Next Article in Journal
N-Acetylcysteine Treatment May Compensate Motor Impairments through Dopaminergic Transmission Modulation in a Striatal 6-Hydroxydopamine Parkinson’s Disease Rat Model
Next Article in Special Issue
Cold Physical Plasma-Mediated Fenretinide Prodrug Activation Confers Additive Cytotoxicity in Epithelial Cells
Previous Article in Journal
Gender Differences in Oxidative Stress in Relation to Cancer Susceptibility and Survival
Previous Article in Special Issue
Synthesis and Biological Evaluation of O6-Aminoalkyl-Hispidol Analogs as Multifunctional Monoamine Oxidase-B Inhibitors towards Management of Neurodegenerative Diseases
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Rational Design of Multifunctional Ferulic Acid Derivatives Aimed for Alzheimer’s and Parkinson’s Diseases

by
Eduardo Gabriel Guzmán-López
1,
Miguel Reina
2,
Luis Felipe Hernández-Ayala
1 and
Annia Galano
1,*
1
Departamento de Química, Universidad Autónoma Metropolitana-Iztapalapa, Av. Ferrocarril San Rafael Atlixco 186, Col. Leyes de Reforma 1A Sección, Alcaldía Iztapalapa, Mexico City 09310, Mexico
2
Departamento de Química Inorgánica y Nuclear, Facultad de Química, Universidad Nacional Autónoma de México, Mexico City 04510, Mexico
*
Author to whom correspondence should be addressed.
Antioxidants 2023, 12(6), 1256; https://doi.org/10.3390/antiox12061256
Submission received: 29 April 2023 / Revised: 31 May 2023 / Accepted: 6 June 2023 / Published: 11 June 2023

Abstract

:
Ferulic acid has numerous beneficial effects on human health, which are frequently attributed to its antioxidant behavior. In this report, many of them are reviewed, and 185 new ferulic acid derivatives are computationally designed using the CADMA-Chem protocol. Consequently, their chemical space was sampled and evaluated. To that purpose, selection and elimination scores were used, which are built from a set of descriptors accounting for ADME properties, toxicity, and synthetic accessibility. After the first screening, 12 derivatives were selected and further investigated. Their potential role as antioxidants was predicted from reactivity indexes directly related to the formal hydrogen atom transfer and the single electron transfer mechanisms. The best performing molecules were identified by comparisons with the parent molecule and two references: Trolox and α-tocopherol. Their potential as polygenic neuroprotectors was investigated through the interactions with enzymes directly related to the etiologies of Parkinson’s and Alzheimer’s diseases. These enzymes are acetylcholinesterase, catechol-O-methyltransferase, and monoamine oxidase B. Based on the obtained results, the most promising candidates (FA-26, FA-118, and FA-138) are proposed as multifunctional antioxidants with potential neuroprotective effects. The findings derived from this investigation are encouraging and might promote further investigations on these molecules.

1. Introduction

Oxidative stress (OS) is a harmful multifaceted phenomenon, often referred to as the “chemical silent killer” since no evident symptoms are associated with it. To this day, there is no available test to detect it. Thus, its damaging effects can evolve without any advice to the affected person. Currently, OS represents a major concern linked to the onset and development of hundreds of illnesses. Among the available strategies to lessen OS risks to human health, chemical protection by antioxidant molecules is one of the most effective and studied approaches. Antioxidants can be seen as sacrificial compounds that prevent oxidants from reaching biomolecules. Antioxidants are produced endogenously by the human body and can be acquired through the intake of food and dietary supplements.
Ferulic acid (4-hydroxy-3-methoxy cinnamic acid, FA, Scheme 1) is one of these valuable molecules. It is found in whole grains, grapes, parsley, rhubarb, spinach, cereal seeds, artichoke, and coffee, among many other natural sources [1]. It is a versatile molecule. There are numerous reports on its antioxidant activity [2,3,4,5,6,7,8,9,10,11,12,13,14] as well as on its anti-inflammatory [15,16], antibacterial [17,18,19,20], antiviral [21], anti-thrombotic [22,23], anti-ageing [24,25,26], and antitumoral effects [27,28,29,30,31,32,33,34,35,36,37,38,39,40,41]. It also acts as a cardio protector [42,43,44,45,46,47,48], neuroprotector [49,50,51,52,53,54,55,56,57,58,59], antihypertensive [60,61,62,63], antidepressant [51,64,65,66,67,68,69], hepatoprotector [70,71,72,73,74,75,76,77,78,79,80,81], and has beneficial effects on diabetes [82,83,84,85,86,87,88] and gentamicin-induced nephrotoxicity [89].
Thus, it is not surprising that many efforts have been devoted to the development of FA derivatives [21,27,29,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117,118,119,120,121,122,123,124,125,126,127,128,129,130,131,132,133,134,135]. Some of their relevant structural modifications and properties are summarized in Table 1. Their bioactivities are diverse, including antioxidant, anticancer, anti-inflammatory, and neuroprotective effects. According to the gathered data, it becomes evident that the ferulic acid molecular framework is a promising choice for developing new molecules with health benefits.
In this work, a systematic and rational search for FA derivatives is presented. For that purpose, a computer-assisted protocol known as CADMA-Chem [136] was used. The goal of the search is to find candidates that behave as multifunctional antioxidants, which are currently recognized as promising candidates to deal with OS-related complex diseases. This kind of antioxidants can scavenge free radicals, chelate metals and inhibit OH production, repair oxidatively damaged biomolecules, and inhibit enzymes involved in the development of health disorders [136,137,138]. In particular, the derivatives designed in the present work are meant to be oral drugs that simultaneously act as neuroprotectors against Parkinson’s and Alzheimer’s diseases, as well as free radical scavengers. It seems worthwhile to emphasize the fact that they are novel structures and that such combined activity has not been previously reported. In pursuit of such a goal, the FA framework was modified through the inclusion of different functional groups at all R1 to R5 sites (Scheme 1). Absorption, distribution, metabolism and excretion (ADME) properties were evaluated, as well as toxicity and synthetic accessibility (SA). Antioxidant activity through electron and H donation was predicted. Polygenic protection was explored by the interaction with enzymes linked to the target diseases. Namely: acetylcholinesterase (AChE), monoamine oxidase type B (MAOB), and catechol-O-methyltransferase (COMT). The inhibition of the first one has been shown to help with Alzheimer’s [139,140,141], while the inhibition of the other two is beneficial for Parkinson’s [142,143,144,145,146,147,148,149]. The obtained results are encouraging and might promote further investigations on the molecules identified as the most promising candidates.

2. Computational Details

2.1. Molecular Properties

For all the designed ferulic acid derivatives (Table S1), physicochemical parameters related to absorption, distribution, metabolism and excretion (ADME) were evaluated (Table S2) with the Molinspiration Property Calculation Service [150] and RDKit software (RDKit: Open-source cheminformatics. https://www.rdkit.org, accessed on 25 May 2023) The computed parameters are employed to confirm if the designed derivatives satisfy the Lipinski’s, Ghose’s, and Veber’s rules [151,152,153]. Compounds violating more than one of Lipinski’s or Veber’s rules are assumed to have difficulties with bioavailability, while those violating Ghose’s may present absorption problems or low permeation. Viable medical drugs also need to fulfill other vital requirements, such as synthetic accessibility (SA) and safety. The SA of the designed compounds was determined with the SYLVIA-XT 1.4 program (Molecular Networks, Erlangen, Germany) [154,155]. It estimates a value between 1 and 10. The smaller the value, the easier it is to synthesize the compound. LD50 and Ames mutagenicity (M) were employed to assess the toxicity of FA and its derivatives. The Toxicity Estimation Software Tool (T.E.S.T.), version 4.1 [156], was employed for that purpose. Selection and elimination scores (Tables S2 and S3), expressed in terms of toxicity, manufacturability and ADME properties, were used for sampling the molecular space. A reference set of molecules, which have been used (to some extent) as neuroprotectors, was used for comparison purposes (Table S4).

2.2. Reactivity Indexes

Gaussian 09 package of programs was employed for electronic structure calculations [157]. The M05-2X/6-311+G(d,p) level of theory was used for geometry optimizations and frequency calculations. The solvation model density (SMD) [158] was used for solvent effects, using water as a solvent. Local minima were identified by the absence of imaginary frequencies, and unrestricted calculations were used for open shell systems. M05-2X is a wide-spectrum functional with good performance for noncovalent interactions, thermochemistry and kinetics [159]. In addition, it has been recommended for modeling open-shell systems [160]. M05-2X functional has also been successfully used to estimate bond dissociation energies (BDE) and the free radical scavenging activity of diverse antioxidants [161,162,163,164,165].
The electron propagator theory (EPT) [166,167] was used to calculate ionization energies (IE) and electron affinities (EA). The partial third-order quasiparticle theory (P3)112 was chosen, within the EPT framework, because it produces lower mean errors than other approaches [168]. Pole strength (PS) values were checked to be larger than 0.80–0.85 (Table S5), which validates the obtained results [169].
For the estimation of BDEs, all sites likely to act as H-atom donors were considered, i.e., the -CH3 in the ether moiety of FA, and the phenolic OH (sites a and b, Scheme 1) and the new groups arising from functionalization of R1 to R5 sites.
Acid constants, expressed as pKa, were calculated with the Marvin suite [170]. This property is of crucial importance for medical drugs since it governs the proportion of neutral species at a particular pH, and these are the species most likely to passively cross biological barriers. The reliability of Marvin estimations was validated. To that purpose, 137 pKa were estimated, which correspond to first and second deprotonations of phenols, amines, carboxylic acids, thiols, and compounds structurally close to ferulic acid. The molecules used for such a validation are those reported in references [171,172,173]. The mean unsigned error (MUE) obtained from comparisons with the corresponding experimental data was found to be 0.42 pKa units. The correlation between calculated and experimental pKa values is provided in Supplementary Figure S1 (slope = 1.01, intercept = 0.04, and R2 = 0.95). This seems to support the reliability of the pKa values estimated here with the Marvin software.

2.3. Enzymatic Interactions

The structures of COMT (PDB ID: 3S68), MAO-B (PDB ID: 2V5Z) and AChE (PDB ID: 4EY7) co-crystallized with recognized neuroprotector drugs, tolcapone, safinamide and donepezil, respectively, were obtained from the protein data bank [174,175,176]. AChE missing loop regions (256-PGGTGG-261 and 493-PKA-496) were fixed using the Modeller web service [177]. Water molecules and species without biological interest were removed with the Discovery Studio software [178]. Ionizable groups of protein were modelled considering the protonation state of lateral chains and charge for D, E, R, K and H amino acids at physiological pH. For ligands, atomic charges are estimated by NBO protocol as single-point calculations with DFT (M05-2X/6-311+G(d,p)) methodology. Docking simulations were carried out using AutoDock Vina software [179]. A gradient optimization algorithm was performed inside of the active site centered at x: −13.50, y: 37.69, z: 61.63 and grid size of 15 × 15 × 15 Å3 for COMT, x: 51.81, y: 156.34, z: 28.15 and grid size of 13 × 13 × 13 Å3 for MAO-B and x: −18.80, y: −43.83, z: 27.67 and grid size of 17 × 13 × 13 Å3 for AChE. Docking scores (ΔGBW) were reported for the best-docked pose and weighted according to the abundance (molar fraction) of the acid-base species at physiological pH. The best conformation was analyzed and drawn with Pymol 2.5.4 software [180].
The redocking RMSD values were 1.8, 1.6 and 2.8 Å, respectively, and redocking scores (7.65, 10.10 and 10.86 kcal/mol) were found for tolcapone (COMT), safinamide (MAO-B) and donepezil (AChE), respectively, which agrees with experimental findings. These results confirm the suitability of the docking methodology. Redocking conformations are obtained with Chimera software [181], and they can be found in Figure S2.

