
notypic and functional maturation of dendritic cells. Toxicol. Appl.
Pharmacol., 223, 206–217.
454. Kavlock, R. etal. (2012) Update on EPA’s ToxCast program: providing
high throughput decision support tools for chemical risk manage-
ment. Chem. Res. Toxicol., 25, 1287–1302.
455. Martin, M.T. etal. (2011) Predictive model of rat reproductive toxicity
from ToxCast high throughput screening. Biol. Reprod., 85, 327–339.
456. Metzler, M. et al. (2001) Chemistry of natural and anthropogenic
endocrine active compounds. In M.Metzler (ed) The Handbook of
Environmental Chemistry Vol. 3, Part L. Endocrine Disruptors–Part I.
Springer, Berlin Heidelberg, pp. 63–80.
457. Gatidou, G. etal. (2007) Simultaneous determination of the endo-
crine disrupting compounds nonylphenol, nonylphenol ethoxy-
lates, triclosan and bisphenol Ain wastewater and sewage sludge
by gas chromatography–mass spectrometry. J. Chromatogr. A, 1138,
32–41.
458. Foran, C. et al. (2000) Developmental evaluation of a potential non-
steroidal estrogen: triclosan. Mar. Environ. Res., 50, 153–156.
459. Ishibashi, H. etal. (2004) Effects of triclosan on the early life stages
and reproduction of medaka Oryzias latipes and induction of hepatic
vitellogenin. Aquat. Toxicol., 67, 167–179.
460. Boyd, G.R. et al. (2003) Pharmaceuticals and personal care products
(PPCPs) in surface and treated waters of Louisiana, USA and Ontario,
Canada. Sci. Total Environ., 311, 135–149.
461. Kolpin, D.W. et al. (2002) Pharmaceuticals, hormones, and other
organic wastewater contaminants in US streams, 1999–2000: a
national reconnaissance. Environ. Sci. Technol., 36, 1202–1211.
462. Vandhana, S. etal. (2013) Biochemical changes accompanying apop-
totic cell death in retinoblastoma cancer cells treated with lipogenic
enzyme inhibitors. Biochim. Biophys. Acta (BBA)-Molecular and Cell
Biology of Lipids, 1831, 1458–1466.
463. Zuckerbraun, H.L. etal. (1998) Triclosan, cytotoxicity, mode of action,
and induction of apoptosis in human gingival cells in vitro. Eur. J.Oral
Sci., 106, 628–636.
S296 | Carcinogenesis, 2015, Vol. 36, Supplement 1
464. Terasaka, H. etal. (2005) Cytotoxicity and apoptosis-inducing activity
of bisphenol Aand hydroquinone in HL-60 cells. Anticancer Res., 25,
2241–2247.
465. The Australian National Industrial Chemicals Notication and
Assessment Scheme (2006) Diethylhexyl Phthalate (DEHP) Factsheet.
CAS: 117-81-7. http://www.nicnas.gov.au/communications/publica-
tions/information-sheets/existing-chemical-info-sheets/diethyl-
hexyl-phthalate-dehp-factsheet (7 May 2015, date last accessed).
466. Jobling, S. etal. (1995) A variety of environmentally persistent chemi-
cals, including some phthalate plasticizers, are weakly estrogenic.
Environ. Health Persp., 103, 582.
467. Hao, C. et al. (2012) Perinatal exposure to diethyl-hexyl-phthalate
induces obesity in mice. Front. Biosci. (Elite edition), 5, 725–733.
468. Moushumi Priya, A. etal. (2012) Induction of Apoptosis and cell cycle
arrest by Bis (2-ethylhexyl) phthalate produced by Marine Bacillus
pumilus MB 40. Chem. Biol. Interact., 195, 133–143.
469. Park, M.A. etal. (2012) Cell growth of BG-1 ovarian cancer cells is
promoted by di-n-butyl phthalate and hexabromocyclododecane via
upregulation of the cyclin D and cyclin-dependent kinase-4 genes.
Mol. Med. Rep., 5, 761–766.
470. Kinoshita, Y. etal. (2003) Induction of aromatase (CYP19) expression
in breast cancer cells through a nongenomic action of estrogen recep-
tor alpha. Cancer Res., 63, 3546–3555.
471. Prins, G.S. (2008) Endocrine disruptors and prostate cancer risk.
Endocr. Relat. Cancer, 15, 649–656.
