You are on page 1of 16

APPLICATION OF COMPUTATIONAL CHEMISTRY METHOD TO STUDY THE

RELATIVE ANTIOXIDANT EFFICIENCY OF A LARGE SERIES OF


CAROTENOIDS

Zaky Al-Fatony 1, Hanggara Sudrajat 2,*, Bongkot Pipoosananakaton 3,


Sudarath Veravong 3, and Nathaya Selphusit 3,4
1
Department of Chemistry, Faculty of Mathematics and Natural Sciences,
Gadjah Mada University, Bulaksumur Yogyakarta 55281, Indonesia
2
Austrian-Indonesian Centre for Computational Chemistry, Department of Chemistry, Faculty
of Mathematics and Natural Sciences, Gadjah Mada University,
Bulaksumur Yogyakarta 55281, Indonesia
3
Department of Chemistry, Faculty of Sciences,
Chulalongkorn University, Bangkok 10330, Thailand
4
Austrian-Thai Centre for Computational Chemistry, Department of Chemistry, Faculty of
Sciences, Chulalongkorn University, Bangkok 10330, Thailand

*Corresponding author. Tel : +6285868077123 ;


E-mail: angga_vanniveau@yahoo.co.id

ABSTRACT
The relative antioxidant efficiency, expressed as electron donating capability, of a large
series of carotenoids has been studied using computational chemistry method at the hybrid
density functional theory with the B3LYP functional and 6-31G(d,p) basis set. Their reactivity
toward nine different radicals has been modeled as well as the electron transfer between pairs
of carotenoids, one of which is present as a radical cation. The influence of the solvent
polarity has also been studied. Torulene was found to be the most easily oxidized carotenoid,
followed by lycopene. This higher reactivity is proposed in the present work for the first time,
and the potential implications of such a finding are discussed. Since torulene has not been
previously studied, compared to other carotenoids in terms of oxidation potentials, further
experimental studies are suggested in order to confirm or reject this prediction. Ionization
potential seems to be a magnitude calculable at low computational cost that correctly predicts
the relative ease of oxidation in a series of carotenoids.

Keywords: computational chemistry method, carotenoids, electron donating capability,


antioxidant efficiency

INTRODUCTION

Carotenoids (Car) are versatile organic molecules, which are of great relevance to living

systems. They serve multiple purposes for their producers as well as for their consumers.

Among their diverse functions [1], their free-radical-scavenging antioxidant properties stand

out [2,3]. However, their broadly defined antioxidant activity strongly depends on a mixture

of variables. Consequently, attempts to uniformly characterize antioxidant activity may be a


futile task. On the other hand, testing antioxidant systems for a specific function may provide

valuable information for a particular application. This work will focus on the electron-transfer

(ET) processes, although hydrogen abstraction and adduct formation are also viable

mechanisms of carotenoids’ antioxidant activity.

Electron transfer: R•+ Car → R-+ Car•+ (I)

Adduct formation: R• + Car → [R-Car]• (II)

Hydrogen abstraction: R• + Car → RH + Car(-H)• (III)

The relative importance of these reaction channels will depend on diverse factors,

including the nature of the reacting free radical and the structural features of the carotenoid

[4,5] and, in biological systems, its location and orientation within the membrane [5].

Generally speaking, the antioxidant properties of carotenoids are related to their excellent

ability to deactivate excited states and to their high electron donation capability. Although

electron donation can immediately deactivate harmful radicals, the antioxidant mechanism

can be inseparably accompanied by the simultaneous formation of oxidized carotenoids such

as Car•+. In addition, redox pairs involving Car/Car •+ conjugates [6] could regenerate the

parent molecule and eliminate the pro-oxidant concerns related to carotenoid radical cations.

The electron-transfer reaction (I) will clearly be favored when the R groups are electron

withdrawing. For example, the reaction of the trichloromethyl peroxyl radical (CCl 3O2) with

β-carotene [7] and several other carotenoids [8] has been suggested to involve electron

transfer to generate the corresponding anion and carotene radical cation. Everett and co-

workers [9] have also shown that β-carotene is oxidized by NO2• in solution via reaction I. In

a recent study using benzyl peroxyl radicals, it was concluded that carotenoids scavenge

peroxyl radicals, which are not highly reactive, by adduct formation and not by electron

transfer, while reaction III seems to be much less important than reaction II [10]. These results

are in line with a previous review article showing that carotenoids react with free radicals by
addition and/or electron-transfer reactions, with the reaction channel distribution varying

according to the nature of the reacting free radical [5].

