Next Article in Journal
Development of In-Needle SPME Devices for Microextraction Applied to the Quantification of Pesticides in Agricultural Water
Previous Article in Journal
Improvement of Bioactive Polyphenol Accumulation in Callus of Salvia atropatana Bunge
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Computational Chemistry Strategies to Investigate the Antioxidant Activity of Flavonoids—An Overview

1
Department of Pharmacology and Personalized Medicine, School of Nutrition and Translational Research in Metabolism (NUTRIM), Maastricht University, 6200 MD Maastricht, The Netherlands
2
School of Pharmacy, Chengdu University, 2025 Chengluo Avenue, Chengdu 610106, China
3
Research Centre Material Sciences, Zuyd University of Applied Science, 6400 AN Heerlen, The Netherlands
4
Hainan University-HSF/LWL Collaborative Innovation Laboratory, College of Food Sciences & Engineering, Hainan University, 58 People Road, Haikou 570228, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2024, 29(11), 2627; https://doi.org/10.3390/molecules29112627
Submission received: 28 April 2024 / Revised: 27 May 2024 / Accepted: 29 May 2024 / Published: 3 June 2024

Abstract

:
Despite several decades of research, the beneficial effect of flavonoids on health is still enigmatic. Here, we focus on the antioxidant effect of flavonoids, which is elementary to their biological activity. A relatively new strategy for obtaining a more accurate understanding of this effect is to leverage computational chemistry. This review systematically presents various computational chemistry indicators employed over the past five years to investigate the antioxidant activity of flavonoids. We categorize these strategies into five aspects: electronic structure analysis, thermodynamic analysis, kinetic analysis, interaction analysis, and bioavailability analysis. The principles, characteristics, and limitations of these methods are discussed, along with current trends.

1. Introduction

1.1. The Enigma of Flavonoids’ Antioxidant Activity

A major desire is to be and stay in good health. This desire has fueled our interest in compounds that have a positive impact on the body. However, the limitations and negative aspects of synthetic drugs are becoming increasingly apparent. This has revived our interest in natural compounds present in our diet that can promote health. A group of natural compounds gaining significant attention are flavonoids. A high dietary intake of flavonoids has been associated with positive effects on health, such as a lower incidence of cardiovascular diseases and cancer [1,2,3]. Based on the paradigm coined by Brown and Fraser in 1868, we assume that the fundament for this beneficial effect lies in the interaction of flavonoid molecules with molecules and processes in our bodies [4].
Surprisingly, after decades of research, the exact molecular mechanism responsible for the antioxidant effect of flavonoids is still enigmatic [5,6]. First of all, this is because measuring antioxidant activity appears to be less straightforward than commonly assumed. Complicating factors affecting the outcome of antioxidant assays are as follows:
  • The myriad of reactive species used in the assays varies from ‘simple’ highly reactive oxygen radicals to complex relatively stable synthetic radicals.
  • The degradation of the unstable compounds during the experiments.
  • Compartmentation, e.g., of lipophilic compounds, in cell culture experiments.
Regarding the latter, it is usually tacitly assumed that in cell experiments, the concentration in all compartments is equal to the initial concentration of the compound in the culture medium, which of course is not the case. Moreover, the existence of the different types of antioxidant assays, along with the procedural differences within specific assays, blur our view on the antioxidant activity of flavonoids.
In addition, the group of flavonoids comprises more than 5000 compounds. In line with the paradigm of Brown and Fraser, we know that a relatively small change in chemical structure can lead to a drastic impact on biological activity mechanism. Nevertheless, often all flavonoids are put under the same umbrella, incorrectly assuming that the molecular mechanisms underlying the antioxidant activity of all flavonoids are nearly identical.

1.2. The Application of Computational Chemistry in Exploring the Antioxidant Activity of Flavonoids

An intriguing alternative for obtaining a more accurate understanding of the biological activity of flavonoids is to leverage computational chemistry. In recent years, our understanding and skills in quantum chemical theory have further deepened causing the expanding landscape of computational chemistry [7,8]. New ‘pathways’ for exploring the antioxidant activity of flavonoids continuously emerge. In this review, we will provide an overview of diverse strategies of computational chemistry used to elucidate and visualize the antioxidant mechanism of flavonoids.

2. Preliminary Statistics on Research Status

To begin, we gathered the literature that utilized computational chemistry to investigate the antioxidant activity of flavonoids over the past five years. In this survey, we found data on 189 flavonoids. Figure 1 illustrates the basic skeleton of the subclasses to which the studied structures belong, along with the number of compounds in each subclass that have been studied over the past five years. This underscores a predominant focus on flavones/flavonols and flavonoid-glycosides.
In this period, computational chemistry has been employed to explore five aspects of flavonoids related to antioxidant activity: electronic structure, thermodynamic enthalpies, the reaction kinetics in the fee radical scavenging reaction, intramolecular and intermolecular interaction, and bioavailability. Among these, electronic structure analysis has been the most extensively studied. Figure 2 illustrates the methods employed in the past five years to investigate the electronic structure of flavonoids, along with the proportion of each method used.
Figure 2 clearly shows that in this period, the classical methods were preferred, including the frontier molecular orbitals, molecular electrostatic potential, and global chemical reactivity descriptors. The more recently developed and intuitive indicators, for example, the density-of-states and donator–acceptor map, have not received significant attention. Apparently, most researchers prefer to follow the already well-known and well-trodden paths. However, it is crucial to acknowledge the limitations of these conventional approaches. At times, the chosen calculation methods may not align with the assumed scenario.
To progress, we have to explore and forge new paths. Ultimately, we should strive to combine diverse methods, as each provides valuable but limited information. Integrating various methods and perspectives will yield a more accurate and comprehensive understanding.
Our goal is to optimize the use of computational chemistry in resolving the antioxidant activity of flavonoids. Therefore, the principles, characteristics, and limitations of the above five calculation strategies are discussed in this review, as well as current trends.

3. Computational Chemistry Strategies to Investigate the Antioxidant Activity of Flavonoids

3.1. Electronic Structure Analysis

Electronic structure analysis strategies giving the electron distribution characteristics of flavonoids have often been used to explain and predict their activity, pinpoint the reaction sites, and elucidate the molecular mechanisms, as well as estimate the stability of complexes between flavonoids with targets, e.g., metals and macromolecules. In the following paragraphs, the electronic structure analysis strategies used in the past five years will be discussed successively.

3.1.1. Frontier Molecular Orbitals (FMOs)

The frontier orbitals are the molecular orbitals that preferentially participate in the reactions. These include (i) the highest occupied molecular orbital (HOMO) and (ii) the lowest unoccupied molecular orbital (LUMO) [83]. The energy levels of the HOMO (EHOMO) and LUMO (ELUMO) specify the electron donating and accepting ability of a molecule, respectively. For an antioxidant, a more negative value of EHOMO implies a higher potency to donate an electron to a radical [84,85]. The localization of the HOMO and LUMO also pinpoints the key pharmacophore of a molecule. The HOMO provides a reliable indication of the location of the outermost, and consequently most energetic, electron in a molecule. Theoretically, this is also the electron most readily donated by an antioxidant. In Figure 3, the HOMO orbital of the quercetin (Q) is shown, which indicates that Q is most likely to donate an electron from the B ring and the double bond in the C ring.
The energy gap of frontier molecular orbital (Egap) is the difference between ELUMO and EHOMO; the higher Egap, the more kinetically inert the molecule is [89,90]. Egap has been suggested to be a more precise descriptor than EHOMO for the activity investigation of an antioxidant [91].
FMO analysis has been the most widely used tool to study the antioxidant activity of flavonoids over the past five years. However, FMO does not provide the full picture of chemical reactivity [92]. Recent evidence shows that the lower-lying molecular orbitals cannot be neglected [93,94,95]. Bulat et al., in 2021, used FMO to predict the reaction site of twelve molecules and only succeeded with five of them. They found that average local ionization energy I(r) is a better descriptor [83]. Average local ionization energy, introduced in 1990, is the average energy needed to remove an electron at any point of a molecule [96]. Apparently, the contribution of all electrons, rather than just the electrons in specific orbitals of the molecule, needs to be taken into account [97,98]. I(r) has been used to explore the activity of antioxidants, such as L-ascorbate and α-tocopherol [99,100]. Despite its potential advantages, I(r) has not been used to study the antioxidant activity of flavonoids in the past five years.

3.1.2. Molecular Electrostatic Potential (MEP/ESP)

MEP is defined as the energy required to move one positive point charge from infinity to a specific point in a molecule. The surface MEP map depicts the electrostatic potential distribution of the molecule with different colors [101]. Typically, the red area signifies a negative MEP surface, indicating a tendency to donate electrons, while the blue region indicates a positive MEP surface, suggesting a propensity to accept electrons. The surface MEP map of Q is shown in Figure 4.
The surface MEP map is an effective tool for predicting reaction sites and has already been applied to flavonoids [58]. There is a strong correlation between the MEP and the acidity of the molecule, indicating that the MEP can be used to estimate the pKa value of the hydroxyl groups at different positions on flavonoids [102]. Given the significant influence of deprotonation on the antioxidant activity of flavonoids (as will be discussed in Section 3.5.1), this warrants further exploration.

3.1.3. Global Chemical Reactivity Descriptors (GCRDs)

GCRDs describe the sensitivity of a compound’s structure to slight changes in the external potential and electrons. These descriptors are used to predict reactivity and structure stability [103]. The Hartree–Fock theory defines the energy involved in extracting and donating electrons from a molecule, as IP and EA respectively [83]. The IP is equal to the difference of electronic energy between the N state (the neutral state of the molecule) and the N-1 state (the state after losing one proton). The adiabatic ionization potential (AIP) is the electronic energy gap between the optimized N and N-1 states, while the vertical ionization potential (VIP) only uses the optimized N state. The lower the IP, the less energy is needed to lose an electron [104,105]. The GCRPs commonly used to measure the electron transfer characteristics of antioxidants are given in Table 1.
GCRPs are important, since they closely correlate with various physicochemical properties like bonding energies, aromaticity, and polarizability [113]. The advantage of its broad correlation comes with the disadvantage that the results of GCRPs usually need to be refined by more specific descriptors.

3.1.4. Natural Bond Orbitals (NBOs)

The NBO theory partially localizes molecular orbitals by diagonalizing the density matrix to facilitate an intuitive understanding of electron orbitals and distribution. NBO analysis is employed in computational chemistry to calculate the electron-occupied orbitals and distribution of electron density between chemical bonds and describe the migration and rearrangement of bonds during chemical reactions [114,115].
Transitions and rearrangements often occur during the antioxidant processes of flavonoids, yet the precise mechanisms remain elusive due to the difficulty in capturing them. Undoubtedly, NBO analysis stands out as a powerful tools to fill this gap. Moreover, it also deserves wider utilization in exploring the impact of substituent groups. Although NBO analysis is powerful, it currently only serves as an auxiliary tool for describing electron distribution in the study of the antioxidant activity of flavonoids [47,54].
It should be noted that traditional NBO analysis results may be unreliable for systems with complex electronic structures and multi-center delocalization. However, an enhanced version is already available, capable of accurately describing the shifts of localized bonds in reaction processes [116]. Moreover, NBO analysis needs to be improved to examine the interaction between metals and non-metals [117]. Therefore, it is not yet suited to explore the metal-chelating activity of flavonoids.

3.1.5. Natural Transition Orbitals (NTOs)

The transition of electrons between the ground state and excited states in molecules can be described as the migration between different NTOs [79]. At present, the NTO analysis of flavonoids focuses on the outer electron going from the ground state (S0) to the first excited state (S1) [118]. In particular, the emphasis in the study revolves around excited state intramolecular proton transfer (ESIPT), the process in which excited molecules relax their energy through tautomerization, the intramolecular transfer of a proton. Chaofan et al. demonstrated, using density functional theory (DFT) and time-dependent density functional theory (TD-DFT), that the rate of the ESIPT reaction has a negative relationship with the antioxidant activity in a series of flavonoids [119].

3.1.6. Spin Density Distribution (SDD)

SDD is a descriptor that shows the delocalization of the unpaired electron in a radical molecule. For flavonoids, it provides information on the reaction kinetics and stability of the intermediates formed during free radical scavenging. A highly delocalized SDD indicates a fast radical scavenging reaction and high stability of the radical intermediate. Here also, intramolecular proton transfer should be considered together, as it can assist the unpaired electron to delocalize and disperse over the molecule [120,121]. For example, as shown in Figure 5, for Q-7-O-radical, a long-conjugated chain can be formed after intramolecular proton transfer, the lone pair electron is delocalized throughout the molecule.
SDD holds a unique position in the study of flavonoids’ ability to directly scavenge free radicals because it actually describes the intermediates formed. It is often used in conjunction with other reactivity descriptors to comprehensively evaluate the antioxidant activity of flavonoids.

3.1.7. Polarity and Dipole Moment

The uneven distribution of charges in covalent bonds of a molecule creates a dipole moment and makes the molecule polar. This uneven distribution is caused by the difference in electronegativity between bonded atoms. The extent of the imbalance is given by the MPI [122].
Although polarity and dipole moment are not descriptors that directly describe the electron distribution, they are also often calculated when studying the activity of antioxidants, as they are a ‘conventional characteristic’ of a molecule [123]. The dipole moment has also been used to study the stereo-conformation and electronic distribution properties of flavonoids [124,125].

3.1.8. Fukui Function (FF)

In computational chemistry, the FF describes the changes in electron density in frontier orbitals caused by the change in the total number of electrons in a molecule [126].
FF descriptors have been developed separately for nucleophilic reactions (f+), electrophilic reactions (f), and/or radical reactions (f0), as shown in the following formulas:
f = ρ N ρ N 1
f   + = ρ ( N + 1 ) ρ ( N )
f 0 = ρ N + 1 + ρ N 1 2
In these equations, ρ (N + 1), ρ (N) and ρ (N − 1) are the electron density in the frontier orbital of the anion, the neutral molecule, and the cation respectively.
Dual descriptor (DD) is a more accurate descriptor derived from the FF, which reflects the comprehensive effect of f and f+ [127].
D D = ρ ( N + 1 ) 2 ρ ( N ) + ρ ( N 1 )
The condensed Fukui function (CFF) concentrates electrons in a molecule to each atom, showing the magnitude of the change in electron density on each atom when a molecule loses or gains electrons [128,129]. The higher the CFF f value of an atom, the higher the possibility that an electron will be donated from that atom.
The FF has been widely used in the prediction of the reactive site and appears to be valuable in this respect [130,131]. Houria Djeradi et al. found a significant correlation between the FF and the antioxidant activity of 36 flavonoids [132].