472. Laville, N. etal. (2006) Modulation of aromatase activity and mRNA by
various selected pesticides in the human choriocarcinoma JEG-3 cell
line. Toxicology, 228, 98–108.
473. Bulun, S.E. etal. (2007) Aromatase excess in cancers of breast, endo-
metrium and ovary. J. Steroid Biochem. Mol. Biol., 106, 81–96.
474. (2012) A Review of Human Carcinogens—Pharmaceuticals IARC Mon-
ographs on the Evaluation of Caricinogenic Risk to Humans. Vol. 100,
World health Organization IARC, Geneva, Switzerland.
475. Paul, A. etal. (2014) The breast cancer susceptibility genes (BRCA) in
breast and ovarian cancers. Front. Biosci. (Landmark Ed), 19, 605–618.
476. Supplementary Guidance for Conducting Health Risk Assessment
of Chemical Mixtures (2000) U.S. Environmental Protection Agency
Report No. EPA/630/R-00/002, Washington, DC.
477. Dellarco, V. et al. (2012) Mode of action: moving toward a more rel-
evant and efcient assessment paradigm. J. Nutr., 142, 2192S–2198S.
478. Meek, M.E. etal. (2003) A framework for human relevance analysis of
information on carcinogenic modes of action. Crit. Rev. Toxicol., 33,
591–653.
479. Boobis, A.R. et al. (2006) IPCS framework for analyzing the rele-
vance of a cancer mode of action for humans. Crit. Rev. Toxicol., 36,
781–792.
480. (2012) OECD Guidance Document 116 On The Conduct And Design Of
Chronic Toxicity And Carcinogenicity Studies, Supporting Test Guide-
lines 451, 452 And 453, 2nd Edition, Series on Testing and Assessment,
ENV/JM/MONO(2011)47. OECD Environment Directorate Joint Meeting
of the Chemicals Committee and the Working Party on Chemicals,
Pesticides and Biotechnology, Paris, France.
481. U.S.E.P.A. (2002) Guidance on Cumulative Risk Assessment of Pesti-
cide Chemicals That Have a Common Mechanism of Toxicity. In Ofce
of Pesticide Programs, Washington, D.C. 20460.
482. EFSA PPR Panel (EFSA Panel on Plant Protection Products and their
Residues) (2013). Scientic opinion on the relevance of dissimilar
mode of action and its appropriate application for cumulative risk
assessment of pesticides residues in food. EFSA J., 11, 40.
483. Driedger, A.A. etal. (1971) Demonstration of two types of DNA repair in
X-irradiated Micrococcus radiodurans. Can. J.Microbiol., 17, 495–499.
484. Mason, P.A. etal. (2003) Mismatch repair activity in mammalian mito-
chondria. Nucleic Acids Res., 31, 1052–1058.
485. Heindorff, K. etal. (1983) Genetic toxicology of ethylenediaminetet-
raacetic acid (EDTA). Mutat. Res., 115, 149–173.
486. Cory-Slechta, D. et al. (2008) Phthalates Cumulative Risk Assess-
ment—The Tasks Ahead. In National Research Council, N.A.o.S., Board
on Environmental Science and Technology, Committee on Phthalates
Health Risks. National Academy Press, Washington, DC, p. 208.
487. Bucala, R. (1996) MIF rediscovered: cytokine, pituitary hormone, and
glucocorticoid-induced regulator of the immune response. FASEB J.,
10, 1607–1613.
488. Bucala, R. et al. (2007) Macrophage migration inhibitory factor: a
probable link between inammation and cancer. Immunity, 26, 281–
285.
489. Grieb, G. etal. (2010) Macrophage migration inhibitory factor (MIF): a
promising biomarker. Drug News Perspect., 23, 257–264.
490. Landesmann, B. etal. (2013) Adverse outcome pathway-based screen-
ing strategies for an animal-free safety assessment of chemicals.
Altern. Lab. Anim., 41, 461–471.
491. Malhotra, J. etal. (2015) Effect of occupational exposures on lung can-
cer susceptibility: a study of gene-environment interaction analysis.
Cancer Epidemiol. Biomarkers Prev., 24, 570–579.
492. Singh, S. etal. (2012) Epigenetic effects of environmental chemicals
bisphenol a and phthalates. Int. J.Mol. Sci., 13, 10143–10153.
493. Vineis, P. etal. (2010) Models of carcinogenesis: an overview. Carcino-
genesis, 31, 1703–1709.