On the other hand, it seems obvious that there must be a relationship between the structure

of carotenoids and their reactivity toward free radicals. There are some reports addressing this

issue. Jeevarajan et al. [11] found that carotenoids substituted with electron donating groups

are more easily oxidized than those with electron accepting substituents. Mortensen and

Skibsted [12] have studied the reactions of eight carotenoids with phenoxyl radicals. These

authors report the following order of reactivity: lycopene (LYC) > β-carotene (BC) >

zeaxanthin (ZEA) > lutein (LUT) > echinenone (ECH). They also found that canthaxanthin

(CAN) and β-apo-8'-carotenal (APO) hardly react, while asthaxanthin (ASTA) does not react

at all. Edge et al. [13] have studied ET reactions between different pairs of carotenoids with

similar outcomes. They found that LYC is the most easily oxidized followed by BC > ZEA >

LUT > CAN > APO > ASTA.

In addition, El-Agamey and McGarvey [14] have recently shown that the profile of

carotenoid radical products formed depends on the polarity of the solvent medium. According

to these authors, in nonpolar solvents, only addition radicals are formed, while in polar

solvents, these adducts decay to carotenoid radical cations.

The main aim of this work is to provide a computational strategy that allows reliable

predictions of carotenoids’ relative antioxidant efficiency, expressed as electron donating

capability, at a reasonable computational cost. The studied species are hydrocarbon

carotenoids and oxygen carotenoids. Their reactivity toward nine different radicals has been

modeled through reaction I, as well as the electron transfer between pairs of carotenoids, one

of which is present as a radical cation. The influence of the solvent polarity has also been

investigated.
COMPUTATIONAL METHODS

For all of the modeled species, full geometry optimizations have been carried out using

the B3LYP hybrid density functional and the 6-31G(d,p) basis set, while frequency

calculations were performed at B3LYP/3-21G level of theory for geometries previously

optimized at the same level. Unrestricted calculations were used for open-shell systems, and

local minima were identified by the number of imaginary frequencies (NIMAG = 0). All of

the electronic calculations were performed with the Gaussian 98 suite of programs.

Thermodynamic corrections at 298 K were included in the calculation of the relative energies.

Solvent effects were included by using the polarizable continuum model (PCM) with water

and benzene as the solvents for polar and nonpolar environments, respectively. The Gibbs free

energies of reaction for an aqueous environment were calculated using a mixed discrete-

continuum model with two water molecules explicitly included, with a surrounding

continuum of bulk solvent for carotenoids + free radical reactions. The corresponding

energies were improved by single point (SP) calculations at the B3LYP/6-311++G(d,p) level

of theory.

RESULTS AND DISCUSSION

Since there is previous knowledge about the relative antioxidant activity of seven

carotenoids [12], they have been included in the present work in order to test the reliability of

the results proposed here for the whole studied series. The structures of all modeled

carotenoids are shown in Tables 1 and 2. The simplest approach to study the ease of ET

processes from different molecules is to analyze their ionization potentials (IP). The IP of an

n-electron system (X), calculated at a given level of theory, implies the following energy

difference

IP = EXn-1(gn-1) – EXn(gn) (1)


where EXn(gn) is the energy of the n-electron neutral system calculated at a geometry gn and

EXn-1(gn-1) is the energy of the (n - 1)-electron ionic species calculated at a geometry gn-1.

Accordingly, to calculate IPs, two separate geometry optimizations are needed, performed

on the neutral and radical cationic species. The IP calculated this way is known as an adiabatic

IP and takes into account the geometry relaxation of the radical

cation after the electron transfer.

Another approximation, known as the vertical IP, is to optimize the geometry of the X

species and then to perform the calculation of the radical cation at this very geometry. This

approach implies also two calculations but only one geometry optimization and one single

point to obtain the electronic energy of the Xn-1 species. A special case of a vertical IP is that

obtained within the framework of the Koopman’s theorem approximation. In this particular

case, the vertical IP is evaluated as minus the energy of the molecular orbital of the neutral

system from which an electron is removed, usually the highest occupied molecular orbital

(HOMO), provided that the potential is defined in such a way that it vanishes at infinity. This

is a very simple, relatively inexpensive way to estimate the IP from a single calculation on the

neutral system.