3. Results and Discussion

3.1. Derivatives and Properties

By inserting -OH, -NH2, -SH and -COOH groups in sites R1 to R5, 185 new FA derivatives were built (Table S1). Twenty of them with one functional group, one hundred and sixty with any possible combination of two functional groups, and five with three functional groups. The latter were constructed from the most promising bi-functionalized species.
A selection score (SS) was computed. It is meant to identify the FA derivatives with the most likely drug-like behavior and corresponds to that included in the CADMA-Chem protocol [136,137,138,182,183,184]. The associated equations are provided in Table S6. The higher the value of SS, the more likely the drug-like behavior. SS takes into account eight ADME properties: water/octanol partition coefficient (logP), topological polar surface area (PSA), number of heavy atoms (XAt), molecular weight (MW), number of H-bond acceptors (HBA), number of H-bond donors (HBD), rotatable bonds (RB), and molar refractivity (MR); two toxicity descriptors: median lethal dose for rats (LD50) and Ames’ mutagenicity (M); and the synthetic accessibility (SA).
The SS for all the designed FA derivatives is presented in Figure 1. The parent molecule and the average SS value for the reference set are included for comparison purposes. The individual values of all the FA derivatives are reported in Table S2, together with those of the above-mentioned descriptors. Higher values of SS suggest better drug-like behavior, lower toxicity, and easier synthesis. The first screening was based on this score, and twelve FA derivatives were selected. However, before moving them forward to the next stage of the investigation (Scheme 2), a double-check analysis was performed using exclusion scores (SE), which allowes to verify if any of the selected molecules significantly deviate (in any of its properties) from the average value of the reference set.
Four exclusion scores were analyzed (SE,ADME2, SE,ADME8, SE,ADMET and SE,ADMETSA). Their equations are provided in Table S7. SE,ADME8, SE,ADMET and SE,ADMETSA are extensions of the well-known SE,ADME2, based on two descriptors (logP and MW) [185,186]. SE,ADME8 uses the same kind of strategy as SE,ADME2, but includes six additional terms (PSA, XAt, HBA, HBD, RB, and MR). SE,ADMET and SE,ADMETSA also include toxicity (LD50 and M) and synthetic accessibility (SA) descriptors.
SE,ADME2 values were previously estimated to be between 1.2 and 1.5 for 1791, 152 and 35 oral drugs [185,186]. For the 12 selected FA derivatives, the average SE,ADME2 value was found to be 1.06, with individual values ranging from 0.56 to 1.42 (Table S4). The estimated average values for the other elimination scores were found to be: SE,ADME8 = 4.86 (ranging from 2.59 to 7.34), SE,ADMET = 8.13 (ranging from 4.53 to 11.54), and SE,ADMETSA = 9.59 (ranging from 5.50 to 12.52). It seems worthwhile mentioning that high values of the exclusion scores might result from either worse or better behavior than the average of the reference drugs. Thus, a detailed analysis is required to determine if any particular candidate should be removed from the selection or not.
According to the gathered results (Figure 2), toxicity is responsible for the largest deviation. Regarding ADME, the six additional descriptors lead to the largest deviation than logP and MW. Synthetic accessibility also has a rather small influence on the deviations from the reference molecules. A more detailed examination, considering the individual contribution of all the investigated descriptors, is presented in Figure 3.
The more important deviations arise from LD50, followed by M, PSA and HBD. The FA derivatives with the largest LD50 deviations from the reference set are FA, FA-173, FA-175 and FA-26. However, they correspond to a lower toxicity to rats than the average of the reference set (LD50 = 960.8), with values of 4742.7, 4471.9, 4040.7, and 3635.2, respectively. Regarding Ames mutagenicity, a similar trend was found. The FA derivatives predicted as the least mutagenic are just those that deviate the most from the reference set average (M = 0.41). They are FA-88, FA-106, FA-115 and FA-142, all with M = 0.01. Thus, these deviations imply that the above-mentioned derivatives have a more desirable behavior than that of the reference set. Accordingly, they were not excluded from the chosen subset.
The largest PSA deviation were found for FA-41, FA-26, FA-88 and FA-173 (124.3, 113.0, 104.1 and 104.1 Å2, respectively). However, their PSA values are all below Veber’s limit: 140 Å2. Thus, these derivatives were also kept in the chosen subset. The largest deviations for HBD correspond to FA-26 with HBD = 5 and FA-8, FA-41 and FA-138 with HBD = 4. Since they do not represent violations of Lipinski’s rule, these candidates were not eliminated.
After carefully examining elimination scores for the 12 FA derivatives with the highest SS values, none of them were excluded from the selection. Thus, they were investigated regarding their antioxidant capacity through electron and H-atom donation. This detailed analysis is important since it allows interpreting deviations for all the used descriptors and prevents the exclusion of suitable candidates for no good reason.

3.2. pKa and Antioxidant Activity

As previously mentioned, acid-base equilibria are crucial for medical drugs intended to passively cross biological barriers. The pKa values and molar fractions (Mf) at physiological pH were estimated for the 12 FA derivatives chosen in the first stage of the investigation as those with the best drug-like behavior (Table 2). Additionally, the corresponding deprotonation routes and distribution diagrams are provided in Figures S2 and S3.
The calculated molar fractions (Table 2) revealed that 7 of the 12 derivatives, selected based on the SS value, would have a negligible population (<10−4) at physiological pH, i.e., at pH = 7.4. Thus, they were excluded as viable candidates. Although the Mf(0) for the other three (FA-26, FA-118, and FA-175) are rather small, they are very similar to that of FA. Since there is abundant data on the biological activities of FA (Table 1), it can be inferred that such fractions are enough. Consequently, five derivatives (FA-8, FA-26, FA-118, FA-138, and FA-175) were further investigated. Among the studied derivatives, FA-138 is the only one that is predicted to have similar fractions of neutral (q = 0) and anionic (q = −1) species. This feature might be relevant to its possible use as a multifunctional antioxidant. The rather large neutral fraction (59.0%) is expected to promote passive crossing through biological membranes, while the anionic fraction (40.3%) is likely to be the key one for the free radical scavenging activity, as it is the case for many phenolic compounds.
The ionization energies (IE), electron affinities (EA), and the lowest bond dissociation energies (BDE) for the acid-base species with a non-negligible population (Mf(q) ≥ 10−4) of FA and its derivatives at pH = 7.4, are reported in Table 3. The complete set of BDEs, i.e., considering all viable H-donating sites, is provided as Supplementary Materials (Table S8). IE and BDE reactivity indexes are related to the viability of electron and H-atom donation. Thus, they were used to compare the efficiency of the derivatives with that of reference antioxidants as free radical scavengers via single electron transfer (SET) and formal hydrogen atom transfer (HAT) mechanisms, respectively.
IE and BDE values were used to build the electron and hydrogen-donating ability map for antioxidants (eH-DAMA, Figure 4). This graphical tool has been recently proposed to simultaneously account for the likeliness of molecules as H donors (formal HAT reaction route) and electron donors (SET reaction route) [92,93]. The dominant acid-base species of the investigated FA derivatives at physiological pH were included in this map, as well as two antioxidant references (Trolox and α-tocopherol), the parent molecule, and the H2O2/O2•−. This pair represents the potential oxidant target. The best radical scavengers are expected to be located at the bottom left, i.e., lower IE and lower BDE. The species in this region are likely to simultaneously act as electron and H-atom donors.
Based on the eH-DAMA (Figure 4), it is predicted that the five FA derivatives included in it should be efficient for scavenging peroxyl radicals through both mechanisms, SET and f-HAT. Their efficiency for that purpose is expected to surpass that of α-tocopherol and ferulic acid. On the contrary, only the anionic form of FA-138 is predicted to be more efficient than Trolox for that purpose. FA-8 may be a better electron donor than Trolox but not as good for donating H-atoms. However, further investigations dealing with other aspects of antioxidant activity, kinetics in particular, are still needed to confirm or refute the foreseen trends.

3.3. Polygenic Activity

To evaluate general neuroprotection activity, a polygenic score (SP) was developed. SP is a measure of the tested compounds' capacity to bind to the enzymes compared with natural substrates (COMT: dopamine (dopa), MAO-B phenylethylamine (pea) and AChE: acetylcholine (ACh). It was defined according to our previous reports [136,137] as:
S P = G B , C O M T W G B , d o p a + G B , M A O B W G B , p e a + G B , A C h E W G B , A C h
The scoring values are presented in Table 4. When the values of SP are examined, it can be predicted that the compounds exhibit neuroprotection activity since their scores are higher than those of the corresponding natural substrates (SP = 3.00), i.e., the investigated ferulic acid derivatives may present stronger affinities towards the enzymes. Among the studied compounds, the FA-26 analog is expected to have the best neuroprotection activity. Interestingly, according to the docking results, the parent molecule (ferulic acid) is also likely to act as a neuroprotector.
The examination of individual ΔGBW values reveals that the studied compounds could be better inhibitors for AChE and MAO-B than they are for the COMT enzyme. Negative values of COMT (blue fragment of the bars in Figure 5) indicate that this enzyme forms more stable complexes with dopamine than with the tested FA derivatives. Only FA-118 shows a slightly higher score than dopamine. Interestingly, this compound has a catechol moiety, which is recognized to exhibit effective COMT inhibition potential [187]. FA-175 presents almost the same score as dopamine (log ΔGWB/ΔGB,sub= −0.001). On the other hand, for MAO-B and AChE (green and red fragments, respectively, Figure 5), the neuroprotection behavior of FA derivatives was evidenced by their positive values. Between these two enzymes, the inhibitor potential of the studied derivatives is expected to be stronger for AChE. The binding energies ΔGB values per acid-base species of the most promised derivatives can be found in Table S9, and the complete set of ΔGBW for the thirteen selected derivatives (see Scheme 2) can be consulted in Table S10, Supplementary Materials.
Molecular docking allows the prediction of the binding conformations between therapeutic targets and small molecules. The analysis of the possible interactions that form the adducts promotes development and drug discovery. We must not lose sight of the limitations of the method, and if the work demands obtaining more realistic conformations, the use of more precise tools such as molecular dynamics or QM protocols is essential. Even so, molecular docking has been shown to provide reliable predictions of non-covalent bonds, such as hydrogen bonding and hydrophobic interactions [187,188]. The main interactions for the complexes with the highest SP are shown in Figure 6. They are FA-26 with AChE (left), FA-26 with MAO-B (middle), and FA-118 with COMT (right). For all of them, FA-26 is in its anionic form, which is the most abundant species at physiological pH (X ~ 0.97). To understand the interactions formed in the protein-ligand complexes, it is important to know the architecture of the enzymes and the function of the key residues. AChE has a highly specialized structure, which allows it to be one of the fastest-known enzymes. The catalytic triad (H447, E334 and S203) is found at the bottom of the enzyme and surrounded by 14 well-conserved aromatic residues [189]. Among them, W83 plays an essential role since it forms a substrate union site, while Y70, Y121, and W279 conform to the anionic peripheric site [189]. Additionally, AChE has a high dipole moment with the axis oriented towards the substrate entry site. It has been suggested that this moment may serve to pull down the cationic substrate of AChE. This dipole is controlled mainly by residues D71, E199, and E443 [190]. The binding and anionic sites are responsible for supporting the cationic substrate acetylcholine by the ammonium group, as well as both quaternary ligands (edrophonium, N-methylacridinium) acting as competitive inhibitors. In the catalytic site, the ester hydrolysis leads to the formation of an acyl group attached to the enzyme and the release of choline. Then, a water molecule assists at residue H447, releasing acetic acid, regenerating free enzyme and ending the function of this neurotransmitter [191]. The pharmacological effect of AChE inhibitors consists of the inactivation of the enzymatic activity resulting in the increase of synaptic ACh and the stimulation of postsynaptic cholinergic receptors in the central and peripheral nervous systems. Therefore, these drugs improve cholinergic neurotransmission and compensate for the loss of brain cells in some conditions, such as Alzheimer’s, providing benefits in all the key symptoms of the disease [192].
FA-26 has several H-bond donors and acceptors and an aromatic ring that contributes to generating intermolecular connections with the AChE key amino acids. In fact, complex FA-26:AChE is formed by several interactions, mainly hydrogen bonds and π-interactions. This derivative is bonded to the active site of AChE through four hydrogen bonds (D71, Y121, F292, and Y338), one π-stacking interaction (Y334) and one p-alkyl interaction (W83). The observed interactions suggest that FA-26, although not bonded to the catalytic triad, can inhibit ACh degradation, blocking the entry and union sites.
MAO-B function involves two hydrophobic pockets, an entry pocket and an active site pocket, with I199 acting as a gatekeeper between two cavities. The catalytic reaction site comprises a redox cofactor, flavin adenine dinucleotide (FAD). The active site is completed by residues Y398 and Y435, orienting the substrate to the proper position [193]. The enzyme promotes the oxidation of amines, generating aldehyde, ammonia and hydrogen peroxide. Although the mechanism has not been fully elucidated, studies with MAO inhibitors suggest that FAD is a key fragment in the transformation of amines [194]. Inhibitors of MAO-B are used to conserve adequate levels of several neurotransmitters as dopamine, norepinephrine, and serotonin, or to increase them. For this reason, MAO-B inhibitors are used to treat depression and alleviate the symptoms of Parkinson’s disease [195].
Four H-bonds involving Q206, L171, and FAD, a p-stacking (F343) and non-conventional C-H bonds stabilize the complex formation. An important feature of the conformation adopted by FA-26 in the complex is the formation of an H-bond with N5 in the FAD moiety. This atom is required for the redox activity of the cofactor [193] and, hence, for the catalytic function of the enzyme. This conformation could not be achieved without the orientation promoted by the L171 and Y398 residues, which suggests that FA-26 could inhibit some enzymes with the same mechanism of action as MAO (type A) or other flavoenzymes as lactate oxidase [196]. According to these findings, FA-26 is predicted to act as a reversible or non-covalent MAO-B inhibitor as Safinamide or Moclobemide [175,197], which are recognized antidepressant drugs. This way of inhibition is preferable since it has been proven to be associated with less toxicity than others [198].
COMT is a selective enzyme that catalyzes the transfer of methyl groups to the 3-OH position of catecholamines. COMT is an Mg-dependent enzyme, with the metal bound to D141, D169, and N170 residues. This enzyme uses the Mg atom to bind the substrate and make it more easily ionizable [199]. The methyl group is transferred by the S-Adenosylmethionine cofactor. The binding substrate site is completed with several hydrophobic residues M40, L198, W143, and the gatekeepers W38 and P174 [199]. COMT is responsible for the selective methylation of catecholamines hydroxyls, including dopamine, epinephrine, and norepinephrine. The inhibition of this protein has become a key strategy to manipulate the levels of these neurotransmitters and other substances that are dopamine precursors, such as L-DOPA or Carbidopa, used to treat Parkinson's disease [187].
According to the docking simulations, FA-118 has a catechol moiety that binds the Mg atom by two metal-donor unions. A hard acid-base interaction (Mg-O) stabilizes the formation of this adduct. In addition, H-bonds between the catechol fragment and the K144 and N170 residues also contribute to the binding energy. Finally, several hydrophobic interactions with key residues of the active site (M40 and P174) complete the stabilization of the FA-118:COMT complex. Such an arrangement explains the good score obtained in the simulations and suggests that FA-118 can be efficient as a COMT inhibitor.
The docking simulations indicate that while all the investigated FA derivatives can act as neuroprotectors of acetylcholine and phenylethylamine (with FA-26 being predicted as the best one for that purpose), only FA-118 would be able to protect dopamine against COMT-induced degradation. Accordingly, FA-118 is proposed as a promising candidate in the context of Alzheimer’s and/or anti-anxiety disorders, while FA-26 was identified as the best candidate (among the studied molecule) for Parkinson’s. All of them certainly deserve further investigations related to their potential as neuroprotectors.