3.1.9. Atom Charge

Atomic charge represents the point charge situated at the center of an atom. It stands as one of the simplest and most intuitive methods to depict the charge distribution in molecules [133]. Currently, multiple methods have been used to calculate atomic charge. Significant disparities exist in the outcomes obtained with these methods. Therefore, the appropriate method must be selected after considering the underlying principles and unique characteristics of each computational method [134].
In the study on the antioxidant activity of flavonoids, the analysis of atomic charge can be used to predict reaction sites and to describe the electrostatic interactions in molecular docking, molecular dynamics simulations as well as Monte Carlo simulations [135,136,137]. The Mulliken charge is currently prevalent in antioxidant research. However, its theoretical foundation is relatively weak, and the outcome is highly dependent on the basic set used. Based on the good results obtained with ligand-receptor interactions in drug design, AM1-BCC charges (Mulliken charge generated by AM1 semi-empirical method, then corrected by bond charge) and MMFF94 charges (Merck Molecular Force Field 94) appear to be more appropriate [138,139].

3.1.10. Redox Potential

Redox potential is considered to be one of the most accurate chemical descriptors of the ability of a compound to gain or lose electrons. Because the redox potential can be determined experimentally relatively easily and unambiguously, it is a relatively rigorous indicator.
The redox potential of flavonoids is often used in studying their antioxidant behavior. For example, Miličević et al. have studied changes in the electronic structure of 20 flavonoids due to electrochemical oxidation and explored the relationship between the antioxidant activity and oxidation potential of 6 flavonoids more extensively [140]. An example of its use in the last 5 years is the attempt to design multi-target flavonoids to combat Alzheimer’s [141].

3.1.11. Density-of-States (DOS)

The DOS function is a tool to visualize the electronic structure of molecules with energy as a variable. It can be adapted to a specific problem, and therefore DOS has numerous forms. When applied to visualize the electronic structure of a molecule, DOS (E) indicates the number of molecular orbitals per energy unit. The DOS versions best suited for studying the antioxidant activity of flavonoids are total DOS (TDOS) and overlap population DOS (OPDOS).
The TDOS curve delineates the density of orbital distribution across different energy regions. TDOS diagrams are valuable for finding all possible sites where bonds can break. Partial DOS (PDOS), also known as ‘local DOS’ or ‘fractional DOS’, elucidates the contribution of specific fragments of a molecule to TDOS.
OPDOS is instrumental in exploring interactions between fragments. Notably, if an OPDOS value is negative within a particular energy range, it indicates that the orbitals within this range promote fragmentation [142,143]. The OPDOS curves stand out as one of the most effective methods for studying the influence of a specific group on the activity of a compound and for constructing structure–activity relationships (SAR).

3.1.12. Donator–Acceptor Map (DAM)

DAM is a tool that directly visualizes the electron-accepting and donating abilities of compounds. In Figure 6, the four distinct zones of the DAM are shown. The value of Rd and Ra was calculated by dividing ω and ω+ of the compound by ωF (3.40) and ωNa+ (3.46), respectively [144].
Various modified versions of DAMs more suitable for visualizing the antioxidant activity of flavonoids have been developed [145,146]. An example is the full electron–donor–acceptor map (FEDAM) [147]. In FEDAM, electron acceptance (REA) and electron donation (RIE) indices are plotted on the X and Y axis. The REA and RIE are calculated from the VIP and VEA. Another example is the electron and hydrogen donating ability map (eH-DAMA). The Y axis of the map corresponds to ω of a compound, while the X axis corresponds to the BDE of that compound. Compounds located in the bottom left region of the eH-DAMA are likely to be good radical scavengers, as they tend to easily donate electrons (low ω) and hydrogen atoms (low BDE).

3.1.13. Electron Localization Function (ELF) and Localization Orbital Locator (LOL)

ELF is a three-dimensional real space function with a value ranging from 0 to 1. The upper limit value of 1 indicates that the electron is completely localized in this range. Electrons are easily delocalized from regions with lower ELF values to other regions.
E L F r = 1 1 + D r D 0 r 2
The D(r) is given by the following equation:
D r = 1 2 × i N i × ψ i r 2 ρ r 2 8 × ρ r
In this equation, N(i) is the orbital occupation number, σ is the gradient operator, ψi is the wave function, and ρ(r) is the total electron density function. The 1/2 × ∑[i]N(i) × |σψi(r)|2 represents the kinetic energy density function of the system (also called G(r) or gradient kinetic energy density), |σρ(r)|2/(8 × ρ(r)) is the Weizsacker functional [148,149].
The D0(r) is an artificial reference standard, so ELF is essentially a relative degree of localization rather than an absolute measure.
D 0 r = 3 10 × 3 × p i 2 2 3 × ρ r 5 3
The characteristics of the atomic shell structure, lone pair electrons, and chemical bonds can be visualized simultaneously in the distribution figure of ELF [150,151]. The ELF function curve can also be used to investigate the changes in the total electron density of different shells. It should be noted that the ELF function has a flaw. When r approaches infinity, D(r) will gradually approach 0, so ELF will be equal to 1 at a position far from the molecule, suggesting complete localization, which is obviously incorrect [152].
LOL and ELF are in essence both reflections of different kinetic energy density functions. Schmider and Becke defined the LOL function, γ(r), in 2000 [153] as follows:
γ r = t r 1 + t r
The parameter t(r) is defined by the following formula:
t r = D 0 r 1 2 × i N i × ψ i r 2
t(r) is the ratio of the kinetic energy density of a uniform electron gas to the current actual kinetic energy density of the system, which has orbital unitary transformation invariance. Therefore, LOL will not become 1 when r approaches infinity. Moreover, LOL offers simpler functional forms and more straightforward graphs compared to those generated by ELF. Consequently, despite the currently predominant use of ELF in studies on the antioxidant properties of flavonoids, LOL appears to be more suitable [154,155].

3.2. Thermodynamic Analysis

3.2.1. Related Enthalpies of Different Mechanisms for Direct Scavenging of Free Radicals

The mechanisms that have been mainly considered in the direct scavenging of free radicals by flavonoids are (i) hydrogen atom transfer (HAT), (ii) sequential proton loss electron transfer (SPLET), and (iii) electron transfer followed by proton transfer (ETPT). Their reaction equations and enthalpies are given in Table 2.
In the first attempts to evaluate the antioxidant activity of flavonoids, HAT was considered to be the preferred mechanism. Therefore, BDE values were usually calculated to estimate the free radical scavenging potency, and mostly the 4′-OH on the B ring was found to be the primary reaction site [156]. Later on, it was found that antioxidant activity also depended on the reaction medium, especially the polarity and pH of the solvent. Incorporating this indicated that the SPLET mechanism was the most relevant in vivo [157,158,159,160]. Nowadays, the enthalpies of all these three mechanisms are widely used to estimate the antioxidant potency and pinpoint reaction sites of both neutral flavonoid molecules and their deprotonated forms [161,162].
Initially, the overall reaction was only characterized by the tendency to lose one hydrogen and one electron. Gradually, more complete free radical scavenging mechanisms were explored, involving several electron and proton loss sequences [163]. For example, Amic et al. found that the enthalpy of the second proton-coupled electron transfer is lower than that of the first one, indicating that also this second reaction is involved in the antioxidant activity of flavonoids [164]. Mittal et al. calculated and compared the 12 enthalpies of possible scavenging mechanisms for chalcone and suggested that mechanisms containing several steps should be considered [165].

3.2.2. The Reorganization Enthalpy (RE) and Hydrogen Abstraction Energy (HAE)

The RE value describes the energy difference between a specific conformation of a molecule and its optimized conformation [166]. It can serve as a predictor of the conformational changes in flavonoids after the hydrogen extraction process [167]. One of the most representative RE indicators is ortho-hydroquinone formation energy (o-HFE). It represents the enthalpy associated with the formation of an ortho-hydroquinone [168]. This process has been extensively investigated as a critical stage of flavonoids to scavenge free radicals [169].
The calculation formulas of HAE and BDE share similarities, but their calculation processes differ. The method proposed by Sroka, Z et al. to calculate the HAE of flavonoids is first to calculate the bond dissociation enthalpies of each site in various conformations that have different orientations of the hydroxyl groups. The minimum value is considered the final HAE for each site of flavonoids [170]. Conversely, to obtain the BDE of flavonoids, the lowest energy conformation is first identified, followed by separate calculations of the bond dissociation enthalpy for each site [58].
Another related descriptor is ΔHf, which has been used to calculate the heat needed to convert a flavonoid into the corresponding radical [171].
After reviewing the literature, we conclude that the calculation of enthalpy values does contribute to understanding the antioxidant process. Sometimes a correlation is found between a particular enthalpy descriptor and the antioxidant activity of flavonoids [161], but usually this is not the case. Apparently, it is impossible to evaluate the antioxidant process using a single enthalpy indicator. There are still many factors that cannot be ignored, such as kinetic aspects and the influence of the reaction environment. For instance, Morteza Jabbari et al. investigated the reaction rate between naringenin and 2,2-diphenyl-1-picrylhydrazyl radical (DPPH) in water-ethanol solutions ranging from 50% to 90% ethanol and found that a higher ethanol content results in lower reaction rates [172]. Similarly, Malgorzata Musiali et al. discovered that the reaction rate of ten flavonoids with the DPPH in alcohols was consistently significantly faster than in acidified alcohols or dioxane [173].
Another reason why the antioxidant activity of flavonoids does not correlate with a single particular enthalpy descriptor, is that the antioxidant mechanism of then does not follow a single path. Instead, it more likely can follow various parallel reaction paths, that can be intertwined, spanning multiple steps [5]. Moreover, these paths may differ for different flavonoids.

3.3. Kinetic Studies

Transition state (TS) means the intermediate structure formed in a reaction that has the highest energy. The conventional transition state theory (TST) has been widely employed to view the details of a reaction, especially for radical reactions [174,175]. For the reaction of flavonoids with radicals, the gradual change in the structure of the reactants during the reaction also can be depicted using TST [176,177].
Current computational chemistry studies on the kinetics of flavonoids’ antioxidant activity primarily focus on calculating rate constants (K) with TST. The reaction rate constant (K) can be calculated using the ΔG (Gibbs free energy of activation) found with TST according to the following equation:
K T = σ × κ × T × k B h × e G R T
In the equation, κ, h, and kB denote the Wigner coefficient and the Plank and Boltzmann constants, respectively. The parameter σ reflects the reaction symmetry, which represents reaction path degeneracy [178].
In general, for flavonoids, the HAT mechanism predominates in the gas phase, whereas the SPLET mechanism is the primary pathway in polar solvents such as methanol and water [176]. In water, the anion of flavonoids tends to get more attention [162]. Calculation of the electron transfer reaction rates of neutral and anionic forms of a flavonoid can indicate which specific reactions are kinetically favored, and thus identify the favorable reaction pathways [179,180]. This is important for elucidating the mechanisms by which flavonoids scavenge free radicals in biological systems.
However, TST can only calculate the reaction rate of flavonoids with a ‘specific’ free radical, and therefore it is not a descriptor of the overall antioxidant activity. This limitation is inevitable, as the structure of each radical and the solvent environment will affect the reaction rate. In addition, the currently widely used TST is not very successful in accurately calculating the reaction rate constant. This is because the calculation of the reaction rate constant requires more in-depth knowledge of other descriptors, such as the potential energy surface [181]. Given that the rate constant can be determined accurately in experiments, further refining TST and validating theoretical outcomes with experimental data is the logical path to take.

3.4. Interaction Analysis

3.4.1. Quantum Theory of Atoms in Molecules (QTAIM)

The QTAIM is a mathematically rigorous theory concerning atomic region division, bonding analysis, and charge analysis. Its theoretical foundation relies entirely on the properties and geometric characteristics of electron density (as a scalar field), thus minimizing the subjectivity of artificial division and yielding highly precise and quantifiable results. For isolated molecular systems, the most conscientious QTAIM analysis program currently available is Multiwfn, which is both powerful and free [142]. Another powerful open-source program capable of QTAIM analysis is Critic 2 [182]. Its primary focus lies in topological analysis related to critical points, while it also incidentally supports various analysis methods based on real space functions.
Currently, QTAIM analysis stands out as one of the most powerful tools for analyzing intramolecular hydrogen bonds of flavonoids [58,183,184]. The type of intramolecular hydrogen bond (IHB) could be identified by studying the electron density ρ(r) and Laplacian ∇2ρ(r) and Hamiltonian kinetic energy H(r) at the bond critical points (BCP). The hydrogen bond interactions can be divided into the following categories: (1) ∇2ρ(r) > 0 and H(r) > 0 indicating a weak hydrogen bond; (2) ∇2ρ(r) > 0 and H(r) < 0 indicating a moderate hydrogen bond; and (3) ∇2ρ(r) < 0 and H(r) < 0 indicating strong hydrogen bond [185]. Another indicator to describe the strength of hydrogen bonds is the hydrogen bonding energy (EHB) [58].
The IHBs are present in the most stable conformation of flavonoids. It reduces the antioxidant activity of hydrogen-bond donors (such as the 5-OH, 3-OH and 3′-OH groups in flavonoids) and enhances the antioxidant activity of hydrogen-bond acceptors (such as the 4′-OH group) [163,186], which is considered crucial for scavenging free radicals by flavonoids [33,78,187]. Moreover, it is closely connected to the ESIPT mechanism. Zhong, Y et al. reported that IHB promotes ESIPT in the S0 and S1 states of flavonoids [188]. Furthermore, Feixiang et al. found that the site where ESIPT occurs, is partly determined by IHB [189].

3.4.2. Reduced Density Gradient (RDG) and Noncovalent Interactions (NCI) Analysis

  • RDG is a dimensionless parameter utilized to characterize electronic nonuniformity in density functional theory. In 2010, Professor Yang introduced the noncovalent interaction (NCI) analysis method based on RDG, to study the interactions between atoms or molecules [190]. While the electron localization function and QTAIM theory are suitable for analyzing chemical bonds, they possess certain limitations when examining non-covalent interactions like hydrogen bonds, steric hindrance, and π–π stacking. NCI analysis is a potent tool to reveal these non-covalent interactions.
  • NCI analysis can intuitively and comprehensively depict the distribution of weak interactions of flavonoid molecules [27,58]. In NCI analysis, the interaction types are visualized using a color gradient, varying from blue to green indicating a gradual weakening of weak interactions, while red denotes increased resistance. An example is given in Figure 7.