Even though the fast development of computers in the last few decades has remarkably

enlarged the potential use of computational quantum chemistry for modeling molecular

systems of increasing size at reliable levels of calculation, carotenoids can still be considered

as large molecules from a computational point of view. Accordingly, it would be of interest to

find some parameters that describe their relative ease of electron transfer in a proper way, at

least qualitatively, and that, at the same time, could be calculated at a reasonable

computational cost. With that purpose in mind, the correlations of the adiabatic IP versus the

Koopman IP and the vertical IP versus the Koopman IP have been tested for all of the studied

carotenoids in polar and nonpolar media. As Figure 1 shows, there is a good correlation in all
of the tested cases. However, the agreement of Koopman’s IP with the more accurate ones is

better in polar solvents. According to these results, Koopman’s IP can be used to qualitatively

compare the ease of electron donation capability among a series of similar compounds. In the

specific case of carotenoids, it can be used as the most economical criteria (computationally

speaking) to predict their relative ease of oxidation.

As the values in Table 3 show, the ionization potentials calculated within the vertical and

adiabatic approaches are lower for polar than those for nonpolar media, as it is expected since

a charged species is formed. In general, this effect was found to be less important for

hydrocarbon carotenoids than that for oxygen carotenoids. These results agree with those

reported by Sliwka et al. [15] who found that electron-transfer processes are enhanced in the

presence of water, especially for hydrophilic carotenoids. Koopman’s IPs do not show this

experimentally validated tendency; therefore, they are good enough to qualitatively compare

the ease in electron donation for a series of carotenoids, but they do not properly describe the

influence of the solvent in such process, at least when the continuum PCM model is used.

Among the different approaches used to calculate ionization potentials, special attention

should be paid to that referred to as the vertical IP. The ET reactions can be assumed to

mainly occur by electron tunneling from one species to another. Since electrons are much

faster than nuclei, the electronic transitions can be considered to take place in fixed nuclear

configurations. This process is expected to occur very fast, in such a way that at the moment

of the transfer, there is no time for geometry relaxation. Accordingly, while adiabatic IPs can

be related to the energy involved in the completion of the reaction, vertical IPs can be

associated with the energetic barrier. Vertical IPs can then be considered as an upper limit of

the activation barrier. According to these assumptions, vertical IPs would be better criteria to

predict which carotenoid would more readily transfer one of its electrons to another species

and then to relatively order a series of carotenoids in terms of their oxidation potentials. To
confirm this hypothesis, it is necessary to compare vertical and adiabatic IPs with the

available experimental data, which indicates the following order of reactivity in terms of the

ease of oxidation: LYC > BC > ZEA > LUT > ECH > CAN > ASTA [12] and LYC > BC >

ZEA > LUT > CAN > APO > ASTA [13]. For this particular sets of carotenoids, their relative

order, based on vertical IPs, is LYC > BC > ZEA > LUT > ECH > CAN > APO > ASTA,

which is in perfect agreement with the experimental findings. For this subset of carotenoids,

the order is the same in both solvents. On the other hand, the order based on adiabatic IPs is

LYC > BC > ZEA > LUT > ECH > APO > CAN > ASTA, which inverts the relative reactivity

of canthaxanthin and β-apo-8'-carotenal. Consequently, for the possible interactions between

pairs of carotenoids, their relative order, in term of oxidation potential, is shown in Figure 2

based on vertical IPs. This is a schematic representation, standing only for the order of the

ease of electron transfer. However, for some of these pairs, the difference may be small,

leading to slow reactions (e.g., DIH/ECH pair). The relative ease of electron transfer between

carotenoid pairs has been provided in polar and nonpolar media since these reactions can be

studied independently of their hydrophilicity only under special circumstances [12, 13].

However, in living systems, carotenoids are not likely to always react either in pure lipid or

water-based environments but at the typical hydrophilic/hydrophobic interphase. Accordingly,

a better knowledge of their behavior in such extreme environments, which is hard to obtain

from experiments, could be helpful since the physicochemical assessment of the carotenoid

radical scavenging behavior in aqueous solutions might be a key property in the formulation

of parental therapeutics.