4. Conclusions

A total of 185 ferulic acid (FA) derivatives were built through a rational in silico design using the CADMA-Chem protocol. The chemical space was sampled using a selection score (SS) that considers ADME properties, toxicity and synthetic accessibility descriptors. Based on the estimated SS values, 12 FA derivatives were identified as the candidates with the best drug-like behavior. For this subset, some reactivity indexes were computed, as well as their pKa values. According to eH-DAMA results, which take into account the free radical scavenging behavior through single electron transfer (SET) and formal hydrogen transfer (HAT) mechanisms, FA-138 seems to be the best candidate to scavenge free radicals. However, FA-8, FA-26, FA-118, and FA-175 derivatives are predicted to be better for that purpose than α-tocopherol and the parent molecule.
On the other hand, docking studies suggest that ferulic acid and some of its derivatives can act as inhibitors of AChE and MAO-B enzymes. FA-26 is predicted as the most efficient one for that purpose. This compound is bound preferably to the entry site of AChE and to the catalytic site of MAO-B, acting as a reversible inhibitor for the latter. On the contrary, FA-118 was the only compound identified as a viable candidate to efficiently inhibit COMT. Accordingly, FA-26 is proposed as the best candidate in the context of Alzheimer’s and/or anti-anxiety disorders and FA-118 for Parkinson’s. At least these two compounds certainly deserve further investigation regarding their potential role as neuroprotectors.
Considering the gathered data altogether, the FA derivatives proposed for further investigations are FA-26, FA-118, and FA-138.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/antiox12061256/s1. Table S1: FA derivatives designed in this work. Table S2: Values of the ADME properties, toxicity descriptors, synthetic accessibility, and selection score (SS) for all designed derivatives. Table S3: Elimination scores for the subset of ferulic acid derivatives chosen as the most promising, according to SS. Table S4: Reference set of molecules with some neuroprotective effects. Table S5: Pole strength values for the EPT approximation (P3) used to calculate ionization energies and electron affinities. Table S6: Equations concerning SS construction. Table S7: Exclusion scores (SE) equations. Table S8: Complete set of BDEs for ferulic acid and its derivatives. Table S9: Complete set of the binding energies for ferulic acid and its derivatives. Figure S2: Redocking simulation: tolcapone in COMT, Safrinamide in MAO-B, and Donopezil in AChE. Figure S3: Deprotonation routes for the subset of ferulic acid derivatives chosen as the most promising from their drug-like behavior. Figure S4: Distribution diagram of the acid-base species of ferulic acid derivatives.

Author Contributions

Conceptualization, A.G.; Investigation, E.G.G.-L., M.R., L.F.H.-A.; Formal Analysis, E.G.G.-L., M.R., L.F.H.-A. and A.G.; Methodology, E.G.G.-L., M.R., L.F.H.-A. and A.G.; Project Administration, A.G.; Supervision, A.G.; Validation, E.G.G.-L., M.R., L.F.H.-A.; Visualization, E.G.G.-L., M.R., L.F.H.-A. and A.G.; Writing—Original Draft Preparation, E.G.G.-L., M.R., L.F.H.-A.; Writing—Review & Editing, A.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article and supplementary materials.

