You are on page 1of 13

Molecular Physics

An International Journal at the Interface Between Chemistry and


Physics

ISSN: (Print) (Online) Journal homepage: https://www.tandfonline.com/loi/tmph20

Theoretical investigation on the structure


and antioxidant activity of (+) catechin and (−)
epicatechin – a comparative study

S. Anitha, S. Krishnan, K. Senthilkumar & V. Sasirekha

To cite this article: S. Anitha, S. Krishnan, K. Senthilkumar & V. Sasirekha (2020)


Theoretical investigation on the structure and antioxidant activity of (+) catechin and
(−) epicatechin – a comparative study, Molecular Physics, 118:17, e1745917, DOI:
10.1080/00268976.2020.1745917

To link to this article: https://doi.org/10.1080/00268976.2020.1745917

View supplementary material Published online: 29 Mar 2020.

Submit your article to this journal Article views: 306

View related articles View Crossmark data

Citing articles: 14 View citing articles

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=tmph20
MOLECULAR PHYSICS
2020, VOL. 118, NO. 17, e1745917 (12 pages)
https://doi.org/10.1080/00268976.2020.1745917

RESEARCH ARTICLE

Theoretical investigation on the structure and antioxidant activity of (+) catechin


and (−) epicatechin – a comparative study

S. Anithaa , S. Krishnanb , K. Senthilkumarb and V. Sasirekhaa


a Department of Physics, Avinashilingam Institute for Home Science and Higher Education for Women, Coimbatore, India; b Department of
Physics, Bharathiar University, Coimbatore, India

ABSTRACT ARTICLE HISTORY


The monomeric flavan-3-ols, namely (+) catechin (CT) and (−) epicatechin (ECT), are two impor- Received 23 January 2020
tant antioxidants available in nature. Density functional theory method has been used to study the Accepted 15 March 2020
homolytic, heterolytic bond cleavages and the associated radical scavenging activities at B3LYP/6- KEYWORDS
311G(2d,2p) level of theory. The radical scavenging mechanism of CT and ECT was studied by (+) catechin; (−)
considering three scavenging mechanisms, namely hydrogen atom transfer (HAT), single-electron epicatechin;
transfer-proton transfer (SET-PT), and sequential proton loss electron transfer (SPLET). Parameters thermodynamical
related to the above mechanisms, such as bond dissociation enthalpy (BDE), ionisation potential (IP), mechanism; global
proton dissociation enthalpy (PDE), proton affinity (PA) and electron transfer enthalpy in gas and sol- parameters; MEP
vent medium (benzene, methanol and water) were studied. The results suggest that 4’ -OH site from
B-ring would play a crucial role in the scavenging activity of both the compounds. We observed
that HAT is a thermodynamically favoured mechanism in the gas phase, whereas SPLET mechanism
is more preferential in the polar medium for both the compounds. The global descriptors, such as
ionisation potential (IPv ), electron affinity (EAv ), chemical hardness (η), softness (S), electronegativity
(χ ) and electrophilic index (ω) also confirm the high antioxidant activity of CT and ECT. Reactive site
of the electrophilic and nucleophilic attack is confirmed and visualised by molecular electrostatic
potential (MEP) map.

Highlights

• The optimised geometrical parameters of (+) catechin (CT) and (−) epicatechin (ECT) have been
calculated using B3LYP with 6-311G(2d,2p) level of theory.
• The antioxidant properties of CT and ECT are compared using three mechanisms, namely hydro-
gen atom transfer (HAT), single-electron transfer-proton transfer (SET-PT), and sequential proton
loss Electron transfer (SPLET) by homolytic and heterolytic bond cleavages.
• The hydroxyl group present at 4’ site from B-ring is the most preferential site for hydrogen
donation in both the compounds.
• The HAT mechanism is thermodynamically preferred in the gas phase, while SPLET mechanism is
more favourable in polar solvents.
• Identification of reactive site by MEP and global descriptors is also confirming the high antioxidant
activity in both compounds at the gas phase

CONTACT V. Sasirekha saserekha@gmail.com


Supplemental data for this article can be accessed here. https://doi.org/10.1080/00268976.2020.1745917

© 2020 Informa UK Limited, trading as Taylor & Francis Group


2 S. ANITHA ET AL.

1. Introduction group of polyphenols and act as preventing agent against


prooxidants [15–17].
Free radicals are produced due to oxidative stress and are
Flavan-3-ols are one of the important class of
termed as reactive oxygen species which include super-
flavonoids identified in grapes, green tea, blue berries and
oxide (O− 2 ), hydrogen peroxide (H2 O2 ), hydroxyl radical cocoa [18–21]. The antioxidant activity of the flavan-3-
(• OH) and nitric oxide (NO– ) [1,2]. Oxidation by free
ols is related to their structural properties. Flavan-3-ols
radicals is the origin for the damage of biological com-
(flavanols) are characterised by resorcinol and catechol
plexes, such as DNA, lipids and proteins. The presence
moiety, namely A ring and B ring, that are interconnected
of free radicals induces various degenerative diseases,
by Pyron ring (C ring) [22,23] refer Figure 1. The radi-
like inflammatory, ischaemic, neuro disorder and can-
cal scavenging properties of flavan-3-ols depend mainly
cer. Antioxidants are scavengers of free radicals, thereby
on the distribution of hydroxyl groups and its H-atom
protecting the human cell from the oxidative damage
donating capacity [24]. The stability of phenoxy radi-
caused by free radicals [3,4]. The study on natural antiox-
cal generated after hydrogen atom transfer (HAT) also
idants that are extracted from natural sources is receiv-
contributes to their scavenging action against reactive
ing more attention in recent years [5,6]. Polyphenols
oxygen radicals [25,26]. Based on the projection angle
are the important natural antioxidants found in plant
of a hydroxyl group bonded to C ring at C3 atom and
sources. Previous studies show that polyphenols are effi-
projection angle of B ring bonded to C ring at C2 atom,
cient radical scavenger than vitamins C, E and β-carotene
catechin has four diastereoisomers, namely (+) catechin
[7–10]. The antioxidant capacity of polyphenolic com-
(2R,3S), (−) epicatechin (2R,3R), (−) catechin (2S,3R)
pounds is mainly depending on their ability to scavenge
and (+) epicatechin (2S,3S) [27]. The numbers in the
free radicals and is widely used to suppress the oxida-
bracket proceeding names of the structure are indicative
tive damage, and is having high biological functionality
of site of a projected bond, while R refers to dashed wedge
that includes anti-inflammatory, antimicrobial, antiviral
bonds (bonds that are projecting out of paper away from
and anticancer activities [11–14]. Flavonoids are essen-
the viewer) and S refers to solid wedge bond (bonds that
tial aromatic secondary metabolites that belong to the
projects out of paper towards the viewer), refer Figure 2.
Among these four stereoisomers, (+) catechin (CT a
trans-isomer) and (−) epicatechin (ECT a cis isomer) are
the most important naturally occurring flavan-3-ols that
are present in grape seed [28–30]. CT and ECT are main
monomers characterised by their tendency to form com-
plex structures, namely procyanidins. Their antioxidant
properties also depend on the degree of polymerisation
[31]. The radical scavenging ability of CT and ECT is
due to their structural characteristics, such as dihydroxyl
group at C-3’ and C-4’ on the B ring, C3-hydroxyl group
on C ring with the absence of 2, 3 double bond, and C5,
Figure 1. Structure of flavan-3-ols. C7 hydroxyl groups on ring A. Here, catechol ring (B) has

Figure 2. Structure of stereoisomers (a) (+) catechin, (−) epicatechin and (b) (−) catechin, (+) epicatechin.
MOLECULAR PHYSICS 3

high electron-donating ability compared to other rings reactive site for electrophilic and nucleophilic attacks in
due to the presence of ortho-dihydroxyl group [32–35]. CT and ECT.
Many studies reported that antioxidant activity of CT
and ECT is mainly due to their high redox properties
2. Computational details
that are responsible for the inhibition of various oxida-
tive free radicals [36–40]. The antioxidant behaviour of The geometry of neutral state and radical of CT and
CT and ECT present in various plant extract was evalu- ECT in the gas and solvent medium was optimised using
ated by various experimental antioxidant assays, such as the DFT method with hybrid exchange-correlation func-
1,1-diphenyl-1-picrylhydrazyl (DPPH. ), 2, 2 -azino-bis tional, B3LYP [56,57] and 6-311G(2d,2p) basis set. The
(3-ethylbenzothiazoline-6-sulfonic acid) (ABTS+ ), nitric B3LYP method is known to provide accurate results
oxide (NO. ),ferric reducing antioxidant power (FRAP) of structure and themodynamical properties of pheno-
and experimental results show that CT and ECT are lic compounds [58,59]. For radical structure, the unre-
powerful antioxidants [41–43]. Mendoza-Wilson et al. stricted method was used. The harmonic vibrational fre-
[44] reported that CT and ECT are similar in molecu- quency calculation was performed for all the optimised
lar structure, but their chemical reactivity properties are structures at the above level of theory to confirm the
different. Leopoldini et al. [45] stated that the antioxi- stationery state of the optimised structure. The solvent
dant ability of CT is depending on its planar geometry. effect has been introduced with polar media, methanol
Hydrogen bond interaction at catechol moiety is also the and water, and non-polar media, benzene using self-
main feature responsible for the antioxidant activity of consistent reaction field method with integral equation
ECT [46]. Vagánek et al. [47] reported that the absence formalism of the polarised continuum model (IEF-PCM)
of C2 = C3 double bond at C ring has a large impact on [47,60]. The electronic structure calculations were per-
the antioxidant ability of ECT. Besides, it has been shown formed using Gaussian 09 program package [61].
that the reactivity of flavan-3-ols is further altered ther- While studying the scavenging ability of the antiox-
modynamically in accordance with the effect of solvents idant we focus on the feasibility of hydrogen abstrac-
[48,49]. tion from antioxidant species. Three antioxidant mech-
Density functional theory has been used to evalu- anisms are applied to analyse the deactivation of free
ate the relationship between structural properties and radical (R· ) by antioxidants CT and ECT (ArOH), as
the antioxidant activity [50–52]. The thermodynamical shown in Figure 3. All the three mechanisms are directed
parameters, such as O–H bond dissociation enthalpy towards neutralising the free radicals by the donation of
(BDE), ionisation potential (IP), proton dissociation the hydrogen atom from antioxidant and thus reducing
enthalpy (PDE), proton affinity (PA) and electron trans- the reactivity of the species. The three mechanisms are
fer enthalpy (ETE), are useful tools to identify an active diversified on the basis of the pathway of hydrogen atom
site and preferred radical scavenging mechanism induced abstraction from antioxidants and addition of H-atom at
in a molecule [53–55]. To the best of our knowledge, free radical. The first mechanism involves direct H-atom
there is no detailed comparative study on the structure transfer (HAT), the second mechanism involves elec-
and radical scavenging activity relationship between CT tron transfer followed by deprotonation (SET-PT) and in
and ECT. In the present investigation, the competence of the third mechanism, anion is formed followed by pro-
antioxidant behaviour of CT and ECT is compared with ton transfer (SPLET) [62,63]. In the HAT mechanism, H
the structure-related properties using the established atom is transferred to free radical (R. ) from OH groups
mechanisms, such as H-atom transfer (HAT), single- of ArOH, as given in Equation (1).
electron transfer-proton transfer (SET-PT) and sequen-
tial proton loss electron transfer (SPLET), the favoura- ArOH + R· → ArO· + RH (1)
bility of mechanism and impact of solvent medium are
The feasibility of this mechanism is studied by the esti-
investigated. Molecular descriptors were evaluated at the
mation of dissociation enthalpy of O–H bond by BDE,
gas medium to predict the antioxidant behaviour of both
The BDE value can be calculated using the following
the compounds. The frontier molecular orbitals (FMO)
equation
and the energy gap of both compounds were computed.
Density of states (DOS) spectrum for occupied and unoc- BDE = H(ArO· ) + H(H· ) − H(ArOH) (2)
cupied molecular orbitals is interpreted. The above find-
ings are helpful in understanding the possible mechanism where H(ArO· ) is the enthalpy of radical formed after
of radical scavenging activity of CT and ECT. Molecu- H-atom abstraction from ArOH, H(H· ) is the enthalpy
lar electrostatic potential map is utilised to identify the of the hydrogen atom, H(ArOH) is the enthalpy of
4 S. ANITHA ET AL.

Figure 3. Antioxidant mechanism of CT and ECT (1)-HAT, (3) and (5)-SET-PT, (7) and (9)-SPLET.

the neutral molecule. In the SET-PT mechanism, cation (Equation (7)).


radical(ArOH.+ ) is formed in the first step through elec-
tron transfer from ArOH to R· (Equation (3)). ArOH → ArO− + H+ (7)

In this step PA is calculated as


ArOH + R· → ArOH.+ + R− (3)
PA = H (ArO− ) + H(H + ) − H(ArOH) (8)
Ionisation Potential (IP) of ArOH is calculated by
where H (ArO− ) is the enthalpy of anion after the elim-
ination abstraction of H+ . In the second step, ArO· is
IP = H(ArOH·+ ) + H(e− ) − H(ArOH) (4) formed due to the loss of an electron from anion, ArO−
and is represented in Equation (9).
where H(ArOH·+ ) is the enthalpy of radical cation
after electron abstraction and H(e– ) is the enthalpy ArO− + R· + H+ → ArO· + RH (9)
of an electron. Equation (5) describes the deproto-
This step corresponds to ETE which is calculated by
nation from ArOH+ , which is the second step in
SET-PT, ETE = H(ArO· ) + H(e− ) − H(ArO− ) (10)

ArOH·+ + R− → ArO· + RH (5) In the present study, the above thermodynamical


quantities are calculated for both CT and ECT at both gas
and solvent medium, respectively. Gas-phase enthalpy of
Hence, PDE is calculated as
hydrogen, proton and electron is −0.499897, 0.002363
and 0.001198 hartree, respectively, as reported in the
PDE = H(ArO· ) + H(H + ) − H(ArOH·+ ) (6) previous study [64]. The enthalpy values H· , H+ , e– in
methanol, benzene and water have been calculated by
where H(H+ ) is the enthalpy of the proton. including the solvent effect [65,66].
In the SPLET mechanism, phenoxide anion (ArO– ) The global descriptors, namely ionisation potential
is formed due to the loss of the proton from ArOH (IPV ), electron affinity (EAV ), chemical hardness (η),
MOLECULAR PHYSICS 5

softness (S), electronegativity (χ) and electrophilic index through electron density distribution. MEP is computed
(ω), are computed by using Equations (11)–(17). Ionisa- from the optimised structure of both the molecules and
tion potential (IPV ) is calculated from vertical energy as is plotted by using Arguslab software [70].
the difference between the energy of cationic form and
neutral molecule. Electron affinity(EAV ) is calculated as
the energy difference between the neutral and anionic 3. Results and discussion
form of the molecule [67,68]. 3.1. Molecular structure
IPV = Ecation − En (11) The optimised structures of neutral state CT and ECT
EAV = En − Eanion (12) in the gas medium are shown in Figure 4. The ori-
entation of C3–OH and C3–H bonds is the chief dif-
From the calculated IPV and EAV , chemical hardness ferentiating aspects between CT and ECT. The torsion
(η), softness (S), electronegativity (χ ) and electrophilic angles, such as 1 (C2–C3–O–H), 2 (O1–C2–C3–H),
index (ω) are calculated as follows. 3 (O1–C2–C3 –O), 4 (C1’–C2–C3–O) and 5 (C1’
–C2–C3–H) concerning the orientation of C3–OH and
IP + EA
Electronegativity (χ ) = (13) C3–H bonds, are analysed for CT and ECT at gas, ben-
2 zene, methanol and water environments and are sum-
IP + EA
Chemical potential (μ) = − (14) marised in Table 1. Out of the above mentioned five
2 torsional angles, 1 (C2–C3–O–H), 2 (O1–C2–C3–H),
IP − EA and 3 (O1–C2–C3–O) are for C ring. The sig-
Chemical hardness (η) = (15)
2 nificant changes observed in two torsion angles 4
1 (C1’–C2–C3–O) and 5 (C1’–C2–C3–H) confirm the
Softness (S) = (16)
2η trans and cis isomerism for CT and ECT, as represented
μ2 in Figure 2 and a previous study [44]. The torsion angle
Electrophilicity index (ω) = (17) between B and C rings (O1–C2–C1’–C2’) is observed as

−36° for CT and −33° for ECT, it is indicative of non-
FMO are plotted for both the molecules and are anal- planarity in CT and ECT. When the polarity of solvent
ysed along with the DOS spectrum using GaussSum (ver- changes, there was no significant change in the torsion
sion 3.0) software [69]. Molecular electrostatic poten- angles 2 , 3 , 4 and 5 of both the molecules. The
tial (MEP) is a three-dimensional map to visualise the slight change is noted only in torsion angle 1 . That
charge distribution and to measure the strength of nearby is, the solvent environment does not affect the struc-
charges at a particular site. MEP is used to predict elec- tural parameter much [71]. The ortho-dihydroxy groups,
trophilic and nucleophilic sites of both the molecules 3’ –OH and 4’ –OH, are linked through intramolecular

Figure 4. Optimised structures of CT and ECT.

Table 1. Dihedral angle () in degrees (°) and relative energy (EE ) in kcal/mol of CT and ECT in the gas and solvent phase, respectively.
CT ECT
Phase 1 2 3 4 5 EE 1 2 3 4 5 EE
Gas −176.7 59.8 178.7 58.2 −60.7 0.00 53.8 −178.4 59.9 −61.3 60.4 0.00
Benzene −176.6 59.4 178.3 57.9 −60.9 −5.72 55.2 −178.5 59.9 −61.3 60.2 −5.53
Methanol −179.1 58.4 177.3 56.9 −61.9 −13.15 60.1 −178.7 59.8 −61.5 59.9 −12.99
Water −179.2 58.3 177.2 56.9 −61.9 −13.59 60.8 −178.7 59.8 −61.5 59.9 −13.46
Note: 1 (C2–C3–O–H), 2 (O1–C2–C3–H), 3 (O1–C2–C3–O), 4 (C1’–C2–C3–O), 5 (C1’–C2–C3–H).
6 S. ANITHA ET AL.

hydrogen bond in a neutral molecule. The hydrogen O–H BDE values in gas, benzene, methanol and water
bond length is found to be 2.14 Å in CT and ECT. The medium are summarised in Table 2. In the gas medium,
presence of hydrogen bond at 4’-O and 3’-OH provides the most stable radical is formed by H atom abstrac-
additional stability for CT and ECT [44]. tion at 4’-OH site in both the molecules. The relative
enthalpy (Hg ) and relative energy (Eg ) were calcu-
lated with respect to 4’-OH site for the remaining radical
3.2. Solvent effects on molecular stability
sites [72]. It is observed that the radical produced by
The stability of CT and ECT was characterised by the the abstraction of H-atom at other sites, 5-OH, 7-OH,
difference in total energy in different solvent environ- 3-OH and 3’-OH has higher enthalpy and energy. High
ments. The ground state energy of CT and ECT at gas BDEg of 98.95 kcal/mol for CT and 100.50 kcal/mol for
medium is almost the same and there is a slight differ- ECT at 3-OH site was observed. This shows that H-atom
ence in their energies at solvent medium. This shows abstraction from the site 3-OH is more difficult than
that the stereochemistry of both these compounds is not the other sites. This is due to the absence of a double
much influential upon their total energy [71]. The rela- bond between C2 and C3 atoms in C ring which sup-
tive energy (EE ) is calculated for both the molecules presses the resonance effect [45,72]. The lower BDEg of
in a different medium with respect to the energy at the 72.33 and 73.27 kcal/mol for CT and ECT, respectively,
gas phase and summarised in Table 1. The value of EE was observed for H-atom abstraction from 4’ -OH site.
observed at water media is −13.59 kcal/mol for CT and This is due to the presence of ortho-dihydroxy groups on
−13.46 kcal/mol for ECT. In methanol, EE value of B-ring in both the compounds. The strong intramolec-
CT is −13.15 kcal/mol and ECT is −12.99 kcal/mol. In ular hydrogen bond (O–H . . . O) is formed between 3’-
the case of benzene, the value of EE for CT is −5.72 OH and 4’-O in the radical of CT and ECT, we infer that
kcal/mol and for ECT EE is −5.53 kcal/mol. That is, the this intramolecular hydrogen bond would stabilise the
stability of CT and ECT is enhanced by an increase in the radical. The stability of catechol moiety brought by the
polarity of the solvent. effects of conjugation, delocalisation and resonance will
influence the antioxidant activity of the molecule [32,73].
The order of stability for CT and ECT is 4’-OH > 5-OH
3.3. Bond dissociation enthalpy (BDE) > 7-OH > 3’-OH > 3-OH which confirms that HAT is
In the HAT mechanism, a hydrogen atom is abstracted easier from 4’-OH site (B ring) than that of other sites.
from the flavon-3-ol hydroxyl groups by the free radical The antioxidant activity of flavan-3ols may depend
through homolytic cleavage. In this case, the antioxi- on the polarisation induced by solvents [64]. The rela-
dant capacity of the compound is evaluated by BDE. The tive energy and relative enthalpy calculated at benzene,
lower BDE value represents the higher antioxidant activ- methanol and water medium with reference to the sta-
ity of the compound. The optimised structure of CT and ble radical 4’ -OH and are summarised in Table 2. From
ECT was used to generate radicals by abstracting H-atom our results, it is observed that hydrogen-donating abil-
from the site, C5–OH, C7–OH, C3–OH, C3’–OH and ity in the non-polar medium is as good as that in the
C4’–OH. These radical sites are represented as 5-OH, 7- gas medium, and the sequence is different in the polar
OH, 3-OH, 3’-OH and 4’-OH, respectively for both the medium. For the water medium, BDE is slightly higher
molecules and are shown in Figures S1 and S2. The rela- than that of the gas medium and BDE is further higher
tive energy (E), relative enthalpy (H) and calculated in the other two solvent mediums. The lowest BDE value

Table 2. The relative energies (E) and relative enthalpies (H) and bond dissociation enthalpy (BDE) values of CT and ECT in the gas
and solvent phase (kcal/mol).
Position Gas phase Benzene Methanol Water
CT Eg Hg BDEg Eb Hb BDEb Em Hm BDEm Ew Hw BDEw
4’-OH 0.00 0.00 72.33 0.00 0.00 74.51 0.00 0.00 75.34 0.00 0.00 73.28
3’-OH 10.23 9.84 82.18 8.47 8.22 82.73 5.96 5.88 81.22 5.81 5.73 79.01
7-OH 7.24 6.73 79.85 8.89 8.15 82.66 7.49 6.95 82.29 7.4 6.87 80.15
5-OH 8.18 7.51 79.06 7.65 7.01 81.52 6.93 6.35 81.70 6.89 6.31 79.59
3-OH 27.5 26.61 98.95 27.33 26.37 100.88 27.64 25.91 101.26 26.64 25.36 98.63
ECT
4’-OH 0.00 0.00 73.27 0.00 0.00 75.30 0.00 0.00 75.73 0.00 0.00 73.64
3’-OH 10.5 10.1 83.36 8.65 8.37 83.68 5.96 5.84 81.50 5.79 5.66 79.30
7-OH 8.34 7.76 81.03 7.52 6.97 82.28 6.82 6.26 81.98 6.79 6.22 79.85
5-OH 7.07 6.49 79.76 6.59 6.02 81.32 6.37 5.77 81.56 6.38 5.77 79.41
3-OH 28.56 27.23 100.50 27.84 26.51 101.81 26.72 25.42 101.15 26.65 25.34 98.98
MOLECULAR PHYSICS 7

is observed for 4’ -OH site in all the mediums. The results 162.91, 142.96, 111.36 and 106.28 kcal/mol in gas, ben-
confirm that HAT from 4’-OH is easier than from other zene, methanol and water medium, respectively. With
sites in both polar and non-polar mediums [66]. In addi- respect to the medium, it is found that IP of CT and ECT
tion to that, the obtained results suggest that the hydro- is decreasing in the order, gas > benzene > methanol >
gen donation ability of CT is marginally higher than water. That is, as the polarity increases, the IP decreases
ECT, implying CT is showing marginally better radical for both the molecules. Notably, in the water medium, the
scavenging ability in terms of HAT mechanism. IP is lesser than that of gas-phase value by ∼ 55 kcal/mol.
The result is in agreement with the previous study [64]
and confirms that polar solvents facilitate the process
3.4. Spin density distribution of electron removal [66]. In all the medium, the calcu-
lated IP of CT and ECT is higher than the O–H bond
The differences in reactivity of O–H sites are studied BDE. The deviation between O–H bond BDE at 4’-OH
through spin density distribution analysis which is given and IP is ∼ 89, ∼ 67, ∼ 35 and ∼ 32 kcal/mol at gas,
in Figure S1 and Figure S2 for CT and ECT, respectively, benzene, methanol and water medium for both CT and
in the supplementary section. If the electron density delo- ECT, respectively. That is, HAT from 4’-OH bond of CT
calisation is higher, then the H-atom abstraction from and ECT to the radical is much easier than the electron
that site is easier [60]. The spin density on the 4’-OH transfer.
site of CT (0.337) is lower than that of other sites, 3-OH PDE determines the deprotonation property of
(0.847), 5-OH (0.361), 7-OH (0.390) and 3’-OH (0.383). molecules. PDE of CT and ECT in gas and solvent
Similarly, for ECT, the lower spin density occurs on the medium is summarised in Table 3. The most stable rad-
site of 4’-OH (0.344) than other sites 3-OH (0.817), 5- ical produced by H-atom abstraction from 4’-OH site
OH (0.362), 7-OH (0.389) and 3’-OH (0.380). That is, is having the lowest PDE of 225.66 kcal/mol in the gas
the spin density is more delocalised at 4’-OH of both medium, 30.17 kcal/mol in benzene, 9.11 and −2.58
compounds due to the intramolecular hydrogen bond. kcal/mol in methanol and water medium for CT, and
By comparing the spin density, we presuppose that CT for ECT the PDE values are 226.28, 31.39, 10.46 and
could have better antioxidant property due to its greater −1.38 kcal/mol in gas, benzene, methanol and water
H-atom abstraction feasibility than ECT. medium, respectively. Thus, proton dissociation ability
is higher at 4’-OH site in B ring. As shown in Table 3,
the PDE is higher for proton dissociation at 3-OH site
3.5. Ionisation potential and proton dissociation
of CT and ECT at all mediums. In the solvent medium,
enthalpy
the PDE value decreases significantly due to high sol-
The ability of electron transfer in title compounds is vation enthalpy of the proton. The PDE in the water
studied through ionisation potential (neutral compound) medium is lower than that of other mediums, and this
values (in relation to SET-PT mechanism). The IP of result is in agreement with the previous result [64].
CT and ECT in gas and solvent medium is summarised Hence, water is the most active medium to dissociate
in Table 3. The calculated IP of CT in gas, benzene, protons easily. From the above discussions, both CT and
methanol and water medium is 162.60, 143.38, 112.32 ECT exhibit similar capability of scavenging via SET-
and 107.13 kcal/mol, respectively and for ECT IP is PT mechanism; however, CT shows marginally better

Table 3. The ionisation potential (IP) and proton dissociation enthalpy (PDE) values (kcal/mol) of CT and ECT at gas and solvent phase,
respectively.
IP PDE SET-PT (IP + PDE)
Position Gas phase Benzene Methanol Water Gas phase Benzene Methanol Water Gas phase Benzene Methanol Water
CT 162.60 143.38 112.32 107.13
4’-OH 225.66 30.17 9.11 −2.58 388.26 173.56 121.43 104.55
3’-OH 235.50 38.40 14.99 3.15 398.10 181.78 127.31 110.27
7-OH 233.17 38.33 16.06 4.29 395.77 181.71 128.38 111.41
5-OH 232.38 37.19 15.46 3.73 394.99 180.57 127.78 110.86
3-OH 252.27 56.55 35.02 22.77 414.87 199.93 147.34 129.90
ECT 162.91 142.96 111.36 106.28
4’-OH 226.28 31.39 10.45 226.28 389.19 174.35 121.81 104.90
3’-OH 236.37 39.77 16.23 236.37 399.29 182.72 127.59 110.57
7-OH 234.04 38.37 16.71 234.04 396.95 181.32 128.07 111.12
5-OH 232.38 37.19 15.46 232.38 395.68 180.37 127.65 110.67
3-OH 252.27 56.55 35.02 252.27 416.42 200.86 147.24 130.24
8 S. ANITHA ET AL.

capability to scavenge the free radical. We also observe thermodynamically favoured at the gas phase, whereas
that 4’ -OH is the most preferential site for homolytic the SPLET mechanism is more preferred in the polar
followed by heterolytic O–H bond cleavage at all medium for both the compounds. The sequence of the
environments. preferential site for deprotonation is the same in all the
three mechanisms, that is H-atom from 4’-OH site of CT
and ECT will cleave and bind with the free radical.
3.6. Proton affinity and electron transfer enthalpy
According to the SPLET mechanism, PA and ETE of CT
3.7. Frontier molecular orbitals
and ECT are calculated and summarised in Table 4 at
gas and solvent environment. In the first step, PA is the FMO, highest occupied molecular orbital (HOMO) and
measure of binding of the proton; lower PA corresponds lowest unoccupied molecular orbital (LUMO) for CT and
to easy deprotonation. The lowest PA of 333.58, 97.39, ECT are shown in Figure 5. Electron donation and elec-
47.75 and 34.33 kcal/mol in gas, benzene, methanol and tron acceptance ability are described by the energy of
water medium, respectively, is observed at 4’-OH site for HOMO and LUMO, respectively. The most active site for
CT and for ECT PA is 336.66, 99.97, 47.21 and 33.94 the nucleophilic attack is characterised by high density at
kcal/mol in gas, benzene, methanol and water medium, HOMO. HOMO of CT clearly shows that electron delo-
respectively. The PA decreases from gas to the solvent calisation is more on ring B than A and C rings [44]. This
medium as the polarity of the solvent increases, which is in agreement with the computed BDE values (Table 2).
is similar to PDE [64]. That is, polar solvents favour In ECT, HOMO is highly localised on ring A. LUMO of
the deprotonation process. ETE is the second step of CT and ECT is mainly localised on ring B [74]. The scarce
the SPLET mechanism; here the electron transfer occurs FMO distribution in C ring of CT and ECT could be
from deprotonated radical. From Table 4 it is observed attributed due to lack of conjugation. HOMO energy of
that ETE of both the compounds is higher in the sol- CT and ECT is −5.69 and −5.67 eV, respectively. It shows
vent medium than the gas medium. ETE of protonated that both CT and ECT have similar electron-donating
radical is lower than the IP of the neutral molecule. This capacity. LUMO energy of CT and ECT is −0.02 eV and
indicates that a single-electron transfer from the depro- −0.33 eV, respectively. It is noticeable that the LUMO of
tonated form is easier than from the neutral form. The ECT is lower than that of CT by 0.31 eV, which is indica-
sum of PA and ETE (SPLET) of each O–H site shows tive that ECT is more electrophilic than CT. The energy
that 4’-OH site of both the compounds has the high gap is found to be 5.67 and 5.34 eV for CT and ECT,
dehydrogenation ability, and the water medium supports respectively.
SET-PT mechanism as well. Also, in terms of SET-PT, CT FMO orbital coefficient is analysed for both CT and
is having better antioxidant property than ECT. ECT (Figure S3). The site 3-OH of CT has lower FMO
Three mechanisms, namely HAT, SET-PT and SPLET, orbital coefficient, which indicates that electron donation
are mainly described by the descriptors BDE, IP and PA, from this site is difficult [75]. FMO orbital coefficient at
respectively. For CT and ECT, gas-phase IP and PA of the 4’ -OH is higher than that of other sites, which shows that
studied compounds are higher than BDE in the gas phase. 4’ -OH is the most active site for donating an electron. In
PA in water and methanol is lower than corresponding ECT, FMO orbital coefficients observed at 5-OH site were
values of BDE and IP. Hence, the HAT mechanism is lower than that of other sites. The high HOMO orbital

Table 4. The proton affinity (PA) and electron transfer enthalpy (ETE) values (kcal/mol) of CT and ECT in the gas and solvent phase,
respectively.
PA ETE SPLET(PA + ETE)
Position Gas phase Benzene Methanol Water Gas phase Benzene Methanol Water Gas phase Benzene Methanol Water
CT
4’-OH 333.58 97.39 47.75 34.33 54.68 76.16 73.68 70.21 388.26 173.56 121.43 104.55
3’-OH 349.48 110.64 57.23 43.56 48.62 71.14 70.09 66.72 398.10 181.78 127.31 110.27
7-OH 345.56 111.21 55.08 41.64 50.21 70.50 73.30 69.78 395.77 181.71 128.38 111.41
5-OH 349.04 107.77 52.96 39.59 45.94 72.80 74.82 71.27 394.99 180.57 127.78 110.86
3-OH 366.37 127.66 70.92 57.35 48.50 72.27 76.43 72.55 414.87 199.93 147.34 129.90
ECT
4’-OH 336.66 99.97 47.21 33.94 52.53 74.38 74.60 70.96 389.19 174.35 121.81 104.90
3’-OH 351.24 112.18 53.71 42.25 48.05 70.55 73.88 68.32 399.29 182.72 127.59 110.57
7-OH 348.91 112.72 55.63 42.11 48.04 68.60 72.43 69.02 396.95 181.32 128.07 111.12
5-OH 349.49 109.99 55.76 40.23 46.19 70.38 71.89 70.44 395.68 180.37 127.65 110.67
3-OH 370.58 131.13 72.72 59.06 45.84 69.73 74.52 71.19 416.42 200.86 147.24 130.24
MOLECULAR PHYSICS 9

Figure 5. FMO distribution of neutral CT and ECT.

coefficient seen at 7-OH site indicates that 7-OH has high Table 5. Global molecular descriptors of CT and ECT (eV) in the
electron-donating ability compared to other sites. DOS gas phase.
spectrum of CT and ECT is depicted in Figure S4. It is Global molecular descriptors CT ECT
observable that the energy states of frontier orbitals are Ionisation potential 7.02 7.04
different for CT and ECT, indicative of distinct chemical Electron affinity −1.21 −0.84
Chemical potential −2.91 −3.10
properties. Chemical hardness 4.11 3.94
Softness 0.12 0.13
Electronegativity 2.91 3.10
Electrophilic index 1.03 1.22
3.8. Global molecular descriptors
The global molecular descriptors, such as chemical hard-
ness, chemical potential and softness, will provide infor-
index and LUMO energy level show the marginally
mation about chemical reactivity and stability of the
higher electrophilic nature of ECT. We observe that the
molecules [74]. The calculated global molecular descrip-
information obtained from global molecular descriptors
tors for CT and ECT are summarised in Table 5. It also
are in accord with the results of HAT, SET-PT, SPLET and
provides information about the tendency to release or
FMO analysis.
capture the electron, which is one of the key aspects
of the antioxidant potential. It is observed that ionisa-
tion potential (IPV ) of CT and ECT is 7.02 and 7.04
3.9. Molecular electrostatic potential
eV, respectively. The electron affinity (EAV ) of CT and
ECT is −0.84 eV and −1.21 eV, respectively. These val- The MEP is another important descriptor for identi-
ues are comparable with ionisation potential and electron fying the electrophilic and nucleophilic active site of a
affinity of quercetin flavonoid, a powerful antioxidant molecule [77]. MEP is a 3D plot of charge distribu-
(IPV = 7.22 eV and EAV = 0.76 eV) [76], which once tion and it is investigated from the optimised geometry
again asserts the antioxidant potential of both the com- of neutral state of CT and ECT in the gas medium at
pounds. The chemical hardness is a measure of resistance B3LYP/6-311G(2d,2p) level of theory using Argus lab
to charge transfer and its inverse gives softness. The neg- software. A plot of 3D MEP surface of CT and ECT
ative value of chemical potential in CT and ECT is indica- is shown in Figure 6. In the ESP map, the maximum
tive of the stability of the molecule. Calculative chemical negative region shown in red colour is indicative of nucle-
hardness and chemical potential show that CT is slightly ophilic attack and electrophilic attack site presented in
less stable than the ECT. Lower stability would corre- blue colour which represents the positive region. The
spond to better hydrogen donation, that needs better electron (nucleophilic) is mainly concentrated on oxygen
antioxidant property, an observation that is reflective of atom at sites, 3’ -OH and 4’ -OH on B-ring, 3-OH and O1
the results of HAT, SET-PT and SPLET. The electrophilic on C-ring, 5-OH and 7-OH on A-ring of both the com-
index defines the affinity of electrons. The electrophilic pounds, indicated by red colour. The electron-deficient
10 S. ANITHA ET AL.

Figure 6. Molecular electrostatic potential map of neutral CT and ECT.

region (electrophilic) is observed on hydrogen atom of B- Disclosure statement


ring for CT and A-ring for ECT and is indicated by blue No potential conflict of interest was reported by the author(s).
colour.

References
4. Conclusions [1] C. Kandaswami and E. MiddletonJr, Free. Radic. Diag.
Med 366, 351–376 (1994).
In the present work, density functional theory method [2] A. Phaniendra, D. Babu and P. Latha, Ind. J. Clin.
with B3LYP/6-311G(2d,2p) has been employed for the Biochem. 30, 11–26 (2015).
[3] P. Iacopini, M. Baldi, P. Storchi and L. Sebastiani, J. Food.
systematic comparison of the antioxidant and scaveng- Comp. Ana. 21, 589–598 (2008).
ing ability between (+) catechin (CT) and (−) epicat- [4] A. Galano, G. Mazzone, R. Alvarez-Diduk, T. Marino,
echin (ECT). The results show that the site, 4’-OH is J.R. Alvarez-Idaboy and N. Russo, Annu. Rev. Food Sci.
the most preferential active site for hydrogen donation Technol. 7, 335–352 (2016).
in both the compounds as per all the three investigated [5] A. Masek, E. Chrzescijanska, M. Latos, M. Zaborski and
A. Podsędek, Int. J. Electrochem. Sci. 12, 6600–6610
scavenging mechanisms. It is found that the intramolec-
(2017).
ular hydrogen bond plays a vital role in deciding the [6] M. Oroian and I. Escriche, Food Res. Int. 74, 10–36
preferential site of the radical attack by altering the elec- (2015).
tron density delocalisation particularly on FMOs. It is [7] D. Bagchi, M. Bagchi, S.J. Stohs, D.K. Das, S.D. Ray, C.A.
observed that 4’-OH and 7-OH sites have more FMO Kuszynski, S.S. Joshi and H.G. Pruess, Toxicology. 148,
distribution for CT and ECT, respectively. The outcome 187–197 (2000).
[8] P. Siddhuraju, Food Chem. 99, 149–157 (2006).
of the HAT mechanism suggested that antioxidant activ- [9] K. Thaipong, U. Boonprakob, K. Crosby, L. Cisneros-
ity is favourable in the gas phase. The polar solvent zevallos and D. Hawkins, J. Food. Comp. Analy. 19,
had a favouring impact on the SPLET mechanism, espe- 669–675 (2006).
cially in water media (which is a prevalent medium in [10] R. Amarowicz, F. Shahidi and W. Wiczkowski, J. Food Lip.
biological systems) it seems to support this mechanism 10, 165–177 (2003).
[11] D. Choi, Y. Lee, J. Tae and H. Lee, Brain Res. Bulletin. 87,
more. Hence one could say aqueous medium provides
144–153 (2012).
a supportive environment for radical scavenging by CT [12] M. Jabbari, H. Mir, A. Kanaani and D. Ajloo, J. Mol. Liq.
and ECT. Our comparative investigation concludes that 196, 381–391 (2014).
both CT and ECT exhibit similar scavenging ability; [13] G. Baeza, B. Sarriá, L. Bravo and R.M. Briz, J. Agric. Food
however, CT shows marginally better free radical scav- Chem. 51, 9663–9674 (2016).
enging capacity to deprotonation and ionisation. From [14] M. Saqib, S. Iqbal, A. Mahmood and R. Akram, Int. J. Food
Prop. 19 (4), 745–751 (2016).
the energy level of LUMO, we observed that ECT has [15] H.F.P. Martins, J.P. Leal, M.T. Fernandez, V.H.C. Lopes
a more electrophilic tendency that would, in turn, be and D.S. Cordeiro, J. Am. Soc. Mass Spec. 15, 848–861
a parameter influencing the possibilities of attracting (2004).
nucleophilic radicals, this could be an influential param- [16] P.G. Pietta, J. Nat. Prod. 63, 1035–1042 (2000).
eter in scavenging nucleophilic radicals in low concentra- [17] E.I. Korotkova, O.A. Voronova and E.V. Dorozhko, J. Solid
State Electrochem. 16, 2435–2440 (2012).
tion. Thereby, the present comparative study sheds light
[18] V.K. Ananingsih, A. Sharma and W. Zhou, Food. Res. Int.
on scavenging mechanisms and antioxidant potentials of 50, 469–479 (2013).
CT and ECT that could provide an insightful value for the [19] P.V. Gadkari and M. Balaraman, Food. Biopro. Proc. 93,
health sector. 122–138 (2015).
MOLECULAR PHYSICS 11

[20] C.G. Fraga, K.D. Croft, D.O. Kennedy and F.A. Tomás- [47] A. Vagánek, J. Rimarčík, K. Dropková, J. Lengyel and E.
Barberán, Food Funct. 10, 514–528 (2019). Klein, Comp. Theo. Chem. 1050, 31–38 (2014).
[21] S. Martini, A. Conte and D. Tagliazucchi, Food Res. Int. [48] Y. Xue, Y. Zheng, L. Zhang, W. Wu, D. Yu and Y. Liu, J.
97, 15–26 (2017). Mol. Model. 19, 3851–3862 (2013).
[22] M. Šeruga and I. Tomac, Int. J. Electrochem. Sci. 12, [49] Y. Chen, H. Xiao, J. Zheng and G. Liang, Plos One. 10 (3),
7616–7637 (2017). e0121276 (2015). doi:10.1371/journal.pone.0121276.
[23] P.M. Aron and J.A. Kennedy, Mol. Nutr. Food Res. 52, [50] D. Özbakır Işın, Mol. Phys. 114 (24), 3578–3588 (2016).
79–104 (2008). [51] D. Amić, D. Davidović-Amić, D. Bešlo, V. Rastija, B.
[24] J. Wang, H. Tang, B. Hou, P. Zhang, Q. Wang, B. Zhang, Lučić and N. Trinajstić, Current Med. Chem. 14, 827–845
Y. Huang, Y. Wang, Z. Xiang, C. Zi, X. Wang and J. Sheng, (2007).
RSC Adv. 7, 54136–54141 (2017). [52] P. Alov, I. Tsakovska and I. Pajeva, Current Topics in Med.
[25] E.N. Bentz, A.B. Pomilio and R.M. Lobayan, J. Mol. Chem. 15, 85–104 (2015).
Model. 20, 2105 (2014). [53] Z. Marković, D. Amić, D. Milenković, J.M. Dimitrić-
[26] E.N. Bentz, A.B. Pomilio and R.M. Lobayan, J. Mol. Marković and S. Marković, Phys. Chem. Chem. Phys. 15,
Model. 20, 2522 (2014). 7370–7378 (2013).
[27] J. Oliveira, N. Mateus and V. De Freitas, Nat. Prod., [54] P. Škorňa, J. Rimarčík, P. Poliak, V. Lukeš, E. Klein and J.
1753–1801 (2013). doi:10.1007/978-3-642-22144-6. Rimarc, Comp. Theo. Chem. 1077, 32–38 (2016).
[28] A.I. Mandic, S.M. Ðilas, G.S. Ćetković, J.M. Čanadanović [55] D.S. Dimić, D.A. Milenković, J.M.D. Marković and Z.S.
and V.T. Tumbas, Int. J. Food Pro. 11, 713–726 (2008). Marković, Mol. Phys. 116 (9), 1166–1178 (2018).
[29] M. Monagas, C. Gómez-Cordovés, B. Bartolomé, O. Lau- [56] A.D. Becke, Phys. Rev. A. 38 (6), 3098–3100 (1988).
reano and J.M.R. Da Silva, J. Agric. Food Chem. 51, [57] C. Lee, W. Yang and R.G. Parr, Phys. Rev. B. 37 (2), 785
6475–6481 (2003). (1988).
[30] I. Ivan, E. Jungfer, C. Ritter, B. Santiago-schübel, B. Thiele, [58] L. Goerigk, A. Hansen, C. Bauer, S. Ehrlich, A. Najibi and
R. Fett and R. Galensa, Food Res. Int. 48, 848–855 S. Grimme, Phys. Chem. Chem. Phys. 19, 32184–32215
(2012). (2017).
[31] A. María, M. Wilson, S.I. Castro-arredondo and R.R. [59] N. Mardirossian and M. Head-gordon, Mol. Phys. 115
Balandrán-quintana, Food Chem. 161, 155–161 (2014). (19), 2315–2372 (2017).
[32] L. Wang and H. Zhang, Bioorg. Chem. 33, 108–115 [60] G. Wang, Y. Xue, L. An, Y. Zheng, Y. Dou, L. Zhang and Y.
(2005). Liu, Food Chem. 171, 89–97 (2015).
[33] S. Antonczak, J. Mol. Struc. Theochem. 856, 38–45 [61] M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria,
(2008). M.A. Robb, J.R. Cheeseman, G. Scalmani, V. Barone,
[34] M. Leopoldini, N. Russo and M. Toscano, Food Chem. B. Mennucci, G.A. Petersson, H. Nakatsuji, M. Cari-
125, 288–306 (2011). cato, X. Li, H.P. Hratchian, A.F. Izmaylov, J. Bloino, G.
[35] E.N. Bentz, A.B. Pomilio and R.M. Lobayan, Comp. Theo. Zheng, J.L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R.
Chem. 1110, 14–24 (2017). Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda,
[36] D. Zhang, L. Chu, Y. Liu, A. Wang, B. Ji, W. Wu, F. Zhou, O. Kitao, H. Nakai, T. Vreven, J.A. Montgomery, Jr., J.E.
Y. Wei, Q. Cheng, S. Cai, L. Xie and G. Jia, J. Agric. Food Peralta, F. Ogliaro, M. Bearpark, J.J. Heyd, E. Brothers,
Chem. 59, 10277–10285 (2011). K.N. Kudin, V.N. Staroverov, R. Kobayashi, J. Normand,
[37] I. Nakanishi, K. Miyazaki, T. Shimada, K. Ohkubo, S. K. Raghavachari, A. Rendell, J.C. Burant, S.S. Iyengar,
Urano, N. Ikota, T. Ozawa, S. Fukuzumi and K. Fukuhara, J. Tomasi, M. Cossi, N. Rega, J.M. Millam, M. Klene,
Phys. Chem. A. 106, 11123–11126 (2002). J.E. Knox, J.B. Cross, V. Bakken, C. Adamo, J. Jaramillo,
[38] W. Li, H.H.S. Fong, K.W. Singletary and J.F. Fitzloff, J. Liq. R. Gomperts, R.E. Stratmann, O. Yazyev, A.J. Austin,
Chrom. Rel. Technol. 25 (3), 397–407 (2002). R. Cammi, C. Pomelli, J.W. Ochterski, R.L. Martin, K.
[39] Y. Katsuda, Y. Niwano, T. Nakashima, T. Mokudai, K. Morokuma, V.G. Zakrzewski, G.A. Voth, P. Salvador, J.J.
Nakamura, S. Oizumi, T. Kanno, H. Kanetaka and H. Dannenberg, S. Dapprich, A.D. Daniels, O. Farkas, J.B.
Egusa, Plos One. 10 (8), (2015). doi:10.1371/journal.pone. Foresman, J.V. Ortiz, J. Cioslowski and D.J. Fox, Gaussian
0134704. 09, Revision B. 01 (Gaussian, Inc., Wallingford, CT, 2009).
[40] A.M. Tarola, F. Milano and V. Giannetti, Anal. Lett. 40, [62] S. Marković and J. Tošović, Food Chem. 210, 585–592
2433–2445 (2007). (2016).
[41] I.I. Rockenbach, L.V. Gonzaga, V.M. Rizelio, A.E.d.S.S. [63] S.M. Reza Nazifi, M.H. Asgharshamsi, M.M. Dehkordi
Gonçalves and M.I. Genovese, R. Fett, Food Res. Int. 44, and K.K. Zborowski, Free Radic. Res. 53 (8), 922–931
897–901 (2011). (2019).
[42] P. Doshi, P. Adsule, K. Banerjee and D. Oulkar, J. Food Sci. [64] Y. Xue, Y. Zheng, L. An, Y. Dou and Y. Liu, Food Chem.
Technol. 52 (1), 181–190 (2015). 151, 198–206 (2014).
[43] D.P. Makris, G. Boskou and N.K. Andrikopoulos, J. [65] A. Urbaniak, M. Szela and M. Molski, Comp. Theo. Chem.
Comp. Anal. 20, 125–132 (2007). 1012, 33–40 (2013).
[44] A.M. Mendoza-Wilson and D. Glossman-mitnik, J. Mol. [66] J. Rimarčík, V. Lukeš, E. Klein and M. Ilčin, J. Mol. Struc.
Struc. Theochem. 761, 97–106 (2006). Theochem. 952, 25–30 (2010).
[45] M. Leopoldini, N. Russo and M. Toscano, J. Agric. Food [67] B. Badhani and R. Kakkar, Struct. Chem. 29, 359–373
Chem. 55, 7944–7949 (2007). (2017).
[46] M. Leopoldini, T. Marino, N. Russo and M. Toscano, J. [68] K. Sadasivam and R. Kumaresan, Mol. Phys. 109 (6),
Phys. Chem. A. 108, 4916–4922 (2004). 839–852 (2011).
12 S. ANITHA ET AL.

[69] N.M. O’Boyle, A.L. Tenderholt and K.M. Langner, J. [74] R. Praveena, K. Sadasivam, R. Kumaresan, V. Deepha and
Comp. Chem. 29, 839–845 (2008). R. Sivakumar, Spectrochim. Acta Part A Mol. Biomol.
[70] M.A. Thompson, ArgusLab 4.0.1 (Planaria Softwae LLC, Spectrosc. 103, 442–452 (2013).
Seattle, 2004). < http://www.ArgusLab.com > . [75] Y. Shang, H. Zhou, X. Li, J. Zhou and K. Chen, New J.
[71] Y. Zheng, Y. Zhou, Q. Liang, D. Chen, R. Guo, C. Xiong, Chem. 43 (39), 15736–15742 (2019).
X. Xu, Z. Zhang and Z. Huang, Dyes Pigm. 141, 179–187 [76] A.M. Mendoza-Wilson and D. Glossman-mitnik, J. Mol.
(2017). Struc. Theochem. 716, 67–72 (2005).
[72] Y. Rong, Z. Wang, J. Wu and B. Zhao, Spectrochim. Acta [77] R. Praveena, K. Sadasivam, V. Deepha and R. Sivakumar,
Part A Mol. Biomol. Spectrosc. 93, 235–239 (2012). J. Mol. Struc. 1061, 114–123 (2014).
[73] A. Vagánek, J. Rimarčík, V. Lukeš and E. Klein, Comput.
Theor. Chem. 991, 192–200 (2012).

You might also like