Download Review of Literature-Based Models for Skin and Eye Irritation and

Document related concepts

Dry eye syndrome wikipedia , lookup

Human eye wikipedia , lookup

Transcript
Review of Literature-Based Models for Skin and
Eye Irritation and Corrosion
Ana Gallegos Saliner, Grace Patlewicz & Andrew P. Worth
2006
EUR 22320 EN
EUROPEAN COMMISSION
DIRECTORATE GENERAL
JOINT RESEARCH CENTRE
Institute for Health and Consumer Protection
Toxicology and Chemical Substances Unit
European Chemicals Bureau
I-21020 Ispra (VA) Italy
Review of Literature-Based Models for Skin and
Eye Irritation and Corrosion
Ana Gallegos Saliner, Grace Patlewicz & Andrew P. Worth
2006
EUR 22320 EN
The mission of the IHCP is to provide scientific support to the development and
implementation of EU policies related to health and consumer protection.
European Commission
Directorate – General Joint Research Centre
Institute for Health and Consumer Protection
Contact information
Address: E. Fermi, 1, 21020-Ispra (VA) Italy
E-mail: [email protected]
Tel.: +39 0 332 78 9566
Fax: +39 0 332 78 6717
http:// http://ecb.jrc.it/QSAR/
http:// ihcp.jrc.cec.eu.int
http://www.jrc.cec.eu.int
LEGAL NOTICE
Neither the European Commission nor any person
acting on behalf of the Commission is responsible for
the use which might be made of this publication.
A great deal of additional information on the
European Union is available on the Internet.
It can be accessed through the Europa server
(http://europa.eu.int)
EUR 22320 EN
© European Communities, 2006
Reproduction is authorised provided the source is acknowledged
Printed in Italy
ABSTRACT
This report reviews the state-of-the-art of in silico and in vitro methods for assessing
dermal and ocular irritation and corrosion. Following a general introduction, the
current EU legislation for the classification and labelling of chemicals causing
irritation and corrosivity is summarised. Then currently available non-animal
approaches are reviewed. The main alternative approaches to assess acute local toxic
effects are: a) in silico approaches, including SARs, QSARs and expert systems
integrating multiple approaches; and b) in vitro test methods. In this review, emphasis
is placed on literature-based (Q)SAR models for skin and eye irritation and corrosion
as well as computer-based expert systems.
TABLE OF CONTENTS
1.
Introduction............................................................................................................1
2.
Regulatory context in the European Union............................................................3
3.
4.
5.
2.1.
REACH ..........................................................................................................3
2.2.
Globally Harmonised System ........................................................................4
2.3.
Cosmetics Directive .......................................................................................5
2.4.
Endpoint – Regulatory definitions .................................................................5
2.4.1.
Skin Irritation .........................................................................................7
2.4.2.
Skin Corrosion .......................................................................................9
2.4.3.
Eye Irritation / Corrosion .....................................................................10
(Q)SAR Models ...................................................................................................14
3.1.1.
Skin Irritation .......................................................................................16
3.1.2.
Skin Corrosion .....................................................................................22
3.1.3.
Eye Irritation / Corrosion .....................................................................29
Alternative test methods ......................................................................................47
4.1.
Use of existing data......................................................................................47
4.2.
In vitro methods ...........................................................................................47
4.2.1.
Skin Irritation .......................................................................................48
4.2.2.
Skin Corrosion .....................................................................................49
4.2.3.
Eye Irritation / Corrosion .....................................................................49
Expert Systems.....................................................................................................51
5.1.
BfR Decision Support System .....................................................................54
5.2.
DEREK ........................................................................................................59
5.3.
HazardExpert ...............................................................................................61
5.4.
MultiCASE ..................................................................................................62
5.5.
TOPKAT......................................................................................................63
5.6.
Other Expert Systems ..................................................................................65
5.6.1.
Cache....................................................................................................65
5.6.2.
SuCCSES .............................................................................................65
7
5.6.3.
5.7.
OASIS ..................................................................................................66
Integrated Testing Exercise..........................................................................66
6.
Conclusions..........................................................................................................68
7.
References............................................................................................................69
8
LIST OF ABBREVIATIONS
3D-QSAR
Three-Dimensional QSAR
3 Rs
Reduction, Refinement, and Replacement (of animal experiments)
AD
Average Deviation
AAD
Absolute Average Deviation
ALS
Adaptive Least Squares
BAIA
Baseline Activity Identification Algorithm
BCOP
Bovine Corneal Opacity and Permeability
BECAM
Bovine Eye Chorio-Allantoic Membrane (in vitro assay)
BfR
(German) Federal Institute for Risk Assessment (former BgVV)
BgVV
(German) Federal Institute for Health Protection of Consumers and
Veterinary Medicine (now BfR)
BLR
Binary Logistic Regression
BNN
Bayesian Neural Network
C / NC
Corrosive / Non-Corrosive
CA
Competent Authority (in EU Member State)
CAChe
Computer-Aided Chemistry
CEC
Commission of the European Communities
CEET
Chicken Enucleated Eye Test
CHIP
Chemicals Hazard Information and Packaging
CI
Confidence Interval
C&L
Classification and Labelling
CM
Classification Model
CMC
Critical Micelle Concentration
COLIPA
European Cosmetic, Toiletry and Perfumery Association
CT
Classification Tree
CSA
Cluster Significance Analysis
DA
Discriminant Analysis
DEREK
Deductive Estimation of Risk from Existing Knowledge
9
DES
Draize Eye Score
DM
Dipole Moment
DRD
Detailed Review Document
DSS
Decision Support System
ECB
European Chemicals Bureau
ECETOC
European Centre for Ecotoxicology and Toxicology of Chemicals
ECM
Embedded Cluster Modelling
ECVAM
European Centre for the Validation of Alternative Methods
EEC
European Economic Community
EHS
Environmental Health and Safety
EIT
Eye Irritation Thresholds
ELINCS
European List of Notified Chemical Substances
EPA
Environmental Protection Agency
EU
European Union
E-value
Electrotopological state value
EWG
Endpoint Working Group (within REACH Implementation Project 3.3)
F
Fischer statistic
FDA
Food and Drug Administration
GFA
Genetic Function Approximation
GHS
Globally Harmonised System
HCE
Human Corneal Epithelium in vitro reconstituted model
HCL
Harmonisation of Classification and Labelling
HET-CAM
Hen's Egg Test on the Chorio-Allantoic Membrane
HOMO
Highest Occupied Molecular Orbital
HSG
Human Serum γ-Globulin
I / NI
Irritant / Non-Irritant
IPCS
International Programme on Chemical Safety
IRAG
Interagency Regulatory Alternatives Group
ITC
Interagency Testing Committee
10
IVSCT
In vitro Skin Corrosivity Test
LDA
Linear Discriminant Analysis
LHASA
Logic and Heuristics Applied to Synthetic Analysis
LMC
Laboratory of Mathematical Chemistry
LogP
Logarithm of the octanol/water partition coefficient
LOO
Leave-One Out
LOO-CV
Leave-One-Out Cross-Validation
LUMO
Lowest Unoccupied Molecular Orbital
LVET
Low Volume Eye Test
MAS
Maximum Average Score (eye irritation measure)
MCI
Molecular Connectivity Index
MDES
Molar Draize Eye Score
MES
Molar-adjusted Eye Score
MI-QSAR
Membrane-Interaction QSAR
MLR
Multidimensional Linear Regression
MMAS
Modified Maximum Average Score (eye irritation measure)
MP
Melting Point
MultiCASE
Multiple Computer Automated Structure Evaluation
MV
Mean Value
MV
Molecular Volume
N
Absolute Hardness
NN
Neural Network
NOAEL
No Observed Adverse Effect Level
OECD
Organisation for Economic Co-operation and Development
OPS
Optimum Prediction Space
PII
Primary (skin/eye) irritation scores
PBBK
Physiologically Based Bio-Kinetic (model)
PC
Principal Component
PCA
Principal Components Analysis
11
PLS
Partial Least Squares
PM
Prediction Model (for converting in vitro to in vivo data)
pKa
-log10 of the acid dissociation constant
(Q)SAR
(Quantitative) Structure Activity Relationship
QSTR
Quantitative Structure-Toxicity Relationship
REACH
Registration, Evaluation and Authorisation of Chemicals
RIP
REACH – Implementation Project
RIVM
Dutch Institute for Public Health and the Environment
SAR
Structure-Activity Relationship
SCCNFP
European Commission’s Scientific Committee on Cosmetics and Non
Food Products
SD
Standard Deviation
SICRET
Skin Irritation Corrosion Rules Estimation Tool
SuCCSES
Substructure-Based Computerized Chemical Selection Expert System
TER
(rat skin) Transcutaneous Electrical Resistance test
TGD
Technical Guidance Document
TOPKAT
TOxicity Prediction by C(K)omputer Assisted Technology
VOC
Volatile Organic Compound
WDS
Weighted Draize Score
X
Electronegativity
12
1. Introduction
This report reviews the state-of-the-art of in silico and in vitro methods for assessing
dermal and ocular irritation and corrosion. Following a general introduction (Chapter
1), the current EU legislation for classification and labelling of chemicals causing
irritation and corrosivity is summarised (Chapter 2). Then currently available nonanimal approaches are reviewed. The main alternative approaches to assess acute
local toxic effects are in silico approaches, particularly SARs, and QSARs (Chapter
3); in vitro test methods (Chapter 4). In this report, emphasis is placed on literaturebased (Q)SAR models for skin and eye irritation and corrosion (Chapter 3), and on
expert systems integrating multiple approaches into computer-based tools (Chapter 5).
Acute local irritancy and corrosivity are mainly assessed in two contexts: in the
hazard classification of industrial chemicals and in the safety assessment of
ingredients and mixtures used in industrial, pharmaceutical and consumer products.
The intended purpose in each context is different, and therefore the considerations
made in the two situations are not equivalent.
In the hazard identification of chemicals, the purpose of testing is to assess the
irritation/corrosion potential according to classification schemes defined by regulatory
authorities. Current regulatory proposals recommend a tiered (stepwise) approach to
hazard identification in which chemicals can be classified as irritating/corrosive on the
basis of results from non-animal methods. Since testing in animals is only required as
a last step to confirm negative results generated by non-animal tests, these testing
strategies contribute to the reduction and refinement of in vivo tests. The non-animal
methods act as partial replacements of the animal test being used to place chemicals
into two or more categories of irritation/corrosion potential, without generating too
many false positive results. In such strategies, there is less concern about the
generation of false negatives because these can be identified by the animal tests
carried out in the last step of the process.
In contrast, in the safety assessment of ophthalmologic and cosmetic ingredients,
mixtures and products, the purpose is to demonstrate that the products will not cause
adverse effects. In this case, the placement of test substances into broad
irritation/corrosion categories is often not sufficient, since it is necessary to establish
the absence of adverse effects at lower concentrations. Although there is an increasing
1
reliance on non-animal methods, it is more challenging to assess the reliability of
predictions for product safety.
Since the skin is often exposed to cosmetic products, the potential for a particular
product/ingredient to cause skin irritation or corrosion needs to be evaluated as part of
the overall safety assessment process. The standard test for assessing skin irritation
and corrosion potential has been for many years the Draize rabbit skin test [1].
The eye can be exposed to cosmetic products and their ingredients either through use
of products intended to be used around the eyes (e.g. mascaras, eye creams) or
through accidental exposure to products that may enter the eye in diluted form during
normal use (e.g. shampoos). The evaluation of eye irritation potential for a cosmetic
product and its ingredients is essential to ensure that a product is safe for consumers.
The conventional test for assessing eye irritation and corrosion potential of chemicals
is the Draize rabbit eye test developed by Draize in 1944 [1].
Although corrosivity should not be caused by cosmetics, occasionally it can occur
after a manufacturing mistake or misuse. In the formulation process, a cosmetic
ingredient that has the intrinsic property to be irritant or corrosive is not necessarily
excluded for use in cosmetics. Irritation and corrosivity are dependent on the final
concentration in the cosmetic product, on the presence of neutralising substances, the
excipients used, the exposure route, and on the conditions of use, among other factors.
2
2. Regulatory context in the European Union
2.1.
REACH
Currently under European legislation, Directive 67/548/EEC [2] on the classification,
packaging and labelling of dangerous substances requires new substances to be tested
and assessed for possible risks to human health and the environment before they are
marketed in volumes starting at 10 kg. Skin and eye irritation and corrosion (acute
dermal toxicity) testing is required for substances produced or imported at levels
above 100 kg per year.
In contrast, existing substances are assessed under the provisions of Regulation
793/93/EEC [3]-[4]. This Regulation requires the identification of priority substances,
which are then subjected to comprehensive risk assessment led by EU member state
regulatory authorities. A consequence of this distinction is that testing of low priority
existing substances has not been required, which means that data on the health and
environmental effects of most existing substances is lacking.
In October 2003, to address this knowledge gap, the European Commission published
a draft proposal for a new chemicals legislation called REACH, which stands for the
Registration, Evaluation and Authorisation of Chemicals in the European Union [5][12]. The proposed REACH legislation aims to obtain the necessary information for
all substances imported or manufactured at volumes greater than 1 tonne, with a
period of 11 years from the entry into force of the legislation. REACH envisages an
extensive system of assessing the physicochemical, (eco)toxicological and
environmental properties of chemicals based mainly on their tonnage. Annexes VII to
X of REACH describe the information requirements for substances manufactured or
imported into the EU in quantities above one tonne per year, with possibilities for
adaptations (including derogations) where justified. The exact testing and data
requirements depend on the volume, but the base set for substances produced or
imported in quantities between 1 and 10 tonnes per year includes physicochemical
properties, toxicological and ecotoxicological data.
Thus, the risk assessment of New and Existing Chemicals will be driven by the
forthcoming REACH legislation, expected to enter into force in 2007. Several studies
reporting the estimated economical benefits and animal life savings of using
3
intelligent testing strategies based on (Q)SARs, read-across, and other waiving
possibilities have been published by the ECB [13]-[14].
2.2.
Globally Harmonised System
In September 1989, the Commission of the European Communities (CEC) raised the
need to evaluate the feasibility of rationalising the different classification schemes
based on acute toxicity of chemicals existing in different countries, trading blocks,
and international organisations [15]-[22]. The schemes for classifying chemicals as
dangerous based on toxicological properties, and the national authorities or
international organisations responsible for the schemes were identified.
In November 1994, the OECD decided to create the Programme on Harmonisation of
Classification and Labelling (HCL). The objectives of this programme were to
develop a harmonised classification system for chemical substances covering nine
hazard endpoints considered in existing classification systems. The objectives also
included the development of a harmonised hazard classification system for chemical
mixtures, based on the same endpoints as considered for substances.
As the work progressed, Detailed Review Documents (DRD) were drafted and
subsequently discussed and agreed upon by OECD ad hoc Expert Groups, endorsed
by the OECD Task Force on HCL. They were used as the basis for consensus building
on the proposals for harmonised classification systems. The various DRD are
available as Environmental Health and Safety (EHS) Monographs in the Series on
Testing and Assessment.
In November 1998, the 28th Joint Meeting recognised the achievements made in the
area of classification of substances and approved the concept of an Integrated
Document, which brings together the harmonised classification systems for each of
the nine recognised hazard categories, including the substances which cause skin
irritation/corrosion, and eye irritation/corrosion. The two endpoints are combined into
a single classification system in the case of mixtures. The Harmonised System for the
Classification of Chemicals which cause Skin Irritation/Corrosion and Eye
Irritation/Corrosion are collected in Chapters 2.2 and 2.3, respectively, while the
Harmonised System for the Classification of Chemical Mixtures which cause Skin
and Eye Corrosion/ Irritation is collected in Chapter 3.3 [20].
4
2.3.
Cosmetics Directive
In the EU, the Cosmetics Directive 76/768/EEC [23] imposes the legal obligation that
cosmetics sold to consumers should be safe. Thus, the responsibility for product
safety lies with the cosmetic manufacturer or the legal person placing a cosmetic
product on the Community market, who must be able to demonstrate that the product
is safe for the consumer.
The Cosmetics Directive does not specify a fixed dataset or methods needed to assess
the safety of cosmetic ingredients [24]; however, the safety assessment of finished
cosmetic products without animal testing is only possible provided that an adequate
toxicological data package on the ingredients is available, according to guidelines
developed by the Commission’s Scientific Committee on Cosmetics and Non Food
Products (SCCNFP) [25]. Toxicological data requirements for cosmetic ingredients
based on the SCCNFP guidelines include eleven endpoints: acute systemic toxicity;
skin irritation/corrosion; eye irritation; skin sensitisation; skin absorption; subacute
and subchronic toxicity; genotoxicity/mutagenicity; UV-induced toxic effects
(phototoxicity, photoallergy, photogenotoxicity); toxicokinetics and metabolism;
carcinogenicity; and reproductive and developmental toxicity.
The safety assessment of a cosmetic product requires taking into account the general
profile of the ingredients, their hazard (which is the intrinsic potential of a substance
to cause damage to human health) and the risk (which is the probability of a substance
to cause damage under relevant use conditions). For all cosmetic ingredients,
appropriate data must be provided to permit an adequate safety assessment, regardless
of tonnage.
2.4.
Endpoint – Regulatory definitions
The classification and labelling of commercial substances within the EU is based on
the use of Risk phases (R-phrases). R-phrase definitions are assigned after the
toxicological evaluation of a wide range of human health hazard effects, including
acute toxicity, repeat dose toxicity, carcinogenicity, mutagenicity, reproductive
toxicity, teratogenicity, irritancy and sensitisation. Thus, R-phrases are used as
indicators of hazard potential for a large number of new and existing substances.
Guidance on the classification and labelling of chemicals is provided under the
5
Chemicals Hazard Information and Packaging for Supply (CHIP) Regulations 1997
[26]-[27]. R-phrase definitions and the grounds upon which chemicals are assigned Rphrases for skin and eye irritation/corrosion are presented in Table 1.
Table 1. R-phrases for skin and eye irritation/corrosion
Risk
Phrase
Effect
Description
R34
C
Causes burns
Full thickness destruction of the skin occurs as a
result of up to 4 hours exposure.
R35
C
Causes severe burns
Full thickness destruction of the skin occurs as a
result of up to 3 minutes exposure.
R36
Xi
Irritating to eyes
Defined reversible changes of cornea, iris or
conjunctiva
R38
Xi
Irritating to skin
Defined reversible inflammation of the skin as a
result of up to four hours of exposure
R41
Xi
Risk of serious
damage to eyes
Defined severe reversible changes or irreversible
changes of cornea, iris or conjunctiva
The protocols recommended for the different required tests are described in the Annex
VII of the dangerous substances directive (Directive 67/548/EEC [2]) and in the
OECD testing guidelines for chemicals. Test methods in Annex VII for dermal and
ocular irritation/corrosion are summarised in Table 2.
Table 2. Test methods for dermal and ocular toxicity in Annex V to Directive 67/548/EEC.
Method
Title
Directive
B.3
Acute toxicity (dermal) [28]
92/69/EEC
B.4
Acute toxicity (skin irritation) [29]
92/69/EEC
B.5
Acute toxicity (eye irritation) [30]
92/69/EEC
B.40
Skin Corrosion [31]
2000/33/EC
The OECD in cooperation with the Coordinating Group on the Harmonisation of
Chemical Classification Systems has agreed on a harmonised system of classification
to be used internationally [20]. Separate categories have been adopted for skin
irritation and for skin corrosion, while eye irritation and corrosion are comprised in a
single category. The available OECD test guidelines for these endpoints are
summarised in Table 3.
6
Table 3. OECD Guidelines for Testing of Chemicals.
No.
Title
Original
Adoption
Most Recent
Update
404
Acute dermal irritation/corrosion [32]
12 May 1981
24 April 2002
405
Acute eye irritation/corrosion [33]
12 May 1981
24 April 2002
430
In vitro Skin Corrosion: Transcutaneous Electrical
Resistance Test (TER) [34]
13 April 2004
-
431
In vitro Skin Corrosion: Human Skin Model Test [35]
13 April 2004
-
The TG 404 was first adopted in 1981 and was revised in 1992 [36]; likewise, TG 405
was adopted in 1981 and revised in 1987. Detailed Review Documents on
Classification Systems for Eye Irritation/Corrosion and Skin Irritation/Corrosion in
OECD Member Countries are gathered in the OECD Series on Testing and
Assessment documents No.14 [37] and No.16 [38], respectively. At the OECD
workshop in Solna, Sweden, in 1996 [16], it was agreed that both TGs should be
updated to include a sequential testing strategy [39]-[40]. The subsequently revised
draft documents were submitted for review in May 2001 [41]-[43] and adopted in
April 2002. The definitions according to the OECD guidelines and to the GHS are
given below in the next sections.
2.4.1. Skin Irritation
Definition according to the OECD TG. Dermal irritation is defined in OECD TG
404 as “the production of reversible damage of the skin following the application of a
test substance for up to 4 hours” [32]. In the EU this effect is assigned the R-phrases
“Xi” and R38.
Pathological characteristics and manifestation of skin irritation. Erythema/eschar
and edema are manifestations of dermal irritation, but further guidance is lacking.
Irritation is initially manifest by redness (erythema), vesicles, serous exudates, serous
scabs (eschar) and various degrees of swelling (edema). Over time, other reactions
may be manifest, like small areas of alopecia, hyperkeratosis, hyperplasia and scaling.
Histopathology is useful in discerning among responses. In most cases inflammation
is well developed within the first 72 hours of observation.
7
Grading of dermal responses. For the grading of dermal responses, there is a
consensus among all the authorities to use the Draize grading scale for dermal lesions
(Table 4).
Table 4. Grading of dermal responses (Draize scale) [32].
Erythema and eschar formation
Grade
No erythema
0
Very slight erythema (barely perceptible)
1
Well defined erythema
2
Moderate to severe erythema
3
Severe erythema (beet redness) to slight eschar formation (injuries in depth)
4
Edema formation
Grade
No edema
0
Very slight edema (barely perceptible)
1
Slight edema (edges of area well defined by definite raising)
2
Moderate edema (raised approximately 1 mm)
3
Severe edema (raised more than 1 mm and extending beyond the area of exposure)
4
Globally Harmonised System. According to the GHS [20], a single category is
adopted for skin irritation. Erythema/eschar and oedema are graded separately; an
animal’s mean score from readings over the first three days after exposure must meet
a defined level to be positive; and at least 2 of 3 tested animals must be positive for
the test to be positive. Positive responses can also be obtained using other less
common criteria. The proportion of test substances expected to be positive by the
proposed irritant category is within the range of positives among existing
classification systems. Some authorities use two hazard categories, i.e. irritants and
mild irritants.
Criteria for classification. Authorities vary as to the existence of criteria for
evaluating irritation; whether mean scores are computed for one time period or across
all grading times; whether endpoint grades are used alone (edema or erythema/eschar)
or grades are pooled (edema plus erythema/eschar). Authorities also vary as to when
animals are observed and scored for dermal lesions and whether animals are followed
to evaluate the reversibility of lesions. Most authorities grade lesions at or around 24,
8
48 and 72 hours after patch removal; fewer of them grade lesions within 1 hour of
patch removal. None use times beyond 72 hours.