3.4.3. Molecular Docking (MD) and Molecular Dynamics Simulations (MDS)

The principles of MD are the ‘lock-and-key theory’ and the ‘inducible fit theory’, which are models to visualize the reversible binding between a ligand and a receptor [191,192]. The strategies can be divided into:
  • Rigid docking: both the conformation of ligands and the receptors are fixed.
  • Semiflexible docking: only the conformation of the ligands can change freely.
  • Flexible docking: the conformation of both the ligands and the receptors can change freely.
Flexible docking offers the closest simulations of the interaction between ligands and macromolecules [193,194].
MD can rapidly simulate molecular interactions such as hydrogen bonds, electronic interaction, and Van der Waals interaction. MDS can acquire diverse positions, velocities, and trajectories of molecules by solving the potential energy and atomic bonds within the system. It is instrumental in further inferring or evaluating receptor–ligand interactions by simulating the lowest energy conformation and kinetic trajectories of the compounds.
Despite the availability of multiple matching patterns, experimental data indicate that the current molecular docking scoring functions are still insufficiently accurate [68]. Moreover, MD-MDS solely considers physical changes simulates microscopic molecular behaviors, and neglects to account for the breaking and formation of chemical bonds, as well as the precise dynamics of electrons [195,196]. In addition, MDS is both computationally intensive and time-consuming.
Flavonoid antioxidants regulate oxidative stress through numerous pathways. MD might be used to predict the potency of flavonoids on a specific indirect antioxidant pathway such as the inhibition of radical-generating enzymes, the induction of antioxidant enzymes, epigenetic modification, and also qualitatively elucidate the most dominant mechanism. In the past five years, the most common docking objectives of flavonoids were Keap1 [197], Xanthine oxidase (XO) [198,199], NADPH [9,69], peroxiredoxin 5 (HP5) [200], Human tyrosinase-related protein [68], Escherichia coli DNA Gyr [201], α-glucosidase [34], β-lactoglobulin (BLG) [198] and DNA [80]. A novel and promising direction is the use of flavonoids bound to polymers, which may hold great potential for the future [202]. It is important to highlight that, when examining the interaction between flavonoids and macromolecules, the emphasis shifts from reaction kinetics and enthalpy, most prominent in the study of free radical scavenging mechanisms, to equilibriums and Gibbs free energy [203,204,205].
Moreover, the integration of semiempirical molecular orbital methods and density functional theory (DFT) is developing. Currently, there have been some promising attempts, such as ONIOM [206].

3.5. Bioavailability Analysis

3.5.1. Dissociation Constant (pKa)

The pKa, defined as the negative logarithm of the equilibrium dissociation constant of an acid into its conjugated base and a proton (Ka), is a crucial property of flavonoids. It determines the charge of the compound in solutions and thus influences for example solubility, lipophilicity, distribution, protein binding, hydrogen-bound formation, and biological activity. Nevertheless, experimental data regarding the pKa values of flavonoids are scarce. Despite continuous progress, the existing experimental data are frequently incomplete or contradictory [207]. This is predominantly caused by the small difference in the pKa of the various hydroxyl groups in flavonoids. The emergence of computational chemistry is a potential remedy for this issue.
The pKa results provided by computational chemistry are based on the thermodynamic cycle shown in Figure 8, using the following equation [208].
p K a = G s o l n R T l n ( 10 ) = Δ G s o l n + Δ n 1.89 R T l n ( 10 )
In this equation, R represents the ideal gas constant (8.314), n denotes the charge difference before and after deprotonation, and ΔGsoln signifies the free energy of the dissociation process.
Alternatively, there is a more convenient method to obtain the pKa of a compound. In some databases, for example, ACD/Percepta and SPARC, the experimental data of the pKa values of a wide variety of compounds have been collected and used to predict the pKa of related compounds. These software predictions of the pKa appear to be accurate [207,209].
Deprotonation of hydroxyl groups in a flavonoid affects its antioxidant activity [210]. Theoretically, a negative correlation between the free radical scavenging activity of flavonoids and their pKa is expected in polar solvents, as low pKa is more conducive to the SPLET mechanism. Nevertheless, there are some studies showing the opposite. For example, Yi Hu et al. found that for licorice flavonoids, a higher pKa value correlated with a higher solvent-mediated antioxidant activity [64]. Apparently, we have not yet found a clear-cut relationship between pKa and free radical scavenging ability, and this is the subject of ongoing research.

3.5.2. Lipid Water Partition Coefficient (LogP and LogD)

The lipophilicity of compounds is typically estimated by its LogP value, which represents the ratio of the concentration of a compound in water versus that in octanol at equilibrium. The partition coefficient generally refers to the concentration ratio of the non-ionized form of the compound. LogP can be determined experimentally, for instance, using reverse phase chromatography or a shaking flask method. Alternatively, it can be calculated, offering advantages such as not requiring compounds of high purity and saving time and costs. The computer algorithms for predicting LogP of compounds have reached a high level of maturity. Examples are the commercial programs Chem3D and ACD lab, and the open-source programs XLOGP3 [211], ALOGPS [212], MarvinSketch [213] and ChemAxon [214].
LogD refers to the sum of the concentrations of all forms of the compound (ionized plus un-ionized) in water with a specific pH relative to that in octanol, at equilibrium. For compounds that cannot donate or accept protons, LogP equals LogD throughout the whole pH range of the water phase. For acids and bases, the fraction of the compound in water that has a charge depends on the pH. In a water/octanol mixture, mostly the charged molecules are almost completely found in the water phase. Therefore, by combining the log P and the pKa value of an acidic flavonoid with the pH of the water phase, the logD of the flavonoid at that pH can be estimated by logDpH = logP − log(1 + 10pKa − pH).
Low values of either LogD or LogP indicate high hydrophilicity and poor lipophilicity, which predicts good water solubility and poor lipid solubility. Most flavonoids enter cells through passive transport. This means that to enter a cell, the flavonoid has to diffuse from the watery environment of the cell into the lipid bilayer of the membrane. Therefore, a low LogP or logD value of a flavonoid indicates that the flavonoid does not easily go into the lipid bilayer. Therefore, the Logp value and—probably more accurately—the logD value serve as crucial parameters for the transport of a flavonoid over membranes [135,215]. A too-high logP value might also be problematic, as these compounds will be sequestrated in fat tissue [216]. According to Lipinski’s Rule of 5, an oral drug should have a LogP value < 5 and ideally between 1.35 and 1.8 for good oral absorption.
As stated in the introduction, usually it is assumed that in cell experiments the concentration of a compound in all compartments of the cell is equal to the initial concentration in the culture medium. Of course, this is not the case. Based on their lipophilicity, notably LogD, flavonoids will compartmentalize in the body over different organs, and in cells over different organelles. This means that LogD should be valued along with quantum chemical calculations to obtain a deeper understanding of the antioxidant activity of flavonoids.

3.5.3. Polar Surface Area (PSA)

The PSA or topological polar surface area (TPSA) is defined as the sum of the surface areas of all electronegative atoms within a molecule (for flavonoids primarily oxygen) along with their attached hydrogen atoms. This means that the calculation principle of PSA is quite straightforward. Currently, the most popular methods to calculate PSA use quantum chemical topology (QCT) analysis, for instance in the QikProp program. It is worth noting that PSA values obtained from different methods may vary, and therefore cannot be directly compared [217].
In addition to LogD, PSA is also used to predict membrane permeability. Molecules with a PSA exceeding 140 square angstroms (Å2) typically exhibit poor membrane permeability. For molecules to traverse the blood-brain barrier and thereby interact with receptors in the central nervous system, a PSA of less than 90 Å2 is generally required [218].
To exert a biological effect, dietary flavonoids have to cross the intestinal membrane. Over the past five years, PSA has increasingly been used to predict the permeability of flavonoids through cell membranes. The PSA of myricetin [219], kaempferol [219], Q [219] and procyanidin B1(PB1) [68] have been calculated and used as supplementary descriptors to estimate their overall in vivo antioxidant potency.

4. Concluding Remarks

Over the past five years, various computational chemistry strategies have significantly advanced our understanding of the antioxidant activity of flavonoids. Moreover, in our drive to progress, each of these strategies is continuously refined to develop even more suitable, accurate and detailed calculation methods.
However, due to the flexibility of the reaction pathway, even the ‘simple’ direct scavenging activity of a flavonoid in a biological system cannot fully be described by only one of the parameters given in this overview. As also stated in the introduction, a minor change in the molecular structure of a flavonoid can have a drastic effect on its antioxidant effect, indicating that each flavonoid has its unique antioxidant profile [220]. Therefore, it is essential to integrate the results of multiple parameters to obtain a comprehensive profile that accurately reflects the antioxidant effects of flavonoids. To address this, we propose a convenient way to visualize the comprehensive antioxidant of flavonoids, displaying various computational descriptors simultaneously in a radar map, as exemplified in Figure 9.
Figure 10 gives an overarching view of the computational landscape of the antioxidant activity of a flavonoid incorporating all discussed descriptors. In this, we follow the paradigm coined by Frazer and Brown that the biological activity of a compound is a function of its molecular structure. Therefore, we have placed the structure activity relationship (SAR) in the center. By integrating the antioxidant profiles of structurally related flavonoids in a SAR, clues can be found to identify the structural elements of the flavonoid molecule that are important for its antioxidant activity, as well as the type of interaction, and the consequence of this interaction, thus enhancing our understanding of the molecular mechanisms underlying flavonoid antioxidant activity, as well as their biological effects and health implications.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29112627/s1, Table S1: The flavonoids studied by computational chemistry theory in the past five years; Table S2: The Electronic structure analysis strategies used to investigate the antioxidants of flavonoids in the past five years.

Author Contributions

As with all achievements in life, this manuscript is the result of concerted action of many people, not restricted to the ones at the top of this manuscript, and therefore even specific roles cannot be accurately claimed by one person. The contributors to this manuscript are: data collection, Y.W. and C.L.; writing—original draft preparation, Y.W.; writing—reviewing and editing, Y.W. and M.Z.; visualization, Y.W.; supervision, M.Z., Z.L., G.J.M.d.H. and M.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by National Natural Science Foundation of China (32201977). Yue Wang and Chujie Li were supported by the China Scholarship Council (CSC, File No. 202109110043 and No. 202108440109).

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

We appreciate the help of Guido R.M.M. Haenen.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Panche, A.N.; Diwan, A.D.; Chandra, S.R. Flavonoids: An Overview. J. Nutr. Sci. 2016, 5, 47. [Google Scholar] [CrossRef] [PubMed]
  2. Ullah, A.; Munir, S.; Badshah, S.L.; Khan, N.; Ghani, L.; Poulson, B.G.; Emwas, A.-H.; Jaremko, M. Important Flavonoids and Their Role as a Therapeutic Agent. Molecules 2020, 25, 5243. [Google Scholar] [CrossRef]
  3. Dias, M.C.; Pinto, D.C.G.A.; Silva, A.M.S. Plant Flavonoids: Chemical Characteristics and Biological Activity. Molecules 2021, 26, 5377. [Google Scholar] [CrossRef] [PubMed]
  4. Brown, A.C.; Fraser, T.R. V.—On the connection between chemical constitution and physiological action. part. i.—On the physiological action of the salts of the ammonium bases, derived from Strychnia, Brucia, Thebaia, Codeia, Morphia, and Nicotia. Earth Environ. Sci. Trans. R. Soc. Edinb. 1868, 25, 151–203. [Google Scholar] [CrossRef]
  5. Li, Z.; Moalin, M.; Zhang, M.; Vervoort, L.; Hursel, E.; Mommers, A.; Haenen, G.R.M.M. The Flow of the Redox Energy in Q during Its Antioxidant Activity in Water. Int. J. Mol. Sci. 2020, 21, 6015. [Google Scholar] [CrossRef] [PubMed]
  6. Li, Z.; Zhang, M.; Haenen, G.R.M.M.; Vervoort, L.; Moalin, M. Flavonoids Seen through the Energy Perspective. Int. J. Mol. Sci. 2022, 23, 187. [Google Scholar] [CrossRef] [PubMed]
  7. Gorbachev, M.; Gorinchoy, N.; Balan, I.; Arsene, I. Electronic Structure-Antioxidant Action Relationships for Chemical Compounds: A Quantum Chemical Study. In Fundamental and Biomedical Aspects of Redox Processes; IGI Global: Hershey, PA, USA, 2023; pp. 143–158. [Google Scholar]
  8. Spiegel, M. Current Trends in Computational Quantum Chemistry Studies on Antioxidant Radical Scavenging Activity. J. Chem. Inf. Model. 2022, 62, 2639–2658. [Google Scholar] [CrossRef] [PubMed]
  9. Imran, M.; Irfan, A.; Khalid, M.; Khalid, N.; Uddin, J.; Hussain, R.; Ali, B.; Hussien, M.; Assiri, M.A.; Al-Sehemi, A.G. In-Vitro and in-Silico Antioxidant, α-Glucosidase Inhibitory Potentials of Abutilins C and D, New Flavonoide Glycosides from Abutilon Pakistanicum. Arab. J. Chem. 2021, 14, 103021. [Google Scholar] [CrossRef]
  10. Roquete Amparo, T.; Cherem Peixoto da Silva, A.; Brandão Seibert, J.; dos Santos da Silva, D.; Martins Rebello dos Santos, V.; Melo de Abreu Vieira, P.; Célio Brandão, G.; Henrique Bianco de Souza, G.; Aloise Maneira Corrêa Santos, B. In Vitro and in Silico Investigation of the Photoprotective and Antioxidant Potential of Protium Spruceanum Leaves and Its Main Flavonoids. J. Photochem. Photobiol. Chem. 2022, 431, 114037. [Google Scholar] [CrossRef]
  11. Anbazhakan, K.; Praveena, R.; Sadasivam, K.; Salgado, G.; Cardona, W.; Gerli, L.; Alvarado-Soto, L.; Ramirez-Tagle, R. Theoretical Insight on Structural Activities and Targets of Kaempferol Glycosides. Afinidad J. Chem. Eng. Theor. Appl. Chem. 2021, 78, 236–239. [Google Scholar]
  12. Ninh The, S.; Do Minh, T.; Nguyen Van, T. Isoflavones and Isoflavone Glycosides: Structural-Electronic Properties and Antioxidant Relations—A Case of DFT Study. J. Chem. 2019, 2019, 4360175. [Google Scholar] [CrossRef]
  13. Yang, X.; Wang, T.; Guo, J.; Sun, M.; Wong, M.W.; Huang, D. Dietary Flavonoids Scavenge Hypochlorous Acid via Chlorination on A- and C-Rings as Primary Reaction Sites: Structure and Reactivity Relationship. J. Agric. Food Chem. 2019, 67, 4346–4354. [Google Scholar] [CrossRef] [PubMed]
  14. de Souza Farias, S.A.; da Costa, K.S.; Martins, J.B.L. Analysis of Conformational, Structural, Magnetic, and Electronic Properties Related to Antioxidant Activity: Revisiting Flavan, Anthocyanidin, Flavanone, Flavonol, Isoflavone, Flavone, and Flavan-3-Ol. ACS Omega 2021, 6, 8908–8918. [Google Scholar] [CrossRef] [PubMed]
  15. Sharma, M.; Ahuja, D. QSAR Studies of Flavonoids Derivatives for Antioxidant and Antimicrobial Activity. J. Drug Deliv. Ther. 2019, 9, 765–773. [Google Scholar]
  16. Bitew, M.; Desalegn, T.; Demissie, T.B.; Belayneh, A.; Endale, M.; Eswaramoorthy, R. Pharmacokinetics and Drug-Likeness of Antidiabetic Flavonoids: Molecular Docking and DFT Study. PLoS ONE 2021, 16, e0260853. [Google Scholar] [CrossRef] [PubMed]
  17. Messaadia, L.; Bekkar, Y.; Benamira, M.; Lahmar, H. Predicting the Antioxidant Activity of Some Flavonoids of Arbutus Plant: A Theoretical Approach. Chem. Phys. Impact 2020, 1, 100007. [Google Scholar] [CrossRef]
  18. Huy, N.T.; Trang, N.V.; Xuyen, N.T.; Hong, C.T.; Anh, V.T.K.; Thai, V.Q.; Cuong, B.V.; Van, T.T.T.; Trang, N.V.; Lam, T.D.; et al. Studies on the Antioxidant Activity of Apigenin, Luteolin and Nevadensin Using DFT. Vietnam. J. Sci. Technol. 2021, 59, 19–29. [Google Scholar] [CrossRef]
  19. Irfan, A.; Imran, M.; Khalid, M.; Sami Ullah, M.; Khalid, N.; Assiri, M.A.; Thomas, R.; Muthu, S.; Raza Basra, M.A.; Hussein, M.; et al. Phenolic and Flavonoid Contents in Malva Sylvestris and Exploration of Active Drugs as Antioxidant and Anti-COVID19 by Quantum Chemical and Molecular Docking Studies. J. Saudi Chem. Soc. 2021, 25, 101277. [Google Scholar] [CrossRef]
  20. Elsharkawy, E.R.; Almalki, F.; Ben Hadda, T.; Rastija, V.; Lafridi, H.; Zgou, H. DFT Calculations and POM Analyses of Cytotoxicity of Some Flavonoids from Aerial Parts of Cupressus Sempervirens: Docking and Identification of Pharmacophore Sites. Bioorg. Chem. 2020, 100, 103850. [Google Scholar] [CrossRef]
  21. Meshram, R.J.; Bagul, K.T.; Pawnikar, S.P.; Barage, S.H.; Kolte, B.S.; Gacche, R.N. Known Compounds and New Lessons: Structural and Electronic Basis of Flavonoid-Based Bioactivities. J. Biomol. Struct. Dyn. 2020, 38, 1168–1184. [Google Scholar] [CrossRef]
  22. Boulebd, H. The Role of Benzylic-Allylic Hydrogen Atoms on the Antiradical Activity of Prenylated Natural Chalcones: A Thermodynamic and Kinetic Study. J. Biomol. Struct. Dyn. 2021, 39, 1955–1964. [Google Scholar] [CrossRef]
  23. Bartella, L.; Mazzotti, F.; Talarico, I.R.; De Luca, G.; Santoro, I.; Prejanò, M.; Riccioni, C.; Marino, T.; Di Donna, L. Structural Characterization of Peripolin and Study of Antioxidant Activity of HMG Flavonoids from Bergamot Fruit. Antioxidants 2022, 11, 1847. [Google Scholar] [CrossRef]
  24. Mustafa, S.R.; Al-Ani, H.N. Calculation of Vibrational Frequencies, Energetic and Some Other Quantum Chemical Parameters for Some Flavonoids. J. Phys. Conf. Ser. 2021, 1999, 012018. [Google Scholar] [CrossRef]
  25. Jalezadeh, A.; Mirjafary, Z.; Rouhani, M.; Saeidian, H. Investigation of Structural, Electronic, and Antioxidant Properties of Calycopetrin and Xanthomicrol as Two Polymethoxylated Flavones Using DFT Calculations. Struct. Chem. 2022, 33, 1241–1250. [Google Scholar] [CrossRef]
  26. Veiko, A.G.; Lapshina, E.A.; Zavodnik, I.B. Comparative Analysis of Molecular Properties and Reactions with Oxidants for Q, Catechin, and Naringenin. Mol. Cell. Biochem. 2021, 476, 4287–4299. [Google Scholar] [CrossRef]
  27. Wang, L.; Yang, Q.; Li, Y.; Wang, S.; Yang, F.; Zhao, X. How the Functional Group Substitution and Solvent Effects Affect the Antioxidant Activity of (+)-Catechin? J. Mol. Liq. 2021, 327, 114818. [Google Scholar] [CrossRef]
  28. Deviani, V.; Hardianto, A.; Farabi, K.; Herlina, T. Flavanones from Erythrina Crista-Galli Twigs and Their Antioxidant Properties Determined through In Silico and In Vitro Studies. Molecules 2022, 27, 6018. [Google Scholar] [CrossRef]
  29. Santos, K.L.B.; Queiroz, A.N.; Lobato, C.C.; Vale, J.K.L.; Santos, C.B.R.; Borges, R.S. A Comparative Theoretical Mechanism on Simplified Flavonoid Derivatives and Isoxazolone Analogous as Michael System Inhibitor. J. Mol. Model. 2021, 27, 26. [Google Scholar] [CrossRef]
  30. Faria, E.C.M.; Duarte, V.S.; Oliveira, A.M. A trimethoxy-chalcone applied as antioxidant and antibacterial additive for diesel and biodiesel blend. In Proceedings of the XXIX International Symposium on Automotive Engineering—SIMEA 2022, São Paulo, Brazil, 17–18 August 2022. [Google Scholar]
  31. Moreira, C.A.; Faria, E.C.; Queiroz, J.E.; Duarte, V.S.; Gomes, M.D.N.; da Silva, A.M.; de Paula, R.L.G.; Franco, C.H.; Cavalcanti, E.H.D.S.; de Aquino, G.L.; et al. Structural Insights and Antioxidant Analysis of a Tri-Methoxy Chalcone with Potential as a Diesel-Biodiesel Blend Additive. Fuel Process. Technol. 2022, 227, 107122. [Google Scholar] [CrossRef]
  32. Xue, Y.; Liu, Y.; Zhang, L. Antioxidant and spectral properties of chalcones and analogous aurones: Theoretical insights. Int. J. Quantum Chem. 2019, 119, e25808. [Google Scholar] [CrossRef]
  33. Shang, C.; Zhang, Y.; Sun, C.; Wang, L. Tactfully Improve the Antioxidant Activity of 2′-Hydroxychalcone with the Strategy of Substituent, Solvent and Intramolecular Hydrogen Bond Effects. J. Mol. Liq. 2022, 362, 119748. [Google Scholar] [CrossRef]
  34. Siddiqa, A.; Tajammal, A.; Irfan, A.; Azam, M.; Munawar, M.A.; Hardy, R.S.; Basra, M.A.R. Synthesis, Molecular Docking, Bio-Evaluation and Quantitative Structure Activity Relationship of New Chalcone Derivatives as Antioxidants. J. Mol. Struct. 2023, 1277, 134814. [Google Scholar] [CrossRef]
  35. Khalili, A.; Baei, M.T.; Ghaboos, S.H.H. Chrysin Flavonoid Adsorbed on B12N12 Nanocage—A Novel Antioxidant Nanomaterial. Vietnam J. Chem. 2021, 59, 211–220. [Google Scholar] [CrossRef]
  36. Szewczuk, N.A.; Duchowicz, P.R.; Pomilio, A.B.; Lobayan, R.M. Resonance Structure Contributions, Flexibility, and Frontier Molecular Orbitals (HOMO–LUMO) of Pelargonidin, Cyanidin, and Delphinidin throughout the Conformational Space: Application to Antioxidant and Antimutagenic Activities. J. Mol. Model. 2022, 29, 2. [Google Scholar] [CrossRef] [PubMed]
  37. Cheng, L.; Wang, F.; Cao, Y.; Cai, G.; Wei, Q.; Shi, S.; Guo, Y. Screening of Potent α-Glucosidase Inhibitory and Antioxidant Polyphenols in Prunella vulgaris L. by Bioreaction–HPLC–Quadrupole-Time-of-Flight-MS/MS and in Silico Analysis. J. Sep. Sci. 2022, 45, 3393–3403. [Google Scholar] [CrossRef] [PubMed]
  38. Chen, B.; Ma, Y.; Li, H.; Chen, X.; Zhang, C.; Wang, H.; Deng, Z. The Antioxidant Activity and Active Sites of Delphinidin and Petunidin Measured by DFT, in Vitro Chemical-Based and Cell-Based Assays. J. Food Biochem. 2019, 43, e12968. [Google Scholar] [CrossRef] [PubMed]
  39. Xiao, Y.; Fu, F.; Wei, Y.; Shi, S.; Shan, Y. Online Extraction–DPPH–HPLC–DAD–QTOF-MS System for Efficient Screening and Identification of Antioxidants from Citrus aurantium L. Var. Amara (Rutaceae): Integrating Sample Preparation and Antioxidants Profiling. Antioxidants 2022, 11, 1014. [Google Scholar] [CrossRef] [PubMed]
  40. Boulmokh, Y.; Belguidoum, K.; Meddour, F.; Amira-Guebailia, H. Investigation of Antioxidant Activity of Epigallocatechin Gallate and Epicatechin as Compared to Resveratrol and Ascorbic Acid: Experimental and Theoretical Insights. Struct. Chem. 2021, 32, 1907–1923. [Google Scholar] [CrossRef]
  41. Boulebd, H. Structure-Activity Relationship of Antioxidant Prenylated (Iso)Flavonoid-Type Compounds: Quantum Chemistry and Molecular Docking Studies. J. Biomol. Struct. Dyn. 2022, 40, 10373–10382. [Google Scholar] [CrossRef]
  42. Menacer, R.; Rekkab, S.; Kabouche, Z. Fisetin and Robinetin Antiradical Activity under Solvent Effect: Density Functional Theory Study. J. Mol. Model. 2022, 28, 240. [Google Scholar] [CrossRef]
  43. Pop, R. Evaluation of the Radical-Scavenging Properties of Various Flavonols in Ethanol Environment: An Ab Initio Study. Croat. Chem. Acta 2019, 92, 337–346. [Google Scholar] [CrossRef]
  44. Ugodi, G.W.; Asogwa, F.C.; Paul, E.; Asogwa, C.J. Reactivity Indexes of Antioxidant Molecules from Psorospermum Febrifugum. J. Pharmacogn. Phytochem. 2023, 12, 14–18. [Google Scholar] [CrossRef]
  45. Ji, W.; Li, Z.; Gu, L.; Zou, X.; Wu, J.; Zhang, S.; Deng, H. Differences in the Composition of Archidendron Clypearia at Different Harvest Periods and Spectrum–Effect Relationship and Structure–Activity Analysis of Antioxidant Components. SSRN Electron. J. 2022, preprint. [Google Scholar] [CrossRef]