Acknowledgments

We gratefully thank the Laboratorio de Visualización y Cómputo Paralelo at Universidad Autónoma Metropolitana-Iztapalapa for computing time. E.G.G.-L. acknowledges CONACyT for the Doctoral fellowship. L.F.H.-A thanks to Estancias Posdoctorales por México (2022) CONACyT program for the postdoctoral grant.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kumar, N.; Pruthi, V. Potential applications of ferulic acid from natural sources. Biotechnol. Rep. 2014, 4, 86–93. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. León-Carmona, J.R.; Alvarez-Idaboy, J.R.; Galano, A. On the peroxyl scavenging activity of hydroxycinnamic acid derivatives: Mechanisms, kinetics, and importance of the acid-base equilibrium. Phys. Chem. Chem. Phys. 2012, 14, 12534–12543. [Google Scholar] [CrossRef] [PubMed]
  3. Chen, X.; Wang, Y.; Chen, D.; Yu, B.; Huang, Z. Dietary ferulic acid supplementation improves intestinal antioxidant capacity and intestinal barrier function in weaned piglets. Anim. Biotechnol. 2022, 33, 356–361. [Google Scholar] [CrossRef]
  4. Han, H.; Dye, L.; Mackie, A. The impact of processing on the release and antioxidant capacity of ferulic acid from wheat: A systematic review. Food Res. Int. 2023, 164, 112327. [Google Scholar] [CrossRef]
  5. Horbury, M.D.; Baker, L.A.; Quan, W.D.; Greenough, S.E.; Stavros, V.G. Photodynamics of potent antioxidants: Ferulic and caffeic acids. Phys. Chem. Chem. Phys. 2016, 18, 17691–17697. [Google Scholar] [CrossRef] [Green Version]
  6. Hwang, H.J.; Lee, S.R.; Yoon, J.G.; Moon, H.R.; Zhang, J.; Park, E.; Yoon, S.I.; Cho, J.A. Ferulic Acid as a Protective Antioxidant of Human Intestinal Epithelial Cells. Antioxidants 2022, 11, 1448. [Google Scholar] [CrossRef]
  7. Itagaki, S.; Kurokawa, T.; Nakata, C.; Saito, Y.; Oikawa, S.; Kobayashi, M.; Hirano, T.; Iseki, K. In vitro and in vivo antioxidant properties of ferulic acid: A comparative study with other natural oxidation inhibitors. Food Chem. 2009, 114, 466–471. [Google Scholar] [CrossRef]
  8. Lima, Â.C.O.; Dias, E.R.; Reis, I.M.A.; Carneiro, K.O.; Pinheiro, A.M.; Nascimento, A.S.; Silva, S.M.P.C.; Carvalho, C.A.L.; Mendonça, A.V.R.; Vieira, I.J.C.; et al. Ferulic acid as major antioxidant phenolic compound of the Tetragonisca angustula honey collected in Vera Cruz-Itaparica Island, Bahia, Brazil. Braz. J. Biol. 2022, 84, e253599. [Google Scholar] [CrossRef]
  9. Rampelotto, C.R.; Pereira, V.G.; da Silva Silveira, L.; Rossato, A.; Machado, A.K.; Sagrillo, M.R.; Gündel, A.; Burger, M.E.; Schaffazick, S.R.; de Bona da Silva, C. Ferulic acid-loaded nanocapsules: Evaluation of mucosal interaction, safety and antioxidant activity in human mononucleated cells. Toxicol. In Vitro 2022, 78, 105259. [Google Scholar] [CrossRef] [PubMed]
  10. Srinivasan, M.; Sudheer, A.R.; Menon, V.P. Ferulic acid: Therapeutic potential through its antioxidant property. J. Clin. Biochem. Nutr. 2007, 40, 92–100. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Truong, D.H.; Nhung, N.T.A.; Dao, D.Q. Iron ions chelation-based antioxidant potential vs. pro-oxidant risk of ferulic acid: A DFT study in aqueous phase. Comp. Theor. Chem. 2020, 1185, 112905. [Google Scholar] [CrossRef]
  12. Wagle, S.; Sim, H.J.; Bhattarai, G.; Choi, K.C.; Kook, S.H.; Lee, J.C.; Jeon, Y.M. Supplemental ferulic acid inhibits total body irradiation-mediated bone marrow damage, bone mass loss, stem cell senescence, and hematopoietic defect in mice by enhancing antioxidant defense systems. Antioxidants 2021, 10, 1209. [Google Scholar] [CrossRef]
  13. Yildiztugay, E.; Ozfidan-Konakci, C.; Karahan, H.; Kucukoduk, M.; Turkan, I. Ferulic acid confers tolerance against excess boron by regulating ROS levels and inducing antioxidant system in wheat leaves (Triticum aestivum). Environ. Exp. Bot. 2019, 161, 193–202. [Google Scholar] [CrossRef]
  14. Zduńska, K.; Dana, A.; Kolodziejczak, A.; Rotsztejn, H. Antioxidant properties of ferulic acid and its possible application. Skin Pharmacol. Physiol. 2018, 31, 332–336. [Google Scholar] [CrossRef]
  15. Mir, S.M.; Ravuri, H.G.; Pradhan, R.K.; Narra, S.; Kumar, J.M.; Kuncha, M.; Kanjilal, S.; Sistla, R. Ferulic acid protects lipopolysaccharide-induced acute kidney injury by suppressing inflammatory events and upregulating antioxidant defenses in Balb/c mice. Biomed. Pharmacother. 2018, 100, 304–315. [Google Scholar] [CrossRef]
  16. Yin, Z.N.; Wu, W.J.; Sun, C.Z.; Liu, H.F.; Chen, W.B.; Zhan, Q.P.; Lei, Z.G.; Xin, X.; Ma, J.J.; Yao, K.; et al. Antioxidant and Anti-inflammatory Capacity of Ferulic Acid Released from Wheat Bran by Solid-state Fermentation of Aspergillus niger. Biomed. Environ. Sci. 2019, 32, 11–21. [Google Scholar] [PubMed]
  17. Amani, F.; Rezaei, A.; Kharazmi, M.S.; Jafari, S.M. Loading ferulic acid into β-cyclodextrin nanosponges; antibacterial activity, controlled release and application in pomegranate juice as a copigment agent. Colloids Surf. Physicochem. Eng. Asp. 2022, 649, 129454. [Google Scholar] [CrossRef]
  18. Borges, A.; Ferreira, C.; Saavedra, M.J.; Simões, M. Antibacterial activity and mode of action of ferulic and gallic acids against pathogenic bacteria. Microb. Drug Resist. 2013, 19, 256–265. [Google Scholar] [CrossRef] [PubMed]
  19. Ordoñez, R.; Atarés, L.; Chiralt, A. Antibacterial properties of cinnamic and ferulic acids incorporated to starch and PLA monolayer and multilayer films. Food Control 2022, 136, 108878. [Google Scholar] [CrossRef]
  20. Tu, Q.B.; Shi, H.C.; Li, P.; Sheng, S.; Wu, F.A. Antibacterial Activity of Ferulic Acid Ester against Ralstonia solanacearum and Its Synergy with Essential Oils. Sustainability 2022, 14, 16348. [Google Scholar] [CrossRef]
  21. Antonopoulou, I.; Sapountzaki, E.; Rova, U.; Christakopoulos, P. Ferulic Acid From Plant Biomass: A Phytochemical With Promising Antiviral Properties. Front. Nutr. 2022, 8, 777576. [Google Scholar] [CrossRef]
  22. Choi, J.H.; Park, J.K.; Kim, K.M.; Lee, H.J.; Kim, S. In vitro and in vivo antithrombotic and cytotoxicity effects of ferulic acid. J. Biochem. Mol. Toxicol. 2018, 32, 22004. [Google Scholar] [CrossRef]
  23. Hong, Q.; Ma, Z.C.; Huang, H.; Wang, Y.G.; Tan, H.L.; Xiao, C.R.; Liang, Q.D.; Zhang, H.T.; Gao, Y. Antithrombotic activities of ferulic acid via intracellular cyclic nucleotide signaling. Eur. J. Pharmacol. 2016, 777, 1–8. [Google Scholar] [CrossRef]
  24. Zduńska-Pęciak, K.; Kołodziejczak, A.; Rotsztejn, H. Two superior antioxidants: Ferulic acid and ascorbic acid in reducing signs of photoaging—A split-face comparative study. Dermatol. Ther. 2022, 35, e15254. [Google Scholar] [CrossRef]
  25. Fukuda, T.; Kuroda, T.; Kono, M.; Hyoguchi, M.; Tanaka, M.; Matsui, T. Augmentation of ferulic acid-induced vasorelaxation with aging and its structure importance in thoracic aorta of spontaneously hypertensive rats. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2015, 388, 1113–1117. [Google Scholar] [CrossRef]
  26. Yang, H.; Qu, Z.; Zhang, J.; Huo, L.; Gao, J.; Gao, W. Ferulic acid ameliorates memory impairment in d-galactose-induced aging mouse model. Int. J. Food Sci. Nutr. 2016, 67, 806–817. [Google Scholar] [CrossRef] [PubMed]
  27. Kumar, N.; Kumar, S.; Abbat, S.; Nikhil, K.; Sondhi, S.M.; Bharatam, P.V.; Roy, P.; Pruthi, V. Ferulic acid amide derivatives as anticancer and antioxidant agents: Synthesis, thermal, biological and computational studies. Med. Chem. Res. 2016, 25, 1175–1192. [Google Scholar] [CrossRef]
  28. Ani, G.; Tanya, T.Y.; Reneta, T. Antitumor and apoptogenic effects of ferulic acid on cervical carcinoma cells. Res. J. Biotechnol. 2021, 16, 6–11. [Google Scholar]
  29. Bakholdina, L.A.; Markova, A.A.; Khlebnikov, A.I.; Sevodin, V.P. Cytotoxicity of New Ferulic-Acid Derivatives on Human Colon Carcinoma (HCT116) Cells. Pharm. Chem. J. 2019, 53, 516–520. [Google Scholar] [CrossRef]
  30. Cao, Y.; Zhang, H.; Tang, J.; Wang, R. Ferulic Acid Mitigates Growth and Invasion of Esophageal Squamous Cell Carcinoma through Inducing Ferroptotic Cell Death. Dis. Markers 2022, 2022, 4607966. [Google Scholar] [CrossRef]
  31. Cui, K.; Wu, H.; Fan, J.; Zhang, L.; Li, H.; Guo, H.; Yang, R.; Li, Z. The Mixture of Ferulic Acid and P-Coumaric Acid Suppresses Colorectal Cancer through lncRNA 495810/PKM2 Mediated Aerobic Glycolysis. Int. J. Mol. Sci. 2022, 23, 12106. [Google Scholar] [CrossRef] [PubMed]
  32. Damasceno, S.S.; Dantas, B.B.; Ribeiro-Filho, J.; Araújo, D.A.M.; Da Costa, J.G.M. Chemical properties of caffeic and ferulic acids in biological system: Implications in cancer therapy. A review. Curr. Pharm. Des. 2017, 23, 3015–3023. [Google Scholar] [CrossRef] [PubMed]
  33. Dodurga, Y.; Eroğlu, C.; Seçme, M.; Elmas, L.; Avcı, Ç.B.; Şatıroğlu-Tufan, N.L. Anti-proliferative and anti-invasive effects of ferulic acid in TT medullary thyroid cancer cells interacting with URG4/URGCP. Tumor Biol. 2016, 37, 1933–1940. [Google Scholar] [CrossRef] [PubMed]
  34. El-Gogary, R.I.; Nasr, M.; Rahsed, L.A.; Hamzawy, M.A. Ferulic acid nanocapsules as a promising treatment modality for colorectal cancer: Preparation and in vitro/in vivo appraisal. Life Sci. 2022, 298, 120500. [Google Scholar] [CrossRef]
  35. ElKhazendar, M.; Chalak, J.; El-Huneidi, W.; Vinod, A.; Abdel-Rahman, W.M.; Abu-Gharbieh, E. Antiproliferative and proapoptotic activities of ferulic acid in breast and liver cancer cell lines. Trop. J. Pharm. Res. 2019, 18, 2571–2576. [Google Scholar]
  36. Eroğlu, C.; Seçme, M.; Bağcı, G.; Dodurga, Y. Assessment of the anticancer mechanism of ferulic acid via cell cycle and apoptotic pathways in human prostate cancer cell lines. Tumor Biol. 2015, 36, 9437–9446. [Google Scholar] [CrossRef]
  37. Fahrioğlu, U.; Dodurga, Y.; Elmas, L.; Seçme, M. Ferulic acid decreases cell viability and colony formation while inhibiting migration of MIA PaCa-2 human pancreatic cancer cells in vitro. Gene 2016, 576, 476–482. [Google Scholar] [CrossRef]
  38. Gao, J.; Yu, H.; Guo, W.; Kong, Y.; Gu, L.; Li, Q.; Yang, S.; Zhang, Y.; Wang, Y. The anticancer effects of ferulic acid is associated with induction of cell cycle arrest and autophagy in cervical cancer cells. Cancer Cell Int. 2018, 18, 102. [Google Scholar] [CrossRef] [Green Version]
  39. Gupta, A.; Singh, A.K.; Loka, M.; Pandey, A.K.; Bishayee, A. Ferulic acid-mediated modulation of apoptotic signaling pathways in cancer. Adv. Protein Chem. Struct. Biol. 2021, 125, 215–257. [Google Scholar]
  40. Luo, L.; Zhu, S.; Tong, Y.; Peng, S. Ferulic acid induces apoptosis of HeLa and caski cervical carcinoma cells by down-regulating the phosphatidylinositol 3-kinase (PI3K)/Akt signaling pathway. Med. Sci. Monit. 2020, 26, e920095. [Google Scholar] [CrossRef]
  41. Zhang, X.; Lin, D.; Jiang, R.; Li, H.; Wan, J.; Li, H. Ferulic acid exerts antitumor activity and inhibits metastasis in breast cancer cells by regulating epithelial to mesenchymal transition. Oncol. Rep. 2016, 36, 271–278. [Google Scholar] [CrossRef] [Green Version]
  42. Alam, M.A.; Sernia, C.; Brown, L. Ferulic acid improves cardiovascular and kidney structure and function in hypertensive rats. J. Cardiovasc. Pharmacol. 2013, 61, 240–249. [Google Scholar] [CrossRef] [PubMed]
  43. Li, C.; Chen, L.; Song, M.; Fang, Z.; Zhang, L.; Coffie, J.W.; Zhang, L.; Ma, L.; Wang, Q.; Yang, W.; et al. Ferulic acid protects cardiomyocytes from TNF-α/cycloheximide-induced apoptosis by regulating autophagy. Arch. Pharm. Re. 2020, 43, 863–874. [Google Scholar] [CrossRef] [PubMed]
  44. Monceaux, K.; Gressette, M.; Karoui, A.; Pires Da Silva, J.; Piquereau, J.; Ventura-Clapier, R.; Garnier, A.; Mericskay, M.; Lemaire, C. Ferulic Acid, Pterostilbene, and Tyrosol Protect the Heart from ER-Stress-Induced Injury by Activating SIRT1-Dependent Deacetylation of eIF2α. Int. J. Mol. Sci. 2022, 23, 6628. [Google Scholar] [CrossRef]