EU Criteria. In the in vivo rabbit Draize test, the chemical is applied to the shaved
skin of 3-6 rabbits and covered for 4 or 24 hours [29]. The chemical is then wiped off
to permit observation for signs of skin irritation, and application sites are examined
after 24, 48, and 72 hours after application for assessment of primary skin irritation
scores, using an empirical formula. For a three-animal test, mean scores across 3
scoring times (24, 48 and 72 hours after patch removal) are calculated for each animal
for edema grades and for erythema/eschar grades, separately. A test is positive when
the mean score is 2 or greater. The test is positive for irritation when at least two
animals are positive for the same endpoint (erythema/eschar or edema). For a sixanimal test, a mean score is calculated across 3 scoring times (as above) and across all
six animals for edema grades and for erythema/eschar grades, separately. A mean
score of 2 or more identifies an irritant. EU criteria also recognise the importance of
reversibility of lesions over time up to 14 days.
2.4.2. Skin Corrosion
Definition according to the OECD TG. Dermal corrosion is defined in OECD TG
404 as “the production of irreversible damage to skin; namely, visible necrosis
through the epidermis and into the dermis, following the application of a test
substance for up to four hours. Corrosive reactions are typified by ulcers, bleeding,
bloody scabs, and, by the end of observation at 14 days, by discolouration due to
blanching of the skin, complete areas of alopecia, and scars” [32]. In the EU this
effect is assigned “C” and either R34 or R35.
Pathological characteristics and manifestation of skin corrosion. In some cases
corrosive manifestations are noted soon after application to the skin, however, it can
take a period of days for the process to reach maximum intensity. Corrosion involves
the destruction of tissue through the epidermis and down into the dermis. When there
is bleeding or bloody scabs (eschar) it is obvious that the epidermal basement
membrane has been breached and the process has entered the dermis. Over time there
is often blanching of the skin due to the loss of pigment cells. Since hair follicles are
embedded in the dermis, areas of alopecia also indicate corrosivity. Histopathology
9
can give a definitive diagnosis. Vesicles are an indication of irritation, not corrosion,
as they represent disturbances at the epidermal-dermal interface.
Global Harmonised System. According to the GHS [20], a single category is
adopted for skin corrosion, although some authorities may subdivide the single
corrosive category into up to three subcategories.
Criteria for classification. The GHS criteria for classification of corrosion are agreed
upon (e.g. at least one of three tested animals must show corrosive effects after
exposure up to a 4 hour duration before an agent is classified as corrosive) [20].
EU Criteria. EU criteria state that if at least one test animal manifests a corrosive
action following exposures of up to 3 minutes or up to 4 hours, then the test material
will be judged to cause severe burns (R35) or to cause burns (R34), respectively.
2.4.3. Eye Irritation / Corrosion
Definition according to the OECD TG. Eye irritation is defined in OECD TG 405 as
“the production of changes in the eye following application of a test substance to the
anterior surface of the eye, which are fully reversible within 21 days of application”
[33]. In the EU this effect is assigned “Xi” and R36.
Eye corrosion is defined in OECD TG 405 as ”the production of tissue damage in the
eye, or serious physical decay of vision, following application of a test substance to
the anterior surface of the eye, which is not fully reversible within 21 days of
application” [33]. In the EU, this effect is assigned “Xi”, and R41.
Grading of ocular responses. Essentially four different approaches to the evaluation
of ocular hazard have been developed, some of which have been used in the
international regulation of chemical products. The classification procedures are
similar since they all are based on the rabbit eye model. The effects of the substances
on the cornea, iris and conjunctivae are graded on individual scales and given
weighted scores, using the Draize scoring system for ranking responses (Table 5).
10
Table 5. Grading of Ocular Responses (Draize scoring system) based on the rabbit eye model
[33].
RESPONSE
Score
CORNEA
Opacity: degree of density (area most dense taken for reading)
No ulceration or opacity
0
Scattered or diffuse areas of opacity (other than slight dulling of normal
lustre), details of iris clearly visible
1
Easily discernible translucent area, details of iris slightly obscured
2
Nacrous area, no details of iris visible, size of pupil barely discernible
3
Opaque cornea, iris not discernible through the opacity
4
IRIS
Normal
0
Markedly deepened rugae, congestion, swelling, moderate circumcorneal
hyperaemia, or injection, any of these or combination of any thereof, iris
still reacting to light (sluggish reaction is positive)
1
No reaction to light, haemorrhage, gross destruction (any or all of these)
2
CONJUNCTIVAE
Redness (refers to palpebral and bulbar conjunctivae, cornea and iris);
blood vessels normal
0
Some blood vessels definitely hyperaemic (injected)
1
Diffuse, crimson colour, individual vessels not easily discernible
2
Diffuse beefy red
3
Chemosis
Swelling (refers to lids and/or nictating membranes)
No swelling
0
Any swelling above normal (includes nictating membranes)
1
Obvious swelling with partial eversion of lids
2
Swelling with lids about half closed
3
Swelling with lids more than half closed
4
The scores are processed differently to decide if a test substance is irritant or not, and
to grade the severity of the irritation. The different features are of varying importance
in each classification scheme [44]: the intensity of the irritant response in different
regions of the anterior eye; the weighting of response scores prior to determining an
irritation index; the level of response during the first three days after exposure; and
the reversibility or irreversibility of a response.
11
Global Harmonised System. A single harmonised hazard group is defined for the
classification of eye irritation that reverses within an appropriate observation time
[20]. Local effects detected in a Draize rabbit test that reverse within 21 days after
instillation of the substance into the eye are classified. Effects on the cornea, effects
on the iris and conjunctival erythema and oedema are graded separately. An animal’s
mean score from readings over the first three days after instillation must meet a
defined level to be positive, and at least 2 of 3 tested animals must be positive for the
test to be positive. Some authorities distinguish between mild and moderate eye
irritants.
Criteria for classification. The criteria for classification are based on mean Draize
scores, and on the duration of effects (Table 6).
Table 6. EU Criteria for the classification of eye irritation and corrosion.
Response
R 36
R 41
Corneal opacity
> 2, < 3
>3
Iris lesion
> 1, < 1.5
> 1.5
Redness of the conjunctivae
> 2.5
Oedema of the conjunctivae
>2
(Chemosis)
The duration of the study should be sufficient to evaluate fully the reversibility or
irreversibility of the effects observed. An extended observation period may be
necessary if there is persistent corneal involvement or other ocular irritation in any
animal in order to determine the progress of the lesions.
Persistent colouration of the eye (this effect demonstrating irreversible penetration of
the chemical into the cells of the eye) and irreversible corneal opacity are also judged
as severe eye damage because these effects diminish crucially the function of the eye
as an optical organ. Animals showing severe and enduring signs of distress and pain
may need to be humanely killed. In that case the substance has to be labelled R41
without taking into account the possible reversibility of effects.
EU Criteria. Generally a Draize eye test is performed by applying the substance to
the eye of a living rabbit [30], using one animal (if severe effects are suspected) or
three animals (if no severe eye irritation is anticipated). Ocular lesions which occur
12
within 72 hours after exposure and which persist for at least 24 hours after instillation
qualify a material to be classified as irritant to the eye, if they also reach certain score
levels.
13
3. (Q)SAR Models
In vivo testing for skin and eye irritation and corrosion is among the most emotive of
all testing procedures. Due to public pressure and legislation, there has been a move to
reduce this testing, and replacements have been sought, mainly by exploring the use
of in vitro methods, but also non-testing methods (e.g. QSAR and read-across). Under
REACH, it is expected that read-across and QSAR data will contribute to the initial
assessment of whether testing is scientifically necessary. The OECD principles for
(Q)SAR validation [45]-[46] identify the types of information that should be
associated with a (Q)SAR model proposed for regulatory use. This section reviews the
most
representative
(Q)SAR
approaches
developed
for
skin
and
eye
irritation/corrosion.
The underlying mechanisms of skin irritation/corrosion and eye irritation/corrosion
are diverse and often not completely understood, and the relevance of the animal tests
as indicators of human effects has been questioned. On the other side, it has been
argued that the current animal tests provide conservative and protective assessments.
For example, there is evidence that animal skin is more permeable to chemicals than
human skin; thus, the use of rabbit Draize test as indicator of potential skin irritation
in humans has been questioned. Draize test have been pointed as being conservative
and protective, i.e. it overestimates the human response. In addition, species-related
shifts in classification threshold have been observed in the corrosive / non-corrosive
threshold when substituting rat skin for human skin.
From the scientific literature, it appears that more emphasis has been placed on the
QSAR modelling of ocular irritation compared with dermal irritation. Dermal and
ocular responses are usually modelled separately, although some similarities have
been established between both endpoints. For instance, the cross-correlation between
primary skin irritation scores (PII) and eye irritation measures (i.e. maximum average
score (MAS), and molar-adjusted eye score (MES)) showed moderate correlations
that suggested both similarities and differences between the skin and eye mechanisms
of action.
Tiered assessment strategies have been applied jointly and separately to both
endpoints. There is strong evidence that chemicals that are corrosive to the skin
14
should also be classified as being corrosive to the eye, especially if the assessment is
made from knowledge of acidity and alkalinity [47]-[48]. In particular, in the EU and
OECD classification schemes, chemicals that have been found to be corrosive to the
skin are automatically considered to be corrosive to the eye as well (and are therefore
labelled as such without animal testing).
For the prediction of skin toxicity, basic physicochemical properties underlie the
irritant response. These properties include the fundamental acidic and basic properties
of the chemical, its ability to penetrate the skin (a function of hydrophobicity and
molecular size), and cytotoxicity. This last property can be described by in vitro
modelling or it can be estimated from physicochemical properties. Thus, skin
irritation/corrosivity is postulated to result from the chemical first penetrating the
skin; then, if the chemical is sufficiently cytotoxic, death of the cells beneath results in
corrosivity. However, irritation and corrosion endpoints have been modelled mostly
separately for the dermal response. This is due to the fact that corrosion is considered
to be more a physical effect (which can be captured by a simple description of acidity
or basicity) than a biologically mediated effect. The description of acidity/basicity can
be derived from a measurement of pH of certain strength of solution or acid/alkali
reserve, or from the presence of certain structural features known to be associated
with acids and bases.
The mechanistic pathway is likely to vary with chemical class, and it has generally
been assumed that multiple class-specific (Q)SAR models are required to predict the
toxic potential of a wide array of chemicals. Thus, (Q)SAR models have generally
been developed for small datasets of specific groups of compounds, although in some
cases more diverse and larger datasets have been also examined. In general, it has
been suggested that basic physicochemical parameters such as acidity, basicity,
hydrophobicity, and molecular size as well as electrophilic reactivity, are useful to
predict the toxic potential of homologous chemicals. Thus, certain parameters may be
used to define cut-offs (e.g. a range of pH values, or molecular weights) so that
compounds with a physicochemical parameter below or above a certain value can be
classified
as
toxic
(or
non-toxic).
Most
models
employ
intramolecular
physicochemical properties associated with solubility. A few models also consider the
intermolecular physicochemical parameters that describe the interactions between the
compound of interest and a biological receptor. For example, some models are based
15
on the molecular simulation of molecule-membrane interactions (MI-QSAR). In
contrast, models intended to predict the toxic potential of heterogeneous groups of
chemicals have emphasised the commonality of structural features.
The simplest (Q)SAR approaches are traditional quantitative regression models and
classification models (CM). Concerning the statistical technique used to derive the
model, linear modelling techniques (linear regression, and linear discriminant analysis
for CM) have generally been employed, although a number of non-linear approaches
(neural networks, and cluster analysis for CM) have also been explored. Non-linear
modelling techniques are thought to be generally useful for the modelling of
heterogeneous datasets. A number of selected reviews have been referenced [49]-[60].
Some of the most comprehensive reviews on acute local effects are the PhD thesis of
Worth [49]; the series of reviews by Schultz et al. [53]-[54], papers by Cronin et al.
[55] and by Patlewicz et al. [56]; and the review of expert systems by Cronin et al.
[57].
3.1.1. Skin Irritation
Following in vivo animal testing (Draize rabbit skin test), chemicals are classified as
being either irritant or corrosive, depending on specific responses to the tests. In the
regulatory assessment of chemicals, relatively few QSAR methods have been applied
to the prediction of skin irritation and corrosion. One of the earliest SAR approaches
to predict skin irritation is based on the TOPKAT methodology [61]. Other local
models have been reported by Berner [62]-[64], Nangia [65], Hayashi [66], Smith
[67]-[68], and Kodithala [69], who reported a MI-QSAR model.
Enslein et al. [61] reported QSARs to predict rabbit skin irritation severity, based on
the Draize assay. The models, based on an extensive dataset of publicly available skin
irritation data, were implemented in the TOPKAT software package (Accelrys, Inc.).
After harmonising the experimental numerical scores into a consistent severity rating
system, the dataset was split into compounds without, and with rings, prior to develop
the models. It was noted that compounds without rings (260 compounds) were located
at the high range of the severity scale, while the compounds with rings (526
compounds) were concentrated near the low range of the scale. Thus, two sets of
independent models were developed for the different categories. For each subset of
16
compounds, two QSAR equations were developed; a first equation to select severe
irritant compounds, and a second one to discriminate mild and moderate irritants from
non-irritant compounds. The models could not distinguish between mild and moderate
irritation. The parameters used for the development of the models were molecular
connectivity indices (MCI), MOLSTAC substructural keys (either individual keys
describing single chemical features, or combination keys), and molecular length
parameters. The QSAR models were developed by stepwise discriminant analysis
(DA). The variable selection process was applied separately first to dichotomous
parameters and after to continuous parameters. Robustness of the DA was checked,
and all the possible regression subsets were calculated to determine the subset of
variables that best fit regression criteria. Validation was carried out employing the
resubstitution method, in which the discriminant function is used to evaluate the
compounds used to develop the equation. The classification accuracy for the QSAR
models ranged from 90% to 95% (Table 7). Indeterminate compounds were defined in
the region around chance probability (0.5), set between p=0.3 and p=0.7. These
QSARs form the basis of the skin irritation module in TOPKAT.
Table 7. Classification results of the discriminant QSAR models for skin irritation.
Models
Non-ring
compounds
Ring
compounds
Subequations
Overall
Accuracy
False
Positives
False
Negatives
Indeterminate
F
(p<<0.001)
Severes vs
Others
94.8%
3.1%
(3.8%)*
2.2%
(11.1%)*
11.9%
15.56
Negative svs
Mild/Moderates
95.0%
1.5%
(13.6%)*
3.5%
(3.9%)*
2.4%
8.33
Severes vs
Others
91.0%
5.6%
(6.6%)*
3.4%
(22.3%)*
15.8%
13.51
Severes vs
Others
89.8%
4.3%
(15.8%)*
5.9%
(8.1%)*
15.7%
11.93
* Percentage of false predictions in relation to the experimental class
Berner et al. published a series of papers [62]-[64] in which they reported that the
severity of irritation from reduced sets of organic acids and bases in humans
correlated with the acid or basic strength as measured by pKa. For example, one study
reports the relationship between pKa and skin irritation in humans for a congeneric
series of benzoic acid derivatives [64]. It was found that skin irritation and pKa were
17
strongly correlated; for pKa higher than 4 skin irritation rapidly increased. Related to
the same study, Nangia et al. [65] published a one-variable model based on 12 organic
bases relating skin irritation with pKa. According to the interpretation of the results,
skin irritation would be triggered by local alterations in pH caused by acidic and basic
chemicals.
Hayashi et al. [66] reported two local QSAR models for predicting the molarweighted skin irritation scores (PII) of a set of 30 phenols. The dataset was split into a
training set (24 phenols) and a test set (6 phenols) used to validate the models. The
descriptors used to build the models were based on reactivity and permeability. The
absolute hardness (N) and electronegativity (X) calculated from the highest occupied
molecular orbital (HOMO), and the lowest unoccupied molecular orbital (LUMO)
energy levels, accounted for reactivity; the logarithm of the octanol/water partition
coefficient (LogP) accounted for permeability. Two different models were obtained
according to the sign of the LUMO energy level (positive or negative), i.e. the
electronic characteristics of phenols. It was found that phenols with positive LUMO
react with amino acids in both the epidermis and the dermis as electron donors,
resulting in irritation. Therefore, their irritation potential was based on their skin
permeability, LogP (r=0.82). On the other hand, phenols with negative LUMO only
reacted with amino acids in the epidermis as electron donors, resulting in relatively
low irritation (r=0.72). This model, independent of skin permeability, was correlated
with absolute hardness (N). The goodness-of-fit for the different regression equations
are summarised in Table 8. The differences between predicted and experimental
scores for the six test phenols showed a good correlation (n=30, r=0.83). This study
suggested that skin irritation caused by phenols is related to their reactivity with
macromolecules and their skin permeability; differences in reactivity among phenols
were dependent on the electron-releasing or electron-attracting characteristics of their
substituents.
18
Table 8. Goodness-of-fit of the linear regression QSAR models for skin irritant phenols.
Model
Positive LUMO
Negative LUMO
N
Equation
r
F
s
13
PII = −178 N + 1130
–0.36
1.27
18.00
13
PII = 230 log P + 25.5
0.82
23.12
12.82
11
PII = −474 N + 2200
0.72
3.86
7.85
11
PII = 22.9 log P + 63.4
0.16
0.25
9.21
Smith et al. [67] reported a local QSAR model for 42 esters (13 irritants and 29 nonirritants), for which human skin irritation data were available. The model was based
on 19 calculated physicochemical parameters, representing transport, electronic and
steric properties. The robustness (stability) of the model was assessed by using
multiple random resampling. The 13 irritant esters were paired with ten random sets
(with replacement) of 13 non-irritants esters, and statistical analyses were conducted
with each test. Best subsets regression was used to select variables for subsequent
inclusion in discriminant models, and to identify candidate models for further
analysis. Linear discriminant analysis (LDA) was used to distinguish between irritant
and non-irritant esters. The sensitivity of the resulting ten models evaluated by leaveone-out (LOO) cross-validation ranged from 0.846 to 0.923; the specificity ranged
from 0.615 to 0.923. The averaged performance results for the ten models is
summarised in Table 9. Compared with non-irritant esters, the irritant ones had lower
density, lower water solubility, lower sum of partial positive charges, higher Hansen
hydrogen bonding parameter, and higher Hansen dispersion parameters. The results
indicate that transport and electronic parameters related to intermolecular interactions
are relevant parameters for modelling skin irritation in humans.
Table 9. Averaged performance of the discriminant SAR models for human irritant esters.
Squared Distance*
Sensitivity
Specificity
Mean
4.9
0.885
0.738
Upper confidence interval
6.9
0.910
0.799
Lower confidence interval
3.0
0.859
0.678
* Degree of separation between the active and inactive group means
19
In a subsequent study, Smith et al. conducted an experimental validation study [68] to
assess the ability of the previously developed QSAR model [67] to predict human skin
irritation by esters commonly used as fragrance ingredients. Thirty four esters
reported to cause irritation in rabbits, located within the predictive space of the QSAR
model, were selected for human testing by using the patch test procedure. Of the 34
rabbit irritants selected, 16 were predicted by the QSAR model to be positive and 18
were predicted to be negative. The comparison of the experimental rabbit Draize test
with human patch testing revealed that only two of the 34 esters were human irritants.
Training set compounds were then queried by using the previously developed ten
QSAR submodels, following a consensus approach [67]. An ester was predicted to be
an irritant if at least five of the submodels predicted it to be positive. The QSAR
model was refined by incorporating the results from the new experimental patch
testing. As before, 19 physicochemical descriptors were calculated; statistical analyses
were performed on 10 subsets of the database, each subset consisting of randomly
selected irritants and non-irritants. Best subsets regression analysis identified
candidate models for subsequent discriminant analysis, and each candidate model was
analysed by using linear regression. Sensitivities and specificities were determined by
the leave-one-out cross-validation (LOO-CV) method. Although all of the 34
investigated esters had been reported to be dermal irritants on rabbit skin, only 16
were predicted to be potentially irritant according to the QSAR models. Then the
dataset of esters was combined with a previously published dataset of 42 esters to
create a more robust dataset. Ten random sets of 10 irritant esters and 50 non-irritant
esters were formed and statistical analyses were conducted on each set. The most
consistent parameters were solubility parameter, water solubility, Hansen dispersion,
Hansen hydrogen bonding, and sum of partial positive charges. The sensitivity and
specificity values for the submodels ranged from 0.50 to 0.80 and from 0.72 to 0.90,
respectively. The averaged results are given in Table 10.
20
Table 10. Averaged predictivity of the discriminant QSAR models for human irritant esters.
Squared Distance*
Sensitivity
Specificity
Mean
2.87
0.66
0.81
95% CI ±
0.60
0.06
0.04
* Degree of separation between the active and inactive group means
CI: Confidence Interval
Kodithala et al. [69] reported a membrane-interaction QSAR (MI-QSAR) study to
predict skin irritation of a set of 20 diverse hydroxy alcohol compounds. This study
combined the use of an in vitro skin irritation model with QSAR techniques. The main
assumption was that skin irritation does not depend on specific interaction receptors
but on the absolute uptake into the phospholipid rich regions of cellular membranes,
which constitute the effective “receptor”. Different phases of this model comprised
modelling of the membrane and molecular dynamics simulations; construction of
solute molecules and docking of the molecules in the membrane; calculation of intra
and intermolecular QSAR descriptors; and construction of QSAR models by using a
genetic function approximation (GFA) as a fitting function in the multidimensional
linear regression (MLR) model. The comparison of different intramolecular QSAR
models (using descriptors such as LogP and HOMO) with MI-QSAR models
suggested that the latter were more robust. After checking the cross-correlation
descriptor matrix for each model, the MI-QSAR model with three descriptors (LogP,
HOMO and electrostatic plus Van der Waals interaction energy (E(vdw+chg)) with
performance of r2=0.76 and r2cv=0.67 was considered the most stable one (see Table
11). The training set was split into two subsets of aliphatic alcohols and phenols;
newly developed three-term MI-QSAR models achieved r2=0.95, and r2=0.94,
respectively (Table 12). In conclusion, skin irritation was predicted to increase with
increasing effective concentration of irritants, with increasing binding of compounds
to the phospholipid-rich regions of the cellular membrane, and with increasing
chemical reactivity, reflected by HOMO and LUMO.
21
Table 11. Goodness-of-fit and robustness of the QSAR and MI-QSAR models for skin irritant
hydroxy alcohols.
QSAR models
MI-QSAR models
N
Equation
r2
r2cv
N
Equation
r2
r2cv
20
PII = 1.84 + 0.90 log P
0.54
0.37
20
PII = 1.96 + 0.93 log P
0.54
0.37
20
PII = 56.3 + 3.63HOMO +
+0.41FOCT
0.61
0.43
20
PII = 34.3 − 0.06 E (vdw + chg )
+2.42 HOMO
0.62
0.48
0.64
0.40
20
PII = 3.51 − 0.26 E (vdw + chg )
−0.82 log P + 1.14 HOMO
0.76
0.67
0.67
0.37
20
0.86
0.72
20
20
PII = 42.4 + 2.57 HOMO + 0.16MR
+0.93FOCT
PII = −19.84 − 0.57Vm + 0.61FOCT
+3.71LUMO + 2.36MR
PII = 68.07 − 0.06MW + 0.66ETO
−0.18E (vdw + chg ) + 4.72 HOMO
FOCT ,Vm, E (vdw + chg ), ETO : intermolecular solute-membrane interaction descriptors
Table 12. Goodness-of-fit and robustness of the MI-QSAR models for skin irritant aliphatic
alcohols and phenols.