  46. Razooqi, M.S.; Al-Ani, H.N. Quantum Mechanical Calculations and Electrochemical Study of Vibrational Frequencies, Energies in Some Flavonoids Molecules. Iraqi J. Sci. 2022, 63, 2331–2344. [Google Scholar] [CrossRef]
  47. Lewandowski, W.; Lewandowska, H.; Golonko, A.; Świderski, G.; Świsłocka, R.; Kalinowska, M. Correlations between Molecular Structure and Biological Activity in “Logical Series” of Dietary Chromone Derivatives. PLoS ONE 2020, 15, e0229477. [Google Scholar] [CrossRef]
  48. Rammohan, A.; Bhaskar, B.V.; Camilo, A.; Gunasekar, D.; Gu, W.; Zyryanov, G.V. In Silico, in Vitro Antioxidant and Density Functional Theory Based Structure Activity Relationship Studies of Plant Polyphenolics as Prominent Natural Antioxidants. Arab. J. Chem. 2020, 13, 3690–3701. [Google Scholar] [CrossRef]
  49. Bağlan, M.; Yildiko, Ü.; Gören, K. Computational Investigation of 5.5’’,7’’-Trihydroxy-3,7-Dimethoxy-4’-4’’-O-Biflavone from Flavonoids Using DFT Calculations and Molecular Docking. Adıyaman Univ. J. Sci. 2022, 12, 283–298. [Google Scholar] [CrossRef]
  50. Anbazhakan, K.; Praveena, R.; Sadasivam, K. Theoretical Insight on Antioxidant Potency of Kanzakiflavone-2 and Its Derivatives. Struct. Chem. 2021, 32, 1451–1458. [Google Scholar] [CrossRef]
  51. Praveena, A.; Prabu, S.; Madi, F.; Rajamohan, R. Theoretical Investigation of Inclusion Complexes of 3-Hydroxyflavone and Q as Guests with Native and Modified β-Cyclodextrins as Hosts. Polycycl. Aromat. Compd. 2023, 43, 141–153. [Google Scholar] [CrossRef]
  52. El-Hadidy, E.; Ali, M. Theoretical Study, Antioxidant Activity and Anti-Cancer Studies of Galangal (Alpinia Galangal). Int. J. Curr. Res. 2020, 7, 101–145. [Google Scholar]
  53. Balanescu, F.; Busuioc, A.C.; Botezatu, A.V.D.; Gosav, S.; Avramescu, S.M.; Furdui, B.; Dinica, R.M. Comparative Study of Natural Antioxidants from Glycine Max, Anethum Graveolensand Pimpinella Anisum Seed and Sprout Extracts Obtained by Ultrasound-Assisted Extraction. Separations 2022, 9, 152. [Google Scholar] [CrossRef]
  54. Son, N.T.; Mai Thanh, D.T.; Van Trang, N. Flavone Norartocarpetin and Isoflavone 2′-Hydroxygenistein: A Spectroscopic Study for Structure, Electronic Property and Antioxidant Potential Using DFT (Density Functional Theory). J. Mol. Struct. 2019, 1193, 76–88. [Google Scholar] [CrossRef]
  55. Wang, R.; Li, W.; Fang, C.; Zheng, X.; Liu, C.; Huang, Q. Identification of New Flavonoid Compounds in Dandelion Taraxacum Mongolicum Hand.-Mazz. and Evaluation of Their Antioxidant Activities. Res. Sq. 2022. submitted. [Google Scholar]
  56. Wang, R.; Li, W.; Fang, C.; Zheng, X.; Liu, C.; Huang, Q. Extraction and Identification of New Flavonoid Compounds in Dandelion Taraxacum Mongolicum Hand.-Mazz. with Evaluation of Antioxidant Activities. Sci. Rep. 2023, 13, 2166. [Google Scholar] [CrossRef]
  57. Sulaiman, G.M.; Waheeba, H.M.; AL-Shmgani, H.; Eassa, H.A.; Al-Amiery, A.A.; Jabir, M.S.; Dewir, Y.H.; Alwahibi, M.S.; Soliman, D.A. Synthesis, Molecular Modeling, DNA Damage Interaction, and Antioxidant Potential of Hesperidin Loaded on Gold Nanoparticles. J. Biomim. Biomater. Biomed. Eng. 2022, 54, 17–29. [Google Scholar] [CrossRef]
  58. Ragi, C.; Muraleedharan, K. Antioxidant Activity of Hibiscetin and Hibiscitrin: Insight from DFT, NCI, and QTAIM. Theor. Chem. Acc. 2023, 142, 30. [Google Scholar] [CrossRef]
  59. Hu, Y.; Liang, P.; Wang, Z.; Jiang, C.; Zeng, Q.; Shen, C.; Wu, Y.; Liu, L.; Yi, Y.; Zhu, H.; et al. Explore the Effect of the Structure-Activity Relationship and Dose-Effect Relationship on the Antioxidant Activity of Licorice Flavonoids. J. Mol. Struct. 2023, 1292, 136101. [Google Scholar] [CrossRef]
  60. Enisoğlu Atalay, V.; Ölüç, İ.B. Antioxidant Activity of the Hazelnut Plant Determination by Computational Chemistry Methods. Main Group Chem. 2021, 19, 273–282. [Google Scholar] [CrossRef]
  61. Vilas-Boas, I.T.; da Silva, A.C.P.; de AF Accioli, C.; Amorim, J.M.; Leite, P.M.; Faraco, A.A.G.; Santos, B.A.M.C.; Scopel, M.; Castilho, R.O. Optimized Baccharis Dracunculifolia Extract as Photoprotective and Antioxidant: In Vitro and in Silico Assessment. J. Photochem. Photobiol. Chem. 2023, 440, 114654. [Google Scholar] [CrossRef]
  62. Santos, S.C.; Fortes, G.A.C.; Camargo, L.T.F.M.; Camargo, A.J.; Ferri, P.H. Antioxidant Effects of Polyphenolic Compounds and Structure-Activity Relationship Predicted by Multivariate Regression Tree. LWT 2021, 137, 110366. [Google Scholar] [CrossRef]
  63. Babiaka, S.B.; Nia, R.; Abuga, K.O.; Mbah, J.A.; Vincent de Paul, N.N.; Paper, D.H.; Ntie-Kang, F. Antioxidant Potential of Flavonoid Glycosides from Manniophyton Fulvum Müll. (Euphorbiaceae): Identification and Molecular Modeling. Sci. Afr. 2020, 8, e00423. [Google Scholar] [CrossRef]
  64. Hu, Y.; Wang, Z.; Shen, C.; Jiang, C.; Zhu, Z.; Liang, P.; Li, H.; Zeng, Q.; Xue, Y.; Wu, Y.; et al. Influence of the pKa Value on the Antioxidant Activity of Licorice Flavonoids under Solvent-Mediated Effects. Arch. Pharm. 2023, 356, 2200470. [Google Scholar] [CrossRef] [PubMed]
  65. Kenouche, S.; Sandoval-Yañez, C.; Martínez-Araya, J.I. The Antioxidant Capacity of Myricetin. A Molecular Electrostatic Potential Analysis Based on DFT Calculations. Chem. Phys. Lett. 2022, 801, 139708. [Google Scholar] [CrossRef]
  66. Rajan, V.K.; Ragi, C.; Muraleedharan, K. A Computational Exploration into the Structure, Antioxidant Capacity, Toxicity and Drug-like Activity of the Anthocyanidin “Petunidin”. Heliyon 2019, 5, e02115. [Google Scholar] [CrossRef] [PubMed]
  67. Rouhani, M. Evaluation of Structural Properties and Antioxidant Capacity of Proxison: A DFT Investigation. Comput. Theor. Chem. 2021, 1195, 113096. [Google Scholar] [CrossRef]
  68. Thuy, P.T.; Quan, P.M.; Duc, D.X.; Son, N.T. The Antioxidative Potential of Procyanidin B1: DFT (Density Functional Theory) and Docking Approaches. J. Mol. Model. 2022, 28, 356. [Google Scholar] [CrossRef] [PubMed]
  69. da Silva Pantoja, L.V.P.; Trindade, S.S.A.; da Silva Carneiro, A.; Silva, J.P.B.; da Paixão, T.P.; Romeiro, C.F.R.; de Moraes, C.S.P.; Pinto, A.C.G.; Raposo, N.R.B.; de Andrade, M.A. Computational Study of the Main Flavonoids from Chrysobalanus icaco L. against NADPH-Oxidase and in Vitro Antioxidant Activity. Res. Soc. Dev. 2022, 11, e5011628542. [Google Scholar] [CrossRef]
  70. Borges, R.S.; Aguiar, C.P.O.; Oliveira, N.L.L.; Amaral, I.N.A.; Vale, J.K.L.; Chaves Neto, A.M.J.; Queiroz, A.N.; da Silva, A.B.F. Antioxidant Capacity of Simplified Oxygen Heterocycles and Proposed Derivatives by Theoretical Calculations. J. Mol. Model. 2023, 29, 232. [Google Scholar] [CrossRef] [PubMed]
  71. Kiraz, A.Ö.; Yalçın, F. Structure-Activity and Antioxidant Properties of Q and Its Co2+ Chelate. Int. J. Second. Metab. 2021, 8, 414–424. [Google Scholar] [CrossRef]
  72. Thao, T.H.D.; Dung, V.T.N.; Dao, D.Q. Antioxidant vs. pro-Oxidant Activities of Q in Aqueous Phase: A Density Functional Theory Study. Vietnam J. Chem. 2019, 57, 696–701. [Google Scholar] [CrossRef]
  73. Aydın, L.; Özpozan, T. NBO, HOMO–LUMO, Conformational and Vibrational Spectroscopic Study of 2-(3,4-Dihydroxyphenyl)-5,7-Dihydroxy-3-[(2S,3R,4S,5R,6R)-3,4,5-Trihydroxy-6-(Hydroxymethyl) Oxan-2-Yl]-Oxychromen-4-One, as the Main Phenolic Component of the Extracts of Walnut Leaves by DFT. J. Iran. Chem. Soc. 2021, 18, 1067–1079. [Google Scholar]
  74. Boshkayeva, A.K.; Omarova, R.A.; Ordabayeva, S.K.; Serikbayeva, A.D.; Umurzakhova, G.G.; Massakbayev, A.J. Modeling of the Structure and Forecasting Properties of DihydroQ Derivatives. Drug Dev. Ind. Pharm. 2022, 48, 52–57. [Google Scholar] [CrossRef]
  75. Son, N.T.; Thuy, P.T.; Van Trang, N. Antioxidative Capacities of Stilbenoid Suaveolensone A and Flavonoid Suaveolensone B: A Detailed Analysis of Structural-Electronic Properties and Mechanisms. J. Mol. Struct. 2021, 1224, 129025. [Google Scholar] [CrossRef]
  76. Siddiqa, A.; Tajammal, A.; Irfan, A.; Munawar, M.A.; Azam, M.; Basra, M.A.R. Synthesis, Antioxidant, in Silico and Computational Investigation of 2,5-Dihydroxyacetophenone Derived Chloro-Substituted Hydroxychalcones, Hydroxyflavanones and Hydroxyflavindogenides. J. Biomol. Struct. Dyn. 2022, 40, 10265–10277. [Google Scholar] [CrossRef]
  77. Martínez-Busi, M.; Arredondo, F.; González, D.; Echeverry, C.; Vega-Teijido, M.A.; Carvalho, D.; Rodríguez-Haralambides, A.; Rivera, F.; Dajas, F.; Abin-Carriquiry, J.A. Purification, Structural Elucidation, Antioxidant Capacity and Neuroprotective Potential of the Main Polyphenolic Compounds Contained in Achyrocline Satureioides (Lam) D.C. (Compositae). Bioorg. Med. Chem. 2019, 27, 2579–2591. [Google Scholar] [CrossRef]
  78. Zhang, Y.; Shang, C.; Sun, C.; Wang, L. Understanding Prominent Effects of the Intramolecular Hydrogen Bond on the Photophysical Properties and Antiradical Abilities of Six Flavonoids. J. Mol. Liq. 2023, 386, 122534. [Google Scholar] [CrossRef]
  79. Abbo, H.S.; Hung Lai, C.; Titinchi, S.J.J. Substituent and Solvent Effects on UV-visible Absorption Spectra of Chalcones Derivatives: Experimental and Computational Studies. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2023, 303, 123180. [Google Scholar] [CrossRef] [PubMed]
  80. Song, X.; Wang, Y.; Gao, L. Mechanism of Antioxidant Properties of Q and Q-DNA Complex. J. Mol. Model. 2020, 26, 133. [Google Scholar] [CrossRef]
  81. Nam, G.; Hong, M.; Lee, J.; Jin Lee, H.; Ji, Y.; Kang, J.; Baik, M.-H.; Hee Lim, M. Multiple Reactivities of Flavonoids towards Pathological Elements in Alzheimer’s Disease: Structure–Activity Relationship. Chem. Sci. 2020, 11, 10243–10254. [Google Scholar] [CrossRef] [PubMed]
  82. Halevas, E.; Matsia, S.; Hatzidimitriou, A.; Geromichalou, E.; Papadopoulos, T.A.; Katsipis, G.; Pantazaki, A.; Litsardakis, G.; Salifoglou, A. A Unique Ternary Ce(III)-Q-Phenanthroline Assembly with Antioxidant and Anti-Inflammatory Properties. J. Inorg. Biochem. 2022, 235, 111947. [Google Scholar] [CrossRef]
  83. Bulat, F.A.; Murray, J.S.; Politzer, P. Identifying the Most Energetic Electrons in a Molecule: The Highest Occupied Molecular Orbital and the Average Local Ionization Energy. Comput. Theor. Chem. 2021, 1199, 113192. [Google Scholar] [CrossRef]
  84. Akhtari, K.; Hassanzadeh, K.; Fakhraei, B.; Fakhraei, N.; Hassanzadeh, H.; Zarei, S.A. A Density Functional Theory Study of the Reactivity Descriptors and Antioxidant Behavior of Crocin. Comput. Theor. Chem. 2013, 1013, 123–129. [Google Scholar] [CrossRef]
  85. Reza Nazifi, S.M.; Asgharshamsi, M.H.; Dehkordi, M.M.; Zborowski, K.K. Antioxidant Properties of Aloe Vera Components: A DFT Theoretical Evaluation. Free Radic. Res. 2019, 53, 922–931. [Google Scholar] [CrossRef] [PubMed]
  86. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.E. Gaussian ~09 Revision D. 01; Gaussian, Inc.: Wallingford, CT, USA, 2009; Volume 32, pp. 5648–5652. [Google Scholar]
  87. Zhao, Y.; Truhlar, D.G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: Two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215–241. [Google Scholar]
  88. Krishnan, R.; Binkley, J.S.; Seeger, R.; Pople, J.A. Self-consistent molecular orbital methods. XX. A basis set for correlated wave functions. J. Chem. Phys. 1980, 72, 650–654. [Google Scholar] [CrossRef]
  89. Fleming, I. Molecular Orbitals and Organic Chemical Reactions; John Wiley & Sons: Hoboken, NJ, USA, 2009; ISBN 978-0-470-74659-2. [Google Scholar]
  90. Huq, F. Molecular Modelling Analysis of the Metabolism of Fentanyl. J. Pharmacol. Toxicol. 2007, 2, 176–182. [Google Scholar] [CrossRef]
  91. Zainuri, D.A.; Razak, I.A.; Arshad, S. Crystal Structure, Spectroscopic Characterization and DFT Study of Two New Linear Fused-Ring Chalcones. Acta Crystallogr. Sect. E Crystallogr. Commun. 2018, 74, 1427–1432. [Google Scholar] [CrossRef]
  92. da Silva, R.R.; Ramalho, T.C.; Santos, J.M.; Figueroa-Villar, J.D. Reply to “Comment on the Paper ‘On the Limits of Highest-Occupied Molecular Orbital Driven Reactions: The Frontier Effective-for-Reaction Molecular Orbital Concept’”. J. Phys. Chem. A 2006, 110, 10653–10654. [Google Scholar] [CrossRef]
  93. Dewar, M.J.S. A Critique of Frontier Orbital Theory. J. Mol. Struct. Theochem. 1989, 200, 301–323. [Google Scholar] [CrossRef]