494. Brash, D. et al. (2009) The mysterious steps in carcinogenesis. Br.
J.Cancer, 101, 379–380.
495. Brash, D. etal. (2009) The mysterious steps in carcinogenesis: adden-
dum. Br. J.Cancer, 101, 1490.
496. Rappaport, S.M. etal. (2010) Epidemiology. Environment and disease
risks. Science, 330, 460–461.
497. Bisson, W.H. (2012) Editorial: computational chemogenomics in drug
design and discovery. Curr. Top. Med. Chem., 12, 1867–1868.
498. Tomasetti, C. et al. (2015) Cancer etiology. Variation in cancer risk
among tissues can be explained by the number of stem cell divisions.
Science, 347, 78–81.
499. Weinberg, R.A. (2014) Coming full circle-from endless complexity to
simplicity and back again. Cell, 157, 267–271.
500. Koutsogiannouli, E. etal. (2013) Complexity in cancer biology: is sys-
tems biology the answer? Cancer Med., 2, 164–177.
501. Alberghina, L. etal. (2004) Systems biology and the molecular circuits
of cancer. Chembiochem, 5, 1322–1333.
502. Carvalho, R.N. et al. (2014) Mixtures of chemical pollutants at Euro-
pean legislation safety concentrations: how safe are they? Toxicol.
Sci., 141, 218–233.
503. Porter, W.P. etal. (1999) Endocrine, immune, and behavioral effects of
aldicarb (carbamate), atrazine (triazine) and nitrate (fertilizer) mix-
tures at groundwater concentrations. Toxicol. Ind. Health, 15, 133–150.
504. Tarone, R.E. etal. (2011) Combating environmental causes of cancer.
N. Engl. J.Med., 364, 2266–2267.
505. Willett, W.C. etal. (2011) Combating environmental causes of cancer.
N. Engl. J.Med., 364, 2266.
506. Richter, E.D. etal. (2004) The precautionary principle, epidemiology
and the ethics of delay. Int. J.Occup. Med. Environ. Health, 17, 9–16.
507. Pohl, H.R. et al. (2010) Chemical risk assessment and uncertainty
associated with extrapolation across exposure duration. Regul. Toxi-
col. Pharmacol., 57, 18–23.
508. Tice, R.R. et al. (2013) Improving the human hazard characteriza-
tion of chemicals: a Tox21 update. Environ. Health Perspect., 121,
756–765.
... The WHO has established that most inhabitants of developing countries such as Africa depend on crop residue, wood, and animal dung for cooking and heating. Since biomass stoves are noted for bio-carcinogen emission, women become exposed to smoke from these stoves and hence stand higher chances of developing CC [57,58]. ...
Article
Full-text available
Cervical and prostate cancer account for 7.1 and 7.3 deaths per 100,000 people globally in 2022. These rates increased significantly to 17.6 and 17.3 in Africa, respectively, making them the second and third leading cause of cancer deaths in Africa, only surpassed by breast cancer. The human papillomavirus is the prime risk factor for cervical cancer infection. On the other hand, pros-tate cancer risks include ageing, genetics, race, geography, and family history. However, these factors alone cannot account for the high mortality rate in Africa, which is more than twice the global mortality rate for the two cancers. We searched PubMed, Embase, Scopus, and Web of Science to select relevant articles using keywords related to microorganisms involved in cervical and prostate cancer and the impact of poor healthcare systems on the mortality rates of these two cancers in Africa by carrying out a detailed synopsis of the studies on microbial agents involved and the contributory factors to the deteriorating healthcare system in Africa. It became apparent that the developed countries come first in terms of the prevalence of cervical and prostate cancer. However, more people per capita in Africa die from these cancers as compared to other continents. Also, microbial infections (bacterial or viral), especially sexually transmitted infections, cause inflammation, which triggers the pathogenesis and progression of these cancers among the African population; this has been linked to the region's deficient health infrastructure, making it difficult for people with microbial infections to access healthcare and hence making infection control and prevention challenging. Taken together, untreated microbial infections, primarily sexually transmitted infections due to the deficient healthcare systems in Africa, are responsible for the high mortality rate of cervical and prostate cancer.
... Environmental pollution is a well-known risk factor for human cancer [1][2][3][4]. It has been reported that the 1976 industrial tragedy at Seveso, Italy, gradually drew public attention to the influence of environmental pollutants on human health [5]. ...