  45. Neto-Neves, E.M.; Filho, C.D.S.M.B.; Dejani, N.N.; de Sousa, D.P. Ferulic acid and cardiovascular health: Therapeutic and preventive potential. Mini-Rev. Med. Chem. 2021, 21, 1625–1637. [Google Scholar] [CrossRef] [PubMed]
  46. Pandi, A.; Raghu, M.H.; Chandrashekar, N.; Kalappan, V.M. Cardioprotective effects of Ferulic acid against various drugs and toxic agents. J. Basic Appl. Sci. 2022, 11, 92. [Google Scholar] [CrossRef]
  47. Salau, V.F.; Erukainure, O.L.; Olofinsan, K.A.; Msomi, N.Z.; Ijomone, O.K.; Islam, M.S. Ferulic acid mitigates diabetic cardiomyopathy via modulation of metabolic abnormalities in cardiac tissues of diabetic rats. Fundam. Clin. Pharmacol. 2023, 37, 44–59. [Google Scholar] [CrossRef]
  48. Zhang, X.X.; Zhao, D.S.; Wang, J.; Zhou, H.; Wang, L.; Mao, J.L.; He, J.X. The treatment of cardiovascular diseases: A review of ferulic acid and its derivatives. Pharmazie 2021, 76, 55–60. [Google Scholar]
  49. Ren, Z.; Li, Y.; Zhang, R.; Li, Y.; Yang, Z.; Yang, H. Ferulic acid exerts neuroprotective effects against cerebral ischemia/reperfusion-induced injury via antioxidant and anti-apoptotic mechanisms in vitro and in vivo. Int. J. Mol. Med. 2017, 40, 1444–1456. [Google Scholar] [CrossRef] [Green Version]
  50. Di Giacomo, S.; Percaccio, E.; Gullì, M.; Romano, A.; Vitalone, A.; Mazzanti, G.; Gaetani, S.; Di Sotto, A. Recent Advances in the Neuroprotective Properties of Ferulic Acid in Alzheimer’s Disease: A Narrative Review. Nutrients 2022, 14, 3709. [Google Scholar] [CrossRef]
  51. Dong, X.; Huang, R. Ferulic acid: An extraordinarily neuroprotective phenolic acid with anti-depressive properties. Phytomedicine 2022, 105, 154355. [Google Scholar] [CrossRef]
  52. Hassanzadeh, P.; Arbabi, E.; Atyabi, F.; Dinarvand, R. Ferulic acid, a phenolic compound with therapeutic effects in neuropsychiatric disorders, stimulates the production of nerve growth factor and endocannabinoids in rat brain. Physiol. Pharmacol. (Iran) 2017, 21, 279–294. [Google Scholar]
  53. Liu, G.; Nie, Y.; Huang, C.; Zhu, G.; Zhang, X.; Hu, C.; Li, Z.; Gao, Y.; Ma, Z. Ferulic acid produces neuroprotection against radiation-induced neuroinflammation by affecting NLRP3 inflammasome activation. Int. J. Radia. Biol. 2022, 98, 1442–1451. [Google Scholar] [CrossRef]
  54. Liu, Y.M.; Shen, J.D.; Xu, L.P.; Li, H.B.; Li, Y.C.; Yi, L.T. Ferulic acid inhibits neuro-inflammation in mice exposed to chronic unpredictable mild stress. Int. Immunopharmacol. 2017, 45, 128–134. [Google Scholar] [CrossRef]
  55. Long, T.; Wu, Q.; Wei, J.; Tang, Y.; He, Y.N.; He, C.L.; Chen, X.; Yu, L.; Yu, C.L.; Law, B.Y.; et al. Ferulic Acid Exerts Neuroprotective Effects via Autophagy Induction in C. elegans and Cellular Models of Parkinson’s Disease. Oxid. Med. Cell. Longev. 2022, 2022, 3723567. [Google Scholar] [CrossRef] [PubMed]
  56. Ojha, S.; Javed, H.; Azimullah, S.; Khair, S.B.A.; Haque, M.E. Neuroprotective potential of ferulic acid in the rotenone model of Parkinson’s disease. Drug Des. Devel. Ther. 2015, 9, 5499–5510. [Google Scholar]
  57. Singh, S.; Arthur, R.; Upadhayay, S.; Kumar, P. Ferulic acid ameliorates neurodegeneration via the Nrf2/ARE signalling pathway: A Review. Pharmacol. Res.-Modern Chinese Med. 2022, 5, 100190. [Google Scholar] [CrossRef]
  58. Thapliyal, S.; Singh, T.; Handu, S.; Bisht, M.; Kumari, P.; Arya, P.; Srivastava, P.; Gandham, R. A Review on Potential Footprints of Ferulic Acid for Treatment of Neurological Disorders. Neurochem. Res. 2021, 46, 1043–1057. [Google Scholar] [CrossRef]
  59. Yin, C.L.; Lu, R.G.; Zhu, J.F.; Huang, H.M.; Liu, X.; Li, Q.F.; Mo, Y.Y.; Zhu, H.J.; Chin, B.; Wu, J.X.; et al. The study of neuroprotective effect of ferulic acid based on cell metabolomics. Eur. J. Pharmacol. 2019, 864, 172694. [Google Scholar] [CrossRef] [PubMed]
  60. Alam, M.A. Anti-hypertensive Effect of Cereal Antioxidant Ferulic Acid and Its Mechanism of Action. Front. Nutr. 2019, 6, 121. [Google Scholar] [CrossRef] [PubMed]
  61. Ardiansyah; Ohsaki, Y.; Shirakawa, H.; Koseki, T.; Komai, M. Novel effects of a single administration of ferulic acid on the regulation of blood pressure and the hepatic lipid metabolic profile in stroke-prone spontaneously hypertensive rats. J. Agric. Food Chem. 2008, 56, 2825–2830. [Google Scholar] [CrossRef] [PubMed]
  62. El-Bassossy, H.; Badawy, D.; Neamatallah, T.; Fahmy, A. Ferulic acid, a natural polyphenol, alleviates insulin resistance and hypertension in fructose fed rats: Effect on endothelial-dependent relaxation. Chem. Biol. Interact. 2016, 254, 191–197. [Google Scholar] [CrossRef] [PubMed]
  63. Suzuki, A.; Yamamoto, M.; Jokura, H.; Fujii, A.; Tokimitsu, I.; Hase, T.; Saito, I. Ferulic Acid Restores Endothelium-Dependent Vasodilation in Aortas of Spontaneously Hypertensive Rats. Am. J. Hypertens. 2007, 20, 508–513. [Google Scholar] [CrossRef] [PubMed]
  64. Lenzi, J.; Rodrigues, A.F.; Rós, A.S.; de Castro, B.B.; de Lima, D.D.; Magro, D.D.D.; Zeni, A.L.B. Ferulic acid chronic treatment exerts antidepressant-like effect: Role of antioxidant defense system. Metab. Brain Dis. 2015, 30, 1453–1463. [Google Scholar] [CrossRef]
  65. Chen, J.; Lin, D.; Zhang, C.; Li, G.; Zhang, N.; Ruan, L.; Yan, Q.; Li, J.; Yu, X.; Xie, X.; et al. Antidepressant-like effects of ferulic acid: Involvement of serotonergic and norepinergic systems. Metab. Brain Dis. 2015, 30, 129–136. [Google Scholar] [CrossRef]
  66. Deng, L.; Zhou, X.; Tao, G.; Hao, W.; Wang, L.; Lan, Z.; Song, Y.; Wu, M.; Huang, J.Q. Ferulic acid and feruloylated oligosaccharides alleviate anxiety and depression symptom via regulating gut microbiome and microbial metabolism. Food Res. Int. 2022, 162, 111887. [Google Scholar] [CrossRef]
  67. Sasaki, K.; Iwata, N.; Ferdousi, F.; Isoda, H. Antidepressant-Like Effect of Ferulic Acid via Promotion of Energy Metabolism Activity. Mol. Nutr. Food Res. 2019, 63, e1900327. [Google Scholar] [CrossRef] [Green Version]
  68. Singh, T.; Kaur, T.; Goel, R.K. Ferulic Acid Supplementation for Management of Depression in Epilepsy. Neurochem. Res. 2017, 42, 2940–2948. [Google Scholar] [CrossRef]
  69. Zheng, X.; Cheng, Y.; Chen, Y.; Yue, Y.; Li, Y.; Xia, S.; Li, Y.; Deng, H.; Zhang, J.; Cao, Y. Ferulic acid improves depressive-like behavior in prenatally-stressed offspring rats via anti-inflammatory activity and HPA axis. Int. J. Mol. Sci. 2019, 20, 493. [Google Scholar] [CrossRef] [Green Version]
  70. Krishnan, D.N.; Prasanna, N.; Sabina, E.P.; Rasool, M.K. Hepatoprotective and antioxidant potential of ferulic acid against acetaminophen-induced liver damage in mice. Comp. Clin. Path. 2013, 22, 1177–1181. [Google Scholar] [CrossRef]
  71. Esmat, M.A.; Osman, A.; Hassan, R.E.; Hagag, S.A.; El-maghraby, T.K. Hepatoprotective effect of ferulic acid and/or low doses of γ-irradiation against cisplatin-induced liver injury in rats. Hum. Exp. Toxicol. 2022, 41, 9603271221136205. [Google Scholar] [CrossRef]
  72. Gerin, F.; Erman, H.; Erboga, M.; Sener, U.; Yilmaz, A.; Seyhan, H.; Gurel, A. The Effects of Ferulic Acid Against Oxidative Stress and Inflammation in Formaldehyde-Induced Hepatotoxicity. Inflammation 2016, 39, 1377–1386. [Google Scholar] [CrossRef]
  73. Hussein, R.M.; Anwar, M.M.; Farghaly, H.S.; Kandeil, M.A. Gallic acid and ferulic acid protect the liver from thioacetamide-induced fibrosis in rats via differential expression of miR-21, miR-30 and miR-200 and impact on TGF-β1/Smad3 signaling. Chem. Biol. Interact. 2020, 324, 109098. [Google Scholar] [CrossRef] [PubMed]
  74. Luo, Z.; Li, M.; Yang, Q.; Zhang, Y.; Liu, F.; Gong, L.; Han, L.; Wang, M. Ferulic Acid Prevents Nonalcoholic Fatty Liver Disease by Promoting Fatty Acid Oxidation and Energy Expenditure in C57BL/6 Mice Fed a High-Fat Diet. Nutrients 2022, 14, 2530. [Google Scholar] [CrossRef]
  75. Ma, Y.; Chen, K.; Lv, L.; Wu, S.; Guo, Z. Ferulic acid ameliorates nonalcoholic fatty liver disease and modulates the gut microbiota composition in high-fat diet fed ApoE −/− mice. Biomed. Pharmacother. 2019, 113, 108753. [Google Scholar] [CrossRef]
  76. Mahmoud, A.M.; Hussein, O.E.; Hozayen, W.G.; Bin-Jumah, M.; Abd El-Twab, S.M. Ferulic acid prevents oxidative stress, inflammation, and liver injury via upregulation of Nrf2/HO-1 signaling in methotrexate-induced rats. Environ. Sci. Pollut. Res. 2020, 27, 7910–7921. [Google Scholar] [CrossRef]
  77. Roghani, M.; Kalantari, H.; Khodayar, M.J.; Khorsandi, L.; Kalantar, M.; Goudarzi, M.; Kalantar, H. Alleviation of liver dysfunction, oxidative stress and inflammation underlies the protective effect of ferulic acid in methotrexate-induced hepatotoxicity. Drug Des. Devel. Ther. 2020, 14, 1933–1941. [Google Scholar] [CrossRef]
  78. Tawfik, M.S.; Saif-Elnasr, M.; Elkady, A.A.; Alkady, M.M.; Hawas, A.M. Protective role of ferulic acid against the damaging effect induced by electromagnetic waves on rat liver and intestine tissues. Int. J. Radiat. Res. 2018, 16, 421–430. [Google Scholar]
  79. Wu, J.; Xue, X.; Fan, G.; Gu, Y.; Zhou, F.; Zheng, Q.; Liu, R.; Li, Y.; Ma, B.; Li, S.; et al. Ferulic Acid Ameliorates Hepatic Inflammation and Fibrotic Liver Injury by Inhibiting PTP1B Activity and Subsequent Promoting AMPK Phosphorylation. Front. Pharmacol. 2021, 12, 754976. [Google Scholar] [CrossRef]
  80. Wu, J.; Zhou, F.; Fan, G.; Liu, J.; Wang, Y.; Xue, X.; Lyu, X.; Lin, S.; Li, X. Ferulic acid ameliorates acetaminophen-induced acute liver injury by promoting AMPK-mediated protective autophagy. IUBMB Life 2022, 74, 880–895. [Google Scholar] [CrossRef] [PubMed]
  81. Xu, T.; Song, Q.; Zhou, L.; Yang, W.; Wu, X.; Qian, Q.; Chai, H.; Han, Q.; Pan, H.; Dou, X.; et al. Ferulic acid alleviates lipotoxicity-induced hepatocellular death through the SIRT1-regulated autophagy pathway and independently of AMPK and Akt in AML-12 hepatocytes. Nutr. Metab. 2021, 18, 13. [Google Scholar] [CrossRef] [PubMed]
  82. Bairagi, U.; Mittal, P.; Singh, J.; Mishra, B. Preparation, characterization, and in vivo evaluation of nano formulations of ferulic acid in diabetic wound healing. Drug Dev. Ind. Pharm. 2018, 44, 1783–1796. [Google Scholar] [CrossRef] [PubMed]
  83. Ghosh, S.; Chowdhury, S.; Sarkar, P.; Sil, P.C. Ameliorative role of ferulic acid against diabetes associated oxidative stress induced spleen damage. Food Chem. Toxicol. 2018, 118, 272–286. [Google Scholar] [CrossRef]
  84. Li, J.; Bai, L.; Ma, H.; Guo, H. Ferulic acid alleviates diabetic cardiomyopathy in mice via decreasing blood glucose, reducing inflammation and down-regulating TLR-4/NF-κB pathway. Latin Am. J. Pharm. 2021, 40, 1445–1450. [Google Scholar]
  85. Li, X.; Wu, J.; Xu, F.; Chu, C.; Li, X.; Shi, X.; Zheng, W.; Wang, Z.; Jia, Y.; Xiao, W. Use of Ferulic Acid in the Management of Diabetes Mellitus and Its Complications. Molecules 2022, 27, 6010. [Google Scholar] [CrossRef]
  86. Panwar, R.; Raghuwanshi, N.; Srivastava, A.K.; Sharma, A.K.; Pruthi, V. In-vivo sustained release of nanoencapsulated ferulic acid and its impact in induced diabetes. Mater. Sci. Eng. C 2018, 92, 381–392. [Google Scholar] [CrossRef]
  87. Salau, V.F.; Erukainure, O.L.; Olofinsan, K.O.; Bharuth, V.; Ijomone, O.M.; Islam, M.S. Ferulic acid improves glucose homeostasis by modulation of key diabetogenic activities and restoration of pancreatic architecture in diabetic rats. Fundam. Clin. Pharmacol. 2023, 37, 324–339. [Google Scholar] [CrossRef]
  88. Zhao, J.; Gao, J.; Li, H. Ferulic acid confers protection on islet β cells and placental tissues of rats with gestational diabetes mellitus. Cell. Mol. Biol. 2020, 66, 37–41. [Google Scholar] [CrossRef]
  89. Hasanvand, A.; Kharazmkia, A.; Mir, S.; Khorramabadi, R.M.; Darabi, S. Ameliorative effect of ferulic acid on gentamicin-induced nephrotoxicity in a rat model; role of antioxidant effects. J. Re. Inj. Prev. 2018, 7, 73–77. [Google Scholar] [CrossRef]