Torulene was found as the most easily oxidized carotenoid in polar and nonpolar

environments. This can be a relevant finding since it implies that TOR is expected to repair

other damaged carotenoids through redox reactions between pairs Car2/Car1•+, regardless of

the environment’s polarity.


Since such repairing processes are relevant to the combined antioxidant activity of

carotenoids, a more detailed study has been performed. The energy evolution associated with

the completion of ET reactions between carotenoid conjugated pairs has been computed as the

corresponding adiabatic Gibbs free energies of such processes at 298 K

ΔG0ET = G(Car1•+, gn-1) + G(Car2, gn) - G(Car1, gn) - G(Car2•+, gn-1) (2)

where gn represents the relaxed geometry of the neutral carotenoid and gn-1 that of the radical

cation.

BC, LYC, and TOR have been modeled as the repairing carotenoids (Car 2). They have

been chosen based on the following criteria: BC is the most representative carotenoid of the

series; LYC has been predicted as the Car most easily oxidized from all previously studied

Car; and TOR has not been comparatively studied before, but according to our results, it

oxidizes even more easily than LYC. The values in Table 4

show that the processes involving electron transfer from neutral TOR to any of the other

studied radical cations are energetically favored in terms of Gibbs free energies, that is, they

are exergonic processes (ΔG < 0). Electron transfers from LYC are mostly exergonic, with the

logical exception of its reaction with TOR. When the electron transfer is modeled from BC,

the calculated ΔGs predict it as able to repair other Car•+ species than TOR, LYC, SAP, and

ZEA.

The possible electron-transfer reactions in nonpolar and polar environments have also

been modeled for carotenoids in Tables 1 and 2 and for the following radicals: hydrogen

peroxyl: HOO• (R1), methyl peroxyl: CH3OO• (R2), methoxyl: CH3O• (R3) benzyl peroxyl:

C6H5CH2OO• (R4), phenoxyl: C6H5O• (R5), acyl peroxyl: CH3C(O)OO• (R6), phenylacetyl

peroxyl: C6H5C(O)OO• (R7), trichloromethyl peroxyl: CCl3OO (R8), nitrite: NO2• (R9).

The adiabatic Gibbs free energies of reaction have been calculated as

ΔG0ET = G(Car1•+, gn-1) + G(Ox-, gn+1) - G(Car, gn) - G(Ox•, gn) (3)
where Ox represent the oxidant species, that is, free radicals. The values obtained that way for

nonpolar media were found to be highly endergonic (ΔG from 25 to 80 kcal/mol) and,

accordingly, very unlikely to occur. On the other hand, the presence of polar solvent

drastically increased the reactivity of carotenoids toward the studied free radicals, through

electrontransfer reactions. Accordingly, it can be stated that in nonpolar media, the studied

radicals and probably most oxygenated radicals are bound to react by adduct formation

instead of electron transfer since the ionic products cannot be solvated, whereas in polar

media, carotenoid radical cations and the corresponding anions can be stabilized by the

surrounding molecules. In biological systems, these charged species could be solvated at the

lipid-water interface.

The carotenoids’ reactions with HOO•, CH3OO•, and C6H5-CH2OO• remain endergonic in

aqueous phase but with ΔG0ET more than 50 kcal/mol lower than those corresponding to

benzene as the solvent. On the other hand, the reactions of CH 3O•, C6H5O•, CH3C(O)OO•,

C6H5C(O)OO•, CCl3OO•, and NO2• become exergonic (Table 5). Accordingly, among the

peroxyl radicals, only those with electron withdrawing groups are able to react with

carotenoids, via electron-transfer reaction. Thus, peroxyl radicals with electron donating

groups are expected to react via adduct formation. All of these findings nicely agree with the

available experimental results. The most energetically favored of the modeled reactions are

those involving trichloromethyl peroxyl, nitro, acyl peroxyl, and phenylacetyl peroxyl

radicals, in that order.

The calculations of the Gibbs free energy of electron transfers from carotenoids to free

radicals are relatively expensive since frequency calculations of the corresponding Car, Car•+,

oxygenated free radical, and its anion are necessary in order to add thermodynamic

corrections to the electronic energies. Since they are rather time-consuming calculations, it

would be useful to find another magnitude that could be calculated at a lower computational
cost and that retains the correct order of relative reactivity. Ionization potentials have been

tested for that purpose. Figure 3 shows the correlations of the ΔG of reaction and the

carotenoids’ ionization potentials calculated using the different approaches discussed above

for their reaction with the acyl peroxyl radical. Even though, for the sake of simplicity, their

reactions with the other studied radicals are not shown, the same tendency was obtained in all

of the cases. Regardless of the approach chosen for computing the ionization potentials, good

correlations were found with the Gibbs energies of reaction. However, and as it was expected,

the best correlation was obtained for the adiabatic IP.

The only carotenoid that seems to have higher antioxidant efficiency than LYC, expressed

as electron donating capability, is TOR. This seems to be the case in both polar and nonpolar

media and in Car + free radical reactions as well as in redox reactions between carotenoid

pairs. This new finding suggests that this Car could be of great use for repairing other

oxidation damaged carotenoids and can be easily explained in terms of the length of the

polyene chain. TOR has 13 conjugated bonds, that is, the largest number among all of the

modeled structures, while LYC and BC have 11, but in the case of BC, two of them are in

cyclohexane rings, which are nonplanar with the rest of the molecular backbone, reducing the

effective length of the conjugated system. The larger reactivity of BC compared to that of

ZEA, however, cannot be explained by the length of the conjugated chain but by the presence

of two OH’s, which comparatively reduces the stability of the carotenoid radical cation. In the

case of ASTA and CAN, the presence of the two carbonyl groups at the end of the conjugated

systems also seems to lower the stability of the products.

CONCLUSIONS

The relative order of 19 different carotenoids, in terms of oxidation potentials, is proposed

for polar and nonpolar solvents. The same order was obtained for Car + oxygenated free
radical and Car2 + Car1 •+ reactions. Ionization potentials seem to be capable of predicting the

correct relative ease of oxidation in a series of carotenoids at a low computational cost.

The nuclear reorganization energy associated with ET reactions has been calculated in a

very simple but apparently efficient way that allows computing of free energy barriers and, at

least, relative rate coefficients in good agreement with the experimental values. TOR is

predicted as the carotenoid with the highest antioxidant efficiency from the modeled set,

expressed as the electron donating capability, in polar and nonpolar media and in Car + free

radical reactions as well as in carotenoid pair redox reactions. The general agreement between

different calculated magnitudes and the corresponding available experimental data supports

the predictions from this work. Since TOR was found to be the most easily oxidized from all

of the studied carotenoids, experimental studies on the antioxidant capabilities of this

carotenoid, relative to other carotenoids would be of great interest, especially relative to LYC,

which up to date has been thought of as the most efficient one. Such experimental studies

might also help to validate the accuracy of the methodology used in this work for describing

electron-transfer reactions. They might also support the utility of this kind of calculations in

the design of new antioxidants with increasing efficiency in terms of electron donating

capabilities, which might potentially help to prevent or repair oxidative damage in living

systems.

REFERENCES

1. van den Berg, H., Faulks, R., Fernando Granado, H., Hirschberg, J., Olmedilla, B.,
Sandmann, G., Southon, S. and Stahl, W., 2000, J. Sci. Food Agric., 80, 880.
2. Krinsky, N. I., 1993, Annu. Rev. Nutr., 13, 561.
3. Edge, R., McGarvey, D. J. and Truscott, T., 1997, J. Photochem. Photobiol., 41, 189.
4. Everett, S. A., Dennis, M. F., Patel, K. B., Maddix, S., Kundu, S. C. and Willson, R. L.,
1996, J. Biol. Chem., 271, 3988.
5. Woodall, A. A., Britton, G. and Jackson, M., 1997, J. Biochim. Biophys. Acta, 1336, 575.
6. Larsen, E., Abendroth, J., Partali, V., Schulz, B., Sliwka, H. R. and Quartey, E. G. K.,
1998, Chem. Eur. J., 4, 113.
7. Packer, J. E., Mahood, J. S., Mora-Arellano, V. O., Slater, T. F., Willson, R. L. and
Wolfenden, R. S., 1981, Biochem. Biophys. Res. Commun., 98, 901.
8. Hill, T. J., Land, E. J., McGarvey, D. J., Schalch, W., Tinkler, J. H. and Truscott, T. G.,
1995, J. Am. Chem. Soc., 117, 8322.
9. Everett, S. A., Kundu, S. C., Maddix, S. and Willson, R. L., 1995, Biochem. Soc. Trans.,
23, 230.
10. Mortensen, A., 2002, Free Radical Res., 36, 211
11. Jeevarajan, A. S., Khaled, M. and Kispert, L. D., 1994, J. Phys. Chem., 98, 1111.
12. Mortensen, A. and Skibsted, L. H., 1997, J. Agric. Food Chem., 45, 2970.
13. Edge, R., Land, E. J., McGarvey, D., Mulroy, L. and Truscott, T. G., 1998, J. Am. Chem.
Soc., 120, 4087.
14. El-Agamey, A. and McGarvey, D. J., 2003, J. Am. Chem. Soc., 125, 3330.
15 Sliwka, H. R., Melø, T-B., Foss, B. J., Abdel-Hafez, S. H., Partali, V., Nadolski, G.,
Jackson, H. and Lockwood, S. F., 2007, Chem.Eur. J., 13, 4458.

Figure 2. Relative order of carotenoids in terms of oxidation potentials in nonpolar and polar solvents; ■ ref 12,
□ ref 13

Figure 1. Correlation between vertical and Figure 3. Correlations between the Gibbs free
adiabatic ionization potentials with Koopman’s IP, energy of electrontransfer reactions (Car + R7) and
for polar and nonpolar media. carotenoids’ IP calculated using different
approaches.
Table 1. Carotenoids Previously Studied by Experimental Techniques and Modeled in the Present
Work
Symbol Formula Structure and Name
O

OH

ASTA C40H52O4
HO

Astaxanthin

CHO

APO C30H40O

β-apo-8'-carotenal
O

CAN C40H52O2

Canthaxanthin
OH

LUT C40H56O2
HO

Lutein
OH

ZEA C40H52O2
HO

Zeaxanthin

BC C40H56

All-trans-β-carotene

LYC C40H56
Lycopene
Table 2. Carotenoids Modeled in the Present Work and with No Previous Data on Their Electron-
Transfer Reactions
Symbol Formula Structure and Name

DIH C40H58

7,7' Dihydro-β-carotene
O
Table 2.

ECH C40H54O

Echinenone
OH

HO

NOS C40H56O4 OH

HO

Nostoxanthin
OCH3

OKE C41H54O2
O

Okenone

SAP C40H56O2 OH
HO

Saproxanthin

TOR C40H54
Torulene
OH

CSAN C40H56O3
O

HO

Capsanthin
OH

CRUB C40H56O4 O

OH

Capsorubin
(Continued)
Symbol Formula Structure and Name
OH

CRY C40H56O2
O

Cryptocapsin
HO

NEO C40H56O4 O

HO OH

Neoxanthin
O

VIO C40H56O4 O
OH

HO

Violaxanthin
OH O

MIT C40H52O4
HO O

Mytiloxanthinone

Table 3. Ionization Potentials (eV) in Nonpolar and Polar Environments


solvent = benzene solvent = water
Koopman vertical adiabatic Koopman vertical adiabatic
ASTA 4.808 5.165 5.038 4.701 4.559 4.438
APO 4.676 5.083 4.914 4.687 4.553 4.406
CAN 4.707 5.069 4.942 4.658 4.518 4.412
LUT 4.493 4.882 4.736 4.495 4.366 4.243
ZEA 4.458 4.836 4.681 4.474 4.346 4.213
BC 4.414 4.790 4.629 4.455 4.328 4.186
LYC 4.321 4.672 4.522 4.388 4.251 4.105
DIH 4.491 4.938 4.779 4.548 4.425 4.271
ECH 4.563 4.922 4.770 4.550 4.415 4.282
NOS 4.472 4.849 4.680 4.489 4.360 4.206
OKE 4.593 4.937 4.790 4.572 4.431 4.300
SAP 4.375 4.721 4.577 4.408 4.276 4.158
TOR 4.313 4.641 4.495 4.361 4.230 4.094
CSAN 4.685 5.064 4.902 4.646 4.511 4.368
CRUB 4.886 5.269 5.142 4.850 4.706 4.600
CRY 4.625 5.003 4.853 4.627 4.491 4.364
NEO 4.444 4.842 4.689 4.509 4.380 4.242
VIO 4.516 4.917 4.778 4.558 4.429 4.310
MYT 4.741 5.124 4.990 4.723 4.590 4.471
Table 4. Gibbs Free Energies of Reaction (kcal/mol) for Electron Transfer between Pairs of
Carotenoids with Respect to BC, LYC, and TOR
solvent = benzene solvent = water
BC LYC TOR BC LYC TOR
ASTA -8.21 -11.65 -12.50 -4.57 -7.42 -7.91
APO -5.32 -8.76 -9.61 -3.81 -6.66 -7.15
CAN -6.17 -9.60 -10.46 -4.16 -7.00 -7.49
LUT -2.35 -5.79 -6.64 -1.20 -4.05 -4.54
ZEA 0.08 -3.36 -4.21 0.65 -2.20 -2.68
BC -3.44 -4.29 -2.85 -3.34
LYC 3.44 -0.85 2.85 -0.49
DIH -2.60 -6.04 -6.89 -1.08 -3.93 -4.41
ECH -1.87 -5.31 -6.16 -0.82 -3.66 -4.15
NOS -1.18 -4.62 -5.47 -0.46 -3.30 -3.79
OKE -2.50 -5.94 -6.79 -1.42 -4.27 -4.76
SAP 1.64 -1.79 -2.65 1.10 -1.75 -2.24
TOR 4.29 0.85 3.34 0.49
CSAN -6.08 -9.52 -10.37 -4.00 -6.85 -7.33
CRUB -11.31 -14.74 -15.60 -9.03 -11.87 -12.35
CRY -4.97 -8.41 -9.26 -3.90 -6.75 -7.24
NEO -0.18 -3.62 -4.47 -0.09 -2.93 -3.42
VIO -2.42 -5.85 -6.71 -1.84 -4.69 -5.18
MYT -6.83 -10.26 -11.12 -5.06 -7.91 -8.40

Table 5. Gibbs Free Energies (kcal/mol)a of Carotenoidsa Electron-Transfer Reactions with Different Radicals in
a Polar Environment (Solvent = Water)
R1 R2 R3 R4 R5 R6 R7 R8 R9
ASTA 9.26 1.55 -10.84 4.11 2.71 -14.46 -11.03 -20.69 -14.50
APO 8.73 1.02 -11.37 3.58 2.18 -14.99 -11.56 -21.22 -15.03
CAN 9.02 1.31 -11.08 3.86 2.46 -14.71 -11.27 -20.93 -14.75
LUT 5.78 -1.92 -14.32 0.63 -0.77 -17.94 -14.51 -24.17 -17.98
ZEA 3.96 -3.75 -16.14 -1.20 -2.59 -19.77 -16.33 -25.99 -19.81
BC 4.48 -3.23 -15.62 -0.67 -2.07 -19.24 -15.81 -25.47 -19.28
LYC 2.07 -5.64 -18.03 -3.08 -4.48 -21.65 -18.22 -27.88 -21.69
DIH 5.59 -2.11 -14.51 0.44 -0.96 -18.13 -14.70 -24.36 -18.17
ECH 5.81 -1.90 -14.29 0.66 -0.74 -17.91 -14.48 -24.14 -17.95
NOS 5.69 -2.01 -14.40 0.54 -0.86 -18.03 -14.60 -24.25 -18.07
OKE 6.31 -1.40 -13.79 1.16 -0.24 -17.41 -13.98 -23.64 -17.45
SAP 3.96 -3.75 -16.14 -1.20 -2.59 -19.77 -16.33 -25.99 -19.81
TOR 1.52 -6.19 -18.58 -3.64 -5.03 -22.21 -18.77 -28.43 -22.25
CSAN 8.84 1.13 -11.26 3.69 2.29 -14.88 -11.45 -21.11 -14.92
CRUB 14.03 6.32 -6.07 8.87 7.47 -9.70 -6.26 -15.92 -9.74
CRY 8.42 0.71 -11.68 3.26 1.87 -15.30 -11.87 -21.53 -15.34
NEO 4.98 -2.73 -15.12 -0.18 -1.58 -18.75 -15.32 -24.97 -18.79
VIO 6.38 -1.33 -13.72 1.23 -0.17 -17.34 -13.91 -23.57 -17.38
MYT 9.92 2.21 -10.18 4.76 3.37 13.80 --10.37 -20.03 -13.84
a
Energies: SP/B3LYP/6-311++G(d,p); geometries: B3LYP/6-31G(d,p); thermodynamic corrections: B3LYP/3-21G.

You might also like