Article
Full-text available
Environmental pollutants are closely linked to lung cancer. The different types of environmental pollutants can be classified as chemical, physical, and biological. The roles of common chemical and physical pollutants such as PM2.5, smoking, radon, asbestos, and formaldehyde in lung cancer have been extensively studied. Notably, the worldwide COVID-19 pandemic raised awareness of the strong link between biological pollution and human health. Allergens such as house dust mites and pollen, as well as bacteria and viruses, are common biological pollutants. A few biological pollutants have been reported to promote lung cancer via inducing inflammatory cytokines secretion, such as IL-1β, IL-6, and TGF-β, as well as suppressing immunosurveillance by upregulating regulatory T (Treg) cells while dampening the function of CD8+ T cells and dendritic cells. However, the correlation between common biological hazards, such as SARS-CoV-2, human immunodeficiency viruses, Helicobacter pylori, and house dust mites, and lung cancer is not fully elucidated, and the underlying mechanisms are still unclear. Moreover, the majority of studies that have been performed in lung cancer and biological carcinogens were not based on the perspective of biological pollutants, which has challenged the systematicity and coherence in the field of biological pollutants in lung cancer. Here, in addition to reviewing the recent progress made in investigating the roles of allergens, viruses, and bacteria in lung cancer, we summarized the potential mechanisms underlying biological pollutants in lung cancer. Our narrative review can shed light on understanding the significance of biological pollutants in lung cancer, as well as inspire and broaden research ideas on lung cancer etiology.
... Ashauer et al. (2017), for instance, observed the possibility of carry-over toxicity when the recovery time needed by the exposed organisms was longer than the time between pesticide applications. Goodson et al. (2015) stated that sequential exposure to pesticides could result in deleterious synergic effects. However, most pesticide risk assessments, including pre-market approval assessments, examine the toxicity of only single compounds on single, standard species based on standard endpoints under controlled conditions (Frische et al., 2018;Streissl et al., 2018). ...
Article
Anthropogenic particles (AP), which include microplastics and other synthetic, semisynthetic, and anthropogenically modified materials, are pollutants of concern in aquatic ecosystems worldwide. Rivers are important conduits and retention sites for AP, and time series data on the movement of these particles in lotic ecosystems are needed to assess the role of rivers in the global AP cycle. Much research assessing AP pollution extrapolates stream loads based on single time point measurements, but lotic ecosystems are highly variable over time (e.g., seasonality and storm events). The accuracy of models describing AP dynamics in rivers is constrained by the limited studies that examine how frequent changes in discharge drive particle retention and transport. This study addressed this knowledge gap by using automated, high‐resolution sampling to track AP concentrations and fluxes during multiple storm events in an urban river (Milwaukee River) and comparing these measurements to commonly monitored water quality metrics. AP concentrations and fluxes varied significantly across four storm events, highlighting the temporal variability of AP dynamics. When data from the sampling periods were pooled, there were increases in particle concentration and flux during the early phases of the storms, suggesting that floods may flush AP into the river and/or resuspend particles from the benthic zone. AP flux was closely linked to river discharge, suggesting large loads of AP are delivered downstream during storms. Unexpectedly, AP concentrations were not correlated with other simultaneously measured water quality metrics, including total suspended solids, fecal coliforms, chloride, nitrate, and sulfate, indicating that these metrics cannot be used to estimate AP. These data will contribute to more accurate models of particle dynamics in rivers and global plastic export to oceans. Practitioner Points Anthropogenic particle (AP) concentrations and fluxes in an urban river varied across four storm events. AP concentrations and fluxes were the highest during the early phases of the storms. Storms increased AP transport downstream compared with baseflow. AP concentrations did not correlate with other water quality metrics during storms.
Article
Full-text available
The objective of this research was to evaluate the cumulative exposure of the population aged 10–24 years to pesticides that have a chronic effect on the thyroid. A consumption study covering fresh fruits and vegetables was collected from 377 respondents. In parallel, 2369 fruit and vegetable samples were chemically analyzed for pesticide residues. As a result, cumulative exposure was calculated for four different scenarios (as is, maximum residue level, 70% of MRL and below detection limit) using two recall methods. The results show that, depending on the scenario, cyprodinil from green lettuce and strawberries, fluopyram blueberries and strawberries, and fluxapyroxad, detected in grapes, contribute most to exposure. More stringent scenarios, with limits at 70% of the MRLs (0.7 MRL) and below the detection limit (“zero residue” approach), show that the estimated total margin of exposure increases by up to 50% in the “0.7 MRL” model, while levels almost triple in the “zero residue” model. The optimization of pesticide use has a beneficial effect on human health and the environment, contributing to the “One Health” approach.
Article
The aim of this study was to investigate whether the application of four bioremediation strategies: mycoaugmentation with Actinomucor elegans LBM 239; biostimulation with carbon, nitrogen, and phosphorus sources; combined mycoaugmentation and biostimulation; and natural attenuation could improve the quality of carbendazim-contaminated soil. The experiments were done in microcosm soil and remediation effectiveness was assessed based on pollutant content, soil characteristics, and ecotoxicological tests. A greenhouse experiment was conducted to evaluate the effect of A. elegans LBM 239 on plant growth promotion. Results revealed that mycoaugmentation, biostimulation, and combined strategy (mycoaugmentation + biostimulation) significantly decreased the concentration of carbendazim, with removal percentages higher than 90%, with a significant toxicity reduction at 15 days. Mycoaugmentation and biostimulation combined strategy improved the content of carbon, nitrogen, and phosphorus in the soil. A. elegans did not demonstrate plant growth-promoting properties and no signs and symptoms of the disease were observed on inoculated tomato seedlings. Finally, the principal component analysis revealed that the combined treatment of mycoaugmentation and biostimulation constitute the most efficient bioremediation alternative to restore carbendazim-contaminated soils.
Article
Breast cancer (BC) is one of the urgent problems of health care, which is due to a constant trend of growth. One of the risk factors for the development of breast cancer is ionizing radiation (IR). Numerous epidemiological and experimental studies have shown the high sensitivity of the mammary gland (MG) to this factor. Consideration of models of absolute and relative risks of the occurrence of radio-induced tumors of the MG in irradiated persons showed the importance of such factors as age at the time of irradiation, multiplicity. frequency of exposure, dose level and concomitant non-neoplastic diseases of the mammary and thyroid gland (TG). Excess radiation-induced cases of cervical cancer were found among irradiated women after the atomic bombings of Hiroshima and Nagasaki. Epidemiological features of the development of breast cancer under the influence of IV are presented in detail, which is one of the environmental factors involved in the formation of the modern carcinogenic situation. In connection with the significant sensitivity of the MG to the carcinogenic effect of IR, this form of neoplasms attracted special attention after the Chornobyl accident. The effect of small doses of radiation after the Chornobyl disaster led to a wave-like change in the incidence of breast cancer in certain periods of the year, and the radiation-induced incidence of this pathology can occur spontaneously. Key words: breast cancer, radiation risk, ionizing radiation, territory contaminated with radionuclides, radiation effect, carcinogenic risk, radiosensitivity.
Article
Full-text available
Several risk factors for the development of laryngeal cancer have been identified, such as smoking and alcohol consumption, but the molecular mechanisms related to the carcinogenesis in the larynx remain under investigation. In this context, deregulations of the cell-cycle-controling mechanisms, Rb-pathway in particular, have been suggested to be involved in the pathogenesis of laryngeal carcinoma. Our purpose was to investigate 13q14 LOH and the expression of Rb protein and their possible prognostic value in laryngeal cancer. The group of 67 patients with laryngeal cancer, surgically treated with minimum 5 years observation, was multi-variously analysed. LOH for Rb was investigated by PCR-based techniques using two microsatellite markers, D13S263 and D13S126, flanking the Rb locus. Amplification products from each polymorphism were fractionated by denaturing gel electrophoresis and detected by audioradiography. Immunohistochemical staining of paraffin specimens of laryngeal cancers was supervised by the use of monoclonal mouse antibodies IgG1 (Anti-Human Retinoblastoma Gene Product of Dako) in dilution of 1:50. Inactivation of Rb protein was assumed to represent the expression in < or =10% tumour cells. The results of each examined individual factor were compared with clinicopathologic features and the results were statistically transformed (Chi-square test with Yates' correction, Mann-Whitney test). The Kaplan and Meier model was used for overall and disease free survival curves. Only p value of less than 0,05 was considered significant. 13q14 LOH was detected in 7/67 (10,4%) of informative tumours. No correlations were found between Rb genetic alteration (LOH) and gender, age, TNM staging, histological differentiation, nodal and local recurrences (p>0.05). There was a strong association between the loss of Rb and supraglottic localisation of tumour in the larynx (p<0.01). By univariate analysis 13q14 LOH proved to be significantly related to the overall survival whereas it was not related to the quicker relapse (p=0.01, p>0.05 respectively). The genetic data were correlated with the expression of the Rb protein (p=0.001). All tumours with Rb-LOH were immunohistochemically Rb-negative. Inactivation of Rb protein was observed in 9/67 cases (13.49%) and was significantly correlated with the polymorphism of cancer cells, but not with the histological grading. We also found the correlation between reduction of Rb protein and the size of primary tumour (T) (p=0.03) and local recurrence (p=0.035). There was no significant dependence between the level of Rb protein and other histopathological and clinical features (p>0.05). To conclude, analysis of 13q14 LOH enables the assessment of biology of laryngeal cancer and it can be a prognostic factor in overall survival. Immunohistochemical analysis of Rb protein expression in neoplastic cells made it easier to evaluate the mechanisms of cancerogenesis in laryngeal cancer and is closely related to genetic alteration in Rb locus.
Article
The hormonal activity of natural estrogens is influenced by the degree to which they bind to serum proteins. In the pregnant female and in the fetus, greater than 99% of estradiol may be bound by serum binding proteins. Therefore, even though total serum levels of estradiol appear very high in fetuses, we have found that in rodent fetuses, there is a very low free concentration of estradiol (0.2 pg/ml). Naturally occurring variation in fetal serum estradiol predicts differences in numerous postnatal traits, including prostate size. In addition, when this low level of free estradiol was experimentally increased from 0.2 to 0.3 pg/ml during the last third of fetal life, treated male mice showed an increase in adult prostate weight. Fetal exposure to low doses of xenobiotic estrogens by feeding to pregnant females, including the compounds methoxychlor (20 and 2000 μg/kg body weight), DES (0.02 to 2 μg/kg body weight) and bisphenol A (2 and 20 μg/kg body weight), also led to increased prostate weight in adulthood. In contrast, fetal doses of natural estradiol and DES above the physiological range of estrogenic activity, and within a toxicological dose range, led to the opposite outcome, a reduction in subsequent adult prostate weight. This indicates that it may be impossible to assess endocrine-disrupting activities in response to low doses within a physiological range of activity by using high, toxic doses of xenoestrogens in testing procedures. We have developed approaches in vitro to predict the potential estrogenic bioactivity of compounds in the physiologically relevant range in animals and humans. We address the following factors in predicting the final observed endocrine-disrupting effect in the animal: (1) the intrinsic estrogenic activity of a given molecule, (2) the effective free concentration determined by how the molecule is carried in serum, (3) partitioning between aqueous and lipid compartments in body and cell lipids, and (4) absorption and metabolism relative to the route of exposure. The studies and strategies we describe are important in developing criteria for a tiered testing system for the detection of estrogenic chemicals as well as endocrine-disrupting chemicals with different modes of action.
Article
The European Food Safety Authority (EFSA) asked the Panel on Plant Protection Products and their Residues (PPR Panel) to develop a Scientific Opinion on the relevance of dissimilar mode of action and its appropriate application for cumulative risk assessment of pesticides residues in food. The present opinion was preceded by three previous opinions of the PPR Panel (EFSA, 2008, 2009, 2013). The purpose of the present opinion was to assess the relevance of dissimilar modes of action (MoA) for cumulative risk assessment, to evaluate the existing methods for assessing chemicals acting by dissimilar MoA and to identify which methods need to be considered. The PPR Panel restricted its considerations of pesticide combinations with dissimilar modes of action to substances that produce a common adverse effect on the same organ/system. The PPR Panel noted that there is no case documented in the scientific literature where independent action provided more conservative predictions of combination effects than dose addition, and where at the same time independent action also produced accurate predictions. The use of independent action as an assessment concept for combination effects requires demonstration that modes of action of individual substances in a mixture are strictly independent, a condition that can rarely be met in practice. The PPR Panel also noted that there is no cumulative risk assessment method derived from independent action. The PPR Panel therefore recommends using cumulative risk assessment methods derived from dose addition also for the assessment of mixtures of pesticides with dissimilar modes of action, provided they produce a common adverse outcome. Pesticides that produce common adverse outcomes on the same target organ/system should be grouped together in CAGs, and their combined effects assessed by using the concept of dose addition as a pragmatic and conservative default approach for the purpose of assessing cumulative risk in relation to MRL setting or risk assessment of chemical mixtures in practice.