  90. Adeyemi, O.S.; Atolani, O.; Banerjee, P.; Arolasafe, G.; Preissner, R.; Etukudoh, P.; Ibraheem, O. Computational and experimental validation of antioxidant properties of synthesized bioactive ferulic acid derivatives. Int. J. Food Prop. 2018, 21, 101–113. [Google Scholar] [CrossRef]
  91. Adeyemi, O.S.; Awakan, O.J.; Atolani, O.; Iyeye, C.O.; Oweibo, O.O.; Adejumo, O.J.; Ibrahim, A.; Batiha, G.E.S. New ferulic acid derivatives protect against carbon tetrachloride-induced liver injury in rats. Open Biochem. J. 2019, 13, 13–22. [Google Scholar] [CrossRef] [Green Version]
  92. Bautista-Aguilera, O.M.; Alonso, J.M.; Catto, M.; Iriepa, I.; Knez, D.; Gobec, S.; Marco-Contelles, J. N-Hydroxy-N-Propargylamide Derivatives of Ferulic Acid: Inhibitors of Cholinesterases and Monoamine Oxidases. Molecules 2022, 27, 7437. [Google Scholar] [CrossRef]
  93. Borgohain, R.; Handique, J.G.; Guha, A.K.; Pratihar, S. A theoretical study on antioxidant activity of ferulic acid and its ester derivatives. J. Theor. Comput. Chem. 2016, 15, 1650028. [Google Scholar] [CrossRef]
  94. Cui, M.Y.; Xiao, M.W.; Xu, L.J.; Chen, Y.; Liu, A.L.; Ye, J.; Hu, A.X. Bioassay of ferulic acid derivatives as influenza neuraminidase inhibitors. Arch. Pharm. 2020, 353, e1900174. [Google Scholar] [CrossRef] [PubMed]
  95. de Paiva, L.B.; Goldbeck, R.; dos Santos, W.D.; Squina, F.M. Ferulic acid and derivatives: Molecules with potential application in the pharmaceutical field. Braz. J. Pharm. Sci. 2013, 49, 395–411. [Google Scholar] [CrossRef] [Green Version]
  96. Drăgan, M.; Stan, C.D.; Iacob, A.; Profire, L. Assessment of in vitro antioxidant and anti-inflammatory activities of new azetidin-2-one derivatives of ferulic acid. Farmacia 2016, 64, 717–721. [Google Scholar]
  97. Drăgan, M.; Stan, C.D.; Iacob, A.T.; Dragostin, O.M.; Boancă, M.; Lupuşoru, C.E.; Zamfir, C.L.; Profire, L. Biological evaluation of azetidine-2-one derivatives of ferulic acid as promising anti-inflammatory agents. Processes 2020, 8, 1401. [Google Scholar] [CrossRef]
  98. Ekowati, J.; Diyah, N.W.; Nofianti, K.A.; Hamid, I.S.; Siswandono. Molecular docking of ferulic acid derivatives on P2Y12 receptor and their ADMET prediction. J. Math. Fundam. Sci. 2018, 50, 203–219. [Google Scholar] [CrossRef]
  99. Hernández-García, L.; Sandoval-Lira, J.; Rosete-Luna, S.; Niño-Medina, G.; Sanchez, M. Theoretical study of ferulic acid dimer derivatives: Bond dissociation enthalpy, spin density, and HOMO-LUMO analysis. Struct. Chem. 2018, 29, 1265–1272. [Google Scholar] [CrossRef]
  100. Jung, J.S.; Yan, J.J.; Li, H.M.; Sultan, M.T.; Yu, J.; Lee, H.S.; Shin, K.J.; Song, D.K. Protective effects of a dimeric derivative of ferulic acid in animal models of Alzheimer’s disease. Eur. J. Pharmacol. 2016, 782, 30–34. [Google Scholar] [CrossRef]
  101. Khatkar, A.; Nanda, A.; Kumar, P.; Narasimhan, B. Synthesis and antimicrobial evaluation of ferulic acid derivatives. Res. Chem. Intermed. 2015, 41, 299–309. [Google Scholar] [CrossRef]
  102. Kikugawa, M.; Tsutsuki, H.; Ida, T.; Nakajima, H.; Ihara, H.; Sakamoto, T. Water-soluble ferulic acid derivatives improve amyloid-β-induced neuronal cell death and dysmnesia through inhibition of amyloid-β aggregation. Biosci. Biotechnol. Biochem. 2016, 80, 547–553. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Kolaj, I.; Wang, Y.; Ye, K.; Meek, A.; Liyanage, S.I.; Santos, C.; Weaver, D.F. Ferulic acid amide derivatives with varying inhibition of amyloid-β oligomerization and fibrillization. Bioorg. Med. Chem. 2021, 43, 116247. [Google Scholar] [CrossRef] [PubMed]
  104. Kumar, N.; Goel, N.; Chand Yadav, T.; Pruthi, V. Quantum chemical, ADMET and molecular docking studies of ferulic acid amide derivatives with a novel anticancer drug target. Med. Chem. Res. 2017, 26, 1822–1834. [Google Scholar] [CrossRef]
  105. Lan, J.S.; Zeng, R.F.; Jiang, X.Y.; Hou, J.W.; Liu, Y.; Hu, Z.H.; Li, H.X.; Li, Y.; Xie, S.S.; Ding, Y.; et al. Design, synthesis and evaluation of novel ferulic acid derivatives as multi-target-directed ligands for the treatment of Alzheimer’s disease. Bioorg. Chem. 2020, 94, 103413. [Google Scholar] [CrossRef]
  106. Li, D.; Rui, Y.X.; Guo, S.D.; Luan, F.; Liu, R.; Zeng, N. Ferulic acid: A review of its pharmacology, pharmacokinetics and derivatives. Life Sci. 2021, 284, 119921. [Google Scholar] [CrossRef]
  107. Li, W.; Li, N.; Tang, Y.; Li, B.; Liu, L.; Zhang, X.; Fu, H.; Duan, J.A. Biological activity evaluation and structure-activity relationships analysis of ferulic acid and caffeic acid derivatives for anticancer. Bioorg. Med. Chem. Lett. 2012, 22, 6085–6088. [Google Scholar] [CrossRef]
  108. Liang, Y.; Xi, X.; Liu, Q.; Huang, P.; Li, J.; Lin, Q. Research progress on the physiological activity and application of ferulic acid and its derivatives. J. Food Sci. Biotechnol. 2018, 37, 449–454. [Google Scholar]
  109. Malik, S.A.; Ali, K.F.; Dawood, A.H. Synthesis, Characterization, and Preliminary Evaluation of Ferulic Acid Derivatives Containing Heterocyclic Moiety. J. Med. Chem. Sci. 2023, 6, 1444–1456. [Google Scholar]
  110. Montaser, A.; Huttunen, J.; Ibrahim, S.A.; Huttunen, K.M. Astrocyte-Targeted Transporter-Utilizing Derivatives of Ferulic Acid Can Have Multifunctional Effects Ameliorating Inflammation and Oxidative Stress in the Brain. Oxid. Med. Cell. Longev. 2019, 2019, 3528148. [Google Scholar] [CrossRef] [Green Version]
  111. Pasquereau, S.; Galais, M.; Bellefroid, M.; Pachón Angona, I.; Morot-Bizot, S.; Ismaili, L.; Van Lint, C.; Herbein, G. Ferulic acid derivatives block coronaviruses HCoV-229E and SARS-CoV-2 replication in vitro. Sci. Rep. 2022, 12, 20309. [Google Scholar] [CrossRef]
  112. Phadke, A.V.; Tayade, A.A.; Khambete, M.P. Therapeutic potential of ferulic acid and its derivatives in Alzheimer’s disease—A systematic review. Chem. Biol. Drug Des. 2021, 98, 713–721. [Google Scholar] [CrossRef] [PubMed]
  113. Pinheiro, P.; Santiago, G.; Da Silva, F.; De Araujo, A.; De Oliveira, C.; Freitas, P.; Rocha, J.; De Araujo Neto, J.; Da Silva, M.; Tintino, S.; et al. Antibacterial activity and inhibition against Staphylococcus aureus NorA efflux pump by ferulic acid and its esterified derivatives. Asian Pac. J. Tro. Biomed. 2021, 11, 405–413. [Google Scholar]
  114. Pinheiro, P.G.; Santiago, G.M.P.; da Silva, F.E.F.; de Araújo, A.C.J.; de Oliveira, C.R.T.; Freitas, P.R.; Rocha, J.E.; Neto, J.B.D.A.; da Silva, M.M.C.; Tintino, S.R.; et al. Ferulic acid derivatives inhibiting Staphylococcus aureus tetK and MsrA efflux pumps. Biotechnol. Rep. 2022, 34, e00717. [Google Scholar] [CrossRef] [PubMed]
  115. Sang, Z.; Pan, W.; Wang, K.; Ma, Q.; Yu, L.; Yang, Y.; Bai, P.; Leng, C.; Xu, Q.; Li, X.; et al. Design, synthesis and evaluation of novel ferulic acid-O-alkylamine derivatives as potential multifunctional agents for the treatment of Alzheimer’s disease. Eur. J. Med. Chem. 2017, 130, 379–392. [Google Scholar] [CrossRef]
  116. Sang, Z.; Wang, K.; Han, X.; Cao, M.; Tan, Z.; Liu, W. Design, Synthesis, and Evaluation of Novel Ferulic Acid Derivatives as Multi-Target-Directed Ligands for the Treatment of Alzheimer’s Disease. ACS Chem. Neurosci. 2019, 10, 1008–1024. [Google Scholar] [CrossRef]
  117. Senthil, R.; Sakthivel, M.; Usha, S. Structure-based drug design of peroxisome proliferator-activated receptor gamma inhibitors: Ferulic acid and derivatives. J. Biomol. Struct. Dyn. 2021, 39, 1295–1311. [Google Scholar] [CrossRef]
  118. Serafim, T.L.; Carvalho, F.S.; Marques, M.P.M.; Calheiros, R.; Silva, T.; Garrido, J.; Milhazes, N.; Borges, F.; Roleira, F.; Silva, E.T.; et al. Lipophilic caffeic and ferulic acid derivatives presenting cytotoxicity against human breast cancer cells. Chem. Res. Toxicol. 2011, 24, 763–774. [Google Scholar] [CrossRef] [Green Version]
  119. Shi, Y.; Chen, X.; Qiang, S.; Su, J.; Li, J. Anti-oxidation and anti-inflammatory potency evaluation of ferulic acid derivatives obtained through virtual screening. Int. J. Mol. Sci. 2021, 22, 11305. [Google Scholar] [CrossRef]
  120. Wang, D.; Guo, D.; Tang, Y.; Qi, M.; Fang, J.; Zhang, Y.; Chai, Y.; Cao, Y.; Lv, D. A multi-omics study of the anti-cancer effect of a ferulic acid derivative FA-30. Mol. Omics 2022, 18, 805–813. [Google Scholar] [CrossRef]
  121. Wang, F.; Yang, L.; Huang, K.; Li, X.; Hao, X.; Stöckigt, J.; Zhao, Y. Preparation of ferulic acid derivatives and evaluation of their xanthine oxidase inhibition activity. Nat. Prod. Res. 2007, 21, 196–202. [Google Scholar] [CrossRef]
  122. Wang, Z.; Xie, D.; Gan, X.; Zeng, S.; Zhang, A.; Yin, L.; Song, B.; Jin, L.; Hu, D. Synthesis, antiviral activity, and molecular docking study of trans-ferulic acid derivatives containing acylhydrazone moiety. Bioorg. Med. Chem. Lett. 2017, 27, 4096–4100. [Google Scholar] [CrossRef]
  123. Wu, J.; Yin, W.; Zhang, Y.; Ye, H.; Li, Y.; Tian, J.; Huang, Z.; Zhang, Y. Design and synthesis of the ring-opened derivative of 3-n-butylphthalide-ferulic acid-glucose trihybrids as potential anti-ischemic agents. Chin. Chem. Lett. 2020, 31, 1881–1886. [Google Scholar] [CrossRef]
  124. Wu, Y.; Shi, Y.G.; Zheng, X.L.; Dang, Y.L.; Zhu, C.M.; Zhang, R.R.; Fu, Y.Y.; Zhou, T.Y.; Li, J.H. Lipophilic ferulic acid derivatives protect PC12 cells against oxidative damage: Via modulating β-amyloid aggregation and activating Nrf2 enzymes. Food Funct. 2020, 11, 4707–4718. [Google Scholar] [CrossRef]
  125. Wu, Z.; Zhang, J.; Chen, J.; Pan, J.; Zhao, L.; Liu, D.; Zhang, A.; Chen, J.; Hu, D.; Song, B. Design, synthesis, antiviral bioactivity and three-dimensional quantitative structure–activity relationship study of novel ferulic acid ester derivatives containing quinazoline moiety. Pest Manag. Sci. 2017, 73, 2079–2089. [Google Scholar] [CrossRef] [PubMed]
  126. Xie, Y.; Liu, Y.; Sun, J.; Zheng, L. Synthesis of mitochondria-targeted ferulic acid amide derivatives with antioxidant, anti-inflammatory activities and inducing mitophagy. Bioorg. Chem. 2022, 127, 106037. [Google Scholar] [CrossRef] [PubMed]
  127. Yuan, T.; Wang, Z.; Lan, S.; Gan, X. Design, synthesis, antiviral activity, and mechanisms of novel ferulic acid derivatives containing amide moiety. Bioorg. Chem. 2022, 128, 106054. [Google Scholar] [CrossRef]
  128. Yuan, T.; Wang, Z.; Liu, D.; Zeng, H.; Liang, J.; Hu, D.; Gan, X. Ferulic acid derivatives with piperazine moiety as potential antiviral agents. Pest Manag. Sci. 2022, 78, 1749–1758. [Google Scholar] [CrossRef]
  129. Yue, S.J.; Zhang, P.X.; Zhu, Y.; Li, N.G.; Chen, Y.Y.; Li, J.J.; Zhang, S.; Jin, R.Y.; Yan, H.; Shi, X.Q.; et al. A ferulic acid derivative FXS-3 inhibits proliferation and metastasis of human lung cancer A549 cells via positive JNK signaling pathway and negative ERK/p38, AKt/mTOR and MEK/ERK signaling pathways. Molecules 2019, 24, 2165. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Zhang, P.X.; Lin, H.; Qu, C.; Tang, Y.P.; Li, N.G.; Kai, J.; Shang, G.; Li, B.; Zhang, L.; Yan, H.; et al. Design, synthesis, and in vitro antiplatelet aggregation activities of ferulic acid derivatives. J. Chem. 2015, 2015, 376527. [Google Scholar] [CrossRef] [Green Version]
  131. Kong, H.; Fu, X.; Chang, X.; Ding, Z.; Yu, Y.; Xu, H.; Wang, R.; Shan, Y.; Ding, S. The ester derivatives of ferulic acid exhibit strong inhibitory effect on the growth of Alternaria alternata in vitro and in vivo. Postharvest Biol. Technol. 2023, 196, 112158. [Google Scholar] [CrossRef]
  132. Pellerito, C.; Emanuele, S.; Ferrante, F.; Celesia, A.; Giuliano, M.; Fiore, T. Tributyltin(IV) ferulate, a novel synthetic ferulic acid derivative, induces autophagic cell death in colon cancer cells: From chemical synthesis to biochemical effects. J. Inorg. Biochem. 2020, 205, 110999. [Google Scholar] [CrossRef] [PubMed]