  94. Anderson, J.S.M.; Melin, J.; Ayers, P.W. Conceptual Density-Functional Theory for General Chemical Reactions, Including Those That Are Neither Charge- nor Frontier-Orbital-Controlled. 2. Application to Molecules Where Frontier Molecular Orbital Theory Fails. J. Chem. Theory Comput. 2007, 3, 375–389. [Google Scholar] [CrossRef]
  95. Svatunek, D.; Denk, C.; Mikula, H. A Computational Model to Predict the Diels–Alder Reactivity of Aryl/Alkyl-Substituted Tetrazines. Monatshefte Chem.-Chem. Mon. 2018, 149, 833–837. [Google Scholar] [CrossRef]
  96. Sjoberg, P.; Murray, J.S.; Brinck, T.; Politzer, P. Average Local Ionization Energies on the Molecular Surfaces of Aromatic Systems as Guides to Chemical Reactivity. Can. J. Chem. 1990, 68, 1440–1443. [Google Scholar] [CrossRef]
  97. Brinck, T.; Liljenberg, M. The Use of Quantum Chemistry for Mechanistic Analyses of SEAr Reactions. In Arene Chemistry; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2015; pp. 83–105. ISBN 978-1-118-75488-7. [Google Scholar]
  98. Brinck, T.; Stenlid, J.H. The Molecular Surface Property Approach: A Guide to Chemical Interactions in Chemistry, Medicine, and Material Science. Adv. Theory Simul. 2019, 2, 1800149. [Google Scholar] [CrossRef]
  99. Jakobušić Brala, C.; Fabijanić, I.; Karković Marković, A.; Pilepić, V. The Average Local Ionization Energy and Fukui Function of L-Ascorbate, the Local Reactivity Descriptors of Antioxidant Reactivity. Comput. Theor. Chem. 2014, 1049, 1–6. [Google Scholar] [CrossRef]
  100. Fabijanić, I.; Jakobušić Brala, C.; Pilepić, V. The DFT Local Reactivity Descriptors of α-Tocopherol. J. Mol. Model. 2015, 21, 99. [Google Scholar] [CrossRef] [PubMed]
  101. Lakshminarayanan, S.; Jeyasingh, V.; Murugesan, K.; Selvapalam, N.; Dass, G. Molecular Electrostatic Potential (MEP) Surface Analysis of Chemo Sensors: An Extra Supporting Hand for Strength, Selectivity & Non-Traditional Interactions. J. Photochem. Photobiol. 2021, 6, 100022. [Google Scholar]
  102. Liu, S.; Pedersen, L.G. Estimation of Molecular Acidity via Electrostatic Potential at the Nucleus and Valence Natural Atomic Orbitals. J. Phys. Chem. A 2009, 113, 3648–3655. [Google Scholar] [CrossRef] [PubMed]
  103. Rajan, V.K.; Muraleedharan, K. A Computational Investigation on the Structure, Global Parameters and Antioxidant Capacity of a Polyphenol, Gallic Acid. Food Chem. 2017, 220, 93–99. [Google Scholar] [CrossRef] [PubMed]
  104. Vikramaditya, T.; Chai, J.-D.; Lin, S.-T. Impact of Non-Empirically Tuning the Range-Separation Parameter of Long-Range Corrected Hybrid Functionals on Ionization Potentials, Electron Affinities, and Fundamental Gaps. J. Comput. Chem. 2018, 39, 2378–2384. [Google Scholar] [CrossRef]
  105. Fassihi, A.; Hasanzadeh, F.; Attar, A.M.; Saghaie, L.; Mohammadpour, M. Synthesis and Evaluation of Antioxidant Activity of Some Novel Hydroxypyridinone Derivatives: A DFT Approach for Explanation of Their Radical Scavenging Activity. Res. Pharm. Sci. 2020, 15, 515–528. [Google Scholar] [CrossRef]
  106. Pearson, R.G. Chemical Hardness: Applications from Molecules to Solids; Wiley: Hoboken, NJ, USA, 1998; ISBN 978-3-527-29482-4. [Google Scholar]
  107. Yang, W.; Parr, R.G. Hardness, Softness, and the Fukui Function in the Electronic Theory of Metals and Catalysis. Proc. Natl. Acad. Sci. USA 1985, 82, 6723–6726. [Google Scholar] [CrossRef]
  108. hossein Asgarshamsi, M.; Fassihi, A.; Hassanzadeh, F.; Saghaei, L.; Attar, A.M.; Mohammad-Beigi, H. Synthesis, Antioxidant Activity, and Density Functional Theory Study of Some Novel 4-[(Benzo[d]Thiazol-2-Ylimino) Methyl] Phenol Derivatives: A Comparative Approach for the Explanation of Their Radical Scavenging Activities. Res. Pharm. Sci. 2020, 16, 35–47. [Google Scholar]
  109. Foresman, J.B. Computational Chemistry:  A Practical Guide for Applying Techniques to Real World Problems by David Young (Cytoclonal Pharmaceutics Inc.). J. Am. Chem. Soc. 2001, 123, 10142–10143. [Google Scholar] [CrossRef]
  110. Sheela, N.R.; Muthu, S.; Sampathkrishnan, S. Molecular Orbital Studies (Hardness, Chemical Potential and Electrophilicity), Vibrational Investigation and Theoretical NBO Analysis of 4-4′-(1H-1,2,4-Triazol-1-Yl Methylene)-Dibenzonitrile Based on Abinitio and DFT Methods. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2014, 120, 237–251. [Google Scholar] [CrossRef]
  111. Gázquez, J.L.; Cedillo, A.; Vela, A. Electrodonating and Electroaccepting Powers. J. Phys. Chem. A 2007, 111, 1966–1970. [Google Scholar] [CrossRef]
  112. Zheng, Y.-Z.; Chen, D.-F.; Deng, G.; Guo, R. The Substituent Effect on the Radical Scavenging Activity of Apigenin. Molecules 2018, 23, 1989. [Google Scholar] [CrossRef]
  113. Pal, R.; Chattaraj, P.K. Chemical Reactivity from a Conceptual Density Functional Theory Perspective. J. Indian Chem. Soc. 2021, 98, 100008. [Google Scholar] [CrossRef]
  114. Weinhold, F. Discovering Chemistry with Natural Bond Orbitals; John Wiley & Sons: Hoboken, NJ, USA, 2012; ISBN 978-1-118-22919-4. [Google Scholar]
  115. Weinhold, F.; Landis, C.R.; Glendening, E.D. What Is NBO Analysis and How Is It Useful? Int. Rev. Phys. Chem. 2016, 35, 399–440. [Google Scholar] [CrossRef]
  116. Glendening, E.D.; Weinhold, F. Resonance Natural Bond Orbitals: Efficient Semilocalized Orbitals for Computing and Visualizing Reactive Chemical Processes. Chem. Theory Comput. 2019, 15, 916–921. [Google Scholar] [CrossRef] [PubMed]
  117. Weinhold, F. Natural Bond Orbital Analysis: A Critical Overview of Relationships to Alternative Bonding Perspectives. Comput. Chem. 2012, 33, 2363–2379. [Google Scholar] [CrossRef] [PubMed]
  118. Pahari, B.; Chakraborty, S.; Chaudhuri, S.; Sengupta, B.; Sengupta, P.K. Binding and Antioxidant Properties of Therapeutically Important Plant Flavonoids in Biomembranes: Insights from Spectroscopic and Quantum Chemical Studies. Chem. Phys. Lipids 2012, 165, 488–496. [Google Scholar] [CrossRef] [PubMed]
  119. Sun, C.; Cao, B.; Yin, H.; Shi, Y. Relationship between ESIPT Properties and Antioxidant Activities of 5-Hydroxyflavone Derivates. Chin. Phys. B 2020, 29, 058202. [Google Scholar] [CrossRef]
  120. Daday, C.; Manolescu, A.; Marinescu, D.C.; Gudmundsson, V. Electronic Charge and Spin Density Distribution in a Quantum Ring with Spin-Orbit and Coulomb Interactions. Phys. Rev. B 2011, 84, 115311. [Google Scholar] [CrossRef]
  121. Brehm, P.C.; Frontera, A.; Streubel, R. On Metal Coordination of Neutral Open-Shell P-Ligands Focusing on Phosphanoxyls, Their Electron Residence and Reactivity. Chem. Commun. 2022, 58, 6270–6279. [Google Scholar] [CrossRef]
  122. Pluta, T.; Kolaski, M.; Medved’, M.; Budzák, Š. Dipole Moment and Polarizability of the Low-Lying Excited States of Uracil. Chem. Phys. Lett. 2012, 546, 24–29. [Google Scholar] [CrossRef]
  123. Glevitzky, I.; Dumitrel, G.-A.; Mirel, G.; Pasca, M.B.; Otřísal, P.; Bungau, S.; Cioca, G.; Carmen, P.; Popa, M. Statistical Analysis of the Relationship Between Antioxidant Activity and the Structure of Flavonoid Compounds. Rev. Chim. 2019, 70, 7497. [Google Scholar] [CrossRef]
  124. Bentz, E.N.; Pomilio, A.B.; Lobayan, R.M. Exploratory Conformational Study of (+)-Catechin. Modeling of the Polarizability and Electric Dipole Moment. J. Mol. Model. 2014, 20, 2522. [Google Scholar] [CrossRef]
  125. Lobayan, R.M.; Bentz, E.N.; Jubert, A.H.; Pomilio, A.B. Structural and Electronic Properties of Z Isomers of (4α→6´´,2α→O→1´´)-Phenylflavans Substituted with R = H, OH and OCH3 Calculated in Aqueous Solution with PCM Solvation Model. J. Mol. Model. 2012, 18, 1667–1676. [Google Scholar] [CrossRef]
  126. Chermette, H. Density Functional Theory: A Powerful Tool for Theoretical Studies in Coordination Chemistry. Coord. Chem. Rev. 1998, 178–180, 699–721. [Google Scholar] [CrossRef]
  127. Martínez-Araya, J.I. Why Is the Dual Descriptor a More Accurate Local Reactivity Descriptor than Fukui Functions? J. Math. Chem. 2015, 53, 451–465. [Google Scholar] [CrossRef]
  128. Yang, Z.-Z.; Wang, C.-S. Atom-Bond Electronegativity Equalization Method and Its Applications Based on Density Functional Theory. J. Theor. Comput. Chem. 2003, 2, 273–299. [Google Scholar] [CrossRef]
  129. Li, Y.; Yang, Y.; Lei, J.; Liu, W.; Tong, M.; Liang, J. The Degradation Pathways of Carbamazepine in Advanced Oxidation Process: A Mini Review Coupled with DFT Calculation. Sci. Total Environ. 2021, 779, 146498. [Google Scholar] [CrossRef]
  130. Bardak, F. Experimental and DFT Analysis of Structural and Spectroscopic Features of Nitroterephthalic Acid, and Computational Insights into Its Molecular Interactions with hER-α via Molecular Docking. J. Mol. Struct. 2019, 1175, 458–470. [Google Scholar] [CrossRef]
  131. Abraham, C.S.; Muthu, S.; Prasana, J.C.; Armaković, S.; Armaković, S.J.; Geoffrey, B. Computational Evaluation of the Reactivity and Pharmaceutical Potential of an Organic Amine: A DFT, Molecular Dynamics Simulations and Molecular Docking Approach. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2019, 222, 117188. [Google Scholar] [CrossRef]
  132. Djeradi, H.; Rahmouni, A.; Cheriti, A. Antioxidant Activity of Flavonoids: A QSAR Modeling Using Fukui Indices Descriptors. J. Mol. Model. 2014, 20, 2476. [Google Scholar] [CrossRef]
  133. Wiberg, K.B.; Rablen, P.R. Atomic Charges. J. Org. Chem. 2018, 83, 15463–15469. [Google Scholar] [CrossRef]
  134. Bultinck, P.; Langenaeker, W.; Lahorte, P.; De Proft, F.; Geerlings, P.; Waroquier, M.; Tollenaere, J.P. The Electronegativity Equalization Method I: Parametrization and Validation for Atomic Charge Calculations. J. Phys. Chem. A 2002, 106, 7887–7894. [Google Scholar] [CrossRef]
  135. Amic, D.; Davidovic-Amic, D.; Beslo, D.; Rastija, V.; Lucic, B.; Trinajstic, N. SAR and QSAR of the Antioxidant Activity of Flavonoids. Curr. Med. Chem. 2007, 14, 827–845. [Google Scholar] [CrossRef]
  136. Rajendran, M.; Ravichandran, R.; Devapiriam, D. Electronic Description of Few Selected Flavonoids by Theoretical Study. Int. J. Comput. Appl. 2013, 77, 18–25. [Google Scholar] [CrossRef]
  137. Šimunková, M.; Biela, M.; Štekláč, M.; Hlinčík, A.; Klein, E.; Malček, M. Cu(II) Complexes of Flavonoids in Solution: Impact of the Cu(II) Ion on the Antioxidant and DNA-Intercalating Properties. J. Mol. Liq. 2022, 359, 119230. [Google Scholar] [CrossRef]
  138. Lu, T.; Chen, F. Comparison of Computational Methods for Atomic Charges. Acta Phys.-Chim. Sin. 2011, 28, 1–18. [Google Scholar]
  139. Ionescu, C.-M.; Sehnal, D.; Falginella, F.L.; Pant, P.; Pravda, L.; Bouchal, T.; Svobodová Vařeková, R.; Geidl, S.; Koča, J. AtomicChargeCalculator: Interactive Web-Based Calculation of Atomic Charges in Large Biomolecular Complexes and Drug-like Molecules. J. Cheminform. 2015, 7, 50. [Google Scholar] [CrossRef] [PubMed]
  140. Miličević, A.; Miletić, G.I.; Novak Jovanović, I. Electrochemical Oxidation of Flavonoids: PM6 and DFT for Elucidating Electronic Changes and Modelling Oxidation Potential (Part II). J. Mol. Liq. 2019, 295, 111730. [Google Scholar] [CrossRef]
  141. Park, S.; Kim, M.; Lin, Y.; Hong, M.; Nam, G.; Mieczkowski, A.; Kardos, J.; Lee, Y.-H.; Hee Lim, M. Designing Multi-Target-Directed Flavonoids: A Strategic Approach to Alzheimer’s Disease. Chem. Sci. 2023, 14, 9293–9305. [Google Scholar] [CrossRef]
  142. Lu, T.; Chen, F. Multiwfn: A Multifunctional Wavefunction Analyzer. J. Comput. Chem. 2012, 33, 580–592. [Google Scholar] [CrossRef] [PubMed]
  143. Agwupuye, J.A.; Louis, H.; Unimuke, T.O.; David, P.; Ubana, E.I.; Moshood, Y.L. Electronic Structure Investigation of the Stability, Reactivity, NBO Analysis, Thermodynamics, and the Nature of the Interactions in Methyl-Substituted Imidazolium-Based Ionic Liquids. J. Mol. Liq. 2021, 337, 116458. [Google Scholar] [CrossRef]
  144. Gázquez, J.L. Perspectives on the Density Functional Theory of Chemical Reactivity. J. Mex. Chem. Soc. 2008, 52, 3–10. [Google Scholar]
  145. Reina, M.; Castañeda-Arriaga, R.; Perez-Gonzalez, A.; Guzman-Lopez, E.G.; Tan, D.-X.; Reiter, R.J.; Galano, A. A Computer-Assisted Systematic Search for Melatonin Derivatives with High Potential as Antioxidants. Melatonin Res. 2018, 1, 27–58. [Google Scholar] [CrossRef]
  146. Castro-González, L.M.; Alvarez-Idaboy, J.R.; Galano, A. Computationally Designed Sesamol Derivatives Proposed as Potent Antioxidants. ACS Omega 2020, 5, 9566–9575. [Google Scholar] [CrossRef] [PubMed]
  147. Martínez, A.; Vargas, R.; Galano, A. What Is Important to Prevent Oxidative Stress? A Theoretical Study on Electron-Transfer Reactions between Carotenoids and Free Radicals. J. Phys. Chem. B 2009, 113, 12113–12120. [Google Scholar] [CrossRef] [PubMed]
  148. Burdett, J.K.; McCormick, T.A. Electron Localization in Molecules and Solids:  The Meaning of ELF. J. Phys. Chem. A 1998, 102, 6366–6372. [Google Scholar] [CrossRef]
  149. Savin, A.; Nesper, R.; Wengert, S.; Fässler, T.F. ELF: The Electron Localization Function. Angew. Chem. Int. Ed. Engl. 1997, 36, 1808–1832. [Google Scholar] [CrossRef]
  150. Silvi, B.; Fourré, I.; Alikhani, M.E. The Topological Analysis of the Electron Localization Function. A Key for a Position Space Representation of Chemical Bonds. Monatshefte Chem.-Chem. Mon. 2005, 136, 855–879. [Google Scholar] [CrossRef]
  151. Matito, E.; Solà, M. The Role of Electronic Delocalization in Transition Metal Complexes from the Electron Localization Function and the Quantum Theory of Atoms in Molecules Viewpoints. Coord. Chem. Rev. 2009, 253, 647–665. [Google Scholar] [CrossRef]
  152. Sadovskiĭ, M.V. Electron Localization in Disordered Systems: Critical Behavior and Macroscopic Manifestations. Sov. Phys. Uspekhi 1981, 24, 96. [Google Scholar] [CrossRef]
  153. Schmider, H.L.; Becke, A.D. Chemical Content of the Kinetic Energy Density. J. Mol. Struct. Theochem. 2000, 527, 51–61. [Google Scholar] [CrossRef]
  154. Jacobsen, H. Localized-Orbital Locator (LOL) Profiles of Chemical Bonding. Can. J. Chem. 2008, 86, 695–702. [Google Scholar] [CrossRef]
  155. Divya, P.; Muthuraja, P.; Dhandapani, M.; Jothy, V.B. Hydrogen Bonding Interactions on Molecular Properties of Pesticidal Compound 4-Nitrophthalic Acid: Experimental Density Functional Theory Computations, Electron Localized Function, Localized Orbital Locator Analysis and Molecular Docking Scrutiny. Spectrosc. Lett. 2022, 55, 362–372. [Google Scholar] [CrossRef]
  156. Amić, D.; Stepanić, V.; Lučić, B.; Marković, Z.; Dimitrić Marković, J.M. PM6 Study of Free Radical Scavenging Mechanisms of Flavonoids: Why Does O–H Bond Dissociation Enthalpy Effectively Represent Free Radical Scavenging Activity? J. Mol. Model. 2013, 19, 2593–2603. [Google Scholar] [CrossRef]
  157. Foti, M.C. Antioxidant Properties of Phenols. J. Pharm. Pharmacol. 2007, 59, 1673–1685. [Google Scholar] [CrossRef]
  158. Olszowy, M. What Is Responsible for Antioxidant Properties of Polyphenolic Compounds from Plants? Plant Physiol. Biochem. 2019, 144, 135–143. [Google Scholar] [CrossRef]
  159. Klein, E.; Rimarčík, J.; Senajová, E.; Vagánek, A.; Lengyel, J. Deprotonation of Flavonoids Severely Alters the Thermodynamics of the Hydrogen Atom Transfer. Comput. Theor. Chem. 2016, 1085, 7–17. [Google Scholar] [CrossRef]
  160. Yang, Y.; Zhang, Y.; Gu, D.; Liu, C.; Wang, Y.; Tang, S.; Yin, Y.; Tian, J. Fermentation of Robinia Pseudoacacia Flower for Improving the Antioxidation: Optimized Conditions, Active Composition, Mechanism, and Biotransformation Process. Prep. Biochem. Biotechnol. 2023, 53, 1224–1236. [Google Scholar] [CrossRef] [PubMed]
  161. Lespade, L.; Bercion, S. Theoretical investigation of the effect of sugar substitution on the antioxidant properties of flavonoids. Free Radic. Res. 2012, 46, 346–358. [Google Scholar] [CrossRef] [PubMed]
  162. Marković, Z.; Milenković, D.; Đorović, J.; Dimitrić Marković, J.M.; Stepanić, V.; Lučić, B.; Amić, D. Free Radical Scavenging Activity of Morin 2′-O− Phenoxide Anion. Food Chem. 2012, 135, 2070–2077. [Google Scholar] [CrossRef]
  163. Zheng, Y.-Z.; Deng, G.; Zhang, Y.-C. Multiple Free Radical Scavenging Reactions of Flavonoids. Dyes Pigment. 2022, 198, 109877. [Google Scholar] [CrossRef]
  164. Amić, A.; Marković, Z.; Dimitrić Marković, J.M.; Stepanić, V.; Lučić, 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. [Google Scholar] [CrossRef] [PubMed]
  165. Mittal, A.; Vashistha, V.K.; Das, D.K. Recent Advances in the Antioxidant Activity and Mechanisms of Chalcone Derivatives: A Computational Review. Free Radic. Res. 2022, 56, 378–397. [Google Scholar] [CrossRef] [PubMed]
  166. Parson, W.W. Reorganization energies, entropies, and free energy surfaces for electron transfer. J. Phys. Chem. B 2021, 125, 7940–7945. [Google Scholar] [CrossRef] [PubMed]
  167. Spiegel, M.; Andruniów, T.; Sroka, Z. Flavones’ and Flavonols’ Antiradical Structure–Activity Relationship—A Quantum Chemical Study. Antioxidants 2020, 9, 461. [Google Scholar] [CrossRef]
  168. Satoh, T.; McKercher, S.R.; Lipton, S.A. Reprint of: Nrf2/ARE-mediated antioxidant actions of pro-electrophilic drugs. Free Radic. Biol. Med. 2014, 66, 45–57. [Google Scholar] [CrossRef]
  169. Zhang, M.; Vervoort, L.; Moalin, M.; Mommers, A.; Douny, C.; den Hartog, G.J.M.; Haenen, G.R.M.M. The Chemical Reactivity of (-)-Epicatechin Quinone Mainly Resides in Its B-Ring. Free Radic. Biol. Med. 2018, 124, 31–39. [Google Scholar] [CrossRef] [PubMed]
  170. Sroka, Z.; Żbikowska, B.; Hładyszowski, J. The Antiradical Activity of Some Selected Flavones and Flavonols. Experimental and Quantum Mechanical Study. J. Mol. Model. 2015, 21, 307. [Google Scholar] [CrossRef] [PubMed]
  171. Parcheta, M.; Świsłocka, R.; Orzechowska, S.; Akimowicz, M.; Choińska, R.; Lewandowski, W. Recent Developments in Effective Antioxidants: The Structure and Antioxidant Properties. Materials 2021, 14, 1984. [Google Scholar] [CrossRef]
  172. Jabbari, M.; Jabbari, A. Antioxidant Potential and DPPH Radical Scavenging Kinetics of Water-Insoluble Flavonoid Naringenin in Aqueous Solution of Micelles. Colloids Surf. A Physicochem. Eng. Asp. 2016, 489, 392–399. [Google Scholar] [CrossRef]
  173. Musialik, M.; Kuzmicz, R.; Pawłowski, T.S.; Litwinienko, G. Acidity of Hydroxyl Groups: An Overlooked Influence on Antiradical Properties of Flavonoids. J. Org. Chem. 2009, 74, 2699–2709. [Google Scholar] [CrossRef] [PubMed]
  174. Trang, N.V.; Thuy, P.T.; Thanh, D.T.M.; Son, N.T. Benzofuran–Stilbene Hybrid Compounds: An Antioxidant Assessment—A DFT Study. RSC Adv. 2021, 11, 12971–12980. [Google Scholar] [CrossRef]
  175. Thuy, P.T.; Trang, N.V.; Son, N.T. Antioxidation of 2-Phenylbenzofuran Derivatives: Structural-Electronic Effects and Mechanisms. RSC Adv. 2020, 10, 6315–6332. [Google Scholar] [CrossRef] [PubMed]
  176. Thuy, P.T.; Son, N.T. Thermodynamic and Kinetic Studies on Antioxidant Capacity of Amentoflavone: A DFT (Density Functional Theory) Computational Approach. Free Radic. Res. 2022, 56, 526–535. [Google Scholar] [CrossRef] [PubMed]
  177. Erdoğan, Ş.; Özbakır Işın, D. A DFT Study on OH Radical Scavenging Activities of Eriodictyol, Isosakuranetin and Pinocembrin. J. Biomol. Struct. Dyn. 2022, 40, 10802–10811. [Google Scholar] [CrossRef]
  178. Vo, Q.V.; Nam, P.C.; Thong, N.M.; Trung, N.T.; Phan, C.-T.D.; Mechler, A. Antioxidant Motifs in Flavonoids: O–H versus C–H Bond Dissociation. ACS Omega 2019, 4, 8935–8942. [Google Scholar] [CrossRef]
  179. Milenković, D.; Dimić, D.; Avdović, E.; Simijonović, D.; Vojinović, R.; Marković, Z. A Thermodynamic and Kinetic HO Radical Scavenging Study and Protein Binding of Baicalein. J. Chem. Thermodyn. 2023, 185, 107110. [Google Scholar] [CrossRef]
  180. Costentin, C. Electrochemical Approach to the Mechanistic Study of Proton-Coupled Electron Transfer. Chem. Rev. 2008, 108, 2145–2179. [Google Scholar] [CrossRef] [PubMed]
  181. Truhlar, D.G.; Garrett, B.C.; Klippenstein, S.J. Current Status of Transition-State Theory. J. Phys. Chem. 1996, 100, 12771–12800. [Google Scholar] [CrossRef]
  182. Otero-de-la-Roza, A.; Johnson, E.R.; Luaña, V. Critic2: A Program for Real-Space Analysis of Quantum Chemical Interactions in Solids. Comput. Phys. Commun. 2014, 185, 1007–1018. [Google Scholar] [CrossRef]
  183. Brovarets’, O.O.; Hovorun, D.M. Energy of the CH⋯O H-Bonds and Others Specific Contacts in the Q Molecule: QM/QTAIM Approximation Formulas. J. Mol. Liq. 2020, 313, 113456. [Google Scholar] [CrossRef]
  184. Šimunková, M.; Valko, M.; Bučinský, L.; Malček, M. Structure Functionality Relationship of Flavonoids (Myricetin, Morin, Taxifolin and 3′,4′-Dihydroxyflavone). A Computational Study via the Cupric Ion Probe. J. Mol. Struct. 2020, 1222, 128923. [Google Scholar] [CrossRef]
  185. Zhang, M.; Zhu, X. Computational Investigation of Flavonol-Based GLP-1R Agonists Using DFT Calculations and Molecular Docking. Comput. Theor. Chem. 2020, 1190, 113005. [Google Scholar] [CrossRef]
  186. Zannou, O.; Koca, I. Greener extraction of anthocyanins and antioxidant activity from blackberry (Rubus spp) using natural deep eutectic solvents. LWT 2022, 158, 113184. [Google Scholar] [CrossRef]
  187. Zheng, Y.-Z.; Deng, G.; Guo, R.; Fu, Z.-M.; Chen, D.-F. The Influence of the H5⋯OC4 Intramolecular Hydrogen-Bond (IHB) on the Antioxidative Activity of Flavonoid. Phytochemistry 2019, 160, 19–24. [Google Scholar] [CrossRef]
  188. Zhong, Y.; Chen, Y.; Feng, X.; Sun, Y.; Cui, S.; Li, X.; Jin, X.; Zhao, G. Hydrogen-Bond Facilitated Intramolecular Proton Transfer in Excited State and Fluorescence Quenching Mechanism of Flavonoid Compounds in Aqueous Solution. J. Mol. Liq. 2020, 302, 112562. [Google Scholar] [CrossRef]
  189. Ji, F.; Guo, Y.; Wang, M.; Wang, C.; Wu, Z.; Wang, S.; Wang, H.; Feng, X.; Zhao, G. New Insights into ESIPT Mechanism of Three Sunscreen Compounds in Solution: A Combined Experimental and Theoretical Study. Colloids Surf. B Biointerfaces 2021, 207, 112039. [Google Scholar] [CrossRef]
  190. Wu, P.; Chaudret, R.; Hu, X.; Yang, W. Noncovalent Interaction Analysis in Fluctuating Environments. J. Chem. Theory Comput. 2013, 9, 2226–2234. [Google Scholar] [CrossRef]
  191. Fischer, E. Einfluss der Configuration auf die Wirkung der Enzyme. Berichte Dtsch. Chem. Ges. 1894, 27, 2985–2993. [Google Scholar] [CrossRef]
  192. Koshland, D.E. Application of a Theory of Enzyme Specificity to Protein Synthesis. Proc. Natl. Acad. Sci. USA 1958, 44, 98–104. [Google Scholar] [CrossRef]
  193. Aderinwale, T.; Christoffer, C.W.; Sarkar, D.; Alnabati, E.; Kihara, D. Computational Structure Modeling for Diverse Categories of Macromolecular Interactions. Curr. Opin. Struct. Biol. 2020, 64, 1–8. [Google Scholar] [CrossRef] [PubMed]
  194. Miller, E.B.; Murphy, R.B.; Sindhikara, D.; Borrelli, K.W.; Grisewood, M.J.; Ranalli, F.; Dixon, S.L.; Jerome, S.; Boyles, N.A.; Day, T.; et al. Reliable and Accurate Solution to the Induced Fit Docking Problem for Protein–Ligand Binding. J. Chem. Theory Comput. 2021, 17, 2630–2639. [Google Scholar] [CrossRef] [PubMed]
  195. Hasan, A.E.Z.; Ambarsari, L.; Hariana, K. Children’s Opinion on Vegetables Consumption: A Qualitative Study on School-Agers in City of Semarang. Indones. J. Appl. Res. IJAR 2021, 2, 126–143. [Google Scholar]
  196. Sulimov, V.B.; Kutov, D.C.; Sulimov, A.V. Advances in Docking. Curr. Med. Chem. 2019, 26, 7555–7580. [Google Scholar] [CrossRef]
  197. Vásquez-Espinal, A.; Yañez, O.; Osorio, E.; Areche, C.; García-Beltrán, O.; Ruiz, L.M.; Cassels, B.K.; Tiznado, W. Theoretical Study of the Antioxidant Activity of Q Oxidation Products. Front. Chem. 2019, 7, 818. [Google Scholar] [CrossRef]
  198. Kamel, E.M.; Bin-Ammar, A.; El-Bassuony, A.A.; Alanazi, M.M.; Altharawi, A.; Ahmeda, A.F.; Alanazi, A.S.; Lamsabhi, A.M.; Mahmoud, A.M. Molecular Modeling and DFT Studies on the Antioxidant Activity of Centaurea Scoparia Flavonoids and Molecular Dynamics Simulation of Their Interaction with β-Lactoglobulin. RSC Adv. 2023, 13, 12361–12374. [Google Scholar] [CrossRef]
  199. Zhang, M.; Li, Y.; Zhu, T. The Theoretical Investigation of Monohydroxy Flavone: A Combined DFT and Molecular Docking Study. J. Mol. Struct. 2022, 1250, 131823. [Google Scholar] [CrossRef]
  200. Erdoğan, Ş.; Özbakır Işın, D. Molecular Docking and A DFT Study on the Antiradical Activity of Naringenin and Hesperetin with Nitric Oxide, Peroxy, and Methoxy Radicals. J. Phys. Org. Chem. 2023, 36, e4479. [Google Scholar] [CrossRef]
  201. Bamdad, F.; Farrokhpour, H.; Ashrafizaadeh, M.; Najafi, B. Decomposition of the Interaction Energy of Several Flavonoids with Escherichia Coli DNA Gyr Using the SAPT (DFT) Method: The Relation between the Interaction Energy Components, Ligand Structure, and Biological Activity. Biochim. Biophys. Acta BBA-Gen. Subj. 2022, 1866, 130111. [Google Scholar] [CrossRef] [PubMed]
  202. Ding, B.; Yu, Y.; Geng, S.; Liu, B.; Hao, Y.; Liang, G. Computational Methods for the Interaction between Cyclodextrins and Natural Compounds: Technology, Benefits, Limitations, and Trends. J. Agric. Food Chem. 2022, 70, 2466–2482. [Google Scholar] [CrossRef] [PubMed]
  203. Raeessi-Babaheydari, E.; Farhadian, S.; Shareghi, B. Evaluation of Interaction between Citrus Flavonoid, Naringenin, and Pepsin Using Spectroscopic Analysis and Docking Simulation. J. Mol. Liq. 2021, 339, 116763. [Google Scholar] [CrossRef]
  204. Singh, A.; Vanga, S.K.; Orsat, V.; Raghavan, V. Application of Molecular Dynamic Simulation to Study Food Proteins: A Review. Crit. Rev. Food Sci. Nutr. 2018, 58, 2779–2789. [Google Scholar] [CrossRef]
  205. Chen, Q.; Xiao, Y.; Zhang, W.; Mu, W. Current Methods and Applications in Computational Protein Design for Food Industry. Crit. Rev. Food Sci. Nutr. 2020, 60, 3259–3270. [Google Scholar] [CrossRef]
  206. Chung, L.W.; Sameera, W.M.C.; Ramozzi, R.; Page, A.J.; Hatanaka, M.; Petrova, G.P.; Harris, T.V.; Li, X.; Ke, Z.; Liu, F.; et al. The ONIOM Method and Its Applications. Chem. Rev. 2015, 115, 5678–5796. [Google Scholar] [CrossRef] [PubMed]
  207. Fuguet, E.; Ràfols, C.; Mañé, M.; Ruiz, R.; Bosch, E. Acidity Constants of Hydroxyl Groups Placed in Several Flavonoids: Two Flavanones, Two Flavones and Five Flavonols. Talanta 2023, 253, 124096. [Google Scholar] [CrossRef]
  208. Toth, A.M.; Liptak, M.D.; Phillips, D.L.; Shields, G.C. Accurate Relative pKa Calculations for Carboxylic Acids Using Complete Basis Set and Gaussian-n Models Combined with Continuum Solvation Methods. J. Chem. Phys. 2001, 114, 4595–4606. [Google Scholar] [CrossRef]
  209. Ribeiro, A.R.; Schmidt, T.C. Determination of Acid Dissociation Constants (pKa) of Cephalosporin Antibiotics: Computational and Experimental Approaches. Chemosphere 2017, 169, 524–533. [Google Scholar] [CrossRef] [PubMed]
  210. Lemańska, K.; Szymusiak, H.; Tyrakowska, B.; Zieliński, R.; Soffers, A.E.M.F.; Rietjens, I.M.C.M. The Influence of pH on Antioxidant Properties and the Mechanism of Antioxidant Action of Hydroxyflavones. Free Radic. Biol. Med. 2001, 31, 869–881. [Google Scholar] [CrossRef] [PubMed]
  211. XLOGP3 (v3.2.2). Available online: http://www.sioc-ccbg.ac.cn/skins/ccbgwebsite/software/xlogp3/ (accessed on 18 March 2024).
  212. On-Line Lipophilicity/Aqueous Solubility Calculation Software. Available online: https://vcclab.org/lab/alogps/ (accessed on 18 March 2024).
  213. Marvin. Available online: https://chemaxon.com/marvin (accessed on 18 March 2024).
  214. Chemaxon. Available online: https://chemaxon.com (accessed on 18 March 2024).
  215. Hitchcock, S.A.; Pennington, L.D. Structure−Brain Exposure Relationships. J. Med. Chem. 2006, 49, 7559–7583. [Google Scholar] [CrossRef] [PubMed]
  216. Berger, T.A.; Berger, B.K.; Kogelman, K. 8.18 Chromatographic Separations and Analysis: Supercritical Fluid Chromatography for Chiral Analysis and Semi-preparative Purification. Compr. Chirality 2012, 8, 354–392. [Google Scholar]
  217. Bytheway, I.; Darley, M.G.; Popelier, P.L.A. The Calculation of Polar Surface Area from First Principles: An Application of Quantum Chemical Topology to Drug Design. ChemMedChem 2008, 3, 445–453. [Google Scholar] [CrossRef] [PubMed]
  218. Markowicz-Piasecka, M.; Markiewicz, A.; Darłak, P.; Sikora, J.; Adla, S.K.; Bagina, S.; Huttunen, K.M. Current chemical, biological, and physiological views in the development of successful brain-targeted pharmaceutics. Neurotherapeutics 2022, 19, 942–976. [Google Scholar] [CrossRef] [PubMed]
  219. Ronsisvalle, S.; Panarello, F.; Longhitano, G.; Siciliano, E.A.; Montenegro, L.; Panico, A. Natural Flavones and Flavonols: Relationships among Antioxidant Activity, Glycation, and Metalloproteinase Inhibition. Cosmetics 2020, 7, 71. [Google Scholar] [CrossRef]
  220. Lemmens, K.J.; Vrolijk, M.F.; Bouwman, F.G.; Van der Vijgh, W.J.; Bast, A.; Haenen, G.R. The minor structural difference between the antioxidants quercetin and 4’O-methylquercetin has a major impact on their selective thiol toxicity. Int. J. Mol. Sci. 2014, 15, 7475–7484. [Google Scholar] [CrossRef]
Figure 1. The basic skeleton of the subclasses to which the investigated compounds belong, along with the number of compounds in each subclass that have been studied over the past five years [9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82]. In Table S1 of the Supplementary Materials, the compound names and corresponding references are given.
Figure 1. The basic skeleton of the subclasses to which the investigated compounds belong, along with the number of compounds in each subclass that have been studied over the past five years [9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82]. In Table S1 of the Supplementary Materials, the compound names and corresponding references are given.
Molecules 29 02627 g001
Figure 2. The methods employed in the past five years to investigate the electronic structure of flavonoids, along with the proportion of each method used [9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82]. (The percentages less than 2% are not marked here). In Table S1 of the Supplementary Materials, the compound names and corresponding references are given.
Figure 2. The methods employed in the past five years to investigate the electronic structure of flavonoids, along with the proportion of each method used [9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82]. (The percentages less than 2% are not marked here). In Table S1 of the Supplementary Materials, the compound names and corresponding references are given.
Molecules 29 02627 g002
Figure 3. The HOMO map of Q. Calculated with Gaussian 09 package [86] using the DFT method at the M062X [87]/6-311G (d,p) [88] level.
Figure 3. The HOMO map of Q. Calculated with Gaussian 09 package [86] using the DFT method at the M062X [87]/6-311G (d,p) [88] level.
Molecules 29 02627 g003
Figure 4. The surface MEP map of Q. Calculated with Gaussian 09 package [86] using the DFT method at the M062X [87]/6-311G (d,p) [88] level.
Figure 4. The surface MEP map of Q. Calculated with Gaussian 09 package [86] using the DFT method at the M062X [87]/6-311G (d,p) [88] level.
Molecules 29 02627 g004
Figure 5. The SDD of the lone pair electron in the Q-7-O-radical: (a) before and (b) after intramolecular proton transfer from the oxygen at the 3-OH to the oxygen at the 4-carbonyl. Calculated with Gaussian 09 package [86] using the DFT method at the M062X [87]/6-311G (d,p) [88] level.
Figure 5. The SDD of the lone pair electron in the Q-7-O-radical: (a) before and (b) after intramolecular proton transfer from the oxygen at the 3-OH to the oxygen at the 4-carbonyl. Calculated with Gaussian 09 package [86] using the DFT method at the M062X [87]/6-311G (d,p) [88] level.
Molecules 29 02627 g005
Figure 6. Schematic diagram of the DAM.
Figure 6. Schematic diagram of the DAM.
Molecules 29 02627 g006
Figure 7. The NCI analysis map of Q. Calculated with Gaussian 09 package [86] using the DFT method at the M062X [87]/6-311G (d,p) [88] level.
Figure 7. The NCI analysis map of Q. Calculated with Gaussian 09 package [86] using the DFT method at the M062X [87]/6-311G (d,p) [88] level.
Molecules 29 02627 g007
Figure 8. The thermodynamic cycle connects the deprotonation of the acid HA dissolved in water (aq) to that as a gas (g). The values of ΔGgas(H+) and ΔGsolv(H+) are known from experiments (with a sum of 269.0 Kcal/mol). The other gas-free energies and the solvation-free energies require relatively complex calculations using experimental data and different basis sets.
Figure 8. The thermodynamic cycle connects the deprotonation of the acid HA dissolved in water (aq) to that as a gas (g). The values of ΔGgas(H+) and ΔGsolv(H+) are known from experiments (with a sum of 269.0 Kcal/mol). The other gas-free energies and the solvation-free energies require relatively complex calculations using experimental data and different basis sets.
Molecules 29 02627 g008
Figure 9. A way to visualize the comprehensive antioxidant profile of a flavonoid. The various electronic characteristics of an antioxidant are related to those of the very potent antioxidant flavonoid Q. The radius of the red circle is the value of each descriptor for Q. The length of each vector starting from the center of the circle can be used to visualize the value of the denoted descriptor of a particular flavonoid relative to that of Q.
Figure 9. A way to visualize the comprehensive antioxidant profile of a flavonoid. The various electronic characteristics of an antioxidant are related to those of the very potent antioxidant flavonoid Q. The radius of the red circle is the value of each descriptor for Q. The length of each vector starting from the center of the circle can be used to visualize the value of the denoted descriptor of a particular flavonoid relative to that of Q.
Molecules 29 02627 g009
Figure 10. An overarching view on the computational landscape of the antioxidant activity of Q.
Figure 10. An overarching view on the computational landscape of the antioxidant activity of Q.
Molecules 29 02627 g010
Table 1. The conventional GCRPs measure the electron transfer characteristics of antioxidants.
Table 1. The conventional GCRPs measure the electron transfer characteristics of antioxidants.
PropertiesFormulaDescription
Hardness(η)
Softness (S)
η = (IP − EA)/2
S = 1/2η
η reflects the reluctance towards deformation or polarization of the electron cloud under slight perturbation [106,107,108].
Chemical potential (μ)−μ = (IP + EA)/2μ indicates the direction of charge flow. Electrons will migrate from high μ to low μ locations [109,110].
Electronegativity (χ)χ = (IP + EA)/2χ gives the tendency of a molecule to attract electrons.
Electrophilicity (ω)ω = μ2/2ηω reflects the electron-donating ability of a molecule.
ω+ω+ = (IP + 3EA)2/16(IP − EA)ω and ω+ reflect respectively the electron-donating and electron-accepting ability of antioxidants [111,112].
ωω = (3IP + EA)2/16(IP − EA)
Table 2. The reaction equations and enthalpies of the three main mechanisms of flavonoids to directly scavenge free radicals.
Table 2. The reaction equations and enthalpies of the three main mechanisms of flavonoids to directly scavenge free radicals.
MechanismReaction EquationEnthalpy
HATArOHoptimized → ArOoptimized + HBond dissociation enthalpy (BDE)
SPLETArOHoptimized → ArOoptimized + H+Proton affinity (PA)
ArOunoptimized → ArOoptimized + eElectron transfer enthalpy (ETE)
ETPTArOHoptimized → ArOH•+optimized + eIonization potential (IP)
ArOH•+unoptimized → ArOoptimized + H+Proton dissociation enthalpy (PDE)
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, Y.; Li, C.; Li, Z.; Moalin, M.; Hartog, G.J.M.d.; Zhang, M. Computational Chemistry Strategies to Investigate the Antioxidant Activity of Flavonoids—An Overview. Molecules 2024, 29, 2627. https://doi.org/10.3390/molecules29112627

AMA Style

Wang Y, Li C, Li Z, Moalin M, Hartog GJMd, Zhang M. Computational Chemistry Strategies to Investigate the Antioxidant Activity of Flavonoids—An Overview. Molecules. 2024; 29(11):2627. https://doi.org/10.3390/molecules29112627

Chicago/Turabian Style

Wang, Yue, Chujie Li, Zhengwen Li, Mohamed Moalin, Gertjan J. M. den Hartog, and Ming Zhang. 2024. "Computational Chemistry Strategies to Investigate the Antioxidant Activity of Flavonoids—An Overview" Molecules 29, no. 11: 2627. https://doi.org/10.3390/molecules29112627

Article Metrics

Back to TopTop