You are on page 1of 13

PERSPECTIVE

pubs.acs.org/crt

Application of the Hard and Soft, Acids and Bases (HSAB) Theory
to ToxicantTarget Interactions
Richard M. LoPachin,*,† Terrence Gavin,*,‡ Anthony DeCaprio,§ and David S. Barber||

Department of Anesthesiology, Montefiore Medical Center, 111 East 210th Street, Bronx, New York 10467, United States

Department of Chemistry, Iona College, New Rochelle, New York 10804, United States
§
Department of Chemistry and Biochemistry, Florida International University, 11200 S.W. Eighth Street, Miami, Florida 33199,
United States
)

Center for Environmental and Human Toxicology, University of Florida, Gainesville, Florida 32611, United States

ABSTRACT: Many chemical toxicants and/or their active metabolites are electrophiles that
cause cell injury by forming covalent bonds with nucleophilic targets on biological macro-
molecules. Covalent reactions between nucleophilic and electrophilic reagents are, however,
discriminatory since there is a significant degree of selectivity associated with these interactions.
Over the course of the past few decades, the theory of Hard and Soft, Acids and Bases (HSAB)
has proven to be a useful tool in predicting the outcome of such reactions. This concept utilizes
the inherent electronic characteristic of polarizability to define, for example, reacting electro-
philes and nucleophiles as either hard or soft. These HSAB definitions have been successfully
applied to chemical-induced toxicity in biological systems. Thus, according to this principle, a
toxic electrophile reacts preferentially with biological targets of similar hardness or softness. The
soft/hard classification of a xenobiotic electrophile has obvious utility in discerning plausible
biological targets and molecular mechanisms of toxicity. The purpose of this perspective is to
discuss the HSAB theory of electrophiles and nucleophiles within a toxicological framework. In
principle, covalent bond formation can be described by using the properties of their outermost or frontier orbitals. Because these
orbital energies for most chemicals can be calculated using quantum mechanical models, it is possible to quantify the relative softness
(σ) or hardness (η) of electrophiles or nucleophiles and to subsequently convert this information into useful indices of reactivity.
This atomic level information can provide insight into the design of corroborative laboratory research and thereby help investigators
discern corresponding molecular sites and mechanisms of toxicant action. The use of HSAB parameters has also been instrumental
in the development and identification of potential nucleophilic cytoprotectants that can scavenge toxic electrophiles. Clearly, the
difficult task of delineating molecular sites and mechanisms of toxicant action can be facilitated by the application of this quantitative
approach.

’ CONTENTS α,β-Unsaturated Carbonyl Derivatives: Relevance of


Introduction 240 HSAB Calculations to in Vivo Toxicokinetics and
HSAB Definitions: Soft and Hard, Electrophiles and Toxicodynamics 246
Nucleophiles 240 Cytoprotective Properties of 1,3-Dicarbonyl Enols:
Physiochemical Principles Determining the Selective Mechanistic Insights Gained by HSAB Quantum
Interactions of Electrophiles and Nucleophiles 240 Mechanical Calculations 247
HSAB Descriptors of ElectrophileNucleophile Adduct Cellular Oxidative Stress: Phytopolyphenol
Formation 241 Cytoprotection 247
Toxicological Application of HSAB Principles 242 1,3-Dicarbonyl Enols: A New Class of Cytoprotectants 247
α,β-Unsaturated Carbonyl Derivatives: 1,3-Dicarbonyl Enols: HSAB Parameters 247
Toxicological Consequences of Exposure 242 Summary 248
α,β-Unsaturated Carbonyl Derivatives: Softness Author Information 248
and Electrophilicity 242 Abbreviations 248
α,β-Unsaturated Carbonyl Electrophilicity: Intervening References 248
Variables That Influence Interpretation 243
α,β-Unsaturated Carbonyl Derivatives: Nucleophilic Received: August 5, 2011
Targets and Molecular Mechanisms of Toxicity 244 Published: November 04, 2011

r 2011 American Chemical Society 239 dx.doi.org/10.1021/tx2003257 | Chem. Res. Toxicol. 2012, 25, 239–251
Chemical Research in Toxicology PERSPECTIVE

’ INTRODUCTION Table 1. Hardness/Softness Values for Selected Electrophilea


Unlike the formation of drugreceptor complexes that primarily
involve short distance forces such as hydrogen-bonding, hydro-
phobic, or van der Waals interactions, many chemical xenobiotics
and/or their active metabolites are electrophiles (electron deficient)
that form covalent bonds (e.g., SN2, Schiff base formation, Michael
addition) with cognate nucleophilic (electron rich) sites on bio-
logical targets.110 Cytotoxicity occurs because the formation of
covalently bonded adducts can impair the functions of enzymes,
DNA, cytoskeletal proteins, and other macromolecules, thereby
leading to defective cell processes. Thus, it is now recognized that
electrophilic α,β-unsaturated carbonyl derivatives of the type-2
alkene chemical class (e.g., acrolein, acrylamide, or methylvinyl
ketone) cause broad organ system toxicity by forming covalent
Michael-type adducts with nucleophilic sulfhydryl groups on func-
tionally critical proteins.1120 However, electrophiles do not react
indiscriminately with nucleophiles, and instead, such interactions
occur along a continuum of relative reactivity. The significant degree
of selectivity that occurs in electrophilenucleophile interactions is
predicted by Pearson's Hard and Soft, Acids and Bases (HSAB)
theory.5,17,21 On the basis of this concept, reactive molecules are
assessed on the basis of their respective polarizability such that
electrophiles and nucleophiles are classified as either soft (relatively
polarizable) or hard (relatively nonpolarizable). A useful theorem
stemming from this principle is that toxic electrophiles react
preferentially with biological targets of similar hardness or softness.
Therefore, the α,β-unsaturated carbonyls can be viewed as soft
electrophiles that cause cytotoxicity by selectively forming adducts
with soft nucleophilic anionic thiolates on functionally critical pro-
teins.9,1416,18Whereas the HSAB theory initially described hard- a
Ground state equilibrium geometries were calculated for each structure
ness and softness in terms of the experimental ionization potential with DF B3LYP-6-31G* in water from 6-31G* initial geometries. LUMO
and electron affinity of the reacting molecules, these parameters also and HOMO energies were generated exclusively from corresponding s-cis
can be related (e.g., by Koopmans theorem) to the respective conformations and were used to calculate η and σ (see text).
energies of the outermost or frontier molecular orbitals (FMOs).
Since small molecule FMO energies can be determined computa- characteristics of hard and soft, and electrophiles and nucleo-
tionally using various quantum mechanical models, HSAB param- philes. As we will point out, establishing structureactivity
eters such as the softness (σ) or hardness (η) of an electrophile relationships (SAR) based on kinetic and HSAB properties can
and its nucleophilic target can be readily derived. Correspondingly, reveal the substructure (e.g., α,β-unsaturated aldehyde) or
the values for σ and η can be used in a more recently developed moiety (e.g., carbonyl group) responsible for the toxicity of the
algorithm to calculate the electrophilic index (ω) of a toxicant, the electrophile. This forms a rational basis for categorizing a
value of which reflects the propensity of the electrophile to form chemical series and for understanding the corresponding toxic
adducts with a given biological nucleophile. Although more com- mechanism.19,22,23 Finally, in this perspective we will present
plex, the respective σ and η values for the nucleophile can also be examples from the literature as a framework for demonstrating
used to derive an index of nucleophilicity (ω, see ahead). On a how quantum chemical calculations can be effectively applied
pragmatic basis, these parameters can be used to predict the toxic toward understanding molecular toxicodynamic issues.
potential of electrophilic xenobiotic chemicals and to facilitate the
identification of corresponding nucleophilic molecular sites of ’ HSAB DEFINITIONS: SOFT AND HARD, ELECTRO-
action.5,8,17 Indeed, our research has shown that, excluding steric PHILES AND NUCLEOPHILES
and other physiochemical issues (see ahead), the calculated values of
these descriptors (σ, ω, and ω) directly corresponded to the Physiochemical Principles Determining the Selective In-
second order rate constants for α,β-unsaturated carbonyl adduc- teractions of Electrophiles and Nucleophiles. The HSAB
tion of cysteine sulfhydryl groups. For individual members of a theory classifies reacting species as either relatively “hard” or
α,β-unsaturated carbonyl series, the kinetic and HSAB parameters “soft”, based on polarizability, i.e., the ease with which electron
were closely correlated to the respective magnitude of in vitro density can be displaced or delocalized to form new covalent
toxicity, i.e., disruption of synaptosomal function.1416,18 This ability bonds. Polarizability is an inherent characteristic of electron
of HSAB parameters (σ, η, ω, and ω) to predict the potency of a distribution in atoms or molecules. Thus, electrons that are far
given chemical toxicant has significant implications for risk asses- from the influence of the nucleus or that occupy a large volume
sment.19,20 In addition, these parameters are a useful tool in the (electron “cloud”) are more likely to be displaced, i.e., redis-
search for macromolecular targets and subsequent elucidation of tributed into a new bonding pattern. For example, the conjugated
molecular mechanisms of toxicity. Therefore, in this perspective, we α,β-unsaturated carbonyl structure of many type-2 alkenes
will discuss the HSAB theory and define the physiochemical is considered to be a soft electrophile (Table 1) because the
240 dx.doi.org/10.1021/tx2003257 |Chem. Res. Toxicol. 2012, 25, 239–251
Chemical Research in Toxicology PERSPECTIVE

Table 2. Hardness/Softness Values for Selected comparable softness, whereas hard electrophiles form adducts
Nucleophilesa with hard nucleophiles (see ahead). The data presented in
Table 2 indicate that, among the biological nucleophiles, the
sulfhydryl thiolate state of cysteine residues is the softest.5,8,17
Therefore, the preferred nucleophilic cellular targets of the soft
electrophilic α,β-unsaturated carbonyls are cysteine sulfhydryl
thiolate groups. In contrast, the neurotoxic n-hexane metabolite,
2, 5-hexanedione is a hard electrophile (Table 1) that preferen-
tially forms adducts with harder nucleophiles such as nitrogen
atoms on the ε-amino group of lysine residues. Additional hard
nucleophilic targets include the secondary imidazole nitrogen
of histidine and the carbon, nitrogen, and oxygen groups of
DNA/RNA.
HSAB Descriptors of ElectrophileNucleophile Adduct
Formation. As indicated above, covalent bond formation be-
tween reacting molecules can be described by using the proper-
ties of their outermost orbitals. The two most important of these
are the highest energy orbital that contains electrons (known by
the acronym HOMO = Highest Occupied Molecular Orbital)
and the lowest energy orbital that is vacant (LUMO = Lowest
Unoccupied Molecular Orbital). Covalent bonding such as
adduct formation occurs when electron density is transferred
between the orbitals of the reacting species (nucleophile to
electrophile) to form the new bonds of the product (the adduct).
FMO energies (EHOMO and ELUMO) for a wide variety of
chemical structures can be calculated from quantum mechanical
models, and the corresponding hardness (η) and softness (σ)
can then be derived using eqs 1 and 2.
η ¼ ½ELUMO  EHOMO =2 ð1Þ
a
Ground state equilibrium geometries were calculated for each structure
with DF B3LYP-6-31G* in water from 6-31G* initial geometries. Values σ ¼ 1=η ð2Þ
obtained were used to calculate σ and η (see text).
In essence, softness measures the ease with which electron
delocalized π electrons are mobile. This mobility can be consid- redistribution takes place during covalent bonding, i.e., nucleo-
ered the extension of an electron cloud over four nuclear centers phile donation of electrons and acceptance by the electrophile.
(CdCCdO) based on interactions between the orbitals of Therefore, with respect to electrophilic species, it is often the case
the electron-withdrawing carbonyl group and those of the that softness (higher σ value) is correlated with ease of adduct
alkene moiety. The extended cloud is easily distorted, hence soft formation (expressed, for example, as a reaction rate). The
(polarizable). In contrast, hard electrophilic toxicants, e.g., chloro- electrophilic index (ω) is a parameter that combines softness
ethylene oxide or vinyl chloride (Table 1), have highly localized, and chemical potential (μ; eq 3).24 The latter parameter, μ, is the
nonextended charge densitites at specific electrophilic centers, e.g., propensity of a species to undergo chemical change. Thus, ω is
the respective C1 carbon atoms. Consequently, these chemicals calculated according to eq 4 and represents a more comprehen-
are characterized by low electron polarizability. The softness or sive measure of electrophilic reactivity.
hardness of a nucleophile is also determined by the polarizability of μ ¼ ½ELUMO þ EHOMO =2Þ ð3Þ
corresponding electrons. Sulfur has a large atomic radius so that its
valence electrons are relatively far from the nucleus and, as such,
are highly polarizable. Upon ionization of the thiol (i.e., SH f S), ω ¼ μ2 =2η ð4Þ
the consequent expansion of the anionic cloud yields the relatively
With respect to the nucleophile, the corresponding molecular
soft (most easily polarizable) thiolate nucleophile, e.g., compare
orbital energies can also be used to calculate nucleophilic softness
the respective thiol- and thiolate-states of the amino acid residue,
and chemical potential as shown above in eqs 2 and 3. In addi-
cysteine (Table 2). In contrast, nitrogen and oxygen nucleophiles
tion, the likelihood that a given nucleophile (A) will form an
have relatively small atomic radii and, accordingly, their electron
adduct with a given electrophile (B) can be predicted by
clouds are less readily distorted. Consequently, such compounds
calculating a nucleophilicity index (ω) using eq 5.25
are much harder nucleophiles, e.g., the methoxide ion and the
primary nitrogen of the amino acid lysine (Table 2). Expansion of ω ¼ ηA ðμA  μB Þ2 =2ðηA þ ηB Þ2  ð5Þ
the electron cloud via nearby conjugation with unsaturated sites
yields a less hard nucleophile, e.g., the secondary amine of This parameter considers the hardness (η) and chemical poten-
histidine, the phenolate ion, or the enolate of 2-acetylcyclopenta- tial (μ) of both the electrophilic and nucleophilic reactants. The
none (Table 2). electrophilic (ω) and nucleophilic (ω) indices have been demon-
According to the selectivity principle of the HSAB model, soft strated to be reliable descriptors for a variety of electrophile
electrophiles preferentially form adducts with nucleophiles of nucleophile interactions.8,18,27,28
241 dx.doi.org/10.1021/tx2003257 |Chem. Res. Toxicol. 2012, 25, 239–251
Chemical Research in Toxicology PERSPECTIVE

Table 3. Calculated and Experimental Parameters for Conjugated Type-2 Alkenes and Non-Conjugated Analogues
type-2 alkene ELUMOa (ev) EHOMO (ev) σ (103 ev1) ω (ev) log k2b (pH 7.4) uptake inhibitionc (log IC50)

acrolein 1.70 6.98 379 3.57 2.596 4.28


MVK 1.46 6.69 382 3.18 2.048 3.48
HNE 1.84 6.93 393 3.78 0.938 3.40
MA 1.01 7.36 315 2.76 1.893 0.34
ACR 1.00 6.78 346 2.62 1.804 0.36
Nonconjugated Analogues
propanald 0.65 6.84 323 2.26
allyl alcohold +0.18 7.06 276 1.63
a
ELUMO and EHOMO were obtained after calculating ground state equilibrium geometries for each compound with DF B3LYP-6-31G* in vacuum from
6-31G* initial geometries. These values were generated exclusively from corresponding s-cis conformations and were used to calculate softness (σ) and
the electrophilic index (ω) of each electrophile as described in LoPachin et al.15 b Second-order reaction rates (k2) were determined for type-2 alkene
reactions with L-cysteine at pH 7.4 (n = 46 experiments). c Inhibition of synaptosomal membrane 3H-DA uptake was determined in striatal
synaptosomes exposed to type-2 alkenes (n = 46 experiments).14 d Does not undergo the Michael Reaction.

To perform the quantum mechanical computations, numerous α,β-Unsaturated Carbonyl Derivatives: Toxicological Con-
computer software packages such as Gaussian (www.gaussian.com), sequences of Exposure. Acrylamide (ACR), acrolein, 4-hydroxy-
HyperChem (www.hyper.com), Q-Chem (www.q-chem.com), 2-nonenal (HNE), and other structurally related derivatives
Spartan (www.wavefun.com) are available commercially. In addi- (Table 1) are characterized by a conjugated α,β-unsaturated
tion, there are many open source programs like GAMESS (www. carbonyl substructure that is formed when an electron-
msg.chem.iastate.edu), deMon (www.demon-software.com) and withdrawing group (e.g., a carbonyl group) is linked to an alkene.
Abinit (www.abinit.org) among others that can be used to make These are therefore soft electrophiles of the type-2 alkene
these types of calculations. While various models exist, density chemical class. Members of this class are used extensively in
functional theory (DFT) provides the most consistently em- the manufacturing, agricultural, and polymer industries, and
ployed approach (the B3LYP model) for the purposes described consequently, occupational human exposure is significant. In
by this perspective. However, in one case (see ahead) we used a addition, these chemicals are recognized as environmental pol-
HartreeFock (HF) model with good results. Although in most lutants and dietary contaminants, e.g., acrolein from the combus-
of the work described here molecular geometries (or energies) tion of petrochemical fuels, methyl vinyl ketone (MVK) in
are calculated in water, we have also reported an application automobile exhaust, and ACR in French fries. α,β-Unsaturated
using gas phase (in vacuum) calculations. The use of either carbonyl derivatives are also prevalent components of cigarette
state can be rationalized, but in our applications, no differe- smoke, e.g., acrolein, ACR, and acrylonitrile. Acute or chronic
nces in the results were observed as long as the method chosen exposure to these type-2 alkenes has been linked to major organ
was consistently employed. Throughout this study, only global system toxicity and to possible carcinogenicity in humans
(i.e., whole molecule) parameters are considered, but the use of and laboratory animals.5,3034 In addition to environmental or
localized functions at sites of reactivity within a molecule has been exogenous intoxication, acrolein, HNE, and certain other α,β-
successfully applied by others in the development of computa- unsaturated aldehydes are highly toxic byproducts of membrane
tional models.23,24 lipid peroxidation associated with cellular oxidative stress. Grow-
ing evidence indicates that these endogenous type-2 alkenes play
’ TOXICOLOGICAL APPLICATION OF HSAB a pathogenic role in many diseases that have oxidative stress as
PRINCIPLES a common molecular etiology, e.g., atherosclerosis, Alzheimer’s
Whereas exceptions exist, most toxic chemicals are electro- disease, and diabetes.5,13,17,35,36 Clearly, the type-2 alkenes are an
philes that produce toxicity by interacting with biological nucleo- important class of chemicals with respect to environmentally
philes,5,6,8,10,29 and therefore, the HSAB concepts should have acquired toxicity and endogenous disease pathogenesis. As
broad applicability to the field of molecular toxicology. In the described in the following subsection, deciphering the mecha-
following sections, we will demonstrate how the application of nism of type-2 alkene cytotoxicity can be aided significantly by
HSAB principles can help identify the sites and mechanisms of application of HSAB concepts and quantum mechanical descriptors.
toxicant action. Thus, we will consider the application of HSAB α,β-Unsaturated Carbonyl Derivatives: Softness and Electro-
principles to recent structuretoxicity studies involving acrolein, philicity. As discussed above, the relative softness and electro-
4-hydroxy-2-nonenal, and other α,β-unsaturated carbonyl and philicity of a α,β-unsaturated carbonyl derivative can be calcu-
aldehyde derivatives. In addition, we will discuss the role of lated using LUMO and HOMO energies determined via quantum
HSAB principles in the development of nucleophilic 1,3-dicar- mechanical models. As the data in Table 3 illustrate, acrolein
bonyl compounds that might be useful in treating disease and HNE are the strongest electrophiles among the type-2
processes involving endogenous α,β-unsaturated carbonyl com- alkenes tested, i.e., the respective gas phase electrophilicity
pounds. Although the HSAB calculations are broadly applicable index (ω) is numerically larger than other constituents in the
and straightforward, our discussions of these studies will also chemical series. In contrast, acrylamide (ACR) and methyl
serve to identify certain physiochemical toxicant characteristics acrylate (MA) are substantially weaker electrophiles, whereas
and intervening biological factors that can influence toxic expres- MVK exhibits intermediate electrophilic reactivity. Overall, the
sion and interpretation of the quantum mechanical data. type-2 alkene electrophilic indices suggest the following rank
242 dx.doi.org/10.1021/tx2003257 |Chem. Res. Toxicol. 2012, 25, 239–251
Chemical Research in Toxicology PERSPECTIVE

order of electrophilicity: acrolein ≈ HNE > MVK . MA g Table 4. Reaction of HNE/ONE Analogues with
ACR. In this type-2 alkene series, the relative differences in N-Acetylcysteine
electrophilicity are determined by substituent functional groups
and their respective contributions to the electron density of the
α,β-unsaturated carbonyl structure. Allyl alcohol and propanal
do not contain extended π electron systems and therefore have
substantially lower values of electrophilicity compared to that of
acrolein and related conjugated analogues (Table 3). Since allyl
alcohol and propanal lack a conjugated structure, neither of these
substances can undergo Michael-type reactions, i.e., conjugate
addition of a nucleophile. Subsequent experimental research
showed that softness and electrophilicity were determinants of
the chemical reactions that mediate type-2 alkene toxicity.1416
Excluding HNE (see below), the rank order of electrophilicity
was closely correlated (r2 > 0.90) to the corresponding second
order rate constants for the adduct reaction with cysteine (log k2)
and to the respective potencies (log IC50) for in vitro nerve
terminal (synaptosomes) toxicity (Table 3). While this corre-
lated data set consists of only four compounds, it should be
recognized that each structure represents a different functional
group (i.e., aldehyde, ketone, ester, and amide) and thus con-
stitutes a significant range of potential variability. Within the
confines of a particular functional group (e.g., esters), electro-
philicity is consistently expressed (see Table 5). Molecular
softness was also correlated to these experimental parameters
but to a slightly lesser extent (r2 ≈ 0.80). These data suggest that
electrophilicity governs the cysteine adduction rate and toxicity
a
of a chemical and that softness also is an important factor. Data are from McGrath et al.39, and shaded entries indicate dominant
α,β-Unsaturated Carbonyl Electrophilicity: Intervening species (see corresponding text). b Half-lives and rate constants from
Variables That Influence Interpretation. The preceding data reaction with N-acetylcysteine at pH 7.4, 37 °C. c ω = μ2/2η (see text).
Orbital energies were generated exclusively from corresponding s-cis
indicate that, although values for HNE electrophilicity and
conformations and using ground state equilibrium geometries computed
softness were higher than those of acrolein and MVK,37 the rank using HF 6-31G* in water starting from 3-21G geometry. d Orbital
order of the respective adduct rate constant (log k2) and toxic energies were generated exclusively from corresponding s-cis conforma-
potency (log IC50) were substantially lower than those predicted tions and were obtained from ground state equilibrium geometries
by the HSAB parameters. This discrepancy would seem to indi- computed using DF B3LYP 6-31G* in water starting from 6-31G*
cate a limited role for electrophilicity in determining toxic expres- geometry. e Failed to converge.
sion. However, the extended alkane tail of HNE (Tables 1 and 4)
imposes significant steric hindrance that can impede access to the
active site of many enzymes and thereby slow the corresponding PTP1B at higher concentrations and/or longer exposure times
kinetics of adduct formation and reduce toxic potency.37 Such that overcome steric hindrance. However, this is not the case for
disagreement between direct chemical measurements and calcu- propanal since it is neither structurally nor electronically capable
lated HSAB parameters should be expected since the respective of irreversible sulfhydryl adduction.
algorithms for σ and ω do not consider steric factors. As an As part of a structuretoxicity analysis of lipid electrophiles
additional illustration of steric influences, Seiner et al.38 demon- generated in vivo during phospholipid peroxidation, McGrath
strated that acrolein and a series of structurally related aldehydes et al.39 determined the corresponding half-lives (t1/2) and second
inhibited protein tyrosine phosphatase 1B (PTP1B) activity via order rate constants (k2) for the reactions of HNE, 4-oxonenal
covalent modification of a specific cysteine residue (Cys215). (ONE) and analogue metabolites (Table 4) with N-acetyl-
The order of potency was acrolein . crotonaldehyde > 3-methyl- cysteine (NAC). As indicated above, these lipid derivatives are
2-butanal ≈ propanal. When the corresponding aqueous α,β-unsaturated carbonyl compounds that mediate cytotoxicity
phase electrophilic indices (ω) were calculated retrospectively, through irreversible modification of macromolecules and disrup-
the greater ability of acrolein to inactivate PTP1B was clearly tion of cellular response pathways.17,40,41 When the correspond-
related to correspondingly higher electrophilicity (ω = 3.82 ev), ing electrophilicity values (ω) for the HNE analogues were
whereas the reduced potency of crotonaldehyde was a function of calculated, the kinetic parameters (t1/2 and k2) were only
lower electrophilicity (ω = 3.27 ev). However, the ω value for qualitatively correlated to ω unless the acidity of certain meta-
3-methyl-2-butanal (ω = 3.31) was significantly higher than that bolites (RCOOH f RCOO) was considered. Specifically,
of propanal (ω = 2.32) and equivalent to that of crotonaldehyde when structures of the dominant anionic species (RCOO) at
(ω = 3.27). Similar to the previous HNE study (Table 4), this pH 7.4 were used in a HF quantum mechanical model, correla-
order of electrophilicity is not consistent with the rank order of tion of both k2 and t1/2 with ω improved from r2 = 0.553 to 0.906
potency. Nonetheless, as the authors indicate, it can be rationa- and r2 = 0.752 to 0.911, respectively. In addition, the lack of
lized by steric hindrance imposed through the bulky tertiary reactivity exhibited by HNEA in this study can be related to the
alkene terminus of 3-methyl-2-butanal. It is noteworthy that, due extremely low value of ω calculated for its conjugate base
to its significant ω value, this compound would likely inactivate (Table 4). It can also be seen from Table 4 that although the
243 dx.doi.org/10.1021/tx2003257 |Chem. Res. Toxicol. 2012, 25, 239–251
Chemical Research in Toxicology PERSPECTIVE

Table 5. Acrylate and Methacrylate Oral Lethality in Mice molecule (i.e., carboncarbon double bond) and might have
little effect on electrophilic reactivity. Thus, it should be evident
LD50a ω
from this example that, when individual differences in solubility
ester (mmol/kg) log Pa log ka (ev)b are considered, the relative electrophilicity (ω) of the acrylates/
methyl acrylate 9.6 0.8 1.602 3.2245 methacrylates determines both the corresponding in vitro rate
ethyl acrylate 17.97 1.33 1.428 3.1981 (log k) of cysteine adduct formation and the expression of in vivo
n-butyl acrylate 58.98 2.36 1.401 3.196067
toxicity.18,29,44
The preceding discussion indicates that a variety of chemical
isobutyl acrylate 47.63 2.22 1.431 3.196067
toxicants can be classified according to the quantitative HSAB
2-hydroxyethyl acrylate 5.177 0.21 1.919 3.20026
descriptors of hardness (η) and softness (σ), as determined by
2-hydroxypropyl acrylate 8.112 0.35 1.758 3.214 quantum mechanical calculations. Furthermore, the degree of
methyl methacrylate 51.97 1.38 0.675 2.9967 electrophilic reactivity (ω) for these compounds can also be
ethyl methacrylate 68.64 1.94 0.403 2.9729 quantitatively determined. This latter parameter is critically
isopropyl methacrylate 67.84 2.25 0.876 2.9708 important since it is directly related the second order rate
n-butyl methacrylate 142.7 2.88 0.655 2.95895 constant (k2) of the target adduct reaction and, therefore, reflects
isobutyl methacrylate 83.14 2.66 0.769 2.9708 toxicological potential. Accordingly, ω has been suggested to
2-hydroxyethyl methacrylate 45.24 0.47 0.273 3.0108 be an in silico surrogate parameter for experimentally derived
2-hydroxypropyl methacrylate 55.24 0.97 0.273 2.9848 rate constants (k2) and, as such, might be a useful tool for
tert-butyl methacrylate 60.12 2.54 0.632 2.9708
predicting the toxicity of industrial chemicals.10,19,20,22,44 The
preceding discussion also pointed out that the HSAB algorithms
a
Data from Tanii and Hashimoto.42 b ω = μ2/2η (see text). Orbital
energies were generated exclusively from corresponding s-cis conforma- do not account for possible intervening physiochemical variables,
tions and were obtained from ground state equilibrium geometries e.g., solubility, acidbase equilibrium and steric hindrance, that
computed using DF B3LYP 6-31G* in water starting from 6-31G* could independently modify the toxicological outcome and lead
geometry. to discrepancies in predicted vs observed results. These issues
have been addressed to some extent, and in at least one case,
electrophilic and steric parameters were combined in a computa-
DF model yielded numerically different values for ω, the same tional algorithm.24
result was obtained, i.e., rate data were well correlated to electro- The α,β-unsaturated ester and carbonyl derivatives discussed
philicity when the relevant anionic species were considered. here form adducts with cysteine and other amino acid targets on
Esters of acrylic acid and methacrylic acid are used extensively proteins via second order (covalent) reactions. The rates of these
in the production of polymers for textiles, latex paints, and reactions are not only dependent upon the toxicant ω but also
medical/dental cements. The acrylates (CH2dCHCO2R) and the nucleophilic status of the target residues. Therefore, in the
methacrylates (CH2dCCH3CO2R) are conjugated α,β-unsatu- next section, we will discuss the application of HSAB parameters
rated ester derivatives. Several lines of evidence suggest that to the identification of amino acid sites of toxicant action.
these compounds produce toxicity (e.g., carcinogenicity and α,β-Unsaturated Carbonyl Derivatives: Nucleophilic Targets
skin sensitization) through adduction of protein sulfhydryl and Molecular Mechanisms of Toxicity. According to the
groups.20,4244 Early studies by Tanii and Hashimoto42 deter- selectivity principle of the HSAB theory, soft electrophiles should
mined the lethal oral doses (LD50) in mice for a broad spectrum preferentially react with soft nucleophiles. There are several
of acrylates/methacrylates. Results indicated that the LD50 was nucleophilic amino acid side chains that might be relevant
related to both solubility (as a partition coefficient, log P) and the protein targets for the α,β-unsaturated carbonyl compounds,
reaction rate (log k) of these esters with glutathione in a buffered i.e., sulfhydryl groups on cysteine (Cys) residues, the primary
physiological solution. When the electrophilic index (ω) was nitrogen atoms of the lysine (Lys) ε-amino group, and the
calculated for the 14 compounds (Table 5), we found that the secondary nitrogen atoms of the imidazole group of histidine
corresponding rate constants (log k) were highly correlated (r2 = (His). Although the type-2 alkenes can form adducts with lysine
0.97) to the ω parameter. In contrast, both log k (r2 = 0.65) and and histidine residues,40,4552 the weight of kinetic and proteo-
ω (r2 = 0.60) were only modestly correlated with the in vivo mic data indicate that cysteine sulfhydryl groups are the prefer-
LD50 values. This discrepancy can, however, be explained by the ential targets for the type-2 alkenes.11,12,14,16,37,40,41,49,5258 On a
variable solubility (log P; Table 5) of the acrylate and methacry- quantitative basis, this cysteine preference is several orders of
late compounds18 and by the fact that toxicant solubility is a magnitude greater than that for the other biological nucleophiles.
physiochemical determinant of tissue distribution and, therefore, This differential is consistent with the fact that the side chain
target accessibility. This is evidenced by comparing the respective amino nitrogen groups of lysine and histidine are harder nucleo-
data for the water-soluble 2-hydroxyethyl acrylate (log P = philes (Table 2) and have inherently lower reactivity for soft
0.21) to the less soluble ethyl acrylate (P = 1.33), i.e., both electrophiles such as the type-2 alkenes. The likelihood that a
have virtually identical ω values (∼3.200 ev) but significantly given nucleophile will form a covalent adduct with a type-2
different reaction rates (log k = 1.92 vs 1.43, respectively) and alkene can be predicted by calculating an index of nucleophilicity
toxic potencies (LD50 = 5.2 mmol/kg vs 18 mmol/kg, (ω, see above). This comprehensive parameter considers the
respectively). Clearly, the chemical factors that establish solubi- hardness (η) and chemical potential (μ) of both the electrophilic
lity do not necessarily produce commensurate changes in (type-2 alkene) and nucleophilic (cysteine, histidine, or lysine)
electronic characteristics such as electrophilicity. This is not reactants.25 As suggested by the respective ω values (Table 6),
surprising since, although the hydroxyl group improves solubility the type-2 alkenes preferentially form adducts with cysteine
(i.e., CH2dCHCO2CH2CH2OH vs CH2dCHCO2CH2CH3), thiolate sites. However, based on a pKa of 8.3, the cysteine
this substituent is distant from the electrophilic site of the sulfhydryl side chain is mostly protonated at physiological
244 dx.doi.org/10.1021/tx2003257 |Chem. Res. Toxicol. 2012, 25, 239–251
Chemical Research in Toxicology PERSPECTIVE

Table 6. Calculated Nucleophilic Indices (ω) for Type-2 (concentration independent) nucleophilic strength.58 Using the
Alkene Reactions with Cysteine Sulfhydryl Groups expression log (kRS  k2) = log k2 + pKa  pH (where kRS
represents the anionic rate constant), a representative series of
electrophile ω Cysa (1) ω Cys (0) log kRSb log IC50c
thiolate rate constants were calculated. For each α,β-unsaturated
acrolein 2.03 0.103 3.417 4.28 carbonyl derivative, the corresponding thiolate rate constants
HNE 1.93 0.083 1.759 3.40 (kRS) were highly correlated to ω (r2 = 0.91; Table 6).
MVK 1.83 0.064 2.953 3.48 Furthermore, the fact that ω and kRS were closely correlated
MA 1.59 0.069 1.011 0.34
to the neurotoxic potencies (IC50’s; Table 3) provided evidence
that thiolate targeting by acrolein and HNE has toxicological
ACR 1.50 0.036 0.767 0.36
relevance.15,16 Similarly, previously reported rate data (k2, kRS)
a
The nucleophilicity index (ω) was calculated as ω = ηA (μA  μB)2/
for the reactions of unhindered cysteine related thiolate nucleo-
2(ηA  ηB)2, where η = (ELUMO  EHOMO)/2, μ = (ELUMO + EHOMO)/2,
A = reacting nucleophile, and B = reacting electrophile.15 FMO energies philes58 with acrylonitrile were well correlated (r2 > 0.97) to our
(ELUMO and EHOMO) were generated exclusively from corresponding corresponding calculations of ω values (Table 7).
s-cis conformations and were derived from equilibrium geometries of the As a more relevant experimental approach to defining the
electrophiles at the ground state calculated using DF B3LYP-6-31G* in interactions of type-2 alkenes with cysteine sulfhydryl groups, we
water starting from 6-31G* geometries. For the sulfhydryl thiolate and determined the effects of acrolein, MVK, and ACR on erythro-
thiol states, the respective ionization-states are presented in parentheses. cyte glyceraldehyde-3-phosphate dehydrogenase (GAPDH)
The nucleophilicity index considers the respective hardness and chemi- activity.18 Consistent with the concepts of the previous section,
cal potential of the electrophilic (type-2 alkene) and nucleophilic the softness (σ) and electrophilicity (ω) values of the selected
(cysteine, histidine, or lysine) reactants and is, therefore, a measure of type-2 alkenes were related to the corresponding second order
the likelihood of subsequent adduct formation. As suggested by the
rate constants (log k2) and potencies (log KI) for GAPDH
respective ω values, the type-2 alkenes preferentially form adducts with
cysteine thiolate sites as opposed to the thiol state. b The k2 values at pH inhibition (Table 8). Whereas earlier proteomic studies demon-
7.4 were corrected for the corresponding cysteine thiolate concentration strated the cysteine preference of type-2 alkene adduct forma-
(kRS) according to the algorithm: log (kRS  k2) = log k2 + pKa  pH. tion,40,41,56,59,60 it is unlikely that these adducts all have toxico-
Inhibition of membrane 3H-dopamine transport was determined in rat logical relevance, i.e., the corresponding cysteine residues all play
striatal synaptosomes exposed in vitro to graded concentrations of each a vital role in protein function. Indeed, other studies indicated
type-2 alkene. c The concentrationresponse data for transport were that the type-2 alkenes impaired protein function by reacting
fitted by nonlinear regression analysis and the respective IC50’s were with specific cysteine residues on important cellular proteins, e.g.,
calculated by the ChengPrusoff equation.14,15 Abbreviations: HNE = N-ethylmaleimide inhibits vesicle (H+)-ATPase activity by
4-hydroxy-2-nonenal, MVK = methyl vinyl ketone, MA = methyl reacting with Cys254;61 and HNE adduct formation at Cys280
acrylate, and ACR = acrylamide.
inhibits mitochondrial SIRT3 acitivity.62 However, it was unclear
why these residues were selectively targeted. Therefore, we used
pH (7.4) and, therefore, exists primarily as the neutral (0) thiol. tandem mass spectrometry to quantify the adduct formation
Also at this pH, the imidazole secondary amine of histidine (pKa associated with graded concentrations of a weak electrophile,
ACR.18 Results indicated that lower in vitro concentrations of
of 6.0) is mostly deprotonated (0), and the primary ε-amino group
ACR inhibited GAPDH activity by selectively forming adducts
amine of lysine (pKa = 10.5) is completely protonated (+1). As
with Cys 152 in the active site of this enzyme, whereas at higher
the corresponding HSAB parameters indicate (Table 2), the concentrations, ACR also reacted with Cys 156 and Cys247.
relative nucleophilicity values for the thiol and nitrogen side Calculations using the PROPKA program revealed a pKa of
chain groups are low, and hence, these side chain states are not 6.03 for Cys152, whereas the pKa values for Cys156 and Cys247
particularly reactive. Instead, a large body of evidence indicates were higher. Furthermore, we found that GAPDH inhibition by
that the α,β-unsaturated carbonyls selectively form adducts with the selected type-2 alkenes was pH-dependent indicating thiolate
the highly nucleophilic thiolate of cysteine sulfhydryl groups. mediation. These data suggest that Cys152 is contained within a
This anionic state can be found within specialized pKa-lowering low pKa microenvironment63,64 and that ACR reacted preferen-
microenvironments such as catalytic triads,5,17 and as the respec- tially with the resulting sulfhydryl thiolate.
tive HSAB descriptors (Tables 2 and 6) demonstrate, cysteine Considered together, our data indicate that α,β-unsaturated
thiolates are a much softer and significantly more reactive carbonyl chemicals produce cytotoxicity by forming Michael-
nucleophile than the Lys, His, or thiol sulfhydryl groups. type adducts with sulfhydryl groups on specific cysteine residues.5,17,59
To provide experimental evidence for the thiolate-state pre- The elucidation of this mechanism was guided by HSAB-based
ference of the α,β-unsaturated carbonyls, we determined the calculations of softness and hardness, which indicated that the
effects of pH on the corresponding rates of adduct formation. type-2 alkenes were electrophiles of varying softness and that, as
The thiolate concentration in solution will vary as a function of such, these chemicals could form covalent bonds (adducts) with
pH, i.e., given a cysteine pKa of 8.3, at physiological pH (7.4) the nucleophiles of comparable softness. Calculation of ω sug-
majority (∼90%) of sulfhydryl groups will be in the nonreactive gested nucleophilic anionic thiolates as potential molecular
thiol-state, whereas at higher pH, the thiolate/thiol ratio will sites of electrophile action. Cysteine thiolates are present in the
become increasingly larger. Thus, we found that the reactions of active sites of many proteins where they play critical roles in
L-cysteine (pKa = 8.3) with selected type-2 alkenes were 1015 modulating protein activities and, therefore, post translational
times faster when the pH was increased from 7.4 to 8.8.1416 modification via adduct formation would cause cytotoxicity by
The observed changes in second-order rate constant (k2) reflect inhibiting the function of important proteins, e.g., N-ethylmaleimide
the increased thiolate concentration at the higher pH value. The sensitive factor (NSF), GAPDH, ubiquitin carboxyl terminal
experimentally determined k2 can be used to derive an anionic hydrolase-L1 (UCH-L1), clathrin, and vesicular-ATPase.1114
rate constant (kRS) as a quantifiable measure of the inherent These predictions were corroborated by direct evidence from a
245 dx.doi.org/10.1021/tx2003257 |Chem. Res. Toxicol. 2012, 25, 239–251
Chemical Research in Toxicology PERSPECTIVE

Table 7. Reaction of Thiolate Nucleophiles with Acrylonitrile at pH 8.1 and 30 °C

a
Data from Friedman et al.58 b Calculated as log (k  k2) = log k2 + pKa  pH. c Orbital energies obtained from ground state equilibrium geometries
computed using DF B3LYP 6-31G* in water starting from 6-31G* geometry.

Table 8. HSAB and Kinetic Parameters for Type-2 Alkene This common mode of action suggests that members of this
Interactions with GAPDH chemical family should, depending upon route of exposure,
produce a similar pattern of organ system toxicity. However,
electrophilea σ ( 103 ev1) ω (ev) log k2 log KI
ACR intoxication in laboratory animals or humans (e.g., occupa-
acroleinb 371 3.82 4.250 4.419 tional exposure or poisoning) produces selective neurotoxicity,
MVK 363 3.38 3.885 4.220 whereas similar intoxication with acrolein, MVK, or other type-2
ACR 315 2.61 0.502 0.607 alkenes causes peripheral organ toxicity.5,31 This seemingly
a
HSAB (σ, ω) and kinetic parameters (k2, KI) were calculated as inconsistent diversity in toxicological responses is not due to
described in Martyniuk et al.18 b On the basis of the HSAB parameters, variations in mechanism but to relative differences in electro-
acrolein and MVK are significantly softer and more reactive electrophiles philicity that correspondingly influences tissue distribution and
than ACR, i.e., larger values of σ and ω, respectively. The rank orders of toxicokinetics. As we have discussed, acrolein is a highly reactive
respective σ and ω values for each type-2 alkene were closely correlated soft electrophile (Table 1) that rapidly forms adducts with soft
to the corresponding rate constants (k2; r2 = 0.9996 and 0.9359, nucleophilic thiolate sites on cysteine residues of proteins
respectively) and relative potencies (KI; r2 = 0.9926 and 0.9004, (Table 6). Following systemic intoxication with a reactive type-
respectively) for the inhibition of GAPDH activity. 2 alkene, the rapid formation of protein adducts essentially limits
tissue distribution. Consequently, the resulting toxic manifesta-
tions are determined by the site of absorption, e.g., acrolein
variety of proteomic and kinetic experimental approaches in both inhalation (cigarette smoke and pollution) produces pulmonary
in vitro and in vivo models of type-2 alkene toxicity.5,17,31,59 toxicity, whereas systemic acrolein intoxication is associated with
Because these thiolate targets are acceptors for nitric oxide (NO) hepatic and/or vascular toxicity.66,67 In contrast to more reac-
and other modulatory electrophiles, we have proposed that toxi- tive type-2 alkenes, ACR is a weak water-soluble electrophile
city induced by α,β-unsaturated carbonyl exposure is mediated (Table 1) that slowly forms adducts with protein thiolate sites
by the disruption of cellular redox signaling pathways.5,17,31,65 As (Table 6). Correspondingly, ACR is less influenced by systemic
will be discussed, a similar approach was used to decipher the “adduct buffering” and therefore has a relatively large volume of
mechanism of protection against cellular oxidative stress pro- distribution that includes the central nervous system.68,69 In
vided by 1,3-dicarbonyl enols. accordance with these concepts, oral acrolein intoxication of rats
α,β-Unsaturated Carbonyl Derivatives: Relevance of HSAB (0.050.5 mg/kg) produced centrolobular liver necrosis and
Calculations to in Vivo Toxicokinetics and Toxicodynamics. vascular hypertension,70 whereas oral ACR (1050 mg/kg)
The evidence reviewed above indicates that the type-2 alkenes caused selective neurotoxicity in this species.71 However, the
cause cytotoxicity through a common molecular mechanism. production of neurotoxicity is not simply a function of brain
246 dx.doi.org/10.1021/tx2003257 |Chem. Res. Toxicol. 2012, 25, 239–251
Chemical Research in Toxicology PERSPECTIVE

accessibility since ACR has an equal propensity to form thiolate Table 9. HSAB and Kinetic Parameters for Reactions of 1,3-
adducts in both nervous and non-nervous organ systems. Instead, Dicarbonyl Compounds with Acrolein
our studies indicate that ACR neurotoxicity is mediated by
compd ω (103 ev)a pKa k2 rate constantb
selective nerve terminal damage and that several unique anato-
mical and functional characteristics predispose this neuronal 2-ACP 401 7.8 0
region to electrophile-induced toxicity.5,17 Specifically, as indi- AcAc 160 8.9 24.9
cated above, we suspect that ACR (and other type-2 alkenes) TFPD 108 4.7 52.1
causes cytotoxicity by reacting with thiolate sites that act as NO DEM 228 12.9 78.7
acceptors. Neurotransmission is a complex multistep process phloretin 139 7.3 1.9
that is highly regulated by NO signaling,59 and therefore, NAC (2) 316 9.5 nd
disruption of this pathway is likely to have a significant impact a
The nucleophilic index (ω) was calculated based on the reaction of
on presynaptic activity. Furthermore, the nerve terminal is acrolein with the selected 1,3-dicarbonyl and was calculated according to
anatomically separated from the cell body and is, as a result, the algorithm provided in the text. FMO energies were generated
devoid of transcriptional and translational abilities. Therefore, exclusively from corresponding s-cis conformations and were calculated
unlike the cell body, the nerve terminal is vulnerable to electro- as equilibrium geometries at the ground state using DF B3LYP-6-31G*
phile attack since it cannot mount transcription-based cytopro- in water starting from 6-31G* geometries. b Second order rate con-
tective mechanisms such as the Nrf2-Keap1 response.72 Finally, stants (k2) for the reactions of 1,3-dicarbonyls with acrolein are derived
relative to proteins in other nerve regions or cell types, the from LoPachin et al.87 and are expressed as nM sulfhydryl loss s1.
turnover of nerve terminal proteins is exceptionally slow.73 Thus,
when these proteins are rendered dysfunctional through cysteine 1,3-Dicarbonyl Enols: A New Class of Cytoprotectants.
adduct formation, they are replaced slowly. This leads to the The search for safe, water-soluble phytopolyphenol analogues
accumulation of nerve terminal adducts,11,12 progressive impair- led to the discovery that 1,3-dicarbonyl enols such as 2-acetyl-
ment of presynaptic processes,13,57 and cumulative neurotoxi- cyclopentanone (2-ACP) and acetylacetone (AcAc; Table 2)
city.71 Thus, although the type-2 alkenes have a common mech- might be cytoprotective.87 The structure of these simple soluble
enols resembles the central diketonebridge of curcumin. Like
anism of cytotoxicity, differences in electrophilic strength among
the phytopolyphenols, the α-carbon hydrogen atoms of these
these chemicals can influence respective tissue distributions and
simple enols can also readily ionize to form highly nucleophilic
therefore their toxicological manifestations.
soft enolates. Indeed, our studies87 showed that the 1,3-dicarbo-
nyl compounds provided levels of cytoprotection in several in
’ CYTOPROTECTIVE PROPERTIES OF 1,3-DICARBO- vitro models of oxidative stress that exceeded those of the
NYL ENOLS: MECHANISTIC INSIGHTS GAINED BY HSAB phytopolyphenols. The evidence suggested at least two mecha-
QUANTUM MECHANICAL CALCULATIONS nisms of 1,3-dicarbonyl cytoprotection in oxidative stress-induced
Cellular Oxidative Stress: Phytopolyphenol Cytoprotec-
cell injury. First, metal ion chelation mediated by bidentate
tion. There is general agreement that cellular oxidative stress coordination through the β-diketone structure. Presumably,
is a common molecular etiology in many chronic disease states the resulting inhibition or modification of the metal-catalyzed
(e.g., Alzheimer’s disease, atherosclerosis) and acute injuries Fenton reaction reduces the cellular free radical burden. Second,
(e.g., stroke and traumatic spinal cord injury). A confluence of the nucleophilic β-dicarbonyl enolates act as soft surrogate
targets for free acrolein and other electrophilic α,β-unsaturated
epidemiological data and clinical trials,74,75 corroborated by
aldehydes. The reduction in toxic aldehyde content prevents
laboratory evidence,7678 indicate that the course of these con-
secondary protein dysfunction in exposed cells. In studies that
ditions can be significantly improved by the administration of
specifically examined the latter cytoprotective component, we
plant-derived polyphenolic compounds such as curcumin (curry
identified the following rank order of 1,3-dicarbonyl analogues
spice), resveratrol (red grapes), or phloretin (apple skins). It has
with respect to protection in acrolein-exposed neuronal cell
been largely assumed that the protective effects of these phyto- culture systems: 2-ACP > AcAc, 1,1,1-trifluoro-2,4-pentanedione
polyphenols were related to their abilities as antioxidants to (TFPD) . diethylmalonate (DEM). This sequence reflects the
trap reactive oxygen species. However, recent evidence indicates differences in second order rate constants (k2) for the reaction of
that polyphenolic cytoprotection involves scavenging of other these 1,3-dicarbonyls with acrolein. The differential reactivity of
electron-deficient species that also mediate oxidative stress, i.e., the 1,3-dicarbonyls can be explained by the acidity (pKa) of the
divalent metal ions (e.g., Fe2+ and Cu2+) and toxic unsaturated parent compound and by the inherent nucleophilicity of the
aldehydes (e.g., acrolein and HNE).7981 Structureactivity corresponding enolate. The pKa values indicate the relative
studies have indicated that the molecular determinant of these extent of ionization (enolate formation) and hence reflect nucleo-
protective actions is resident enol moieties.8284 For example, philic concentration as a factor contributing to reaction rate.
the α-carbon hydrogen atoms on the central β-diketone bridge of 1,3-Dicarbonyl Enols: HSAB Parameters. As discussed pre-
curcumin are acidic, and upon ionization, in physiological buffers, viously, the nucleophilic strength of an enolate carbanion can be
a carbanionic enolate is formed as the corresponding conjugate quantified by ω.5,17 The ω values computed for reactions
base. Moreover, the characteristic phenols of flavonoids such as between acrolein and various nucleophiles are presented in
phloretin are enols since they consist of a hydroxyl group Table 9. A nucleophile of significant reactivity (high ω value)
attached to a carboncarbon double bond. These enols will also will increase the reaction rate (k2) with a given electrophile by
form nucleophilic enolate ions. Despite therapeutic potential, lowering the energy of the transition state. Therefore, ω values
however, the clinical utility of the phytopolyphenols could should correlate with the relative rate at which a given concen-
be limited by poor bioavailability, toxicity, and chemical tration of enolate reacts with a specific type-2 alkene (Table 9).
instability. 36,85,86 Inherent differences in the respective ω and known pKa values
247 dx.doi.org/10.1021/tx2003257 |Chem. Res. Toxicol. 2012, 25, 239–251
Chemical Research in Toxicology PERSPECTIVE

provide a molecular explanation of the corresponding adduct less reactive protoxicants. However, despite these obvious benefi-
reaction rates and, hence, cytoprotective potency. Thus, 2-ACP is cial applications of the HSAB calculations, in vivo and in vitro
reasonably acidic (3040% ionized at pH 7.4; pKa = 7.8) and model systems and their presumed macromolecular targets are
forms an enolate that is a powerful nucleophile (relative ω = highly complex, and accordingly, accurate formulation of kinetic
401). Correspondingly, it exhibited superior protective capacity or mechanistic predictions is dependent upon the knowledge of
in all toxicity models. In contrast, AcAc is less reactive toward additional chemical features, such as steric hindrance, that might
electrophiles, not only because it ionizes to a lesser extent (2% at independently influence adduct reaction rates. Furthermore, it
pH 7.4; pKa = 8.9) but also because it is a weaker nucleophile should be recognized that the relative electrophilicity of a
(ω = 160). Although TFPD is highly acidic (pKa = 4.7), the toxicant will influence corresponding in vivo toxicokinetics and
resulting higher concentration of enolate at physiological pH is the organ systems affected, as evidenced in our study of type-2
offset by its weakness as a nucleophile (ω = 108). Conversely, alkene toxicity. Regardless, HSAB calculations provide molecular
the enolate form of DEM is a significant nucleophile (ω = 228). level information that can offer guidance for corroborative
However, the low acidity (pKa = 12.9) of this chemical precludes laboratory research and thereby help investigators discern corre-
any potential capability as a protectant since less than 0.001% sponding electrophile molecular mechanisms and sites of tox-
would be in the enolate state at physiological pH. NAC (pKa = icant action. The use of quantum chemical calculations has
9.5) ionizes poorly in biological solutions, but is protective also been instrumental in the development and identifica-
against acrolein-induced thiol loss and toxicity because the tion of potential nucleophilic antagonists, which can scavenge
nucleophilicity of the thiolate anion is relatively high (ω = electron-deficient species that mediate many diseases with a
316). Phloretin has several enolizable sites, and although the common etiology of cellular oxidative stress. Clearly, the difficult
nucleophilicity of these sites is modest (e.g., the relative ω of task of delineating molecular sites and mechanisms of toxicant
the C-3 enol on the A ring = 139), the corresponding pKa value of action can be advanced by the application of quantum mechanical
7.3 indicates significant (>50%) ionization. Therefore, at phy- approaches.
siological pH the high concentration of enolate and multiple site
potential enable phloretin to provide significant thiol preserva- ’ AUTHOR INFORMATION
tion through acrolein adduction. Together, these results indicate Corresponding Author
that the ability of 1,3-dicarbonyl compounds to act as acrolein *(R.M.L.)Tel: (718) 920-5054. Fax: (718) 515-4903. E-mail:
targets is a predictable function of enolate formation and the lopachin@einstein.yu.edu. (T.G.) Department of Chemistry,
inherent nucleophilicity of these carbanions. We are currently Iona College, 402 North Ave., New Rochelle, NY 10804. Tel:
using the ability to quantify nucleophilicity (ω), in conjunction (914) 633-2301. E-mail: tgavin@iona.edu.
with a high level of chemical understanding (i.e., conformational
flexibility, pKa, and nucleophilicity requirements) derived from Funding Sources
extensive experimentation, to design and identify efficacious This research was supported by NIH grants from the National
cytoprotective 1,3-dicarbonyl analogues. Institutes of Environmental Health Sciences to R.M.L. (NIEHS
RO1 ES03830-24; NIEHS RO1 07912-13).
’ SUMMARY
’ ABBREVIATIONS
In the field of toxicology, the irreversible covalent interaction
of a toxic electrophile and its cognate nucleophilic target is HSAB, hard and soft, acids and bases; FMOs, frontier molecular
recognized as a basic reaction mediating chemical-induced cell orbitals; SAR, stuctureactivity relationship; LUMO, lowest
injury. It is now understood that electrophilenucleophile unoccupied molecular orbital; HOMO, highest occupied molecular
reactions exhibit a significant degree of selectivity, in that a given orbital; DFT, density-functional theory; HTT, Hartree
electrophile preferentially reacts with specific nucleophiles of Fock theory; ACR, acrylamide; HNE, 4-hydroxy-2-nonenal;
comparable softness or hardness. This selectivity is based on MVK, methylvinyl ketone; MA, methyl acrylate; PTP1B, protein
electronic and structural characteristics that constitute the soft tyrosine phosphate 1B; NAC, N-acetylcysteine; LD50, lethal
and hard classifications of the HSAB theory. In this perspective, oral dose for 50% of the population; GAPDH, glyceraldehydes
we have shown that the softsoft interactions of electrophilic 3-phosphate dehydrogenase; SIRT3, mitochondrial sirtuin3;
α,β-unsaturated carbonyl and aldehyde derivatives with their NSF, N-ethylmaleimide sensitive factor; UCH-L1, ubiquitin carboxyl
nucleophilic sulfhydryl thiolate targets can be understood and terminal hydrolase-L1; NO, nitric oxide; Nrf2/Keap1, nuclear
predicted through the calculation of HSAB parameters. Hard factor erythroid 2-related factor 2/kelch-like erythroid cell-
hard interactions can also be quantitatively described and used to derived protein with CNS homology-associated protein 1; 2-
clarify toxicodynamics. Accordingly, we have employed HSAB ACP, 2-acetylcyclopentanone; AcAc, acetylacetone; TFPD,
calculations to describe the interactions of 2,5-hexanedione, the 1,1,1-trifluoro-2,4-pentanedione; DEM, diethylmalonate
hard electrophilic metabolite of the parent compound, n-hexane,
with possible hard nucleophilic targets in nervous tissue, e.g., ’ REFERENCES
ε-amino groups on lysine residues of axon cytoskeletal proteins.88
Similar to the n-hexane/2,5-hexanedione relationship, many (1) Cohen, S. D., Pumford, N. R., Khairallah, E. A., Boekelheide, K., Pohl,
L. R., Amouzadeh, H. R., and Hinson, J. A. (1997) Selective protein covalent
toxicants are metabolic derivatives of a less reactive parent or binding and target organ toxicity. Toxicol. Appl. Pharmacol. 143, 1–12.
protoxicant, e.g., desulfuration of the relatively soft electrophilic (2) Coles, B. (198485) Effects of modifying structure on electrophilic
chlorpyrifos parent to the highly reactive hard electrophilic oxon reactions with biological nucleophiles. Drug Met. Rev. 15, 1307–1334.
metabolite (chlorpyrifos oxon). Therefore, on a theoretical basis, (3) Hinson, J. A., and Roberts, D. W. (1992) Role of covalent and
HSAB calculations could also be used to determine the electro- noncovalent interactions in cell toxicity: effects on proteins. Annu. Rev.
philic reactivity and potential toxicity of metabolites derived from Pharmacol. Toxicol. 32, 471–510.

248 dx.doi.org/10.1021/tx2003257 |Chem. Res. Toxicol. 2012, 25, 239–251


Chemical Research in Toxicology PERSPECTIVE

(4) Liebler, D. C. (2008) Protein damage by reactive electrophiles: read-across for risk assessment: toxicity of α,β-unsaturated carbonyl
targets and consequences. Chem. Res. Toxicol. 21, 117–128. compounds. Chem. Res. Toxicol. 21, 2300–2312.
(5) LoPachin, R. M., Barber, D. S., and Gavin, T. (2008) Molecular (24) Chattaraj, P. K., Sarkar, U., and Roy, D. R. (2006) Electro-
mechanisms of the conjugated α,β-unsaturated carbonyl derivatives: philicity index. Chem. Rev. 106, 2065–2091.
relevance to neurotoxicity and neurodegenerative diseases. Toxicol. Sci. (25) Jaramillo, P., Periz, P., Contreras, R., Tiznada, W., and Fuentealba,
104, 235–249. P. (2006) Definition of a nucleophilicity scale. J. Phys. Chem. A 110,
(6) LoPachin, R. M., and DeCaprio, A. P. (2005) Protein adduct 8181–8187.
formation as a molecular mechanism in neurotoxicity. Toxicol. Sci. (26) Cronin, M. T. D., Manga, N., Seward, J. R., Sinks, G. D., and
86, 214–225. Schultz, T. W. (2001) Parametrization of electrophilicity for the
(7) Nelson, S. D., and Pearson, P. G. (1990) Covalent and non- prediction of the toxicity of aromatic compounds. Chem. Res. Toxicol.
covalent interactions in acute lethal cell injury caused by chemicals. 14, 1498–1505.
Annu. Rev. Pharmacol. Toxicol. 30, 169–195. (27) Geiss, K. T., and Frazier, J. M. (2001) QSAR modeling of
(8) Schultz, T. W., Carlson, R. E., Cronin, M. T. D., Hermens, oxidative stress in vitro following hepatocyte exposures to halogenated
J. L. M., Johnson, R., O’Brien, P. J., Roberts, D. W., Siraki, A., Wallace, methanes. Toxicol. in Vitro 15, 557–563.
K. B., and Veith, G. D. (2006) A conceptual framework for predicting the (28) Maynard, A. T., Huang, M., Rice, W. G., and Covell, D. G.
toxicity of reactive chemicals: modeling soft electrophilicity. SAR QSAR (1998) Reactivity of the HIV-1 nucleocapsid protein p7 zinc finger
Eviron. Res. 17, 413–428. domains from the perspective of density-functional theory. Proc. Natl.
(9) Schultz, T. W., Netzeva, T. I., Roberts, D. W., and Cronin, Acad. Sci. U.S.A. 95, 11578–11583.
M. T. D. (2005) Structure-toxicity relationships for the effects to Tetra- (29) Aptula, A. O., and Roberts, D. W. (2006) Mechanistic applic-
hymena pyriformis of aliphatic, carbonyl-containing α,β-unsaturated ability domains for nonanimal-based prediction of toxicological end
chemicals. Chem. Res. Toxicol. 18, 330–341. points: general principles and application to reactive toxicity. Chem. Res.
(10) Schwobel, J. A. H., Koleva, Y. K., Enoch, S. J., Bajot, F., Hewitt, Toxicol. 19, 1097–1105.
M., Madden, J. C., Roberts, D. W., Schultz, T. W., and Cronin, M. T. D. (30) Friedman, M. (2003) Chemistry, biochemistry and safety of
(2011) Measurement and estimation of electrophilic reactivity for acrylamide. A review. J. Agric. Food Chem. 51, 4504–4526.
predictive toxicology. Chem. Rev. 111, 2562–2596. (31) LoPachin, R. M., and Gavin, T. (2008) Acrylamide-induced
(11) Barber, D. S., and LoPachin, R. M. (2004) Proteomic analysis of nerve terminal damage: relevance to neurotoxic and neurodegenerative
acrylamide-protein adduct formation in rat brain synaptosomes. Toxicol. mechanisms. J. Agric. Food Chem. 56, 5994–6003.
Appl. Pharmacol. 201, 120–136. (32) Stedman, R. L. (1968) The chemical composition of tobacco
(12) Barber, D. S., Stevens, S., and LoPachin, R. M. (2007) and tobacco smoke. Chem. Rev. 68, 153–207.
Proteomic analysis of rat striatal synaptosomes during acrylamide (33) Parzefall, W. (2008) Minireview on the toxicity of dietary
intoxication at a low dose-rate. Toxicol. Sci. 100, 156–167. acrylamide. Food Chem. Toxicol. 46, 1360–1364.
(13) LoPachin, R. M., Schwarcz, A. I., Gaughan, C. L., Mansukhani, (34) Stevens, J. F., and Maier, C. S. (2008) Acrolein: sources,
S., and Das, S. (2004) In vivo and in vitro effects of acrylamide on metabolism and biomolecular interactions relevant to human health
synaptosomal neurotransmitter uptake and release. NeuroToxicology and disease. Mol. Nutr. Food Res. 52, 7–25.
25, 349–363. (35) Butterfield, D. A., Bader Lange, M. L., and Sultana, R. (2010)
(14) LoPachin, R. M., Barber, D. S., Geohagen, B. C., Gavin, T., He, Involvements of the lipid peroxidation product, HNE, in the patho-
D., and Das, S. (2007) Structure-toxicity analysis of Type-2 alkenes: in genesis and progression of Alzheimer’s disease. Biochim. Biophys. Acta
vitro neurotoxicity. Toxicol. Sci. 95, 136–146. 1801, 924–929.
(15) LoPachin, R. M., Gavin, T., Geohagen, B. C., and Das, S. (2007) (36) Singh, M., Arseneault, M., Sanderson, T., Murthy, V., and
Neurotoxic mechanisms of electrophilic type-2 alkenes: soft-soft inter- Ramassamy, C. (2008) Challenges for research on polyphenols from
actions described by quantum mechanical parameters. Toxicol. Sci. foods in Alzheimer’s disease: bioavailability, metabolism, and cellular
98, 561–570. and molecular mechanisms. J. Agric. Food Chem. 56, 4855–4873.
(16) LoPachin, R. M., Gavin, T., Geohagen, B. C., and Das, S. (2009) (37) Friedman, M., and Wall, J. S. (1966) Additive linear free-energy
Synaptosomal toxicity and nucleophilic targets of 4-hydroxy-2-nonenal. relationships in reaction kinetics of amino groups with α,β-unsaturated
Toxicol. Sci. 107, 171–181. compounds. J. Org. Chem. 31, 2888–2894.
(17) LoPachin, R. M., Gavin, T., Petersen, D. R., and Barber, D. S. (38) Seiner, D. R., LaButti, J. N., and Gates, K. S. (2007) Kinetics and
(2009) Molecular mechanisms of 4-hydroxy-2-nonenal and acrolein mechanism of protein tyrosine phosphatase 1B inactivation by acrolein.
toxicity: nucleophilic targets and adduct formation. Chem. Res. Toxicol. Chem. Res. Toxicol. 20, 1315–1320.
22, 1499–1508. (39) McGarth, C. E., Tallman, K. A., Proter, N. A., and Marnett,
(18) Martyniuk, C. J., Fang, B., Koomen, J. M., Gavin, T., LoPachin, L. J. (2011) Structure-activity analysis of diffusible lipid electrophiles
R. M. Barber, D. S. (2011) Molecular mechanisms of α,β-unsaturated associated with phospholipid peroxidation: 4-hydroxnonenal and
carbonyl toxicity: cysteine-adduct formation correlates with loss of 4-oxononenal analogues. Chem. Res. Toxicol. 24, 357–370.
enzyme function. Chem. Res. Toxicol. In press. (40) Doorn, J. A., and Petersen, D. R. (2002) Covalent modification
(19) Enoch, S. J., Cronin, M. T. D., Schultz, T. W., and Madden, J. C. of amino acid nucleophiles by the lipid peroxidation products 4-hydro-
(2008) Quantitative and mechanistic read across for predicting the skin xynonenal and 4-oxo-2-nonenal. Chem. Res. Toxicol. 15, 1445–1450.
sensitization potential of alkenes acting via michael addition. Chem. Res. (41) Doorn, J. A., and Petersen, D. R. (2003) Covalent adduction of
Toxicol. 21, 513–520. nucleophilic amino acids by 4-hydroxy-2-nonenal and 4-oxononenal.
(20) Schwobel, J. A. H., Wondrousch, D., Koleva, Y. K., Madden, Chem.-Biol. Interact. 143/144, 93–100.
J. C., Cronin, M. T. D., and Schuurmann, G. (2010) Prediction of (42) Tanii, H., and Hashimoto, K. (1982) Structure-toxicity rela-
Michael-type acceptor reactivity toward glutathione. Chem. Res. Toxicol. tionship of acrylates and methacrylates. Toxicol. Lett. 11, 125–129.
23, 1576–1585. (43) McCarthy, T. J., Hayes, E. P., Schwartz, C. S., and Witz, G.
(21) Pearson, R. G. (1990) Hard and soft acids and bases  the (1994) The reactivity of selected acrylate esters toward glutathione and
evolution of a chemical concept. Coord. Chem. Rev. 100, 403–425. deoxyribonucleosides in vitro: structure-activity relationships. Fundam.
(22) Wondrousch, D., Bohme, A., Thaens, D., Ost, N., and Schuurmann, Appl. Toxicol. 22, 543–548.
G. (2010) Local electrophilicity predicts the toxicity-relevant reactivity of (44) Bohme, A., Thaens, D., Pascke, A., and Schuurmann, G. (2009)
Michael acceptors. J. Phys. Chem. Lett. 1, 1605–1610. Kinetic glutathione chemoassay to quantify thiol reactivity of organic
(23) Koleva, Y. K., Madden, J. C., and Cronin, M. T. D. (2008) electrophiles  application to α,β-unsturated ketones, acrylates and
Formation of categories from structure-activity relationships to allow propiolates. Chem. Res. Toxicol. 22, 742–750.

249 dx.doi.org/10.1021/tx2003257 |Chem. Res. Toxicol. 2012, 25, 239–251


Chemical Research in Toxicology PERSPECTIVE

(45) Aldini, G., Gamberoni, L., Orioli, M., Beretta, G., Regazzoni, L., (65) LoPachin, R. M., Gavin, T., and Barber, D. S. (2008) Type-2
Facino, R. M., and Carini, M. (2006) Mass spectrometric characteriza- alkenes mediate synaptotoxicity in neurodegenerative diseases. Neuro-
tion of covalent modification of human serum albumin by 4-hydroxy- Toxicology 29, 871–882.
trans-2-nonenal. J. Mass Spec. 41, 1149–1161. (66) Beauchamp, R. O., Andjelkovich, D. A., Kligerman, A. D.,
(46) Friguet, B., Stadtman, E. R., and Szweda, L. I. (1994) Modifica- Morgan, K. T., and Heck, H.d’A. (1985) A critical review of the literature
tion of glucose-6-phophate dehydrogenase by 4-hydroxy-2-nonenal. on acrolein toxicity. Crit. Rev. Toxicol. 14, 309–380.
J. Biol. Chem. 269, 21639–21643. (67) Kehrer, J. P., and Biswal, S. S. (2000) The molecular effects of
(47) Liu, Z., Minkler, P. E., and Sayre, L. M. (2003) Mass spectro- acrolein. Toxicol. Sci. 57, 6–15.
scopic characterization of protein modification by 4-hydroxy-2-(E)- (68) Barber, D. S., Hunt, J. R., Ehrich, M., Lehning, E. J., and
nonenal. Chem. Res. Toxicol. 16, 901–911. LoPachin, R. M. (2001) Metabolism, toxicokinetics and hemoglobin
(48) Szweda, L. I., Uchida, K., Tsai, L., and Stadtman, E. R. (1993) adduct formation in rats following subacute and subchronic acrylamide
Inactivation of glucose-66phophate dehydrogenase by 4-hydroxy-2- dosing. NeuroToxicology 22, 341–353.
nonenal. J. Biol. Chem. 268, 3342–3347. (69) Miller, M. J., Carter, D. E., and Sipes, I. G. (1982) Pharmaco-
(49) Uchida, K., and Stadtman, E. R. (1992) Modification of kinetics of acrylamide in Fisher-334 rats. Toxicol. Appl. Pharmacol.
histidine residues in proteins by reaction with 4-hydroxynonenal. Proc. 63, 36–44.
Natl. Acad. Sci. U.S.A. 89, 4544–4548. (70) Green, M. A., and Egle, J. L. (1983) The effects of acetaldehyde
(50) Uchida, K., and Stadtman, E. R. (1993) Covalent attachment and acrolein on blood pressure in guanethidine-pretreated hypertensive
of 4-hydroxynonenal to glyceraldehyde-3-phosphate dehydrogenase. rats. Toxicol. Appl. Pharmacol. 69, 29–34.
J. Biol. Chem. 268, 6388–6393. (71) LoPachin, R. M., Ross, J. F., Reid, M. L., Dasgupta, S.,
(51) Uchida, K., Kanematsu, M., Morimitsu, Y., Noriko, T., Noguchis, Mansukhani, S., and Lehning, E. J. (2002) Neurological evaluation of
N., and Niki, E. (1998) Acrolein is product of lipid peroxidation reaction. toxic axonopathies in rats: acrylamide and 2,5-hexanedione. NeuroTox-
J. Biol. Chem. 273, 16058–16066. icology 23, 95–110.
(52) Uchida, K., Kanematsu, M., Sakai, K., Matsuda, T., Hattori, N., (72) Zhang, L., Gavin, T., Barber, D. S., and LoPachin, R. M. (2010)
Mizuno, Y., Suzuki, D., Miyata, T., Noguchi, N., and Niki, E. (1998) Protein- Role of the Nrf2-ARE pathway in acrylamide neurotoxicituy. Toxicol.
bound acrolein: Potential markers for oxidative stress. Proc. Natl. Acad. Sci. Lett. 205, 1–7.
U.S.A. 95, 4882–4887. (73) Calakos, N., and Scheller, R. H. (1996) Synaptic vesicle
(53) Ishii, T., Tatsuda, E., Kumazawa, S., Nakayama, T., and Uchida, biogenesis, docking and fusion: A molecular description. Physiol. Rev.
K. (2003) Molecular basis of enzyme inactivation by an endogenous 76, 1–29.
electrophile 4-hydroxy-2-nonenal: identification of modification sites in (74) Hatcher, H., Planalp, R., Cho, J., Torti, F. M., and Torti, S. V.
glyceraldehyde-3-phosphate dehydrogenase. Biochemistry 42, 3473–3480. (2008) Curcumin: from ancient medicine to current clinical trials. Cell.
(54) Aldini, G., Dalle-Donne, I., Vitoli, G., Facino, R. M., and Carini, Mol. Life Sci. 65, 1631–1652.
M. (2005) Covalent modification of actin by 4-hydroxy-trans-2-nonenal (75) Mandel, S., and Youdim, M. B. H. (2004) Catechin poly-
(HNE): LC-ESI-MS/MS evidence for Cys374 Michael adduction. phenols: neurodegeneration and neuroprotection in neurodegenerative
J. Mass Spec. 40, 946–954. diseases. Free Radical Biol. Med. 37, 304–317.
(55) Dalle-Donne, I., Vistoli, G., Gamberoni, L., Giustarini, D., (76) Lim, G. P., Chu, T., Yang, F., Beech, W., Frautschy, S. A., and
Colmbo, R., Facino, R. M., Rossi, R., Milzani, A., and Aldini, G. Cole, G. M. (2001) The curry spice curcumin reduces oxidative damage
(2007) Actin Cys374 as a nucleophilic target of α,β-unsaturated and amyloid pathology in an Alzheimer trangenic mouse. J. Neurosci.
aldehydes. Free Radical Biol. Med. 42, 583–598. 21, 8370–8377.
(56) LoPachin, R. M., Barber, D. S., He, D., and Das, S. (2006) (77) Liu, Y., Dargusch, R., Maher, P., and Schubert, D. (2008) A
Acrylamide inhibits dopamine uptake in rat striatal synaptic vesicles. broadly neuroprotective derivative of curcumin. J. Neurochem.
Toxicol. Sci. 89, 224–234. 105, 1336–1345.
(57) Van Iersel, M. L. P. S., Ploemen, J-P. H.T.M., LoBello, (78) Kim, J., Lee, H. J., and Lee, K. W. (2010) Naturally occurring
M., Federici, G., and van Bladeren, P. J. (1997) Interactions of phytochemicals for the prevention of Alzheimer’s disease. J. Neurochem.
α,β-unsaturated aldehydes and ketones with human glutathione 112, 1415–1430.
S-transferase P11. Chem-Biol. Interact. 108, 67–78. (79) Awasthi, S., Srivatava, S. K., Piper, J. T., Singhal, S. S., Chaubey,
(58) Friedman, M., Cavins, J. F., and Wall, J. S. (1965) Relative M., and Awasthi, Y. (1996) Curcumin protects against 4-hydroxy-2-
nucleophilic reactivities of amino groups and mercaptide ions in addition nonenal-induced cataract formation in rat lenses. Am. J. Clin. Nutr.
reactions with α,β-unsaturated compounds. J. Am. Chem. Soc. 87, 64, 761–766.
3672–3682. (80) Jiao, Y., Wilkinson, J., Pietsch, E. C., Buss, J. L., Wang, W.,
(59) LoPachin, R. M., and Barber, D. S. (2006) Synaptic cysteine Planalp, R., Torti, F. M., and Torti, S. V. (2006) Iron chelation in the
sulfhydryl groups as targets of electrophilic neurotoxicants. Toxicol. Sci. biological activity of curcumin. Free Radical Biol. Med. 40, 1152–1160.
94, 240–255. (81) Zhu, Q., Zheng, Z., Cheng, M K., Wu, J., Zhang, S., Tang, Y. S.,
(60) Cai, J., Bhatnagar, A., and Pierce, W. M. (2009) Protein Sze, K., Chen, J., Chen, F., and Wang, M. (2009) Natural polyphenols
modification by acrolein: formation and stability of cysteine adducts. as direct trapping agents of lipid peroxidation-derived acrolein and
Chem. Res. Toxicol. 22, 708–716. 4-hydroxy-2-nonenal. Chem. Res. Toxicol. 22, 1721–1729.
(61) Feng, Y., and Forgac, M. (1992) Cysteine 254 of the 73-kDa A (82) Begum, A. N., Jones, M. R., Lim, G. P., Morihara, T., Kim, P.,
subunit is responsible for inhibition of the coated vesicle (H+)-ATPase Heath, D. D., Rock, C. L., Pruitt, M. A., Yang, R., Hudspeth, B., Hu, S.,
upon modification by sulfhydryl reagents. J. Biol. Chem. 267, 5817–5822. Faull, K. F., Teter, B., Cole, G. M., and Frautschy, S. A. (2008) Curcumin
(62) Fritz, K. S., Galligan, J. J., Smathers, R. L., Roede, J. R., structure-function, bioavailability and efficacy in models of neuroin-
Shearn, C. T., Peigan, P., and Petersen, D. R. (2011) 4-Hydroxynonenal flammation and Alzheimer’s disease. J. Pharmacol. Exp. Ther. 326,
inhibits SIRT3 via thiol-specific modification. Chem. Res. Toxicol. 196–208.
24, 651–662. (83) Vajragupta, O., Boonchoong, P., Morris, G. M., and Olson, A. J.
(63) Mercer, W. D., Winn, S. I., and Watson, H. C. (1976) Twinning (2005) Active site binding modes of curcumin in HIV-1 protease and
in crystals of human skeletal muscle D-glyceraldehyde-3-phophate integrase. Bioorg. Med. Chem. Lett. 15, 3364–3368.
dehydrogenase. J. Mol. Biol. 104, 277–283. (84) Weber, W. M., Hunsaker, L. A., Gonzales, A. M., Heynekamp, J. J.,
(64) Thomas, J. A., Poland, B., and Hanzatka, R. (1995) Protein Orlando, R. A., Deck, L. M., and Vander Jagt, D. L. (2006) TPA-induced
sulfhydryls and their role in the antioxidant function of protein up-regulation of activator protein-1 can be inhibited or enhanced by analogs
S-thiolation. Arch. Biochem. Biophys. 319, 1–9. of the natural product curcumin. Biochem. Pharmacol. 72, 928–940.

250 dx.doi.org/10.1021/tx2003257 |Chem. Res. Toxicol. 2012, 25, 239–251


Chemical Research in Toxicology PERSPECTIVE

(85) Galati, G., and O’Brien, P. J. (2004) Potential toxicity of


flavonoids and other dietary phenolics: significance for their chemo-
preventive and anticancer properties. Free Radical Biol. Med. 37,
287–303.
(86) Lambert, S. D., Sang, S., and Yang, C. S. (2007) Possible
controversy over dietary polyphenols: benefits vs risks. Chem. Res.
Toxicol. 20, 583–585.
(87) LoPachin, R. M., Gavin, T., Geohagen, B. C., Zhang, L., Casper,
D., Lekhraj, R., and Barber, D. S. (2011) β-Dicarbonyl enolates: a new
class of neuroprotectants. J. Neurochem. 116, 132–143.
(88) Zhang, L., Gavin, T., DeCaprio, A. P., and LoPachin, R. M.
(2010) γ-Diketone axonopathy: analyses of cytoskeletal motors and
highways in CNS myelinated axon. Toxicol. Sci. 117, 180–189.

251 dx.doi.org/10.1021/tx2003257 |Chem. Res. Toxicol. 2012, 25, 239–251

You might also like