  133. Wang, F.; Peng, Q.; Liu, J.; Alolga, R.N.; Zhou, W. A novel ferulic acid derivative attenuates myocardial cell hypoxia reoxygenation injury through a succinate dehydrogenase dependent antioxidant mechanism. Eur. J. Pharmacol. 2019, 856, 172417. [Google Scholar] [CrossRef]
  134. Gan, X.; Zhang, W.; Lan, S.; Hu, D. Novel Cyclized Derivatives of Ferulic Acid as Potential Antiviral Agents through Activation of Photosynthesis. J. Agric. Food Chem. 2023, 71, 1369–1380. [Google Scholar] [CrossRef] [PubMed]
  135. Machado, K.C.; Oliveira, G.L.S.; Islam, M.T.; Junior, A.L.G.; De Sousa, D.P.; Freitas, R.M. Anticonvulsant and behavioral effects observed in mice following treatment with an ester derivative of ferulic acid: Isopentyl ferulate. Chem. Biol. Interact. 2015, 242, 273–279. [Google Scholar] [CrossRef]
  136. Guzman-Lopez, E.G.; Reina, M.; Perez-Gonzalez, A.; Francisco-Marquez, M.; Hernandez-Ayala, L.F.; Castañeda-Arriaga, R.; Galano, A. CADMA-Chem: A Computational Protocol Based on Chemical Properties Aimed to Design Multifunctional Antioxidants. Int. J. Mol. Sci. 2022, 23, 13246. [Google Scholar] [CrossRef]
  137. Pérez-González, A.; Castañeda-Arriaga, R.; Guzmán-López, E.G.; Hernández-Ayala, L.F.; Galano, A. Chalcone Derivatives with a High Potential as Multifunctional Antioxidant Neuroprotectors. ACS Omega 2022, 7, 38254–38268. [Google Scholar] [CrossRef]
  138. Reina, M.; Guzmán-López, E.G.; Galano, A. Computational design of rasagiline derivatives: Searching for enhanced antioxidant capability. Int. J. Quantum Chem. 2023, 123, e72011. [Google Scholar] [CrossRef]
  139. Marucci, G.; Buccioni, M.; Ben, D.D.; Lambertucci, C.; Volpini, R.; Amenta, F. Efficacy of acetylcholinesterase inhibitors in Alzheimer’s disease. Neuropharmacology 2021, 190, 108352. [Google Scholar] [CrossRef]
  140. Moreta, M.P.G.; Burgos-Alonso, N.; Torrecilla, M.; Marco-Contelles, J.; Bruzos-Cidón, C. Efficacy of acetylcholinesterase inhibitors on cognitive function in alzheimer’s disease. Review of reviews. Biomedicines 2021, 9, 1689. [Google Scholar] [CrossRef]
  141. Uddin, M.S.; Al Mamun, A.; Kabir, M.T.; Ashraf, G.M.; Bin-Jumah, M.N.; Abdel-Daim, M.M. Multi-Target Drug Candidates for Multifactorial Alzheimer’s Disease: AChE and NMDAR as Molecular Targets. Mol. Neurobiol. 2021, 58, 281–303. [Google Scholar] [CrossRef] [PubMed]
  142. Finberg, J.P.M. Inhibitors of MAO-B and COMT: Their effects on brain dopamine levels and uses in Parkinson’s disease. J. Neural Transm. 2019, 126, 433–448. [Google Scholar] [CrossRef]
  143. Jost, W.H. A critical appraisal of MAO-B inhibitors in the treatment of Parkinson’s disease. J. Neural Transm. 2022, 129, 723–736. [Google Scholar] [CrossRef] [PubMed]
  144. Özdemir, Z.; Alagöz, M.A.; Bahçecioğlu, Ö.F.; Gök, S. Monoamine oxidase-B (MAO-B) inhibitors in the treatment of alzheimer’s and parkinson’s disease. Curr. Med. Chem. 2021, 28, 6045–6065. [Google Scholar] [CrossRef] [PubMed]
  145. Parambi, D.G.T. Treatment of parkinson’s disease by MAO-B inhibitors, new therapies and future challenges-A mini-review. Comb. Chem. High Throughput Screen. 2020, 23, 847–861. [Google Scholar] [CrossRef]
  146. Müller, T. Catechol-O-methyltransferase inhibitors in Parkinson’s disease. Drugs 2015, 75, 157–174. [Google Scholar] [CrossRef]
  147. Nakamagoe, K.; Tsuji, H.; Ishii, K.; Tamaoka, A. Remarkable clinical responses of non-fluctuating Parkinson’s disease (PD) after alternating catechol O-methyltransferase inhibitors: Case series switching from entacapone 200~300 mg/day to opicapone 25 mg/day. Neurol. Sci. 2021, 42, 4813–4814. [Google Scholar] [CrossRef]
  148. Fabbri, M.; Ferreira, J.J.; Rascol, O. COMT Inhibitors in the Management of Parkinson’s Disease. CNS Drugs 2022, 36, 261–282. [Google Scholar] [CrossRef]
  149. St. Onge, E.; Vanderhoof, M.; Miller, S. Opicapone (Ongentys): A New COMT Inhibitor for the Treatment of Parkinson’s Disease. Ann. Pharmacother. 2021, 55, 1159–1166. [Google Scholar] [CrossRef]
  150. Calculation of molecular properties and bioactivity score. Available online: https://www.molinspiration.com/cgi-bin/properties (accessed on 15 January 2023).
  151. Lipinski, C.A.; Lombardo, F.; Dominy, B.W.; Feeney, P.J. Experimental and computational approaches to estimate solubility and permeability in drug discovery and development settings. Adv. Drug Delivery Rev. 2001, 46, 3–26. [Google Scholar] [CrossRef]
  152. Ghose, A.K.; Viswanadhan, V.N.; Wendoloski, J.J. A Knowledge-Based Approach in Designing Combinatorial or Medicinal Chemistry Libraries for Drug Discovery. 1. A Qualitative and Quantitative Characterization of Known Drug Databases. J. Comb. Chem. 1999, 1, 55–68. [Google Scholar] [CrossRef] [PubMed]
  153. Veber, D.F.; Johnson, S.R.; Cheng, H.Y.; Smith, B.R.; Ward, K.W.; Kopple, K.D. Molecular Properties That Influence the Oral Bioavailability of Drug Candidates. J. Med. Chem. 2002, 45, 2615–2623. [Google Scholar] [CrossRef] [PubMed]
  154. Boda, K.; Seidel, T.; Gasteiger, J. Structure and reaction based evaluation of synthetic accessibility. J. Comput.-Aided Mol. Des. 2007, 21, 311–325. [Google Scholar] [CrossRef]
  155. Bonnet, P. Is chemical synthetic accessibility computationally predictable for drug and lead-like molecules? A comparative assessment between medicinal and computational chemists. Eur. J. Med. Chem. 2012, 54, 679–689. [Google Scholar] [CrossRef]
  156. Zhu, H.; Tropsha, A.; Fourches, D.; Varnek, A.; Papa, E.; Gramatica, P.; Öberg, T.; Dao, P.; Cherkasov, A.; Tetko, I.V. Combinatorial QSAR Modeling of Chemical Toxicants Tested against Tetrahymena pyriformis. J. Chem. Inf. Mode. 2008, 48, 766–784. [Google Scholar] [CrossRef] [Green Version]
  157. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16 Rev. C.01; Gaussian Inc.: Pittsburgh, PA, USA, 2016. [Google Scholar]
  158. Marenich, A.V.; Cramer, C.J.; Truhlar, D.G. Universal solvation model based on solute electron density and on a continuum model of the solvent defined by the bulk dielectric constant and atomic surface tensions. J. Phys. Chem. B 2009, 113, 6378–6396. [Google Scholar] [CrossRef]
  159. Zhao, Y.; Schultz, N.E.; Truhlar, D.G. Design of Density Functionals by Combining the Method of Constraint Satisfaction with Parametrization for Thermochemistry, Thermochemical Kinetics, and Noncovalent Interactions. J. Chem. Theory Comput. 2006, 2, 364–382. [Google Scholar] [CrossRef] [PubMed]
  160. Wu, W.h.; Lei, P.; Liu, Q.; Hu, J.; Gunn, A.P.; Chen, M.s.; Rui, Y.f.; Su, X.y.; Xie, Z.p.; Zhao, Y.F. Sequestration of Copper from β-Amyloid Promotes Selective Lysis by Cyclen-Hybrid Cleavage Agents. J. Biol. Chem. 2008, 283, 31657–31664. [Google Scholar] [CrossRef] [Green Version]
  161. Milenković, D.; Dorović, J.; Jeremić, S.; Dimitrić Marković, J.M.; Avdović, E.H.; Marković, Z. Free Radical Scavenging Potency of Dihydroxybenzoic Acids. J. Chem. 2017, 2017, 5936239. [Google Scholar] [CrossRef] [Green Version]
  162. Amić, A.; Marković, Z.; Dimitrić Marković, J.M.; Lučić, B.; Stepanić, V.; Amić, D. The 2H+/2e- free radical scavenging mechanisms of uric acid: Thermodynamics of N-H bond cleavage. Comput. Theor. Chem. 2016, 1077, 2–10. [Google Scholar] [CrossRef]
  163. Dorović, J.; Marković, J.M.D.; Stepanić, V.; Begović, N.; Amić, D.; Marković, Z. Influence of different free radicals on scavenging potency of gallic acid. J. Mol. Model. 2014, 20, 2345. [Google Scholar] [CrossRef]
  164. Marković, Z.; Crossed, D.; Signorović, J.; Dekić, M.; Radulović, M.; Marković, S.; Ilić, M. DFT study of free radical scavenging activity of erodiol. Chem. Pap. 2013, 67, 1453–1461. [Google Scholar] [CrossRef]
  165. Galano, A.; Alvarez-Idaboy, J.R.; Francisco-Márquez, M. Physicochemical Insights on the Free Radical Scavenging Activity of Sesamol: Importance of the Acid/Base Equilibrium. J. Phys. Chem. B 2011, 115, 13101–13109. [Google Scholar] [CrossRef] [PubMed]
  166. Ortiz, J.V. Toward an Exact One-Electron Picture of Chemical Bonding. Adv. Quantum Chem. 1999, 35, 33–52. [Google Scholar]
  167. Ortiz, J.V. Electron propagator theory: An approach to prediction and interpretation in quantum chemistry. Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2013, 3, 123–142. [Google Scholar] [CrossRef]
  168. Pérez-González, A.; Galano, A.; Ortiz, J.V. Vertical ionization energies of free radicals and electron detachment energies of their anions: A comparison of direct and indirect methods versus experiment. J. Phys. Chem. A 2014, 118, 6125–6131. [Google Scholar] [CrossRef]
  169. Ortiz, J.V. The Electron Propagator Picture of Molecular Electronic Structure. In Computational Chemistry: Reviews of Current Trends; Computational Chemistry: Reviews of Current Trends; World Scientific: Singapore, 1997; Volume 2, pp. 1–61. [Google Scholar]
  170. Marvin; 23.4.0; Chemaxon Ltd.: Budapest, Hungary, 2023.
  171. Galano, A.; Pérez-González, A.; Castañeda-Arriaga, R.; Muñoz-Rugeles, L.; Mendoza-Sarmiento, G.; Romero-Silva, A.; Ibarra-Escutia, A.; Rebollar-Zepeda, A.M.; León-Carmona, J.R.; Hernández-Olivares, M.A.; et al. Empirically Fitted Parameters for Calculating pKaValues with Small Deviations from Experiments Using a Simple Computational Strategy. J. Chem. Inf. Model. 2016, 56, 1714–1724. [Google Scholar] [CrossRef]
  172. Pérez-González, A.; Castañeda-Arriaga, R.; Verastegui, B.; Carreón-González, M.; Alvarez-Idaboy, J.R.; Galano, A. Estimation of empirically fitted parameters for calculating pK a values of thiols in a fast and reliable way. Theor. Chem. Acc. 2018, 137, 5. [Google Scholar] [CrossRef]
  173. Ozkorucuklu, S.P.; Beltrán, J.L.; Fonrodona, G.; Barrón, D.; Alsancak, G.; Barbosa, J. Determination of dissociation constants of some hydroxylated benzoic and cinnamic acids in water from mobility and spectroscopic data obtained by CE-DAD. J. Chem. Eng. Data 2009, 54, 807–811. [Google Scholar] [CrossRef]
  174. Ellermann, M.; Lerner, C.; Burgy, G.; Ehler, A.; Bissantz, C.; Jakob-Roetne, R.; Paulini, R.; Allemann, O.; Tissot, H.; Grünstein, D.; et al. Catechol-O-methyltransferase in complex with substituted 3′-deoxyribose bisubstrate inhibitors. Acta Crystallogr. Sect. D. Biol. Crystallogr. 2012, 68, 253–260. [Google Scholar] [CrossRef]
  175. Binda, C.; Wang, J.; Pisani, L.; Caccia, C.; Carotti, A.; Salvati, P.; Edmondson, D.E.; Mattevi, A. Structures of human monoamine oxidase B complexes with selective noncovalent inhibitors: Safinamide and coumarin analogs. J. Med. Chem. 2007, 50, 5848–5852. [Google Scholar] [CrossRef]
  176. Cheung, J.; Rudolph, M.J.; Burshteyn, F.; Cassidy, M.S.; Gary, E.N.; Love, J.; Franklin, M.C.; Height, J.J. Structures of human acetylcholinesterase in complex with pharmacologically important ligands. J. Med. Chem. 2012, 55, 10282–10286. [Google Scholar] [CrossRef] [PubMed]
  177. Šali, A.; Blundell, T.L. Comparative protein modelling by satisfaction of spatial restraints. J. Mol. Biol. 1993, 234, 779–815. [Google Scholar] [CrossRef] [PubMed]
  178. BIOVIA. Available online: https://www.3ds.com/products-services/biovia/ (accessed on 22 February 2023).
  179. Trott, O.; Olson, A.J. AutoDock Vina: Improving the speed and accuracy of docking with a new scoring function, efficient optimization, and multithreading. J. Comput. Chem. 2010, 31, 455–461. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. The PyMOL Molecular Graphics System; 2.0; DeLano Scientific LLC: New York, NY, USA, 2015.
  181. Pettersen, E.F.; Goddard, T.D.; Huang, C.C.; Couch, G.S.; Greenblatt, D.M.; Meng, E.C.; Ferrin, T.E. UCSF Chimera--a visualization system for exploratory research and analysis. J. Comput. Chem. 2004, 25, 1605–1612. [Google Scholar] [CrossRef] [Green Version]
