Subscriber access provided by GAZI UNIV
Review
Antioxidant activity/capacity measurement: I. Classification, physicochemical principles, mechanisms and electron transfer (ET)-based assays Re#at Apak, Mustafa Özyürek, Kubilay Guclu, and Esra Capanoglu J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.5b04739 • Publication Date (Web): 04 Jan 2016 Downloaded from http://pubs.acs.org on January 8, 2016
Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.
Journal of Agricultural and Food Chemistry is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.
Page 1 of 105
Journal of Agricultural and Food Chemistry
Antioxidant activity/capacity measurement: I. Classification, physicochemical principles, mechanisms and electron transfer (ET)-based assays
Reşat Apak1, Mustafa Özyürek1, Kubilay Güçlü1, Esra Çapanoğlu2
1
Department of Chemistry, Faculty of Engineering, Istanbul University, Avcilar 34320,
Istanbul-Turkey
2
Department of Food Engineering, Faculty of Chemical and Metallurgical Engineering,
Istanbul Technical University, Maslak 34469, Istanbul-Turkey
* Corresponding Author. Tel.: +90 212 4737070 Fax: +90 212 473 7180 E-mail:
[email protected] 1 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 2 of 105
ABSTRACT
1 2 3
Since there is no widely adopted “total antioxidant parameter” as a nutritional index for
4
labeling food and biological fluids, it is desirable to establish and standardize methods that
5
can measure the total antioxidant capacity (TAC) level directly from plant based food extracts
6
and biological fluids. In this review, we (i) present and classify the widely used analytical
7
approaches (e.g., in vitro and in vivo, enzymatic and non-enzymatic, electron transfer (ET)−
8
and hydrogen atom transfer (HAT)−based, direct and indirect assays) for evaluating
9
antioxidant capacity/activity; (ii) discuss total antioxidant capacity/activity assays in terms of
10
chemical kinetics and thermodynamics, reaction mechanisms, analytical performance
11
characteristics, together with advantages and drawbacks; (iii) critically evaluate electron
12
transfer (ET)-based methods for analytical, food chemical, biomedical/clinical and
13
environmental scientific communities so that they can effectively use these assays in their
14
correct places to meet their needs.
15 16
Keywords: Antioxidant activity; Total antioxidant capacity; Electron transfer-based
17
assays; Antioxidant mechanisms; Food analytical methods.
18 19
2 ACS Paragon Plus Environment
Page 3 of 105
20
Journal of Agricultural and Food Chemistry
1. INTRODUCTION
21 22
1.1. Scope
23 24
Oxidative stress is a pathological state in which reactive oxygen/nitrogen species (ROS/RNS)
25
overwhelm antioxidative defenses of the organism, leading to oxidative modification of
26
biological macromolecules (i.e., lipid, protein, DNA), tissue injury, and accelerated cellular
27
death1 as the foundation of many diseases. Measuring the antioxidant activity/capacity levels
28
of biological fluids and foods is carried out for the diagnosis and treatment of oxidative stress-
29
associated diseases in clinical biochemistry, for meaningful comparison of foods in regard to
30
their antioxidant content, and for controlling variations within or between products.
31
Antioxidant measurements have not yet been adapted to standard protocols in medical
32
diagnosis and treatment in spite of evidence-based changes of serum antioxidant capacity in
33
certain diseases (e.g., overall, total antioxidant capacity (TAC) of human serum was increased
34
in dialysis patients of chronic renal failure due to high urate values, but there was a marked
35
reduction after hemodialysis; antioxidant activity (AOA) of serum was markedly lower in
36
acute myocardial infarction and chronic lymphocytic leukaemia patients compared to those of
37
controls). So, ideally one should be able to detect/screen certain oxidative stress-originated
38
diseases and monitor the course of medical treatments by considering the changes in TAC
39
values of intracellular fluids and blood plasma/serum of a given individual measured by
40
standardized methods. The complexities of food and physiological applications of
41
antioxidants, separately and combined, require rigorous consideration and analysis of all
42
aspects of the chemistry, reaction mechanisms, and reaction/radical/target specificity in
43
various test systems, as well as careful and accurate quantitation of all reactants and products
44
involved. Current literature clearly state that there is no widely adopted “total antioxidant
3 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 4 of 105
45
parameter” as a nutritional index available for the labeling of food and biological fluids due to
46
the lack of standardized quantitation methods. Therefore it is desirable to establish and
47
standardize methods that can measure the TAC level directly from plant based food extracts
48
and biological fluids. Ideally, agreement on standardized methods for antioxidant testing (i)
49
requires general guidance for the application of assays, including exact description of test
50
conditions, experimental apparatus and stability of reagents; (ii) all new methods need to be
51
properly validated within their framework, associated with intra- and inter-laboratory
52
validation (including the production of repeatability, reproducibility and recovery data),
53
standardization, internal quality control, proficiency testing and analytical quality assurance;
54
(iii) the results should enable a reasonable comparison of the antioxidant content of foods,
55
pharmaceuticals and other commercial products; (iv) it may provide a measure for meeting
56
the need of quality standards for regulatory issues and health claims.2,3
57
To date, many in vitro tests are available from a chemical assay performed in a
58
homogenous solution to more complex methods using exogenic probes to detect oxidation.
59
Although many literature methods do not distinguish between ‘antioxidant capacity’ and
60
‘antioxidant activity’ of non-enzymatic antioxidants, end-point assays measuring the
61
efficiency of antioxidant action (i.e. reactive species inactivation) should be differentiated
62
from kinetic-based assays measuring reaction rate. TAC assays may be broadly classified as
63
electron transfer (ET)− and hydrogen atom transfer (HAT)−based assays, though in some
64
cases, these two mechanisms may not be differentiated with distinct boundaries (such as those
65
utilizing 2,2’-azino-bis(3-ethylbenzothiazoline-6-sulfonic acid (ABTS) and 2,2-di(4-tert-
66
octylphenyl)-1-picrylhydrazyl (DPPH) radical reagents). In fact, most non-enzymatic
67
antioxidant activity (e.g., scavenging of free radicals, inhibition of lipid peroxidation, etc.) is
68
mediated by redox reactions. In addition to these two basic classes focusing on mechanism,
69
ROS/RNS scavenging assays will also be taken into account. Complementary to existing
4 ACS Paragon Plus Environment
Page 5 of 105
Journal of Agricultural and Food Chemistry
70
methods, novel approaches have recently been developed such as CUPric Reducing
71
Antioxidant Capacity (CUPRAC) TAC assay (introduced by our research group to world
72
literature in 2004), its modified ROS scavenging assays and other modifications (e.g.,
73
antioxidant sensor, postcolumn online HPLC technology). The current direction of CUPRAC
74
methodology can be best described as a self-sufficient and integrated train of measurements
75
providing a useful “antioxidant and antiradical assay package”.
76
The complexity and diversity of research topics investigated have led to the
77
development of a multitude of tests, but unfortunately none has gained universal acceptance.
78
Thus, one of the major challenges in antioxidant testing is to know which method is best
79
suited for a particular application. Since antioxidants may exert their effect through various
80
mechanisms such as scavenging radicals, sequestering transition metal ions, decomposing
81
hydrogen peroxide or hydroperoxides, quenching active prooxidants, and repairing biological
82
damage, it should be clarified which function of antioxidants is being measured, and
83
accordingly, the antioxidant assay method should be selected considering the function to be
84
evaluated.4 Because of the wide divergence of results for natural antioxidants in food systems,
85
more valid and rigorous guidelines and assay protocols are required to bring some order and
86
agreement to this important field. Naturally, it should be remembered that TAC and AOA are
87
not like elemental analysis parameters for which the analyst has to obtain more or less the
88
same result from different techniques (e.g., calcium in a milk sample measured by different
89
techniques should yield identical results within tolerable limits), because it is possible to get
90
quite different TAC or AOA results using the same probe under different experimental
91
conditions. Our understanding of the effects of antioxidant compounds can only be improved
92
with a deeper chemical insight if more specific methodology is used and is capable of
93
defining what products are formed and inhibited by antioxidants depending on conditions,
94
systems, and targets of protection.
5 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
95 96 97
Page 6 of 105
1.2. Technical Issues with TAC Assays
98 99 100
The technical issues associated with current ‘state-of-the-art’ TAC assays that require special consideration can be summarized as follows:
101
i. Almost total lack of standardization in experimental procedures and expressing results.
102
ii. Too many assays not having a demonstrated clear chemistry give rise to inconsistent
103
results, inappropriate application and interpretation of assays, and improper specification of
104
TAC. Naturally, statistical validation (e.g., using a certified reference material) of a given
105
assay is not possible in view of the different reaction kinetics and chemistry of the assays
106
producing different results.
107
iii. Most assays developed for TAC screening are devoid of detailed investigations related
108
to underlying initiators, targets, antioxidant interactions, kinetics, effects of solvent,
109
concentration and pH, etc. In earlier work regarding TAC assays, there seems to be some lack
110
of relevant classification according to chemical reaction mechanisms and kinetics. Molecular
111
accessibility problems of relevant reagents (such as steric hindrance, inner- and outer-sphere
112
electron transfer mechanisms of transition metal−based chromophores/fluorophores, etc.)
113
have not been discussed.
114
iv. Several studies on natural phytochemical compounds produced conflicting results
115
because of the non-specific “one-dimensional” character of methods used to evaluate
116
antioxidant activity. Since most natural antioxidants and phytochemicals are multifunctional
117
(i.e. due to variations in system composition, type of oxidizable substrate, media of initiation
118
and acceleration of oxidation, methods to assess oxidation and to quantify antioxidant
119
activity), a reliable antioxidant protocol requires the measurement of more than one property
120
relevant to either foods or biological systems.5
6 ACS Paragon Plus Environment
Page 7 of 105
Journal of Agricultural and Food Chemistry
121
v. A noteworthy question is that, if the protective action of an antioxidant is being
122
assayed in a selected test, what biomolecule (lipid, protein, DNA, etc.) is the antioxidant
123
supposed to protect, and how? Is the tested antioxidant at a significantly lower concentration
124
than that of the protected substrate? Are relevant ROS/RNS utilized in assaying the
125
antioxidant?6 In AOA and TAC assays covering a simulated oxidant-antioxidant reaction,
126
chromophore or luminescent probes capable of accepting H-atoms or electrons from
127
antioxidants do not necessarily protect relevant substrates (such as lipid, protein, DNA) from
128
oxidation, and therefore, these TAC assays results may not consequently reflect the capacity
129
to retard or suppress oxidation.7
130
vi. Antioxidant activity (i.e. related to the kinetics of antioxidant action for quenching
131
reactive species, usually expressed as reaction rates or scavenging percentages per unit time)
132
and antioxidant capacity (i.e. thermodynamic conversion efficiency of reactive species by
133
antioxidants, such as the number of moles of reactive species scavenged by one mole of
134
antioxidant during a fixed time period) are both important in antioxidant research, and care
135
must be exercised to distinguish between these two terms which are often used
136
interchangeably and therefore confused.8 Since the term ‘total antioxidant capacity’ (TAC) is
137
indicative of the collaborative (additive and possibly synergistic/antagonistic) action of all
138
antioxidants present in a complex sample, it is considered by most researchers as a more
139
useful parameter to assess the antioxidative defenses in food or plasma than separately
140
determining the concentrations of individual antioxidant constituents.
141
vii. Other than reactivity toward ROS/RNS, several factors such as concentration,
142
distribution, localization, fate of antioxidant−derived radical, interaction with other
143
antioxidants, and metabolism should be evaluated.4 The question of bioavailability and the
144
fate of metabolites of the antioxidant components (e.g., whether they undergo further redox
145
cycles) must be addressed in the case of in vivo assays.9 Also, in vitro antioxidant assays
7 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 8 of 105
146
carried out at unrealistic pH values (i.e. far from physiological pH, either in alkaline or acidic
147
range) cannot have much meaning for in vivo estimations of antioxidant action. In tests
148
carried out at weakly-acidic pH, most oxidation reactions for phenolic antioxidants are
149
incomplete within the protocol time of the assay.
150
viii. Due to the inadequacies in analytical methodology, the antioxidant activity of
151
proteins (i.e. first line defense elements against ROS attack in the human body) and
152
specifically of protein thiols are often ignored in most antioxidant assays, because proteins are
153
separated by precipitation from the main matrix and their contribution to serum TAC is left
154
unmeasured.10 Thus, the precise determination of serum TAC incorporating the “antioxidant
155
gap” originating from protein components is believed to be potentially useful to biochemists
156
investigating the diagnosis, treatment, prognosis, and follow-up of diseases utilizing TAC
157
measurement.11
158
ix.
Possible interactions of antioxidants among themselves (i.e. synergistic or
159
antagonistic effects) should be taken into account so as to better visualize the collaborative
160
action of antioxidants in the organism.
161
x.
Prooxidant effects of antioxidants, especially dependent on the composition of
162
medium in which antioxidant measurements are made, should also be considered, e.g.,
163
transition metal (especially iron) complex−based TAC assay reagents have the possibility of
164
redox cycling at their lower oxidation states that may falsify antioxidant test results.12
165
xi.
Antioxidant capacity assay results have not been effectively correlated to tests
166
measuring oxidative damage, such as thiobarbituric acid−reactive substances (TBARS) test in
167
lipids or carbonyl test in proteins. Normally, the difference in oxidative status of a medium
168
undergoing ROS/RNS attack, to which a food or biological extract was added, should
169
correlate to TAC of the extract under investigation.13 Additionally, antioxidant activity needs
170
to be correlated to electrochemical behaviour of antioxidants.
8 ACS Paragon Plus Environment
Page 9 of 105
Journal of Agricultural and Food Chemistry
171
As required parameters, a standardized TAC method; (i) utilizes a suitable oxidant
172
which actually is or a simulator of a biologically relevant reactive species; (ii) is simple,
173
practical and versatile; (iii) uses a method with a defined end-point and a clear chemical
174
mechanism; (iv) has readily available and preferably low-cost instrumentation; (v) is
175
reproducible with good within-run and between-run precision; (vi) can assay both hydrophilic
176
and lipophilic antioxidants; (vii) does not generate new reactive species that may cause under-
177
or over-estimated TAC readings; (viii) is suitable for “high-throughput” analysis for routine
178
quality control of food and biological extracts.2,14
179 180
1.3. Purpose
181 182
The aims of these comprehensive review series are; (i) to present a brief panorama of the most
183
widely used methods and of new analytical approaches for evaluating antioxidant
184
capacity/activity; (ii) to discuss TAC/AOA assays in terms of chemical kinetics and
185
thermodynamics, reaction mechanisms, analytical performance characteristics (linear
186
concentration range, recovery, repeatability, reproducibility, and recognition of interfering
187
substances, etc.), and advantages and drawbacks; (iii) to bring in terms of definitions or
188
definition-like characterization and classification of the chemical and biochemical methods of
189
antioxidant assays as well as related antioxidant chemistry; and finally (iv) to provide a
190
critical evaluation of this topic to analytical, food chemical, biomedical/clinical and
191
environmental scientific communities so that they can effectively use these assays in their
192
correct places to meet their needs. The literature gap summarized in the above ‘Technical
193
issues with TAC assays’ is endeavoured to be filled. The basic criteria of classification of
194
antioxidant assays, such as in vitro and in vivo, enzymatic and non-enzymatic, electron
195
transfer (ET)− and hydrogen atom transfer (HAT)−based, direct and indirect assays have been
9 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 10 of 105
196
addressed. The essential ET−based assays with possible advantages/drawbacks have been
197
critically evaluated, and at the same time, various methodologies of the main and modified
198
CUPRAC procedures regarding TAC and ROS/RNS-scavenging assays have been unified and
199
summarized. However in the first review of this series, the stress will be upon ET assays, and
200
HAT and mixed-mode (ET/HAT) assays, together with lipid peroxidation assays, ROS/RNS
201
scavenging assays and oxidative stress biomarkers are the subjects of the following reviews.
202 203 204
2. CLASSIFICATION OF AOA AND TAC ASSAYS
205 206
‘Antioxidants’ are natural or synthetic substances that may prevent or delay oxidative cell
207
damage caused by physiological ‘oxidants’ having distinctly positive reduction potentials,
208
covering ROS/RNS and free radicals (i.e. unstable molecules or ions having unpaired
209
electrons). The terms ‘oxidant’ and ‘antioxidant’ have complementary meanings in the sense
210
that these compounds neutralize the effects of each other. On the other hand, a ‘prooxidant’
211
may not have a large reduction potential by itself, but may induce oxidative damage to
212
various biological targets like DNA (e.g., nucleic base modification and single/double strand
213
breaks), lipids (e.g., structural changes in fatty acid composition and lipid peroxidation) and
214
proteins (e.g., protein carbonylation and oxidation of certain amino acid moieties).15 For
215
example, transition metal ions at their lower oxidation states are not oxidant species by
216
themselves, but may induce the generation of ROS/RNS with hydrogen peroxide or molecular
217
oxygen, thereby acting as prooxidants. Antioxidants may be broadly defined as ‘substances
218
that, when present at relatively low concentrations compared with those of the oxidizable
219
substrates, significantly delay or inhibit oxidation of those substrates’.16,17 Although the term
220
‘oxidizable substrate’ includes every type of molecule found in vivo, it is generally
221
understood as biomacromolecules like lipid, protein and DNA. This definition emphasizes the
10 ACS Paragon Plus Environment
Page 11 of 105
Journal of Agricultural and Food Chemistry
222
importance of the selected damage target and the source of ROS/RNS used when antioxidants
223
are tested and their actions are examined.18 However, Finley et al.19 points out to the
224
ambiguity of this definition in that the term “antioxidant” may have different connotations to
225
different audiences, such as the capability of quenching metabolically generated ROS (to
226
biochemists and nutritionists), the functionality of retarding food oxidation (to food
227
scientists), or the property of yielding high TAC values in ET− and HAT−based in vitro
228
assays (presumably to a wider community of food science, commerce and industry). For
229
convenience, antioxidants have been traditionally divided into two classes; primary or chain-
230
breaking antioxidants (mainly acting by ROS/RNS scavenging), and secondary or
231
preventative antioxidants (usually acting by transition metal ion chelation).20 Thus, an
232
antioxidant may act directly by scavenging reactive species, or inhibiting their generation. It
233
may also act indirectly, e.g., by up-regulating endogenous antioxidant defenses.6,21 This work
234
is concerned with the corresponding methods of antioxidant capacity/activity assay capable of
235
measuring chain-breaking or preventive antioxidant ability. In lipid peroxidation experiments,
236
antioxidants acting only by metal chelation are essentially maintained, whereas chain-
237
breaking antioxidants are consumed. Most of the time, this difference is reflected in a lag
238
(retardation) phase of the peroxidation process by chain-breaking antioxidants, compared to a
239
constant inhibition by metal-binding antioxidants throughout the reaction.6,21 However, the
240
scope of this review is limited to the integrated activity of non-enzymatic antioxidants,
241
meaning that endogenous antioxidative enzymes such as superoxide dismutase (scavenging
242
superoxide anion radicals), catalase and glutathione peroxidase (able to remove hydrogen
243
peroxide) are not considered.
244
Chain-breaking mechanisms regarding the breaking of the oxidation chain of lipid
245
radicals (L•, LOO•, or LO•) involve the sacrificial consumption of antioxidants (AH) to
246
produce antioxidant radicals (A•) protecting lipid molecules (L):
11 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
247
L• + AH → LH + A• … (Eq. 1)
248
LOO• + AH → LOOH + A•. … (Eq. 2)
249
LO• + AH → LOH + A• … (Eq. 3)
Page 12 of 105
250
Thus, radical initiation (by reacting with a lipid radical) or propagation (by reacting
251
with lipid peroxyl or alkoxyl radicals) steps are inhibited. Since chain-breaking antioxidants
252
exert their action through either hydrogen atom (H•) and electron (e-) donation or both (i.e.
253
proton-coupled electron transfer), such AOA measurement methods are commonly classified
254
as hydrogen atom transfer (HAT)− and electron transfer (ET)−based assays according to
255
mechanism.
256
On the other hand, secondary (or preventive) antioxidants retard or prevent lipid
257
oxidation. For example, transition metal ion (e.g., Fe(II) or Cu(I)) chelator antioxidants may
258
inhibit Fenton-type reactions that produce hydroxyl radicals which may cause oxidative
259
degradation of biological macromolecules (lipids, proteins, DNA, etc.):
260
Fe(II) + H2O2 → Fe(III) + •OH + OH- … (Eq. 4)
261
Cu(I) + H2O2 → Cu(II) + •OH + OH- … (Eq. 5)
262 263
Therefore, preventive antioxidant assay methods should measure transition metal ion chelating ability.
264
The transition metal chelation functionality of antioxidants covers a neutralization
265
reaction between a Lewis base (antioxidant) and a Lewis acid (metal ion), without involving
266
the donation of H-atoms or electrons by the antioxidant. Therefore, the measurement of such
267
preventive AOA should be covered under a different category. On the other hand, hindering
268
the formation of ROS/RNS should be considered as a preventive action,7 while scavenging of
269
ROS/RNS is closely related to HAT and ET mechanisms for the measurement of chain-
270
breaking AOA. Nevertheless, techniques for the measurement of scavenging activity of
271
reactive species (e.g., ROS essentially cover hydroxyl radical: •OH, superoxide anion radical:
12 ACS Paragon Plus Environment
Page 13 of 105
Journal of Agricultural and Food Chemistry
272
O2•-, singlet oxygen: 1O2, hydrogen peroxide: H2O2, and hypochlorous acid (HOCl); RNS
273
mainly cover peroxynitrite: ONOO-, and nitric oxide: •NO) are investigated under a separate
274
subtitle.
275
Enzymatic antioxidants are either reductase enzymes (e.g., superoxide dismutase
276
(SOD), glutathione peroxidase (GSH-Px), glutathione reductase (GSH-Rx), and catalase) and
277
their cofactors, which limit the cellular concentration of free radicals and prevent excessive
278
oxidative damage, or oxidase enzyme inhibitors. Therefore, the corresponding enzymatic
279
AOA assays should either measure the enzymatic reduction ability of ROS or the inhibition of
280
oxidases (e.g., xanthine oxidase, NADPH oxidase) capable of producing reactive species. On
281
the other hand, non-enzymatic AOA/TAC assays utilize a relevant probe for simulating the
282
antioxidative action toward oxidant species. Since antioxidant and oxidant-regenerating
283
enzymes in blood cells and the blood vessel wall have a profound impact on the antioxidant
284
properties of blood plasma (which is not reflected in the in vitro assays of isolated plasma),
285
the term ‘TAC’ has been claimed to measure only a part of antioxidant capacity, usually
286
excluding enzymatic activities, and therefore 'non-enzymatic antioxidant capacity' (NEAC)
287
has been suggested as a more relevant term than TAC.22 For example, non-enzymatic plasma
288
antioxidants usually cover albumin and other related proteins containing thiols and other
289
antioxidative amino acid residues, as well as small molecules like α-tocopherol, bilirubin,
290
ascorbic acid, uric acid, and reduced glutathione (GSH). In the literature, much more effort
291
has been spent on developing non-enzymatic antioxidant assays covering a wide range of
292
HAT− and ET−based assays, and methods for measuring ROS/RNS scavenging activity.
293
Recently, synthetic and natural phenolic antioxidants have been summarized, together with
294
their mode of action, health effects, degradation products and toxicology.23 A general and
295
update overview of methods available for measuring antioxidant activity and the chemistry
13 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 14 of 105
296
behind them has been provided.24 Advantages and limitations of common testing methods
297
have been listed, with a preference for method selection serving different needs.25
298 299
2.1. Measurement of non-Enzymatic Chain-Breaking Antioxidant Activity/Capacity
300 301
Non-enzymatic chain-breaking antioxidant ability can be measured by finding the rate (kinetic
302
AOA methods) or thermodynamic conversion efficiency (TAC methods) for the reaction of a
303
suitable oxidant probe with the antioxidant. AOA assays like TRAP (total peroxyl radical
304
trapping antioxidant parameter),10,26 crocin bleaching,27,28 ORAC (oxygen radical absorbance
305
capacity),29,30 TOSC (total oxyradical scavenging capacity),31,32 etc. are usually competitive
306
(Figure 1) and work on a HAT mechanism, whereas TAC measurement methods are usually
307
non-competitive (Figure 2) ET and mixed-mode (ET/HAT) assays. In a competitive
308
inhibition (or scavenging) assay, the oxidant reacts with a probe leading to changes in its
309
absorbance, fluorescence, luminescence, or any other measurable property, where
310
antioxidants compete with the probe for the oxidant and repair the oxidized probe.33 Due to
311
the competition between the probe and antioxidants for reactive species, the probe undergoes
312
less oxidative conversion by ROS/RNS in the presence of antioxidants. A major criticism
313
directed at HAT-based competitive assays using fluorescent probes is that the concentration
314
of the target species (i.e. assay probe simulating a biological substrate) is usually smaller than
315
that of tested antioxidants,14 which contradicts with the basic ‘definition of antioxidant’,
316
because in real life, antioxidants exert their protective effects even when they are at much
317
lower concentration than that of the biological (oxidizable) substrate.15 On the other hand, in
318
ET−based assays, the probe undergoing reduction with the antioxidant is either converted to a
319
colored, fluorescent, chemiluminescent etc. species, or the initial absorbance/fluorescence is
320
attenuated as a result of the antioxidation reaction. ET−based assays have been criticised
14 ACS Paragon Plus Environment
Page 15 of 105
Journal of Agricultural and Food Chemistry
321
mainly because the utilized probe acting as the oxidizing agent does not involve
322
physiologically important oxidants like ROS/RNS, causing a discrepancy between the
323
simulated assay and the real-life antioxidant action.
324
Figure 1
325
Figure 2
326 327
The most widely used chromophores in ET−based spectrophotometric TAC assays of (2,2’-azino-bis(3-ethylbenzothiazoline-6-sulfonic
328
Folin-Ciocalteu,34,35
329
acid/Trolox®-equivalent antioxidant capacity),36,37 DPPH (2,2-di(4-tert-octylphenyl)-1-
330
picrylhydrazyl),38,39 CUPRAC (cupric reducing antioxidant capacity),40,41 FRAP (ferric
331
reducing antioxidant power),42-45 ferricyanide,46,47 ferric-phenanthroline,48 and ferric-
332
ferrozine49 assays are phospho-tungsto-molybdate(V), ABTS•+ radical cation, DPPH• radical,
333
cuprous neocuproine: [Cu(Nc)2]+ chelate, ferrous tripyridyltriazine: [Fe(TPTZ)3]2+ chelate,
334
Prussian blue: K[Fe(Fe(CN)6] heteropoly acid salt, ferrous phenanthroline: [Fe(phen)3]2+
335
chelate, and [Fe(FZ)3]2+ chelate, respectively. When transition metal ion−based probes such
336
as ferric and cupric chelates oxidize the to-be-assayed antioxidant compounds, the metal ion
337
placed at the coordination center of the complex is reduced to the lower oxidation number and
338
its chelate has strong intermolecular charge-transfer interactions, i.e. visible light absorbed by
339
the chelate causes a partial transfer of electronic charge from the metal ion to the ligand
340
(giving rise to very high molar extinction coefficients for these chromophores, thereby raising
341
the sensitivity for antioxidant determinations). Some authors classify DPPH and ABTS tests
342
as mixed-mode (i.e. having both ET- and HAT mechanisms) assays.2 TAC assays using
343
Cr(VI) as chromate50 and Ce(IV)51-53 as oxidant are also ET-based methods where
344
antioxidants reduce these species to Cr(III) and Ce(III), respectively; naturally, certain
345
measures have to be taken to decrease the oxidizing ability of these reagents so that
ABTS/TEAC
15 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 16 of 105
346
antioxidants but not other organic compounds (e.g., citric acid and common sugars in food
347
and beverages) are oxidized in the assays. When gold nanoparticles (Au-NPs) generated from
348
HAuCl4 upon reduction with phenolic acid antioxidants are used as colored probes exhibiting
349
localized surface plasmon resonance (LSPR) absorption, the highest capacity of reducing
350
gold(III) to elemental gold nanoparticles corresponds to the highest antioxidant activity,
351
consistent with the tendency of phenolic antioxidants to donate electrons.54 The same
352
reasoning applies for the formation of silver nanoparticles (Ag-NPs) coatings generated from
353
AgNO3 with phenolic antioxidants onto citrate-stabilized silver seeds, where the enlargement
354
of preformed Ag-NPs gave rise to enhanced LSPR absorption linearly dependent on
355
polyphenol concentration.55
356 357
2.1.1. HAT−Based Methods
358 359
HAT−based assays measure the capability of an antioxidant to quench free radicals (generally
360
peroxyl radicals) by H-atom donation. Peroxyl radicals are generally chosen as the reactive
361
species in these assays because of their higher biological relevance and longer half-life
362
(compared to hydroxyl and superoxide radicals). The HAT mechanism of antioxidant action,
363
in which the hydrogen atom (H•) of a phenol (Ar-OH) is transfered to an ROO• radical, can be
364
summarized by the reaction:
365
ROO• + AH/ArOH → ROOH + A• / ArO• … (Eq. 6)
366
where the aryloxy radical (ArO•) formed from the reaction of antioxidant phenol with peroxyl
367
radical is usually stabilized by resonance. The AH and ArOH species denote the protected
368
biomolecules and antioxidants, respectively. Effective phenolic antioxidants need to react
369
faster than biomolecules with free radicals to protect the latter from oxidation.56 As major
370
criticisms directed at HAT–based assays having a ‘lag-phase’ approach (i.e. measuring the
371
retardation time for the initiation of oxidative probe conversion) to quantification of 16 ACS Paragon Plus Environment
Page 17 of 105
Journal of Agricultural and Food Chemistry
372
antioxidant capacity; (i) not every antioxidant possesses an apparent lag-phase (i.e. in order to
373
have a distinct lag-time, an antioxidant should have a rate constant for the tested radical much
374
higher than that of the probe such that the antioxidant should be consumed up by the time
375
when the probe oxidation is observable),22 (ii) ambiguity in end-point observation makes
376
inter-laboratory comparison of generated data difficult, and (iii) the antioxidant capacity
377
profile of samples following the lag-phase is disregarded.15 Since in HAT–based antioxidant
378
assays using a fluorescent probe, both the probe and antioxidants react with ROO•, the
379
antioxidant activity can be determined from competition kinetics (Figure 1) by measuring the
380
fluorescence decay curve of the probe in the absence and presence of antioxidants, and
381
integrating the area under these curves (AUC approach). The AUC difference between a
382
sample and a reagent blank is then related to antioxidant concentration in the sample.2,14
383
2.1.2. ET−Based Methods
384 385
The ET mechanism of antioxidant action with a biologically relevant radical is based on the
386
reactions:
387 388 389 390 391 392 393
ROO• + AH/ArOH → ROO- + AH•+/ArOH•+ … (Eq. 13)
394
assays (though for metal-complex probes capable of outer-sphere electron transfer, this
395
assumption was shown to be invalid as a function of solvent type), and are both solvent– and
396
pH–dependent. The pH-dependency is apparent from the above reaction sequence of ET
397
mechanism. For example, phenolic compounds (Ar-OH) having weakly acidic -OH groups
398
dissociate to a greater extent at higher pH, and become more susceptible to oxidation. Thus,
399
most ET reactions occur at a higher rate at higher pH. The reactivity of the ABTS•+ cation
AH•+/ArOH•+ + H2O ↔ A• / ArO• + H3O+
… (Eq. 14)
ROO- + H3O+ ↔ ROOH + H2O … (Eq. 15) where the reactions are generally assumed to be relatively slower than those of HAT– based
17 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 18 of 105
400
radical toward ascorbic acid at neutral pH is expressed with a second-order rate constant of
401
8x106 M-1s-1, whereas in acidic pH, this rate constant decreases by almost two orders-of-
402
magnitude.57 Generally, ionization potentials of phenolic antioxidants decrease with
403
increasing pH, which causes an increase in electron-donating capacity concomitant with
404
deprotonation.2 Iron(III)-based antioxidant assays conforming to ET mechanism (excluding
405
the sole hexacyanoferrate(III) complex not in combination with ferric ion) have to be carried
406
out at an acidic pH in order to prevent the hydrolysis of trivalent iron (e.g., giving rise to
407
FeOH2+ and further hydrolysis complexes).48 The aryloxy radical (ArO•) is subsequently
408
oxidized to the corresponding quinone (Ar=O). The more stabilized the aryloxy radical is, the
409
easier will be the oxidation from ArOH to Ar=O due to reduced redox potential,56 and the
410
stronger is the antioxidant.
411
Electron transfer (ET)‒based (or reduction‒based) assays can be better understood
412
considering the fact that antioxidants are also good reducing agents capable of reductive
413
quenching of ROS/RNS. However, these assays do not necessarily use biologically relevant
414
reactive species (such as peroxyl radicals); instead they use artificial probes that change color
415
or fluorescence when reduced by antioxidants (Figure 2).14,56 For example, the cupric- or
416
ferric-reducing ability measured for a biological sample may indirectly but efficiently reflect
417
the total antioxidant power of the sample even though no radical species are involved in the
418
assay. The change in absorbance or fluorescence of the probe (at a pre-specified wavelength)
419
upon reduction with antioxidants is a measure of the total concentration of antioxidants in a
420
sample, or TAC. This TAC is usually expressed in terms of a reference compound, such as
421
Trolox for hydrophilic antioxidants, α-tocopherol (vitamin E) for lipophilic antioxidants, and
422
gallic acid for aqueous solutions of polyphenols, where trolox-equivalent and gallic acid-
423
equivalent TAC values are denoted as TE and GAE, respectively. The TEAC coefficient is a
424
unitless value, defined as the reducing potency −in Trolox® mM equivalents− of 1 mM
18 ACS Paragon Plus Environment
Page 19 of 105
Journal of Agricultural and Food Chemistry
425
antioxidant solution under investigation. Usually in spectrophotometric TAC assays, this
426
TEAC coefficient is found from the ratio of the slope of the calibration curve (drawn as
427
absorbance versus concentration) of the tested compound to that of Trolox obtained under
428
identical conditions. The TAC value, in mM-TE units, of a complex antioxidant mixture
429
(comprised of antioxidants: 1,2,3,…, i, …n) is the sum of the products obtained by
430
multiplying the mM concentration (Ci) of each antioxidant with its TEAC coefficient
431
(TEACi): TAC(mixture) = ΣCi(TEAC)i
432
Naturally this equation is valid as long as the principle of additivity of absorbances is
433
retained for a complex mixture conforming to Beer’s law of spectrophotometry, i.e. the TAC
434
of a complex antioxidant mixture is the sum of the TAC values of individual antioxidants
435
comprising the mixture.
436
In general, ET–based TAC assays have good precision, because the difference in
437
absorbance or fluorescence intensities of the reduced and original probes can be directly
438
measured. Since most of the time a single reduced species is produced from the probe upon
439
chemical reduction with antioxidants, the absorbance changes of a single chromogenic
440
product at a fixed pH and wavelength usually vary almost perfectly linearly with total
441
antioxidant concentration (in TE units). Thus, within the linear concentration range obeying
442
Beer’s law, additivity of absorbances (and therefore additivity of TAC values of individual
443
constituents forming an antioxidant mixture) is usually attained. Reduction–based assays have
444
generally been criticised for not involving biologically relevant radicals. However, their
445
simulated conditions can well imitate the media in food and biological fluids such as redox
446
potential, pH, and lipophilic/hydrophilic balance of solvent mixtures and microemulsions.
447
With the exception of the Folin-Ciocalteu reagent having an indefinite redox potential, most
448
ET reagents have standard potentials in the useful range of 0.6-0.7 V relevant for food and
449
biological fluids, i.e. they can oxidize important antioxidants to measure their TAC values.
19 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 20 of 105
450
The reaction conditions of reduction-based TAC assays should be well defined so as to
451
maintain reproducibility, since inter-assay results without detailed description of conditions
452
cannot be compared. The time and temperature of measurements should always be indicated
453
in regard to repeatability of reaction kinetics, because some oxidation reactions of ET probes
454
with antioxidants may not reach saturation within the pre-specified protocol time of the
455
assays. For example, high-spin Fe(III) having half-filled d-orbitals may show a kinetic
456
inertness to thiols, some phenolic acids and flavonoids such that the envisaged oxidation
457
cannot be completed within the protocol time period of the FRAP assay and similar Fe(III)-
458
based assays.48
459
When ET-based reagents, especially metal complexes utilizing Fe(III) or Cu(II), are to
460
be used in the TAC determination of plasma or serum, it should be remembered that these
461
samples should be preserved in the cold with either heparin or citrate and not EDTA, because
462
EDTA usually stabilizes the higher oxidation state (such as Fe(III) of Cu(II)) in preference to
463
the lower one (e.g., Fe(II) or Cu(I)) by forming a more stable complex, and this decreases the
464
Nernst potential of the concerned redox couple in the presence of a chromogenic ligand (e.g.,
465
tripyridyltriazine or neocuproine), thereby weakening the oxidizing power of the reagent
466
against certain plasma antioxidants and causing negative errors in TAC measurement. Bartosz
467
recommended the use of human plasma in TAC studies rather than serum, which should be
468
analyzed immediately after blood collection; he also recommended the use of citrate or
469
heparin rather than EDTA as anticoagulant for serum samples (EDTA may only be
470
permissible at low concentrations), and pointed out to cases where serum yielded higher TAC
471
values than plasma due to the possible release of antioxidants from blood platelets during the
472
clotting process.33
473
There is lack of correlation between activities determined by the same antioxidant by
474
different assays and between activities determined by the same assay in different
20 ACS Paragon Plus Environment
Page 21 of 105
Journal of Agricultural and Food Chemistry
475
laboratories.58 Especially assays based on the inhibition of lipid peroxidation are known to
476
correlate poorly with either HAT- or ET-based assays, meaning that an antioxidant with a
477
high TEAC value in routine TAC assays may not perform well in preventing/retarding lipid
478
peroxidation. When the TAC results found by different methods for complex samples are
479
compared, there may only be a good linear correlation (i.e. r2 ≈ 1) but not a one-to-one
480
identity between them. The reason is that every assay has its own unique thermodynamic and
481
kinetic characteristics, and the oxidizing power of each TAC reagent against a given
482
antioxidant within a fixed time is naturally different from one another. Even the same assay
483
with slight differences in reagent preparation may give rise to serious differences in results:
484
for example, ORAC assay using two different probes, β-phycoerythrin and fluorescein, may
485
give entirely different TEAC values for quercetin as 2.07±0.05 and 7.28±0.22, respectively;
486
ABTS radical yields 0.55 and 0.94 TEAC values for glutathione using the ABTS/H2O2/HRP
487
and ABTS•+ decolorization methods, respectively; the TEAC value for ascorbic acid may
488
change three-fold depending on the type of oxidant used for generating ABTS•+ radical, i.e.
489
MnO2 versus persulfate.33 Generally ET-based assays correlate well among themselves due to
490
the similarity in mechanism of action, and therefore some antioxidant researchers consider the
491
application of a series of ET-based assays redundant. On the other hand, it is well known that
492
HAT- and ET-based assays like ORAC, ABTS/TEAC and FRAP give none or weak
493
correlations for plasma samples.33 In this regard, the antioxidant activities of common
494
vegetables (total sample size: 927) collected from the U.S. market, analyzed using the ORAC
495
and FRAP procedures, did not correlate well.59 Cao and Prior observed a weak linear
496
correlation between serum ORAC and serum FRAP, but no correlation either between serum
497
ORAC and serum ABTS/TEAC, or between serum FRAP and serum ABTS/TEAC.60 Prior et
498
al. recommended the evaluation of overall antioxidant capacity by using multiple assays to
499
generate an “antioxidant profile” encompassing reactivity toward both aqueous and
21 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 22 of 105
500
lipid/organic radicals directly via radical quenching and radical reducing mechanisms and
501
indirectly via metal complexing.2 It may generally be recommended to use a variety of assays
502
with different mechanisms (such as ET-, HAT- and lipid peroxidation-based assays) for
503
complex samples in order to see the whole picture for antioxidant action.
504 505
2.1.2.1. Spectroscopic Methods
506 507
2.1.2.1.1. Folin-Ciocalteu (FC) Assay
508 509
The Folin−Ciocalteu (FC) method was initially intended for the analysis of proteins, taking
510
advantage of the reagent’s activity toward protein tyrosine (containing a phenol group)
511
residue, where tyrosine reacted with the heteropoly reagent to give a blue color proportionate
512
to the protein content.34 Much later, Singleton et al.35 extended this assay to the analysis of
513
total phenols in wine. Fundamentally, the FC assay is based on the oxidation of phenol
514
compounds in alkaline (carbonate) solution with a molybdotungstophosphate heteropolyanion
515
reagent (3H2O-P2O5-13WO3-5MoO3-10H2O), yielding a colored product with a broad band
516
having an absorbance maximum (λmax) laying between 750 and 765 nm. Although a similar
517
phosphomolybdenum blue formation method, i.e. without tungstate in the reagent, was also
518
reported for antioxidant capacity determination in acidic medium at elevated temperature,61 it
519
was not tested for a wide variety of antioxidants and its molar absorptivity for vitamin E was
520
rather low, which possibly restricted further practice of the method.
521
The possible simultaneous occurrence of several reduced species in the FC heteropoly-
522
chromophore can account for the broad peaks. Although the color may be developed more
523
quickly at warmer temperature, the loss of color with time is greater at higher temperatures.
524
As tannic acid from different preparations of wines and spirits could vary, and other tannins
22 ACS Paragon Plus Environment
Page 23 of 105
Journal of Agricultural and Food Chemistry
525
covered a wide range of color yield per unit weight, Singleton et al.35 replaced tannin (as a
526
reference compound) with GAE in reporting FC results; the minimum detectable amount of
527
phenols was at the order of 3 mg GAE/liter, depending on optical cuvette thickness. The
528
gallic acid added to wine was recovered quantitatively, and the absorbance produced from a
529
mixture of natural phenols of different classes was equivalent to the sum of their individual
530
contributions, meaning that chemical deviations from Beer’s law was essentially absent in the
531
FC system.
532
Neither the exact chemical nature nor the redox potential of the FC phenol reagent is
533
definitely known. It is a strong oxidizing agent, and can non-specifically oxidize many non-
534
phenolic reducing compounds including weak reductants (e.g., aromatic amines, sulphite,
535
ascorbic acid, Cu(I), Fe(II), etc.) along with phenolics.15 Due to this oxidative ability, FC was
536
proposed to be used as a TAC reagent in the reducing capacity assay for antioxidants, where
537
the molybdenum center in the complex reagent is reduced from Mo(VI) to Mo(V) with an
538
electron donated by an antioxidant to produce a blue color.14 Unfortunately, the detailed
539
molecular and electronic structures of the blue reduction products are unclear, but it is known
540
that molybdates are more easily reduced than tungstates in heteropoly salts. FC assay has
541
certain advantages over some other TAC assays in that it is simple, fast, robust, does not
542
require specialized equipment, and the long-wavelength absorption of the chromophere
543
minimizes interference from the sample matrix. However, a drawback of the FC assay is that
544
reducing agents such as ascorbic acid, citric acid, simple sugars or certain amino-acids can
545
interfere with the analysis and thus overestimate the content of phenolic compounds.
546
Tryptophan, indoles, purines, guanine, xanthine, and uric acid were also reported to react with
547
the FC reagent to yield molybdenum blue.35 Another disadvantage is that its commercially
548
available reagent is to be used for reproducible TAC assays, since its preparation by separate
549
laboratories is rather cumbersome. The conventional FC reagent is only applicable to water-
23 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 24 of 105
550
soluble antioxidants, and operates at an unrealistically high pH. To avoid the possible air
551
oxidation of the tested phenols before the color reaction, the FC reagent should be added
552
before alkali. Nevertheless, FC is routinely practiced in antioxidant research laboratories for
553
testing food and plant extracts. Its fully automated-continuous flow 40-sample/hour procedure
554
was also adapted.62
555
Since most phenolic compounds are in dissociated form (as conjugate bases, mainly
556
phenolate anions) at the working pH of the assay (pH ≈10), they can be more easily oxidized
557
with the FC reagent, possibly giving rise to an overestimated TAC value.14,56 The FC
558
chromophore, the molybdotungstophosphate heteropolyanion (PMoW11O404-), does not have
559
an affinity toward organic solvents owing to its quadruple negative charge2 giving rise to
560
strong ion−dipole interactions with solvent water molecules. Thus, the FC method was
561
modified and standardized by Berker et al.63 so as to enable simultaneous measurement of
562
lipophilic and hydrophilic antioxidants in NaOH-added isobutanol−water medium. The
563
modified procedure was successfully applied to the total antioxidant capacity assay of Trolox,
564
quercetin, ascorbic acid, gallic acid, catechin, caffeic acid, ferulic acid, rosmarinic acid,
565
glutathione, and cysteine, as well as of lipophilic antioxidants such as α-tocopherol (vitamin
566
E), butylated hydroxyanisole, butylated hydroxytoluene, tert-butylhydroquinone, lauryl
567
gallate, and β-carotene.63
568 569
2.1.2.1.2. FRAP Assay
570 571
The FRAP assay, first introduced by Benzie and Strain to determine the antioxidant capacity
572
of plasma and later modified for application to other matrices such as tea and wine,42-45 is
573
based on the reduction of Fe(III) to Fe(II) by antioxidants in the presence of tripyridyltriazine
574
tridentate ligand forming a colored complex with Fe(II):
24 ACS Paragon Plus Environment
Page 25 of 105
575 576 577 578
Journal of Agricultural and Food Chemistry
Fe(TPTZ)23+ + ArOH → Fe(TPTZ)22+ + ArO• + H+ … (Eq. 16) where TPTZ denotes 2,4,6-tripyridyl-s-triazine ligand, and the absorption maximum lies at a
579
wavelength of λmax = 593 nm. Although Fe(III)-Fe(II) reduction may equally well produce
580
colored chelates in the presence of ortho- and batho-phenanthrolines yielding reduction
581
potentials exceeding 1.0 V, TPTZ was carefully selected, because EoFe(III)-Fe(II) in the presence
582
of TPTZ is only 0.70 V which may selectively oxidize most antioxidants but not citric acid
583
and simple sugars. The FRAP assay is simple, practical, inexpensive, and may offer a putative
584
index of antioxidant capacity.15 Since the FRAP reaction with antioxidants produces a single
585
colored product, i.e. Fe(II)-TPTZ, the FRAP absorbance versus concentration curves are well
586
linear over a wide range, and TAC additivity is usually observed in mixtures ‒ except for
587
those containing small-molecular and protein thiols.11,64
588
Pulido et al.45 compared the antioxidant efficiency of a number of antioxidant
589
compounds with the use of equivalent concentrations (EC1) defined as the concentration of
590
antioxidant with a reducing effect equivalent to 1 mmol/L Fe(II), and found that polyphenols
591
had lower EC1 values, and therefore higher reducing power, than ascorbic acid and Trolox.
592
Tannic acid and quercetin had the highest antioxidant capacity followed by gallic and caffeic
593
acids, while resveratrol showed the lowest reducing effect and carotenoids had no ferric
594
reducing ability. This inability of the FRAP method to assay thiols64 and carotenoids,45
595
possibly due to the kinetic inertness of high-spin Fe(III) and the problems associated with
596
mutual solubility of reagent and analyte in the same solvent medium, respectively, has also
597
been criticized in various research articles by other users of the method. Especially the
598
inability of the FRAP reagent to effectively oxidize biothiols make the method rather
599
ineffective in evaluting the TAC of intracellular fluids and human plasma/serum.11,64,65
600
Magalhaes et al.15 attribute this specific inadequacy of FRAP (i.e. an ET-reagent) to the mode
601
of action of thiols and carotenoids (i.e., antioxidants mainly acting by H-atom transfer). 25 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 26 of 105
602
However, if the problem had merely arisen from H-atom transfer, CUPRAC as another well-
603
known ET-based method would not have responded to carotenoids in aqueous acetone
604
solution.66 High-spin iron(III) in the reagent has half-filled d-orbitals responsible for kinetic
605
inertness, and thiol (RSH) oxidations usually proceed through thiyl (RS•) radical
606
intermediates which do not form appreciably at the acidic pH of the FRAP protocol.67
607
Another fact worthy of comment is that polyphenols with slow kinetic behaviors such as
608
caffeic acid, tannic acid, ferulic acid, p-coumaric acid, and quercetin cannot be fully oxidized
609
within the protocol time (i.e. typically 4 min) of the FRAP assay.14 FRAP reactions are
610
carried out in acidic medium in order to suppress Fe(III) hydrolysis (i.e. at pH 3.6) where
611
phenolic antioxidants are not dissociated, and therefore lower results than actual TAC are to
612
be expected because phenolates are oxidized much faster than the corresponding phenols.
613
FRAP and similar Fe(III)-reduction based ET assays were also criticized for producing Fe(II)
614
as the reduction product, which could give rise to the generation of reactive species (such as
615
hydroxyl radicals) upon Fenton-like reactions with H2O2, thereby causing ‘redox cycling’ of
616
phenolics and yielding erroneous TAC results.30 Antolovich et al.68 are of the opinion that the
617
measured ferric reducing capacity does not necessarily reflect antioxidant activity, but instead
618
provides a very useful ‘total’ antioxidant concentration, without measurement and summation
619
of the concentration of all antioxidants involved. Pulido et al.45 concluded that the antioxidant
620
efficiency of polyphenols depended on the extent of hydroxylation and conjugation.
621
Specifically for flavonoids, scientists investigating structure-activity relationships of
622
antioxidants suggested that the free radical scavenging ability increases when the following
623
criteria are met: (i) the presence of a 3’,4’-dihydroxy structure in the B ring; (ii) the presence
624
of a 2,3-double bond in conjunction with the 4-oxo group in the heterocycle, allowing for
625
conjugation between the A and B rings; (iii) the presence of 3- and 5-hydroxyl groups in the
626
A ring together with a 4-oxo function in the A and C rings (Figure 3).69,70 The results of
26 ACS Paragon Plus Environment
Page 27 of 105
Journal of Agricultural and Food Chemistry
627
Pulido et al.45 with flavonoids agreed with these criteria, e.g., quercetin, meeting all the listed
628
three conditions, was more potent than rutin (a flavonoid glycoside) and catechin (lacking
629
coplanarity due to the absence of 2,3-double bond, and therefore having hindered
630
conjugation/resonance stabilization over the whole molecule).
631 632 633 634 635
Figure 3
2.1.2.1.3. CUPRAC Assay
636 637
The CUPRAC assay of TAC determination is only 11-years old,40 but has branched into
638
various modified methods of antioxidant capacity/activity measurement associated with
639
Cu(II)-Cu(I) reduction in the presence of selective Cu(I)-stabilizing ligand, neocuproine (2,9-
640
dimethyl-1,10-phenanthroline). It has also been applied to various matrices containing both
641
hydrophilic and lipophilic antioxidants.
642
The main method is based on the absorbance measurement of the CUPRAC
643
chromophore, Cu(I)-neocuproine (Nc) chelate, formed as a result of the redox reaction of
644
antioxidants with the CUPRAC reagent, bis(neocuproine)copper(II) cation [Cu(II)-Nc], where
645
absorbance is recorded at the maximal light absorption wavelength of 450 nm (Figure 4).
646 647 648 649 650
Figure 4
651
with n-electron reductant antioxidants (AOX), according to Eq. 17:
The chromogenic oxidizing reagent of the developed CUPRAC method, Cu(II)-Nc, reacts
652 653 654 655 656
nCu(Nc)2 2+ + n-electron reductant (AOX) ↔ nCu(Nc)2+ + n-electron oxidized product + nH+ … (Eq. 17)
657
to the corresponding quinones (Ar=O), of ascorbic acid to dehdroascorbic acid, and of thiols
In this reaction, the oxidation of reactive Ar–OH groups of polyphenolic antioxidants
27 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 28 of 105
658
to the corresponding disulfides occur, while Cu(II)-Nc is reduced to the yellow-orange
659
colored Cu(Nc)2+ chelate. Although the concentration of Cu2+ ions is in stoichiometric excess
660
of that of Nc in the CUPRAC reagent for driving the redox equilibrium reaction to the right,
661
the actual oxidant is the Cu(Nc)22+ species and not the sole Cu2+, because the standard redox
662
potential of the Cu(II/I)-Nc is 0.6 V, much higher that of the Cu2+/Cu+ couple (0.17 V).71 The
663
reason is that Cu(I)-Nc is perfectly tetrahedral owing to the d10-electronic configuration of
664
Cu(I) having sp3 hybridization, while two molecules of neocuproine give a distorted
665
tetrahedral structure to Cu(II) having d9-configuration (i.e. known as Jahn-Teller effect in
666
coordination chemistry, which is enhanced when solvent molecules are also attached to
667
CuII(Nc)2 in octahedral coordination), thereby selectively stabilizing Cu(I) over Cu(II) (i.e. the
668
logarithmic stability constants of CuII(Nc)2 and CuI(Nc)2 are 12 and 19, respectively).72 As a
669
result, polyphenols are oxidized much more rapidly and efficiently with Cu(II)-Nc than with
670
Cu2+, and the amount of colored product [i.e., Cu(I)-Nc chelate] emerging at the end of the
671
redox reaction is equivalent to that of reacted Cu(II)-Nc.71 The liberated protons are buffered
672
in ammonium acetate medium, which provides a pH of 7.0 basically conforming to
673
physiological conditions. The highest antioxidant capacities in the CUPRAC method were
674
observed for epicatechin gallate, rosmarinic acid, epigallocatechin gallate, quercetin, fisetin,
675
epigallocatechin, catechin, caffeic acid, epicatechin, gallic acid, rutin, and chlorogenic acid in
676
this order,73,74 in accordance with theoretical expectations, because the number and position of
677
the hydroxyl groups as well as the degree of conjugation of the whole molecule are important
678
for easy electron transfer.70
679
Among the three substituted-phenanthrolines ‒used before in antioxidant or protein
680
assays‒ capable of selective stabilization of Cu(I) thereby increasing the Cu(II,I) reduction
681
potential, namely, Nc: neocuproine, BCS: 2,9-dimethyl-4,7-diphenyl-1,10-phenantroline
682
disulfonic acid,75 and bicinchoninic acid (BCA: 2-(4-carboxyquinolin-2-yl)quinoline-4-
28 ACS Paragon Plus Environment
Page 29 of 105
Journal of Agricultural and Food Chemistry
683
carboxylic acid),76 only neocuproine with the CUPRAC method has found wide use as an
684
effective tool in antioxidant research.40,41 Certainly, there are a number of reasons for this
685
choice due to the more diverse areas of usage of the cupric-neocuproine reagent. For example,
686
it is known that Cu(I)-BCS has a higher overall charge than Cu(I)-Nc due to the presence of
687
negatively-charged sulfonate groups on the phenanthroline ring, giving rise to stronger ion-
688
dipole interaction of the former with water molecules and subsequent cell membrane
689
impermeability. As a result, the Cu(I)-BCS method is expected to be less useful than
690
CUPRAC for the TAC assay of tissue homogenates. ET-based antioxidant assays may show
691
significant solvent dependencies and differences in proton-coupled electron transfer rate,77
692
and it has been shown by Çelik et al.78 that the cupric-BCS assay is not competent with
693
conventional CUPRAC using cupric-neocuproine reagent in regard to reaction kinetics and
694
response to lipophilic plasma antioxidants (e.g., β-carotene, α-tocopherol). The standard
695
reduction potential of the Cu(II,I)-BCS couple was reported to be Eo=0.844 V,79 a little higher
696
than those of most widely used ET-reagents, possibly affecting selectivity towards antioxidant
697
compounds. On the other hand, although BCA provides a longer wavelength (558 nm) than
698
Nc for cuprous chelate absorption which may seem advantageous at first glance for
699
preventing the possible interferences of plant pigments during measurement, Marques et al.80
700
recently discovered that in the BCA assay, the concentration of free Cu2+ ions cannot be
701
maintained in excess (i.e. required for the completion of certain oxidation reactions) because
702
of the precipitation of the complex.81 To guarantee a complete complexation without
703
precipitation, the BCA ligand must be added in stoichiometric ratio or in excess. Therefore,
704
the potential complexation of other metal ions that might be present in the sample (for
705
example Fe) cannot be ignored, and this situation limits the applicability of BCA.80
706
Owing to the favorable redox potential in neutral medium, the CUPRAC method has a
707
better chance of simulating physiologically important redox reactions of antioxidant
29 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 30 of 105
708
compounds, including serum antioxidants. In the normal CUPRAC method (CUPRACN), the
709
oxidation reactions of most food/biological antioxidants are essentially complete within 30
710
min. Flavonoid glycosides require acid hydrolysis to their corresponding aglycons for fully
711
exhibiting their antioxidant potency. Slow reacting antioxidants may need elevated
712
temperature incubation so as to complete their oxidation with the CUPRAC reagent.40,41
713
Although the protocol time period of the CUPRAC assay is set at 30 min, few antioxidants
714
with high redox potentials like naringin, naringenin and bilirubin may not reach absorbance
715
saturation so easily. On the other hand, since 80-90 % of the peak absorbances for a great
716
majority of antioxidants is reached within the first few minutes, online HPLC-post column
717
detection and voltammetric modifications of CUPRAC may allow a reaction time of 1 min for
718
antioxidants with the Cu(II)-neocuproine reagent.
719
The CUPRAC method of TAC assay has been successfully applied to antioxidants in
720
food plants, human serum, and to hydroxyl and superoxide radical scavengers. In the assay of
721
human serum antioxidants, hydrophilic antioxidants were measured after precipitation of
722
proteins with perchloric acid (trichloroacetic acid, ammonium sulfate and organic solvents are
723
other known protein-precipitation reagents), while lipophilic ones like α-tocopherol and β-
724
carotene were determined by n-hexane extraction, evaporation, followed by color
725
development in dichloromethane (DCM) of the Cu(I)-Nc charge-transfer complex formed
726
from their CUPRAC reaction.41 Since the CUPRAC chromophore, i.e. Cu(I)-neocuproine
727
cation, has a large molecular size, it has a very weak hydration sphere and therefore is easily
728
extracted into an organic solvent like DCM due to its low hydration energy (for divalent
729
chromophores
730
chromophore: phospho-tungsto-molybdate anion, this extraction is not so easy because of
731
enhanced ion-dipole interactions with solvent water molecules). This feature of the Cu(I)-Nc
732
chelate provides a great advantage for the CUPRAC method to be applicable to lipophilic
like
Fe(II)-tripyridyltriazine
cation
or
30 ACS Paragon Plus Environment
tetravalent
Folin-Ciocalteu
Page 31 of 105
Journal of Agricultural and Food Chemistry
733
antioxidants along with hydrophilic ones. In a miniaturized CUPRAC method without
734
preliminary separation of lipophilic and hydrophilic serum antioxidants, serum samples were
735
centrifuged after 10 % TCA precipitation, and CUPRAC was directly applied to the
736
supernate.82
737
Since essentially flavones and flavonols (and other flavonoids to a lesser extent) could
738
be chelated with lanthanum(III) in the form of basically nonpolar complexes, and ascorbic
739
acid (AA) either did not complex or formed very weak hydrophilic complexes under the same
740
conditions, AA assay with a high redox equilibrium constant of the CUPRAC reaction with
741
preliminary extraction of flavonoids as their La(III) complexes was possible.83 Although in
742
principle, ascorbate could be assayed by measuring the CUPRAC absorbance of a real
743
mixture before and after treating with the substrate-specific ascorbate oxidase enzyme, many
744
constituents in plant extracts may inhibit this enzyme causing erroneous results. Lipophilic
745
and hydrophilic antioxidants (e.g., β-carotene, α-tocopherol, AA, quercetin, etc.) could be
746
simultaneously assayed with a modified CUPRAC method in the same solvent medium of
747
acetone- water (9:1, v/v) with the aid of their inclusion complexes formed with 2 % methyl-β-
748
cyclodextrin (M-β-CD), because this oligosaccharide can form inclusion complexes (Figure
749
5) with lipophilic antioxidants in the interior part while the outer part binds hydrophilic
750
antioxidants.84
751 752 753 754
Figure 5 The CUPRAC assay could be further modified to fit the needs of scavenging activity
755
determinations of reactive oxygen and nitrogen species (ROS/RNS) such as hydrogen
756
peroxide, superoxide anion and hydroxyl radicals, where either the original probe or the
757
converted product had CUPRAC reactivity. In measuring the hydroxyl radical scavenging
758
activity of certain water-soluble compounds (metabisulfite, thiourea, glucose, lysine, etc.), the
759
probes of p-aminobenzoate, 2,4-dimethoxy-benzoate, and 3,5-dimethoxybenzoate were used 31 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 32 of 105
760
to detect hydroxyl radicals, and the •OH scavenging rate constants of these compounds were
761
determined by competition kinetics.85 In the measurement of hydroxyl radical scavenging
762
activities of polyphenolics, special measures were taken so as to prevent the redox cycling of
763
phenolic compounds that could otherwise produce unrealistic results. For this purpose, the
764
Fenton reaction was stopped at the end of the 10th minute with the addition of catalase to
765
annihilate hydrogen peroxide and cease •OH production, and the dihydroxybenzoates formed
766
from the salicylate probe under hydroxyl radical attack (Figures 6a and 6b) were measured
767
with the CUPRAC method, rate constants being calculated with competition kinetics.12
768 769 770 771 772
Figure 6
773
with xanthine–xanthine oxidase (X–XO), and the inhibition of the enzyme was measured
774
upon addition of polyphenolics to the system (Figure 7).86
775 776 777 778 779 780 781 782
In another modified CUPRAC method, the superoxide anion radical was generated
Figure 7. Xanthine oxidase (XO) inhibitors obstruct the formation of CUPRAC-reactive uric acid from xanthine, thereby enabling a modified CUPRAC assay for measuring XO-inhibiting antioxidant activity.
The hydrogen peroxide scavenging (HPS) activity of the polyphenolics was measured
783
in the presence of Cu(II) (as catalyst) with the HPS-CUPRAC method.87 A low-cost optical
784
antioxidant sensor (CUPRAC sensor) was developed by immobilizing the Cu(II)-Nc reagent
785
onto a perfluorosulfonate cation-exchange polymer membrane matrix (Nafion®),73 and the
786
colored Cu(I)-Nc cation was produced on the membrane without diffusing into solution. This
787
membrane sensor provided great ease and convenience to TAC determinations, like a
788
sensitive pH paper immersed in solution for hydrogen ion activity determination (Figure 8).
789 790 791
Figure 8. The cationic chelate of the CUPRAC reagent, Cu(Nc)22+, is electrostatically held on the sulfonate groups of a cation-exchange membrane (Nafion), producing a CUPRAC 32 ACS Paragon Plus Environment
Page 33 of 105
Journal of Agricultural and Food Chemistry
792 793 794 795
antioxidant sensor; antioxidants reduce this membrane-held chelate to the cuprous chromophore, Cu(Nc)2+, showing maximal absorption at 450 nm.
796
determination of polyphenols in complex plant matrices. This method combines
797
chromatographic separation, constituent analysis, and post-column identification of
798
antioxidants (Figure 9) in plant extracts. Antioxidant polyphenols in complex samples can be
799
separated on a C18 HPLC column (diode-array detected at 280 nm) and further react with the
800
Cu(II)-Nc reagent in a post-column reactor to yield the Cu(I)-Nc chromophore detected at 450
801
nm. Thus, twice as much information can be extracted from the same sample using these two
802
chromatograms, i.e. their separability through a C18 column and their CUPRAC reactivity in
803
the post-column. This robust online chromatographic method enables individual
804
detection/quantitation of antioxidant constituents as well as total measurement of TAC;
805
moreover, non-antioxidants are not detected in the post-column chromatogram, increasing
806
selectivity of the method.88
807 808 809 810 811 812 813 814
Figure 9. The chromatogram of green tea extract showing conventional HPLC (with 280-nm detection: positive trace chromatogram) and on-line post-column HPLC-CUPRAC assay (with 450-nm detection: negative-trace chromatogram); notice that gallic acid (GA) and tea catechins (CT, EC, EGC, ECG, EGCG) are identified in both chromatograms, while the CUPRAC non-reactive caffeine, symbolized as (C), gave a peak only in the upper chromatogram but not in the lower one, because it is not an antioxidant compound.
815
ammonium acetate, should be replaced with urea at the same pH in order to prevent re-
816
precipitation of dissolved proteins. CUPRAC in urea buffer also responded to thiol-containing
817
proteins in food and serum.11,64 Another modified CUPRAC method is comprised of a tert-
818
butylhydroquinone (TBHQ) probe with the phenazine methosulphate/β-nicotinamide adenine
819
dinucleotide (PMS/NADH) non-enzymic O2•– generating system for superoxide radical
820
scavenging activity (SRSA) assay of thiol-type antioxidants (e.g., GSH, cysteine), amino
821
acids (e.g., serine, threonine), plasma antioxidants (e.g., bilirubin, albumin), and other
A novel online HPLC-CUPRAC method was developed for the selective
In protein TAC determination, the classical buffer of the CUPRAC reaction, i.e.
33 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 34 of 105
822
antioxidants (e.g., methionine); the SRSA method is based on the measurement of the
823
CUPRAC absorbance of the remaining TBHQ in the reaction medium (TBHQ is CUPRAC-
824
reactive while its oxidation product is not (Figure 7), and this probe is isolated by ethyl
825
acetate extraction from other CUPRAC-reactive interferents remaining in the aqueous
826
phase).89
827 828
Figure 7
829 830
The CUPRAC assay together with its modifications for ROS scavenging
831
measurements have been summarized, and the methodology has been demonstrated to have
832
certain advantages over other similar ET-based TAC methods, mentioned in a comprehensive
833
review by Özyürek et al.:8
834
(i) The CUPRAC reagent, being an outer-sphere electron-transfer agent, is capable of rapidly
835
oxidizing thiol-type antioxidants, whereas iron(III)-based ET-methods like FRAP may only
836
measure limited thiols like GSH with serious negative error, possibly due to the kinetic
837
inertness of high-spin Fe(III) and to the inadequate formation of thiyl radicals (i.e.
838
intermediary species of thiol oxidation) in acidic medium. Since the redox potential of
839
oxidized and reduced forms of glutathione (EoGSSG/2GSH) is the basic indicator of the biological
840
conditions of a cell, and GSH acts as reconstituent of intercellular ascorbic acid from the
841
dehydroascorbic acid, an ET assay should be capable of accurately measuring glutathione
842
among plasma antioxidants.
843
(ii) The redox potential of Cu(II,I) chelated to necuproine is only 0.6 V, close to that of
844
ABTS•+ /ABTS (Eo=0.68 V) and FRAP (Eo=0.70 V), all versus NHE. Simple sugars and citric
845
acid –which are not classified as antioxidants– are not oxidized with the CUPRAC reagent.
34 ACS Paragon Plus Environment
Page 35 of 105
Journal of Agricultural and Food Chemistry
846
On the other hand, the ferric/ferrous potential in the presence of ortho-phenanthroline or
847
batho-phenanthroline-type ligands is much higher, adversely affecting selectivity.
848
(iii) The reagent is much more stable and readily accessible than the chromogenic radical
849
reagents
850
dihydrochloride (DMPD)).
851
(iv) The CUPRAC method shows versatility to the determination of both hydrophilic and
852
lipophilic antioxidants, because the CUPRAC chromophore, bis(neocuproine)copper(I)
853
chelate, has unipositive charge having less ion-dipole interaction with water, and the chelate
854
rings are essentially hydrophobic. Thus it is compatible with aqueous and organic solvents
855
(alcohols, acetone, dichloromethane, etc.) and alcohol-water mixtures. The robustness of the
856
CUPRAC assay to different solvents has been recently confirmed by Christodouleas et al.90 in
857
the TAC assay development study for edible oils.
858
(v) The redox reaction giving rise to the Cu(I)-Nc chromophore is relatively insensitive to a
859
number of parameters (e.g., air, sunlight, humidity, and to a certain extent pH) adversely
860
affecting radical reagents such as DPPH.
861
(vi) The CUPRAC reagent can be adsorbed on a perfluorosulfonate cation-exchanger
862
membrane enabling the manufacture of a low-cost, linear-response antioxidant sensor.
863
CUPRAC is also adaptable to online-HPLC applications with the use of a post-column
864
reactor, enabling sensitive determination of antioxidants individually.
865
(vii) CUPRAC usually gives perfectly linear absorbance/concentration curves (r ≈ 0.999) over
866
a wide concentration range, as opposed to certain other methods yielding polynomial curves.
867
The molar absorptivity (extinction coefficient) for n-electron reductants, i.e. (7.5–9.5x103 n)
868
M-1cm-1, is sufficiently high to enable sensitive determination of most phenolic antioxidants.
869
(viii) The CUPRAC spectrophotometric method obeys Beer’s law in regard to the additivity
870
of absorbances due to individual antioxidant constituents, because a single chromophore,
(e.g.,
ABTS,
DPPH,
galvinoxyl
and
N,N-Dimethyl-p-phenylenediamine
35 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 36 of 105
871
cuprous-neocuproine, is formed upon reduction of the CUPRAC reagent with antioxidants.
872
Consequently, the CUPRAC-TAC values of antioxidants in complex mixtures are perfectly
873
additive (e.g., the TAC of a phenolic mixture is equal to the sum of individual antioxidant
874
capacities of its constituent polyphenols).
875
(ix) CUPRAC operates at nearly physiological pH (pH 7 of ammonium-acetate buffer) as
876
opposed to the unrealistic acidic conditions (pH 3.6) of FRAP or alkaline conditions (pH 10)
877
necessary for phenols to dissociate protons in the Folin-Ciocalteu assay. At more acidic
878
conditions than the physiological pH, the reducing capacity may be suppressed due to
879
protonation on antioxidant compounds, whereas in more basic conditions, deprotonation of
880
phenolics enhances a sample’s reducing capacity, thereby causing unrealistic TAC
881
measurements.
882
(x) Since the Cu(I) ion emerging as a product of the CUPRAC redox reaction is in a
883
coordinatively saturated state (i.e. two molecules of neocuproine tetrahedrally coordinate the
884
cuprous ion), it cannot act as a prooxidant that may cause oxidative damage to biological
885
macromolecules in body fluids. Fe(III)-based assays were criticized for producing Fe2+, which
886
may act as a prooxidant to produce •OH radicals as a result of its reaction with H2O2, and
887
subsequently cause a ‘redox cycling’ of antioxidants during the assay, yielding unreliable
888
results. Ferric or ferrous iron, even in full octahedrally-coordinated (e.g., EDTA-chelated)
889
state, can catalyze the decomposition of hydrogen peroxide to reactive species.91 On the other
890
hand, it was experimentally shown that the stable Cu(I)-Nc chelate did not react with H2O2,
891
but the reverse reaction (i.e. oxidation of H2O2 with Cu(II)-Nc) is possible, excluding the
892
possibility of redox cycling of antioxidants with the Cu(I)-Nc product.
893
(xi) The CUPRAC method has recently been implemented in microplate and flow modes,
894
allowing the in vitro assessment of antioxidant capacity of endogenous and dietary molecules
895
as well as the TAC determination of human biological samples.92
36 ACS Paragon Plus Environment
Page 37 of 105
Journal of Agricultural and Food Chemistry
896
(xii) As noteworthy experiences of CUPRAC users, Gorinstein research groups stated that, as
897
an advantage over other electron-transfer-based assays, the CUPRAC test gave reproducible
898
values that were acceptable in regard to its realistic pH close to physiological pH in various
899
food extracts (garlic,93 onion, kiwi, etc.). Bean et al.65 made a comparative evaluation of
900
antioxidant reactivity within obstructed (i.e. resulting from the attack of ROS/RNS in cyclic
901
ischemia and reperfusion) and control rabbit urinary bladder tissue using FRAP and CUPRAC
902
assays, and found that that CUPRAC, but not FRAP, could detect a significant decrease in the
903
reactivity of antioxidants found within the obstructed bladder tissue as compared to the
904
control bladder tissue in both the muscle and mucosa. Bean et al.65 concluded that, as the
905
CUPRAC assay was responsive to hydrophilic, lipophilic, and thiol-containing antioxidants at
906
physiological pH, it was a much better tool to analyze the reactivity found within tissues.
907 908 909
2.1.2.1.4. Ferricyanide (Hexacyanoferrate(III))-Prussian Blue Assay
910 911
The ferricyanide-Prussian blue assay is based on the following chemical reactions:
912 913
Fe(CN)63− + ArOH → Fe(CN)64− + ArO• + H+ … (Eq. 18)
914
Fe(CN)64− + Fe3+ + K+ → KFe[Fe(CN)6] … (Eq. 19) with λmax=700 nm.
915 916
In the conventional method,46 the hexacyanoferrate(III) (common name: ferricyanide) reagent
917
is first incubated in (H2PO4-/HPO42-) buffer at pH 6.6 with antioxidants (at 50°C for 20 min),
918
and the reduction product, hexacyanoferrate(II) (common name: ferrocyanide), combines with
919
the later added Fe(III) to produce Prussian blue, suspended in the medium. The method is also
920
referred to as the ‘reducing power’ assay.94
37 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 38 of 105
921
Although the Fe(III,II) standard reduction potential is 0.77 V causing nonspecific
922
oxidation of any species having a redox potential smaller than this value,59 suitable selection
923
of ligands may bring this value close to the range of common food and biological antioxidants
924
(i.e., having standard potentials in the range of 0.2–0.6 V). In the case of cyanide complexes
925
of iron in hexacyanoferrate complex ions, the logarithmic stability constant of Fe(III)
926
complex is greater than that for Fe(II) (i.e. Log β6 for Fe(CN)63− and Fe(CN)64− are 31 and 24,
927
respectively). Therefore, according to the Nernst equation, the reduction potential for
928
hexanocyanoferrate (III,II) couple is 0.36 V versus NHE, less than that of Fe(III,II). This
929
potential may not be sufficient for oxidizing certain antioxidants having a reduction potential
930
(of the ArO•/ArOH couple) Eo > 0.4 V, and therefore modification of the conventional
931
ferricyanide assay46 was a requirement for comprising diverse antioxidants. The wavelength
932
of maximum absorption shifts to longer wavelengths in the order of ferric-complexing
933
ligands: ortho-phen < batho-phen < tripyridyltriazine (FRAP) < ferricyanide. The
934
bathochromic shift in λmax and band broadening was strongest for ‘Prussian blue’ in the
935
ferricyanide method, because the bonding and antibonding energy levels were closest in
936
Fe[Fe(CN)6]− due to the interchanging oxidation states of iron centers in this complex salt.
937
Longer wavelengths almost always constitute an important advantage in spectrophotometric
938
method selection, because most plant pigments as well as some antioxidants show significant
939
absorption at shorter wavelengths of the visible region, close to the UV range of the visible
940
spectrum.48
941
The conventional ferricyanide assay was first modified by incorporating Fe(III) to the
942
ferricyanide reagent in acidic medium and incubating at elevated temperature (30 min
943
incubation at 50°C), but gave rise to overoxidation of certain antioxidants (especially of
944
hydroxycinnamic acids like caffeic and ferulic acids) causing distinct deviations from
945
linearity of calibration curves.48 Then the authors introduced a second modification of the
38 ACS Paragon Plus Environment
Page 39 of 105
Journal of Agricultural and Food Chemistry
946
method, by adding Fe(III) at the start and optimizing the acidity of the incubation medium as
947
pH=1.6 (so as to prevent Fe(III) hydrolysis), and by adding the anionic surfactant SDS to
948
stabilize the negatively-charged Prussian blue complex ion: Fe[Fe(CN)6]-.47 Compared to the
949
original method,46 stabilization with SDS enabled a color development time of 30 min at
950
room temperature (instead of the 2-min reaction time of Oyaizu to avoid Prussian blue
951
precipitation) and a wavelength (λmax) shift from 700 nm to 750 nm. The modified procedure
952
was named as the ‘ferric-ferricyanide assay’, because it was not clear whether Fe(III) or
953
ferricyanide was the actual oxidant. Consequently, the oxidation equilibria for antioxidants by
954
the combined (Fe(III) + Fe(CN)63−) reagent significantly shifted to the right, possibly
955
increasing the potential above that of Fe(III,II) couple, owing to the formation of insoluble
956
Prussian blue, i.e. if the actual oxidant was Fe(III), then its reduced form, Fe(II), would again
957
produce the same Prussian blue product with the ferricyanide constituent of the reagent:
958 959 960 961 962 963
Fe3+ + ArOH → Fe2+ + ArO• + H+ … (Eq. 20) Fe2+ + Fe(CN)63− + K+ → KFe[Fe(CN)6] … (Eq. 21) with λmax=750 nm. Finally, with the third modification of the method, Berker et al.95 were able to measure
964
lipophilic (e.g., α-tocopherol, BHT, and β-carotene) and hydrophilic antioxidants in the same
965
solution comprising 1:9 (v/v) H2O–acetone with or without 2% methyl-β-cyclodextrin; this
966
was necessary, as the original ferricyanide method could only assay hydrophilic antioxidants.
967
This modified assay was not adversely affected from citric acid and simple sugars, and proved
968
to be additive for TAC values of complex mixtures.95
969 970
2.1.2.1.5. ET-based Spectrophotometric Assays Involving Strongly Oxidizing Reagents
971
(Ce(IV), Cr(VI), and Mn(VII) Assays)
972
39 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 40 of 105
973
Strong oxidizing agents such as Ce(IV), Cr(VI) and Mn(VII) may be used as chromogenic
974
TAC reagents only if their oxidizing power are decreased to the level of specifically oxidizing
975
common antioxidants but not other organic substances (i.e. their redox potentials should be
976
brought to roughly 0.6−0.7 V of widely used TAC reagents). This can usually be
977
accomplished by increasing the working pH (e.g., the extent of chromate(VI) and
978
permanganate(VII) oxidations is a function of H+-ion concentration) and/or by selectively
979
stabilizing the higher oxidation state of the redox couple (e.g., by preferential complexation of
980
Ce(IV) over Ce(III) with sulfate) so as to decrease the Nernst potential (E = Eo + (RT/nF) Ln
981
[cox/cred]), where E and Eo are the instantaneous and standard values of the reduction potential,
982
respectively, R: universal gas constant, T: absolute temperature, n: number of electrons
983
involved in half-cell reaction, F: Faraday’s constant, and cox and cred are the concentrations of
984
the oxidized and reduced forms, respectively, of the redox-active constituent of the TAC
985
reagent). Another disadvantage of the use of strong oxidizing agents in TAC assays is their
986
hydrophilicity, i.e. they cannot be applied to lipophilic antioxidant testing.
987
Özyurt et al.51 developed a simple, sensitive, and low-cost indirect spectrophotometric
988
method for the determination of Ce(IV) reducing antioxidant capacity (CERAC) of plant
989
extracts, based on the oxidation of antioxidants with cerium(IV) sulfate in dilute sulfuric acid
990
at room temperature. The Ce(IV) reducing antioxidant capacity (CERAC) of the sample was
991
measured under controlled conditions of oxidant concentration and pH such that antioxidants
992
but not other organic compounds would be oxidized. The spectrophotometric determination of
993
the remaining Ce(IV) at 320 nm was performed after all antioxidants in solution were
994
oxidized. Blank correction of significantly absorbing plant extracts at 320 nm could be made
995
with the aid of spectrophotometric titration. The trolox equivalent antioxidant capacities
996
(TEAC coefficients) of the tested antioxidant compounds were correlated to those found by
997
ABTS and CUPRAC methods. Since the TEAC coefficients (found by CERAC) of naringin–
40 ACS Paragon Plus Environment
Page 41 of 105
Journal of Agricultural and Food Chemistry
998
naringenin and rutin–catechin pairs were close to each other, this assay was advantageous to
999
accomplish the simultaneous hydrolysis of flavonoid glycosides to the corresponding
1000
aglycones and their subsequent oxidation such that the hydrolysis products exhibed
1001
antioxidant capacities roughly proportional the number of –OH groups in a phenolic
1002
molecule. Özyurt et al.52 further developed the CERAC method by employing a medium of
1003
(0.3 M H2SO4 + 0.7 M Na2SO4) to finely tune the Nernst potential of the Ce(IV)/Ce(III)
1004
couple by maintaining sufficient acidity while selectively complexing Ce(IV) with sulfate in
1005
preference to Ce(III) so as to oxidize true antioxidants but not citric acid or simple sugars.
1006
The TEAC coefficients of this modified CERAC procedure for antioxidants were in the order
1007
of quercetin > rutin > gallic acid > catechin > caffeic acid ≥ ferulic acid > naringenin ≥
1008
naringin > trolox ≥ ascorbic acid, in accordance with those found by other antioxidant assays.
1009
It was also possible for Özyurt et al.53 to measure the fluorescence of the reduction product,
1010
Ce(III), with excitation at 256 nm and emission at 360 nm, and then correlate this
1011
fluorescence intensity to the TAC value of a sample. By measuring the produced Ce(III)
1012
fluorometrically instead of the remaining Ce(IV) spectrophotometrically, the linear range was
1013
widened (e.g., 5.0 × 10−7–1.0 × 10−5 M for quercetin) and the possible interferences of plant
1014
pigments for absorbance measurement at 320 nm were eliminated (however, the linear
1015
correlation coefficient in fluorescence was lower than in spectrophotometry due to the
1016
fluorescence-quenching effect of Ce(IV)).
1017
Işık et al.50 developed the Cr(VI) reducing antioxidant capacity (CHROMAC) assay,
1018
involving the reduction of chromate(VI) with antioxidants to Cr(III) in acidic solution at pH
1019
2.8, and after 50-min reaction time, the remaining Cr(VI) was spectrophotometrically
1020
measured with 1,5-diphenylcarbazide (DPC) at 540 nm. The in situ-formed Cr(III) complex
1021
with the oxidation product of DPC (i.e. diphenylcarbazone) indicated the remaining chromate,
1022
and Cr(VI) consumption was correlated to antioxidant concentration in the sample. The
41 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 42 of 105
1023
authors found comparable results to ABTS and CUPRAC for a number of plant extracts.
1024
However, the TAC order of common antioxidants in CHROMAC did not well agree with
1025
those of other ET-based assays (e.g., rosmarinic acid and quercetin were found to be less
1026
potent than ascorbic acid, based on the comparison of TEAC values). If no measures were
1027
taken for regulating the Nernst potential of Cr(VI)/Cr(III) redox couple, as in chemical
1028
oxygen demand (COD) tests performed in water treatment, the acidic dichromate reagent
1029
would oxidize all the organic load of solution without discriminating antioxidant compounds.
1030
Potassium permanganate is a strong oxidant widely used in analytical chemistry for
1031
redox titrations without an external indicator because of the intrinsic violet color of the
1032
reagent. A critical and comprehensive review of acidic potassium permanganate
1033
chemiluminescence was presented, together with a discussion on reaction conditions, the
1034
relationship between analyte structure and chemiluminescence intensity, and its application to
1035
determine a variety of compounds including antioxidants.96 Francis et al.97 reported the use of
1036
acidic potassium permanganate as a chemiluminescence reagent to rapidly assess the
1037
antioxidant status of fruit juices, teas and other beverages. In the acidic KMnO4
1038
spectrophotometric assay of reducing capacity for antioxidants,98 the sample was oxidized
1039
with acidic permanganate leading to sample discoloration until no color was observed;
1040
subsequent decrease of potassium permanganate concentration was determined with the use
1041
of a calibration curve of absorbance at 535 nm versus concentration.99 In the original assay,
1042
Cacig et al.98 also observed MnO2 particles in suspension, showing the non-stoichiometric
1043
character of oxidation (instead of a neat Mn(VII)-Mn(II) reduction). Although the results of
1044
this method were claimed to correlate with those of other reducing assays and acidic
1045
permanganate was assumed to oxidize phenol by forming phenoxyl-radical and manganic
1046
acid (H2MnO4) in the slowest step of a series of oxidation reactions, it is obvious that
42 ACS Paragon Plus Environment
Page 43 of 105
Journal of Agricultural and Food Chemistry
1047
permanganate in sulfuric acid medium would non-specifically oxidize any organic substance
1048
and the measured parameter would be ‘total organic status’ of a sample rather than its TAC.
1049 1050
2.1.2.2. Electrochemical Methods
1051 1052
Direct electrochemical sensing methods for in vitro antioxidant capacity assessment have
1053
been reviewed by Blasco et al.100 Cyclic voltammetry (CV) is an electrochemical technique
1054
in which a sample is introduced to a reaction chamber with three electrodes: working
1055
electrode (such as glassy carbon (GC)), reference electrode (e.g., Ag/AgCl), and auxiliary
1056
electrode (e.g., Pt wire), where an increasing potential is applied to the working electrode and
1057
current intensity versus potential is recorded.33 Although not all antioxidants can give their
1058
electrons to the GC electrode at an appreciable rate, most antioxidants are CV-active reducing
1059
agents and can therefore be assayed by CV on the basis of their redox potentials. For example,
1060
in cyclic voltammetry utilizing a GC electrode, chlorogenic and caffeic acids showed well-
1061
defined reversible waves in weakly acidic solution, while their electrooxidation assumed a
1062
quasi-reversible character at higher pH due to possible polymerization reactions at the
1063
electrode surface.101 Chevion et al.102 made some pioneering studies on the electrochemical
1064
determination of antioxidant capacity, and defined cyclic voltammetry as a simple, sensitive
1065
and reliable method for determining the TAC of human plasma originating from low
1066
molecular-weight antioxidants (LMWA). Half-wave potential (E1/2) indicated the specific
1067
constituent of the LMWA and its ability to donate electron(s), whereas current intensity at
1068
half-wave potential (Ia) indicated the concentration of this constituent. The E1/2 and Ia values
1069
reflected the antioxidant capacity of the plasma, while the change of Ia upon exposure to
1070
copper ions, ionizing radiation and peroxyl radicals represented the severity of the oxidative
1071
stress induced. In their later work,103 the authors suggested another parameter better
43 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 44 of 105
1072
correlating with antioxidant concentration, namely the AUC around the anodic wave
1073
potential: E1/2. Using the AUC approach, LMWA constituents of human plasma and animal
1074
tissues were identified and further validated by reconstruction of the CV tracing and by
1075
HPLC-electrochemical detection. As the changes for individual constituents within a single
1076
AC wave could be different, the changes in AUC would better represent the residual
1077
antioxidant capacity and allow better quantitation of the loss in a specific component upon
1078
oxidative stress. Chevion et al.102 identified two critical E1/2 values for human plasma, namely
1079
420 ± 25 mV (mainly derived from ascorbate and urate constituents of LMWA) and 920 ± 25
1080
mV (derived from unidentified constituents, excluding simple and protein thiols). Kohen et al.
1081
later showed that this group of mostly unidentified LMWA giving rise to the high-potential
1082
anodic wave could be related to the oxidized form of lipoic acid.104 Spiking brain tissue
1083
samples with relatively high concentrations (0.5 mM) of other antioxidants (carnosine,
1084
tryptophan, and melatonin) yielded an increase in the amplitude of the second anodic wave.103
1085
The authors also found that when the wave was comprised of several constituents
1086
characterized by close but different E1/2 values, a change in the relative concentrations of the
1087
constituents would cause a small shift in the E1/2 of their combined anodic wave.102 Plasma
1088
samples should be preserved with heparin when necessary, but not with EDTA which gave a
1089
specific anodic wave at E1/2 > 900 mV. The major disadvantages of CV are low sensitivity for
1090
antioxidants (usually in several micromolar range) and the necessity for frequent electrode
1091
cleansing before each measurement due to adsorption of proteins and other macromolecules
1092
from biological fluids onto the glassy carbon working electrode.33
1093
Piljac-Žegarac et al.105 used CV to study the electrochemical properties of antioxidants
1094
in fruit tea infusions as well as to estimate the antioxidant capacity. The easiest oxidized
1095
compound (at approximately 130 mV) was ascribed to the oxidation of the ene-diol of
1096
ascorbic acid. A pronounced anodic current peak (corresponding to a quasi-reversible redox
44 ACS Paragon Plus Environment
Page 45 of 105
Journal of Agricultural and Food Chemistry
1097
process, reflecting oxidation to a stable quinone) observed at 440 mV in all analyzed fruit teas
1098
indicated that ortho-dihydroxy-phenol and gallate groups were the major contributors to the
1099
antioxidant capacity of investigated teas. The less oxidizable compounds presented an
1100
irreversible oxidation process between 670 and 700 mV, which was ascribed to the oxidation
1101
of the monophenol group or the meta-diphenols on the A-ring of flavonoids that led to a
1102
phenoxy radical or a phenoxonium ion undergoing successive secondary reactions. The
1103
antioxidant capacity of these fruit teas was determined by estimating the integrated area
1104
(AUC) under the peak up to 600 mV.
1105
Electrochemical techniques of antioxidant characterization consist of CV, differential
1106
pulse (DPV) and square-wave voltammetry (SWV), proposed as useful tools to investigate the
1107
electrochemical behavior of phenolic compounds in different food samples in conjunction
1108
with carbon, diamond and graphite electrodes.106,107 DPV is one of the most sensitive
1109
techniques, and has a received a great deal of attention in recent years.
1110
CV methods have also been described to detect ascorbic acid, citric acid, and sugars in
1111
both food products and pharmaceuticals and to consider their influence on the oxygen
1112
reduction process, but this approach is restricted due to the electrochemically inactive nature
1113
of these compounds in the potential range of O2 reduction.108 Qualitative assessment of wine
1114
phenolics based on reducing strength was also realized by using CV at a GC electrode;
1115
although in this approach, the AUC, namely the charge passed to 0.5 V potential during CV
1116
recording, is understood as a better estimate of the concentration of polyphenols with low
1117
oxidation potentials, quantitation efforts of all reducing compounds may only provide a
1118
qualitative picture in a complex mixture such as wine mainly because the magnitude of the
1119
response is not identical on a molar basis for dissimilar compounds.109 The antioxidant
1120
capacities of several drugs containing acetylsalicylic acid were evaluated using a SOD-based
1121
biosensor method in comparison to ORAC-fluorometric and DPV methods, but although the
45 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 46 of 105
1122
DPV-based method generally seemed to be sufficiently in line with the results of the other
1123
two methods, its precision was rather poor (RSD > 10%).110 Novak et al. used square-wave
1124
voltammetry for investigating the electrochemical behavior of major green tea compounds
1125
such as epigallocatechin gallate, epigallocatechin and gallic acid.106 Magarelli et al. developed
1126
and validated a sensitive DPV method using a GC electrode for the determination of total
1127
phenolic acids in cotton cultivars;107 however, the authors studied only four polyphenolics
1128
(i.e. caffeic, chlorogenic, gallic and gentisic acids) and presented one anodic peak at
1129
approximately 0.4 V that was assigned to the oxidation of phenolic hydroxyls leading to the
1130
formation of o-quinone via semiquinone form, and therefore, it may be argued whether these
1131
four easily-oxidized phenolic acids are a true representative of ‘total phenolic acids’ having
1132
diverse oxidation potentials.
1133
There is very limited study about the usage of GC electrodes as electrochemical TAC
1134
sensors. Milardović et al.111 developed an amperometric method, based on the reduction of
1135
DPPH• at the GC electrode, for measuring the antioxidant activity of pure compounds and
1136
their real mixtures such as tea, wine and other beverages; since the cyclic voltammograms of
1137
a number of water- and ethanol-soluble common antioxidants gave either irreversible or
1138
undefined oxidation peaks, electrochemical reduction of DPPH• was conducted in the
1139
presence and absence of antioxidants, but potential selection was critical due to the
1140
electrochemical interferences of caffeic acid and trolox. Milardović et al.111 further concluded
1141
that amperometric and spectrophotometric DPPH•-based TAC methods had similarities, since
1142
both methods are based on the same reaction thermodynamics and kinetics. The
1143
electrochemical ABTS•+ assay is based on measuring catalytic voltammetric currents caused
1144
by antioxidants acting as reductant toward the electrochemically generated ABTS 2+ (thereby
1145
enabling reoxidation of ABTS•+ on the electrode) in edible oils; this oxidative voltammetric
1146
current intensity increased with an increase in antioxidant concentration, enabling TAC
46 ACS Paragon Plus Environment
Page 47 of 105
Journal of Agricultural and Food Chemistry
1147
determination of edible oils. This method produced rather high blank values in the Trolox
1148
calibration curve, and Trolox addition to sunflower oil matrices did not yield perfectly linear
1149
responses.112 In an attempt to adapt the Ce(IV)-reducing TAC assay51 to electroanalytical
1150
chemistry, Ferreira and Avaca113 measured the ability of eight different compounds in
1151
reducing Ce4+ by chronoamperometric measurement of the remaining Ce3+ species, and found
1152
the TAC order as: tannic acid >> quercetin > rutin > gallic acid ≈ catechin > ascorbic acid >
1153
BHA > Trolox, agreeing well with FRAP. The electrochemical behavior of the Cu(Nc)22+
1154
complex was recently studied by cyclic voltammetry at a GC electrode.114 The
1155
electroanalytical method was based on the reduction of Cu(Nc)22+ to Cu(Nc)2+ by antioxidants
1156
and electrochemical detection of the remaining Cu(II)–Nc (unreacted complex), the difference
1157
being correlated to antioxidant capacity of the analytes. The calibration curves of individual
1158
compounds comprising polyphenolics and vitamin C were constructed, and their response
1159
sensitivities and linear concentration ranges were determined. The reagent on the GC
1160
electrode retained its reactivity toward antioxidants, and the measured TEAC values of
1161
various antioxidants suggested that the reactivity of the Cu(II)–Nc reagent is comparable to
1162
that of the solution-based spectrophotometric CUPRAC assay. This electroanalytical method
1163
better tolerated sample turbidity and provided higher sensitivity (i.e. lower detection limits) in
1164
antioxidant determination than the spectrophotometric assay.114
1165
In a review work, Prieto-Simon et al.115 made a generalization that electroanalytical
1166
biosensor-originated antioxidant activity/capacity measurements based on the reduction of
1167
hazard caused by O2•− essentially involve the use of (i) Cyt c heme protein, (ii) SOD enzyme,
1168
and (iii) DNA. Superoxide anion determination using a Cyt c-based sensor was reported not
1169
to be so selective mainly because this heme protein is not specific for O2•− and can
1170
simultaneously reduce endogeneous H2O2 in biological systems, whereas SOD-based
1171
biosensors were claimed to be more selective and sensitive.115 When using DNA-based
47 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 48 of 105
1172
bioelectrosensors, the oxidizibility of DNA bases (i.e. maintaining DNA integrity) in the
1173
absence and presence of antioxidants was taken as a measure of antioxidant activity.
1174
(i) O2•− was produced by the xanthine-XOD enzyme system, using a Cyt c-modified
1175
electrode, where Cyt c was reported to communicate with the nano-Au electrode through self-
1176
assembled monolayers of short chain alkanethiols. The immobilised Cyt c was reduced by
1177
O2•− and rapidly regenerated at the surface of the electrode polarized at the oxidation
1178
potential, where the current generated by electron transfer from the radical to the electrode
1179
through Cyt c was proportional to the radical concentration.116 In this process, antioxidants,
1180
when present, quenched the radicals and decreased their concentration reflected in a decrease
1181
of oxidation current. Using this bioelectrosensor, an antioxidant activity sequence was
1182
established for flavonoids in decreasing order: flavanols > flavonols > flavones > flavonones
1183
> isoflavonones.117
1184
(ii)
1185
monolayers) Au nanoparticles-coated carbon fiber microelectrodes, enabling direct electron
1186
transfer with O2•−,118 catalyzing their dismutation to O2 and H2O2 :
1187 1188 1189 1190 1191 1192
SOD enzyme was immobilized onto cysteine-functionalized (via self-assembled
SOD (Cu2+) + O2•− → SOD (Cu+) + O2 … (Eq. 22) Eo = + 0.3 V (vs Ag/AgCl) SOD (Cu+) + O2•− +2H+ → SOD (Cu2+) + H2O2 … (Eq. 23) Eo = - 0.2 V (vs Ag/AgCl) Naturally antioxidants, when present, would cause a decrease in O2•− concentration.
1193
Both oxidation and reduction of O2•− could be measured with high sensitivity and selectivity.
1194
Both dismutation products (i.e. H2O2 and O2) could be easily detected simultaneously using
1195
amperometric transducers,115 e.g., the former with an electrodeposited polyprrole covered
1196
glassy carbon microelectrode modified with horseradish peroxidase and the latter with the
1197
same electrode modified with superoxide dismutase enzyme embedded in polymer layers.119
1198
(iii) DNA-based bioelectrosensors worked on the principle of immobilizing (usually calf
1199
thymus−originated) double stranded (ds)-DNA on screen-printed carbon electrodes (SPE), 48 ACS Paragon Plus Environment
Page 49 of 105
Journal of Agricultural and Food Chemistry
1200
followed by the the detection of the guanine oxidation peak between +0.8 and +1.0 V (vs.
1201
Ag/AgCl) by square wave voltammetry. Since the peak current intensity was proportional to
1202
the concentration of DNA base, guanine, the immersion of the DNA-modified electrode into a
1203
Fenton solution (such as Fe(II)+H2O2) produced a signal decrease in the peak current
1204
intensity, whereas the presence of antioxidants restored the signal (demonstrating DNA
1205
integrity) due to hydroxyl radical quenching.115 In order to measure this signal alteration more
1206
effectively, screen-printed carbon was doped with TiO2 nanoparticles, creating a porous
1207
surface structure on which ds-DNA adsorbed better because of specific DNA phosphate-TiO2
1208
interactions.120 A redox mediator, namely tris-2,2’-bipyridine (bipy) ruthenium(II), was
1209
electrooxidized on the electrode surface to subsequently oxidize both the adsorbed ds-DNA
1210
and the antioxidants in solution. Divalent and trivalent ruthenium-bipy species, i.e. RuDNA(II)
1211
and RuDNA(III), represented those redox mediators that were generated in the vicinity of TiO2
1212
nanoclusters. Here, the oxidative damage was produced by [Ru(bipy)3]3+, an efficient oxidant
1213
for the DNA bases, guanine and adenine, that are most sensitive to oxidation. The oxidative
1214
damage on adsorbed DNA was determined by square wave voltammetry via measuring the
1215
current intensity for the oxidation of [Ru(bipy)3]2+ catalyzed by the remaining ds-DNA on the
1216
electrode; antioxidants, when present, had a protective role reducing oxidative damage on
1217
DNA, and were subsequently assayed through competition kinetics by comparing the rate
1218
constants for RuDNA(III) oxidation of antioxidant (in bulk solution) and of adsorbed DNA.120
1219
The recently developed direct current (DC) polarographic assay for AOA estimation is
1220
based on the measurement of anodic current obtained by dropping mercury electrode (DME)
1221
in hydrogen peroxide solution upon the addition of antioxidant compounds. In this DC
1222
polarographic antioxidant assay, the decrease in anodic limiting current of the
1223
[Hg(O2H)(OH)] (hydroxoperhydroxo-mercury(II) complex), formed in alkaline solution of
1224
H2O2, at the potential of mercury oxidation, varied with the concentration of added
49 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 50 of 105
1225
antioxidant capable of H2O2 scavenging, and the method was validated against DPPH and
1226
Folin spectrophotometric assays for wines,121 and for teas and herbal infusions.122 Although
1227
the signal decrease originated from the interaction of antioxidant with hydroperoxo-radicals,
1228
the antioxidant did not directly scavenge H2O2 in the absence of Hg, and the decrease of Hg2+
1229
cathodic current agreed with that of anodic current, leading to the assumption that mercury(II)
1230
reduction caused a decrease in concentration of Hg2+ available for hydroxoperhydroxo-
1231
mercury(II) complex formation, bringing about the decrease in its anodic current.123
1232 1233
2.1.2.3. Nanotechnological Methods
1234 1235
Nanotechnological methods of colorimetric TAC assay usually make use of either the
1236
formation or enlargement of noble metal (Au, Ag, etc.) nanoparticles, abbreviated as AuNPs
1237
or AgNPs, upon reaction of Au(III) or Ag(I) salts with antioxidant compounds. The standard
1238
reduction potentials for Au(III)-Au(0) and Ag(I)-Ag(0) redox couples are 1.5 and 0.8 V,
1239
respectively, and therefore many phenolic antioxidants can be favourably oxidized by
1240
simultaneous reduction of Au(III) and Ag(I) to the corresponding noble metal nanoparticles
1241
(i.e., AuNPs and AgNPs). However, the thermodynamic favourability of these redox reactions
1242
does not guarantee perfect heterogeneous-phase kinetics, and therefore enlargement of
1243
previously formed NP seeds (e.g., citrate-stabilized AgNP seeds) via reaction with
1244
antioxidants is usually preferred, because coating of these NP seeds gives rise to better
1245
linearity of absorbance/concentration responses. The reason that nanomaterial-based methods
1246
have found little use in food science (specifically antioxidant research) is probably the
1247
substoichiometric character of the concerned reduction reactions by antioxidants leading to
1248
NP formation.55 There are also other food constituents (in addition to antioxidants) causing
1249
AuNP or AgNP formation that may give rise to interferences in the assays.
50 ACS Paragon Plus Environment
Page 51 of 105
Journal of Agricultural and Food Chemistry
1250
When AgNPs are dispersed in liquid media, they exhibit a strong UV-Vis absorption
1251
band not present in the spectrum of the bulk metal. This surface plasmon resonance (SPR)
1252
absorption is attributed to the collective oscillation of electrons in the conduction band of
1253
these particles in resonance with the wavelength of incoming light, with a periodic change in
1254
electron density at the surface. The SPR absorption of nano-sized particles having near-zero
1255
dielectric constant gives rise to a localized SPR (LSPR) band. Known LSPR sensors typically
1256
monitor shifts in the peak position or absorption in response to local refractive index changes
1257
in the close vicinity of the NP surface. Although AgNPs have very high molar absorptivities
1258
(ε ≈ 3 × 1011 M−1 cm−1)124 and are expected to allow higher sensitivity in optical detection,
1259
this ε is a corrected value calculated on the basis of the molar amount of nanoparticles per unit
1260
volume of bulk metal. Furthermore, ideal sensitivity in NP-based TAC assays cannot be
1261
achieved because of various factors affecting LSPR absorption such as reaction stoichiometry,
1262
particle size and shape, and dielectric constants of both the metal and the surrounding
1263
medium. Since the properties of surface plasmons can be tailored as a result of altering the NP
1264
surface and, specifically, the shell thickness, the seed-mediated particle growth technique was
1265
adopted (Figures 8 & 9) for developing a AgNPs-based TAC assay.55
1266 1267
Figure 8
1268
Figure 9
1269 1270
Since biothiols play a significant biological role among these compounds due to their strong
1271
reductive ability and capacity to quench reactive oxygen and nitrogen species (ROS/RNS),
1272
and Since the redox potential of the oxidized/reduced forms of glutathione (GSSG/2GSH) is a
1273
basic indicator of the redox environment within a cell, selective quantification of biothiols is
1274
an important topic in bioanalytical chemistry.125 An optical sensor for biologically important
51 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 52 of 105
1275
thiol quantification was designed with the use of Ellman’s reagent (DTNB)-adsorbed gold
1276
nanoparticles (AuNPs) (DTNB-Au-NP) in a colloidal solution.126 5,5’-Dithio-bis(2-
1277
nitrobenzoic acid) (DTNB), a well-documented thiol-selective compound, was adsorbed
1278
through non-covalent interaction onto AuNPs, and the absorbance changes associated with the
1279
formation of the yellow-colored 5-thio-2-nitrobenzoate (TNB2-) anion as a result of reaction
1280
with biothiols was measured at 410 nm (Figure 10). The DTNB reagent could be selectively
1281
desorbed from the derivatized AuNPs surface to give Ellman’s reaction with thiols due to
1282
preferential adsorbabilities of thiols over disulfides127 and to thermodynamic/kinetic
1283
favorability of the thiol-exchange reaction. The linear response of the sensor for cysteine,
1284
glutathione, homocysteine, cysteamine, dihydrolipoic acid and 1,4-dithioerythritol was better
1285
than that of Nile Red dye-derivatized AuNPs sensor for thiol determination.128 Common
1286
biological sample ingredients like amino acids, flavonoids, vitamins, and plasma antioxidants
1287
did not interfere with thiol sensing (Figure 10).126 DTNB-adsorbed AuNPs probes provided
1288
higher sensitivity (i.e., lower detection limits, though not at an order-of-magnitude) in biothiol
1289
determination (Figure 10) than conventional DTNB reagent.110
1290 1291
Figure 10
1292
As H2O2 is a cell membrane-permeable biological oxidant, noble metal nanoparticle-
1293
based methods were also developed to measure H2O2 scavenging antioxidant activity. A gold
1294
nanoshell (GNS)-based optical nanoprobe was developed for assessing hydrogen peroxide
1295
scavenging activity of antioxidants. H2O2 was found to enlarge the AuNPs on the surface of
1296
gold nanoshells (GNS)s precursor nanocomposites (SiO2/AuNPs), and the pre-adsorbed
1297
AuNPs served as nucleation sites for Au deposition. AuNPs on the SiO2 cores were enlarged
1298
by increasing concentrations of H2O2, concomitant with spectral changes corresponding to
1299
AuNP plasmon absorption bands, and antioxidants restricted H2O2-mediated formation of
52 ACS Paragon Plus Environment
Page 53 of 105
Journal of Agricultural and Food Chemistry
1300
GNSs, enabling the determination of their inhibitive activity.129 In another example, H2O2-
1301
induced growth of GNSs was inhibited by the addition of phenolic acids, which affected the
1302
peak wavelength of surface plasmon absorption. Among the tested antioxidants, caffeic acid
1303
was found to be the most efficient H2O2-scavenger, whereas trans-cinnamic acid exhibited the
1304
weakest activity.130
1305 1306
2.1.3. Mixed-Mode (ET‒ and HAT‒Based) Methods
1307 1308
Mixed-mode assays are generally based on the scavenging of a stable radical chromophore
1309
(like ABTS•+ and DPPH•) or fluorophore by antioxidants, in which HAT, ET and PCET
1310
(proton-coupled electron transfer) mechanisms may play different roles to varying extents,
1311
depending on pH, solvent, and other reaction conditions. Schaich and coworkers have
1312
recently directed extensive criticisms to ABTS and DPPH assays131-133 on the grounds that
1313
these assays use sterically hindered, N-centered free radicals as targets to antioxidants rather
1314
than biologically active short-lived radicals (e.g., hydroxyl, superoxide and lipid oxyl radicals
1315
having lifetimes ranging from nano- to ten- seconds) and that their action as radical scavenger
1316
should be irrelevant in vivo.134
1317 1318
2.1.3.1. ABTS/TEAC Assay
1319 1320
ABTS/TEAC assays use intensely-colored cation radicals of ABTS•+ as useful colorimetric
1321
probes accepting hydrogen atoms or electrons supplied by antioxidant compounds. Although
1322
lag-phase assays were preferred at the initial stage of method development, the requirements
1323
for reproducibility and minimizing errors later evolved the assay to a decolorization strategy,
1324
where the initially formed (by H2O2/peroxidase, MnO2 or persulfate) and stabilized ABTS•+
53 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 54 of 105
1325
radical was let to react with the added antioxidant, causing an absorbance decrease at the
1326
characteristic wavelengths. Antioxidant capacity is measured as the ability of the test
1327
compound (e.g., Ph-OH) to decrease ABTS•+ color by intercepting initial oxidation and
1328
preventing ABTS•+ production, or reacting directly with the preformed radical cation. Even
1329
when a fixed-time ABTS assay is preferred, the results may greatly vary for the same
1330
compound (e.g., GSH) depending on the oxidizing agent used to generate the stable colored
1331
radical.33
1332 1333 1334 1335 1336 1337 1338
ABTS + oxidizing agent (such as K2S2O8) → ABTS•+ … (Eq. 24) (λmax=734 nm) ABTS•+ + PhOH → ABTS + PhO• + H+ … (Eq. 25) 2.1.3.2. DPPH• Radical Scavenging Assay
1339 1340
The stable chromogen radical DPPH• was first proposed for quantitating antioxidant content
1341
nearly half a century ago, when Blois used the thiol-containing amino acid cysteine as his
1342
model antioxidant.38 Later it was used as a phenol reagent.135 The more recently introduced
1343
method of Brand-Williams and colleagues136 has been used as a reference point by several
1344
groups of workers.137,138 Reaction with DPPH was adapted for measuring radical quenching
1345
kinetics139,140 and since then numerous variations in methods and time for following the
1346
reaction as well as for calculating relative antioxidant action by reaction stoichiometry have
1347
evolved.131,136 The reaction equation can be formulated with respect to HAT-mechanism,
1348
although proton-coupled ET-mechanism cannot be excluded, especially in phenol-ionizing
1349
solvents:
1350 1351 1352 1353
DPPH• + PhOH → DPPH2 + PhO• … (Eq. 27) where DPPH• is a stable chromogen radical with λmax=515 nm.
1354 54 ACS Paragon Plus Environment
Page 55 of 105
Journal of Agricultural and Food Chemistry
1355 1356
2.1.3.3. DMPD Radical Scavenging Assay
1357 1358
In the presence of ferric iron or reactive species such as hydroxyl radicals, N,N-Dimethyl-p-
1359
phenylenediamine dihydrochloride (DMPD) is converted to the colored DMPD•+ radical
1360
cations, which are scavenged by antioxidant molecules present in test samples, forming the
1361
principle of the DMPD•+ assay. Antioxidant compounds which are able to transfer a hydrogen
1362
atom (or an electron) to DMPD+ cause rapid decolorization of the solution (manifested by an
1363
absorbance descrease at λmax=505 nm) with a stable end-point. The use of DMPD+ has been
1364
widely extended to evaluate the antioxidant capacity of different food products such as fruits,
1365
vegetables and wine.141-143 However, less reproducibility was obtained in the presence of
1366
hydrophobic antioxidants such as tocopherol or BHT.144 Mehdi and Rizvi modified the
1367
DMPD-based method for the measurement of plasma oxidative capacity during human
1368
aging.145 Recently, in a population-based cohort study from Germany, Schöttker et al.
1369
measured the derivatives of reactive oxygen metabolites (d-ROM) in human sera as a proxy
1370
for ROS concentration with the use of DMPD-based d-ROM colorimetric kit.146 Çekiç et al.
1371
simultaneously measured the oxidative status (OS) and antioxidant activity with the aid of a
1372
sensor technique by retaining the pink-colored, positively-charged chromophore of DMPD-
1373
quinone (resulting from the reaction between DMPD and ROS) on a Nafion membrane, where
1374
the 514 nm-absorbance of the sensor membrane was a measure of OS (e.g., derived from
1375
hydroperoxides and labile iron compounds of sera) while antioxidants caused an absorbance
1376
decrease on the membrane due to their ROS scavenging action.147
1377
Other radical probes used for free radical scavenging activity measurement are
1378
Fremy’s salt (galvinoxyl radical: potassium nitrosodisulfonate)148 and the more recently
1379
developed
aroxyl
radical
(2,6-di-tert-butyl-4-(4′-methoxyphenyl)phenoxyl
55 ACS Paragon Plus Environment
radical)
Journal of Agricultural and Food Chemistry
Page 56 of 105
1380
methods,149 but these techniques have been much less frequently preferred than the widely
1381
used ABTS and DPPH assays.
1382 1383
2.1.4. ROS/RNS Scavenging Methods
1384 1385
ROS is a collective term often used to include oxygen radicals [superoxide (O2•-), hydroxyl
1386
(•OH), peroxyl (ROO•) and alkoxyl (RO•)] and certain nonradicals that are either oxidizing
1387
agents and/or are easily converted into radicals, such as HOCl, ozone (O3), peroxynitrite
1388
(ONOO-), singlet oxygen (1O2) and H2O2. RNS is a similar collective term that includes nitric
1389
oxide radical (•NO), ONOO-, nitrogen dioxide radical (•NO2), other oxides of nitrogen and
1390
products arising when NO reacts with O2•-, RO• and ROO•.150 Although ROS and RNS are
1391
essential to human health, an unbalanced excess of these reactive species may cause oxidative
1392
stress‒related diseases.7,151
1393
In ROS/RNS scavenging activity assays, the reactive species generated enzymatically
1394
or by redox-active chemical reagents are allowed to attack a probe, and the subsequent
1395
conversion of this probe is measured spectroscopically or electrochemically, where the extent
1396
of conversion on the probe is a measure of ROS/RNS concentration and its attenuation
1397
indicates the scavenging activity of the tested putative antioxidants. Most conventional
1398
fluorescence (FL) probes for ROS detection react by a free radical mechanism and are rather
1399
non-selectively converted into FL-enhanced or FL-quenched end products, whereas the more
1400
recently developed boronate probes may act more selectively as they undergo nucleophilic
1401
attack by oxidant species to release the hidden fluorescence of a fluorophore containing a
1402
blocking group. Biological macromolecules such as lipids, proteins, sugars and DNA have
1403
been subjected to the attack of biologically relevant reactive species and used as oxidative
1404
stress biomarkers, where various oxidative transformations on these macromolecules have
56 ACS Paragon Plus Environment
Page 57 of 105
Journal of Agricultural and Food Chemistry
1405
been measured. For example, ethane and pentane, conjugated dienes, hydroperoxides,
1406
aldehydes and ketones from lipids; nitro-, chloro-, and bromo-amino acid residues, as well as
1407
carbonyls, -SS-, SOH, and -SOOH compounds, dimers, cross-linked, modified and cleavage
1408
products from proteins; nitrated, chlorinated, brominated and 8-hydroxylated nucleic bases
1409
from DNA can be formed upon oxidation.7,58 In the case of biomarkers, the activity of an
1410
antioxidant can be indirectly mesured by its attenuation of the oxidative hazard formed on the
1411
marker macromolecule. For example, antioxidants may decrease the color intensity of ferric
1412
thiocyanate or ferric-xylenol orange complexes used to measure lipid hydroperoxides, or of
1413
the TBARS chromophore used to measure lipid-derived aldehydes and ketones (e.g.,
1414
malondialdehyde) with questionable specificity.
1415
The selectivity of ROS/RNS assays has been questioned, because frequently there are
1416
more than one reactive species capable of probe conversion, and absorptimetric measurements
1417
in the UV-range (such as the hydrogen peroxide scavenging assay carried out at the intrinsic
1418
maximal absorption wavelength of H2O2 at λ=230 nm)152 suffer from the interference of many
1419
organic compounds absorbing at the same wavelength, thereby yielding inconsistent blanks.
1420
The redundancy of hydroxyl radical scavenging assays have occasionally been discussed in
1421
literature because almost any biological molecule (not necessarily an antioxidant) can react
1422
with this radical at extremely high rates. The possible drawbacks of ROS/RNS scavenging
1423
assays may be listed as follows: (i) if the tested reactive species are produced enzymatically
1424
(e.g., superoxide anion radical by xanthine/xanthine oxidase, reactive species from H2O2 by
1425
peroxidase, HOCl from H2O2 and chloride by myeloperoxidase, reduction of residual nitrate
1426
to nitrite in the course of nitric oxide radical scavenging by NADH-dependent nitrate
1427
reductase), then it is not clear how to differentiate between the ROS/RNS scavenging and
1428
enzyme inhibition actions of antioxidants, and therefore ESR/spin trap methods should
1429
accompany conventional spectroscopic methods in such ambiguous cases; (ii) the tested
57 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 58 of 105
1430
reactive species should not react too rapidly with the selected probe, because in that case, a
1431
wide range of antioxidants reacting with quite different rates cannot be measured. For
1432
example, superoxide radical reacts with the cytochrome c probe much faster than it does with
1433
NBT, rendering the competition of certain antioxidants with the former probe much less
1434
efficient;153 (iii) the tested antioxidant or its oxidation product can enter a direct redox
1435
reaction with the probe via incompletely reacting with the subject ROS/RNS. For example,
1436
ascorbic acid can easily reduce ferricytochrome c in superoxide radical scavenging assay.21
1437
Another example is that in the HOCl scavenging assay in which TNB is oxidized by HOCl to
1438
DTNB, thiol-type antioxidants react with DTNB to produce excessive TNB chromophore;154
1439
(iv) the probe or its conversion product may be unstable, or may itself generate reactive
1440
species during the course of measurement. As a result, such assays should be cautiously
1441
interpreted because they are prone to errors from the above mentioned and other sources,155
1442
and quantification of ROS/RNS scavenging is additionally complicated by the esssentially
1443
non-linear character of either ROS production or consumption with respect to concentration.
1444
Another challenge for the synthesis of future ROS/RNS probe molecules is tailoring them at a
1445
suitable redox potential to specifically quench a given reactive species but not others.
1446 1447
2.1.5. Cellular Antioxidant Activity Assays
1448 1449
Since cellular antioxidant activity assays (CAA) are performed within the cell medium, they
1450
are assumed to better consider certain physico-chemical aspects of the medium such as the
1451
uptake, distribution and metabolism of antioxidants within cells.156 López-Alarcóna and
1452
Denicola157 have recently reviewed cellular antioxidant activity assays in comparison to
1453
classical chemical assays, and to compensate for the possible complexities of antioxidant
1454
action, have recommended the use of CAA to assess the genuine antioxidant activity of a
58 ACS Paragon Plus Environment
Page 59 of 105
Journal of Agricultural and Food Chemistry
1455
compound or extract. To measure CAA, Wolfe and Liu156 used the peroxyl radical oxidation
1456
reaction of the non-fluorescent probe 2’,7’-dichlorofluorescin (DCFH) entrapped in human
1457
hepatocarcinoma HepG2 cells, where the presence of antioxidants attenuated the cellular
1458
fluorescence of the oxidation product, dichlorofluorescein (DCF). This probe may exhibit
1459
several disadvantages such as photochemical instability, incomplete trapping by cells, or
1460
decreased oxidation in the presence of endogenous antioxidants.156 Also, the classical order of
1461
antioxidant effectiveness was not followed in these assays, and the results did not correlate
1462
with those of ORAC.156,158 Halliwell and Whiteman18 directed certain criticisms to cellular
1463
assays, such as interference of enzymic/non-enzymic endogenous antioxidants to the
1464
measurement procedure, intrinsic oxidative stress generation by the cell culture process,
1465
difficulty in distinguishing between intracellular and extracellular fluorescence from chemical
1466
reactions in the culture medium, and inability of DCF fluorescence measurement to
1467
specifically differentiate several reactive species. Effective use of cellular probes also
1468
necessitate a full understanding of the involved mechanistic pathways, environmental factors
1469
(e.g., O2 and pH), distribution and possible intermediary products of the probe, combined with
1470
instrumental artefacts, and actual competition of the measured antioxidants with the probe for
1471
reactive species.159
1472 1473
3. PHYSICO-CHEMICAL ASPECTS OF ANTIOXIDANT ACTION
1474 1475
3.1. Kinetic Solvent Effects and Structure-Activity Relationships
1476 1477
Antioxidant activity/capacity assays should be tested with several reactive species and
1478
oxidizing probes, because the reactivity of known antioxidants toward different ROS/RNS are
1479
quite different. For example, N-acetylcysteine is a powerful scavenger of HOCl, and also
59 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 60 of 105
1480
reacts with hydroxyl radical with a rate constant of 1.36 × 1010 M−1s−1, but only reacts slowly
1481
with H2O2, and does not react at all with superoxide anion radical.160 Antioxidant activity also
1482
strongly depends on the solubility/localization and distribution of antioxidants between
1483
different phases; for example, restricted diffusion of α-tocopherol may reduce its antioxidant
1484
activity in membranes, and a synergistic effect between ascorbic acid and α-tocopherol was
1485
observed under conditions of inhibited peroxidation of linoleate in SDS micelles. A water-
1486
soluble form of α-tocopherol complexed with bovine serum albumin (α-toc:BSA) proved to
1487
be an effective antioxidant for hindering the autoxidation of linoleate in SDS micelles,
1488
whereas (α-toc:BSA) required a long equilibration time with liposomes before α-toc was
1489
transferred to the liposomes to provide effective antioxidant action.161 Antioxidants shown to
1490
be very effective in in vitro TAC assays may exhibit a conflicting order of antioxidant
1491
potency in different systems of varying lipophilic/hydrophilic balance, depending on their
1492
differential abilities to penetrate and interact with the lipid bilayers. A hierarchic order (i.e.
1493
pecking order) of antioxidants in relation to their free radical scavenging ability can be
1494
predicted, based on the comparison of one-electron reduction potentials of the scavenged
1495
reactive species with the corresponding reduction potentials of aryloxy/phenol redox couples.
1496
For example, the formal reduction potentials (at pH=7) of reactive species such as •OH (2.31
1497
V), alkoxyl radical (1.60 V), alkyl peroxyl radical (1.00 V), superoxide anion radical (0.94
1498
V), singlet oxygen (0.65 V), unsaturated fatty acid radical (0.60 V), and H2O2 (0.32 V) are
1499
listed in a comprehensive review.162 Hydrogen peroxide and superoxide anion radical can act
1500
as both an oxidant and reductant, depending on the substrate and medium. Moreover, the
1501
average lifetimes of radicals commonly active in living tissues and in foods vary over ten
1502
orders of magnitude, i.e. •OH (10-9 s), •OR (10-6 s), and ROO• (10 s).132 Thus, disregarding the
1503
distribution between phases, it is natural that different antioxidants would show different
1504
ROS/RNS scavenging abilities based on their redox potentials. In spite of their different
60 ACS Paragon Plus Environment
Page 61 of 105
Journal of Agricultural and Food Chemistry
1505
distribution between lipid membrane and aqueous phases, the fact that vitamin C (ascorbic
1506
acid) can physiologically regenerate vitamin E (α-tocopherol) can be understood by
1507
considering the reduction potentials of α-tocopheroxyl (0.50 V) and ascorbate (0.282 V)
1508
radicals, because α-tocopheroxyl radical can oxidize ascorbate, and is itself converted back to
1509
α-tocopherol. Likewise among flavonoids, quercetin and tea catechins, having reduction
1510
potentials of the corresponding aryloxy radical/phenol couple less than 0.4 V, should be able
1511
to regenerate α-tocopherol from α-tocopheroxyl radical.163 However, it should be borne in
1512
mind that, although a relatively large difference between the standard reduction potentials of
1513
the oxidant and reductant ensures a high equilibrium constant for a nearly complete redox
1514
reaction between the concerned antioxidant and the reactive species it scavenges, it does not
1515
guarantee a high kinetic rate in a given medium, since equilibrium constant is the ratio of the
1516
rate constant of the forward reaction to that of the backward reaction.
1517
The rate and extent of lipid peroxidation can be measured from oxygen uptake,
1518
substrate consumption, or product formation.58 Some secondary oxidation products (e.g.,
1519
hexanal and 2,4-decadienal) or transition metal ions can act as prooxidants and catalyze lipid
1520
oxidation at initial stages, causing early consumption of antioxidants.164 In Cu(II)-induced
1521
lipid peroxidation reactions, the kinetic profile of peroxidation is characterized by three major
1522
parameters: the ‘lag time’ preceding rapid oxidation, the maximal rate of oxidation (Vmax),
1523
and the maximal accumulation of oxidation products (ODmax), and a distinction between
1524
various antioxidants with respect to their mechanism of action can be made based on
1525
observing the different impacts upon these three parameters; for example, antioxidative
1526
effects due to copper-chelating or blockade of copper binding sites can be distinguished from
1527
the effects of free radical quenching.165 The effects of three different flavonoids of similar
1528
structure, i.e. quercetin, morin, and catechin, as potential antioxidant protectors were studied
1529
in a linoleic acid emulsion to yield an inhibitive order of : morin > catechin ≥ quercetin, i.e.
61 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 62 of 105
1530
agreeing with that of formal reduction potentials versus NHE at pH 7, i.e. 0.60, 0.57 and 0.33
1531
V for morin, catechin, and quercetin, respectively. Morin showed antioxidant effect at all
1532
concentrations whereas catechin and quercetin showed both antioxidant and prooxidant
1533
effects depending on their concentrations. The structural requirements for antioxidant activity
1534
in flavonoids interestingly coincided with those for Cu(II)-induced prooxidant activity,
1535
because as the reducing power of a flavonoid increases, Cu(II)–Cu(I) reduction is facilitated
1536
that may end up with the production of reactive species.166
1537
The thermodynamics and kinetics of antioxidant action cannot be properly understood
1538
without understanding solvent effects on HAT- and ET-based reactions, especially on the
1539
latter.167-169 Although HAT-based reactions have been claimed to be relatively rapid at least at
1540
their initial stages, hydrogen bonding in polar solvents may induce dramatic changes in the H-
1541
atom donor activities of phenolic antioxidants and consequently affect the measured reducing
1542
antioxidant capacity.170,171 The rates of oxidation reactions of phenolic compounds (either by
1543
HAT- or proton-coupled electron-transfer (PCET) mechanism) by ROS/RNS are deeply
1544
influenced by H-bond accepting (HBA) and anion solvation abilities of solvents, as well as by
1545
the nature and position of phenol ring substituents (for instance, the rate constants (k, M-1s-1)
1546
for oxidation of phenoxide anions by ClO2 radical, having a standard potential of 0.94 V
1547
versus NHE, were in the descending order of 109, 108, 107, 105 and 103 for resorcinol, p-
1548
methoxyphenol, simple phenol, p-nitrophenol, and p-cyanophenol, respectively, progressively
1549
decreasing with the removal of electron-donating substituents and/or with the increase in
1550
electron-withdrawing substituents on the phenolic ring);172 there is a delicate balance between
1551
the three different, nonexclusive mechanisms of antioxidant action, namely HAT, proton-
1552
coupled ET, and sequential proton loss ET, depending on both the environment and the
1553
reactants.77,173 Choe and Min also discussed the effects of substituents and steric crowding
1554
around the phenolic hydroxyl groups as important parameters determining antioxidant
62 ACS Paragon Plus Environment
Page 63 of 105
Journal of Agricultural and Food Chemistry
1555
activity.174 As an example of the HBA ability of solvent on TAC, in the study of the effects
1556
of polar (e.g., acetonitrile and tert-butyl alcohol) and nonpolar (e.g., cyclohexane) solvents on
1557
the peroxyl-radical-trapping antioxidant activity of some flavonoids, catechol derivatives,
1558
hydroquinone, and monophenols, phenols with two ortho-hydroxyls were the most active
1559
antioxidants, with inhibition rate constants (kinh) in the range of (3−15) × 105 M-1s-1 (in
1560
cyclohexane), whereas in the strong H-bond acceptor solvent tert-butyl alcohol, these rate
1561
constants dramatically declined; however, in the weaker H-bond acceptor solvent acetonitrile,
1562
kinh values were restored close to the values in cyclohexane.175 3,5-Di-tert-butylcatechol, a
1563
very active H-atom donor to DPPH in hexane, showed a dramatic loss of activity in HBA
1564
solvents, especially acetone.176 In general, the aryloxy radicals formed from the oxidation of
1565
catechol (o-dihydroxy phenol) moieties of phenolic compounds are stabilized in non- or weak
1566
hydrogen-bonding solvents by intra-molecular H-bonding, through the interaction of two
1567
adjacent substituents on catechol, i.e., –C(O•)…(HO)C–, bringing about the delocalization of
1568
the odd electron over the whole molecule. An extended delocalization and conjugation of the
1569
π-electrons, enhanced by resonance effects and planarity, favor the lowering of both
1570
ionization potentials (IP) and PhO-H bond dissociation energies (BDE), and affect strongly
1571
the capacity of antioxidants to donate a single electron.177 It should be emphasized that the
1572
relative magnitudes of BDE and IP determine whether the HAT- or ET-mechanism is
1573
predominant for a given PhOH (and these values may show significant variations in polar or
1574
nonpolar solution and gas phases), as a low BDE is required for a strong phenolic antioxidant
1575
essentially acting by H-atom donation whereas a low IP is necessary for a strong antioxidant
1576
functioning essentially via electron-donation.177 The intra-molecular H-bonding stabilization
1577
of the one-electron oxidized catechols will also lower the standard redox potential of the
1578
aryloxy radical/catechol couple, making the phenolic compound a stronger antioxidant in ET-
1579
reactions.173 The strong antioxidant properties of catechol or pyrogallol moieties of
63 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 64 of 105
1580
polyphenols may be predominantly attributed to the intra-molecular H-bonding stabilization
1581
of aryloxy radicals produced from one-electron oxidation of these moieties,178 whereas the
1582
inter-molecular H-bonding abilities of phenolic hydroxyl groups with HBA solvent molecules
1583
lowers antioxidant activity. The antioxidant activities of o-methoxyphenols were decreased in
1584
hydrogen bond accepting media.179 Thus, in evaluating H-atom transfer kinetics of
1585
polyphenols, the ability to form a linear H-bond (with a solvent (S) molecule, in the form of
1586
ArOH…S) should be distinguished from a non-linear H-bond (intra-molecular H-bond),
1587
because only the former may restrict H-atom abstraction from a phenolic compound by a free
1588
radical.180 Most of the PhO-H bond energy differences in H-bonding and non-bonding
1589
solvents (calculated from the enthalpy of the reaction between di-tert-butyl peroxide and
1590
phenol) could be accounted for from the known hydrogen bonding equilibrium between the
1591
solvents and the phenol.181 For strong antioxidant activity, the presence of o-dihydroxy phenol
1592
(catechol) moiety in an antioxidant compound has an additional advantage of chelating
1593
transition metal ions such as iron and copper, thereby hindering transition metal-initiated
1594
Fenton-type reactions generating reactive species. On the other hand, the abnormally high rate
1595
constants in alcohols for H-atom abstraction from 13 hindered and nonhindered phenols by
1596
DPPH• were attributed to the partial ionization of the phenols (Ph-OH) in alcohols and a very
1597
fast electron transfer from phenoxide anion (Ph-O−) to DPPH• ; this also applies for low pKa
1598
phenols in non-hydroxylic polar solvents like di-n-butyl ether, acetonitrile, tetrahydrofurane
1599
and dimethyl sulfoxide.182 The rate constants for DPPH oxidation of phenols in alcohols were
1600
increased by the addition of sodium methoxide and were decreased by added acetic acid. The
1601
initial fast chemistry of the oxidation of curcumin with DPPH• was attributed to the presence
1602
of curcumin anions present at low equilibrium concentrations in alcohols, whereas upon rapid
1603
depletion of these ‘preformed’ anions, ionization of curcumin became partially rate-limiting.77
1604
In this regard, the TEAC coefficients (with respect to the CUPRAC method) of the
64 ACS Paragon Plus Environment
Page 65 of 105
Journal of Agricultural and Food Chemistry
1605
antioxidant compounds: quercetin, catechin, and BHT were higher in pure MeOH than in pure
1606
EtOH, probably due to facilitated electron-transfer in ionizing solvents capable of anion
1607
(phenolate) solvation, because MeOH is the alcohol that best supports ionization.173 The
1608
structure of the B ring in flavonoids seemed to be the primary determinant of antioxidant
1609
activity when studied through fast reaction kinetics with an oxidizing reagent such as
1610
ABTS•+.183 When the H-atom donating ability of 15 flavonoids were studied by ESR, the
1611
second-order reaction rates, primarily governed by O−H bond dissociation energies, were
1612
myricetin > morin > quercetin > fisetin catechin > kaempferol ≈ luteolin > rutin > d-α-
1613
tocopherol > taxifolin > tamarixetin > myricetin 3’,4’,5’-trimethyl ether > datiscetin >
1614
galangin > hesperitin ≈ apigenin.184 Antioxidant effectiveness and reactivity were highly
1615
dependent on the configuration of -OH groups on the flavonoid B and C rings, with very
1616
minor contribution from the A ring. The rate constants for the free radical scavenging action
1617
of flavone, chrysin, and flavonol were low, indicating that the reactivities of 5- and 7-OH
1618
groups at A-ring and 3-OH group at C-ring were very weak and almost negligible; rutin and
1619
quercetin with 3’- and 4’-OH groups at B-ring showed high reactivity, indicating that the o-
1620
dihydroxyl (catechol) structure in the B-ring was the obvious radical target site for
1621
flavonoids.185 Highest reaction rates and stoichiometries were observed with flavonols
1622
capable of being oxidized to ortho-quinones or extended para-quinones.184 The stabilizing
1623
effect of electron-donating groups near phenolic hydroxyls may be exemplified in the much
1624
lower reactivity of 4-methoxy-2,3,5,6-tetramethylphenol (TMMP) than that of α-tocopherol
1625
against peroxyl radicals, because in TMMP, the methoxy group is twisted out of the plane of
1626
the aromatic ring by steric forces and consequently, the p-type lone pair on the methoxyl
1627
oxygen cannot help stabilize the phenoxy1 formed upon abstraction of the phenolic hydrogen,
1628
whereas in tocopherols, the chroman ring system holds the ethereal oxygen’s p-type lone pair
1629
nearly perpendicular to the aromatic ring, thereby providing additional stabilization for the
65 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 66 of 105
1630
resultant phenoxyls.186 The structural requirements for strong antioxidant action with respect
1631
to both thermodynamic efficiency and kinetic rate have been excellently reviewed by Rice-
1632
Evans et al.70 The three criteria for effective radical scavenging of flavonoids were
1633
established as the o-dihydroxy structure in the B ring, the 2,3-double bond in conjugation with
1634
a 4-oxo function in the C ring, and the 3- and 5-OH groups with 4-oxo function in A and C
1635
rings, as characteristically demonstrated in the strong antioxidant flavonoid: quercetin.69,187
1636
Flavanonols and flavanones, due to the lack of conjugation enhancing resonance stabilization
1637
over the entire molecule, are weak antioxidants.163 The half-peak oxidation potentials (Ep/2)
1638
of flavonoids less than 0.2 V are defined as readily oxidizable and therefore good scavengers
1639
of reactive species.163 The reduction potentials (at pH=7) of the aryloxy radical/phenol couple
1640
for quercetin (0.33 V) and myricetin (0.36 V) are lower than those for catechin (0.57 V),
1641
luteolin (0.60 V) and kaempferol (0.75 V), making the former two flavonoids stronger
1642
antioxidants in most in vitro tests. The high antioxidant activities of phenolic and
1643
hydroxycinnamic acids may be attributed to the number and position of phenolic hydroxyls,
1644
the presence of electron-donating substituents (e.g., methoxy groups) in ortho- and para-
1645
position relative to the phenolic –OH thereby stabilizing the produced aryloxy radicals via
1646
oxidation, and the double bonds in side chains. For example, the order of antioxidant
1647
effectiveness of hydroxycinnamates on the induction period of autoxidizing fats was found as:
1648
caffeic > ferulic > p-coumaric acid,188 in accordance with the ET-based CUPRAC results (but
1649
not with the ABTS/TEAC results, in which caffeic acid proved to be less effective than ferulic
1650
and p-coumaric acids).
1651
Recent theoretical analyses of antioxidant action revealed that the kinetics for the free
1652
radical scavenging of polyphenols (formerly assumed to consist of merely HAT and ET
1653
modes) can be further subdivided into three basic mechanisms involving H-atom, proton (H+)
1654
and electron (e−) transfer:189
66 ACS Paragon Plus Environment
Page 67 of 105
Journal of Agricultural and Food Chemistry
1655
(i) Pure HAT and PCET (proton-coupled electron transfer): in pure HAT, the phenolic proton
1656
and electron of the donated H-atom (PhO-H) are transferred to the same atomic orbital of the
1657
free radical whereas PCET involves several molecular orbitals; however, in both HAT and
1658
PCET, the proton and electron are transferred in one kinetic step, shown by the equation.
1659
PhOH + R• → PhO• + RH
1660
In fact, HAT may be visualized as a special case of concerted PCET involving electronically
1661
adiabatic proton transfer. The thermochemistry of PCET and its implications have been
1662
excellently reviewed by Warren et al.190
1663
(ii) ET-PT (electron transfer−proton transfer) is a two-step mechanism started by an e−
1664
transfer and followed by a H+ transfer:
1665
PhOH + R• → PhOH•+ + R−
1666
PhOH•+ + R− → PhO• + RH
1667
In case when PT is very fast, ET-PT is reduced to a HAT process.
1668
(iii) SPL-ET (sequential proton loss−electron transfer), is quite different from HAT, and
1669
occurs (in three consecutive steps) by following the reverse order of ET-PT: it starts with a
1670
proton release (acidic dissociation) forming a phenolate anion, which subsequently transfers
1671
an electron. SPL-ET is favored when the phenolate anion (PhO−) remains stable during the ET
1672
step before reprotonation.
1673
PhOH ↔ PhO− + H+
1674
PhO− + R• → PhO• + R−
1675
R− + H+ → RH
1676
Although these three mechanisms are thermodynamically equivalent (in terms of Gibbs free
1677
energy change), the competition between the different mechanisms is governed by the
1678
kinetics of the limiting step of each mechanism (atom transfer for PCET and electron transfer
1679
for both ET-PT and SPL-ET). By performing quantum chemical calculations on the primary
67 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 68 of 105
1680
roles of the o-dihydroxy catechol moiety and the 3-OH group of quercetin as scavengers of
1681
different types of free radicals, Di Meo et al. reached the conclusion that in nonpolar
1682
environments (e.g., lipid bilayer membranes) and at low pH (e.g., as in the stomach), PCET is
1683
the only possible mechanism, whereas in polar solvents and at pH where quercetin is partly
1684
deprotonated (e.g., in blood plasma), the faster and therefore more predominant process
1685
(SPLET) effectively competes with PCET.189 Thus, the contribution of SPLET to the overall
1686
antioxidant activity of quercetin (also extrapolable to those of other flavonoids) was reported
1687
to increase with a rise in pH where phenols are dissociated to phenolates.189 Amić et al.
1688
theoretically investigated the double-PCET (i.e. 1H+/1e− and 2H+/2e−) processes of free
1689
radical scavenging by flavonoids, and stressed that the significant contribution of the second
1690
PCET mechanism, resulting in the oxidative formation of a quinone/quinone methide, can
1691
effectively distinguish the active flavonoids from inactive ones.191 In other words, the double-
1692
PCET descriptors such as the second O–H bond dissociation enthalpy related to HAT
1693
mechanism in the gas phase and lipid media, and the second electron transfer enthalpy related
1694
to SPLET mechanism in an aqueous (polar) medium were found to better describe the
1695
quantitative structure-activity relationships of the studied set of 21 flavonoids.191 Recently,
1696
Bakhouche et al. theoretically investigated the antioxidant activity of four forms of tocopherol
1697
using the HAT, ET-PT and SPLET mechanisms, and calculated the O-H bond dissociation
1698
free energy (BDFE), ionization potential (IP), proton dissociation free energy (PDFE), proton
1699
affinity (PA) and electron transfer free energy (ETFE) parameters in the gas phase and solvent
1700
media.192 Alpha-tocopherol was shown to be the most reactive form in gas phase with the
1701
lowest BDFE, IP, and PA values. In the gas phase and non-polar media, HAT mechanism was
1702
predominant due to the fact that BDFE was lower than PA and IP, whereas in aqueous and
1703
polar solvents (e.g., DMSO and MeOH), SPLET was more effective because the PA of all
1704
forms of tocopherol was considerably lower than BDFE and IP.192
68 ACS Paragon Plus Environment
Page 69 of 105
Journal of Agricultural and Food Chemistry
1705
In ET-based TAC assays, kinetic solvent effects work in a different way. If the redox
1706
couple of the TAC assay reagent is a coordinatively saturated metal complex (involving
1707
different oxidation states of a given metal ion in the same ligand environment such as
1708
bis(neocuproine)copper(II,I), tris(1,10-phenanthroline)iron(III,II), hexacyanoferrate(III,II))
1709
capable of outer-sphere electron-transfer with the polyphenol,193 then a minor reorientation
1710
(but not substitution) of the already existing ligands around the central metal ion may
1711
expected in the formation of a transient intermediate during the ET process, and consequently,
1712
the reaction rate may only be affected to a limited extent by solvent polarity. In this case, the
1713
magnitude of solvent effects may be determined by the extent of geometric rearrangement
1714
around the coordination center during ET. However, inner-sphere ET reactions of the assay
1715
reagent (e.g., hexaaqua-solvated Fe3+) with phenolics will naturally be affected by the
1716
hydrogen-bonding behavior of the solvent due to stabilization or inhibition of the
1717
intermediary complex formed during electron-transfer. When other factors are disregarded or
1718
assumed to remain constant, HAT-based TAC assays (e.g., ORAC, TRAP, and mixed-mode
1719
ABTS) are generally affected to a greater extent by the solvent behaviour (polarity, HBA,
1720
etc.) than ET-based methods relying on outer-sphere e-transfer (e.g., CUPRAC, ferricyanide,
1721
and FRAP), provided that the ET-reagent chromophore is soluble at effective concentrations
1722
in the solvent of concern.173
1723 1724
3.2. Partitioning of Antioxidants and the Polar Paradox Hypothesis
1725 1726
The capacity for inhibition of lipid peroxidation was measured in many different systems such
1727
as homogeneous solutions, aqueous miscelle dispersions, liposomal membranes, isolated low-
1728
density lipoproteins (LDLs), red blood cells and plasma. Other than the major parameters of
1729
antioxidant activity/capacity, localization of antioxidant, fate of antioxidant-derived radical,
69 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 70 of 105
1730
interaction with other antioxidants, and mobility of antioxidant at the microenvironment are
1731
also important.58
1732
The polar paradox (PP) hypothesis describing the varying behavior of an antioxidant at
1733
different locations can provide limited information about the actually complex behaviour of
1734
antioxidants in oil, lipid and fatty acid emulsions (in water) due to a number of constraints:
1735
(i) The PP hypothesis simply states that due to different locations of polar and nonpolar
1736
antioxidants in bulk oil and water, nonpolar antioxidants should be more effective than their
1737
polar homologues in oil-in-water emulsions because of their better enrichment at the oil-water
1738
interface where oxidation primarily takes place. Even though this was basically confirmed in
1739
earlier cases, recent experiments carried out by varying the polarity of antioxidants by
1740
esterification with various alkyl chain lengths showed that esterified phenolic antioxidants
1741
(such as those of chlorogenic, rosmarinic and dehydrocaffeic acids, tyrosol and
1742
hydroxytyrosol fatty acids, and rutin) essentially obeyed the PP hypothesis up to C8-C12
1743
carbon atoms chain length, beyond which antioxidant activity sharply declined (called the
1744
“cutoff effect”), not conforming to expectations.194-196 Bulky sized phenolic antioxidants
1745
(such as those with long side chains), despite their hydrophobic character, may also show
1746
lower mobility because of steric hindrance, and thus decreased diffusibility toward reactive
1747
centers at the interface.197 It is also probable that the longer chain antioxidants may form
1748
mixed micelles with emulsifiers used to prepare the emulsions resulting in their migration
1749
away from the emulsion droplets.194-196 Emulsifiers may alter the physical location of
1750
antioxidants and the activity of prooxidants; they can also compete with antioxidants for
1751
localization at the interface.197
1752
(ii) The PP hypothesis disregards the possible prooxidant effects of antioxidants under pH-
1753
and concentration-dependent conditions. The concentration dependency is so important that
1754
the hypothesis may only be applicable when the antioxidant reaches a critical concentration so
70 ACS Paragon Plus Environment
Page 71 of 105
Journal of Agricultural and Food Chemistry
1755
that interfacial phenomena are predominant over solubility effects.197 Some polyphenolics
1756
may act as prooxidants within certain concentration ranges by reducing transition metal ions
1757
to their lower oxidation states thereby enhancing Fenton-type oxidation reactions,4 that may
1758
give rise to ‘false antioxidant activity’ measurement with respect to polarity.197
1759
(iii) Excluding surfactant effects, antioxidants can exhibit unexpected interactions with
1760
oxidation promoters and prooxidants (redox reactions) and with trace metal ions
1761
(complexation or chelation), or even among themselves (e.g., through H-bonding and
1762
association), resulting in synergistic or antagonistic effects to change the kinetics of lipid
1763
peroxidation198 and thus deviate from PP hypothetical expectations.
1764
(iv) If the changes in polarity of the phenolic antioxidant is introduced by adding or removing
1765
ring substituents, then such an addition/removal may actually change the O-H bond
1766
dissociation enthalpy (BDE) and hence the antioxidant potency of the molecule.197
1767 1768
3.3. Possible Interactions Among Antioxidant Constituents
1769 1770
The coexistence of multiple antioxidants usually result in additive effects, whereas synergistic
1771
or antagonistic interactions may appear in extreme cases. Synergism or antagonism occur
1772
when the antioxidative protection/scavenging of two antioxidants is greater or smaller than
1773
the sum of their individual effects, respectively. Choe and Min discussed the possible reasons
1774
of synergism in antioxidant activity.174
1775
Synergy and antagonism are two important issues in food and nutrition science
1776
because of the requirement to choose food combinations exhibiting maximal synergism and
1777
minimal antagonism in antioxidant activity. Antioxidative synergy may arise when two
1778
antioxidants (extendable to a greater number of antioxidants) in admixture show the following
1779
behaviors: (i) they display regenerative action, where the primary (stronger) antioxidant is
71 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 72 of 105
1780
regenerated by the weaker (secondary) one; for instance, carotenoids may be regenerated by
1781
α-tocopherol, which in turn may be regenerated by ascorbate, because the order of reduction
1782
potentials for their corresponding oxidized forms (radicals) are: Ecarotenoid ˃ Etocopherol ˃
1783
Eascorbate.162,199 It should be added that this protective action may also suppress the possible
1784
prooxidative action of tocopherol, because the tocopheroxyl radical at relatively high
1785
concentrations with respect to lipids may abstract an H-atom from lipids and produce lipid
1786
radicals, whereas ascorbate may obstruct this pathway by regenerating tocopherol from
1787
tocopheroxyl radical.200 Likewise, inhibited oxygen uptakes of a combination of α-tocopherol
1788
and flavonoids measured in tert-butyl alcohol confirmed the existence of an antioxidant
1789
interaction which produced a more effective inhibition of lipid peroxidation, and the extended
1790
lag-times provided evidence for the regeneration of α-tocopherol by quercetin;201 (ii) they
1791
exhibit cooperative mechanisms of action, for instance, quercetin may act as a metal chelator
1792
while α-tocopherol functions as a radical scavenger.202 On the other hand, although there is no
1793
systematic theory of antagonistic interactions among antioxidant compounds, it is thought to
1794
arise from the regeneration of the minor (less effective) antioxidant from the major (more
1795
effective) one, competition between the formation of antioxidant-radical adducts, and
1796
changing of the microenvironment suitable for one antioxidant in the presence of the other,174
1797
as observed in binary mixtures of α-tocopherol with either caffeic acid or rosmarinic acid, or
1798
in binary mixtures of caffeic acid with either catechin or quercetin.203
1799
From our own laboratory experience with spectrophotometric ET-based TAC assays,
1800
we observed that if antioxidant concentrations in mixtures are appropriately selected to obey
1801
Beer’s law, then the principle of additivity of absorbances (due to the color change of ET-
1802
reagent) would lead to the addivity of total antioxidant capacities of mixture constituents:
1803
TAC(x1, x2 mixture) = TAC(x1) + TAC(x2) in mmol (or micromol) trolox-equiv./L
72 ACS Paragon Plus Environment
Page 73 of 105
Journal of Agricultural and Food Chemistry
1804
Only in this case, real synergistic or antagonistic interactions (if any) would manifest
1805
themselves. If the results of spectrophotometric ET-based (e.g., CUPRAC, FRAP,
1806
ferricyanide) or mixed-mode (e.g., ABTS and DPPH radical scavenging) assays relying on a
1807
fixed-time end-point are evaluated with the above approach, one would see that most data
1808
assumed to originate from synergistic or antagonistic interactions (via calculations involving
1809
quasi-linear combinations of lag-time or inhibition percentage) are in fact additive,11,64
1810
because in accordance with Beer’s law, absorbance varies linearly with concentration within a
1811
reasonable range, which is not valid for the essentially nonlinear parameters of lag-time or
1812
inhibition percentage;204,205 The synergy observed with the CUPRAC method in (BHT+BHA)
1813
or (BHT+TBHQ) mixtures173 has not yet been fully interpreted, but thought to arise from
1814
enhanced solubilization and stabilization of the semi-quinone radical adducts formed in the
1815
course of oxidation via inter-molecular H-bonding and dimerization. Thiol mixtures
1816
seemingly exhibiting synergy in FRAP and ABTS assays may involve further oxidations in
1817
the presence of multiple thiols (e.g., the partially reversible oxidation reaction expected to
1818
generate disulfides may go further beyond that stage, giving rise to sulfenic and sulfinic
1819
acids).
1820
Synergy and antagonism still remain to be hardly quantifiable behaviors of diverse
1821
antioxidant mixtures. Recently, extensive efforts are being spent for quantifying synergy or
1822
antagonism in antioxidant activity using different reagents and calculation methods. For
1823
instance, Skroza et al. observed synergy in the binary mixtures of resveratrol with catechin
1824
and gallic acid.206 Prieto et al. used a dose-dependent mathematical model based on
1825
probability functions, where the interactive effects between antioxidants were introduced into
1826
the model with simple auxiliary functions that describe the parametric variations induced by
1827
each antioxidant in the presence of the other and a collective index of parametric responses
73 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 74 of 105
1828
was generated, and experimentally demonstrated this effect using DPPH and ABTS assays,
1829
extendable to other ET-assays.207
1830 1831
3.4. Enhanced Solubilization of Antioxidants via Coupled Oxidation Using TAC
1832
Reagents
1833 1834
Since dietary fibre-phenolic compounds (DF-PC) are made up of bound hydroxycinnamic
1835
acids, TAC of cereal fibre has been disputed, and therefore the actual TAC of DF-PC has
1836
been largely underestimated because of their low water and organic solvent solubility. In fact,
1837
a proper assessment of DF-PC antioxidant capacity requires a multiple-step extraction and an
1838
appropriate chemical hydrolysis to release PC and to permit them to exert antioxidant activity
1839
in the in vitro assays. For many years, it was not clear whether the water insoluble DF-PC
1840
fraction could exert any antioxidant action itself, i.e. without any chemical hydrolysis.208
1841
Serpen et al.209 were the first researchers to demonstrate the concept of antioxidant activity of
1842
insoluble material measured in a different way. In reality, the slow and continuous release of
1843
DF-PC from the insoluble material (known to survive for a considerable time) in the human
1844
gastrointestinal tract has been established to occur, and in particular, DF-PC may favorably
1845
act in vivo quenching the soluble radicals that are continuously formed in the intestinal
1846
tract.208 Tufan et al.210 determined the TAC of cereals (i.e. barley, wheat, rye, oat) by the
1847
‘QUENCHER procedure’ (involving forced solubilization of bound phenolics by oxidizing
1848
TAC reagents) with the direct use of copper(II)-neocuproine (Cu(II)-Nc) reagent of the
1849
CUPRAC assay; the assay operated without a need for completely solubilizing or extracting
1850
antioxidant constituents of insoluble matrices before the CUPRAC reaction, because the
1851
driving force for the CUPRAC oxidation of bound phenolics was greater than that for their
1852
solubilization, making the whole coupled-oxidation process thermodynamically favourable.
74 ACS Paragon Plus Environment
Page 75 of 105
Journal of Agricultural and Food Chemistry
1853
The authors effectively measured the TAC values of cereals by their proposed QUENCHER-
1854
CUPRAC assay, and their results linearly correlated with those of reference methods (ABTS
1855
and DPPH). They found additive responses of polyphenolic mixtures in a cellulose matrix
1856
with their proposed method.210 Solvent effects in the QUENCHER approach were recently
1857
discussed.211 Enhanced solubilization of antioxidant compounds via coupled oxidation is an
1858
active area of research, with many potential contributions in food analytical chemistry that
1859
may allow the building of unique databases for determining the antioxidant capacity of foods
1860
having insoluble moieties, e.g., the polysaccharide matrix.212
1861 1862
4. A PRACTICAL GUIDE TO ANTIOXIDANT ASSAY SELECTION
1863 1864
In conclusion, we also propose a rather subjective guide (primarily stemming from our
1865
laboratory’s experiences) for assay selection, briefly summarizing which assay can be chosen
1866
(or not to be preferred) to serve a specific purpose. Ideally, the chosen test should be
1867
sensitive, selective, robust, reproducible, use conventionally available reagents and
1868
instruments, and measure a wide variety of antioxidant types including both lipophilic and
1869
hydrophilic antioxidants, but unfortunately these criteria are not met by a single specific
1870
assay. It is obvious that one has to decide at the start whether antioxidant activity (basically
1871
reaction rate) or antioxidant capacity (stoichiometric conversion efficiency) is to be measured.
1872
Antioxidant activity tests having an area-under-curve (AUC) approach toward fluorescence
1873
decay of a probe attacked by reactive species in the presence of antioxidants (like ORAC) can
1874
simultaneously measure the lag phase, initial rate, and total extent of inhibition of
1875
antioxidants. On the other hand, such tests may not differentiate between reaction rate and
1876
radical scavenging efficiency, and may give positively-biased results for slowly reacting
1877
antioxidants. It is also recommended that test probe selection should be made in accordance
75 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 76 of 105
1878
with the classical definition of an ‘antioxidant’, i.e. the concentration of the tested antioxidant
1879
should be much smaller than that of the probe, a requirement which is not strictly obeyed in
1880
certain HAT-based assays using fluorescent probes.
1881
(a) The working pH is an important parameter of antioxidant assay selection. FRAP,
1882
CUPRAC and Folin methods can be used for acidic (pH 3.6), neutral (pH 7.0) and alkaline
1883
(pH 10) media measurements, respectively. It should be borne in mind that phenolic
1884
antioxidants are undissociated, partly dissociated, and predominantly dissociated at the
1885
working pH values of FRAP, CUPRAC and Folin methods, respectively, possibly reflecting
1886
an underestimation (in the case of FRAP) or overestimation (in the case of Folin) of the in
1887
vitro antioxidant action simulating physiological conditions. Classical ‘ferric reducing assay’
1888
utilizing ferricyanide incubation with the antioxidant followed by ferric ion addition at the
1889
end of the reaction (prior to color development) works ar pH 6.6, while modified ferricyanide
1890
assay initially incorporating Fe(III) in the incubation solution has a pH of 1.7. A mixed-mode
1891
(HAT/ET) free radical scavenging assay like DPPH may also be very sensitive to pH, as the
1892
rate-determining step of proton-coupled electron transfer may be the acidic dissociation of a
1893
phenolic proton, followed by a fast electron-transfer from the phenolate anion to DPPH. In
1894
general, a rise in pH significantly enhances both the rate and efficiency of phenolics oxidation
1895
in ET- and mixed-mode assays, owing to deprotonation.
1896
(b) Use of natural materials in the selected assay may adversely affect the reproducibility of
1897
results. For example, the natural pigment, crocin, or the natural protein, β-phycoerythrin, may
1898
contain impurities and may vary from lot-to-lot in production. The same applies for lipids
1899
(e.g., purified LDL) in testing antioxidative action against lipid peroxidation.
1900
(c) Multi-charged chromophores of TAC redox probes strongly interact with solvent water
1901
molecules by ion-dipole interaction, and therefore respond less to hydrophobic antioxidants.
1902
CUPRAC and ABTS methods, having mono-positive charged chromophores, can
76 ACS Paragon Plus Environment
Page 77 of 105
Journal of Agricultural and Food Chemistry
1903
simultaneously measure hydrophilic and lipophilic antioxidants, while ORAC, FRAP,
1904
ferricyanide and Folin methods having either hydrophilic or multi-charged chromophores
1905
require the use of cyclodextrins (CDs) or micellar solutions to enhance the solubilization of
1906
lipophilics. Unfortunately, cyclodextrins form inclusion complexes with antioxidants, which
1907
may adversely affect subsequent reaction with a redox probe if antioxidant-CD complex is too
1908
stable at relatively high concentrations of CDs necessarily used for solubilization.
1909
(d) Colorimetric reagents having chromophores absorbing distinctly in the visible range of the
1910
electromagnetic spectrum (i.e. as far as possible to the UV range) are the appropriate methods
1911
of choice to deal with the background color of plant pigments and other UV-absorbing
1912
substances. The background color arising from the matrix shows maximal adverse effects on
1913
the precision of color-fading reactions (like DPPH and ABTS) rather than on that of color-
1914
forming reactions (like FRAP, CUPRAC, ferricyanide). Antioxidant activity assays based on
1915
diene formation or hydrogen peroxide scavenging, all involving UV measurements, may be
1916
adversely affected from UV-absorbing interferents. This applies well to most plant extracts
1917
and pharmaceutical preparates, where many constituents strongly absorb in the UV range.
1918
(e) Redox potential of the selected probe is important in antioxidant assay selectivity.
1919
CUPRAC, ABTS and FRAP methods use oxidant probes having 0.60, 0.68 and 0.70 V
1920
reduction potentials (vs NHE), which are at the right potential to oxidize many common
1921
antioxidants lying in the 0.2-0.6 V redox potential range, while conventional ferricyanide test
1922
has an insufficient potential (0.36 V vs NHE) to oxidize certain antioxidants and Folin test has
1923
an indefinitely high potential (enabling the oxidation of citric acid, reducing sugars and some
1924
amino acids, which are not ‘true’ antioxidants). The newly developed colorimetric redox
1925
reagents containing Ce(IV), Cr(VI) and Mn(VII) centers have to be tested for longer periods
1926
of time to see the exact interference results, but it may be speculated that their high redox
77 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 78 of 105
1927
potentials, even when restricted by certain measures, would not enable highly selective TAC
1928
assays as other organics would be co-oxidized.
1929
(f) When metal ion‒containing probes (such as those of FRAP and CUPRAC) are used,
1930
strong chelating or complexing agents may interfere with the assay. For example, we have
1931
observed that EDTA at sufficiently low concentrations to preserve serum samples does not
1932
interfere with the CUPRAC assay, but at higher concentrations, it may partly displace the
1933
neocuproine ligand from the coordination sphere of Cu(II) and decrease the redox potential of
1934
the Cu(II,I) couple, thereby preventing the oxidation of certain antioxidants. The same applies
1935
for FRAP when strongly complexing ligands displace the tripyridyltriazine (TPTZ) ligand
1936
from the coordination sphere of Fe(III) ion.
1937
(g) Metal ion‒containing probes may also suffer from undesired redox cycling of
1938
antioxidants, causing serious errors in TAC calculation. For example, the relative
1939
concentration of free Fe3+ ion in the FRAP reagent is higher than stoichiometrically required
1940
for the 1:2 Fe(III)-TPTZ complex. As the close standard potentials of Fe3+,2+ and
1941
FeIII,II(TPTZ)2 redox couples are 0.77 V and 0.70 V, respectively, significant amunts of Fe2+
1942
may be formed along with FeII(TPTZ)2 upon reduction reaction with antioxidants, and may
1943
subsequently give rise to Fenton-type reactions with H2O2 or dissolved O2. Production of
1944
reactive species may give rise to redox cycling of antioxidants causing erroneous TAC results.
1945
On the other hand, the standard potentials of Cu2+,1+ and CuII,I(Nc)2 redox couples are quite
1946
different in magnitude, being 0.17 V and 0.60 V, respectively, and therefore no redox cycling
1947
is observed because the reduction product of the CUPRAC reagent with antioxidants is
1948
definitely CuI(Nc)2, which may not be reoxidized with H2O2.
1949
(h) The active constituent of FRAP reagent is the ferric-TPTZ complex bearing high-spin iron
1950
which exhibits slow kinetics, and therefore FRAP does not sufficiently respond to thiols
1951
mainly due to kinetic rather than thermodynamic reasons. As a result, due to its inefficiency
78 ACS Paragon Plus Environment
Page 79 of 105
Journal of Agricultural and Food Chemistry
1952
in oxidizing biothiols (including both small-molecule and protein thiols), FRAP method may
1953
not be recommended for biological samples. We have observed in our laboratory that
1954
CUPRAC and ABTS methods give satisfactory results for measuring the TAC values of
1955
serum samples, due to their capability of simultaneously measuring lipophilic and hydrophilic
1956
serum antioxidants together with thiols. The results of ABTS test for biothiols should be
1957
carefully interpreted, because the TEAC coefficients of ABTS test for biothiols vary roughly
1958
between 1.2-1.6 depending on the method of ABTS•+ radical preparation, hinting to the fact
1959
that thiol (RSH) oxidation in ABTS assay proceeds further beyond disulfides (i.e. RSSR,
1960
encountered in CUPRAC test with a TEAC coefficient of roughly 0.5 for RSH) to higher
1961
oxidation products.
1962
(i) Structural limitations may be responsible for the slow kinetics of phenolic antioxidants
1963
having bulky substituents. For example, mixed-mode assays using ABTS and DPPH hindered
1964
radicals may suffer from steric accessibility problems caused by polymeric phenols having
1965
multiple hydroxyl groups and ring adducts (e.g., tannins). Use of assays using outer-sphere
1966
electron-transfer agents (such as CUPRAC, FRAP and ferricyanide) may be more appropriate
1967
in such samples. It should always be remembered that reaction kinetics is a function of
1968
antioxidant structure, pH, temperature and solvent type, and many ET- (e.g., FRAP) or
1969
mixed-mode (ABTS and DPPH) assays cannot go to completion within the protocol time of
1970
the assay.
1971
(j) In competitive assays, one has to avoid ‘short-circuit’ reactions. The tested antioxidant
1972
may react directly with the probe (e.g., thiol-NBT reaction in superoxide scavenging test)
1973
instead of competing with the radical. Similar reactions may occur in enzyme-mediated
1974
generation of reactive species (e.g., peroxidase may directly oxidize ABTS in hydrogen
1975
peroxide scavenging test without adding H2O2).
79 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 80 of 105
1976
(k) Several assays, preferably working in different modes (i.e. HAT, ET, mixed-mode), may
1977
be used together to reveal the whole picture of antioxidative action. It should be borne in
1978
mind that a fast-reacting antioxidant may stay behind a slow-reacting one in antioxidant
1979
ranking when an end-point assay is used because of the fact that the latter quenches a greater
1980
number of free radicals per molecule. Thus it may be advisable to use both AOA and TAC
1981
assays in biologically relevant antioxidant evaluation.
1982 1983
ACKNOWLEDGMENTS
1984 1985
The authors would like to express their gratitude to Istanbul University-Application &
1986
Research Center for the Measurement of Food Antioxidants (Istanbul Universitesi Gida
1987
Antioksidanlari Olcumu Uygulama ve Arastirma Merkezi). One of the authors (R. Apak)
1988
wishes to thank the Istanbul University Research Fund (BAP) for sponsoring his participation
1989
in the Pittcon Analytical Chemistry Meeting (8-12 March, 2015; New Orleans, USA) under
1990
the research project UDP-51475.
1991 1992 1993 1994 1995 1996 1997 1998 1999 2000 2001 2002 2003 2004 2005 2006 2007 2008 2009
LIST OF ABBREVIATIONS AAPH = 2,2’azobis (2-methylpropionamidine) dihydrochloride ABAP = 2,2'-azobis-(2-amidinopropane) dihydrochloride ABTS = 2,2’-azino-bis(3-ethylbenzothiazoline-6-sulfonic acid AOA = antioxidant activity AOX = antioxidants Ag-NPs = silver nanoparticles AUC = area under curve Au-NPs = gold nanoparticles β-PE = β-phycoerythrin CERAC = Ce(IV) reducing antioxidant capacity CHROMAC = Cr(VI) reducing antioxidant capacity CL = Chemiluminescence CUPRAC = cupric reducing antioxidant capacity Cu(II)-Nc = bis (neocuproine) copper(II) cation [Cu(Nc)2]+ = cuprous neocuproine 80 ACS Paragon Plus Environment
Page 81 of 105
2010 2011 2012 2013 2014 2015 2016 2017 2018 2019 2020 2021 2022 2023 2024 2025 2026 2027 2028 2029 2030 2031 2032 2033 2034 2035 2036 2037 2038 2039 2040 2041 2042 2043 2044 2045 2046 2047 2048 2049 2050 2051 2052 2053 2054 2055 2056 2057 2058 2059
Journal of Agricultural and Food Chemistry
CV = cyclic voltammetry DCM = dichloromethane DHBA = dihydroxybenzoic acid DMPD = N,N-dimethyl-p-phenylenediamine dihydrochloride DPPH = 2,2-di(4-tert-octylphenyl)-1-picrylhydrazyl DPV = differential pulse voltammetry DTNB = 5,5’-Dithio-bis(2-nitrobenzoic acid) TNB = 5-Thio-2-nitrobenzoate ECL = electrochemiluminescence EPR = electron paramagnetic resonance ESR = electron spin resonance ET = electron transfer FRAP = ferric reducing antioxidant power GAE = gallic acid-equivalent GC = glassy carbon GSH-Px = glutathione peroxidase GSH-Rx = glutathione reductase HAT = hydrogen atom transfer H2O2 = hydrogen peroxide HOCl = hypochlorous acid HPS = hydrogen peroxide scavenging LMWA = low molecular-weight antioxidants LSPR = localized surface plasmon resonance KMBA = -keto-γ-methiolbutyric acid M-β-CD = methyl-β-cyclodextrin NBT = nitroblue tetrazolium Nc = neocuproine NO = nitric oxide radical NO = nitrogen dioxide radical 2 1 O2 = singlet oxygen O2− = superoxide anion radical OH = hydroxyl radical ONOO− = peroxynitrite anion ORAC = oxygen radical absorbance capacity PCA = principal component analysis PLS = partial least squares PP = polar paradox RO = alkoxyl radical ROO = peroxyl radical ROS = reactive oxygen species RNS = reactive nitrogen species SOD = superoxide dismutase SPE = screen-printed carbon electrode SPR = surface plasmon resonance SRSA = superoxide radical scavenging activity SWV = square-wave voltammetry TAC = total antioxidant capacity TBARS = thiobarbituric acid-reactive substances TBHQ = tert-butylhydroquinone TE = trolox-equivalent 81 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
2060 2061 2062 2063 2064 2065 2066 2067 2068 2069 2070 2071 2072 2073 2074 2075 2076 2077 2078 2079 2080 2081 2082 2083 2084 2085 2086 2087 2088 2089 2090 2091 2092 2093 2094 2095 2096 2097 2098 2099 2100 2101 2102 2103 2104 2105 2106 2107
Page 82 of 105
TEAC = trolox-equivalent antioxidant capacity TOSC = total oxyradical scavenging capacity TPTZ = 2,4,6-tripyridyl-s-triazine TRAP = total peroxyl radical trapping antioxidant parameter XOD = xanthine oxidase
REFERENCES (1) Trevisan, M.; Browne, R.; Ram, M.; Freudenheim, J.; Carosella, A. M.; Armstrong, D. Correlates of markers of oxidative status in the general population. Am. J. Epidemiol. 2001, 154, 348-356. (2) Prior, R. L.; Wu, X.; Schaich, K. Standardized methods for the determination of antioxidant capacity and phenolics in foods and dietary supplements. J. Agric. Food Chem. 2005, 53, 4290–4302. (3) Taverniers, I.; De Loose, M; Van Bockstaele, E. Trends in quality in the analytical laboratory. II. Analytical method validation and quality assurance. Trends Anal. Chem. 2004, 23, 535-552. (4) Niki, E.; Noguchi, N. Evaluation of antioxidant capacity. What capacity is being measured by which method? IUBMB Life 2000, 50, 323-329. (5) Frankel, E. N.; Meyer, A. S. The problems of using one dimensional methods to evaluate multifunctional food and biological antioxidants. J. Sci. Food Agric. 2000, 80, 1925-1941. (6) Halliwell, B. Antioxidant characterization. Methodology and mechanism. Biochem. Pharmacol. 1995, 49, 1341-1348. (7) Niki, E. Assesment of Antioxidant Capacity in vitro and in vivo. Free Radical Biol. Med. 2010, 49, 503-515. (8) Özyürek, M.; Güçlü, K.; Apak, R. The main and modified CUPRAC methods of antioxidant measurement. Trends Anal. Chem. 2011, 30, 652-664. (9) Aruoma, O. I. Methodological considerations for characterizing potential antioxidant actions of bioactive components in plant foods. Mutat. Res. 2003, 523-524, 9-20. (10) Wayner, D. D. M.; Burton, G. W.; Ingold, K. U.; Locke, S. Quantitative measurement of the total peroxyl radical trapping antioxidant capability of human blood plasma by controlled peroxidation. The important contribution made by plasma proteins. FEBS Lett. 1985, 187, 33-37. (11) Demirci Çekiç, S.; Kara, N.; Tütem, E.; Sözgen Başkan, K.; Apak, R. Protein−incorporated serum total antioxidant capacity measurement by a modified CUPRAC (CUPric Reducing Antioxidant Capacity) method. Anal. Lett. 2012, 45, 754-763. (12) Özyürek, M.; Bektaşoğlu, B.; Güçlü, K.; Apak, R. Hydroxyl radical scavenging assay of phenolics and flavonoids with a modified cupric reducing antioxidant capacity (CUPRAC) method using catalase for hydrogen peroxide degradation. Anal. Chim. Acta 2008, 616, 196-206. (13) Demirci Çekiç, S.; Çetinkaya, A.; Avan, A. N.; Apak, R. Correlation of total antioxidant capacity with reactive oxygen species (ROS) consumption measured by oxidative conversion, J. Agric. Food Chem., 2013, 61, 5260-5270. (14) Huang, D.; Ou, B.; Prior, R. L. The chemistry behind antioxidant capacity assays J. Agric. Food Chem. 2005, 53, 1841-1856.
82 ACS Paragon Plus Environment
Page 83 of 105
2108 2109 2110 2111 2112 2113 2114 2115 2116 2117 2118 2119 2120 2121 2122 2123 2124 2125 2126 2127 2128 2129 2130 2131 2132 2133 2134 2135 2136 2137 2138 2139 2140 2141 2142 2143 2144 2145 2146 2147 2148 2149 2150 2151 2152 2153 2154 2155 2156 2157
Journal of Agricultural and Food Chemistry
(15) Magalhaes, L. M.; Segundo, M. A.; Reis, S.; Lima, J. L. F. C. Methodological aspects about in vitro evaluation of antioxidant properties. Anal. Chim. Acta 2008, 613, 1-19. (16) Halliwell, B. How to characterize a biological antioxidant. Free Radical Res. Commun. 1990, 9, l-32. (17) Gutteridge, J. M. C. Lipid peroxidation and antioxidants as biomarkers of tissue Damage. Clin. Chem. 1995, 41, 1819-1828. (18) Halliwell, B.; Whiteman, M. Measuring reactive species and oxidative damage in vivo and in cell culture: how should you do it and what do the results mean? Br. J. Pharmacol. 2004, 142, 231-255. (19) Finley, J. W.; Kong, A. N.; Hintze, K. J.; Jeffery, E. H.; Ji, L. L.; Lei, X. G. Antioxidants in foods: State of the science important to the food industry. J. Agric. Food Chem. 2011, 59, 6837-6846. (20) Madhavi, D. L.; Deshpande, S. S.; Salunkhe, D. K. Introduction. In Food Antioxidants: Technological, Toxicological, and Health Perspectives; Madhavi, D. L., Deshpande, S. S., Salunkhe, D. K., Eds.; Marcel Dekker, Inc.: New York, 1996; pp. 1-4. (21) Halliwell, B.; Murcia, M. A.; Chirico, S.; Aruoma, O. I. Free radicals and antioxidants in food and in vivo: what they do and how they work. Crit. Rev. Food Sci. Nutr. 1995, 35, 7-20. (22) Bartosz, G. Non-enzymatic antioxidant capacity assays: Limitations of use in biomedicine. Free Radical Res. 2010, 44, 711-720. (23) Shahidi, F.; Ambigaipalan, P. Phenolics and polyphenolics in foods, beverages and spices: Antioxidant activity and health effects – A review, J. Funct. Foods 2015, 18, 820-897. (24) Shahidi, F.; Zhong, Y. Measurement of antioxidant activity, J. Funct. Foods 2015, 18, 757-781. (25) Amorati, R.; Valgimigli, L. Advantages and limitations of common testing methods for antioxidants, Free Radical Res. 2015, 49, 633-649. (26) Wayner, D. D. M.; Burton, G. W.; Ingold, K. U.; Barclay, L. R.; Locke, S. J. The relative contributions of vitamin E, urate, ascorbate and proteins to the total peroxyl radical-trapping antioxidant activity of human blood plasma. Biochim. Biophys. Acta. 1987, 924, 408-419. (27) Bors, W.; Heller, W.; Michel, C.; Saran, M. Flavonoids as antioxidants: Determination of radical scavenging efficiencies. Methods Enzymol. 1990, 186, 343355. (28) Tubaro, E.; Ghiselli, A.; Rapuzzi, E.; Maiorino, M.; Ursini, E. Analysis of plasma antioxidant capacity by competition kinetics. Free Radical Biol. Med. 1998, 24, 1228-1234. (29) Cao, G.; Alessioo, H. M.; Cutler, R. G. Oxygen-radical absorbance capacity assay for antioxidants. Free Radical Biol. Med. 1993, 14, 303-311. (30) Ou, B.; Hampsch-Woodill, M.; Prior, R. L. Development and validation of an improved oxygen radical absorbance capacity assay using fluorescein as the fluorescent probe. J. Agric. Food Chem. 2001, 49, 4619-4626. (31) Winston, G. W.; Regoli, F.; Dugas, A. J. J.; Fong, J. H.; Blanchard, K. A. A rapid gas chromatographic assay for determining oxyradical scavenging capacity of antioxidants and biological fluids. Free Radical Biol. Med. 1998, 24, 480-493. (32) Regoli, E. Total oxyradical scavenging capacity (TOSC) in polluted and translocated mussels: A predictive biomarker of oxidative stress. Aquat. Toxicol. 2000, 50, 351-361. 83 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
2158 2159 2160 2161 2162 2163 2164 2165 2166 2167 2168 2169 2170 2171 2172 2173 2174 2175 2176 2177 2178 2179 2180 2181 2182 2183 2184 2185 2186 2187 2188 2189 2190 2191 2192 2193 2194 2195 2196 2197 2198 2199 2200 2201 2202 2203 2204 2205
Page 84 of 105
(33) Bartosz, G. Total antioxidant capacity. In Advances in Clinical Chemistry, 1st Edition; Spiegel, H. E., Ed.; Academic Press: New York, 2003; Vol. 37, pp. 219292. (34) Folin, O.; Ciocalteu, V. On tyrosine and tryptophane determinations in proteins. J. Biol. Chem. 1927, 73, 627-650. (35) Singleton, V. L.; Orthofer, R.; Lamuela-Raventos, R. M. Analysis of total phenols and other oxidation substrates and antioxidants by means of folin-ciocalteu reagent. Methods Enzymol. 1999, 299, 152-178. (36) Pellegrini, N.; Re, R.; Yang, M.; Rice-Evans, C. Screening of dietary carotenoids and carotenoid-rich fruit extracts for antioxidant activities applying 2,2'azinobis (3- ethylenebenzothiazoline-6-sulfonic acid) radical cation decolorization assay. Methods Enzymol. 1999, 299, 379-389. (37) Re, R.; Pellegrini, N.; Proteggente, A.; Pannala, A.; Yang, M.; Rice-Evans, C. Antioxidant activity applying an improved ABTS radical cation decolorization assay. Free Radical Biol. Med. 1999, 26, 1231-1237. (38) Blois, M. S. Antioxidant determinations by the use of a stable free radical. Nature 1958, 181, 1199-1200. (39) Sanchez-Moreno, C.; Larrauri, J. A.; Saura-Calixto, F. A. A procedure to measure the antiradical efficiency of polyphenols. J. Sci. Food Agric. 1998, 76, 270276. (40) Apak, R.; Güçlü, K.; Özyürek, M.; Karademir, S. E. A novel total antioxidant capacity index for dietary polyphenols, vitamins C and E, using their cupric ion reducing capability in the presence of neocuproine: CUPRAC method. J. Agric. Food Chem. 2004, 52, 7970-7981. (41) Apak, R.; Güçlü, K.; Özyürek, M.; Karademir, S. E.; Altun, M. Total antioxidant capacity assay of human serum using copper(II)-neocuproine as chromogenic oxidant: the CUPRAC method. Free Radical Res. 2005, 39, 949-961. (42) Benzie, I. F. F.; Strain, J. J. The ferric reducing ability of plasma (FRAP) as a measure of ‘antioxidant power’: The FRAP assay. Anal. Biochem. 1996, 239, 70-76. (43) Benzie, I. F.; Strain, J. J. Ferric reducing/antioxidant power assay: Direct measure of total antioxidant activity of biological fluids and modified version for simultaneous measurement of total antioxidant power and ascorbic acid concentration. Methods Enzymol. 1999, 299, 15-27. (44) Benzie, I. F. F.; Szeto, Y. T. Total antioxidant capacity of teas by the ferric reducing/antioxidant power assay. J. Agric. Food Chem. 1999, 47, 633-636. (45) Pulido, R.; Bravo, L.; Saura-Calixto F. Antioxidant activity of dietary polyphenols as determined by a modified ferric reducing/antioxidant power assay. J. Agric. Food. Chem. 2000, 48, 3396-3402. (46) Oyaizu, M. Studies on products of browning reaction: antioxidative activity of products of browning reaction prepared from glucosamine. Jpn. J. Nutr. 1986, 44, 307-315. (47) Berker, K. I.; Güçlü, K.; Tor, I.; Demirata Öztürk, B.; Apak, R. Total antioxidant capacity assay using optimized ferricyanide / Prussian blue method. Food Anal. Methods 2010, 3, 154-168. (48) Berker, K. I.; Güçlü, K.; Tor, I.; Apak, R. Comparative evaluation of Fe(III) reducing power-based antioxidant capacity assays in the presence of phenanthroline, batho-phenanthroline, tripyridyltriazine (FRAP), and ferricyanide reagents. Talanta 2007, 72, 1157-1165.
84 ACS Paragon Plus Environment
Page 85 of 105
2206 2207 2208 2209 2210 2211 2212 2213 2214 2215 2216 2217 2218 2219 2220 2221 2222 2223 2224 2225 2226 2227 2228 2229 2230 2231 2232 2233 2234 2235 2236 2237 2238 2239 2240 2241 2242 2243 2244 2245 2246 2247 2248 2249 2250 2251 2252 2253 2254 2255
Journal of Agricultural and Food Chemistry
(49) Berker, K. I.; Güçlü, K.; Demirata, B.; Apak, R. A novel antioxidant assay of ferric reducing capacity measurement using ferrozine as the colour forming complexation reagent. Anal. Methods 2010, 2, 1770-1778. (50) Işık, E.; Şahin, S.; Demir, C. Development of a new chromium reducing antioxidant capacity (CHROMAC) assay for plants and fruits. Talanta 2013, 111, 119-124. (51) Özyurt, D.; Demirata, B.; Apak, R. Determination of total antioxidant capacity by a new spectrophotometric method based on Ce(IV) reducing capacity measurement. Talanta 2007, 71, 1155-1165. (52) Özyurt, D.; Demirata Öztürk, B.; Apak, R. Modified cerium(IV)-based antioxidant capacity (CERAC) assay with selectivity over citric acid and simple sugars. J. Food Compos. Anal. 2010, 23, 282-288. (53) Özyurt, D.; Demirata, B.; Apak, R. Determination of total antioxidant capacity by a new spectrofluorometric method based on Ce(IV) reduction: Ce(III) fluorescence probe for CERAC assay. J. Fluoresc. 2011, 21, 2069-2076. (54) Scampicchio, M.; Wang, J.; Blasco, A. J.; Arribas, A. S.; Mannino, S.; Escarpa, A. Nanoparticle-based assays of antioxidant activity. Anal. Chem. 2006, 78, 2060-2063. (55) Özyürek, M.; Güngör, N.; Baki, S.; Güçlü, K.; Apak, R. Development of a silver nanoparticle-based method for the antioxidant capacity measurement of polyphenols. Anal. Chem. 2012, 84, 8052-8059. (56) Apak, R.; Güçlü, K.; Demirata, B.; Özyürek, M.; Çelik, S. E.; Bektaşoğlu, B.; Berker, K. I.; Özyurt, D. Comparative evaluation of total antioxidant capacity assays applied to phenolic compounds, and the CUPRAC assay. Molecules 2007, 12, 14961547. (57) Wolfenden, B. S.; Wilson, R. L. Radical-cations as reference chromogens in kinetic studies of one-electron transfer reactions: Pulse radiolysis studies of 2,2'azinobis-(3-ethylbenzthiazoline-6- sulphonate). J. Chem. Soc. Perkin Trans. 1982, 2, 805-812. (58) Niki, E. Antioxidant capacity: Which capacity and how to assess it? J. Berry Res. 2011, 1, 169-176. (59) Ou, B.; Huang, D.; Hampsch-Woodill, M.; Flanagan, J. A.; Deemer, E. K. Analysis of antioxidant activities of common vegetables employing oxygen radical absorbance capacity (ORAC) and ferric reducing antioxidant power (FRAP) assays: a comparative study. J. Agric. Food Chem. 2002, 50, 3122-3128. (60) Cao, G.; Prior, R. L. Comparison of different analytical methods for assessing total antioxidant capacity of human serum. Clin. Chem. 1998, 44, 1309-1315. (61) Prieto, P.; Pineda, M.; Aguilar, M. Spectrophotometric quantitation of antioxidant capacity through the formation of a phosphomolybdenum complex: specific application to the determination of vitamin E. Anal. Biochem. 1999, 269, 337-341. (62) Slinkard, K.; Singleton, V. L. Total phenol analysis: Automation and comparison with manual methods Am. J. Enol. Vitic. 1977, 28, 49-55. (63) Berker, K. I.; Olgun, F. A. Ö.; Özyurt, D.; Demirata, B.; Apak, R. Modified Folin−Ciocalteu antioxidant capacity assay for measuring lipophilic antioxidants J. Agric. Food Chem. 2013, 61, 4783-4791. (64) Demirci Çekiç, S.; Sözgen Başkan, K.; Tütem, E.; Apak, R. Modified cupric reducing antioxidant capacity (CUPRAC) assay for measuring the antioxidant capacities of thiol−containing proteins in admixture with polyphenols. Talanta 2009, 79, 344-351. 85 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
2256 2257 2258 2259 2260 2261 2262 2263 2264 2265 2266 2267 2268 2269 2270 2271 2272 2273 2274 2275 2276 2277 2278 2279 2280 2281 2282 2283 2284 2285 2286 2287 2288 2289 2290 2291 2292 2293 2294 2295 2296 2297 2298 2299 2300 2301 2302 2303 2304
Page 86 of 105
(65) Bean, H.; Radu, F.; De, E.; Schuler, C.; Leggett, R. E.; Levin, R. M. Comparative evaluation of antioxidant reactivity within obstructed and control rabbit urinary bladder tissue using FRAP and CUPRAC assays. Mol. Cell. Biochem. 2009, 323, 139-142. (66) Sözgen Başkan, K.; Tütem, E.; Özer, N.; Apak, R. Spectrophotometric and chromatographic assessment of contributions of carotenoids and chlorophylls to the total antioxidant capacities of plant foods. J. Agric. Food Chem. 2013, 61, 1137111381. (67) Karoui, H.; Hogg, N.; Frejaville, C.; Tordo, P.; Kalyanaraman, B. Characterization of sulfur-centered radical intermediates formed during the oxidation of thiols and sulfite by peroxynitrite. J. Biol. Chem. 1996, 271, 6000-6009. (68) Antolovich, M.; Prenzler, P. D.; Patsalides, E.; McDonald, S.; Robards, K. Methods for testing antioxidant activity. Analyst 2002, 127, 183-198. (69) Bors, W.; Heller, W.; Michel, C.; Saran, M. Flavonoids as antioxidants: Determination of radical scavenging efficiencies. Methods Enzymol. 1990, 186, 343355. (70) Rice-Evans, C. A.; Miller, N. J.; Paganga, G. Structure‒antioxidant activity relationships of flavonoids and phenolic acids. Free Radical Biol. Med. 1996, 20, 933-956. (71) Tütem, E.; Apak, R.; Baykut, F. Spectrophotometric determination of trace amounts of copper(I) and reducing agents with neocuproine in the presence of copper(II). Analyst 1991, 116, 89-94. (72) Lappin, A. G.; Youngblood, M. P.; Margerum, D. W. Electron-transfer reactions of copper(I) and copper(III) complexes. Inorg. Chem. 1980, 19, 407-413. (73) Bener, M.; Özyürek, M.; Güçlü, K.; Apak, R. Development of a low-cost optical sensor for cupric reducing antioxidant capacity measurement of food extracts. Anal. Chem. 2010, 82, 4252-4258. (74) Apak, R.; Güçlü, K.; Özyürek, M.; Çelik, S. E. Mechanism of antioxidant capacity assays and the CUPRAC (cupric ion reducing antioxidant capacity) assay. Microchim. Acta 2008, 160, 413-419. (75) Campos, C.; Guzmán, R.; López-Fernandez, E.; Casado, A. Evaluation of the copper(II) reduction assay using bathocuproinedisulfonic acid disodium salt for the total antioxidant capacity assessment: The CUPRAC–BCS assay. Anal. Biochem. 2009, 392, 37-44. (76) Smith, P. K.; Krohn, R. I.; Hermanson, G. T.; Mallia, A. K.; Gartner, F. H.; Provenzano, M. D.; Fujimoto, E. K.; Goeke, N. M.; Olson, B. J.; Klenk, D. C. Measurement of protein using bicinchoninic acid. Anal. Biochem. 1985, 150, 76-85. (77) Litwinienko, G.; Ingold, K. U. Solvent effects on the rates and mechanisms of reaction of phenols with free radicals. Acc. Chem. Res. 2007, 40, 222-230. (78) Çelik, S. E.; Özyürek, M., Güçlü, K.; Apak, R. Differences in responsivity of original cupric reducing antioxidant capacity and cupric-bathocuproine sulfonate assays to antioxidant compounds. Anal. Biochem. 2012, 423, 36-38. (79) Zhou, F.; Millhauser, G. L. The rich electrochemistry and redox reactions of the copper sites in the cellular prion protein. Coord. Chem. Rev. 2012, 256, 22852296. (80) Marques, S. S.; Magalhaes, L. M.; Tóth, I. V.; Segundo, M. A. Insights on antioxidant assays for biological samples based on the reduction of copper complexes‒the importance of analytical conditions. Int. J. Mol. Sci. 2014, 15, 11387-11402.
86 ACS Paragon Plus Environment
Page 87 of 105
2305 2306 2307 2308 2309 2310 2311 2312 2313 2314 2315 2316 2317 2318 2319 2320 2321 2322 2323 2324 2325 2326 2327 2328 2329 2330 2331 2332 2333 2334 2335 2336 2337 2338 2339 2340 2341 2342 2343 2344 2345 2346 2347 2348 2349 2350 2351 2352 2353
Journal of Agricultural and Food Chemistry
(81) Huang, X.; Cuajungco, M. P.; Atwood, C. S.; Hartshorn, M. A.; Tyndall, J. D. A.; Hanson, G. R.; Stokes, K. C.; Leopold, M.; Multhaup, G.; Goldstein, L. E.; Scarpa, R. C.; Saunders, A. J.; Lim, J.; Moir, R. D.; Glabe, C.; Bowden, E. F.; Masters, C. L.; Fairlie, D. P.; Tanzi, R. E.; Bush, A. I. Cu(II) potentiation of Alzheimer Aβ neurotoxicity: Correlation with cell-free hydrogen peroxide production and metal reduction. J. Biol. Chem. 1999, 274, 37111-37116. (82) Apak, R.; Güçlü, K.; Özyürek, M.; Bektaşoğlu, B.; Bener, M. CUPRAC (cupric ion reducing antioxidant capacity) assay for antioxidants in human serum, and for hydroxyl radical scavengers. In Advanced Protocols in Oxidative Stress II (Series: Methods in Molecular Biology), Armstrong, D.; Ed.; Humana PressSpringer: Gainesville, FL, USA 2010, Vol. 594, pp. 215-239. (83) Özyürek, M.; Güçlü, K.; Bektaşoğlu, B.; Apak, R. Spectrophotometric determination of ascorbic acid by the modified CUPRAC method with extractive separation of flavonoids-La(III) complexes. Anal. Chim. Acta 2007, 588, 88-95. (84) Özyürek, M.; Bektaşoğlu, B.; Güçlü, K.; Güngör, N.; Apak, R. Simultaneous total antioxidant capacity assay of lipophilic and hydrophilic antioxidants in the same acetone-water solution containing 2% methyl-beta-cyclodextrin using the Cupric Reducing Antioxidant Capacity (CUPRAC) method. Anal. Chim. Acta 2008, 630, 28-39. (85) Bektaşoğlu, B.; Çelik, S. E.; Özyürek, M.; Güçlü, K.; Apak, R. Novel hydroxyl radical scavenging antioxidant activity assay for water-soluble antioxidants using a modified CUPRAC method. Biochem. & Biophys. Res. Commun. 2006, 345, 11941200. (86) Özyürek, M.; Bektaşoğlu, B.; Güçlü, K.; Apak, R. Measurement of xanthine oxidase inhibition activity of phenolics and flavonoids with a modified cupric reducing antioxidant capacity (CUPRAC) method. Anal. Chim. Acta 2009, 636, 4250. (87) Özyürek, M.; Bektaşoğlu, B.; Güçlü, K.; Güngör, N.; Apak, R. A novel hydrogen peroxide scavenging assay of phenolics and flavonoids using cupric reducing antioxidant capacity (CUPRAC) methodology J. Food Compos. Anal. 2010, 23, 689-698. (88) Çelik, S. E.; Özyürek, M.; Güçlü, K.; Apak, R. Determination of antioxidants by a novel on-line HPLC-cupric reducing antioxidant capacity (CUPRAC) assay with post-column detection. Anal. Chim. Acta 2010, 674, 79-88. (89) Bekdeşer, B.; Özyürek, M.; Güçlü, K.; Apak, R. tert-Butylhydroquinone as a spectroscopic probe for the superoxide radical scavenging activity assay of biological samples. Anal. Chem. 2011, 83, 5652-5660. (90) Christodouleas, D. C.; Fotakis, C.; Papadopoulos, K.; Calokerinos, A. C. Evaluation of total reducing power of edible oils. Talanta 2014, 130, 233-240. (91) Walling, C. Kinetics of the decomposition of H2O2 catalyzed by ferric EDTA complexes. Proc. Natl. Acad Sci. USA 1975, 72, 140-142. (92) Ribeiro, J. P. N.; Magalhaes, L. M.; Reis, S.; Lima, J. L. F. C.; Segundo, M. A. High-throughput total cupric ion reducing antioxidant capacity of biological samples determined using flow injection analysis and microplate-based methods. Anal. Sci. 2011, 27, 483-488. (93) Gorinstein, S.; Leontowicz, M.; Leontowicz, H.; Najman, K.; Namiesnik, J.; Park, Y. S.; Jung, S. T.; Kang, S. G.; Trakhtenberg, S. Supplementation of garlic lowers lipids and increases antioxidant capacity in plasma of rats. Nutr. Res. 2006, 26, 362368.
87 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
2354 2355 2356 2357 2358 2359 2360 2361 2362 2363 2364 2365 2366 2367 2368 2369 2370 2371 2372 2373 2374 2375 2376 2377 2378 2379 2380 2381 2382 2383 2384 2385 2386 2387 2388 2389 2390 2391 2392 2393 2394 2395 2396 2397 2398 2399 2400 2401 2402 2403
Page 88 of 105
(94) Yen, G. C.; Duh, P. D. Antioxidative properties of methanolic extracts from peanut hulls. J. Am. Oil Chem. Soc. 1993, 70, 383-386. (95) Berker, K. I.; Demirata, B.; Apak, R. Determination of total antioxidant capacity of lipophilic and hydrophilic antioxidants in the same solution by using ferric– ferricyanide assay. Food Anal. Methods 2012, 5, 1150-1158. (96) Adcock, J. L.; Francis, P. S.; Barnett, N. W. Acidic potassium permanganate as a chemiluminescence reagent-A Review. Anal. Chim. Acta 2007, 601, 36-67. (97) Francis, P. S.; Costin, J. W.; Conlan, X. A.; Bellomarino, S. A.; Barnett, J. A.; Barnett, N. W. A rapid antioxidant assay based on acidic potassium permanganate chemiluminescence. Food Chem. 2010, 122, 926-929. (98) Cacig, S. I.; Szabo, M. I.; Lupea, A. X. D. Spectrophotometric method for the study of the antioxidant activity applied on Ziziphus jujoba and Hydrangea paniculata aqueous extracts. Proc. Nat. Sci, Matica Srpska Novi Sad. 2006, 110, 8793. (99) Popovic, B. M.; Štajner, D.; Slavko, K.; Sandra, B. Antioxidant capacity of cornelian cherry (Cornus mas L.) – Comparison between permanganate reducing antioxidant capacity and other antioxidant methods. Food Chem. 2012, 134, 734741. (100) Blasco, A. J.; Crevillen, A. G.; Gonzalez, M. C.; Escarpa, A. Direct electrochemical sensing and detection of natural antioxidants and antioxidant capacity in vitro systems. Electroanalysis 2007, 19, 2275-2286. (101) Hotta, H.; Sakamoto, H.; Nagano, S.; Osakai, T.; Tsujino, Y. Unusually large numbers of electrons for the oxidation of polyphenolic antioxidants. Biochim. Biophys. Acta 2001, 1526, 159-167. (102) Chevion, M.; Berenshtein, E.; Stadtman, E. R. Human studies related to protein oxidation: protein carbonyl content as a marker of damage. Free Radical Res. 2000, 33, 99-108. (103) Chevion, S.; Berry, E. M.; Kitrossky, N.; Kohen, R. Evaluation of plasma low molecular weight antioxidant capacity by cyclic voltammetry. Free Radical Biol. Med. 1997, 22, 411-421. (104) Kohen, R.; Vellaichamy, E.; Hrbac, J.; Gati, I.; Tirosh, O. Quantification of the overall reactive oxygen species scavenging capacity of biological fluids and tissues. Free Radical Biol. Med. 2000, 28, 871-879. (105) Piljac-Žegarac, J.; Valek, L.; Stipčević, T. T.; Martinez, S. Electrochemical determination of antioxidant capacity of fruit tea infusions. Food Chem. 2010, 121, 820-825. (106) Novak, I.; Šeruga, M.; Komorsky-Lovrić, Š. Electrochemical characterization of epigallocatechin gallate using square-wave voltammetry. Electroanal. 2009, 21, 1019-1025. (107) Magarelli, G.; Da Silva, J. G.; De Sousa Filho, I. A.; Dourado Lopes, I. S.; Rodrigues SouzaDe, J.; Hoffmann, L. V.; De Castro, C. S. P. Development and validation of a voltammetric method for determination of total phenolic acids in cotton cultivars. Microchem. J. 2013, 109, 23-28. (108) Korotkova, E. I.; Karbainov, Y. A.; Shevchuk, A.V. Study of antioxidant properties by voltammetry. J. Electroanal. Chem. 2002, 518, 56-60. (109) Kilmartin, P. A.; Zou, H.; Waterhouse, A. L. A cyclic voltammetry method suitable for characterizing antioxidant properties of wine and wine phenolics. J. Agric. Food Chem. 2001, 49, 1957-1965. (110) Campanella, L.; Bonanni, A.; Bellantoni, D.; Favero, G.; Tomassetti, M. Comparison of fluorimetric, voltammetric and biosensor methods for the 88 ACS Paragon Plus Environment
Page 89 of 105
2404 2405 2406 2407 2408 2409 2410 2411 2412 2413 2414 2415 2416 2417 2418 2419 2420 2421 2422 2423 2424 2425 2426 2427 2428 2429 2430 2431 2432 2433 2434 2435 2436 2437 2438 2439 2440 2441 2442 2443 2444 2445 2446 2447 2448 2449 2450 2451
Journal of Agricultural and Food Chemistry
determination of total antioxidant capacity of drug products containing acetylsalicylic acid. J. Pharm. Biomed. Anal. 2004, 36, 91-99. (111) Milardović, S.; Ivekovic, D.; Grabarić, B. S. A novel amperometric method for antioxidant activity determination using DPPH free radical. Bioelectrochem. 2006, 68, 175-180. (112) Gulaboski, R.; Mirčeski, V.; Mitrev, S. Development of a rapid and simple voltammetric method to determine total antioxidative capacity of edible oils. Food Chem. 2013, 138, 116-121. (113) Ferreira, R. Q.; Avaca, L. A. Electrochemical determination of the antioxidant capacity: The ceric reducing/antioxidant capacity (CRAC) assay. Electroanal. 2008, 20, 1323-1329. (114) Tufan, A. N.; Baki, S.; Güçlü, K.; Özyürek, M.; Apak, R. A novel differential pulse voltammetric (dpv) method for measuring the antioxidant capacity of polyphenols-reducing cupric neocuproine complex. J. Agric. Food Chem. 2014, 62, 7111-7117. (115) Prieto-Simon, B.; Cortina, M.; Campas, M.; Calas-Blanchard, C. Electrochemical biosensors as a tool for antioxidant capacity assessment. Sensors Actuat. B 2008, 129, 459-466. (116) Tammeveski, K.; Tenno, T. T.; Mashirin, A. A.; Hillhouse, E. W.; Manning, P.; McNeil, C. J. Superoxide electrode based on covalently immobilized cyt c: modelling studies. Free Radical Biol. Med. 1998, 25, 973-978. (117) Ignatov, S.; Shishniashvili, D.; Ge, B.; Scheller, F. W.; Lisdat, F. Amperometric biosensor based on a functionalized gold electrode for the detection of antioxidants. Biosens. Bioelectron. 2002, 17, 191-199. (118) Tian, Y.; Mao, L.; Okajima, T.; Ohsaka, T. A carbon fiber microelectrodebased third-generation biosensor for superoxide anion. Biosens. Bioelectron. 2005, 21, 557-564. (119) Lvovich, V.; Scheeline, A. Amperometric sensors for simultaneous superoxide and hydrogen peroxide detection. Anal. Chem. 1997, 69, 454-462. (120) Liu, J.; Su, B.; Lagger, G.; Tacchini, P.; Girault, H. H. Antioxidant redox sensors based on DNA modified carbon screen-printed electrodes. Anal. Chem. 2006, 78, 6879-6886. (121) Gorjanovic, S. Z.; Novakovic, M. M.; Potkonjak, N. I.; Sužnjevic, D. Z. Antioxidant activity of wines determined by a polarographic assay based on hydrogen peroxide scavenge. J. Agric. Food Chem. 2010, 58, 4626-4631. (122) Gorjanović, S.; Komes, D.; Pastor, F. T.; Belscak-Cvitanović, A.; Pezo, L.; Hečimović, I.; Sužnjevic, D. Antioxidant capacity of teas and herbal infusions: polarographic assessment. J. Agric. Food Chem. 2012, 60, 9573-9580. (123) Sužnjević, D.; Petrović, M.; Pastor, F. T.; Veljović, M.; Zlatanović, S.; Antić, M.; Gorjanović, S. Reduction of Hg2+ by individual phenolics and complex samples and its application in polarographic antioxidant assay. J. Electrochem. Soc. 2015, 162, H428-H433. (124) Jensen, T. R.; Duval Malinsky, M.; Haynes, C. L.; Van Duyne, R. P. Nanosphere lithography: Tunable localized surface plasmon resonance spectra of silver nanoparticles. J. Phys. Chem. B 2000, 104, 10549-10556. (125) Schafer, F. Q.; Buettner, G. R. Redox environment of the cell as viewed through the redox state of the glutathione disulfide/glutathione couple. Free Radical Biol. Med. 2001, 30, 1191-1212.
89 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
2452 2453 2454 2455 2456 2457 2458 2459 2460 2461 2462 2463 2464 2465 2466 2467 2468 2469 2470 2471 2472 2473 2474 2475 2476 2477 2478 2479 2480 2481 2482 2483 2484 2485 2486 2487 2488 2489 2490 2491 2492 2493 2494 2495 2496 2497 2498 2499 2500
Page 90 of 105
(126) Güçlü, K.; Özyürek, M.; Güngör, N.; Baki, S.; Apak, R. Selective optical sensing of biothiols with Ellman’s reagent: 5,5’-dithio-bis(2-nitrobenzoic acid)-modified gold nanoparticles. Anal. Chim. Acta 2013, 794, 90-98. (127) Bain, C. D.; Biebuyck, H. A.; Whitesides, G. M. Comparison of self-assembled monolayers on gold: Coadsorption of thiols and disulfides. Langmuir 1989, 5, 723727. (128) Chen, S. J.; Chang H. T. Nile red-adsorbed gold nanoparticles for selective determination of thiols based on energy transfer and aggregation. Anal. Chem. 2004, 76, 3727-3734. (129) Li, H.; Ma, X.; Dong, J.; Qian, W. Development of methodology based on the formation process of gold nanoshells for detecting hydrogen peroxide scavenging activity. Anal. Chem. 2009, 81, 8916-8922. (130) Ma, X.; Li, H.; Dong, J.; Qian, W. Determination of hydrogen peroxide scavenging activity of phenolic acids by employing gold nanoshells precursor composites as nanoprobes. Food Chem. 2011, 126, 698-704. (131) Apak, R.; Gorinstein, S.; Böhm, V.; Schaich, K. M.; Özyürek, M.; Güçlü, K. Methods of measurement and evaluation of natural antioxidant capacity/activity. Pure Appl. Chem. 2013, 85, 957-998. (132) Tian, X.; Schaich, K. M. Effects of molecular structure on kinetics and dynamics of the trolox equivalent antioxidant capacity assay with ABTS+•. J. Agric. Food Chem. 2013, 61, 5511-5519. (133) Xie, J.; Schaich, K. M. Re-evaluation of the 2,2-diphenyl-1-picrylhydrazyl free radical (DPPH) asay for antioxidant activity. J. Agric. Food Chem. 2014, 62, 42514260. (134) Schaich, K. M.; Tian, X.; Xie, J. Hurdles and pitfalls in measuring antioxidant efficacy: A critical evaluation of ABTS, DDPH, and ORAC assays, J. Funct. Foods 2015, 14, 111-125. (135) Papariello, G. J.; Janish, M. A. M. Diphenylpicrylhydrazyl as an organic analytical reagent in the spectrophotometric analysis of phenols. Anal. Chem. 1966, 38, 211-214. (136) Brand-Williams, W.; Cuvelier, M. E.; Berset, C. Use of a free radical method to evaluate antioxidant activity. LWT-Food Sci. Technol. 1995, 28, 25-30. (137) Gomez-Alonso, S.; Fregapane, G.; Salvador, M. D.; Gordon, M. H. Changes in phenolic composition and antioxidant activity of virgin olive oil during frying. J. Agric. Food Chem. 2003, 51, 667-672. (138) Lebeau, J.; Furman, C.; Bernier, J. L.; Duriez, P.; Teissier, E.; Cotelle, N. Antioxidant properties of di-tert-butylhydroxylated flavonoids. Free Radical Biol. Med. 2000, 29, 900-912. (139) McGowan, J. C.; Powell, T.; Raw, R. The rates of reaction of α,α-diphenyl-βpicrylhydrazyl with certain amines and phenols. J. Chem. Soc. 1959, 1959, 31033110. (140) Hogg, J. S.; Lohmann, D. H.; Russell, K. S. The kinetics of reaction of 2,2diphenyl-1-picrylhydrazyl with phenols. Can. J. Chem. 1961, 39, 1588-1594. (141) Beltran-Orozco, M. C.; Oliva-Coba, T. G.; Gallardo-Velazquez, T. ; OsorioRevilla, G. Ascorbic acid, phenolic content, and antioxidant capacity of red, cherry, yellow and white types of pitaya cactus fruit (Stenocereus stellatus riccobono). Agrociencia 2009, 43, 153-161. (142) Corral-Aguayo, R. D.; Yahia, E. M.; Carrillo-Lopez, A.; Gonzalez-Aguilar, G. Correlation between some nutritional components and the total antioxidant capacity
90 ACS Paragon Plus Environment
Page 91 of 105
2501 2502 2503 2504 2505 2506 2507 2508 2509 2510 2511 2512 2513 2514 2515 2516 2517 2518 2519 2520 2521 2522 2523 2524 2525 2526 2527 2528 2529 2530 2531 2532 2533 2534 2535 2536 2537 2538 2539 2540 2541 2542 2543 2544 2545 2546 2547 2548
Journal of Agricultural and Food Chemistry
measured with six different assays in eight horticultural crops. J. Agric. Food Chem. 2008, 56, 10498-10504. (143) Fogliano, V.; Verde, V.; Randazzo, G. Method for measuring antioxidant activity and its application to monitoring the antioxidant capacity of wines. J. Agric. Food Chem. 1999, 47, 1035-1040. (144) Ak, T.; Gülçin, I. Antioxidant and radical scavenging properties of curcumin. Chem.-Biol. Interact. 2008, 174, 27-37. (145) Mehdi, M. M.; Rizvi, S. I. N,N-Dimethyl-p-phenylenediamine dihydrochloride-based method for the measurement of plasma oxidative capacity during human aging. Anal. Biochem. 2013, 436, 165-167. (146) Schöttker, B.; Saum, K.-U.; Jansen, E. H. J. M.; Boffetta, P.; Trichopoulou, A.; Holleczek, B.; Dieffenbach, A. K.; Brenner, H. Oxidative stress markers and allcause mortality at older age: A population-based cohort study, J. Gerontol. A Biol. Sci. Med. Sci. 2015, 70, 518-524. (147) Demirci Çekiç, S.; Avan, A. N.; Uzunboy, S.; Apak, R. A colourimetric sensor for the simultaneous determination of oxidative status and antioxidant activity on the same membrane: N,N-Dimethyl-p-phenylene diamine (DMPD) on Nafion, Anal. Chim. Acta 2015, 865, 60-70. (148) Schwarz, K.; Bertelsen, G.; Nissen, L. R.; Gardner, P. T.; Heinonen, M. I.; Hopia, A.; Huynh-Ba, T.; Lambelet, P.; McPhail, D.; Skibsted, L. H.; Tijburg, L., Investigation of plant extracts for the protection of processed foods against lipid oxidation. Comparison of antioxidant assays based on radical scavenging, lipid oxidation and analysis of the principal antioxidant compounds, Eur. Food Res. Technol. 2001, 212, 319-328. (149) Nagaoka, S.-I.; Nagai, K.; Fujii, Y.; Ouchi, A.; Mukai, K. Development of a new free radical absorption capacity assay method for antioxidants: Aroxyl radical absorption capacity (ARAC), J. Agric. Food Chem. 2013, 61, 10054-10062. (150) Wiseman, H.; Halliwell, B. Damage to DNA by reactive oxygen and nitrogen species: Role in inflammatory disease and progression to cancer. Biochem. J. 1996, 313, 17-29. (151) Bourdon, E.; Blache, D. The importance of proteins in defense against oxidation. Antioxid. Redox Signal. 2001, 3, 293-311. (152) Beers, R. F.; Sizer, I. W. A spectrophotometric method for measuring the breakdown of hydrogen peroxide by catalase. J. Biol. Chem. 1952, 195, 133-140. (153) Aruoma, O. I.; Murcia, A.; Butler, J.; Halliwell, B. Evaluation of the antioxidant and prooxidant actions of gallic acid and its derivatives. J. Agric. Food Chem. 1993, 41, 1880-1885. (154) Ching, T. L.; de Jong, J.; Bast, A. A Method for screening hypochlorous acid scavengers by inhibition of the oxidation of 5-thio-2-nitrobenzoic acid: Application anti-asthmatic drugs. Anal. Biochem. 1994, 218, 377-381. (155) Bartosz, G. Use of spectroscopic probes for detection of reactive oxygen species. Clin. Chim. Acta. 2006, 368, 53-76. (156) Wolfe, K. L.; Liu, R. H. Cellular antioxidant activity (CAA) assay for assessing antioxidants, foods, and dietary supplements. J. Agric. Food Chem. 2007, 55, 8896-8907. (157) López-Alarcóna, C.; Denicola, A. Evaluating the antioxidant capacity of natural products: A review on chemical and cellular-based assays. Anal. Chim. Acta. 2013, 763, 1-10.
91 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
2549 2550 2551 2552 2553 2554 2555 2556 2557 2558 2559 2560 2561 2562 2563 2564 2565 2566 2567 2568 2569 2570 2571 2572 2573 2574 2575 2576 2577 2578 2579 2580 2581 2582 2583 2584 2585 2586 2587 2588 2589 2590 2591 2592 2593 2594 2595 2596 2597
Page 92 of 105
(158) Wolfe, K. L.; Liu, R. H. Structure-activity relationship of flavonoids in the cellular antioxidant activity (CAA) assay. J. Agric. Food Chem. 2008, 56, 84048411. (159) Wardman, P. Fluorescent and luminescent probes for measurement of oxidative and nitrosative species in cells and tissues: Progress, pitfalls, and prospects. Free Radical Biol. Med. 2007, 43, 995-1022. (160) Aruoma, O. I.; Halliwell, B.; Hoey, B. M.; Butler, J. The antioxidant action of N-acetylcysteine: its reaction with hydrogen peroxide, hydroxyl radical, superoxide, and hypochlorous acid. Free Radical Biol. Med. 1989, 6, 593-597. (161) Barclay, L. R. C. 1992 Syntex Award Lecture. Model biomembranes: quantitative studies of peroxidation, antioxidant action, partitioning, and oxidative stress. Can. J. Chem. 1993, 71, 1-16. (162) Buettner, G. R. The pecking order of free radicals and antioxidants: lipid peroxidation, α-tocopherol, and ascorbate. Arch. Biochem. Biophys. 1993, 300, 535543. (163) Pietta, P.-G. Flavonoids as antioxidants. J. Nat. Prod. 2000, 63, 1035-1042. (164) El-Magoli, S. B.; Karel, M.; Yong, S. Acceleration of lipid oxidation by volatile products of hydroperoxide decomposition. J. Food Biochem. 1979, 3, 111123. (165) Pinchuk, I.; Lichtenberg, D. The mechanism of action of antioxidants against lipoprotein peroxidation, evaluation based on kinetic experiments. Prog. Lipid Res. 2002, 41, 279-314. (166) Yıldogan Beker, B.; Bakır, T.; Sonmezoglu, I.; Imer, F.; Apak, R. Antioxidant protective effect of flavonoids on linoleic acid peroxidation induced by copper(II)/ascorbic acid system. Chem. Phys. Lipids. 2011, 164, 732-739. (167) Marcus, R. A. Transfer reactions in chemistry. Theory and experiment. Pure Appl. Chem. 1997, 69, 13-30. (168) Finotti, E.; Di Majo, D. Influence of solvents on the antioxidant property of flavonoids. Nahrung/Food. 2003, 47, 186-187. (169) Perez-Jimenez, J.; Saura-Calixto, F. Effect of solvent and certain food constituents on different antioxidant capacity assays. Food Res. Int. 2006, 39, 791800. (170) Pedrielli, P.; Pedulli, G. F.; Skibsted, L. H. Antioxidant mechanism of flavonoids. Solvent effect on rate constant for chain-breaking reaction of quercetin and epicatechin in autoxidation of methyl linoleate. J. Agric. Food Chem. 2001, 49, 3034-3040. (171) Pinelo, M.; Manzocco, L.; Nunez, M. J.; Nicoli, M. C. Solvent effect on quercetin antioxidant capacity. Food Chem. 2004, 88, 201-207. (172) Alfassi, Z. B.; Huie, R. E.; Neta, P. Substituent effects on rates of one-electron oxidation of phenols by the radicals ClO2, NO2, and SO3-. J. Phys. Chem. 1986, 90, 4156-4158. (173) Çelik, S. E.; Özyürek, M.; Güçlü, K.; Apak, R. Solvent effects on the antioxidant capacity of lipophilic and hydrophilic antioxidants measured by CUPRAC, ABTS/persulphate and FRAP methods. Talanta. 2010, 81, 1300-1309. (174) Choe, E.; Min, D. B. Mechanisms of antioxidants in the oxidation of foods, Comp. Rev. Food Sci. Food Safety 2009, 8, 345-357. (175) Foti, M.; Ruberto, G. Kinetic solvent effects on phenolic antioxidants determined by spectrophotometric measurements. J. Agric. Food Chem. 2001, 49, 342-348.
92 ACS Paragon Plus Environment
Page 93 of 105
2598 2599 2600 2601 2602 2603 2604 2605 2606 2607 2608 2609 2610 2611 2612 2613 2614 2615 2616 2617 2618 2619 2620 2621 2622 2623 2624 2625 2626 2627 2628 2629 2630 2631 2632 2633 2634 2635 2636 2637 2638 2639 2640 2641 2642 2643 2644 2645
Journal of Agricultural and Food Chemistry
(176) Barclay, L. R. C.; Edwards, C. E.; Vinqvist, M. R. Media effects on antioxidant activities of phenols and catechols. J. Am. Chem. Soc. 1999, 121, 6226-6231. (177) Leopoldini, M.; Marino, T.; Russo, N.; Toscano, M. Antioxidant properties of phenolic compounds: H-atom versus electron transfer mechanism. J. Phys. Chem. A. 2004, 108, 4916-4922. (178) Lucarini, M.; Mugnaini, V.; Pedulli, G. F. Bond dissociation enthalpies of polyphenols: The importance of cooperative effects. J. Org. Chem. 2002, 67, 928931. (179) Barclay, L. R. C.; Vinqvist, M. R.; Mukai, K.; Goto, H.; Hashimoto, Y.; Tokunaga, A.; Uno, H. On the antioxidant mechanism of curcumin: classical methods are needed to determine antioxidant mechanism and activity. Org. Lett. 2000, 2, 2841-2843. (180) De Heer, M. I.; Mulder, P.; Korth, H.; Ingold, K. U.; Lusztyk, J. Hydrogen atom abstraction kinetics from intramolecularly hydrogen bonded ubiquinol-0 and other (poly)methoxy phenols. J. Am. Chem. Soc. 2000, 122, 2355-2360. (181) Wayner, D. D.; Lusztyk, E.; Page, D.; Ingold, K. U.; Mulder, P.; Laarhoven, L. J. J.; Aldrichs, H. S. Effects of solvation on the enthalpies of reaction of tert-butoxyl radicals with phenol and on the calculated O-H bond strength in phenol. J. Am. Chem. Soc. 1995, 117, 8738-8744. (182) Litwinienko, G.; Ingold, K. U. Abnormal effects on hydrogen atom abstractions. 1. The reactions of phenols with 2,2-diphenyl-1-picrylhydrazyl (DPPH) in alcohols. J. Org. Chem. 2003, 68, 3433-3438. (183) Pannala, A. S.; Chan, T. S.; O’Brien, P. J.; Rice-Evans, C. A. Flavonoid B-ring chemistry and antioxidant activity: Fast reaction kinetics. Biochem. Biophys. Res. Commun. 2001, 282, 1161-1168. (184) McPhail, D. B.; Hartley, R. C.; Gardner, P. T.; Duthie, G. G. Kinetic and stoichiometric assessment of the antioxidant activity of flavonoids by electron spin resonance spectroscopy. J. Agric. Food Chem. 2003, 51, 1684-1690. (185) Mukai, K.; Oka, W.; Watanabe, K.; Egawa, Y.; Nagaoka, S. Kinetic study of free-radical-scavenging action of flavonoids in homogeneous and aqueous triton X100 micellar solutions. J. Phys. Chem. A. 1997, 101, 3746-3753. (186) Burton, G. W.; Ingold, K. U. Autoxidation of biological molecules. 1. Antioxidant activity of vitamin E and related chain-breaking phenolic antioxidants in vitro. J. Am. Chem. Soc. 1981, 103, 6472-6477. (187) Sichel, G.; Corsaro, C.; Scalia, M.; Di Bilio, A. J.; Bonomo, R. P. In vitro scavenger activity of some flavonoids and melanins against O2•-. Free Radical Biol. Med. 1991, 11, 1-8. (188) Shahidi, F.; Wanasundara, P. K. J. Phenolic antioxidants. Crit. Rev. Food Sci. Nutr. 1992, 32, 67-103. (189) Di Meo, F.; Lemaur, V.; Cornil, J.; Lazzaroni, R.; Duroux, J.-L.; Olivier, Y.; Trouillas, P. Free radical scavenging by natural polyphenols: Atom versus electron transfer, J. Phys. Chem. A 2013, 117, 2082-2092. (190) Warren, J. J.; Tronic, T. A.; Mayer, J. M. Thermochemistry of proton-coupled electron transfer reagents and its implications. Chem. Rev. 2010, 110, 6961-7001. (191) Amić, A.; Marković, Z.; Marković, J. M. D.; Stepanić, V.; Lucić, B.; Amić, D. Towards an improved prediction of the free radical scavenging potency of flavonoids: The significance of double PCET mechanisms. Food Chem. 2014, 152, 578-585.
93 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
2646 2647 2648 2649 2650 2651 2652 2653 2654 2655 2656 2657 2658 2659 2660 2661 2662 2663 2664 2665 2666 2667 2668 2669 2670 2671 2672 2673 2674 2675 2676 2677 2678 2679 2680 2681 2682 2683 2684 2685 2686 2687 2688 2689 2690 2691 2692 2693 2694 2695
Page 94 of 105
(192) Bakhouche, K.; Dhaouadi, Z.; Jaidane, N.; Hammoutène, D. Comparative antioxidant potency and solvent polarity effects on HAT mechanisms of tocopherols, Comput. Theoretic. Chem. 2015, 1060, 58-65. (193) Basolo, F.; Pearson, R. G. Mechanism of Inorganic Reactions-A Study of Metal Complexes in Solution; John Wiley & Sons: New York, 1960; pp. 303-331. (194) Laguerre, M.; Giraldo, L. J. L.; Lecomte, J.; Figueroa-Espinoza, M. C.; Barea, B.; Weiss, J.; Decker, E. A.; Villeneuve, P. Chain length affects antioxidant properties of chlorogenate esters in emulsion: The cutoff theory behind the polar paradox. J. Agric. Food Chem. 2009, 57, 11335-11342. (195) Laguerre, M.; Giraldo, L. J. L.; Lecomte, J.; Figueroa-Espinoza, M. C.; Barea, B.; Weiss, J.; Decker, E. A.; Villeneuve, P. Relationship between hydrophobicity and antioxidant ability of “phenolipids” in emulsion: A parabolic effect of the chain length of rosmarinate esters. J. Agric. Food Chem. 2010, 58, 2869-2876. (196) Lucas, R.; Comelles, F.; Alcántara, D.; Maldonado, O. S;, Curcuroze, M.; Parra, J. L.; Morales, J. C. Surface-active properties of lipophilic antioxidants tyrosol and hydroxytyrosol fatty acid esters: A potential explanation for the nonlinear hypothesis of the antioxidant activity in oil-in-water emulsions. J. Agric. Food Chem. 2010, 58, 8021-8026. (197) Shahidi, F.; Zhong, Y. Revisiting the polar paradox theory: A critical overview. J. Agric. Food Chem. 2011, 59, 3499-3504. (198) Panya, A. Strategies to Improve the Performance of Antioxidants in Oil-inWater Emulsions. Ph.D. Thesis, University of Massachusetts, 2012. (199) A. Mortensen, L. H. Skibsted, A. Willnow, S. A. Everett, Re-appraisal of the tocopheroxyl radical reaction with beta-carotene: evidence for oxidation of vitamin E by the beta-carotene radical cation, Free Radical Res. 28 (1998) 69-80. (200) Bisby, R. H.; Parker, A. W. Reaction of ascorbate with the α-tocopheroxyl radical in micellar and bilayer membrane systems. Arch. Biochem. Biophys. 1995, 317, 170-178. (201) Pedrielli, P.; Skibsted, L. H. Antioxidant synergy and regeneration effect of quercetin, (−)-epicatechin, and (+)-catechin on alpha-tocopherol in homogeneous solutions of peroxidating methyl linoleate. J. Agric. Food Chem. 2002, 50, 71387144. (202) Hudson, B. J. F.; Lewis, J. I. Polyhydroxy flavonoid anti-oxidant for edible oils-structural criteria for activity, Food Chem. 1983, 10, 147-155. (203) Peyrat-Maillard, M. N.; Cuvelier, M. E.; Berset, C. Antioxidant activity of phenolic compounds in 2,2’-azobis (2-amidinopropane) dihydrochloride (AAPH)induced oxidation: synergistic and antagonistic effect, J. Am. Oil Chem. Soc. 2003, 80, 1007-1012. (204) Hidalgo, M.; Sánchez-Moreno, C.; de Pascual-Teresa, S. Flavonoid-flavonoid interaction and its effect on their antioxidant activity, Food Chem. 2010, 121, 691696. (205) Wang, S.; Meckling, K. A.; Marcone, M. F.; Kakuda, Y.; Tsao, R. Synergistic, additive, and antagonistic effects of food mixtures on total antioxidant capacities, J. Agric. Food Chem. 2011, 59, 960-968. (206) Skroza, D.; Mekinić, I. G.; Svilović, S.; Simat, V.; Katalinić, V. Investigation of the potential synergistic effect of resveratrol with other phenolic compounds: A case of binary phenolic mixtures, J. Food Compos. Anal. 2015, 38, 13-18. (207) Prieto, M. A.; Curran, T. P.; Gowen, A.; Vázquez, J. A. An efficient methodology for quantification of synergy and antagonism in single electron transfer antioxidant assays, Food Res. Int. 2015, 67, 284-298. 94 ACS Paragon Plus Environment
Page 95 of 105
2696 2697 2698 2699 2700 2701 2702 2703 2704 2705 2706 2707 2708 2709 2710
Journal of Agricultural and Food Chemistry
(208) Vitaglione, P.; Napolitano, A.; Fogliano, V. Cereal dietary fibre: a natural functional ingredient to deliver phenolic compounds into the gut. Trends Food Sci. Technol. 2008, 19, 451-463. (209) Serpen, A.; Capuano, E.; Fogliano, V.; Gökmen, V. A new procedure to measure the antioxidant activity of insoluble food components. J. Agric. Food Chem. 2007, 55, 7676-7681. (210) Tufan, A. N.; Çelik S. E.; Özyürek, M.; Güçlü, K.; Apak, R. Direct measurement of total antioxidant capacity of cereals: QUENCHER-CUPRAC method. Talanta. 2013, 108, 136-142. (211) Serpen, A.; Gökmen, V.; Fogliano, V.; Solvent effects on total antioxidant capacity of foods measured by direct QUENCHER procedure, J. Food Compos. Anal. 2012, 26, 52-57. (212) Gökmen, V.; Serpen, A.; Fogliano, V. Direct measurement of the total antioxidant capacity of foods: The ‘QUENCHER’ approach. Trends Food Sci. Technol. 2009, 20, 278-288.
95 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 96 of 105
Figure Captions: Figure 1. Direct (competitive) antioxidant assay, involving a fluorogenic or chromogenic probe and biologically relevant ROS/RNS. Figure 2. Indirect (non-competitive) antioxidant assay, in which physiological redox reactions (i.e., oxidant-antioxidant interactions) are simulated on an artificial probe without biologically relevant ROS/RNS. Figure 3. Characteristic functional groups having a key role in high antioxidant capacity of a flavonoid (like quercetin). Figure 4. Reaction scheme for the CUPRAC antioxidant capacity assay (liberated protons are buffered by ammonium acetate). Figure 5. Methyl-β-cyclodextrin oligosaccharide, retaining lipophilic antioxidants in the hydrophobic inner core, while holding hydrophilic antioxidants on the outer surface. Figure 6. (a) Salicylate exposed to hydroxyl radicals generated in a Fenton system is converted to highly CUPRAC-reactive species (CUPRAC-rs), i.e. dihydroxy-benzoate (DHBA) isomers, enabling a turn-on colorimetric assay for hydroxyl radicals. (b) Primary hydroxylation products of salicylate. Figure 7. The superoxide anion radicals generated by a non-enzymatic PMS-NADH system attack the CUPRAC-reactive TBHQ probe, converting it to non-reactive TBB-quinone (TBBQ); this conversion is less in the presence of antioxidants, enabling a modified CUPRAC assay for superoxide scavenging antioxidants. Figure 8. Enlargement of silver nanoparticles by antioxidant addition to a silver nitrate colloidal solution of citrate-stabilized Ag-nanoparticles (via seed-mediated particle growth to generate core-shell AgNPs). Figure 9. Surface plasmon resonance absorption of citrate-stabilized AgNPs is intensified by the addition of apple juices, corresponding to seed-mediated particle growth. Figure 10: AuNPs−adsorbed Ellman’s reagent (DTNB-disulfide) is exchanged for thiols (SH) in solution, enabling a colorimetric thiol assay.126
96 ACS Paragon Plus Environment
Page 97 of 105
Journal of Agricultural and Food Chemistry
Figure 1
97 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Figure 2
98 ACS Paragon Plus Environment
Page 98 of 105
Page 99 of 105
Journal of Agricultural and Food Chemistry
Figure 3
Figure 4
99 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Figure 5
100 ACS Paragon Plus Environment
Page 100 of 105
Page 101 of 105
Journal of Agricultural and Food Chemistry
Figure 6
101 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Figure 7
102 ACS Paragon Plus Environment
Page 102 of 105
Page 103 of 105
Journal of Agricultural and Food Chemistry
Figure 8
103 ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Figure 9
Figure 10
104 ACS Paragon Plus Environment
Page 104 of 105
Page 105 of 105
Journal of Agricultural and Food Chemistry
TOC Graphic
105 ACS Paragon Plus Environment