Submodel
N
MI-QSAR Equation
r2
r2cv
LSE
Aliphatic
Alcohols
13
PII = 49.60 − 0.09E (vdw + chg ) + 3.46HOMO + 0.22EHB
0.95
0.84
0.35
Phenols
9
PII = −19.04 − 0.61E (vdw + chg ) + 3.11LUMO + 2.44 log P − 0.04SA
0.94
0.81
0.42
E (vdw + chg ), EHB, SA : intermolecular solute-membrane interaction descriptors
3.1.2. Skin Corrosion
Various models for skin corrosion have been reported by Barratt and collaborators
[70]-[75]. Most of these studies do not provide explicit classification models but
principal component (PC) analyses based on physicochemical properties, which allow
visual discrimination between corrosive (C) and non-corrosive (NC) chemicals. The
models are based on homogeneous datasets, i.e. acids, bases, electrophiles or neutral
organics. In addition to PC analyses, discriminant analysis (DA), stepwise regression,
and neural network (NN) analyses are also reported. In these studies, Barratt also
points out the potential of using combined in vitro cytotoxicity data together with
more conventional physicochemical parameters.
Tiered testing strategies for skin corrosion have been developed and assessed by
Worth and Cronin [76]-[78]. The evaluation of a two-step strategy, based on the
22
sequential use of pH measurements and in vitro data, indicated that the combined use
of these data improved the ability to predict corrosion potential [76]. Subsequently, a
three-step strategy was reported, based on the sequential use of QSARs, pH
measurements and in vitro data [79]. A separate study by Worth confirmed the
usefulness of pH as a predictor of skin corrosion potential, and provided a new
prediction model (PM) for identifying corrosive chemicals by a pH-dependent
mechanism [78].
Barratt et al. [70]-[71] reported a study to predict the presence or absence of skin
corrosivity potential for four datasets containing 30 mixed organic chemicals (acids,
bases, and neutral), 20 organic acids, 21 organic bases, and 15 phenols, respectively.
The descriptors were selected taking into account postulated mechanisms for skin
corrosion. It was hypothesised that the chemical must first penetrate into the skin, and
then it can kill the cells beneath the stratum corneum if it is sufficiently cytotoxic.
Thus, LogP values, molecular volume (MV), and melting point (MP) was used to
model percutaneous absorption, while the intrinsic acidity or basicity of chemicals
expressed by the pKa value was used to account for cytotoxicity. Classification data
and structure-corrosivity relationships were investigated separately for each subset of
chemicals, and an exploratory similarity analysis was carried out by using principal
components analysis (PCA). For each group of chemicals, plots of the first two
principal components (PC) for the four descriptors were presented. In all cases, the
chemicals classified as corrosive were clearly separated from those classified as noncorrosive. The counterbalancing effects of the different variables on corrosivity were
also illustrated by the plots. Chemicals with low LogP values, large molecular
volumes or high melting points (features associated with low skin permeability and
low solubility) were likely to be found in non-corrosive areas of the plots, unless they
were particularly acidic or basic (very low pKa values). On the other hand, the
mechanism of cytotoxicity of phenols was associated with hydrophobicity and the
proton-donating ability; this mechanism was characterised by the pKa value.
Barratt also reported models relating skin irritation and corrosivity data of neutral and
electrophilic organic chemicals [72]. As before [70], the parameters used modelled
skin permeability (LogP, and molecular volume), reactivity (dipole moment (DM))
and compensated for the fact that skin irritation/corrosivity testing is carried out with
23
a fixed mass or volume of chemical. Datasets were analysed by using stepwise
regression, discriminant analysis, and principal components analysis. Cross validation
was carried out by using a fixed deletion pattern of three subsets of data; three trials
were carried out with each subset omitted once. Although the models were reported in
the same paper, skin irritation and skin corrosion were modelled separately. Finally, a
stepwise approach to the skin corrosivity/irritation classification of neutral and
electrophilic organic chemicals was outlined.
Skin irritation. For the set of 52 neutral and electrophilic organic chemicals,
discriminant analysis between irritant and non-irritant compounds based on the above
parameters was able to discriminate well for only 73.1% of the dataset, and 67.3%
when cross-validated (Table 13). The poor discrimination at the irritant/non-irritant
classification boundary was attributed to biological variability. Stepwise regression
analysis of the Primary Irritation Index (PII) for the same dataset showed a poor
correlation (r2= 0.422; r2cv= 0.201) with a positive dependence on LogP and dipole
moment, and a negative dependence on molecular volume, which confirmed the
putative model for skin irritation (Equation 1). When the QSAR equation was used to
predict the PII values, the 17 misclassified chemicals formed a band with predicted
PII values lying between 2.07 and 3.48. This range of predicted PII values represented
the range of uncertainty for the discriminant analysis, for which no firm predictions
could be made.
Table 13. Classification results for each step of the discriminant model for skin irritation.
Step
Entering
Variable
Total fraction
well classified
Unclassified
Irritant
well classified
1
MV
0.615
0.645
0.571
2
DM
0.615
0.613
0.619
3
LogP
0.712
0.742
0.667
4
1/MW
0.731
0.742
0.714
0.673
0.677
0.667
Cross-Validation
Equation 1. Goodness-of-fit and robustness of the QSAR model for skin irritation.
PII = 1.047 log P − 0.244MV + 0.888 DM + 0.353
n = 52; r 2 = 0.422; rcv2 = 0.201; s = 1.376; F = 11.70
24
Skin corrosion. For the set of 40 electrophiles and two pro-electrophiles, discriminant
analysis based on LogP, MV and DM, was able to discriminate reasonably well
between corrosive and non-corrosive chemicals (92.9% correctly classified). The
addition of the third parameter, DM, did not increase the overall discrimination (Table
14). A plot of the first two principal components of the same parameters, accounting
for the 89.9% of the total variance, showed a clear demarcation between corrosive and
non-corrosive electrophiles. The predominant parameter in determining the skin
corrosivity of electrophilic organic chemicals appeared to be the molar dose at which
they are tested, because skin corrosivity testing is conducted using a fixed mass or
volume of chemical.
Table 14. Classification results for each step of the discriminant model for skin corrosion.
Step
Entering
Variable
Total fraction
well classified
Non-corrosive
well classified
Corrosive
well classified
1
LogP
0.833
0.867
0.750
2
MV
0.929
0.933
0.917
3
DM
0.929
0.900
1.0
0.929
0.933
0.917
Cross-Validation
In another study [73], the corrosive potential of a series of six fatty acids was
investigated in the in vitro skin corrosivity test (IVSCT), using both rat and human
skin. The variables used were the same as those used in the previously reported
studies, i.e. LogP, MV, and MP for percutaneous absorption, and pKa for cytotoxicity.
The dataset previously reported in [70], was extended with the six additional fatty
acids, and consisted of 27 organic acids. A plot of the first two principal components,
containing 84.3% of the variance of the original data for organic acids, was able to
discriminate between corrosive and non-corrosive chemicals on the basis of effects on
rabbit skin. However, for human skin there was a shift in the corrosive/non-corrosive
threshold. This was explained in terms of the effect of LogP: although skin
permeability increases almost linearly with 1ogP, at a certain point it begins to level
off. As LogP increases further, there is evidence that skin permeability of fatty acids
starts to decrease in human skin [47]. The consequence of this latter effect, together
with the increase in melting point with increasing alkyl chain length, is that fatty acids
with alkyl chains longer than 9 carbons are not corrosive to human skin. Due to
25
differences in barrier properties (stratum corneum thickness and lipid composition)
between rabbit and human skin, the shift in the corrosive/non-corrosive threshold
between animal and human skin was rationalised in terms of changes in the skin
permeability parameters. Chemicals on the borderline could be expected to be those in
which the classification was most sensitive to changes in skin type. Thus, the
classification of chemicals located in the boundary could be equivocal as a
consequence of variability of biological data. In contrast, the classification of
chemicals that were away from boundary regions were least likely to be affected by
such changes.
The datasets of organic acids, bases and phenols published in earlier works were
extended in a subsequent study by Barratt [74]. As in the previous studies, PCA on
LogP, MV, MP, and pKa confirmed that the analysis was able to discriminate between
corrosive and non-corrosive chemicals. In addition to PCA, the datasets were also
analysed employing neural networks (NN), using the same parameters as inputs. NN
were trained in the range 0.0 to 1.0, with non-corrosive chemicals being assigned the
value 0 and corrosive chemicals the value 1. The accuracies of the NN were 97.02%,
93.86%, and 95.85% for the acids, bases and phenols, respectively. As well as
yielding classification predictions in agreement with those in the training sets,
predicted outputs in the 0 to 1 range gave an indication of the confidence of the
predicted classification, especially in regions near to the classification boundaries,
where there can be uncertainty in the prediction. The regions of biological uncertainty
at the classification threshold between corrosive and non-corrosive chemicals were
also examined.
In a subsequent study, Barratt et al. [75] derived structure-corrosivity relationships
relating the severity of skin corrosivity of 27 acids (from [74]) designated by the EU
risk phrases R34 and R35, to descriptors for skin permeability and cytotoxicity. Skin
permeability was modelled by the usual parameters (LogP, MV and MP) whereas the
cytotoxicity was accounted for by pKa values and an in vitro cytotoxicity parameter
for the effects of the sodium salts of the acids on mouse embryo cells (pEC50). The
dataset was analysed by using PCA and back-propagation neural network (NN)
analysis. Plots of the first two PCs using only the conventional physicochemical
descriptors failed to distinguish the acids classified as R34 from those as classified as
R35. A plot of PCs 2 and 3 gave better resolution, but with some embedded clustered
26
data, possibly due to the presence of electrophilic groups conferring additional
cytotoxicity to that expressed by the pKa parameter. When in vitro cytotoxicity
parameters, expressed as pEC50 values, were included in the PCA plot, PCs 2 and 3
showed the best discrimination between acids classified as R34 and R35. Backpropagation networks contained a single hidden layer with four nodes. The values of
the inputs were transformed so that the lower limit was 0.0 and the upper limit 1.0.
The accuracy for predicted classifications ranged from 84 to 87%. The classification
predictions were in agreement for 26 of the 27 acids; when an outlier was omitted, the
accuracy achieved for classification increased to 96.26%. PCA and NN analysis were
both able to discriminate well between chemicals with the EU classifications R34, and
R35.
Tiered assessment schemes for the prediction of skin irritation and corrosion have
been designed and evaluated by Worth [49]. The first step of the process consisted of
classifying corrosivity based on melting point (MP) and molecular weight (MW). This
was followed by the use of a classification model (CM) based on pH (Step 2).
Subsequently, in vitro data from the EPISKIN assay were used (Step 3). If the
compound was ultimately classified as non-corrosive in the third step, a similar
iterative process was performed to identify skin irritants. Ultimately, if the compound
was predicted non-irritant, the in vivo Draize skin test was applied. Thus, the purpose
of such a scheme was to classify chemicals and to confirm negative classifications
with the use of animal tests.
In a subsequent study, Worth et al. [76] evaluated the uncertainty associated with the
predictive abilities of two-group classification models (CM), expressed in terms of
their Cooper statistics. Standard and percentile bootstrap resampling techniques were
used to judge whether predicted classifications were significantly better than the
predictions made by a different CM, or whether the performance of a CM exceeded
predefined performance criteria in a statistically significant way. This method was
illustrated by constructing 95% confidence intervals (CI) for the Cooper statistics of
four alternative skin corrosivity tests (TER, EPISKIN, Skin2, and CORROSITEX), as
well as two-step sequences in which each in vitro test was used in combination with a
physicochemical test for skin corrosion based on pH measurements. The PMs were
applied to a dataset of 60 chemicals already published [79]. Cooper statistics were
used to determine whether the four two-step sequences, with sensitivities greater than
27
or equal to 70%, were significantly more predictive than the four stand-alone in vitro
tests. This study showed that the performances of the TER, EPISKIN, Skin2 and
CORROSITEX tests in combination with the pH test were better than the individual
performances of the in vitro tests (Table 15).
Table 15. Classification results of the individual in vitro tests and the in vitro tests in sequence
with the pH test.
Bootstrap mean estimates
In vitro test
Concordance
Sensitivity
Specificity
TER
78
83*
87
96*
71
73*
EPISKIN
81
88*
83
96*
80
82*
Skin2
75
83*
45
70*
100
94*
CORROSITEX
73
74*
71
72*
76
78*
* Use of the in vitro test in sequence with the pH test
The bootstrap resampling technique was subsequently applied to assess the variability
of Draize tissue scores to estimate acute dermal and ocular effects [77]. This
technique was used to estimate biological variability arising from the use of different
animals, and temporal variability arising from different time-points. The estimates of
variability where then used to determine the extent to which Draize skin and eye test
tissue scores could be predicted. A dataset of 143 ECETOC chemicals was used for
the variability of Draize skin scores, while for eye scores the dataset consisted of 92
ECETOC chemicals. The results indicated that the variability in Draize skin scores
was such that no model for predicting PII could be expected to have r2>0.57; in
contrast, the variability in Draize eye scores was such that no model for predicting
MMAS could be expected to have r2> 0.81.
Finally, a paper by Walker et al. reports the so-called Skin Irritation Corrosion Rules
Estimation Tool (SICRET) that was developed to estimate whether chemicals are
likely to cause skin irritation or skin corrosion [80]. SICRET is a tiered approach that
uses physicochemical property limits, structural alerts and in vitro tests to classify
chemicals.
28
3.1.3. Eye Irritation / Corrosion
In comparison to skin irritation and corrosion, a more significant number of (Q)SAR
analyses have been reported to model local eye toxicity. The first reported models
were incorporated into or derived from expert systems. The model published by
Enslein et al. [81], based on structural fragments, was implemented in TOPKAT. In
contrast, the models published by Klopman et al. [82]-[84] used chemical
substructures implemented in MultiCASE (MultiCASE Inc.).
Enslein et al. [81] reported a QSAR study for the estimation of rabbit eye irritation
scores, using the same basis as the study performed on skin irritation [61]. For
statistical and modelling purposes, the original dataset (consisting of 592 compounds)
was divided into compounds without rings and compounds with at least one ring. The
descriptors used were Kier and Hall molecular connectivity indices and computergenerated MOLSTAC substructural keys, which account for the contribution of
structural fragments to eye irritation. Separate models were developed for the two
subsets by using discriminant analysis (DA); furthermore, two sub-equations were
developed for each model in order to separate the various classes of irritancy. After
collinearity analyses and detection of outliers, the subset of variables which explained
most of the variance was determined by an all-possible regressions algorithm, which
generates all the possible combinations for a predefined number of variables. DA
produces probabilities ranging from 0 to 1; the region from 0.3 to 0.7 was considered
as indeterminate, and therefore a region in which predictions should not be used. For
non-ring models, it was not possible to separate the mild irritants from the moderate
ones. The classification accuracy of the equation which separated negative
compounds from all others was 89.8%, with an indeterminate rate of 15%, while the
model separating severe irritant compounds from all others yielded an accuracy of
90.2%, with an indeterminate rate of 30.9%. In contrast, for ring-models, there was
92.7% accuracy for the discrimination between negative/mild irritants and
moderate/severe ones, and a 87.8% accuracy for the separation of severe irritants from
all others. The classification accuracies of the QSAR models is summarised in Table
16. In general, the false negative rate turned out to be substantially higher than the
false positive one.
29
Table 16. Classification results of the discriminant QSAR models for eye irritation.
Models
Non-ring
compounds
Ring
compounds
Subequations
Overall
Accuracy
False
Positives
False
Negatives
Indeterminate
F
(p<0.001)
Negatives vs
Others
89.8%
3.2%
(22.5%)*
7.0%
(8.2%)*
15.0%
25.3
Severes vs Others
90.2%
2.7%
(4.3%)*
7.1%
(19.1%)*
30.9%
15.2
Negative/Milds vs
Moderate/Severes
92.7%
2.9%
(8.1%)*
4.45
(6.9%)*
29.3%
8.63
Severes vs Others
87.8%
6.6%
(9.9%)*
5.6%
(16.7%)*
24.3%
10.0
* Percentage of false predictions in relation to the experimental class
The Multi-CASE SAR method was used to develop a predictive model of ocular
irritation based on the evaluation of the significance of chemical structure in the
induction of eye irritation [82]-[83]. The original Draize dataset was augmented and a
new QSAR model was reported [84]. Multi-CASE was trained with a non-congeneric
database of 186 organic chemicals previously evaluated in a modified Draize test for
their ability to cause eye irritation in vivo (by Klopman et al. [82]). The eye toxicity of
the compounds was ranked by giving an index of 15 to non-toxic molecules, 25 to
those listed as marginally toxic, 35 to toxic molecules, 45 to very toxic molecules, and
55 to those found to be extremely toxic. A total of 37 functionalities and substructural
attributes were found to be relevant for eye irritation potential, including carboxylate
and sulphonate groups (in anionic surfactants), phenolic moieties (in non-ionic
surfactants), amino groups, halogenated and unsaturated structures, anhydrides and
epoxides. Ester groups were recognised as having a low irritation potential. The
Multi-CASE program was then used to predict the eye irritation potential of 21 test
chemicals, chemical mixtures, and polymers. Five of the compounds contained
substructural fragments not encountered in the training set, but the remaining 18
compounds were successfully predicted.
The eye irritation database of the previous study consisting of 186 compounds,
extended with 21 compounds, was studied by using a Baseline Activity Identification
Algorithm (BAIA) [83] that uncovers the underlying relations between the biological
activity of a diverse set of molecules and global parameters. Chemicals were ranked
by a qualitative score, according to the irritation potential. Following elimination of
40 molecules, the correlation between LogP and the remaining 167 molecules yielded
30
r=0.60, and F=93.2 (Equation 2). Correlations between irritancy and LogP revealed
that hydrophilic molecules tend to be more toxic. However, the variation in activity
with changes in LogP was rather small. The program identified 20 substructures
accounting for the toxic activity of the molecules (biophores).
Equation 2. Goodness-of-fit and robustness of the QSAR model for eye irritation.
eye scores = 34.5 − 3.155 log P
n = 167; r = 0.60; F = 93.2; SD = 13.637
A different SAR study based on the results of 297 chemicals tested in the Draize eye
irritation assay was developed with Multi-CASE [84]. Prior to the SAR analyses, the
major differences between the physicochemical properties of ocular irritants (I) and
non-irritants (NI) were determined: in comparison with NI chemicals, I chemicals had
significantly lower molecular weights, higher aqueous solubilities, and lower
lipophilicities. The SAR analysis indicated that chemical reactivity was not associated
with irritancy; in fact, I were significantly less electrophilic (lower LUMO), less
nucleophilic (lower HOMO), less electronegative, and less reactive (greater electron
gap) than NI. The authors used the Multi-CASE expert system to identify biophores
(substructures which occur with a significantly greater frequency in irritants than in
non-irritants) and biophobes (substructures which occur significantly more frequently
in non-irritants than in irritants). The major structural determinants included primary,
secondary and tertiary amine groups (i.e. basic groups), as well as carboxylate,
organosulphate and sulphonate groups (i.e. acidic groups). Multi-CASE analyses
resulted in the development of SAR models which explained the 98.8% and 99.3 % of
the activity, respectively. However, the activities of many compounds (n=22) were
explained by unique biophores that occurred only once. This finding suggested that
chemicals in the training set had a low informational content (41.5%), and a restricted
chemical coverage. The predictivity of the SAR submodels derived from the dataset
was determined by a ten-fold cross-validation procedure. The overall predictivity
determined for the individual SAR submodels ranged from 66.3% to 69.7%, yielding
69.7% for the consensus model (Table 17). The predictive performance determined
after removing chemicals with unique biophores (out of the applicability domain)
improved to 74.2% for the consensus model. Thus, the major structural determinants
included hydrophilicity, alkalinity, acidity, and specific receptor-binding interactions.
31
Table 17. Predictivity of the SAR submodels and consensus model for eye irritation.
SAR submodel
Cut-off
Sensitivity
Specificity
Concordance
MultiCASE potency
18 units
0.47
0.56 *
0.85
0.86 *
66.7%
72.4% *
MultiCASE prob.
49%
0.49
0.58 *
0.83
0.83 *
66.3%
72.0% *
CASE potency
18 units
0.56
0.59 *
0.83
0.82 *
69.7%
71.3% *
CASE prob.
66%
0.46
0.51 *
0.87
0.87 *
67.0%
70.6% *
Consensus Model
50%
0.60
0.69 *
0.80
0.79 *
69.7%
74.2% *
* without unique fragments
MultiCASE is the updated version of CASE, which includes CASE algorithm as an option; the new
program MultiCASE includes the ability to deal with subtle geometrical differences, with
synergies, and with logical analysis and hierarchical decision making.
Later, Sugai and coworkers published two papers: a QSAR model developed by
applying the adaptive least squares (ALS) method to a heterogeneous dataset [85], and
a QSAR for a reduced group of salicylates [86]. The first paper [85] reported a study
of the relationship between the structural features of a set of 131 heterogeneous
chemicals and the primary eye irritation in rabbits. The primary eye irritation ratings
were classified into three classes on the basis of the recovery time of corneal and
conjunctival damages: class I (recovering of damages within 24 hours); class II
(damages persisting for more than 24 hours but recovering within 21 days); and class
III (damages not recovering within 21 days). Thirty six substructural features were
used to describe the molecules, i.e. a physicochemical descriptor, 34 descriptors
accounting for the presence/absence/number of occurrences of structural fragments,
and an indicator variable accounting for structural fragments and moieties. A QSAR
model was developed by using the nonparametric pattern classifier adaptive least
squares (ALS) method. After a preliminary selection of descriptors by backward
stepwise elimination, several subsets of descriptors were selected by using an everypossible-combination technique that minimised the number of misclassified
chemicals. The derived three-class discriminant function included 18 descriptors. The
descriptors with high positive contributions in the discrimination function (rings with
one or two N atoms, –SO2–, –NO2, C=O bonds, and –COOH and –OH on benzene
ring, and halogens) indicated the substructural features which contribute to the
enhancement of the affinity of chemicals for proteins or lipids in ocular tissue due to
electronic effects, resulting in irritation. In contrast, the descriptors with negative
32
contributions (R–O–R’, and –NH2) were associated with a reduced likelihood of
irritation. The accuracy in classifying the chemicals into three classes was 86.3% in
the recognition step, and 74.0% in the LOO prediction step. A prediction of irritation
potential of an external set of 28 diverse chemicals supported the reliability of the
discriminant function, with an accuracy of 72.4% and all the misclassified chemicals
assigned to the nearest neighbouring classes (Table 18).
Table 18. Classification results of the discriminant models for eye irritation.
N. misclassifications
N. misclassifications
(by 2 grades)
Spearman rank
correlation coeff.
Accuracy
Recognition
18
0
0.869
86.3%
LOO prediction
34
0
0.736
74.0%
External prediction
18
0
-
72.4%
In the second part of the previously reported study [85], the primary eye irritation
potential of ten salicylates was assessed by in vivo and in vitro studies [86]. The
effects of salicylates on human serum γ-globulin (HSG) and lipid matrix were
investigated with an in vitro model for the prediction of eye irritation. A nondestructive structural analysis of the corneal surface treated with salicylates was
performed with IR spectroscopy. QSARs were derived for the irritation potential and
in vitro effects of salicylates. A good correlation was obtained between the ability of
salicylates to produce HSG aggregation and protein denaturation; in addition,
statistical correlations were obtained between HSG aggregation of and the corneal and
conjunctival scores. However, poor correlations were obtained between the Draize
score and the lipid-substance interaction parameter. The physicochemical descriptors
(acid dissociation constants and partition coefficients) were correlated with in vivo
and in vitro data (Table 19). For in vivo data, pKa values were correlated with the
cornea scores (r=–0.73) but not with the conjunctiva scores (r=–0.37). Salicylates
with low pKa values led to an increase in the cornea score. LogP values did not
correlate with either the cornea scores or the scores of conjunctiva. For in vitro data, a
significant correlation was found between the aggregation of HSG and pKa values
(r=–0.82), and salicylates with low pKa values led to an increase in the production of
HSG aggregates. A poor correlation was obtained between the aggregation of HSG
and the LogP values. Thus, for salicylates, increasing dissociation potential and
electron withdrawing properties were important factors in inducing corneal
33
opacification and HSG aggregation. The corneal irritation and protein aggregation
potential of salicylates were found to correlate with the acid dissociation constant
more closely than the octanol/water partiion coefficient. Thus, eye irritation caused by
salicylates was mainly attributed to the denaturation of proteins in ocular tissue, while
effects on proteins depended on the dissociation potential of molecules.
Table 19. Goodness-of-fit of the linear regression QSAR models for eye irritant salicylates.
Correlation
N. chemicals
Equation
r (P<0.05)
pKa vs S (cornea)
10
S = −31.0 pKa + 144.1
–0.725
pKa vs S (conjunctiva)
10
S = −1.53 pKa + 11.7
–0.365
LogP vs S (cornea)
10
S = −3.72 log P + 60.0
–0.195
LogP vs S (conjunctiva)
10
S = −0.184 log P + 6.80
–0.100
Log(1/C) vs pKa
10
log(1 C ) = −0.277 pKa + 2.84
–0.820
Log(1/C) vs LogP
10
log(1 C ) = 0.023 log P + 1.98
0.155
S: in vivo Draize scores; log(1/C): in vitro aggregation of HSG
Lipnick et al. [87] analysed a dataset of 42 aliphatic alcohols and ketones, consisting
of eye scores on an integer scale of l-10. Parabolic and bilinear models using LogP
were fitted to model the eye score; the bilinear model yielded a better fit.
Cronin started a series of papers analysing different statistical methods applied to an
eye irritation dataset. He attempted to find linear regression relationships for a
physically heterogeneous set of compounds [88]. However, the impossibility of
deriving meaningful (Q)SAR models suggested the need to restrict the analysis to
pure bulk liquids, developing submodels for limited datasets of alcohols and acetates.
Discriminant analysis revealed an embedded cluster of irritant chemicals within the
non-irritant chemicals, which was investigated further by Cronin [89]. Subsequently,
the method of cluster significance analysis (CSA), which can be used to determine the
significance of embedded clustering, was applied to the eye irritation dataset, and the
technique of Embedded Cluster Modelling (ECM) was developed to generate elliptic
prediction models for the embedded clusters [90]. The series of papers is closed with a
comparative analysis [91] that illustrated the use of several linear statistical techniques
to develop classification models (CM), i.e. linear discriminant analysis (LDA), binary
logistic regression (BLR) and classification tree (CT) analysis.
34
Cronin et al. [88] reported a number of QSAR investigations based on an ECETOC
dataset of 53 chemical diverse compounds with measured Draize eye irritation scores
(DES). These numerical pseudo-quantitative parameters based on the summation of
subjective scores were adjusted with the molarity of pure chemicals into molar eye
scores (MES). The dataset was first analysed by using a set of 23 physicochemical
descriptors accounting for hydrophobicity, size and shape, and electronic
characteristics (Table 20). Three QSAR models obtained by stepwise regression
analysis to all compounds in the dataset (n=38) and to the 23 physicochemical
descriptors yielded poor statistical results: from r2=0.16 (using the square of LogP) to
r2=0.35 (adding LUMO and the zero-order Kappa-alpha index). These results
suggested that different chemical groups were driven by varying mechanisms of toxic
action. QSAR analyses were attempted for defined chemical classes (alcohols and
acetates) expected to follow a similar mechanism of action. When the dataset was
split into subclasses, the QSAR model for 9 alcohols correlating the molar eye scores
with LogP and HOMO achieved r2=0.80. For the subgroup of 7 acetates a correlation
with LogP yielded r2=0.74. Stepwise linear discriminant analysis on the entire dataset
with the 23 descriptors found no significant variables that could discriminate between
irritants and non-irritants. This suggested that there was no linear distribution of the
irritants with any of the variables considered. Principal components analysis of the 32
parameters revealed that four components were sufficient to account for 92.6% of the
original information. As a pattern recognition technique, the pairwise score plots of
the four PCs for the 23 variables revealed that active compounds formed an embedded
cluster surrounded by the diffuse cluster of inactive chemicals.
Table 20. Goodness-of-fit of the linear regression QSAR models for eye irritants.
Compounds
All
compounds
Alcohols
Acetates
r2
S
F
MES = −0.118(log P ) + 3.33
0.16
2.56
8.2
38
MES = −0.125(log P ) + 0.713LUMO + 2.23
0.27
2.4
7.7
38
MES = −0.156(log P ) + 1.13LUMO + 0.397κα 0 − 0.44
0.35
2.26
7.6
9
MES = 2.06 log P + 4.13
0.72
1.89
21.3
9
MES = 1.72 log P + 9.04 HOMO + 102
0.80
1.58
17.3
7
MES = −1.37 log P + 2.97
0.74
0.52
18.2
n
Equation
38
2
2
2
MES: Molar Eye Score
35
The presence of an embedded cluster was corroborated by an further paper [89],
which identified the parameters that formed an embedded cluster of eye irritants
among non irritants. Out of the total of 23 physicochemical descriptors, the five most
significant were LogP, LogP2, heat of formation, dipole moment, and zero-order
valence corrected molecular connectivity. A representation of heat of formation
against LogP did not achieve full separation. However, PCA on the significant
parameters allowed the embedded cluster to be visualised.
Worth and Cronin reported another example of an embedded cluster prediction model
derived from the ECETOC dataset [90]. The statistical significance of the embedded
cluster was assessed by cluster significance analysis (CSA). The embedded cluster
modelling (ECM) method allows the generation of elliptic models of two or more
dimensions that can be used to identify embedded data. This paper reported a twodimensional ECM model for eye irritation potential based on 73 diverse organic
ECETOC chemicals, using LogP (which models chemical diffusion) and a sizeindependent molecular connectivity index (dV1, which accounts for the degree of
branching and cyclicity). The sensitivity and specificity were 73% and 78%,
respectively. The irritant region of parameter space was defined by a LogP range
between 0.7 and 1.43, and an intermediate degree of branching or cyclicity.
An further investigation by Worth and Cronin [91] illustrates the use of several linear
statistical techniques to develop classification models (CM), i.e. linear discriminant
analysis (LDA), binary logistic regression (BLR) and classification tree (CT) analysis,
for predicting the presence or absence of eye irritation. The different methods were
applied to a dataset of 119 organic liquids classified as I or NI according to EU
classification criteria. The CMs were developed by applying LDA, BLR, and CT
analyses, using a single predictor variable (molecular weight), and assigning equal
probabilities for the two classes (I/NI). The cut-off values below which a chemical
should be predicted to be irritating to the eye were 121, 77, and 137 g/mol, in the
LDA, BLR, and CT classification models, respectively (Table 21). Although all the
statistical techniques yielded comparable models, with concordances between 61 and
76%, the most appropriate model was developed by the non-parametric CT analysis,
since the assumptions of LDA and BLR had not been met. Some differences outlined
for the three methods were that: a) LDA and BLR make a number of assumptions
about the underlying data, whereas CT analysis is a non-parametric technique; b)
36
LDA and BLR can be used to generate a distinct probability of group membership for
individual chemicals, whereas CT analysis only produces average probabilities for the
different groups; c) LDA and BLR use all (statistically significant) predictor variables
simultaneously in the model, whilst CT analysis uses the predictor variables in a
hierarchical and recursive manner; d) LDA and BLR models assign all objects to a
pre-defined group in a single step, whereas CT have the flexibility of assigning the
objects in one or more steps.
Table 21. Classification results of the different models of eye irritancy.
CM (p<0.01)
Cut-off value
Sensitivity
Specificity
Accuracy
Linear Discriminant
Analysis (LDA)
if MW ≤ 121 g/mol, then predict
I; otherwise, predict NI
73
62
65
Binary Logistic
Regression (BLR)
if MW ≤ 77 g/mol, then predict
I; otherwise, predict NI
27
93
76
Classification Tree
(CT)
if MW ≤ 137 g/mol, then predict
I; otherwise, predict NI
97
49
61
Following the same PCA approach used to assess skin irritation and skin corrosion
[70], Barratt developed a series of “models” for eye irritation. He reported principal
component analysis (PCA) able to discriminate between I and NI, among reduced
series of neutral organic chemicals [92]-[93]. In the first paper [92], Barratt used a
variation of the extended ECETOC dataset of neutral organic chemicals previously
used by Cronin [88], composed of 46 compounds. The parameters used to model eye
irritation potential, i.e. LogP, the minor principal inertial axes (Ry and Rz) and dipole
moment (DM), were selected according to mechanistic considerations. The choice of
DM as a reactivity parameter was based on the premise that low molecular-weight
neutral molecules can alter electrical resistance and thus the ion permeability of cell
membranes. It was hypothesised that changes in electrical resistance in the
membranes of eye might lead to cytotoxicity and result in irritation. The choice of
LogP and parameters accounting for molecular dimensions was based on the
expectation that chemicals should be able to partition into the membrane and hence
should possess appropriate hydrophobic/hydrophilic properties. All three principal
inertial axes were used in the initial studies, but the size and shape parameters which
gave the best discrimination were Ry and Rz, which model the cross-sectional area of
37
the long axis of molecules. The four descriptors were analysed by PCA. The plot of
the first two PCs of the four parameters was able to discriminate between irritant and
non-irritant chemicals in the dataset. Thus an eye irritant would be a chemical
hydrophobic enough to allow partitioning into a biological membrane whilst also
possessing a moderate dipole moment.
To support the validity of the original four-parameter eye irritation model for neutral
organic chemicals, the previous study of Barratt [92] was followed up by additional
analyses [93]. In the latter study, the same descriptors used to predict the eye irritation
potential of neutral organic chemicals (LogP, the minor principal inertial axes (Ry and
Rz) and DM) were applied to a dataset of 34 aliphatic alcohols previously reported by
Lipnick et al. [87], with measured eye irritation scores on an integer scale of 1-10. In
addition, the previous dataset of neutral organic chemicals was extended to 57
chemicals, including some aliphatic alcohols to increase the chemical parameter space
covered by the prediction model. As in other studies [74]-[75], for the first dataset of
aliphatic alcohols, back propagation NN analysis was carried out by using the four
selected parameters as input, and using an integer scale from 1 to 10 for the eye
irritation scores output. The most important variables turned out to be Rz and LogP,
with the accuracy of the network being 97.27%. A plot of predicted values of the eye
score against actual eye scores showed a strong correlation (r=0.9). However, despite
this good fit, there was a notable uncertainty in the prediction of eye score values,
possibly due to biological variability. For the extended dataset of 57 neutral organic
chemicals and aliphatic alcohols used in the previous paper [92], PCA based on the
same four physicochemical parameters concentrated 89.7% of the variance into the
first two components, which showed a clear discrimination between irritant and nonirritant chemicals. The dataset was also investigated by NN analysis, using the same
inputs and an integer scale from 0 (for NI) to 1 (for I). The accuracy was 97.37%,
with the dipole moment being the most important variable.
Abraham et al. [94] used parameters derived from chromatography experiments to
model a dataset comprising 38 of the 53 organic liquids previously analysed by
Cronin et al. [88]. Subsequently, Abraham et al. [95]-[97] extended the dataset and
studied the relationship between the mechanism of action in the Draize test and in
humans. In Cronin et al. [88], it was concluded that the lack of success of the QSAR
analysis was possibly due to varying mechanisms of toxic action of the different
38
chemical groups, and to the choice of descriptors. In the study by Abraham et al. [94],
a transport-driven mechanism was hypothesised for the process of transfer of an
irritant from the state of a pure organic liquid to a dilute solution in an organic solvent
biophase (model for the biological structure of the eye). In order to obtain a response
related to the transfer process, DES was related to the activity coefficient (γo), which
is the equilibrium constant for the transport mechanism, by means of a vapour
solubility in the solvent phase (L) and the saturated vapour pressure of the pure liquid
(Po). The descriptors used for the QSAR model were compound (or solute)
parameters: excess molar refraction (R2), polarizability/ dipolarity ( π 2H ), effective
hydrogen bond acidity and basicity ( ∑ α 2H and
∑β
H
2
), and the logarithm of the
vapour solubility in hexadecane (logL16). Two QSARs were modelled (Table 22).
When using the DES, a very poor equation was obtained (r2=0.27). However, a
reasonable correlation was achieved (r2=0.89) when using the modified log (DES/Po);
the results improved with the removal of an outlier and the removal of a nonsignificant descriptor (r2=0.95). On transforming the calculated log (DES/Po) values
back to calculated DES, the correlation between log(DES) observed and calculated
yielded a good agreement r2=0.77.
Table 22. Goodness-of-fit of QSAR models for eye irritant pure organic liquids.
N
Descriptors
r2
SD
F
38
log(DES/Po) vs R2, π 2H , ∑ α 2H , ∑ β 2H and logL16
0.894
0.46
54.0
0.953
0.32
125.5
37
37
log(DES/Po) vs π 2H , ∑ α 2H , ∑ β 2H and logL16
0.951
0.32
155.9
37
log(DES) observed vs log(DES) calculated
0.771
0.30
117.6
In a subsequent study, Abraham et al. [95] combined Draize eye scores (DES) and
human eye irritation thresholds (EIT) into a single QSAR, assuming that the main
feature of both in vivo responses was the transport of the compound into the biophase.
They extended the original dataset of measured DES for 38 pure organic liquids with
a dataset of 17 volatile organic compounds (VOC) for which EIT had been measured.
The two scales of irritation, DES and EIT, were compared, establishing 0.66
logarithmic units of difference. The models found in the previous study were redeveloped for the new extended dataset of 54 compounds (Table 23), obtaining
r2=0.928 and r2=0.924 for the models with and without the R2 variable. Finally, EIT
39
scores were predicted from DES values, and they were compared with experimental
EIT.
Table 23. Goodness-of-fit of QSAR models for the eye irritancy of pure organic liquids.
r2
r2cv
SD
F
log(DES/P ) vs R2, π 2H , ∑ α 2H , ∑ β 2H and logL
0.928
0.913
0.363
123.8
log(DES/Po) vs π 2H , ∑ α 2H , ∑ β 2H and logL16
0.924
0.910
0.368
149.6
N.
Descriptors
54
54
o
16
In a more recent study, Abraham et al. [97] extended the connection between the eye
irritation effects on rabbits (MMAS) modified by vapour pressure, and the effects on
humans (EIT) [95]. They combined two different datasets into an extended dataset of
91 compounds. The total dataset for 91 compounds was analysed by the same general
solvation equation (Table 24). To establish the relationship between MMAS and EIT
and to determine whether the two datasets were compatible, an indicator variable (I)
to account for the dataset was set. The same model for eye irritation used in the first
paper of the series [94] was re-developed, and a single QSAR covering a wide range
of compound types fit the two sets with r2=0.94. When the equation was applied to
the subset with calculated MMAS values, comparable statistics were obtained
(r2=0.95). To test the predictive power of the model, the dataset was split into a
training set of 45 compounds; the corresponding equation for the training set
compounds (r2=0.921) was used to predict the remaining 46 test set compounds. The
mechanistic assumption underlying this combined model was that rabbit DES and
human EIT result from the passive transfer of the compound from the bulk liquid or
the vapour phase to a biophase. This passive transfer is assumed to be non-specific,
unless there is some particular interaction between the substituents for which there is a
transfer from the biophase to a receptor. In such a case, the shape of a compound
could be crucial, so to investigate whether QSARs might be improved by the
inclusion of other types of descriptors, three shape descriptors were included by using
an energy minimised conformation: three-dimensional volume (VG), the longest
length of the minimum energy conformation (LG), and a shape descriptor that takes
into account branching and the distance of atoms from a centre of gravity of a
molecule (DPO1). The additional shape descriptors had almost no effect (r2=0.94).
40
Table 24. Goodness-of-fit of QSAR models for eye irritants.
N.
Descriptors
r2
SD
AD
AAD
F
91
log(DES/Po) vs R2, π 2H , ∑ α 2H , ∑ β 2H ,
0.936
0.433
0.000
0.340
204.5
68
log(DES/Po) vs R2, π 2H , ∑ α 2H , ∑ β 2H
0.954
0.398
0.000
0.318
255.8
45
log(DES/Po) vs R2, π 2H , ∑ α 2H , ∑ β 2H
0.921
0.469
0.000
0.354
74.0
log(DES/Po) vs R2, π 2H , ∑ α 2H , ∑ β 2H ,
0.941
0.422
0.000
0.334
144.0
91
logL16, I
and logL16
and logL16 – internal prediction
logL16, I, DPO1, VG, LG
AD: Average Deviation; AAD: Absolute Average Deviation
DPO1, VG, LG: shape descriptors
Kulkarni and Hopfinger developed a new approach, called Membrane-Interaction
QSAR (MI-QSAR) analysis [98], which combines structure-based design techniques
with the traditional QSAR paradigm. Molecular dynamic simulations (MDS) are used
to generate membrane–solute interaction properties for the interaction of organic
chemicals with model phospholipid membranes. This methodology allows
investigating structurally diverse training sets using a genetic algorithm approach. MIQSAR was applied to the neutral organic chemicals within the ECETOC eye irritation
dataset [99].
The seminal study of Kulkarni et al. [98] reports a MI-QSAR model based on a
hypothesised mechanism of eye irritation and developed by using a diverse Draize eye
irritation dataset of 38 compounds. Molecular dynamic simulation was used to
generate intermolecular membrane-solute interaction properties. These intermolecular
properties were combined with intramolecular physicochemical properties and
features of the solute (irritant) to construct QSAR models by using multi-dimensional
linear regression and the genetic function approximation (GFA) algorithm. The
principal correlation descriptors, solute aqueous solvation free energy and solutemembrane interaction energies, were selected from a trial set of 95 descriptors for 16
structurally diverse compounds fully representative of the whole dataset (aliphatic and
aromatic hydrocarbons, and aliphatic ketones, alcohols and acetates). A QSAR based
41
on three parameters (two energy terms and a topological index) showed a good fit
(r2= 0.92).
Kulkarni et al. [99] reported a second MI-QSAR study to predict molar-adjusted
Draize scores for the ECETOC dataset of 38 structurally diverse compounds. The
mechanistic hypothesis for eye irritation relied on the uptake and diffusion of an
irritant into the keratocytes of the corneal epithelium. Thus, uptake and interaction of
each solute molecule with a model phospholipid membrane were simulated. The
membrane-solute and solute-solvent intermolecular interaction properties estimated
from molecular simulations were added to intramolecular physicochemical property
descriptors to provide an extended set of descriptors. Thus, the descriptors used were
physicochemical parameters associated with the solute molecule and with the
proposed receptor (intramolecular properties), as well as parameters describing the
interaction between both (intermolecular solute-membrane properties). Linear MIQSAR models for the 38 compounds were built including the outliers from the
previous study [98] (Table 25); the MES was correlated with the aqueous solvation
free energy, and additively with the electrostatic solute-membrane interaction energy,
obtaining r2=0.15 for both cases. In contrast, a GFA parabolic non-linear model based
on the previous descriptors, Van der Waals solute-membrane interaction energy, and
octanol solvation free energy, yielded r2=0.78. However, when removing the only
two irritants from the model, a good-linear MI-QSAR model (r2=0.71) could be
achieved. Eye irritancy was predicted to increase with the aqueous solubility of the
molecule, with the binding energy of the solute to the phospholipid regions of the
membrane, and with LUMO (as reactivity increases).
42
Table 25. Goodness-of-fit and robustness of the linear and non-linear MI-QSAR models for eye
irritants.
Model
N.
Model
r2
r2cv
LSE
Linear
38
MES = 1.77 − 0.18F (H 2 O )
0.15
–0.18
6.40
Linear
38
MES = 1.74 − 0.004 E (chg ) − 0.18F (H 2O )
0.15
–0.20
6.30
Nonlinear
GFA
38
0.78
0.73
1.52
(F=32.4)
Linear
36
0.71
0.65
1.15
(F=38.9)
{
MES = 1.76 − 0.06[F (H 2O ) − 6.5] + 0.05[F (OCT ) − 4.5]
2
2
}
− 0.01[E (vdw) − 3.9]2 −0.13E (chg ) + 0.05LUMO 2
MES = −0.66 − 0.01E (chg + vdw) − 0.48F (H 2O ) + 0.39 LUMO
F (H 2O ), E (chg ), F (OCT ), E (vdw), E (chg + vdw) : intermolecular solute-membrane interaction descriptors
A more recent MI-QSAR study has been published by Li et al. [100]. QSAR models
were developed to predict drug permeability coefficients across cornea and its
component layers (epithelium, stroma, and endothelium). From a training set of 25
structurally diverse drugs, QSAR models were constructed and compared for the
permeability of the component layers. Cornea permeability was found to be dependent
on the measured distribution coefficient, the cohesive energy of the drug, the total
potential energy of the drug-membrane "complex" and three other energy refinement
descriptor terms. The endothelium might be a more important barrier in cornea
permeation than the stroma. Thirteen structurally diverse drugs, whose molar-adjusted
eye irritation scores (MES) were measured by using the Draize test, were chosen as an
eye irritation comparison set. A poor correlation (r2= 0.02) between the measured
MES and the predicted cornea permeability coefficients for the drugs in the eye
irritation set suggested there was no significant relationship between eye irritation
potency and the corneal permeability.
Patel [101] reported a method that considers mixtures of chemicals, called
Quantitative Component Analysis of Mixtures (QCAM). The QCAM method was
used to analyze an existing set of product formulation data to determine if the
irritating ingredients in the mixtures could be identified [101]. Four eye irritation
scores, based on a rat model, were analysed by QCAM for 18 mixtures having a
composite total of 37 components. QCAM permits the construction of quantitative
models for extracting the individual toxicity profile of each of the components of a
mixture from knowledge of their relative concentrations and the net toxicity of the
entire mixture. Thus, QCAM relates a net toxicity measure of a mixture to the
43
toxicities of the individual components through linear, quadratic, and pairwise crosscomponent concentration-dependent interactions. A correlation model was established
by using a genetic function approximation (GFA) algorithm employing either
multidimensional linear regression (MLR) or partial least-squares (PLS) regression. A
correlation analysis of the eye irritation scores was carried out (Table 26). Each of the
standard deviations (SD) was correlated with its respective mean value (MV): cornea
and average SD correlated significantly with MV, and the larger the MV (greater eye
irritation score), the larger the SD. This indicated that corneal and average eye
irritation scores define NI compounds. Linear correlation equations between eye
irritation scores and selected components (Table 27) showed that corneal eye irritation
and average eye irritation are well-explained in terms of a linear model containing, at
most, three components over the set of mixtures (r2= 0.83). It was also found that
extensive cornea and average eye irritations were due to only one of these three
components of the mixtures. In addition, one of the three significant components was
predicted to decrease the extent of eye irritation, and subsequently identified as an
anti-irritant in contact lens solutions. A reasonable linear correlation model could also
be developed for conjunctival irritation, but no significant iris irritation model could
be constructed. The addition of quadratic and/or cross-component concentration terms
to a linear correlation model did not statistically improve the overall resultant model.
The QCAM models permitted estimation of the intrinsic (self) toxicity of each of the
components of a mixture.
44
Table 26. Linear Correlations between Standard Deviations (SD) and Mean Values (MV) for four
eye irritation scores.
Eye irritation scores
Linear Equations
r2
Cornea
SD = 0.81MV – 1.2
0.91
Iris
SD = 0.17MV – 1.1
0.08
Conjunctiva
SD = – 0.03MV + 0.3
Average
SD = 0.77MV – 3.8
0.26
0.90
SD: Standard Deviation; MV: Mean Value
Table 27. Goodness-of-fit and robustness of QCAM models for eye irritation.
Cornea Irritation
Scores
Iris Irritation
Scores
Conjunctiva
Irritation Scores
Average Eye
Irritation Scores
r2
r2cv
r2
r2cv
r2
r2cv
r2
r2cv
1
0.71
0.60
0.23
0.09
0.52
0.41
0.74
0.65
2
0.83
0.74
0.42
0.29
0.63
0.36
0.84
0.77
3
0.88
0.83
0.62
0.10
0.71
0.38
0.88
0.81
N. components
Patlewicz et al. [102] used a non-linear neural network (NN) analysis applied to
cationic surfactants (Hayashi et al. also reported a QSAR for surfactants [103]). They
reported a non-linear QSAR model based on 29 in vivo rabbit eye irritation tests for
19 different cationic surfactants, and derived by back propagation NN analysis [102].
The parameters used were LogP and molecular volume to model the partitioning of
the surfactants into the membranes of the eye; the logarithm of the critical micelle
concentration (logCMC), which is a measure of the reactivity of the surfactants with
the eye; and surfactant concentration. MAS scores showed strongly positive, nonlinear correlations with surfactant concentration, and logCMC showed a strongly
negative, non-linear correlation with LogP. The accuracy of the NN was 91.18%,
while the Pearson correlation between the actual and predicted values of MAS was
0.838, showing that around 70% (r2=0.702) of the variance in the dataset was
explained by the model.
Patlewicz et al. [104] also proposed a refined QSAR model for prediction of MAS,
using the same cationic surfactant dataset. In this case, the non-linear approach used
was a Bayesian neural network (BNN) that provided a robust and accurate model
(r2=0.89). The model was used to explore the relationship between MAS score and
45
molecular properties. The simulated results were used to identify the range of
molecular properties of potential molecules with a probability greater than 75% of
causing more than mild irritancy. Within the parameter space of the model, the
probability of causing severe irritation was highest for molecules with MV < 320 Å3,
and contained within the following ranges: - 2 < log P < 14 , and - 6 < log CMC < 3 .
46
4. Alternative test methods
An alternative method is any method used to replace, reduce or refine (3 Rs) the use
of animal testing methods for hazard and risk assessment. In relation to alternative
methods, Council Directive 86/609/EEC states that “An experiment shall not be
performed if another scientifically satisfactory method of obtaining the result sought,
not entailing the use of an animal, is reasonably and practicably available” [105].
The objectives of this directive are to diminish the number of materials needing any
animal testing, to decrease the number of animals per test, to minimize animal
distress, to decrease ancillary testing and to ensure development of information
suitable for hazard classification.
Annex XI of REACH describes possibilities for waiving testing requirements, on the
basis of exposure considerations, the use of existing data, (Q)SARs and chemical
grouping, and in vitro tests in order to minimise the need for animal testing [10].
4.1.
Use of existing data
The use of existing data includes the use of data on physicochemical properties,
human and animal data, and non-guideline studies. Concerning the use of existing
human data, an assessment of available human data is requested for all toxicological
endpoints, including skin irritation/corrosion, and eye irritation. In addition, Annex XI
states that historical human data, such as epidemiological studies on exposed
populations, accidental or occupational exposure data and clinical studies, shall be
considered.
In addition, articles from peer-reviewed literature as well as study reports from
companies can constitute a resource for the safety evaluation of chemicals and
substances.
4.2.
In vitro methods
The number of stand-alone in vitro tests that have been validated and incorporated
into an official test guideline, replacing an in vivo procedure, is limited. Recently,
OECD test guidelines based on the use of in vitro tests on skin corrosion were
adopted. In these cases in vitro methods may be used to replace the animal testing
47
currently required for hazard and risk assessment. For ocular and dermal irritation
several well-studied in vitro tests are available [106]- [108] but none of them has yet
been formally validated or adopted as an OECD guideline.
Although the in vitro tests have not been formally included in OECD test guidelines,
positive results obtained in several non-adopted in vitro tests for eye irritation are
often considered sufficient for classification and labelling. In other cases, in vitro tests
provide supplementary information, which may be used in interpretation of the
relevance for humans of data from animal studies or for a better understanding of the
toxicological mechanism of action of the substance. In addition, they are used in
tiered testing strategies and weight of evidence approaches, in order to reduce the
need to test further or to reduce the number of animals needed at a final confirmatory
stage.
4.2.1. Skin Irritation
In vitro alternatives vary from simple models such as keratinocyte cultures to more
complex organotypic cultures and reconstituted human skin models. Although a
number of in vitro alternatives have been proposed in the literature, to date there are
no validated in vitro tests for predicting acute skin irritation and for replacing the
classical Draize test [1]. Several prevalidation efforts on in vitro methods have been
published [109]-[112]. In these studies, some of the more promising in vitro methods
have been evaluated: EpiDerm™, EPISKIN™, PREDISKIN™, the non-perfused pig
ear model, and the in vitro mouse skin integrity function test (SIFT). As a follow-up
to these prevalidation efforts, an ECVAM validation study should be completed
during 2006. Two tests (EpiDerm and EPISKIN) are being validated for the purpose
of classification and labelling on the basis of skin irritation.
At present there is no OECD TG for in vitro skin irritation. Until validated in vitro
tests for skin irritation are available or guidance is provided on the use of non-adopted
tests, in vivo skin irritation tests should be performed to meet the information
requirements in the REACH proposal. However, it is foreseen that suitable in vitro
tests could be used for identifying positives (skin irritants).
48
4.2.2. Skin Corrosion
The traditional animal test used in the past to assess skin corrosion has been replaced
by alternative in vitro validated methods [71],[112]-[114], which have been included
in Annex V of the Dangerous Substances Directive [2]. These are the in vitro skin
corrosion rat skin transcutaneous electrical resistance (TER) test, which uses excised
rat skin as a test system and its electrical resistance as an endpoint; and the human
skin models such as EpiDermTM and EPISKINTM, which are reconstructed human
epidermal equivalents and use the cell viability as an endpoint. OECD TGs for in vitro
skin corrosion testing have also been adopted [31],[34]-[35].
4.2.3. Eye Irritation / Corrosion
The Draize eye irritation tests has been criticised for animal welfare reasons and
scientific reasons, including the subjectivity in allocation of the response and
differences in sensitivity to tested substances between the rabbit and man [115]-[122].
Some modifications have been proposed to reduce the pain endured by the animal and
achieve higher rates of recovery, such as the low volume eye test (LVET), where only
a tenth of the sample volume used in the standard Draize test is applied to the rabbit's
cornea and the eyes are not held shut after instillation of the test material.
In vitro methods for eye irritation comprise organotypic models such as isolated eyes
or components thereof, tissue and cell culture systems and physicochemical tests.
Several in vitro validation studies to identify alternative tests capable of replacing the
Draize rabbit eye irritation test have taken place, e.g. the EC/HO study [115], and the
COLIPA study [116]. Although no test was found capable of replacing the Draize
rabbit eye test, some of the assays showed considerable promise as screens for ocular
irritancy [120]-[121].
At present there is no OECD TG based on the use of in vitro eye irritation tests. Until
validated in vitro tests are available, it was agreed at the 64th Competent Authority
(CA) meeting in November 2002 that a substance can be considered a severe eye
irritant (R41) if it gives a positive result in one of several non validated in vitro tests,
whilst negative results require confirmation by further testing in vivo. The same
approach is expressed by Annex XI of the proposed REACH legislation. Promising
non validated in vitro eye irritation tests include: two isolated organ-based methods
(Bovine Corneal Opacity and Permeability (BCOP) test, Chicken Enucleated Eye Test
49
(CEET)); an organotypic method (Hen's Egg Test on the Chorio-Allantoic Membrane
(HET-CAM)); and two tridimensional methods (EpiOcular™ and in vitro
reconstituted Human Corneal Epithelium model (HCE)).
50
5. Expert Systems
Expert systems are computer programs that guide hazard assessment and that are able
to predict toxicity endpoints of certain chemical structures based on the available
information. These systems can be used to provide an indication of the potential toxic
activity of chemicals, for either screening or assessment purposes. Expert systems for
the prediction of chemical toxicity are built upon experimental data representing one
or more toxic manifestations of chemicals in biological systems (database), and/or
rules derived from such data (rulebase). Ideally, an expert system should be heuristic
(able to reason both with theories and with expert knowledge); transparent (able to
motivate the line of reasoning); and flexible (able to integrate new knowledge into the
existing base).
Expert systems are used to make predictions on the basis of prior information, and in
some cases mimic the ways in which groups of human experts solve problems. They
are either knowledge-based systems, which use specific rules to relate existing
physicochemical properties to specific toxicity endpoints, or automated rule induction
systems, which aim to discover patterns within data.
Some of the more commonly used expert systems to predict skin and eye irritation
and corrosion are included in the OECD's Database on Chemical Risk Assessment
Models [123]. Several reviews make comparative analyses of different expert systems
[124]-[132]. The paper by Hulzebos et al. [132] assesses the results of three expert
systems by applying the OECD Principles for (Q)SAR Validation.
The most widely used expert systems for assessing irritation and corrosion are the
BfR-DSS, DEREK for Windows (DEREKfW), MULTICASE, and TOPKAT, and
will be described in the following sections.
Within a rulebase, rules can be based on different approaches: explicit rules derived
from knowledge and human expert judgement (expert rules and knowledge-based
rules); and rules based on mathematical induction (induced rules), i.e. rules derived by
statistical correlation. Induced rules have the advantage of extending existing
knowledge without being biased toward particular mechanisms of toxic action;
however, they can be empirical relationships devoid of biological meaning. In
contrast, expert rules or knowledge-based rules are likely to have a mechanistic basis,
but they are expressions of existing knowledge rather than of new knowledge.
51
Typically, (Q)SARs are induced rules, whereas expert rules are often based on
reactive chemistry. An expert system based on entirely expert rules is referred to as a
knowledge-based system (e.g. BfR-DSS, and DEREK), whereas a system based on
induced rules is called an automated rule-induction system or a correlative system
(e.g. MultiCASE and TOPKAT) [127]. Table 28 summarises the main characteristics
of these expert systems [123].
52
Table 28. Characteristics of DSS that model skin and eye irritation.
Features
BfR – DSS
DEREK
HazardExpert
MultiCASE
TOPKAT
Model Name
Decision Support System
(DSS) for Local Lesions
DEREK
HazardExpert
MCASE, MC4PC,
TOXPC, META
TOPKAT
Developer / Supplier
(German) Federal Institute
for Risk Assessment
LHASA Ltd.
CompuDrug Ltd.
MultiCASE Inc.
Accelrys Inc.
Country
Germany
United Kingdom
Hungary
United States
United States
Qualitative
Type of Hazard
information
Provided
Qualitative
Qualitative
Qualitative
Categorical
Quantitative
Range
Point estimate
Categorical
Qualitative
Quantitative
Point estimate
QSARs
Model Approach
SARs
SARs
SARs / QSARs
Deterministic or
probabilistic
Artificial intelligence
expert system
Model Applicability
(Inside)
Organic compounds (with
purity >95%)
Organic salts
Industrial chemicals
Can not evaluate
Organic compounds
No physicochemical
properties;
Model Applicability
(Outside)
Inorganic substances;
Organic chelates; Diluted
substances; Mixtures
Model Format
(Computerised or
not )
Computerised algorithms
No searchable database
No searchable database
Defaults and
assumptions are
provided for the user
Computerised model
Computerised model
Input
Structural and empirical
formulas, melting point,
molecular weight, aqueous
solubility, octanol/water
partition coefficient.
Inorganic compounds
Evaluation /
Validation of Model
Computerised
algorithms
Computerised
algorithms
Searchable database
Conditions are defined
by the user
Searchable database
Computerised model
Model Availability
Defaults and
assumptions selected /
by default
Chemical structures
(SMILES codes, MOL
files, SDF files)
Graphical interface
Further requirements of in
vivo testing
Toxicophore (i.e.
substituent with
potential toxic effect),
pertinent endpoint and
literature references
Toxicophores and
substructures causing
modulatory effects;
Physicochemical
parameters
Predicted activity,
rational for conclusions
and documentation
Evaluated with measured
data;
Evaluated by several
companies using inhouse experimental data
Evaluated in the
literaure
Validated using multiple
10%-off and recreation
of model
Proprietary
Proprietary
Validated using 400
datasets
Non-proprietary, but not
yet publicly available
Peer-reviewed
evaluation and
validation
Proprietary
Organometallics;
Inorganics
Mixtures.
Structural information
(MOL,SDF, ISIS files).
EU risk phrases for skin
corrosion or serious
damage to eyes;
Organic molecules
Polymers;
Lipid solubility, surface
activity and pH, vapour
pressure
Output
Salts;
Metallorganics;
Pharmaceuticals;
Peptides and Protein;
Molecules >200 atoms
Computerised
algorithms
Organic compounds
Availability or low
chemical diversity of
datasets
No rank chemicals by
activity;
No 3D information
Small molecules;
QSPR/QSAR
(Optimum Prediction
Space (OPS))
53
Computerised
algorithms
Defaults and
assumptions provided
(inflexible)
Searchable database
Computerised model
Molecular structure
(SMILES strings or SD
files)
In formation on the 16
Physical, Chemical,
Fate, and Hazard
endpoints;
Rabbit Skin Irritancy;
Eye Irritancy
Evaluated with
measured data;
Validated using "leaveone out" crossvalidation
Proprietary
5.1.
BfR Decision Support System
The German Federal Institute for Risk Assessment (BfR), which is the former Federal
Institute for Health Protection of Consumers and Veterinary Medicine (BgVV), has
developed a Decision Support System (DSS) for predicting local lesions. It is used to
assess chemicals for skin/eye irritation/corrosion potential in support of regulatory
decisions (e.g. on classification and labelling). The new substances data within the
database were derived by using standardised test protocols and were evaluated by
regulators. The system also recommends when a specific alternative test should be
used and/or when to conduct an animal test.
The DSS provides categorical predictions of hazard. The predictions are based on the
physicochemical data of the query chemical, by applying physicochemical exclusion
rules, and on the structural formula of the chemical, by applying structural inclusion
rules (i.e. SARs).. The system evaluates the irritation/corrosion caused by single
contact of a chemical with skin and/or eyes. The DSS predict whether a chemical: a)
is likely to cause serious damage or corrosive effects (in which case no testing should
be necessary); b) might have corrosive effects (in which case in vitro testing is
recommended); or c) is unlikely to produce any effect or only marginal effects (in
which case, animal testing may be necessary to confirm the negative prediction). The
DSS applies the appropriate EU risk phrase according to EU regulations, providing
reliable data for legal classification and labelling.
The system can be applied to evaluate organic compounds with structural formulae
including only atoms of carbon, hydrogen, oxygen, nitrogen, sulphur, phosphorus,
silicon and halogens. Structural formulae of organic salts can also be evaluated.
However, inorganic substances, organic chelates, diluted substances, and mixtures
cannot be dealt with.
A number of papers on the BfR-DSS have been reported by the same group
[80],[133]-[139], focusing on the different aspects of the expert system, including the
development of a relational database [134], the one-dimensional representation of
molecular structures [135], the creation of the rulebase [137] (physicochemical
property limit rules [138], and structural alerts [139]), the application of self-learning
classifiers [136], and the skin irritation corrosion rules estimation tool (SICRET) [80].
54
Database. A computerised database was developed from datasets based on
toxicological test protocols relating to substance properties responsible for skin and
eye
irritation/corrosion
[134].
The
electronic
relational
database
contains
physicochemical and toxicological data for more than 1,000 proprietary chemical
substances with purity > 95% submitted according to the EU Notification of New
Substances procedure and the German Chemical Substances act [136]. Eleven of the
physicochemical parameters of this database were used to construct the DSS: specific
particular substructures within the structural formula; pH value of a saturated aqueous
solution at 20º; molecular weight (MW); aqueous solubility at 20º; hydrolytic
properties; lipid solubility at 37º; n-octanol water partition coefficient at 20º; melting
point; boiling point at atmospheric pressure; surface tension of a saturated aqueous
solution at 20º; and vapour pressure at 20º. The overall toxicological assessments of
local corrosive or irritant properties of a chemical substance are represented within the
database by risk EU phrases for skin and eye irritancy/corrosivity. The exclusion rules
were developed by using only measured values of the physicochemical data, and the
structural rules were based on the neighbourhood relations of atoms within the
structural formula without any mechanistic hypothesis.
Representation of molecular structures. The computer-based method used to
analyse and interpret structural formulae characterising chemical molecules from onedimensional representations applies recursive principles and identifies partial
isomorphic graphs [135].
Rulebase rationale. In order to develop the DSS that uses information extracted from
a database, chemical groups were categorised on the basis of their empirical formulae.
Reactive chemical substructures relevant to the formation of local lesions and rules
for the prediction of the absence of skin irritation potential were then identified. In the
evaluation of physical data, relationships were established between local
irritation/corrosion and physicochemical properties.
Two types of rules can be clearly differentiated:
Physicochemical exclusion rules. The data were examined by applying statistical
software and correlations were established between the measured physicochemical
properties and the toxicological endpoints of skin/eye irritation/corrosion. The rules
were checked against all data in the database. The rules were implemented as
exclusion rules of the kind:
55
IF (physicochemical property) A THEN not (toxic) Effect B
(Example: IF MW> 1200 g/mol THEN no effect or only marginal effects to the skin
are expected)
The examination of all the physicochemical parameters led to the definition of
physicochemical exclusion rules, derived from the combination of a single
physicochemical parameter and a particular substructure that serves to subdivide the
total of all chemical substances into different chemical classes. The process of
prediction was made reliable by covering most chemicals by more than one exclusion
rule; i.e. only the rules which covered at least four chemicals were accepted.
More recently, limit values for specific physicochemical properties that are
appropriate for identifying chemical substances that have no skin irritation or
corrosion potential have been described within clearly defined chemical groups. For
skin irritation/corrosion, the physicochemical rules have been updated by evaluating a
regulatory (confidential) database compiled with data submitted within German as
well as within EU chemicals regulations made of 1833 chemicals [138]. The
physicochemical properties include melting point, molecular weight, octanol-water
partition coefficient, surface tension, vapour pressure, aqueous solubility and lipid
solubility. Physicochemical properties limit values were defined to determine that
when the physicochemical properties for a chemical were either greater or smaller
than these limits the chemical would have no skin irritation or corrosion potential. The
application domains of these limits were constructed to correspond with the EU Rphrases for chemicals classified for skin irritation/corrosion, in order to facilitate
classification and labelling.
Structural inclusion rules. Substructures that might be responsible for severe
damaging effects caused by chemicals to skin and eyes were detected. Structural
inclusion rules were derived from correlations between specific local effects and the
submitted structural formulae in the form:
IF (substructure) A THEN Effect B
(Example: IF (phenol) THEN “corrosive”)
These rules are statistically and chemically motivated. In order to make the prediction
process more reliable, nearly a quarter of the labelled chemicals were covered by
more than one exception rule. To ensure the validity of structural rules, they were
cross- checked against the whole database.
56
More recently, structural alerts for acute skin lesions were categorised as irritation or
corrosion alerts or as a combination of corrosion/irritation alerts according to the
mechanisms of action [139]. The rules, which were evaluated for causing skin
irritation/corrosion by
using 1654 substances, were also
connected
with
physicochemical property limits and their domain of applicability was characterised.
Combination of rules. The physicochemical exclusion rules and the structural rules
should be used in combination to form a final prediction of the irritant and corrosive
properties of a chemical to skin and eyes. To achieve this, the rules were divided into
two parts, predicting: the irritant/corrosive properties to skin (R34/R35, R38, and no
statement); and the irritant/corrosive properties to eyes (R36, R41, and no statement).
The system predicts not only irritant or corrosive properties, but also whether a
chemical exhibits no chemical irritation, or whether no prediction can be given if an
insufficient result is obtained.
The system is designed to avoid conflicting rules. To avoid the conflict of structural
rules, the more-severe effect outweighs minor effects. Similarly, the exclusion rules
themselves cannot conflict because they are defined to produce a unique result.
Classifiers. Procedures for the extension of the DSS for the prediction of the local
irritation/corrosion potential of chemicals by using self-learning classifiers were
investigated [136]. Different approaches (decision trees, distance examinations in the
multidimensional space, and k-nearest neighbour methods) have been implemented,
tested and evaluated independently. A combination of all of the established extension
approaches was also developed and tested. In contrast to knowledge-based expert
systems, self-learning classifiers are automatically computed. Using these classifiers
increased significantly the predictive power of the extended DSS; however,
automatically calculated results of self-learning classifiers were not able to motivate
the toxicological relevance of the results obtained, which was not considered adequate
for regulatory hazard assessment purposes.
Evaluation. To evaluate the DSS, it was trained with more than 1,000 substances and
the resulting DSS prototype was evaluated within the current EU notification
procedure by predicting local irritant/corrosive effects of new chemicals not included
in the training set [133]. The results of this initial evaluation showed that the majority
of the tested substances that cause skin corrosion (66.7%) were predicted by the
SARs; a high percentage (20.8%) should be confirmed with a suitable in vitro method,
57
and thus a total of 87.5 % of all skin corrosive substances could be classified without
animal testing. The “no statement” rate was 8.3%. For the prediction of serious
damage to eyes the prediction rates were similar; 73.7% of substances were correctly
predicted to be causing serious eye damage. The no statement rate was 21.1% but the
false prediction rate was low (1.8%).
Recently, evaluations of the DSS physicochemical exclusion rules according to the
OECD principles for (Q)SAR validation were reported. Skin irritation and corrosion
physicochemical property limits have been evaluated by RIVM [140], whilst skin
irritation and corrosion has been assessed by ECB [141]. Assessment of the structural
inclusion rules has also been carried out by ECB [142]-[143].
SICRET. The Skin Irritation Corrosion Rules Estimation Tool (SICRET) [80] is a
tiered approach that uses physicochemical property limits, structural alerts and in vitro
tests to identify and classify chemicals that cause skin irritation or skin corrosion
without animal testing. SICRET uses physicochemical property limits to identify
chemicals with no skin corrosion or skin irritation potential; if the exclusion
physicochemical rules do not identify the chemicals with no skin corrosion or skin
irritation potential, then the structural alerts are used to identify chemicals with skin
corrosion or skin irritation potential. If a chemical does not contain structural alerts
that indicate it has skin corrosion or skin irritation potential, then in vitro skin
corrosion or skin irritation testing is evaluated. If the in vitro test is positive, then the
data are included in feedback loops for development of new structural alerts to
identify chemicals with skin corrosion or skin irritation potential. If the in vitro test is
negative then the data are included in feedback loops for development of new
physicochemical property limits to identify chemicals with no skin corrosion or skin
irritation potential.
SICRET is a tiered approach that it has been proposed to complement the current
OECD skin corrosion and skin irritation testing strategy as described in TG 404 for
acute dermal irritation/corrosion [39]. A significant difference between SICRET and
the strategy described in the OECD TG 404 is that the former only uses information
from in vitro tests, whereas the latter also uses in vivo tests information. Although the
in vitro corrosion testing proposed by SICRET has been adopted by the OECD
member states, in-vitro skin irritation tests have not been yet validated. After the
external validation of the physicochemical property limits and the structural alerts,
58
SICRET software will be coded to allow users determining whether a chemical is
likely to cause either skin irritation or skin corrosion by providing a probability.
5.2.
DEREK
The DEREK system, originally referring to the “Deductive Estimation of Risk from
Existing Knowledge”, [144]-[148] is a computer and knowledge-based expert system
developed by LHASA (Logic and Heuristics Applied to Synthetic Analysis) Ltd and
its members at the School of Chemistry, University of Leeds, UK. The software
running under Windows is called DEREK for Windows and abbreviated as
DEREKfw.
DEREK is used for the qualitative prediction of possible toxic action of compounds
on the basis of their chemical structure. The system contains a number of structural
rules that allow the identification of the structural features of a chemical that may
result in the manifestation of toxicity. These structural rules aim to encode structuretoxicity information with an emphasis on mechanisms. It includes 40 human health
endpoints, mostly related to mammalian toxicity, i.e. mutagenicity, neurotoxicity,
developmental toxicity, oncogenicity, skin sensitisation, and skin and eye irritancy
and corrosivity.
DEREK has several generic rulebases, based on sets of related chemicals rather than
on specific chemicals, and derived from mechanistic organic chemistry. They consist
of descriptions of molecular substructures (structural alerts), which have been
associated with toxic endpoints on the basis of existing knowledge.
To predict toxicity, the program checks whether any alerts within the query structure
match previously characterised toxicophores (substructure with potential toxic effect)
in the knowledge base. The reasoning engine then assesses the likelihood of a
structure being toxic, and a message indicating the nature of the toxicological hazard
is provided together with relevant literature references. There are nine levels of
confidence: certain, probable, plausible, equivocal, doubted, improbable, impossible,
open, contradicted. The reasoning model considers the following information: the
toxicological endpoint, the alerts that match toxicophores in the query structure, the
physicochemical property values calculated for the query structure, and the presence
of an exact match between the query structure and a supporting example within the
knowledge base. The presence of several toxicophores in the molecule means that
59
there is a greater risk, but the additivity of the risks is decided by the user. Structures
can be drawn directly on a computer graphics terminal or automatically retrieved from
an in-house database. The output of the model provides the identification of the
toxicophore together with the endpoint prediction and literature references.
Some limitations of the model are that physicochemical properties cannot be included,
compounds cannot be ranked according to their activity because quantitative
information is not available, metabolic aspects are not uniformly realised in the
models, and three-dimensional information is not considered. The model cannot
evaluate peptides or proteins. The internal performance of the system is biased
towards positive structural alerts.
This rule-based system is used for research purposes, and to estimate human health
hazard. Since DEREK is rule based, it relies on experimental data, and it has been
tested and evaluated by several companies using in-house experimental data.
Validation exercises have been carried out by using as input a class of specified test
structures with a known toxicophore. The criteria for the validation were whether the
program was able to detect them. However, external validation exercises have been
limited because, as the data are proprietary, the test set can contain chemicals of the
training set. Only a few external validations for skin and eye irritation and corrosion
potential in DEREK have been published [127],[132],[147].
The DEREK rulebase has nine rules for the prediction of irritancy; two rules are
specific to eye irritancy, but none is specific to skin irritancy or corrosivity. To assess
the performance of the DEREK rulebase, 300 chemical structures were assessed by
using the irritancy information published in the EU classifications for new and
existing chemicals, the ECETOC reference data banks for eye and skin irritation, and
in vivo data [127]. Of the chemicals labelled as skin and eye irritants, which
constituted distinct but overlapping subsets, 23% and 30%, respectively, were
predicted to be irritants. Of the new chemicals not labelled as skin or eye irritants, 19
out of 45 were predicted to be irritants. DEREK predicted 20-40% of the corrosive
chemicals to be irritants. Thus, probably due to the fact that physicochemical
properties are not taken into account, the sensitivity of DEREK for irritant and
corrosive effects is low.
The external performance of DEREK was assessed by predicting the irritancy of a
number of chemicals not included in the DEREK training set [147]. DEREKfW 5.01
60
contains 248 structural fragments (toxicophores) and rules that identify adverse
effects; 27 of the 248 toxicophores could be used to assess 54 randomly selected and
new notified chemicals for which experimental data were available. DEREK predicted
skin irritation for one out of 54 chemicals, but the rationale for selection of the
toxicophore was not referenced. The 53 remaining experimental results were
compared with structural fragments for corrosion/irritation derived from literature: 17
fragments from the literature predicted corrosion/irritation, while 24 positive test
results were found. The Cooper statistics yielded 14% false negatives, 45%
specificity, 33% sensitivity; and 8% false positives. The authors, Hulzebos and
Posthumus, concluded that DEREK identified irritation poorly, and when structural
irritation fragments were included, the number of false negatives decreased.
5.3.
HazardExpert
HazardExpert is a rule-based software tool developed by CompuDrug (CompuDrug
Chemistry Ltd., Hungary) [149] for predicting the toxicity of organic compounds in
humans and in animals [150]. HazardExpert predictions are based on known toxic
fragments collected from in vivo experiments and reported by the US EPA. The
knowledge base was developed based on the list of toxic fragments reported by more
than 20 experts, but it can also be used as an open architecture knowledge base, where
the user can add new fragments. The input of the rules can be performed by means of
a graphical interface, and the toxicity estimation results are also displayed by a graph.
It contains multiple toxicity endpoints combined with a capability for estimating
toxicokinetic effects, such as bioaccumulation and bioavailability, on the basis of
predicted physicochemical values. Human health hazard effects include oncogenicity,
mutagenicity, teratogenicity, membrane-irritation, sensitisation, immunotoxicity, and
neurotoxicity. The user defines the species, route of administration, dose level, and
duration of exposure, and each endpoint is predicted as one of four levels of toxicity,
taking into account the effects of bioavailability and bioaccumulation. Substructures
which may exert a positive or negative modulatory effect are identified. In addition,
several physicochemical parameters are calculated.
HazardExpert works by searching the query structure for known toxicophores held in
the literature-based knowledge base of toxic fragments. The identification of a
toxicophore leads to estimates of the toxicity endpoints by triggering rules in the
61
knowledge-base. The rules describe toxic segments and their effects on various
biological systems, and are based on the combined use of toxicological knowledge,
expert judgement, QSAR models, and fuzzy logics (which simulates the effects of
different exposure conditions). The system is also able to examine potential
metabolites of a compound.
An evaluation of HazardExpert has been published by Lewis et al. [151].
5.4.
MultiCASE
The Computer Automated Structure Evaluation (CASE) program, upgraded to
Multiple CASE (MultiCASE) [152]-[157], is an automated rule induction tool, based
on an artificial intelligence expert system that is capable of automatically identifying
molecular fragments likely to be relevant to the activity of molecules. It is also able to
assess the importance of these fragments in relation to the potency of the molecules
that contain them, without previous a priori assumptions or knowledge about the
mechanism involved in the toxic phenomenon. No pre-classification of the chemicals
is required with respect to chemical classes or nature of use.
The program is been designed to find structural entities that discriminate between
active molecules from inactive ones in non-congeneric datasets. It works by
generating all possible substrates of a defined dimension in each compound and finds
substructure characteristics of the toxic activity on the basis of the recurrence of the
residue within the series of toxic compounds; similarly, inactivating structures are
also identified. MultiCASE descriptors are automatically generated in an unbiased
way. The molecule of interest is divided into chemical fragments of 2–10 heavy
atoms, and a statistical distribution is performed to determine which fragments are
biophores or biophobes, according to whether they are associated with the biological
activity of interest, or no activity, respectively.
This knowledge-based expert system has been used for both scientific and regulatory
purposes by the Danish EPA, the US FDA, and Health Canada authorities.
The knowledge base consists of a collection of empirically derived heuristics (rules of
thumb) relating chemical structural features with the likely areas of toxic action, and
an inference engine which is able to identify the structural features present in the
target molecule and relate these to the information stored in the knowledge base. The
62
structural commonality between molecules is based on a global as well as local
rational evaluation of their substructures rather than on a common substructure.
MultiCASE is able to learn or update itself automatically from user data. It selects its
own descriptors from the practically infinite number of possible structural assemblies
and creates automatically an ad hoc library of fragments. It is able to deal with subtle
geometrical differences, with the potency of fragments to cause toxicity (synergies),
with logical analysis and hierarchical decision making.
MultiCASE can be used to generate SARs and QSARs, in a deterministic or
probabilistic way, providing qualitative and quantitative information, ranges and point
estimates. The Bayesian MultiCASE has been used in the prediction of
physicochemical and environmental fate properties, and in environmental and human
health
hazard
effects,
including
mutagenicity,
oncogenicity,
reproductive,
developmental and systemic toxicity, and skin and eye irritation [82],[84].
MultiCASE requires a large number of high quality, heterogeneous biological data for
the endpoint in question, covering a wide range of activity. The data are then
converted by the CASE system into scalar units in the range of 10-99. For the creation
and detection of structural alerts, the structure of each molecule is divided up into all
possible fragments, starting from two heavy atoms. Statistical methods are then used
to classify the fragments as biophores or biophobes. The fragments are then combined
to give a QSAR equation. The output of the program provides the predicted activity,
and a rationale for conclusions, together with documentation.
The model is able to deal mainly with small organic molecules, but the application to
salts, organometallics, polymers, and mixtures is limited.
The model has been evaluated with measured data, and validated by using a leavemany-out cross validation process for several endpoints [82], [84].
5.5.
TOPKAT
TOxicity Prediction by C(K)omputer Assisted Technology (TOPKAT, Accelrys Inc.)
[158]-[164] is a PC-based system for the in silico estimation of toxicological
properties of compounds from their molecular structure. TOPKAT allows the
prediction of physicochemical and environmental fate properties, and environmental
and human health hazard effects, including mutagenicity, developmental and systemic
63
toxicity, oncogenicity, and rabbit skin and eye irritation. In particular, the TOPKAT
models for skin [61] and eye irritation/corrosion [81] were developed by Enslein and
have been updated successively.
TOPKAT provides qualitative and quantitative information, and point estimates.
TOPKAT QSPR/QSAR models are based on descriptors for electronic structure,
connectivity, shape, and substructures. Substructure descriptors are selected from a
library of about 3,000 molecular fragments. QSARs are derived using multiple
regression equations for continuous endpoints, and two-group linear discriminant
regression functions for dichotomous endpoints. It is a modular integrated package,
each module consisting of a specific database and several chemical subclass QSAR
models, which perform automatic rule induction. The system first scans the database
for molecules with substructures overlapping with the query compound, and then it
reports the fragments found. The accuracy of the toxicity assessment is estimated by a
patented Optimum Prediction Space (OPS) technology that assesses whether or not
the query compound is within the space defined by the training set of compounds, and
the distance of the query structure from other compounds in the data base (degree of
structural coverage). The predictions are based solely on physical chemistry, not
biochemical mechanisms.
TOPKAT generates a quantitative structure-toxicity relationship model (QSTR) for
each toxicological endpoint, using two-dimensional descriptors of electrotopological
state values (E-values), symmetry and shape descriptors. These descriptors describe
the electronic, shape and symmetry attributes of the molecule. The output of the
program is a point estimate of toxicity. The reliability of the estimated toxicity can be
assessed by ascertaining whether all of the fragments of the query structure are
represented in the database compounds that were used to develop the model, whether
the submitted structure fits within, or near the periphery of the Optimum Prediction
Space (OPS) of the equation, and whether the model accurately computes the
experimental values of the data base compounds most similar to the query structure.
In addition to the toxicity point estimate, the output also provides additional
information that allows evaluation of the reliability of the prediction, including:
fragment coverage of the dataset used to calculate the model with respect to the query
compound presented for evaluation; query structure relationship to the Optimum
Prediction Space as defined by the training dataset; probability scores which
64
contribute to identification of a degree of accuracy for the prediction; and literature
references.
Even if a considerable variety of structures have been included in each model in order
to maximise the chemical space covered by the models, the limitations of the model
are inevitably defined by the contents of the dataset used for the model development.
For example, the program has some limitations in dealing with organometallics and
inorganics.
5.6.
Other Expert Systems
5.6.1. Cache
The Computer-Aided Chemistry (CAChe) program [165]-[167] is a software tool
developed by Fujitsu for creating new models to predict a number of physical,
chemical, and biological properties, and for generating a wide range of descriptors for
QSAR including quantum chemistry descriptors. CAChe includes the quantitative
prediction of physicochemical and environmental fate properties, and environmental
and human health hazard effects.
CAChe is able to deal with either organic or inorganic molecules, and predicts both
structure and properties. The input structures can be provided from X-ray data or in
the form of standard database formats. It is able to dock molecules in reaction sites,
and to perform conformational searches to identify structure-energy relationships.
CAChe can provide a visualisation of reactive sites on molecules, and UV-Visible and
IR spectra, which allow locating transition states and reaction intermediates. CAChe
also includes tools for regression analysis, and the output information for the QSAR
models includes correlation coefficients and cross validated statistics.
One limitation of the model is that it is based on calibration data and statistical
analysis results, and it is unable to deal with any organic or inorganic chemical
containing elements up to Lawrencium.
5.6.2. SuCCSES
The
Substructure-Based
Computerized
Chemical
Selection
Expert
System
(SuCCSES) [168] was developed to facilitate the US EPA review of large groups of
chemicals with similar substructures and modes of action. SuCCSES has been
65
described previously [57],[169]-[173]. SuCCSES was developed based on historical
information and expert opinions. Historical information was obtained from the US
EPA’s InterAgency Testing Committee (ITC) scoring exercises performed from 1978
to 1983. For health effects, numerous international experts were sent a questionnaire
listing more than 100 different chemical substructures and were asked to predict
(based on their toxicological expertise and knowledge of modes or mechanisms of
action) the potential for chemicals containing any of the substructures to cause acute,
chronic, mutagenic, carcinogenic, developmental, reproductive, or neurotoxic effects
or membrane irritation. Opinions from these health effects experts were converted to
codes that identified chemical substructures and indicated the potential of chemicals
containing one or more substructures to cause specific health effects.
The substructures in SuCCSES that were associated with membrane irritation were
included in [58]. SuCCSES is applied in accordance with the ITC's statutory mandate
to use SARs before the US EPA recommends chemicals for testing. SuCCSES is not
publicly available because it contains confidential business information.
5.6.3. OASIS
OASIS is a database platform developed by the Laboratory of Mathematical
Chemistry of University of Bourgas for developing three-dimensional QSAR (3DQSAR) models [174]-[176]. The system provides selection of active conformers as
3D structure representatives, including options for conformation multiplication, and
quantum-chemical optimisation. The system also assesses chemical similarity, taking
account of the flexibility of chemical structures. Calculated steric, electronic and
physicochemical descriptors of selected conformers can be calculated. The database
stores a number of (Q)SAR models on environmental fate properties and
environmental and human health effects, including the MultiCASE model for severe
skin irritation. The output of the program assesses if the (Q)SAR predictions are
inside or outside the applicability domain of the model.
5.7.
Integrated Testing Exercise
An example of application of expert systems to the prediction of skin and eye
irritation for the safety evaluation of new and existing chemicals has been reported in
an integrated testing exercise [131]. The integrated approach consisted of the use of
66
existing non-animal data and the generation of alternative data by using in silico and
in vitro methods. These methods were applied to ten substances listed on ELINCS and
associated with a complete base-set. The hazard assessment for these substances was
performed on the basis of available non-animal data, QSAR, PBBK-modelling and
additional in vitro testing was applied to the following endpoints: acute fish toxicity,
skin and eye irritation, acute oral, dermal and inhalatory toxicity, sensitisation, and
systemic toxicity after repeated dosing. The predictions were compared with the
outcome of the in vivo tests.
The methods used for this endpoint were (Q)SARs using DEREKfw (v7.0) for the
qualitative estimation, and TOPKAT (v6.1) for the quantitative prediction; and the in
vitro BECAM assay. Skin and eye irritation predictions were based on (Q)SARs and
the outcome of the BECAM test in a weight of evidence approach, assuming that any
substance not irritating to the eyes was not irritating to the skin. After the predictions
were made by the integrated method, they were made available to four expert groups,
to check the bias of predictions. Experts associated a reliability score to their own
predictions (from 0 to 5 in order of increasing reliability), and the scores were
combined to calculate and overall prediction value. However, no record on the
method of assessment of the endpoints by the experts was made.
After the results of the in vivo tests for the 10 substances were released, two
comparisons of the predicted values with the in vivo outcome were made: a direct
comparison with the in vitro and in vivo predictions and a comparison with the overall
prediction value from the expert predictions. When a predicted value (or predicted
range) would lead to the same classification according to the in vivo data, a prediction
was considered to be correct. Skin and eye irritation were correctly predicted in 70%
of the cases, with one false negative for both endpoints. However, experts reached 9
correct predictions in skin irritation and 8 correct predictions in eye irritation. For this
endpoint, a test in one animal would be sufficient to confirm the results of the QSAR
and in vitro assay (only if negative) and to prevent false negatives. This approach is
similar to that promoted by the current OECD guidelines for skin and eye irritation.
67
6. Conclusions
In this report, we have provided a detailed review of alternative approaches for
assessing skin and eye irritation/corrosion based on the use of in silico methods
(SARs, QSARs and expert systems). Literature-based models tend to be applicable to
specific groups of chemicals, whereas expert systems tend to have broader
applicability domains. Some of these systems are already well characterised (e.g. the
BfR rulebases and the DEREK system) and may therefore be suitable for regulatory
use.
In general, we conclude that the further development, validation and documentation of
in silico systems for local toxicity to the skin and eye are necessary. A considerable
effort to promote the availability of valid (Q)SARs is already being carried out by the
ECB [45]-[46]. This includes the need to present information on the characteristics of
the models in a transparent way. At present, so-called “QSAR Model Reporting
Formats” (QMRFs) are being developed for this purpose, and are being made
available from the ECB website.
The most effective approach is to integrate all appropriate information to make a
weight-of-evidence-based assessment of the chemical hazard and risk. Integrated
Testing Strategies combine all possible sources of information from (Q)SARs, expert
systems, read-across and other grouping approaches, and test methods (especially in
vitro tests). Current work within the context of one of the REACH-Implementation
Projects (RIP 3.3-2) is developing tiered approaches for the assessment of skin and
eye irritation/corrosion potential
As increasing experience is gained in the regulatory use of (Q)SARs and expert
systems, it is expected that specific research needs will be identified concerning the
development of new, tailor-made models, specifically designed for inclusion within
testing strategies.
68
7. References
[1]
Draize, J.H., Woodard, G., Calvery, H.O. (1944) Methods for the study of irritation
and toxicity of substances applied topically to the skin and mucous membranes.
Journal of Pharmacology and experimental Therapeutics 82, 377-390.
[2]
European Economic Community (1967) Council Directive 67/548/EEC of 27 June
1967 on the approximation of laws, regulations and administrative provisions relating
to the classification, packaging and labelling of dangerous substances. Official
Journal of European Communities 196, 16/08/1967. Latest updates: 9th Amendment
by Directive 1999/33/EC; Adapting to Technical Progress for the 29th time by
Directive 2004/73/EC.
[3]
European Economic Community (1993) Council Regulation (EEC) No 793/93 of 23
March 1993 on the evaluation and control of the risks of existing substances. Official
Journal of European Communities L 84, 5/4/1993.
[4]
Commission of the European Communities (2005) Commission Regulation (EC) No
642/2005 of 27 April 2005 imposing testing and information requirements on the
importers or manufacturers of certain priority substances in accordance with Council
Regulation (EEC) No 793/93 on the evaluation and control of the risks of existing
substancesText with EEA relevance. Official Journal of European Communities L
107, 28/04/2005.
[5]
The Complete REACH documentation.
http://europa.eu.int/comm/enterprise/reach/overview.htm
[6]
Commission of the European Communities (2002) White Paper on the Strategy for a
future Chemicals Policy, COM(2001)88. European Commission, Brussels, Belgium.
Available at: http://europa.eu.int/comm/environment/chemicals/whitepaper.htm
[7]
Commission of the European Communities (2003) Proposal for a Regulation of the
European Parliament and of the Council concerning the Registration, Evaluation,
Authorisation and Restriction of Chemicals (REACH), establishing a European
Chemicals Agency and amending Directive 1999/45/EC and Regulation (EC) {on
Persistent
Organic
Pollutants}.
Available
at:
http://europa.eu.int/comm/enterprise/chemicals/chempol/whitepaper/reach.htm
[8]
European Economic Community (1986) Council Directive of 24 November 1986 on
the approximation of laws, regulations and administrative provisions of the Member
69
States regarding the protection of animals used for experimental and other scientific
purposes (86/609/EEC). Official Journal of European Communities L 358,
18.12.1986, pp. 1-32.
[9]
TAPIR (2005). Three point three – A Project for the Information Requirements of
REACH. Lot 3: Scoping study on the Development of a Technical Guidance
Document on Information requirements on Intrinsic Properties of Substances (RIP
3.3-1). Final TAPIR Report RIP 3.3-1. Submitted to European Chemical Bureau,
Joint Research Centre, European Commission under Contract No. – 22553-2004-12
F1SC ISP BE.
[10]
TAPIR (2005) Final report appendix 8, Report of Information Working Group 2
“Testing” (IWG-2). REACH Implementation Project 3.3-1.
[11]
TAPIR (2005) Final report appendix 9, Report of Endpoint Working Group 3 “No
Testing Considerations”. REACH Implementation Project 3.3-1.
[12]
TAPIR Final report appendix 12 (2005) Report of Endpoint Working Group 4
“Irritation / Corrosion”. REACH Implementation Project 3.3-1.
[13]
Pedersen, F., de Bruijn, J., Munn, S., van Leeuwen, K. (2003) Report “Assessment of
additional testing needs under REACH. Effects of (Q)SARS, risk based testing and
voluntary industry initiatives”. Ispra, Italy: European Commission, Joint Research
Centre.
[14]
van der Jagt, K., Munn, S., Tørsløv, J., de Bruijn, J. (2004). Addendum to the report
“Assessment of additional testing needs under REACH. Effects of (Q)SARS, risk
based testing and voluntary industry initiatives”. EUR 21405 EN. Ispra, Italy:
European Commission, Joint Research Centre.
[15]
Walker A.P. (1990): A proposal for the harmonisation of classification schemes for
chemicals. Contract No. B6680/90/6065 for the Commision of the European
Communities, DG XI-A-2.
[16]
Organisation for Economic Co-operation and Development (1996). Final Report of
the OECD Workshop on Harmonisation of Validation and Acceptance Criteria for
Alternative Toxicological Test Methods. OECD, Paris, France. Held in Solna,
Sweden,
22-24
January
1996,
62pp.
Available
at:
http://www.oecd.org/ehs/test/background.htm
[17]
Organisation for Economic Co-operation and Development (1998) Harmonized
Integrated Hazard Classification System for Human Health and Environmental
70
Effects of Chemical Substances, as endorsed by the 28th Joint Meeting of the
Chemicals Committee and the Working Party on Chemicals, November 1998.
Available at: http://www.oecd.org/ehs/Class/HCL6.htm
[18]
Organisation for Economic Co-operation and Development (1998). Revised Proposal
for the Harmonisation of Hazard Classification Based on Eye Irritation/Corrosion.
ENV/MC/CHEM/HCL(98)5, 10pp. OECD, Paris, France.
[19]
Organisation for Economic Co-operation and Development (1998). Harmonised
Integrated Classification System for Human Health and Environmental Hazards of
Chemical Substances and Mixtures, ENV/JM/HCL(2001)1, OECD Paris, France.
Available at: http://www.oecd.org/ehs/Class/hcl/finaw.pdf
[20]
Organisation for Economic Co-operation and Development (2001). Series on Testing
and Assessment No.33: Harmonised Integrated Classification System for Human
Health and Environmental Hazards of Chemical Substances and Mixtures
ENV/JM/MONO(2001)6.
Available
at:
http://www.olis.oecd.org/olis/2001doc.nsf/LinkTo/env-jm-mono(2001)6
[21]
United Nations (2001). Report of the first session of the Sub-Committee of Experts
on the Globally Harmonized System of Classification and Labelling of Chemicals
ST/SG/AC.10/C.4/2001/4
[22]
United Nations (2003). Globally Harmonized System of Classification and Labelling
of Chemicals (GHS). Part 3. Health and Environmental Hazards, 122pp. New York,
NY,
USA,
and
Geneva,
Switzerland:
United
Nations.
Available
at:
http:// www.unece.org/trans/danger/publi/ghs
[23]
European Economic Community (1976) Council Directive 76/768/EEC of 27 July
1976 on the approximation of the laws of the Member states relating to cosmetic
products. Official Journal of European Communities L 262, 27/9/1976, p. 169.
Update: 7th amendment by Directive 2003/15/EC of the European Parliament and of
the Council of 27 February 2003; Commission Directive 2005/80/EC of 21
November 2005 amending Council Directive 76/768/EEC, concerning cosmetic
products, for the purposes of adapting Annexes II and III thereto to technical
progress. Official Journal of European Communities L 303, 22/11/2005 pp. 0032 –
0037.
[24]
ECVAM (2004) Report for establishing the timetable for phasing out animal testing
for the purpose of the cosmetics directive.
71
[25]
SCCNFP/0690/03, Final: The Scientific Committee on Cosmetic Products and NonFood Products Intended for Consumers Notes of Guidance For Testing of Cosmetic
Ingredients For Their Safety Evaluation. 5th Revision, adopted by the SCCNFP
during the plenary meeting of 20 October 2003.
[26]
http://www.hse.gov.uk/chip/phrases.htm
[27]
HSC (1997) Approved Guide to the Classification and Labelling of Substances and
Preparations Dangerous for Supply (Third Edition) Chemicals (Hazard Information
and Packaging for Supply) (Amendment) Regulations 1997 (L100), Sudbury, UK,
HSE Books.
[28]
European Economic Community (1992) Annex V-Methods for the determination of
toxicity. B.3. Acute toxicity (Dermal). Official Journal of European Communities
L383 A 35.
[29]
European Economic Community (1992) Annex V-Methods for the determination of
toxicity. B.4. Acute toxicity (Skin Irritation). Official Journal of European
Communities L383 A 35.
[30]
European Economic Community (1992) Annex V-Methods for the determination of
toxicity. B.5. Acute toxicity (Eye Irritation). Official Journal of European
Communities L383 A 35.
[31]
Commission of the European Communities (2000) Directive 2000/33/EC of 25 April
2000 Adapting to Technical Progress for the 27th time the Council Directive
67/548/EEC on the Approximation of Laws, Regulations and Administrative
Provisions Relating to the Classification, Packaging and Labelling of Dangerous
Substances. Annex 1. B.40. Skin corrosion. Official Journal of European
Communities L 136.
[32]
Organisation for Economic Co-operation and Development (1981). Guideline for
Testing of Chemicals No. 404: Acute dermal irritation/corrosion. OECD, Paris,
France, adopted on 12 May 1981, revised in 1992. Updated Guideline, adopted on 24
April 2002.
[33]
Organisation for Economic Co-operation and Development (1981). Guideline for
Testing of Chemicals No. 405: Acute Eye Irritation/Corrosion. OECD, Paris, France,
adopted on 12 May 1981, revised on 24 February 1987. Updated Guideline, adopted
on 24 April 2002.
72
[34]
Organisation for Economic Co-operation and Development (2004) Guideline for
Testing of Chemicals 430 In Vitro Skin Corrosion: Transcutaneous Electrical
Resistance Test (TER). OECD, Paris, France, adopted on 13 April 2004.
[35]
Organisation for Economic Co-operation and Development (2004) Guideline for
Testing of Chemicals 431 In Vitro Skin Corrosion: Human Skin Model Test. OECD,
Paris, France, adopted on 13 April 2004.
[36]
Organisation for Economic Co-operation and Development (2001) Brief Chronology
of the Draft Test Guidelines 404: Acute Dermal Irritation/Corrosion, 405: Acute Eye
Irritation/Corrosion, and 429: Skin Sensitisation: Local Lymph Node Assay.
[37]
Organisation for Economic Co-operation and Development (1999). Series on Testing
and Assessment No.14. Detailed Review Document on Classification Systems for Eye
Irritation/Corrosion in OECD Member Countries. ENV/JM/MONO(99)4. Available
at: http://www.olis.oecd.org/olis/1999doc.nsf/LinkTo/env-jm-mono(99)4
[38]
Organisation for Economic Co-operation and Development (1999). Series on Testing
and Assessment No.16: Detailed Review Document on Classification Systems for
Skin Irritation/Corrosion in OECD Member Countries ENV/JM/MONO(99)6.
Available at: http://www.olis.oecd.org/olis/1999doc.nsf/LinkTo/env-jm-mono(99)6
[39]
Organisation for Economic Co-operation and Development (2001) OECD
Attachment to the Test Guideline 404: A Sequential Testing Strategy for Skin
Irritation and Corrosion. OECD, Paris, France.
[40]
Organisation for Economic Co-operation and Development (2001) OECD
Attachment to the Test Guideline 405: A Sequential Testing Strategy for Eye
Irritation and Corrosion. OECD, Paris, France.
[41]
Organisation for Economic Co-operation and Development (2001) Environment
Directorate, Joint meeting of the Chemicals Committee and the working party on
chemicals, pesticides and biotechnology. Test Guidelines Programme. Revised
proposals for updated test guidelines 404 and 405: dermal and eye corrosion/
irritation studies. ENV/JM/TG (2001) 2, 112 pp, Paris, France, OECD.
[42]
Organisation for Economic Co-operation and Development (2001) OECD Guideline
for Testing of Chemicals – Draft Revised Guideline 404: Acute Dermal
Irritation/Corrosion. OECD, Paris, France.
73
[43]
Organisation for Economic Co-operation and Development (2001) OECD Guideline
for
Testing
of
Chemicals—Draft
Revised
Guideline
405:
Acute
Eye
Irritation/Corrosion. OECD, Paris, France.
[44]
European Centre for Ecotoxicology and Toxicology of Chemicals (1988) Eye
Irritation Testing. Monograph No. 11. ECETOC, Brussels, Belgium.
[45]
Organisation for Economic Co-operation and Development (2004) “The Report From
The Expert Group On (Quantitative) Structure-Activity Relationships [(Q)Sars] On
The Principles For The Validation Of (Q)SARs” 2nd Meeting of the ad hoc Expert
Group on QSARs OECD Headquarters, 20-21 September, 2004.
[46]
Worth, A.P., Bassan, A., Gallegos, A., Netzeva, T.I., Patlewicz, G., Pavan, M.,
Tsakovska, I., Vracko, M. (2005) The Characterisation of (Quantitative) StructureActivity Relationships: Preliminary Guidance. JRC Technical EUR Report. EUR
21866 EN, 95pp. Ispra, Italy: European Commission, Joint Research Centre.
[47]
Hinz R.S., Lorence. C.R., Hodson C.D., Hansch C., Hall L.L. and Guy R.H. (1991)
Percutaneous penetration of para-substituted phenols in vitro. Fundamental and
Applied Toxicology 17, 575-583.
[48]
Young, J.R., How, M.J., Walker, A.P., Worth W.M.H. (1988) Classification as
Corrosive or Irritant to Skin of Preparations Containing Acidic or Alkaline Substance
Without Testing on Animals. Toxicol. In vitro, 2, 19 - 26.
[49]
Worth, A.P. (2000) The integrated use of physicochemical and in vitro data for
predicting chemical toxicity. PhD thesis. Liverpool John Moores University,
Liverpool, UK.
[50]
Barratt, M.D. (2000) Prediction of toxicity from chemical structure. Cell Biology and
Toxicology 16, 1-13.
[51]
Barratt, M.D., Rodford, R. (2001) The computational prediction of toxicity. Current
Opinion in Chemical Biology 5, 383-388.
[52]
European Chemical Industry Council (2002) Report of the CEFIC Workshop on
Regulatory Acceptance of QSARs for Human Health and the Environment. Held in
Setubal, Portugal, 4-6 March, 2002, pp. 60.
[53]
Schultz, T.W., Cronin, M.T.D., Walker, J.D., Aptula, A.O. (2003) Quantitative
structure–activity relationships (QSARs) in toxicology: a historical perspective.
Journal of Molecular Structure (Theochem) 622, 1-22.
74
[54]
Schultz, T.W., Cronin, M.T.D., Netzeva, T.I. (2003) The present status of QSAR in
toxicology. Journal of Molecular Structure (Theochem) 622, 23-38.
[55]
Cronin, M.T.D., Dearden, J.C., Walker, J.D., Worth, A.P. (2003) Quantitative
Structure-Activity Relationships for Human Health Effects: Commonalities with
other Endpoints. Environmental Toxicology and Chemistry 22, 1829-1843.
[56]
Patlewicz, G., Rodford, R., Walker, J.D. (2003) Quantitative Structure-Activity
Relationships for Predicting Skin and Eye Irritation. Environmental Toxicology and
Chemistry 22, 1862-1869.
[57]
Cronin, M.T.D., Jaworska, J.S., Walker, J.D., Comber, M.H.I., Watts, C.D., Worth,
A.P. (2003) Use of QSARs in International Decision-Making Frameworks to Predict
Health Effects of Chemical Substances. Environmental Health Perspectives 111(10),
1391-1401.
[58]
Hulzebos, E.M., Maslankiewicz, L., Walker, J.D. (2003) Verification of literaturederived SARs for skin irritation and corrosion. QSAR & Combinatorial Science 22,
351-363.
[59]
Gerner, I., Spielmann, H., Hoefer, T., Liebsch, M., Herzler, M. (2004). Regulatory
use of (Q)SARs in Toxicological Hazard Assessment Strategies. SAR and QSAR in
Environmental Research 15(5-6), 359-366.
[60]
Walker, J.D., Gerner, I., Hulzebos, E., Schlegel, K. (2004) (Q)SARs for Predicting
Skin Irritation and Corrosion: Mechanisms, Transparency and Applicability of
Predictions. QSAR & Combinatorial Science 23, 721-725.
[61]
Enslein, K., Borgstedt, H.H., Blake, B.W., Hart, J.B. (1987) Prediction of rabbit skin
irritation severity by structure activity relationships. In vitro Toxicology 1, 129-147.
[62]
Berner, B., Wilson, D.R., Guy, R.H., Mazzenga, G.C., Clarke, F.H., Maibach, H.I.
(1988) The relationship of pKa and acute skin irritation in man. Pharmaceutical
Research 5, 660-663.
[63]
Berner, B., Wilson, D.R., Steffens, R.J., Mazzenga, G.C., Hinz, R., Guy, R.H.,
Maibach, H.I. (1990) The relationship between pKa and acute skin irritation for a
series of basic penetrants in man. Fundamental and Applied Pharmacology 15, 760–
766.
[64]
Berner, B., Wilson, D.R., Guy, R.H., Mazzenga, G.C., Clarke, F.H., Maibach, H.I.
(1989/1990) Relationship of pKa and acute skin irritation in humans. Journal of
Toxicology-Cutaneous and Ocular Toxicology 9, 481–492.
75
[65]
Nangia, A., Anderson, P.H., Berner, B., Maibach, H.I. (1996). High dissociation
constants (pKa) of baasic permeants are associated with in vivo skin irritation in man.
Contact Dermatitis i34, 237-242.
[66]
Hayashi, M., Nakamura, Y., Higashi, K., Kato, H., Kishida, F., Kaneko, H. (1999) A
Quantitative Structure-Activity Relationship Study of the Skin Irritation Potential of
Phenols. Toxicology in vitro 13, 915-922.
[67]
Smith, J.S., Macina, O.T., Sussman, N.B., Luster, M.I., Karol, M.H. (2000) A Robust
Structure-Activity Relationship (SAR) Model for Esters that Cause Skin Irritation in
Humans. Toxicological Sciences 55, 215-222.
[68]
Smith, J.S., Macina, O.T., Sussman, N.B., Karol, M.H. Maibach, H.I. (2000)
Experimental validation of a structure-activity relationship model of skin irritation by
esters. Quantitative Structure-Activity Relationships 19, 467–474.
[69]
Kodithala, K., Hopfinger, A.J., Thompson, E.D., Robinson, M.K. (2002) Prediction
of Skin Irritation from Organic Chemicals Using Membrane-Interaction QSAR
Analysis. Toxicological Sciences 66, 336-346.
[70]
Barratt, M.D. (1995) Quantitative structure activity relationships for skin corrosivity
of organic acids, bases and phenols. Toxicology Letters 75, 169-176.
[71]
Botham, P., Chamberlain, M., Barratt, M. D., Curren, R.D., Esdaile, D.J., Gardner,
J.R., Gordon, V.C., Hildebrand, B., Lewis, R.D., Liebsch, M., Logemann, P.,
Osborne, R., Ponec, M., Régnier, J.-F., Steiling, W., Walker, A.P., Balls, M. (1995)
A Prevalidation Study on In vitro Skin Corrosivity Testing. The Report and
Recommendations of ECVAM Workshop 6. ATLA – Alternatives to Laboratory
Animals 23, 219-255.
[72]
Barratt, M.D. (1996) Quantitative Structure-Activity Relationships for Skin Irritation
and Corrosivity of Neutral and Electrophilic Organic Chemicals. Toxicology in vitro
10, 247-256.
[73]
Whittle, E., Barratt, M.D., Carter, J.A., Basketter, D.A., Chamberlain, M. (1996) Skin
Corrosivity Potential of Fatty Acids: In vitro Rat and Human Skin Testing and QSAR
Studies. Toxicology in vitro 10, 95-100.
[74]
Barratt, M.D. (1996) Quantitative Structure-Activity Relationships for Skin
Corrosivity of Organic Acids, Bases and Phenols: Principal Components and Neural
Network Analysis of Extended Sets. Toxicology in vitro 10, 85-94.
76
[75]
Barratt, M.D., Dixit, M.B., Jones, P.A. (1996) The Use of In vitro Cytotoxicity
Measurements in QSAR Methods for the Prediction of the Skin Corrosivity Potential
of Acids. Toxicology in vitro 10, 283-290.
[76]
Worth, A.P., Cronin, M.T.D. (2001) The Use of Bootstrap Resampling to Assess the
Uncertainty of Cooper Statistics. ATLA – Alternatives to Laboratory Animals 29,
447-459.
[77]
Worth, A.P., Cronin, M.T.D. (2001) The use of bootstrap resampling to assess the
variability of Draize tissue scores. ATLA – Alternatives to Laboratory Animals 29,
557-573.
[78]
Worth, A.P., Cronin, M.T.D. (2001). The use of pH measurements to predict the
potential of chemicals to cause acute dermal and ocular toxicity. Toxicology 169,
119-131.
[79]
Worth, A.P., Fentem, J., Balls, M., Botham, P.A., Curren, R.D., Earl, L.K., Esdaile,
D.J., Liebsch, M. (1998) An Evaluation of the Proposed OECD Testing Strategy for
Skin Corrosion. ATLA – Alternatives to Laboratory Animals 26, 709-720.
[80]
Walker, J.D., Gerner, I., Hulzebos, E., Schlegel, K. (2005) The Skin Irritation
Corrosion Rules Estimation Tool (SICRET). QSAR & Combinatorial Science 24,
378-384.
[81]
Enslein, K., Blake, B.W., Tuzzeo, T.M., Borgstedt, H.H., Hart, J.B., Salem, H. (1988)
Estimation of rabbit eye irritation scores by structure activity equations. In vitro
Toxicology 2, 1-14.
[82]
Klopman, G., Ptchelintsev, D., Frierson, M., Pennisi, S., Renskers, K., Dickens, M.
(1993) Multiple computer automated structure evaluation methodology as an
alternative to in vitro eye irritation testing. ATLA – Alternatives to Laboratory
Animals 21, 14-27.
[83]
Klopman, G. (1998) The MultiCASE Program II. Baseline Activity Identification
Algorithm (BAIA). Journal of Chemical Information and Computer Sciences 38, 7881.
[84]
Rosenkranz, H.S., Zhang, Y.P., Klopman, G. (1998) The Development and
Characterisation of a Structure-activity Relationship Model of the Draize Eye
Irritation Test. ATLA – Alternatives to Laboratory Animals 26, 779-809.
[85]
Sugai, S, Murata, K, Kitagaki, T, Tomita, I. (1990) Studies on eye irritation caused by
chemicals in rabbits—1. A quantitative structure-activity relationships approach to
77
primary eye irritation of chemicals in rabbits. The Journal of Toxicological Sciences
15, 245-262.
[86]
Sugai, S., Murata, K., Kitagaki, T., Tomita, I. (1991) Studies on eye irritation caused
by chemicals in rabbits—II. Structure-activity relationships and in vitro approach to
primary eye irritation of salicylates in rabbits. The Journal of Toxicological Sciences
16, 111-130.
[87]
Lipnick, R.L., Bentley, J.L., Bentley, D.L. (1992) QSAR analysis of ocular toxicity.
Abstracts, The 5th International Workshop on QSAR in Environmental Toxicology,
Duluth, MN, USA.
[88]
Cronin, M.T.D., Basketter, D.A., York, M. (1994) A quantitative structure- activity
relationship (QSAR) investigation of a Draize eye irritation database. Toxicology in
vitro 8, 21-28.
[89]
Cronin, M.T.D. (1996) The use of cluster significance analysis to identify asymmetric
QSAR datasets in toxicology. An example with eye irritation data. SAR and QSAR in
Environmental Research 5, 167-175.
[90]
Worth, A.P., Cronin, M.T.D. (2000) Embedded cluster modelling: a novel
quantitative structure-activity relationhip for generating elliptic models of biological
activity. In, Progress in the Reduction, Refinement and Replacement of Animal
Experimentation. Balls, M., van Zeller, A.-M., Halder, M.E. (Eds.) Elsevier Science,
Amsterdam, The Netherlands, pp. 479-491.
[91]
Worth, A.P., Cronin, M.T.D. (2003) The use of discriminant analysis, logistic
regression and classification tree analysis in the development of classification models
for human health effects. Journal of Molecular Structure (Theochem) 622, 97-111.
[92]
Barratt, M.D. (1995) A quantitative structure-activity relationship for the eye
irritation potential neutral organic chemicals. Toxicology Letters 80, 69-74.
[93]
Barratt, M.D. (1997) QSARS for the Eye Irritation Potential of Neutral Organic
Chemicals. Toxicology in vitro 11, 1-8.
[94]
Abraham, M.H., Kumarsingh, R., Cometto-Muñiz, J.E., Cain, W.S. (1998) A
Quantitative Structure-Activity Relationship (QSAR) for a Draize Eye Irritation
Database. Toxicology in vitro 12, 201-207.
[95]
Abraham, M.H., Kumarsingh, R., Cometto-Muñiz, J.E., Cain, W.S. (1998) Draize
Eye Scores and Eye Irritation Thresholds in Man Combined into one Quantitative
Structure-Activity Relationship. Toxicology in vitro 12, 403-408.
78
[96]
Abraham, M.H., Kumarsingh, R., Cometto-Muñiz, J.E., Cain, W.S. (1998) Draize eye
scores and eye irritation thresholds in man can be combined into one QSAR.
Olfaction and Taste XII 855, 652-656.
[97]
Abraham, M.H, Hassanisadi, M., Jalali-Heravi, M., Ghafourian, T., Cain, W.S.,
Cometto-Muñiz, J.E. (2003). Draize Rabbit Eye Test Compatibility with Eye
Irritation Thresholds in Humans: A Quantitative Structure-Activity Relationship
Analysis. Toxicological Sciences 76, 384-391.
[98]
Kulkarni, A.S., Hopfinger, A.J. (1999). Membrane-interaction QSAR analysis:
application to the estimation of eye irritation by organic compounds. Pharmaceutical
Research 16, 1245-1253.
[99]
Kulkarni, A., Hopfinger, A.J., Osborne, R., Bruner, L.H., Thompson, E.D. (2001)
Prediction of Eye Irritation from Organic Chemicals Using Membrane-Interaction
QSAR Analysis. Toxicological Sciences 59, 335-345.
[100]
Li, Y., Liu, J., Pan, D., Hopfinger, A.J., (2005) A study of the relationship between
cornea permeability and eye irritation using membrane-interaction QSAR analysis.
Toxicological Sciences 88, 434-446.
[101]
Patel, H.C., Duca, J.S., Hopfinger, A.J., Glendening, C.D., Thompson, E.D. (1999)
Quantitative component analysis of mixtures for risk assessment: Application to eye
irritation. Chemical Research in Toxicology 12, 1050–1056.
[102]
Patlewicz, G.Y., Rodford, R.A., Ellis, G., BArratt, M.D. (2000) A QSAR Model for
the Eye Irritation of Cationic Surfactants. Toxicology in vitro 14, 79-84.
[103]
Hayashi, T., Itagaki, H., Fukuda, T., Tamura, U., Kato, S. (1993) Quantitative
structure-activity relationship of surfactants on eye irritation predicted by hemoglobin
denaturation. Alternatives to Animal Test Experiments 2, 49-55.
[104]
Patlewicz GY, El-Deredy W. 1999. Identification of molecular properties of cationic
surfactants causing eye irritation: A QSAR Bayesian neural network model.
Proceedings, Predictive Toxicology of Chemicals: Experiences and Impact of AI
Tools AAAI Spring Symposium, Menlo Park, CA, USA, March 22–24, pp 123-127.
[105]
European Parliament resolution on Council Directive 86/609/EEC on the protection
of animals used for experimental and other scientific purposes (2001/2259(INI)).
Official Journal of European Communities C 27 E, 30/01/2004, pp. 0151-0155.
[106]
Balls, M., De Klerck, W., Baker, F., van Beek, M., Bouillon, C., Bruner, L.,
Carstensen, J., Chamberlain, M., Cottin, M., Curren, R., Dupuis, J., Fairweather, F.,
79
Faure, U., Fentem, J., Fisher, C., Galli, C., Kemper, F., Knaap, A., Langley, G.,
Loprieno, G., Loprieno, N., Pape, W., Pechovitch, G., Spielmann, H., Ungar, K.,
White, I., Zuang, V. (1995) Development and Validation of Non-Animal Tests and
Testing Strategies: the Identification of a Coordinated Response to the Challenge and
the Opportunity Presented by the Sixth Amendment to the Cosmetics Directive
(76/768/EEC). ATLA – Alternatives To Laboratory Animals 23, 398-409.
[107]
Barratt, M.D., Castell, J.V., Chamberlain, M., Combes, R.D., Dearden, J.C., Fentem,
J.H., Gerner, I., Giuliani, A., Gray, T.J.B., Livingstone, D.J., Provan, W.M., Rutten,
F.A.J.J.L., Verhaar, H.J.M., Zbinden, P. (1995) The Integrated Use of Alternative
Approaches for Predicting Toxic Hazard. Alternatives To Laboratory Animals 23,
410-429.
[108]
Liebsch, M., Spielmann, H. (2002) Currently available in vitro methods used in
regulatory toxicology. Toxicology Letters 127, 127-134.
[109]
van de Sandt, J., Roguet, R., Cohen, C., Esdaile, D., Ponec, M., Corsini, E., Barker,
C., Fusenig, N., Liebsch, M., Benford, D., de Brugerolle de Fraissinette, A., Fartasch,
M. (1999) The Use of Human Keratinocytes and Human Skin Models for Predicting
Skin Irritation. The Report and Recommendations of ECVAM Workshop 38. ATLA –
Alternatives to Laboratory Animals 27, 723-743.
[110]
Harbell, J.W., Curren, R.D. (2001) In vitro methods for the prediction of ocular and
dermal toxicity. In, Handbook of Toxicology, 2nd Ed. M.J. Derelanko and M.A.
Hollinger ED., CRC Press, Boca Raton, pp. 835-866.
[111]
Fentem, J., Botham, P.A. (2004) Update on the Validation and Regulatory
Acceptance of Alternative Tests for Skin Corrosion and Irritation. ATLA –
Alternatives to Laboratory Animals 32 (Supplement 1), 29-40.
[112]
Zuang, V., Alonso, M.A., Botham, P.A., Eskes, C., Fentem, J., Liebsch, M., van de
Sandt, J.J.M. (2005) ECVAM Report: Subgroup 2. Skin Irritation/Corrosion. ATLA –
Alternatives to Laboratory Animals 33 (Supplement 1), 35-46.
[113]
Barratt M.D., Brantom P.G., Fentem J.H., Gerner I., Walker, J.D., Worth A.P. (1998)
The ECVAM International validation study on In vitro Tests for Skin Corrosivity. I.
Selection and distribution of the test chemicals. Toxicology in vitro 12, 471-482.
[114]
Fentem J.H., Archer G.E.B., Balls M., Botham P.A., Curren R.D., Earl L.K., Esdaile
D.J., Holzhütter H.G., Liebsch M. (1998) The ECVAM international validation study
on in vitro tests for skin corrosivity. II. Results and evaluation by the Management
Team. Toxicology in vitro 12, 483-524.
80
[115]
Balls, M., Botham,PA., Bruner, LH., Spielmann,H. (1995) The EC/HO international
validation study on alternatives to the Draize eye irritation test. Toxicology in vitro 9,
871-929.
[116]
Brantom, P.G., Bruner, L.H., Chamberlain, M., Desilva, O., Dupuis, J., Earl, L.K.,
Lovell, D.P., Pape, W.J.W., Uttley, M., Bagley, D.M., Baker, F.W., Brachter, M.,
Courtellemont, P., Declercq, L., Freeman, S., Steiling, W., Walker, A.P., Carr, G.J.,
Dami, N., Thomas, G., Harbell, J., Jones, P.A., Pfannenbecker, U., Southee, J.A.,
Tcheng, M., Argembeaux, H., Castelli, D., Clothier, R., Esdaile, D.J., Itigaki, H.,
Jung, K., Kasai, Y., Kojima, H., Kristen, U., Larnicol, M., Lewis, R.W., Marenus, K.,
Moreno, O., Peterson, A., Rasmussen, E.S., Robles, C., Stern, M. (1997) A summary
report of the COLIPA international validation study on alternatives to the Draize
rabbit eye irritation test. Toxicology in vitro 11(N1-2), 141-179.
[117]
Bagley, D.M., Bruner, L.H., de Silva, O, Cottin, M., O'Brien, K.A.F., Uttley, M.,
Walker, A.P. (1992) An evaluation of five potential alternatives in vitro to the rabbit
eye irritation test in vivo. Toxicology in vitro 6, 136- 149.
[118]
Sina, J.F., Galer, D.M., Sussman, R.G., Gautheron, P.D., Sargent, E.V., Leong, B.,
Shah, P.V., Curren, R.D., Miller K. (1995) A collaborative evaluation of seven
alternatives to the Draize eye irritation test using pharmaceutical intermediates.
Fundamental and applied toxicology 26(1), 20-31.
[119]
Bradlaw, J., Gupta, K., Green, S., Hill, R., Wilcox, N. (1997) Practical application of
non –whole animal alternatives: summary IRAG workshop on eye irritation. Food
and chemical toxicology 35, 175-178.
[120]
Balls, M., Berg, N., Bruner, L.H., Curren, R., de Silva, O., Earl, L.K., Esdaile, D.J.,
Fentem, J.H., Liebsch, M., Ohno, Y., Prinsen, M.K., Spielmann, H., Worth, A.P.
(1999) Eye irritation testing: the way forward. The report and recommendations of
ECVAM workshop 34. Alternatives to laboratory animals. ATLA – Alternatives to
Laboratory Animals 27, 53-77.
[121]
Balls, M., Brantom, P.G., Cassidy, S., Esdaile, D., Fentem, J., Liebsch, M.,
McPherson, J., Pfannenbecker, U., Prinsen, M. (1999) Preliminary evaluation of the
application of reference standards in the prevalidation and validation of in vitro test
for eye irritation. In: D.G. Clark, S.G. Lisansky, R. Macmillan: Alternatives to animal
testing II, proceedings, Brussels. CPL Press, Berkshire, 201-204.
[122]
Eskes, C., Bessou, S., Bruner, L., Curren, R., Harbell, J., Jones, P., Kreiling, R.,
Liebsch, M., McNamee, P., Pape, W., Prinsen, M.K., Seidle, T., Vanparys, P., Worth,
81
A., Zuang, V. (2005) ECVAM Report: Subgroup 3. Eye Irritation. ATLA –
Alternatives to Laboratory Animals 33 (Supplement 1), 47-81.
[123]
Organisation for Economic Co-operation and Development (2002) OECD’s Database
on Chemical Risk Assessment Models. Paris: Organisation for Economic Cooperation
and
Development.
OECD,
Paris,
France.
Available
at:
http://webdomino1.oecd.org/comnet/env/models.nsf
[124]
Cronin, M.T.D., Dearden, J.C. (1995) QSAR in toxicology. 4. Prediction of nonlethal mammalian toxicological endpoints and expert systems for toxicity prediction.
Quantitative Structure-Activity Relationships 14, 518-523.
[125]
Combes, R.D.; Judson, P. (1995) The use of artificial intelligent expert systems for
the prediction of toxicity. Pesticide Science 45, 179-194.
[126]
Benfenati, E., Gini, G. (1997) Computational predictive programs (expert systems) in
toxicology. Toxicology 119, 213-225.
[127]
Dearden, J.C., Barratt, M.D., Benigni, R., Bristol, D.W., Combes, R.D., Cronin,
M.T.D., Judson, P.N., Payne, M.P., Richard, A.M., Tichy, M., Worth, A.P., Yourick,
J.J. (1997) The Development and Validation of Expert Systems for Predicting
Toxicity. The Report and Recommendations of an ECVAM/ECB Workshop
(ECVAM Workshop 24). ATLA – Alternatives to Laboratory Animals 25, 223-252.
[128]
Darvas, F.; Papp, Á.; Allardyce, A.; Benfenati, E.; Gini, G.; Tichy, M.; Sobb, N.;
Citti, A. (1999) Overview of Different Artificial Intelligence Approaches Combined
with a Deductive Logic-based Expert System for Predicting Chemical Toxicity. Oral
presentation at the "Predictive Toxicology of Chemicals: Experiences and Impact of
AI Tools Symposium" within the American Association for Artificial Intelligence
Spring Symposium Series, Stanford University, CA, pp. 94-99.
[129]
Cronin, M.T.D. (2000) Computational methods for the prediction of drug toxicity.
Current Opinion in Drug Discovery & Development 3, 292-297.
[130]
Bhogal, N.; Grindon, C.; Combes, R.; Balls, M. (2005) Toxicity testing: creating a
revolution based on new technologies. Trends in Biotechnology 23, 299-307.
[131]
Gubbels-van Hal, W.M., Blaauboer, B.J., Barentsen, H.M., Hoitink, M.A., Meerts,
I.A., van der Hoeven, J.C. (2005) An alternative approach for the safety evaluation of
new and existing chemicals, an exercise in integrated testing. Regulatory Toxicology
and Pharmacology 42, 284-295.
82
[132]
Hulzebos, E., Sijm, D., Traas, T., Posthumus, R., Maslankiewicz, L. (2005) Validity
and validation of expert (Q)SAR systems. SAR and QSAR in Environmental Research
16, 385-401.
[133]
Zinke, S., Gerner, I., Graetschel, G., Schlede, E. (1999) Local Irritation/Corrosion
Testing Strategies: Development of a Decision Support System for the Introduction
of Alternative Methods. ATLA – Alternatives to Laboratory Animals 28, 29-40.
[134]
Gerner, I., Graetschel, G., Kahl, J., Schlede, E. (2000) Development of a Decision
Support System for the Introduction of Alternative Methods into Local
Irritation/Corrosivity Testing Strategies. Development of a Relational Database.
ATLA – Alternatives to Laboratory Animals 28, 11-28.
[135]
Zinke, S., Gerner, I. (2000) A computer-based structure activity relationship method
for predicting the toxic effects of organic chemicals from one-dimensional
representations of their molecular structures. ATLA – Alternatives to Laboratory
Animals 28, 609-620.
[136]
Zinke, S., Gerner, I. (2000) Local irritation/corrosion testing strategies: Extending a
decision support system by applying self-learning classifiers. ATLA – Alternatives to
Laboratory Animals 28, 651-663.
[137]
Gerner, I., Zinke, S., Graetschel, G., Schlede, E. (2000) Development of a Decision
Support System for the Introduction of Alternative Methods into Local
Irritation/Corrosivity Testing Strategies. Creation of Fundamental Rules for a
Decision Support System. ATLA – Alternatives to Laboratory Animals 28, 665-698.
[138]
Gerner, I., Schlegel, K., Walker, J.D., Hulzebos, E. (2004) Use of Physicochemical
Property Limits to Develop Rules for Identifying Chemical Substances with no Skin
Irritation or Corrosion Potential. QSAR & Combinatorial Science 23, 726-733.
[139]
Hulzebos, E., Walker, J.D., Gerner, I., Schlegel, K. (2005) Use of structural alerts to
develop rules for identifying chemical substances with skin irritation or skin
corrosion potential. QSAR & Combinatorial Science 24, 332-342.
[140]
Rorije, E., Hulzebos, E. (2005) Evaluation of (Q)SARs for the prediction of Skin
Irritation/Corrosion Potential. Physico-chemical exclusion rules. Report submitted by
the National Institute of Public Health and Environment (RIVM) to the European
Chemicals Bureau (ECB). September 2005.
83
[141]
Tsakovska, I., Netzeva, T., Worth, A. (2006) Evaluation of (Q)SARs for the
prediction of Eye Irritation/Corrosion Potential. Structural Inclusion rules. EUR
technical report. Ispra, Italy: European Commission, Joint Research Centre.
[142]
Tsakovska, I., Gallegos, A., Netzeva, T., Pavan, M., Worth, A. (2006) Evaluation of
structure-activity relationships for the prediction of eye irritation and corrosion. SAR
and QSAR in Environmental Research. Accepted.
[143]
Gallegos Saliner, A., Tsakovska, I., Pavan, M., Patlewicz, G., Worth, A. (2006)
Evaluation of SARs for the prediction of skin irritation/corrosion potential - structural
inclusion rules in the BfR decision support system. SAR and QSAR in Environmental
Research. Accepted.
[144]
LHASA Ltd. (2005). DEREK for Windows. Leeds, U.K. Available at:
http://www.chem.leeds.ac.uk/luk/derek
[145]
Sanderson, D.M., Earnshaw, C.G. (1991) Computer Prediction of Possible Toxic
Action from Chemical Structure; The DEREK System. Human & Experimental
Toxicology 10, 261-273.
[146]
Greene, N., Judson, P.N., Langowski, J.J., Marchant, C.A. (1999) Knowledge-based
expert systems for toxicity and metabolism prediction: DEREK, StAR, METEOR.
SAR and QSAR in Environmental Research 10, 299-314.
[147]
Hulzebos, E.M., Posthumus, R. (2003) (Q)SARs: Gatekeepers against risk on
chemicals? SAR and QSAR in Environmental Research 14, 285-316.
[148]
Combes, R.D.; Rodford, R. (2004) The Use of Expert Systems for Toxicity Prediction
- Illustrated With Reference to the DEREK Program. In, Predicting Toxicity and
Fate. Cronin, M.T.D.; Livingstone, D.J. (Eds.) CRC Press, pp. 193-204.
[149]
CompuDrug
Chemistry
Ltd.
Budapest,
Hungary.
Available
at:
http://www.compudrug.com
[150]
Smithing, M.P., Darvas, F. (1992) HazardExpert: An Expert System for Predicting
Chemical Toxicity. In, Food Safety Assessment. Finlay, S.F., Robinson, S.F.,
Armstrong, D.J. (Eds.). American Chemical Society, Washington, DC, pp. 191-200.
[151]
Lewis, D.F., Bird, M.G., Jacobs, M.N. (2002) Human carcinogens: an evaluation
study via the COMPACT and HazardExpert procedures. Human & Experimental
Toxicology 21, 115-122.
84
[152]
MultiCASE
Inc.
MultiCASE.
Cleveland,
OH,
USA.
Available
at:
http://www.multicase.com
[153]
Klopman, G. (1984) Artificial intelligence approach to structure-activity studies.
Computer automated structure evaluation of biological activity of organic molecules.
Journal of the American Chemical Society 106, 7315-7321.
[154]
Klopman, G. (1992) MULTICASE.1. A Hierarchical Computer Automated Structure
Evaluation Program. Quantitative Structure-Activity Relationships 11, 176-184.
[155]
Klopman, G., Ptchelinstev, D., Frierson, M., Pennisi, S., Renskers, K., Dickens, M.
(1993) Multiple computer automated structure evaluation methodology as an
alternative to in vivo eye irritation testing. ATLA – Alternatives to Laboratory
Animals 21, 14-27.
[156]
Rosenkranz, H.S., Cunningham, A.R., Zhang, Y.P., Klopman, G. (1999) Applications
of the case/multicase SAR method to environmental and public health situations. SAR
and QSAR in Environmental Research 10, 263-276.
[157]
Rosenkranz, H.S., Cunningham, A.R., Zhang, Y.P., Claycamp, H.G., Macina, O.T.,
Sussman, N.B., Grant, S.G., Klopman, G. (1999) Development, characterization and
application of predictive-toxicology models. SAR and QSAR in Environmental
Research 10, 277-298.
[158]
Accelrys Inc. (2005) TOPKAT. Cambridge, UK. (Developed by: Health Design Inc.
&
Oxford
Molecular
Group,
Inc.)
Available
at:
http://www.accelrys.com/products/topkat
[159]
Enslein, K. (1988) An overview of structure-activity relationships as an alternative to
testing in animals for carcinogenicity, mutagenicity, dermal and eye irritation, and
acute oral toxicity. Toxicology and Industrial Health 4, 479-498.
[160]
Gombar, V.K., and Enslein, K. (1990) Quantitative structure-activity relationship
(QSAR) studies using electronic descriptors calculated from topological and
molecular orbital (MO) methods. Quantitative Structure-Activity Relationships 8,
321-325.
[161]
Enslein, K. (1993) The future of toxicity prediction with QSAR. In vitro Toxicology
6, 163-169.
[162]
Gombar, V.K. (1997) Quantitative structure-activity relationships in toxicology: from
fundamentals to applications. In, Advances In Molecular Toxicology, pp. 125-139.
85
[163]
Enslein, K. (1997) QSTR applications in acute, chronic, and developmental toxicity,
and carcinogenicity. In, Adv. In Molecular Toxicology, 141-164.
[164]
Gombar, V.K., Enslein, K., Blake. B.W., Reid, D.A. (1997) Quantitative StructureActivity Relationships In Predicting Chemical Toxicity, When is the Prediction
Reliable? In The proceedings of the Workshop on QSARs in Environmental Sciences
27, 399-411.
[165]
http://www.CACheSoftware.com/cache/
[166]
Check, C.E.; Gilbert, T.M. (2000) CAChe: Computer-Aided Chemistry for Macintosh
and Windows. Version 3.2 for Windows; Version 4.1 for Macintosh. Journal of the
American Chemical Society (Computer Software Review) 122, 9566-9567.
[167]
Rajamani, R. (2004) CAChe 6.1 WorkSystem Pro plus Active Site Add-On. CAChe
Group, Fujitsu America, Inc., 15244 Greenbriar Parkway, Beaverton, OR 97006.
http://www.cachesoftware.com/. Contact CAChe group for pricing information.
Journal of the American Chemical Society (Computer Software Review) 126, 88568857.
[168]
Walker, J.D., Gray, D.A. (2001) The Substructure-based Computerized Chemical
Selection Expert System (SuCCSES): Providing Chemical Right-to-Know (CRTK)
Information on Potential Actions of Structurally-Related Chemical Classes on the
Environment and Human Health. In, Quantitative Structure Activity Relationships
(QSARs) for Predicting Effects of Chemicals on Environmental-Human Health
Interactions. Walker, J. D. (Ed). SETAC Press, Pensacola, FL.
[169]
Walker, J.D., Brink, R.H. (1989) New cost-effective, computerized approaches to
selecting chemicals for priority testing consideration. In, Aquatic Toxicology and
Environmental Fate, Vol 11, ASTM STP 1007. Suter, G.W.II, Lewis, M.A. (Eds).
Philadelphia American Society for Testing and Materials, pp. 507-536.
[170]
Walker, J.D. (1991) Chemical selection by the TSCA Interagency Testing
Committee: use of computerized substructure searching to identify chemical groups
for health effects, chemical fate and ecological effects testing. Science of the Total
Environment 109/110, 691-700.
[171]
Walker, J.D. (1993) The TSCA Interagency Testing Committee, 1977 to 1992:
creation, structure, functions and contributions. In, Environmental Toxicology and
Risk Assessment, Vol 2. Gorsuch, J.W., Dwyer, F.J., Ingersoll, C.G., La Pointe, T.W.,
(Eds.). ASTM STP 1216. American Society for Testing and Materials, Philadelphia,
PA, pp. 451-509.
86
[172]
Walker, J.D. (1993) The TSCA Interagency Testing Committee’s approaches to
screening and scoring chemicals and chemical groups: 1977-1983. In, Access and
Use of Information Resources in Assessing Health Risks from Chemical Exposure.
Lu, P.Y. (Ed). Oak Ridge National Laboratories, Oak Ridge, TN, pp. 77-93.
[173]
Walker, J.D. (1995) Estimation methods used by the TSCA Interagency Testing
Committee to prioritize chemicals for testing: exposure and biological effects scoring
and structure activity relationships. Toxicological Modelling 1, 123-141.
[174]
Laboratory of Mathematical Chemistry. (2004) University of Bourgas, Bourgas,
Bulgaria. Available at: http://www.oasis-lmc.org
[175]
Mekenyan, O.G., Karabunarliev, S., Bonchev, D. (1990) The microcomputer OASIS
system for predicting of the biological activity of chemical compounds. Computers &
Chemistry 14, 193-200.
[176]
Mekenyan, O.G., Karabunarliev, S., Ivanov, J., Dimitrov, D. (1994) A new
development of the OASIS computer system for modelling of molecular properties.
Computers & Chemistry 18, 173-187.
87
European Commission
EUR 22320 EN – DG Joint Research Centre,
Institute for Health and Consumer Protection
Luxembourg: Office for Official Publications of the European Communities
2006 – 87 pp. – 21x 29.7 cm
Scientific and Technical Research series
Abstract
This report reviews the state-of-the-art of in silico and in vitro methods for assessing
dermal and ocular irritation and corrosion. Following a general introduction, the
current EU legislation for the classification and labelling of chemicals causing
irritation and corrosivity is summarised. Then currently available non-animal
approaches are reviewed. The main alternative approaches to assess acute local toxic
effects are: a) in silico approaches, including SARs, QSARs and expert systems
integrating multiple approaches; and b) in vitro test methods. In this review, emphasis
is placed on literature-based (Q)SAR models for skin and eye irritation and corrosion
as well as computer-based expert systems.
Mission of the JRC
The mission of the JRC is to provide customer-driven scientific and technical support for the
conception, development, implementation and monitoring of EU policies. As a service of the
European Commission, the JRC functions as a reference centre of science and technology for
the Union. Close to the policy-making process, it serves the common interest of the Member
States, while being independent of special interests, whether private or national.