  182. Reina, M.; Guzmán-López, E.G.; Romeo, I.; Marino, T.; Russo, N.; Galano, A. Computationally designed: P -coumaric acid analogs: Searching for neuroprotective antioxidants. New J. Chem. 2021, 45, 14369–14380. [Google Scholar] [CrossRef]
  183. 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] [Green Version]
  184. Reina, M.; Castañeda-Arriaga, R.; Pérez-González, A.; Guzmán-López, E.G.; Tan, D.X.; Reiter, R.; Galano, A. A computer-assisted systematic search for melatonin derivatives with high potential as antioxidants. Melatonin Res. 2018, 1, 27–58. [Google Scholar] [CrossRef] [Green Version]
  185. Gleeson, M.P.; Hersey, A.; Montanari, D.; Overington, J. Probing the links between in vitro potency, ADMET and physicochemical parameters. Nat. Rev. Drug Discov. 2011, 10, 197–208. [Google Scholar] [CrossRef]
  186. Zhong, H.; Mashinson, V.; Woolman, T.; Zha, M. Understanding the molecular properties and metabolism of top prescribed drugs. Curr. Top. Med. Chem. 2013, 13, 1290–1307. [Google Scholar] [CrossRef]
  187. Kiss, L.E.; Soares-Da-Silva, P. Medicinal chemistry of catechol O -methyltransferase (COMT) inhibitors and their therapeutic utility. J. Med. Chem. 2014, 57, 8692–8717. [Google Scholar] [CrossRef]
  188. Kaur, T.; Madgulkar, A.; Bhalekar, M.; Asgaonkar, K. Molecular docking in formulation and development. Curr. Drug Disc. Technol. 2019, 16, 30–39. [Google Scholar] [CrossRef]
  189. Shafferman, A.; Kronman, C.; Flashner, Y.; Leitner, M.; Grosfeld, H.; Ordentlich, A.; Gozes, Y.; Cohen, S.; Ariel, N.; Barak, D.; et al. Mutagenesis of human acetylcholinesterase. Identification of residues involved in catalytic activity and in polypeptide folding. J. Biol. Chem. 1992, 267, 17640–17648. [Google Scholar] [CrossRef]
  190. Felder, C.E.; Botti, S.A.; Lifson, S.; Silman, I.; Sussman, J.L. External and internal electrostatic potentials of cholinesterase models. J. Mol. Graph. Model. 1997, 15, 318–327. [Google Scholar] [CrossRef]
  191. Quinn, D.M. Acetylcholinesterase: Enzyme structure, reaction dynamics, and virtual transition states. Chem. Rev. 1987, 87, 955–979. [Google Scholar] [CrossRef]
  192. Colovic, M.B.; Krstic, D.Z.; Lazarevic-Pasti, T.D.; Bondzic, A.M.; Vasic, V.M. Acetylcholinesterase Inhibitors: Pharmacology and Toxicology. Curr. Neuropharmacol. 2013, 11, 315–335. [Google Scholar] [CrossRef] [Green Version]
  193. Edmondson, D.E.; Mattevi, A.; Binda, C.; Li, M.; Hubálek, F. Structure and mechanism of monoamine oxidase. Curr. Med. Chem. 2004, 11, 1983–1993. [Google Scholar] [CrossRef]
  194. Cai, Z. Monoamine oxidase inhibitors: Promising therapeutic agents for Alzheimer’s disease (Review). Mol. Med. Rep. 2014, 9, 1533–1541. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Finberg, J.P.M.; Rabey, J.M. Inhibitors of MAO-A and MAO-B in Psychiatry and Neurology. Front. Pharmacol. 2016, 7, 340. [Google Scholar] [CrossRef] [Green Version]
  196. Massey, V.; Ghisla, S.; Kieschke, K. Studies on the reaction mechanism of lactate oxidase. Formation of two covalent flavin-substrate adducts on reaction with glycollate. J. Biol. Chem. 1980, 255, 2796–2806. [Google Scholar] [CrossRef] [PubMed]
  197. Korn, A.; Eichler, H.G.; Fischbach, R.; Gasic, S. Moclobemide, a new reversible MAO inhibitor-interaction with tyramine and tricyclic antidepressants in healthy volunteers and depressive patients. Psychopharmacology 1986, 88, 153–157. [Google Scholar] [CrossRef] [PubMed]
  198. Fowler, J.S.; Logan, J.; Azzaro, A.J.; Fielding, R.M.; Zhu, W.; Poshusta, A.K.; Burch, D.; Brand, B.; Free, J.; Asgharnejad, M.; et al. Reversible inhibitors of monoamine oxidase-A (RIMAs): Robust, reversible inhibition of human brain MAO-A by CX157. Neuropsychopharmacology 2010, 35, 623–631. [Google Scholar] [CrossRef] [PubMed]
  199. Ma, Z.; Liu, H.; Wu, B. Structure-based drug design of catechol-O-methyltransferase inhibitors for CNS disorders. Br. J. Clin. Pharmacol. 2014, 77, 410–420. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Scheme 1. Ferulic acid (FA, R1 = R2 = R3 = R4 = R4 = H) structure and site numbering used in this work.
Scheme 1. Ferulic acid (FA, R1 = R2 = R3 = R4 = R4 = H) structure and site numbering used in this work.
Antioxidants 12 01256 sch001
Figure 1. Selection score (SS) for the FA derivatives designed in this work. Vertical lines mark the arithmetic mean of the reference set (red) and the value for the parent molecule (FA, green).
Figure 1. Selection score (SS) for the FA derivatives designed in this work. Vertical lines mark the arithmetic mean of the reference set (red) and the value for the parent molecule (FA, green).
Antioxidants 12 01256 g001
Scheme 2. Structure and SS values of FA and the derivatives selected for the next stage of the investigation.
Scheme 2. Structure and SS values of FA and the derivatives selected for the next stage of the investigation.
Antioxidants 12 01256 sch002
Figure 2. Elimination score (SE) for the most promising FA derivatives, according to SS. Columns are divided to show the influence of the new contributions included in each score, with respect to the previous one.
Figure 2. Elimination score (SE) for the most promising FA derivatives, according to SS. Columns are divided to show the influence of the new contributions included in each score, with respect to the previous one.
Antioxidants 12 01256 g002
Figure 3. Individual contributions to the elimination score (SE), for the most promising FA derivatives.
Figure 3. Individual contributions to the elimination score (SE), for the most promising FA derivatives.
Antioxidants 12 01256 g003
Figure 4. The electron and hydrogen donating ability map for antioxidants (eH-DAMA), including the dominant acid-base species of FA derivatives, the parent molecule, Trolox, α-tocopherol, and the H2O2/O2•− oxidant pair.
Figure 4. The electron and hydrogen donating ability map for antioxidants (eH-DAMA), including the dominant acid-base species of FA derivatives, the parent molecule, Trolox, α-tocopherol, and the H2O2/O2•− oxidant pair.
Antioxidants 12 01256 g004
Figure 5. Polygenic scores of the ferulic acid and its derivatives.
Figure 5. Polygenic scores of the ferulic acid and its derivatives.
Antioxidants 12 01256 g005
Figure 6. Interactions in FA-26:AChE (left), FA-26:MAOB (middle) and FA-118:COMT complexes (right). FA-26 and FA-118 are shown using the ball and stick model. Interactions are presented as dotted lines: conventional hydrogen bonds (green), π-stacking (red), π-alkyl (magenta), and C-H non-conventional bonds (cyan).
Figure 6. Interactions in FA-26:AChE (left), FA-26:MAOB (middle) and FA-118:COMT complexes (right). FA-26 and FA-118 are shown using the ball and stick model. Interactions are presented as dotted lines: conventional hydrogen bonds (green), π-stacking (red), π-alkyl (magenta), and C-H non-conventional bonds (cyan).
Antioxidants 12 01256 g006
Table 1. Structural modifications and properties of some FA derivatives.
Table 1. Structural modifications and properties of some FA derivatives.
FunctionalizationBioactivityRef.
3-n-butylphthalide + glucoseAnti-ischemic. [123]
Alkyl estersβ-amyloid aggregation inhibition.[124]
AmideAntiviral[127]
AmideAntioxidant, inflammatory, mitophagy enhancing.[126]
Amideβ-amyloid oligomerization and fibrillization inhibition.[103]
AmideAntioxidant, anticancer.[27]
Amide + pyrazoleAntioxidant and myocardial cell hypoxia reoxygenation.[133]
Amino acidAnti-inflammatory, antioxidant.[110]
AnilineAntimicrobial.[101]
Azetidine-2-oneAnti-inflammatory, antioxidant.[96,97]
Benzyl and phenylethyl estersAnticancer.[29]
Benzylamino, and carbamylβ-amyloid aggregation inhibition, antioxidant, AChE inhibition.[105]
CyclizedAntiviral.[134]
Different ringsImprovement of scopolamine-induced memory deficit in mice.[116]
DimerNeuroprotection.[100]
Dimethylthiazol + diphenyltetrazolium bromideAnticancer.[129]
EsterAntibacterial.[113,114]
EsterAntifungal.[131]
EsterAntithrombotic.[130]
EsterAnticancer.[107]
EsterXanthine oxidase inhibition.[121]
Ester and amideAnticancer.[118]
Ester and amideAntioxidant.[90]
Glycerol and diglycerolβ-amyloid aggregation inhibition.[102]
HeterocyclicAnticancer.[109]
Isopentyl Anticonvulsant.[135]
N-Hydroxy-N-PropargylamideFree radical scavenging, AChE inhibition, Cu(II) quelation.[92]
O-alkylaminesAntioxidant, butyrylcholinesterase inhibition.[115]
OH + OMe group + amide Neuraminidase inhibition.[94]
Phthalate and maleateHepatoprotection.[91]
PiperazineAntiviral.[128]
Tributyltin(IV)Anticancer. [132]
Table 2. Estimated pKa values and molar fractions, Mf(q), at pH = 7.4. The (q) in the acronym represents the charge of the acid-base species.
Table 2. Estimated pKa values and molar fractions, Mf(q), at pH = 7.4. The (q) in the acronym represents the charge of the acid-base species.
pKa1pKa2pKa3pKa4Mf(+1)Mf(0)Mf(−1)Mf(−2)Mf(−3)Mf(−4)
FA4.010.0-- -4 × 10−40.9970.003--
FA-83.14.45.810.8<10−40.0230.9764 × 10−4--
FA-123.75.610.311.3- <10−40.0150.9840.001<10−4
FA-262.53.910.111.8<10−43×10−40.9980.002<10−4-
FA-412.74.49.711.1-<10−40.0010.9940.005<10−4
FA-881.95.25.910.5-<10−42 × 10−40.0330.9660.001
FA-1063.75.710.311.2-<10−40.0190.9800.001<10−4
FA-1153.85.88.610.9-<10−40.0230.9180.059<10−4
FA-1183.68.59.513.0-1 × 10−40.9210.0780.001<10−4
FA-1384.07.69.9-2 × 10−40.5960.4030.001--
FA-1423.15.810.412.2-<10−40.0220.9770.001<10−4
FA-1733.64.210.0--<10−40.0010.9970.003-
FA-1753.89.611.0--3 × 10−40.9940.006<10−4-
Table 3. First ionization energy (IE, eV), electron affinities (EA, eV), and lowest bond dissociation energies (BDE, kcal/mol) for FA and the selected subset of derivatives.
Table 3. First ionization energy (IE, eV), electron affinities (EA, eV), and lowest bond dissociation energies (BDE, kcal/mol) for FA and the selected subset of derivatives.
IEEABDEBDE-Site *
q = 1
FA-13811.643.5489.30b (OH)
q = 0
FA8.36−0.2885.15b (OH)
FA-88.75−0.1380.24R5 (SH)
FA-268.350.3483.06R1 (OH)
FA-1188.31−0.9580.18b (OH)
FA-1388.580.2983.97b (OH)
FA-1758.12−0.2380.09R2 (OH)
q = −1
FA4.85−2.9582.48b (OH)
FA-83.75−2.9577.44R5 (SH)
FA-264.33−3.1277.29R1 (OH)
FA-1184.43−2.9878.79b (OH)
FA-1383.89−3.1174.88b (OH)
FA-1754.66−3.0578.54b (OH)
q = −2
FA−0.06−6.0096.96a (OCH3)
FA-80.20−5.5882.04b (OH)
FA-26−0.34−6.0471.30b (OH)
FA-118−1.09−4.9774.87b (OH)
FA-138−0.55−5.7797.02a (OCH3)
FA-175−0.31−5.3275.90b (OH)
q= −3
FA-118−3.78−7.5971.01b (OH)
* The labels correspond to those shown in Scheme 1.
Table 4. Polygenic score (SP) values for ferulic acid and its derivatives.
Table 4. Polygenic score (SP) values for ferulic acid and its derivatives.
CompoundDGBW (kcal/mol)SP
COMTMAO-BAChE
Ferulic Acid−5.28−7.19−7.373.78
FA-9−5.14−6.91−6.433.50
FA-26−5.09−7.63−7.883.93
FA-118−5.90−7.09−6.953.79
FA-138−5.12−7.17−7.403.76
FA-175−5.42−7.02−7.013.70
ΔGB, dopa = −5.44 kcal/mol in COMT; ΔGB, pea = −6.01 kcal/mol in MAO-B; ΔGB, Ach = 4.56 kcal/mol in AChE. For natural substrates SP = 3.00.
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

Guzmán-López, E.G.; Reina, M.; Hernández-Ayala, L.F.; Galano, A. Rational Design of Multifunctional Ferulic Acid Derivatives Aimed for Alzheimer’s and Parkinson’s Diseases. Antioxidants 2023, 12, 1256. https://doi.org/10.3390/antiox12061256

AMA Style

Guzmán-López EG, Reina M, Hernández-Ayala LF, Galano A. Rational Design of Multifunctional Ferulic Acid Derivatives Aimed for Alzheimer’s and Parkinson’s Diseases. Antioxidants. 2023; 12(6):1256. https://doi.org/10.3390/antiox12061256

Chicago/Turabian Style

Guzmán-López, Eduardo Gabriel, Miguel Reina, Luis Felipe Hernández-Ayala, and Annia Galano. 2023. "Rational Design of Multifunctional Ferulic Acid Derivatives Aimed for Alzheimer’s and Parkinson’s Diseases" Antioxidants 12, no. 6: 1256. https://doi.org/10.3390/antiox12061256

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop