Next Article in Journal
Overview of Cardiac Arrhythmias and Treatment Strategies
Previous Article in Journal
Study of Oncolytic Virus Preservation and Formulation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Bioinformatics Tools for the Analysis of Active Compounds Identified in Ranunculaceae Species

by
Cătălina Mareş
1,
Ana-Maria Udrea
2,3,
Nicoleta Anca Şuţan
4,* and
Speranţa Avram
1
1
Department of Anatomy, Animal Physiology and Biophysics, University of Bucharest, 91-95 Splaiul Independentei, 050095 Bucharest, Romania
2
Laser Department, National Institute for Laser, Plasma and Radiation Physics, Atomistilor 409, 077125 Magurele, Romania
3
Research Institute of the University of Bucharest-ICUB, University of Bucharest, 91-95 Splaiul Independentei, 050095 Bucharest, Romania
4
Department of Natural Sciences, University of Piteşti, 1 Targul din Vale Str., 110040 Pitesti, Romania
*
Author to whom correspondence should be addressed.
Pharmaceuticals 2023, 16(6), 842; https://doi.org/10.3390/ph16060842
Submission received: 15 May 2023 / Revised: 30 May 2023 / Accepted: 31 May 2023 / Published: 5 June 2023

Abstract

:
The chemical compounds from extracts of three Ranunculaceae species, Aconitum toxicum Rchb., Anemone nemorosa L. and Helleborus odorus Waldst. & Kit. ex Willd., respectively, were isolated using the HPLC purification technique and analyzed from a bioinformatics point of view. The classes of compounds identified based on the proportion in the rhizomes/leaves/flowers used for microwave-assisted extraction and ultrasound-assisted extraction were alkaloids and phenols. Here, the quantifying of pharmacokinetics, pharmacogenomics and pharmacodynamics helps us to identify the actual biologically active compounds. Our results showed that (i) pharmacokinetically, the compounds show good absorption at the intestinal level and high permeability at the level of the central nervous system for alkaloids; (ii) regarding pharmacogenomics, alkaloids can influence tumor sensitivity and the effectiveness of some treatments; (iii) and pharmacodynamically, the compounds of these Ranunculaceae species bind to carbonic anhydrase and aldose reductase. The results obtained showed a high affinity of the compounds in the binding solution at the level of carbonic anhydrases. Carbonic anhydrase inhibitors extracted from natural sources can represent the path to new drugs useful both in the treatment of glaucoma, but also of some renal, neurological and even neoplastic diseases. The identification of natural compounds with the role of inhibitors can have a role in different types of pathologies, both associated with studied and known receptors such as carbonic anhydrase and aldose reductase, as well as new pathologies not yet addressed.

1. Introduction

1.1. Overview of Ranunculaceae Active Compounds

Interest in Ranunculaceae species, the perennial plant that presents bioactive chemical compounds, has been raised due to the toxic properties that species possess [1]. Herbal medicine has a long history, and lately, it is becoming more popular globally [2,3]. Even if the source of these medicinal compounds is natural, this does not mean that it is safe and has fewer adverse effects [4]. For example, ingestion of aconitine can cause severe cardiotoxicity manifested by ventricular arrhythmia [5].
Aconitum, Helleborus and Anemone are genera of the Ranunculaceae family, comprising valuable species with pharmacological use, such as aconite roots, which have the highest concentration of aconitine (AC). Symptoms due to accidental ingestion of a large amount of the substance are neurological, gastrointestinal and cardiovascular symptoms [6].
The pharmacological effects of AC that have been identified are antitumor, anti-inflammatory, analgesic and local anesthetic effects [7]. These results make AC a potential therapeutic agent for diseases such as cancer, rheumatoid arthritis and chronic pain [8]. In Table 1, we summarize the main study about the pharmacological effects of AC.
The high toxicity of this compound has determined numerous cases of poisoning, whereby the oral administration of AC can result in mild intoxication to cardiac arrest and death depending on the dose [9]. AC toxicity has become an acute problem, and that is why we are trying to obtain the maximum therapeutic effects at the lowest possible doses of the compound [10].
The pharmacological mechanisms and pharmacokinetic properties of AC have not been fully understood, but the evidence of its utility in different pathologies could help in the management of different diseases [1]. The pharmacological effects of AC, but also of similar alkaloids such as hypaconitine and mesaconitine, are given by the chemical skeleton of the compounds. In Figure 1, the chemical structure of AC can be identified, with the skeleton being found in C19-diterpenoid alkaloids [11].
Countless studies on tissues of rats and mice show tissue degeneration and necrosis, especially the heart and brain, following the administration of the solution obtained from the decoction of Aconitum sp. The data were evaluated according to the dose applied and the incubation time [12,13,14,15].
Figure 1. Chemical structure in 2D format of AC (PubChem CID 245005) [16].
Figure 1. Chemical structure in 2D format of AC (PubChem CID 245005) [16].
Pharmaceuticals 16 00842 g001
Natural compounds can be useful in the treatment of central nervous system diseases such as depression, anxiety and inflammatory diseases. Quercetin is a flavonoid with antidepressant activity observed in mice [17]. The anxiolytic and antidepressant effect [17] of quercetin regulates cholinergic and serotoninergic functions [18].
Studies have shown the importance of polyphenols in minimizing neurodegenerative effects [19]. The study by Zhao et al. showed the importance of the therapy of the combination of quercetin and malvidin for the protection of neurons against cognitive dysfunction due to sleep deprivation. Sleep deprivation of mice showed a decrease in memory consolidation capacity, but the mixture of polyphenols improved the negative effects on the studied organisms [20].
Polyphenols are natural compounds with biological effects aimed at their antioxidant effect, improving oxidative stress by activating antioxidant enzymes and inhibiting reactive oxygen species [21]. The relevance of polyphenolic compounds catechin, epicatechin [22], caffeic acid, coumaric acid, chlorogenic acid [23], delphinidin is given by the applicability of these natural compounds in complex diseases such as cardiovascular diseases, neurodegenerative diseases, obesity and cancer [24].
Genistein and gallic acid are part of the family of polyphenols with the role of inhibiting the binding of the nuclear factor of activated B cells (NF-kB) by the target DNA. Polyphenols have been reported in the literature as strong inhibitors of NF-kB, thus being useful in the treatment of various types of dementia, including dementia of the Alzheimer type or Parkinson’s disease [25,26]. Natural compounds such as genistein and gallic acid inhibit the transmission of pro-inflammatory cytokines, neuroinflammation and oxidative stress, which are considered causes of neuronal loss [27].
Syringic acid has a neuroprotective effect, due to the antioxidant, anti-inflammatory and antidepressant properties that polyphenols have [28].
Naringenin and naringin are citrus flavonoids with antioxidant, anti-proliferative and anti-inflammatory activities, thus representing a potential therapeutic effect in preventing and alleviating the symptoms of neurological disorders [29].
Rutin is a powerful antioxidant involved in Alzheimer’s disease and other neurodegenerative pathologies. The pharmacological applications of this compound are mainly due to the antioxidant and anti-inflammatory activity that this compound possesses [17]. The action mechanisms through which rutin can act are conducted via eliminating and inhibiting the production of reactive oxygen species [30].
Magnoflorine is fat soluble and can cross the blood–brain barrier, thus manifesting the antidepressant effect. In vivo studies on model animals have shown that magnoflorine can increase the expression of antioxidant proteins. Magnoflorine is an alkaloid with the function of protecting the cardiovascular system and regulating the immune system, and it has very good antioxidant properties [31]. The medical properties of magnoflorine suggest its antidepressant effect [32].
Hyperoside can improve memory and learning in mice by potentiating synapses and ameliorating memory disorders induced by neurodegenerative diseases [33]. It was also observed, following long-term treatment with hyperoside in mice, the reduction in beta-amyloid plaque and tau protein phosphorylation, the attenuation of neuroinflammation and stress oxidative, thus being a potential drug for Alzheimer’s disease [34].
The chemical structures of the compounds that showed the most important antioxidant activity can be visualized in Figure 2.
The computational approach in biology and medicine plays an essential role both in the actual diagnosis of diseases and in the development of medicines based on computational methods. The steric and electronic molecular descriptors, the number of atoms and the type of chemical bonds represent essential information in studying the biological activity of compounds based on their chemical structure [36].
The purpose of the current study is to characterize from a bioinformatic point of view the compounds identified from an extract of Aconitum sp. and to try to identify those compounds that have a truly biologically active role.
Genistein has proven to be useful both in the case of diabetes and the complications caused by this disease [37] and also in other disorders such as cataracts, cystic fibrosis, and non-alcoholic fatty liver disease. This compound shows high binding affinity to aldose reductase and inhibits its activity [38].
Table 1. Pharmacological effects and mechanism of action of different organisms/cell type determinate by AC.
Table 1. Pharmacological effects and mechanism of action of different organisms/cell type determinate by AC.
Pharmacological EffectsCell LinesMechanism of ActionReference
Anti-inflammatory activityMouse leukemic monocyte/macrophage cell line RAW264.7Suppressing NF-kB and NFATc1 activation and DC-STAMP expression[39]
Anti-rheumatic activitiesRheumatoid arthritis HFLS-RA fibroblast-like synoviocytes[40]
Analgesia activityMale wildtype FVB mice and Male Mdr1a−/− FVB miceMdr1a deficiency[41]
the rat chronic constriction injury of an infraorbital nerve model.N-methyl-D-aspartate receptor[42]
Mice pain models caused by hot plate, acetic acid, formalin, and CFA-[14]
Anti-cancer activityPancreatic cancer cell lines Miacapa-2 and PANC-1Suppressing cancer cell growth and increasing cell apoptosis[43]
Human cervical carcinoma HeLa cellsUpregulating mRNA expression levels of eIF2α, ATF4, IRE1, XBP1, ATF6, PERK[44]
Human breast cancer cell line MDA-MB-231BOInhibiting cancer cell invasion by an alteration of the TGF-β/Smad signaling pathway and down-regulating of NF-κB and RANK expressions.[45]
Human OVCA A2780 cell lineAdjusting ERβ-mediated apoptosis, DNA damage and migration[8]
For a better prediction of the mechanisms, we selected the compounds that showed satisfactory results and performed molecular docking. Molecular docking is a quick method for predicting the interaction of a specific target and a ligand [46,47]. In this study, we will predict the lowest estimated free energy of binding (LEFEB) between quercetin, caffeic acid, chlorogenic acid, coumaric acid, hyperoside, and rutin when they interact with aldose reductase (AKR1B1).
In the current study, the bioinformatics testing of the compounds obtained from the HPLC extraction is conducted with the aim of identifying the antioxidant and anti-inflammatory effect of the compounds. The study wants to identify the most useful compounds for the treatment of anxiety, depression, and neurodegenerative diseases.

1.2. The Choice of Natural Compounds

Our study followed the identification of the compounds obtained from microwave and ultrasound extraction of rhizomes, leaves, and/or flowers of Aconitum toxicum Rchb., Anemone nemorosa L., and Helleborus odorus Waldst. & Kit. ex Willd. and the bioinformatics testing of each of them to highlight their role in the extracts [48].
In the present study, we performed an in silico analysis of the compounds identified by HPLC from hydroalcoholic extracts of A. toxicum, A. nemorosa, and H. odorus. The twenty-one compounds identified in the extracts were aconitine, hypaconitine, mesaconitine, magnoflorine, gallic acid, catechin, caffeic acid, ferulic acid, chlorogenic acid, epicatechin, delphinidin, coumaric acid, daidzein, hyperoside, rutin, naringin, malvidin, quercetin, naringenin, genistein, and syringic acid.
Starting from the literature on natural compounds with effects on the central nervous system, we focus our attention on compounds suitable for the analysis of pathologies in the psychiatric spectrum. Thus, the predicted molecular targets of the studied compounds are generally statistically significant for nervous system pathologies [49].
The study by Avram et al. provides essential information about the molecular descriptors and QSAR equations useful in inducing the antidepressant activity of the different compounds [50].

1.3. Molecular Targets Obtain Using Prediction Target Database

The nervous system presents a particular complexity, but with the help of bioinformatics techniques, it is possible to predict the activity of certain compounds for the treatment of anxiety, depression, and even some neurodegenerative diseases [50,51]. Pharmaceutical descriptors (electrostatic, hydrophobic, hydrogen donor/acceptor bond, etc.) can predict the biological activity of the compounds and their effect in the body [50].
The main molecular targets obtained with the help of prediction databases are from the family of aldose reductase and carbonic anhydrase enzymes.
Carbonic anhydrases (CA) are enzymes (16 isoforms) with important roles in the optimal functioning of the body. Their pharmacological inhibition has been reported as a treatment for many diseases (glaucoma, neuropathic pain, tumors) [52]. CA catalyzes the transition from carbon dioxide to bicarbonate and protons [53]. CA inhibition has good prospects for understanding the protein–drug interaction mechanism and designing useful pharmacological agents [54]. CA are metalloenzymes that become active when the pH is basic, the active locus of the enzyme being the hydrophobic pocket in which the zinc ion is found, which is crucial for the catalytic reaction, thus regulating the concentration of CO2 in the test. In mammals, CA functions are represented by respiration, pH regulation, electrolyte secretion, and metabolic processes dependent on HCO3 [55,56].
CA II is an isoform that is identified in the brain and is expressed in neurons, oligodendrocytes, and the choroid plexus [57], and isoform VII is found in the hippocampus and cortex [58]. The scientific literature indicates the involvement of these enzymes in neuropathological processes [56]. CA inhibitors can function as anticonvulsants in animal models of epilepsy and in diagnosed patients. Current studies approach new structures that can inhibit different isoforms of anhydrases to minimize the side effects on patients of existing inhibitors [59].
Aldose reductase is a cytosolic enzyme that belongs to the aldose-keto-reductase family and is found in most mammalian cells. Although the roles of this enzyme in the polyol pathway, the one that transforms glucose into sorbitol, are well known, recent studies suggest the involvement of this enzyme in detoxification processes under conditions of oxidative stress. This enzyme is mainly involved in the complications of diabetes such as retinopathy, cataracts, nephropathy, and neuropathy [60]. Among the useful compounds that inhibit aldose reductase are polyphenolic compounds such as curcumin, quercetin, and kaempferol. These natural compounds possess a strong inhibitory effect on enzymes both in in vitro and in vivo studies [61].
Neurodegeneration caused by Alzheimer’s disease is achieved by the precipitation of beta-amyloid in the brain, which produces inflammation [62]. A study by Huang et al. showed in a cell culture of neurons and microglia that inhibition of aldose reductase with sorbinil as a therapeutic agent prevented cell migration and phagocytosis. Thus, the neuronal death in the culture was mitigated, and it was proven that the inhibition of aldose reductase is effective in the neuronal degeneration induced by beta-amyloid [63].
These results are useful for us in the search for new aldose reductase inhibitors that may also have a useful role in the case of Alzheimer’s disease.

2. Results

2.1. Drug-Likeness, Pharmacokinetics, and Pharmacogenomics Profiles of Compounds

To evaluate the character of a possible drug, the compounds were processed in the Expasy database [64] and tested to comply with the rules of medical chemistry—the Lipinski [65], Veber [66], Egan [67], and Muegge [68] rules. Violations of drug design rules for the analyzed compounds are in Figure 3.
For the mentioned compounds, the physico-chemical properties are calculated, such as molecular weight, hydrophobicity, the share of hydrogen bond donor/acceptor atoms, the share of the number of rotatable bonds, polar molecular surface, and so on. Chemical compounds for which excess is identified are considered not to meet the drug-like condition [50,68].
Through the bioinformatics calculation of the drug-like character, we found that the following compounds respect the drug-like profile: magnoflorine, gallic acid, catechin, ferulic acid, caffeic acid, chlorogenic acid, epicatechin, delphinidin, coumaric acid, daidzein, malvidin, quercetin, naringenin, genistein, and syringic acid.
Bioavailability is a very important indicator in the absorption of drugs. A high bioavailability means that the compound will reach the systemic circulation more easily when administered orally. A higher bioavailability means that more nutrients are absorbed by the conventional method. Ferulic acid and coumaric acid show the best availability of the series of compounds obtained from the aconitine mixture. The bioavailability score depends on the Lipinski rule [65], and repeated violations of the parameters that form this rule cause the bioavailability score to decrease. At the same time, the bioavailability of 0.55, as shown by most compounds in the series, means that at physiological pH, 55% of the compound is expected to reach the circulation in unchanged or active form [69]. The results of the bioavailability score can be identified in Figure 4.
Quercitin, genistein, and naringin represent only part of the natural compounds with an excellent role in enhancing the bioavailability of medicines [70].
AC, hypaconitine, mesaconitine, chlorogenic acid, hyperoside, rutin, and naringin are studied compounds that present a low bioavailability score, and the violation of all studied drug design rules can also be observed. This can be determined by the molecular mass and volume of these compounds.

2.2. Identification of Physico-Chemical Properties for the Pharmacological Profile

The pharmacological profile of the compounds can be identified in Table 2. The results indicate that flexibility is high, especially in the series of compounds from the alkaloid class—aconitine (11), hypaconitine (10), mesaconitine (10), gallic acid (6), rutin (6), and naringin (6) all show a high flexibility; on the other hand, for compounds such as catechin, epicatechin, delphinidin, quercetin, naringenin, and genistein, the flexibility is low.
The compounds have an average hydrophobicity, this parameter varying between −1.96 (malvidin) and 4.31 (hypaconitine), and the compounds have an average hydrophilic character.
The TPSA in drug design rules can be associated with bioavailability, the optimal value of this parameter being between 20 and 130 Å 2 [68].

2.3. Pharmacokinetic, Pharmacogenomic Profile, and Toxicity of Natural Compounds (ADME-Tox)

The predictive results for the ADME-Tox profile presented in Table 3 show that compounds with good absorption at the intestinal level (exception—magnoflorine). Oral bioavailability is increased for ferric acid, caffeic acid, delphinidin, and syringic acid. Permeability at the level of the blood–brain barrier (BBB) is increased for AC, hypaconitine, mesaconitine, magnoflorine, gallic acid, ferulic acid, malvidin, and syringic acid, and the studied compounds are not inhibitors for OCT2/OCT1 receptors.
The predictive results in the case of the pharmacogenomic profile (Table 4) show that the compounds do not act as CYP3A4 cytochrome inhibitors; instead, they are substrates for this cytochrome (aconitine, hypaconitine, mesaconitine, magnoflorine, chlorogenic acid, hyperoside, rutin, naringin, and quercitrin). For cytochrome CYP2C9, most of the compounds act as inhibitors but not as substrates, and a reduced activity as a substrate/inhibitor is recorded at the site of cytochrome CYP2D6.
The results presented in Table 5 show that most of the compounds do not induce mutagenicity (Ames negative), do not develop toxic character for birds and bees, and are toxic for fish.
At the level of the human body, these compounds develop hepatotoxicity (exceptions: hypaconitine, catechin, ferulic acid, caffeic acid, chlorogenic acid, epicatechin, coumaric acid, and syringic acid), and they are toxic at the mitochondrial level (exceptions: gallic acid, ferulic acid, caffeic acid, coumaric acid, syringic acid, rutin, and naringin).
The compounds do not show nephrotoxicity (exceptions: daidzein coumaric acid delphinidin) and cardiac toxicity. However, most show toxicity at the mitochondrial level (exceptions: gallic acid, caffeic acid, ferulic acid, and coumaric acid).

2.4. Pharmacodynamics Profiles of Studied Compounds

The data obtained after running SwissTargetPrediction are presented in Table 6. From a pharmacokinetic point of view, there is a high probability that the compounds thus analyzed bind to different CA isoforms or aldose reductase.

2.5. Molecular Docking Results

A compound may bind to a specific target if the estimated free energy of binding is lower than −6 Kcal/mol [71,72]. The molecular docking results indicate that quercetin (−7.85 kcal/mol) has the LEFEB when it interacts with AKR1B1 (Table 7). Caffeic acid has an LEFEB of −6.21 kcal/mol, indicating a potential binding interaction with AKR1B1 (Table 7).
According to our molecular docking predictions, chlorogenic acid, coumaric acid, hyperoside, and rutin do not bind to AKR1B1 (Table 7).

3. Materials and Methods

3.1. Molecular Modelling of Chemical Compounds

In the first stage, the compounds identified in the bioinformatics databases to select their most suitable formats for the bioinformatics study.
The study analyses twenty-one natural compounds obtained by HPLC from a mixture of aconitine. The separate effect of these chemical structures are analyzed bioinformatically to identify the main medicinal compound. The effect we are looking for in the analyzed compounds is anti-inflammatory, antineoplastic and, more than that, the involvement of the compounds in the nervous system.
Studies in the literature describe the usefulness of polyphenols and some alkaloids in the central nervous system, having a beneficial effect in a series of diseases from anxiety, depression, Alzheimer’s disease, and Parkinson’s disease. Studies suggest the neuroprotective effect of the compounds by reducing inflammation in the brain and by neutralizing oxidases [18,28,29,30,34,73,74].
The SMILES file and the molecular weight of the compounds obtained from the PubChem database can be found in Table 8.

3.2. Prediction of Compounds Drug- and Lead-Likeness Features

The rules of drug design as well as the bioavailability score represent a useful tool in predicting the character of a compound to be a medicine. Natural compounds should meet a series of chemical structure design rules to be able to act optimally in the body. The rule of Lipinski [65], Veber [75], Ghose [76], and Egan [67] can be predicted with online tools such as SwissADME [68]. Although the drug design rules are very well calibrated and quite permissive, there are several drugs under study that violate these rules [77]. The Lipinski rule provides for a molecular mass lower than 500 Daltons, a maximum of 10 donor hydrogen bonds, a maximum of 5 acceptor hydrogen bonds, and a log octanol/water (Log P(o/w)) lower than 5 [50,65].
Ghose’s rule shows a molecular mass value between 160 and 480 Daltons, but the Log P(o/w) value must be between 0.4 and 5.6. In addition to Lipinski’s rule, Ghose’s rule also considers the refractivity, which must be between 40 and 130, and the total number of atoms, which must be between 40 and 130, as a molecular descriptor [68,78].
The data used were processed by online bioinformatics software that presents several predictive parameters. These calculation methods use algorithms based on vast training sets and cross-validation accuracy. For example, with regard to SwissADME, the BBB (blood–brain barrier) model was built starting from a training set of 260 permeable or impermeable molecules, obtaining an accuracy of 88% [68].

3.3. Identification of Important Physico-Chemical Properties for the Pharmacological Profile

The chemical structures, in SMILES format, were loaded into the MOE software and converted to mol2 files [79]. These files were used for the calculation of physicochemical properties such as flexibility expressed in the number of rotatable bonds, refractivity (Å2), polar molecular surface area (Å2), hydrophobicity, and water solubility (Table 3).
The chemical structure of the compounds influences their pharmacological activity [80]. The three-dimensional arrangement of the atoms in the molecule can play an important role in the activity of the compounds both in the accessibility to the active sites and for the interaction of the drug with the targeted receptor [18]. The polar molecular surface shows the effective space occupied by the atoms of the molecule [81]. This property is useful in predicting the accessibility of the compound [68].
Flexibility is a characteristic often expressed by the number of their rotatable bonds, being the most common descriptor predicted with the help of bioinformatics tools. From the point of view of flexibility, the studied molecule should not have more than nine rotatable bonds [82].
Solubility (lipophilicity) is the most important physical property of the drug, having a special role in absorption, distribution, and elimination [83].
The partition coefficient (LogP) represents the hydrophobic character and is calculated bioinformatically based on the different solubility of the compound between the polar and the non-polar phase. Hydrophobicity is used to estimate the distribution of drugs in the body [84].
The molecular refraction represents the real volume of the molecules, but also the forces acting on the interaction between the drug and the receptor. This molecular descriptor is calculated by the sum of the atomic refraction of each atom in the molecule [85].

3.4. Evaluation of the Pharmacokinetic and Pharmacogenomic Profile of Natural Compounds

pkCSM is a pharmacokinetic properties prediction platform composed of 28 regression and classification models. The models used for prediction with this web server use a Pearson correlation with a coefficient between 0.6 and 0.9 and cross-validation of the data set [86].
An important part of this study consisted of evaluating the pharmacokinetic profile, pharmacogenomics, and toxicity. Absorption, Distribution, Metabolism, Excretion and Toxicity (ADME-T) properties were expressed as intestinal absorption (human), oral bioavailability (human), permeability of the compounds at the level of the blood–brain barrier, percentage of binding to plasma proteins, and inhibition of the OCT receptor at the renal level [87].
Intestinal absorption (human) predicts the proportion of compounds that can be absorbed in the human small intestine. Bioavailability represents the ratio between the amount of active substance administered and the amount absorbed that manifests its biological effect [50]. Permeability at the level of the blood–brain barrier (BBB) represents an important parameter for minimizing the side effects and toxicity of a compound [88]. Computational algorithms predict the ability of a compound to penetrate the BBB [50]. The binding percentage to the plasma proteins is a good indicator of the drugs; they bind to the circulating proteins in the blood and thus cross the cell membranes more easily [18]. The substrate OCT1 (Organic cation transporter (1) is a hepatic transporter with a role in the hepatic clearance of drugs, and OCT2 (Organic cation transporter (2) is a renal transporter with a role in the elimination of drugs and endogenous compounds. Both OCT1 and OCT2 have a special role in the body’s clearance [18].
Cytochrome P450 is a detoxification enzyme in the body, which is found in the liver and helps to excrete xenobiotics from the body, but it can also deactivate many drugs. Inhibitors of these enzymes affect the metabolism of drugs and therefore the ability of compounds to inhibit cytochrome P450 is evaluated. Cytochrome P450 substrate or activator predicts the likelihood that a molecule will be metabolized by P450. CYP1A2, CYP2C19, CYP2C9, CYP2D6, and CYP3A4 are cytochrome P450 isoforms and can be either inhibitors or substrates for different chemical compounds [18,86].

3.5. Toxicity Profile of Natural Compounds

A significant number of items representing toxicity were analyzed for natural compounds—Ames (mutagenesis), carcinogenicity, toxicity to different species (crustacea, bees, fish), nephrotoxicity, hepatotoxicity, cardiotoxicity, mitochondrial toxicity, toxicity to nuclear receptors, and so on [80].
Ames toxicity for a compound follows its mutagenic character; if it turns out to be positive, then in this sense, the compound is mutagenic [18,89].
Carcinogenicity is predicted to identify the possible carcinogenic character of a compound [80].
Aquatic toxicity for fish and crustaceans is also a common indicator for toxicity, with positive compounds for this category inhibiting the growth of half of the studied fish/crustacean population. This type of toxicity refers to the toxicity of the environment, with some drugs not being completely metabolized, being eliminated in their active state or their active metabolites and ending up in domestic water through excretion and in the soil, thus affecting the balance of the environment [80,86].

3.6. Pharmacodynamics Profile of Compounds

The pharmacological effects of AC and its derivatives are presented in Table 1. AC binds to certain receptors in the body to determine the beneficial effect for the body [90,91]. To model compounds, it is necessary to know the molecular target to which these compounds can bind. That is why we will use molecular target prediction software for the studied compounds.
SwissTargetPrediction is an online bioinformatics tool hosted by the Swiss Institute of Bioinformatics SIB that predicts their protein targets based on the similarity of small molecules, based on a library of 370,000 active substances on known targets [92].

3.7. Molecular Modeling

The structure of quercetin, caffeic acid, chlorogenic acid, coumaric acid, hyperoside, and rutin was imported from PubChem and optimized using our usual protocol [93]. The molecules were minimized using Hamiltonian Forcefield MMFF94x at a 0.05 gradient with Gasteiger (PEOE) partial charges [94]. The structures were exported as “.mol2” and converted as pdbqt for the molecular docking analysis using Open Babel software [95].
The structure of AKR1B1 was imported from the RCSB Protein Data Bank (PDB ID: 2ACQ) [96] and optimized for the molecular docking studies according to our usual protocol, we deleted the water molecules and the residues, added the hydrogen atoms, merged the non-polar hydrogen atoms, and applied the Kollman charges [47,72].

3.8. Molecular Docking

Using AutoDock 4.2.6 and following our standard procedure (we used the Lamarckian Genetic Algorithm search parameter, and we generated 100 runs for each protein-ligand complex), we predicted the interaction between natural compounds and selected target proteins [71,72,97]. We ran blind docking simulations with the following grid-box parameters:
The number of specified grid points is 126 (x, y, z); the grid point spacing is 0.375 angstroms; and the coordinates of the central grid point of maps are 15.064, 32.982, and 68.073).

4. Discussion

Active compounds from Aconitum toxicum Rchb., Anemone nemorosa L., and Helleborus odorus Waldst. & Kit. ex Willd were analyzed from a bioinformatics point of view, both pharmacokinetically and pharmacodynamically.
Alkaloids with more complex chemical structures (AC, hypaconitine, mesaconitine, hyperoside, rutin, and naringin) show violations of drug design rules (Lipinski, Ghose, Egan, and Muegge rules), while the compounds magnoflorine, catechin, epicatechin, daidzein, malvidin, quercetin, naringenin, and genistein do not violate any of the drug design rules. The bioavailability score varies between 0.11 (chlorogenic acid) and 0.85 (ferulic acid and coumaric acid).
The pharmacological profile of a compound shows high values of flexibility and refractivity, and TPSA can be observed for alkaloids AC, hypaconitine, and mesaconitine, but also for hyperoside, naringin, and rutin.
All the analyzed compounds can be absorbed at the intestinal level, while only ferulic acid, delphinid, caffeic acid, daidzein, and syringic acid are bioavailable.
AC, hypaconitine, mesaconitine, magnoflorine, gallic acid and ferulic acid, malvidin, and syringic acid can cross the BBB. All the compounds mentioned above have a higher binding affinity at the level of plasma proteins, the rest of the compounds being predicted to circulate freely in the blood in higher proportions. Additionally, except for magnoflorine (OCT1 inhibitor), no compound is an inhibitor of OCT1 and OCT2.
From the point of view of mutagenesis and toxicity to aquatic crustaceans and birds, predominantly, the compounds are predicted to be non-toxic. Instead, the predictions show that all the compounds analyzed are toxic to fish, and most of them show hepatotoxicity and toxicity at the level of mitochondria. These data correlate with the data from the literature, from the study by Wang et al., who stated that aconitine affects signaling pathways and mitochondria, causing apoptosis. Mitochondrial damage can have multiple negative effects based on the length of time exposure to the toxic compound. By treating H9c2 hippocampal cells with aconitine, mitochondrial dysfunction was induced, cytochrome Bax was upregulated, Caspaseno3 was cleaved, and Bcl 2 was decreased, thus providing a possible mechanism of apoptosis mediated by mitochondria [98]. In the case of the study by Ravindran et al., apoptosis depended both on the compound concentration to which the cells were exposed and on their incubation time with the actual compound [99].
For hepatotoxicity, there is not much evidence in humans, but studies on animal models such as rats show a high tendency to accumulate alkaloids in organs such as the liver and kidneys, but this is achieved after long-term oral administration. The study by Xiaoyu Ji et al. does not specify the identified period that can determine the accumulation of these alkaloids in animal tissues. At the same time, the optimal administration time and the recommended dose for rats are not known, so as not to determine liver damage induced by the consumption of the medicinal compound [12].
Making the prediction of molecular targets, we observe a prevalence of carbonic anhydrase to which the compounds bind with high affinity [100]. In these predictions, several isoforms of the enzyme with the role of regulating the acid–base balance appear, but CA II and CA VII are the most frequently encountered.
Supuran’s review showed a high affinity of coumarin for CA, inhibiting their activity. These natural compounds have a binding affinity to CA in areas inaccessible for other compounds [52].
Naringin is a natural phenol, and although it was predicted to have a binding affinity to cytochrome P450, studies in the literature prove the important inhibitory activity of naringin not only on the CA II isoform, but also on other important receptors in the central nervous system such as αnoglucosidase (αnoGly), acetylcholinesterase (AChE), and butyrylcholinesterase (BChE) [101]. Caffeic acid and other natural phenols have been identified as CA II inhibitors [102,103].
Quercetin has the lowest predicted energy of binding when compared to the other evaluated natural compounds. According to our computer simulations, quercetin and AKR1B1 interact favorably and form the conventional hydrogen-bound, carbon-hydrogen-bound, pi-cation, pi-pi stacked, and pi-alkyl interactions (Figure 5).

5. Conclusions

The studies in the literature confirm the predictions made with the bioinformatics tools for the compounds obtained from extracts of three Ranunculaceae species, Aconitum toxicum Rchb., Anemone nemorosa L., and Helleborus odorus Waldst. & Kit. ex Willd. by the HPLC method. The analyzed compounds that inhibit CA II can be used for testing new treatment modalities in cases of glaucoma, neoplasms, and neurodegenerative diseases, diseases in which a great improvement was observed following the inhibition of the CA enzyme [105].
The results obtained following the prediction of molecular targets are not significant for AC, gallic acid, delphinidin, malvidin, and mesaconitine.
Magnoflorine and rutin have targets on metabotropic receptors, and genistein, which is a non-specific compound, has a probability of binding both on nuclear and metabotropic receptors.
The results obtained from this study show the prevalence of natural compounds obtained from Aconitum toxicum Rchb., Anemone nemorosa L., and Helleborus odorus Waldst. & Kit. ex Willd of binding at the level of the carbonic anhydrase family and the aldo-keto reductase family. Quercitin presented the best binding energy and the best values of the molecular descriptors. At the same time, additional studies are needed on AC, hypaconitine, magnoflorine, gallic acid, ferulic acid, malvidin and syringic acid due to the permeability of these compounds to cross the blood–brain barrier.

Author Contributions

Conceptualization, S.A. and N.A.Ş.; methodology, S.A.; software, C.M. and S.A.; validation, S.A., N.A.Ş. and A.-M.U.; formal analysis, S.A., N.A.Ş., A.-M.U. and C.M.; investigation, S.A., N.A.Ş., A.-M.U. and C.M.; resources, S.A.; data curation, S.A. and N.A.Ş.; writing—original draft, C.M. and A.-M.U.; writing—review and editing, S.A.; visualization, C.M., A.-M.U. and S.A.; supervision, S.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Romanian Ministry of Research, Innovation, and Digitization project numbers: PN-III-P4-ID-PCE-2020-0620, PN-III-P1-1.1-PD-2021-0225, Nucleu Program LAPLAS VII—contract no. 30N/2023, and PN-III-P2-2.1-PED-2021-2866.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Li, S.; Yu, L.; Shi, Q.; Liu, Y.; Zhang, Y.; Wang, S.; Lai, X. An Insight into Current Advances on Pharmacology, Pharmacokinetics, Toxicity and Detoxification of Aconitine. Biomed. Pharmacother. 2022, 151, 113115. [Google Scholar] [CrossRef] [PubMed]
  2. Xie, Y.; Mai, C.-T.; Zheng, D.-C.; He, Y.-F.; Feng, S.-L.; Li, Y.-Z.; Liu, C.-X.; Zhou, H.; Liu, L. Wutou Decoction Ameliorates Experimental Rheumatoid Arthritis via Regulating NF-KB and Nrf2: Integrating Efficacy-Oriented Compatibility of Traditional Chinese Medicine. Phytomedicine 2021, 85, 153522. [Google Scholar] [CrossRef]
  3. Wang, J.-J.; Lou, H.-Y.; Liu, Y.; Han, H.-P.; Ma, F.-W.; Pan, W.-D.; Chen, Z. Profiling Alkaloids in Aconitum Pendulum N. Busch Collected from Different Elevations of Qinghai Province Using Widely Targeted Metabolomics. Phytochemistry 2022, 195, 113047. [Google Scholar] [CrossRef]
  4. Mares, C.; Udrea, A.-M.; Buiu, C.; Staicu, A.; Avram, S. Therapeutic Potentials of Aconite-like Alkaloids - Bioinformatics and Experimental Approaches. Mini Rev. Med. Chem. 2023. [Google Scholar] [CrossRef]
  5. Liao, Y.-P.; Shen, L.-H.; Cai, L.-H.; Chen, J.; Shao, H.-Q. Acute Myocardial Necrosis Caused by Aconitine Poisoning: A Case Report. World J. Clin. Cases 2022, 10, 12416–12421. [Google Scholar] [CrossRef]
  6. Chan, T.Y.K. Aconite Poisoning. Clin. Toxicol. Phila. Pa 2009, 47, 279–285. [Google Scholar] [CrossRef]
  7. Mi, L.; Li, Y.-C.; Sun, M.-R.; Zhang, P.-L.; Li, Y.; Yang, H. A Systematic Review of Pharmacological Activities, Toxicological Mechanisms and Pharmacokinetic Studies on Aconitum Alkaloids. Chin. J. Nat. Med. 2021, 19, 505–520. [Google Scholar] [CrossRef] [PubMed]
  8. Wang, X.; Lin, Y.; Zheng, Y. Antitumor Effects of Aconitine in A2780 Cells via Estrogen Receptor Β-mediated Apoptosis, DNA Damage and Migration. Mol. Med. Rep. 2020, 22, 2318–2328. [Google Scholar] [CrossRef] [PubMed]
  9. Singhuber, J.; Zhu, M.; Prinz, S.; Kopp, B. Aconitum in Traditional Chinese Medicine—A Valuable Drug or an Unpredictable Risk? J. Ethnopharmacol. 2009, 126, 18–30. [Google Scholar] [CrossRef]
  10. Gao, X.; Hu, J.; Zhang, X.; Zuo, Y.; Wang, Y.; Zhu, S. Research Progress of Aconitine Toxicity and Forensic Analysis of Aconitine Poisoning. Forensic Sci. Res. 2018, 5, 25–31. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Hu, J.; Wu, Q.; Li, Q.; Lv, T.; Peng, T.-F.; Yin, S.; Jin, H.-Z. Antinociceptive Diterpenoid Alkaloids from the Roots of Aconitum Austroyunnanense. J. Asian Nat. Prod. Res. 2023, 25, 132–138. [Google Scholar] [CrossRef]
  12. Ji, X.; Yang, M.; Or, K.H.; Yim, W.S.; Zuo, Z. Tissue Accumulations of Toxic Aconitum Alkaloids after Short-Term and Long-Term Oral Administrations of Clinically Used Radix Aconiti Lateralis Preparations in Rats. Toxins 2019, 11, 353. [Google Scholar] [CrossRef] [Green Version]
  13. Lu, H.; Mei, L.; Guo, Z.; Wu, K.; Zhang, Y.; Tang, S.; Zhu, Y.; Zhao, B. Hematological and Histopathological Effects of Subacute Aconitine Poisoning in Mouse. Front. Vet. Sci. 2022, 9, 874660. [Google Scholar] [CrossRef] [PubMed]
  14. Deng, J.; Han, J.; Chen, J.; Zhang, Y.; Huang, Q.; Wang, Y.; Qi, X.; Liu, Z.; Leung, E.L.-H.; Wang, D.; et al. Comparison of Analgesic Activities of Aconitine in Different Mice Pain Models. PLoS ONE 2021, 16, e0249276. [Google Scholar] [CrossRef]
  15. Zhang, Y.; Zong, X.; Wu, J.-L.; Liu, Y.; Liu, Z.; Zhou, H.; Liu, L.; Li, N. Pharmacokinetics and Tissue Distribution of Eighteen Major Alkaloids of Aconitum Carmichaelii in Rats by UHPLC-QQQ-MS. J. Pharm. Biomed. Anal. 2020, 185, 113226. [Google Scholar] [CrossRef] [PubMed]
  16. PubChem Aconitine. Available online: https://pubchem.ncbi.nlm.nih.gov/compound/245005 (accessed on 29 May 2023).
  17. Seiman, D.D.; Batalu, A.; Seiman, C.D.; Ciopec, M.; Udrea, A.M.; Motoc, M.; Negrea, A.; Avram, S. Pharmacological Effects of Natural Compounds Extracted from Urtica Dioica Evaluated by in Silico and Experimental Methods. Rev Chim Buchar. 2018, 69, 2377–2381. [Google Scholar] [CrossRef]
  18. Avram, S.; Stan, M.S.; Udrea, A.M.; Buiu, C.; Boboc, A.A.; Mernea, M. 3D-ALMOND-QSAR Models to Predict the Antidepressant Effect of Some Natural Compounds. Pharmaceutics 2021, 13, 1449. [Google Scholar] [CrossRef] [PubMed]
  19. Iban-Arias, R.; Sebastian-Valverde, M.; Wu, H.; Lyu, W.; Wu, Q.; Simon, J.; Pasinetti, G.M. Role of Polyphenol-Derived Phenolic Acid in Mitigation of Inflammasome-Mediated Anxiety and Depression. Biomedicines 2022, 10, 1264. [Google Scholar] [CrossRef]
  20. Zhao, W.; Wang, J.; Bi, W.; Ferruzzi, M.; Yemul, S.; Freire, D.; Mazzola, P.; Ho, L.; Dubner, L.; Pasinetti, G.M. Novel Application of Brain-Targeting Polyphenol Compounds in Sleep Deprivation-Induced Cognitive Dysfunction. Neurochem. Int. 2015, 89, 191–197. [Google Scholar] [CrossRef] [Green Version]
  21. de la Rosa, L.A.; Alvarez-Parrilla, E.; Shahidi, F. Phenolic Compounds and Antioxidant Activity of Kernels and Shells of Mexican Pecan (Carya Illinoinensis). J. Agric. Food Chem. 2011, 59, 152–162. [Google Scholar] [CrossRef]
  22. Addepalli, V.; Suryavanshi, S.V. Catechin Attenuates Diabetic Autonomic Neuropathy in Streptozotocin Induced Diabetic Rats. Biomed. Pharmacother. 2018, 108, 1517–1523. [Google Scholar] [CrossRef] [PubMed]
  23. Agunloye, O.M.; Oboh, G.; Ademiluyi, A.O.; Ademosun, A.O.; Akindahunsi, A.A.; Oyagbemi, A.A.; Omobowale, T.O.; Ajibade, T.O.; Adedapo, A.A. Cardio-Protective and Antioxidant Properties of Caffeic Acid and Chlorogenic Acid: Mechanistic Role of Angiotensin Converting Enzyme, Cholinesterase and Arginase Activities in Cyclosporine Induced Hypertensive Rats. Biomed. Pharmacother. 2019, 109, 450–458. [Google Scholar] [CrossRef] [PubMed]
  24. de Araújo, F.F.; de Paulo Farias, D.; Neri-Numa, I.A.; Pastore, G.M. Polyphenols and Their Applications: An Approach in Food Chemistry and Innovation Potential. Food Chem. 2021, 338, 127535. [Google Scholar] [CrossRef] [PubMed]
  25. Chandrasekhar, Y.; Phani Kumar, G.; Ramya, E.M.; Anilakumar, K.R. Gallic Acid Protects 6-OHDA Induced Neurotoxicity by Attenuating Oxidative Stress in Human Dopaminergic Cell Line. Neurochem. Res. 2018, 43, 1150–1160. [Google Scholar] [CrossRef]
  26. Sakalauskas, A.; Ziaunys, M.; Smirnovas, V. Gallic Acid Oxidation Products Alter the Formation Pathway of Insulin Amyloid Fibrils. Sci. Rep. 2020, 10, 14466. [Google Scholar] [CrossRef]
  27. Seo, E.-J.; Fischer, N.; Efferth, T. Phytochemicals as Inhibitors of NF-ΚB for Treatment of Alzheimer’s Disease. Pharmacol. Res. 2018, 129, 262–273. [Google Scholar] [CrossRef]
  28. Ogut, E.; Armagan, K.; Gül, Z. The Role of Syringic Acid as a Neuroprotective Agent for Neurodegenerative Disorders and Future Expectations. Metab. Brain Dis. 2022, 37, 859–880. [Google Scholar] [CrossRef]
  29. Nouri, Z.; Fakhri, S.; El-Senduny, F.F.; Sanadgol, N.; Abd-ElGhani, G.E.; Farzaei, M.H.; Chen, J.-T. On the Neuroprotective Effects of Naringenin: Pharmacological Targets, Signaling Pathways, Molecular Mechanisms, and Clinical Perspective. Biomolecules 2019, 9, 690. [Google Scholar] [CrossRef] [Green Version]
  30. Enogieru, A.B.; Haylett, W.; Hiss, D.C.; Bardien, S.; Ekpo, O.E. Rutin as a Potent Antioxidant: Implications for Neurodegenerative Disorders. Oxid. Med. Cell. Longev. 2018, 2018, 6241017. [Google Scholar] [CrossRef] [Green Version]
  31. Xie, J.; Guo, L.; Pang, G.; Wu, X.; Zhang, M. Modulation Effect of Semen Ziziphi Spinosae Extracts on IL-1β, IL-4, IL-6, IL-10, TNF-α and IFN-γ in Mouse Serum. Nat. Prod. Res. 2011, 25, 464–467. [Google Scholar] [CrossRef]
  32. Li, B.; Han, L.; Cao, B.; Yang, X.; Zhu, X.; Yang, B.; Zhao, H.; Qiao, W. Use of Magnoflorine-Phospholipid Complex to Permeate Blood-Brain Barrier and Treat Depression in the CUMS Animal Model. Drug Deliv. 2019, 26, 566–574. [Google Scholar] [CrossRef] [Green Version]
  33. Yi, J.H.; Moon, S.; Cho, E.; Kwon, H.; Lee, S.; Jeon, J.; Park, A.Y.; Lee, Y.H.; Kwon, K.J.; Ryu, J.H.; et al. Hyperoside Improves Learning and Memory Deficits by Amyloid Β1-42 in Mice through Regulating Synaptic Calcium-Permeable AMPA Receptors. Eur. J. Pharmacol. 2022, 931, 175188. [Google Scholar] [CrossRef] [PubMed]
  34. Chen, L.; Zhou, Y.-P.; Liu, H.-Y.; Gu, J.-H.; Zhou, X.-F.; Yue-Qin, Z. Long-Term Oral Administration of Hyperoside Ameliorates AD-Related Neuropathology and Improves Cognitive Impairment in APP/PS1 Transgenic Mice. Neurochem. Int. 2021, 151, 105196. [Google Scholar] [CrossRef] [PubMed]
  35. Kim, S.; Chen, J.; Cheng, T.; Gindulyte, A.; He, J.; He, S.; Li, Q.; Shoemaker, B.A.; Thiessen, P.A.; Yu, B.; et al. PubChem in 2021: New Data Content and Improved Web Interfaces. Nucleic Acids Res. 2021, 49, D1388–D1395. [Google Scholar] [CrossRef] [PubMed]
  36. Buiu, C.; Putz, M.V.; Avram, S. Learning the Relationship between the Primary Structure of HIV Envelope Glycoproteins and Neutralization Activity of Particular Antibodies by Using Artificial Neural Networks. Int. J. Mol. Sci. 2016, 17, 1710. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Demir, Y.; Durmaz, L.; Taslimi, P.; Gulçin, İ. Antidiabetic Properties of Dietary Phenolic Compounds: Inhibition Effects on α-Amylase, Aldose Reductase, and α-Glycosidase. Biotechnol. Appl. Biochem. 2019, 66, 781–786. [Google Scholar] [CrossRef]
  38. Garbiec, E.; Cielecka-Piontek, J.; Kowalówka, M.; Hołubiec, M.; Zalewski, P. Genistein-Opportunities Related to an Interesting Molecule of Natural Origin. Mol. Basel Switz. 2022, 27, 815. [Google Scholar] [CrossRef]
  39. Zeng, X.; He, L.; Wang, S.; Wang, K.; Zhang, Y.; Tao, L.; Li, X.; Liu, S. Aconine Inhibits RANKL-Induced Osteoclast Differentiation in RAW264.7 Cells by Suppressing NF-ΚB and NFATc1 Activation and DC-STAMP Expression. Acta Pharmacol. Sin. 2016, 37, 255–263. [Google Scholar] [CrossRef] [Green Version]
  40. Zhang, L.; Siyiti, M.; Zhang, J.; Yao, M.; Zhao, F. Anti-Inflammatory and Anti-Rheumatic Activities in Vitro of Alkaloids Separated from Aconitum Soongoricum Stapf. Exp. Ther. Med. 2021, 21, 493. [Google Scholar] [CrossRef]
  41. Zhu, L.; Wu, J.; Zhao, M.; Song, W.; Qi, X.; Wang, Y.; Lu, L.; Liu, Z. Mdr1a Plays a Crucial Role in Regulating the Analgesic Effect and Toxicity of Aconitine by Altering Its Pharmacokinetic Characteristics. Toxicol. Appl. Pharmacol. 2017, 320, 32–39. [Google Scholar] [CrossRef]
  42. Çankal, D.; Akkol, E.K.; Kılınç, Y.; İlhan, M.; Capasso, R. An Effective Phytoconstituent Aconitine: A Realistic Approach for the Treatment of Trigeminal Neuralgia. Mediators Inflamm. 2021, 2021, 6676063. [Google Scholar] [CrossRef] [PubMed]
  43. Ji, B.-L.; Xia, L.-P.; Zhou, F.-X.; Mao, G.-Z.; Xu, L.-X. Aconitine Induces Cell Apoptosis in Human Pancreatic Cancer via NF-ΚB Signaling Pathway. Eur. Rev. Med. Pharmacol. Sci. 2016, 20, 4955–4964. [Google Scholar] [PubMed]
  44. Li, X.-M.; Liu, J.; Pan, F.-F.; Shi, D.-D.; Wen, Z.-G.; Yang, P.-L. Quercetin and Aconitine Synergistically Induces the Human Cervical Carcinoma HeLa Cell Apoptosis via Endoplasmic Reticulum (ER) Stress Pathway. PLOS ONE 2018, 13, e0191062. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Guo, B. Effects of Osthole, Psoralen, Aconitine on Breast Cancer MDA-MB-231BO Cell Line Inhibition in Vitro. J. Chin. Integr. Med. 2011, 9, 1110–1117. [Google Scholar] [CrossRef] [PubMed]
  46. Udrea, A.-M.; Dinache, A.; Staicu, A.; Avram, S. Target Prediction of 5,10,15,20-Tetrakis(4′-Sulfonatophenyl)-Porphyrin Using Molecular Docking. Pharmaceutics 2022, 14, 2390. [Google Scholar] [CrossRef]
  47. Tozar, T.; Santos Costa, S.; Udrea, A.-M.; Nastasa, V.; Couto, I.; Viveiros, M.; Pascu, M.L.; Romanitan, M.O. Anti-Staphylococcal Activity and Mode of Action of Thioridazine Photoproducts. Sci. Rep. 2020, 10, 18043. [Google Scholar] [CrossRef]
  48. Liu, M.; Cao, Y.; Lv, D.; Zhang, W.; Zhu, Z.; Zhang, H.; Chai, Y. Effect of Processing on the Alkaloids in Aconitum Tubers by HPLC-TOF/MS. J. Pharm. Anal. 2017, 7, 170–175. [Google Scholar] [CrossRef]
  49. Avram, S.; Milac, A.-L.; Mihailescu, D. 3D-QSAR Study Indicates an Enhancing Effect of Membrane Ions on Psychiatric Drugs Targeting Serotonin Receptor 5-HT1A. Mol. Biosyst. 2012, 8, 1418–1425. [Google Scholar] [CrossRef]
  50. Avram, S.; Buiu, C.; Duda-Seiman, D.; Duda-Seiman, C.; Borcan, F.; Mihailescu, D. Evaluation of the Pharmacological Descriptors Related to the Induction of Antidepressant Activity and Its Prediction by QSAR/QRAR Methods. Mini Rev. Med. Chem. 2012, 12, 467–476. [Google Scholar] [CrossRef]
  51. Avram, S.; Mernea, M.; Limban, C.; Borcan, F.; Chifiriuc, C. Potential Therapeutic Approaches to Alzheimer’s Disease By Bioinformatics, Cheminformatics And Predicted Adme-Tox Tools. Curr. Neuropharmacol. 2020, 18, 696–719. [Google Scholar] [CrossRef]
  52. Supuran, C.T. Carbonic Anhydrase Activators. Future Med. Chem. 2018, 10, 561–573. [Google Scholar] [CrossRef] [PubMed]
  53. Supuran, C.T. Carbonic Anhydrase Inhibitors. Bioorg. Med. Chem. Lett. 2010, 20, 3467–3474. [Google Scholar] [CrossRef]
  54. Avram, S.; Milac, A.L.; Carta, F.; Supuran, C.T. More Effective Dithiocarbamate Derivatives Inhibiting Carbonic Anhydrases, Generated by QSAR and Computational Design. J. Enzyme Inhib. Med. Chem. 2013, 28, 350–359. [Google Scholar] [CrossRef] [PubMed]
  55. Mishra, C.B.; Tiwari, M.; Supuran, C.T. Progress in the Development of Human Carbonic Anhydrase Inhibitors and Their Pharmacological Applications: Where Are We Today? Med. Res. Rev. 2020, 40, 2485–2565. [Google Scholar] [CrossRef]
  56. Ciccone, L.; Cerri, C.; Nencetti, S.; Orlandini, E. Carbonic Anhydrase Inhibitors and Epilepsy: State of the Art and Future Perspectives. Molecules 2021, 26, 6380. [Google Scholar] [CrossRef] [PubMed]
  57. Kida, E.; Palminiello, S.; Golabek, A.A.; Walus, M.; Wierzba-Bobrowicz, T.; Rabe, A.; Albertini, G.; Wisniewski, K.E. Carbonic Anhydrase II in the Developing and Adult Human Brain. J. Neuropathol. Exp. Neurol. 2006, 65, 664–674. [Google Scholar] [CrossRef] [Green Version]
  58. Aspatwar, A.; Tolvanen, M.E.E.; Ortutay, C.; Parkkila, S. Carbonic Anhydrase Related Proteins: Molecular Biology and Evolution. Subcell. Biochem. 2014, 75, 135–156. [Google Scholar] [CrossRef]
  59. Mishra, C.B.; Kumari, S.; Angeli, A.; Bua, S.; Tiwari, M.; Supuran, C.T. Discovery of Benzenesulfonamide Derivatives as Carbonic Anhydrase Inhibitors with Effective Anticonvulsant Action: Design, Synthesis, and Pharmacological Evaluation. J. Med. Chem. 2018, 61, 3151–3165. [Google Scholar] [CrossRef]
  60. Balestri, F.; Moschini, R.; Mura, U.; Cappiello, M.; Corso, A.D. In Search of Differential Inhibitors of Aldose Reductase. Biomolecules 2022, 12, 485. [Google Scholar] [CrossRef]
  61. Veeresham, C.; Rama Rao, A.; Asres, K. Aldose Reductase Inhibitors of Plant Origin. Phytother. Res. PTR 2014, 28, 317–333. [Google Scholar] [CrossRef]
  62. Grewal, A.S.; Bhardwaj, S.; Pandita, D.; Lather, V.; Sekhon, B.S. Updates on Aldose Reductase Inhibitors for Management of Diabetic Complications and Non-Diabetic Diseases. Mini Rev. Med. Chem. 2016, 16, 120–162. [Google Scholar] [CrossRef] [PubMed]
  63. Huang, Y.-K.; Liu, C.-C.; Wang, S.; Cheng, H.-C.; Meadows, C.; Chang, K.-C. The Role of Aldose Reductase in Beta-Amyloid-Induced Microglia Activation. Int. J. Mol. Sci. 2022, 23, 15088. [Google Scholar] [CrossRef] [PubMed]
  64. SIB Swiss Institute of Bioinformatics | Expasy. Available online: https://www.expasy.org/ (accessed on 8 February 2023).
  65. 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 Deliv. Rev. 2001, 46, 3–26. [Google Scholar] [CrossRef]
  66. 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]
  67. Egan, W.J.; Merz, K.M.; Baldwin, J.J. Prediction of Drug Absorption Using Multivariate Statistics. J. Med. Chem. 2000, 43, 3867–3877. [Google Scholar] [CrossRef] [PubMed]
  68. Daina, A.; Michielin, O.; Zoete, V. SwissADME: A Free Web Tool to Evaluate Pharmacokinetics, Drug-Likeness and Medicinal Chemistry Friendliness of Small Molecules. Sci. Rep. 2017, 7, 42717. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Martin, Y.C. A Bioavailability Score. J. Med. Chem. 2005, 48, 3164–3170. [Google Scholar] [CrossRef]
  70. Kesarwani, K.; Gupta, R. Bioavailability Enhancers of Herbal Origin: An Overview. Asian Pac. J. Trop. Biomed. 2013, 3, 253–266. [Google Scholar] [CrossRef] [Green Version]
  71. Nistorescu, S.; Gradisteanu Pircalabioru, G.; Udrea, A.-M.; Simon, A.; Pascu, M.L.; Chifiriuc, M.-C. Laser-Irradiated Chlorpromazine as a Potent Anti-Biofilm Agent for Coating of Biomedical Devices. Coatings 2020, 10, 1230. [Google Scholar] [CrossRef]
  72. Udrea, A.-M.; Dinache, A.; Pagès, J.-M.; Pirvulescu, R.A. Quinazoline Derivatives Designed as Efflux Pump Inhibitors: Molecular Modeling and Spectroscopic Studies. Molecules 2021, 26, 2374. [Google Scholar] [CrossRef]
  73. Pervin, M.; Unno, K.; Takagaki, A.; Isemura, M.; Nakamura, Y. Function of Green Tea Catechins in the Brain: Epigallocatechin Gallate and Its Metabolites. Int. J. Mol. Sci. 2019, 20, E3630. [Google Scholar] [CrossRef] [Green Version]
  74. Supuran, C.T. Coumarin Carbonic Anhydrase Inhibitors from Natural Sources. J. Enzyme Inhib. Med. Chem. 2020, 35, 1462–1470. [Google Scholar] [CrossRef] [PubMed]
  75. Bickerton, G.R.; Paolini, G.V.; Besnard, J.; Muresan, S.; Hopkins, A.L. Quantifying the Chemical Beauty of Drugs. Nat. Chem. 2012, 4, 90–98. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. 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]
  77. Avram, S.; Mernea, M.; Borcan, F.; Mihailescu, D. Evaluation of the Therapeutic Properties of Mastoparan- and Sifuvirtide- Derivative Antimicrobial Peptides Using Chemical Structure-Function Relationship—in Vivo and in Silico Approaches. Curr. Drug Deliv. 2016, 13, 202–210. [Google Scholar] [CrossRef] [PubMed]
  78. Daina, A.; Zoete, V. A BOILED-Egg To Predict Gastrointestinal Absorption and Brain Penetration of Small Molecules. ChemMedChem 2016, 11, 1117–1121. [Google Scholar] [CrossRef] [Green Version]
  79. Moret, M.; Pachon Angona, I.; Cotos, L.; Yan, S.; Atz, K.; Brunner, C.; Baumgartner, M.; Grisoni, F.; Schneider, G. Leveraging Molecular Structure and Bioactivity with Chemical Language Models for de Novo Drug Design. Nat. Commun. 2023, 14, 114. [Google Scholar] [CrossRef]
  80. Dumitrascu, F.; Udrea, A.-M.; Caira, M.R.; Nuta, D.C.; Limban, C.; Chifiriuc, M.C.; Popa, M.; Bleotu, C.; Hanganu, A.; Dumitrescu, D.; et al. In Silico and Experimental Investigation of the Biological Potential of Some Recently Developed Carprofen Derivatives. Molecules 2022, 27, 2722. [Google Scholar] [CrossRef] [PubMed]
  81. Putz, M.V.; Duda-Seiman, C.; Duda-Seiman, D.; Putz, A.-M.; Alexandrescu, I.; Mernea, M.; Avram, S. Chemical Structure-Biological Activity Models for Pharmacophores’ 3D-Interactions. Int. J. Mol. Sci. 2016, 17, 1087. [Google Scholar] [CrossRef] [Green Version]
  82. Caron, G.; Digiesi, V.; Solaro, S.; Ermondi, G. Flexibility in Early Drug Discovery: Focus on the beyond-Rule-of-5 Chemical Space. Drug Discov. Today 2020, 25, 621–627. [Google Scholar] [CrossRef]
  83. Raevsky, O.A.; Grigorev, V.Y.; Polianczyk, D.E.; Raevskaja, O.E.; Dearden, J.C. Aqueous Drug Solubility: What Do We Measure, Calculate and QSPR Predict? Mini Rev. Med. Chem. 2019, 19, 362–372. [Google Scholar] [CrossRef] [PubMed]
  84. Vraka, C.; Mijailovic, S.; Fröhlich, V.; Zeilinger, M.; Klebermass, E.-M.; Wadsak, W.; Wagner, K.-H.; Hacker, M.; Mitterhauser, M. Expanding LogP: Present Possibilities. Nucl. Med. Biol. 2018, 58, 20–32. [Google Scholar] [CrossRef] [PubMed]
  85. Chomicki, D.; Kharchenko, O.; Skowronski, L.; Kowalonek, J.; Kozanecka-Szmigiel, A.; Szmigiel, D.; Smokal, V.; Krupka, O.; Derkowska-Zielinska, B. Physico-Chemical and Light-Induced Properties of Quinoline Azo-Dyes Polymers. Int. J. Mol. Sci. 2020, 21, 5755. [Google Scholar] [CrossRef]
  86. Pires, D.E.V.; Blundell, T.L.; Ascher, D.B. PkCSM: Predicting Small-Molecule Pharmacokinetic and Toxicity Properties Using Graph-Based Signatures. J. Med. Chem. 2015, 58, 4066–4072. [Google Scholar] [CrossRef] [PubMed]
  87. Udrea, A.M.; Gradisteanu Pircalabioru, G.; Boboc, A.A.; Mares, C.; Dinache, A.; Mernea, M.; Avram, S. Advanced Bioinformatics Tools in the Pharmacokinetic Profiles of Natural and Synthetic Compounds with Anti-Diabetic Activity. Biomolecules 2021, 11, 1692. [Google Scholar] [CrossRef] [PubMed]
  88. Avram, S.; Udrea, A.M.; Negrea, A.; Ciopec, M.; Duteanu, N.; Postolache, C.; Duda-Seiman, C.; Duda-Seiman, D.; Shaposhnikov, S. Prevention of Deficit in Neuropsychiatric Disorders through Monitoring of Arsenic and Its Derivatives as Well as Through Bioinformatics and Cheminformatics. Int. J. Mol. Sci. 2019, 20, E1804. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Avram, S.; Mernea, M.; Mihailescu, D.; Duda-Seiman, D.; Duda-Seiman, C. Advanced QSAR Methods Evaluated Polycyclic Aromatic Compounds Duality as Drugs and Inductors in Psychiatric Disorders. Curr. Org. Chem. 2013, 17, 2880–2890. [Google Scholar] [CrossRef]
  90. Hikino, H.; Konno, C.; Takata, H.; Yamada, Y.; Yamada, C.; Ohizumi, Y.; Sugio, K.; Fujimura, H. Antiinflammatory Principles of Aconitum Roots. J. Pharmacobiodyn. 1980, 3, 514–525. [Google Scholar] [CrossRef] [Green Version]
  91. Bai, L.; Li, X.; He, L.; Zheng, Y.; Lu, H.; Li, J.; Zhong, L.; Tong, R.; Jiang, Z.; Shi, J.; et al. Antidiabetic Potential of Flavonoids from Traditional Chinese Medicine: A Review. Am. J. Chin. Med. 2019, 47, 933–957. [Google Scholar] [CrossRef]
  92. Daina, A.; Michielin, O.; Zoete, V. SwissTargetPrediction: Updated Data and New Features for Efficient Prediction of Protein Targets of Small Molecules. Nucleic Acids Res. 2019, 47, W357–W364. [Google Scholar] [CrossRef] [Green Version]
  93. Negrea, E.; Oancea, P.; Leonties, A.; Ana Maria, U.; Avram, S.; Raducan, A. Spectroscopic Studies on Binding of Ibuprofen and Drotaverine with Bovine Serum Albumin. J. Photochem. Photobiol. Chem. 2023, 438, 114512. [Google Scholar] [CrossRef]
  94. Molecular Operating Environment (MOE). 2022.02 Chemical Computing Group ULC, 1010 Sherbooke St. West, Suite #910, Montreal, QC, Canada, H3A 2R7. 2022. Available online: https://www.chemcomp.com/Research-Citing_MOE.htm (accessed on 15 May 2023).
  95. Open Babel: An Open Chemical Toolbox | Journal of Cheminformatics | Full Text. Available online: https://jcheminf.biomedcentral.com/articles/10.1186/1758-2946-3-33 (accessed on 16 March 2022).
  96. Harrison, D.H.; Bohren, K.M.; Ringe, D.; Petsko, G.A.; Gabbay, K.H. An Anion Binding Site in Human Aldose Reductase: Mechanistic Implications for the Binding of Citrate, Cacodylate, and Glucose 6-Phosphate. Biochemistry 1994, 33, 2011–2020. [Google Scholar] [CrossRef] [PubMed]
  97. Morris, G.M.; Huey, R.; Olson, A.J. Using AutoDock for Ligand-Receptor Docking. Curr. Protoc. Bioinforma. 2008, 24, 8–14. [Google Scholar] [CrossRef] [PubMed]
  98. Gao, X.; Zhang, X.; Hu, J.; Xu, X.; Zuo, Y.; Wang, Y.; Ding, J.; Xu, H.; Zhu, S. Aconitine Induces Apoptosis in H9c2 Cardiac Cells via Mitochondria-mediated Pathway. Mol. Med. Rep. 2018, 17, 284–292. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Ravindran, J.; Gupta, N.; Agrawal, M.; Bala Bhaskar, A.S.; Lakshmana Rao, P.V. Modulation of ROS/MAPK Signaling Pathways by Okadaic Acid Leads to Cell Death via, Mitochondrial Mediated Caspase-Dependent Mechanism. Apoptosis Int. J. Program. Cell Death 2011, 16, 145–161. [Google Scholar] [CrossRef]
  100. Anjum, F.; Ali, F.; Mohammad, T.; Shafie, A.; Akhtar, O.; Abdullaev, B.; Hassan, I. Discovery of Natural Compounds as Potential Inhibitors of Human Carbonic Anhydrase II: An Integrated Virtual Screening, Docking, and Molecular Dynamics Simulation Study. Omics J. Integr. Biol. 2021, 25, 513–524. [Google Scholar] [CrossRef]
  101. Taslimi, P.; Caglayan, C.; Gulcin, İ. The Impact of Some Natural Phenolic Compounds on Carbonic Anhydrase, Acetylcholinesterase, Butyrylcholinesterase, and α-Glycosidase Enzymes: An Antidiabetic, Anticholinergic, and Antiepileptic Study. J. Biochem. Mol. Toxicol. 2017, 31. [Google Scholar] [CrossRef]
  102. Aggul, A.G.; Uzun, N.; Kuzu, M.; Taslimi, P.; Gulcin, I. Some Phenolic Natural Compounds as Carbonic Anhydrase Inhibitors: An in Vitro and in Silico Study. Arch. Pharm. (Weinheim) 2022, 355, e2100476. [Google Scholar] [CrossRef]
  103. Sethi, K.K.; Sahoo, S.K.; Pichikala, J.N.; Suresh, P. Carbonic Anhydrase I and II Inhibition with Natural Products: Caffeine and Piperine. J. Enzyme Inhib. Med. Chem. 2012, 27, 97–100. [Google Scholar] [CrossRef]
  104. BIOVIA. Dassault Systèmes, [Discovery Studio Visualizer], [V21.1.0.20298]; Dassault Systèmes: San Diego, CA, USA, 2021. [Google Scholar]
  105. Chahal, V.; Kakkar, R. A Combination Strategy of Structure-Based Virtual Screening, MM-GBSA, Cross Docking, Molecular Dynamics and Metadynamics Simulations Used to Investigate Natural Compounds as Potent and Specific Inhibitors of Tumor Linked Human Carbonic Anhydrase IX. J. Biomol. Struct. Dyn. 2022, 23, 1–16. [Google Scholar] [CrossRef]
Figure 2. The chemical structures of the compounds that showed the most important antioxidant activity, quercetin (PubChem CID5280343), rutin (PubChem CID5280805), magnoflorine (PubChem CID73337) [35].
Figure 2. The chemical structures of the compounds that showed the most important antioxidant activity, quercetin (PubChem CID5280343), rutin (PubChem CID5280805), magnoflorine (PubChem CID73337) [35].
Pharmaceuticals 16 00842 g002
Figure 3. Compliance with the drug-like rules—the Lipinski, Ghose, Veber, Egan, and Muegge rules.
Figure 3. Compliance with the drug-like rules—the Lipinski, Ghose, Veber, Egan, and Muegge rules.
Pharmaceuticals 16 00842 g003
Figure 4. Bioavailability score values for natural compounds obtained from Aconitum toxicum Rchb., Anemone nemorosa L., and Helleborus odorus Waldst. & Kit. ex Willd.
Figure 4. Bioavailability score values for natural compounds obtained from Aconitum toxicum Rchb., Anemone nemorosa L., and Helleborus odorus Waldst. & Kit. ex Willd.
Pharmaceuticals 16 00842 g004
Figure 5. A 2D visualization of the interactions between quercetin and AKR1B1 amino acid residues Discovery Studio Visualizer was used to obtain the image [104].
Figure 5. A 2D visualization of the interactions between quercetin and AKR1B1 amino acid residues Discovery Studio Visualizer was used to obtain the image [104].
Pharmaceuticals 16 00842 g005
Table 2. The pharmacological profile of natural compounds using descriptors such as flexibility, refractivity, topological polar surface area (TPSA), hydrophobicity, and solubility.
Table 2. The pharmacological profile of natural compounds using descriptors such as flexibility, refractivity, topological polar surface area (TPSA), hydrophobicity, and solubility.
CompoundFlexibilityRefractivityTPSAHydrophobicitySolubility
Aconitine11165.5153.453.61−3.39
Hypaconitine10159.53133.224.3−3.65
Mesaconitine10160.69153.453.56−3.14
Magnoflorine2101.8758.92−0.66−3.91
Gallic acid6119.1564.993.97−6.25
Catechin174.33110.381.47−2.22
Ferulic acid351.6366.761.62−2.11
Caffeic acid247.1677.760.97−1.89
Chlorogenic acid583.5164.750.96−1.62
Epicatechin174.33110.381.47−2.22
Delphinidin175.26134.191.35−2.35
Coumaric acid245.1357.530.95−2.02
Daidzein171.9770.671.77−3.53
Hyperoside4110.16210.512.11−3.04
Rutin6141.38269.431.58−3.3
Naringin6134.91225.062.38−2.98
Malvidin387.13112.52−1.96−3.6
Quercetin178.03131.361.63−3.16
Naringenin171.5786.991.75−3.49
Genistein173.9990.91.91−3.72
Syringic acid348.4175.991.54−1.84
Table 3. Pharmacokinetic properties of natural compounds.
Table 3. Pharmacokinetic properties of natural compounds.
CompoundIntestinal Absorption (Human)BioavailabilityBBB PermeabilityFraction Unbound (Human)OCT2 InhibitorOCT1 Inhibitor
ACyesnoyes0.52nono
Hypaconitineyesnoyes0.78nono
Mesaconitineyesnoyes0.53nono
Magnoflorinenonoyes0.19noyes
Gallic acidyesnoyes0.99nono
Catechinyesnono1.01nono
Ferulic acidyesyesyes0.72nono
Caffeic acidyesyesno0.73nono
Chlorogenic acidyesnono0.63nono
Epicatechinyesnono1.01nono
Delphinidinyesyesno0.85nono
Coumaric acidyesnono0.50nono
Daidzeinyesyesno0.96Nono
Hyperosideyesnono0.79nono
Rutinyesnono0.96nono
Naringinyesnono0.73nono
Malvidinyesnoyes0.88nono
Quercetinyesnono1.16nono
Naringeninyesnono0.93nono
Genisteinyesnono1.09nono
Syringic acidyesyesyes0.55nono
Table 4. Pharmacogenomic profile predicted for analyzed compounds.
Table 4. Pharmacogenomic profile predicted for analyzed compounds.
CompoundsCYP1A2 InhibitorCYP2C19 InhibitorCYP2C9 InhibitorCYP2C9 SubstrateCYP2D6 InhibitorCYP2D6 SubstrateCYP3A4 InhibitorCYP3A4 Substrate
ACnononononononoyes
Hypaconitinenononononononoyes
Mesaconitinenononononononoyes
Magnoflorinenononononoyesnoyes
Gallic acidnoyesyesnonononono
Catechinnononononoyesnono
ferulic acidnononononononono
Caffeic acidnononononononono
Chlorogenic acidnononoyesnononoyes
Epicatechinnononononoyesnono
Delphinidinyesyesyesnononoyesno
Coumaric acidnononononononono
Daidzeinyesyesyesnonononono
Hyperosidenononononononoyes
Rutinnononononononoyes
Naringinnononononononoyes
Malvidinyesyesyesnononoyesno
Quercetinyesnononononoyesyes
Naringeninyesyesyesnononoyesno
Genisteinyesyesyesnononoyesno
Syringic acidnononononononono
Table 5. Natural compounds predicted toxicity.
Table 5. Natural compounds predicted toxicity.
CompoundsAmes MutagenesisAvian ToxicityCrustacea Aquatic ToxicityToxicities
Fish
HepatotoxicityMitochondrial Toxicity
ACnonoyesyesyesyes
Hypaconitinenonoyesyesnoyes
Mesaconitinenononoyesyesyes
Magnoflorinenonoyesyesyesyes
Gallic acidnononoyesyesno
Catechinyesnoyesyesnoyes
Ferulic acidnononoyesnono
Caffeic acidnononoyesnono
Chlorogenic acidnononoyesnoyes
Epicatechinyesnoyesyesnoyes
Delphinidinnononoyesyesyes
Coumaric acidnononoyesnono
Daidzeinnononoyesyesno
Hyperosideyesnonoyesyesyes
Rutinyesnonoyesyesno
Naringinnononoyesyesno
Malvidinnonoyesyesyesyes
Quercetinyesnonoyesyesyes
Naringeninnononoyesyesyes
Genisteinnononoyesyesyes
Syringic acidnononoyesnono
Table 6. Targets obtain using SwissTargetPrediction tool.
Table 6. Targets obtain using SwissTargetPrediction tool.
Target NameTarget AbbreviationBinding Probability
Caffeic acid
Carbonic anhydrase IICA20.72
Arachidonate 5nolipoxygenaseALOX50.72
Carbonic anhydrase VIICA70.72
Carbonic anhydrase ICA10.7
Carbonic anhydrase VICA60.7
Chlorogenic acid
Aldose reductaseAKR1B10.87
Aldonoketo reductase family 1 member B10AKR1B100.74
Coumaric acid
Aldose reductaseAKR1B11
Carbonic anhydrase IICA21
Carbonic anhydrase VIICA71
Estrogen receptor betaESR21
Carbonic anhydrase ICA11
Carbonic anhydrase IIICA31
Carbonic anhydrase VICA61
Carbonic anhydrase XIICA121
Carbonic anhydrase XIVCA141
Carbonic anhydrase IXCA91
Carbonic anhydrase IVCA41
Carbonic anhydrase VBCA5B1
Carbonic anhydrase VACA5A1
Ferulic acid
Carbonic anhydrase IICA20.93
Carbonic anhydrase VIICA70.93
Carbonic anhydrase ICA10.93
Carbonic anhydrase VICA60.93
Carbonic anhydrase XIICA120.9
Carbonic anhydrase XIVCA140.901
Carbonic anhydrase IXCA90.901
Carbonic anhydrase VACA5A0.9
Gallic acid
without statistical significance
Syringic acid
Carbonic anhydrase IICA21
Carbonic anhydrase VIICA71
Carbonic anhydrase ICA11
Carbonic anhydrase IIICA31
Carbonic anhydrase VICA61
Carbonic anhydrase XIICA121
Carbonic anhydrase XIVCA141
Carbonic anhydrase IXCA91
Carbonic anhydrase VACA5A1
AC
without statistical significance
Daidzein
Aldehyde dehydrogenaseALDH21
Estrogen receptor alphaESR11
Carbonic anhydrase VIICA71
Estrogen receptor betaESR21
Carbonic anhydrase XIICA121
Carbonic anhydrase IVCA41
Delfinidin
without statistical significance
Genistein
ThromboxanenoA synthaseTBXAS11
Monoamine oxidase AMAOA1
Epidermal growth factor receptor erbB1EGFR1
Estrogen receptor alphaESR11
MaltasenoglucoamylaseMGAM1
Serotonin 2a (5noHT2a) receptorHTR2A1
Serotonin 2c (5noHT2c) receptorHTR2C1
Adenosine A1 receptor (by homology)ADORA11
Estrogen receptor betaESR21
Adenosine A2a receptorADORA2A1
Estradiol 17nobetanodehydrogenase 1HSD17B11
Estrogennorelated receptor alphaESRRA1
Estrogennorelated receptor betaESRRB1
ATPnobinding cassette subnofamily G member 2ABCG21
Hypaconitine
Dopamine transporter (by homology)SLC6A30.093576
HERGKCNH20.074565
Hyperoside
Aldose reductaseAKR1B11
Carbonic anhydrase IICA21
Carbonic anhydrase VIICA71
Carbonic anhydrase XIICA121
Carbonic anhydrase IVCA41
Magnoflorine
Dopamine D2 receptorDRD20.548778
Neuronal acetylcholine receptor; alpha4/beta2CHRNA4 CHRNB20.390156
Dopamine D3 receptorDRD30.306473
Malvidin
without statistical significance
Glyoxalase IGLO10.128899
Xanthine dehydrogenaseXDH0.112748
Mesaconitine
without statistical significance
Naringenin
Cytochrome P450 19A1CYP19A10.91
Carbonic anhydrase VIICA70.91
Multidrug resistance no associated protein 1ABCC10.91
Estradiol 17nobetanodehydrogenase 1HSD17B10.91
Carbonic anhydrase XIICA120.91
Test is no specific androgen no binding proteinSHBG0.91
Carbonic anhydrase IVCA40.91
Cytochrome P450 1B1CYP1B10.9
Carbonyl reductase [NADPH] 1CBR10.9
Naringin
Cytochrome P450 19A1CYP19A11
Quercetin
NADPH oxidase 4NOX41
Vasopressin V2 receptorAVPR21
Aldose reductaseAKR1B11
Xanthine dehydrogenaseXDH1
Monoamine oxidase AMAOA1
Insulinnolike growth factor I receptorIGF1R1
Rutin
NeuromedinnoU receptor 2NMUR21
Alphano2a adrenergic receptorADRA2A1
Adrenergic receptor alphano2ADRA2C1
AcetylcholinesteraseACHE1
Aldose reductaseAKR1B10.60286
Table 7. The LEFEB between AKR1B1 and quercetin, caffeic acid, chlorogenic acid, coumaric acid, hyperoside, and rutin.
Table 7. The LEFEB between AKR1B1 and quercetin, caffeic acid, chlorogenic acid, coumaric acid, hyperoside, and rutin.
CompoundLEFEB
Quercetin−7.85 kcal/mol
Caffeic acid−6.21 kcal/mol
Chlorogenic acid−5.81 kcal/mol
Coumaric acid−5.60 kcal/mol
Hyperoside−5.59 kcal/mol
Rutin−4.54 kcal/mol
Table 8. Chemical compounds name, SMILES format, Molecular Formula and Molecular Weight of the analyzed structure.
Table 8. Chemical compounds name, SMILES format, Molecular Formula and Molecular Weight of the analyzed structure.
NameSMILESFormulaMW (Da)
ACCOCC12CN(CC)C3C4(C2C(OC)C3C2(C3C4CC(C3OC(=O)c3ccccc3)(C(C2O)OC)O)OC(=O)C)C(CC1O)OCC34H47NO11645.74
HypaconitineCOCC12CCC(C34C2C(OC)C(C3N(C1)C)C1(C2C4CC(C2OC(=O)c2ccccc2)(C(C1O)OC)O)OC(=O)C)OCC33H45NO10615.71
MesaconitineCOCC12CN(C)C3C4(C2C(OC)C3C2(C3C4CC(C3OC(=O)c3ccccc3)(C(C2O)OC)O)OC(=O)C)C(CC1O)OCC33H45NO11631.71
MagnoflorineCOc1ccc2c(c1O)c1c(O)c(OC)cc3c1C(C2)[N+](C)(C)CC3C20H24NO4+342.41
Gallic acidO=C(c1cc(O)c2c(c1)OC(O2)(c1ccccc1)c1ccccc1)OCc1ccccc1C27H20O5424.44
CatechinOc1cc2OC(c3ccc(c(c3)O)O)C(Cc2c(c1)O)OC15H14O6290.27
Ferulic acidCOc1cc(C=CC(=O)O)ccc1OC10H10O4194.18
Caffeic acidOC(=O)C=Cc1ccc(c(c1)O)OC9H8O4180.16
Chlorogenic acidO=C(OC1CC(O)(CC(C1O)O)C(=O)O)C=Cc1ccc(c(c1)O)OC16H18O9354.31
EpicatechinOc1cc2OC(c3ccc(c(c3)O)O)C(Cc2c(c1)O)OC15H14O6290.27
DelfinidinO=c1cc2oc(c3cc(O)c(c(c3)O)[O-])c(cc2c(c1)O)OC15H9O7-301.23
Coumaric acidOC(=O)C=Cc1ccc(cc1)OC9H8O3164.16
DaidzeinOc1ccc(cc1)c1coc2c(c1=O)ccc(c2)OC15H10O4254.24
HyperosideOCC1OC(Oc2c(oc3c(c2=O)c(O)cc(c3)O)c2ccc(c(c2)O)O)C(C(C1O)O)OC21H20O12464.38
RutinOc1cc(O)c2c(c1)oc(c(c2=O)OC1OC(COC2OC(C)C(C(C2O)O)O)C(C(C1O)O)O)c1ccc(c(c1)O)OC27H30O16610.52
NaringinOCC1OC(Oc2cc(O)c3c(c2)OC(CC3=O)c2ccc(cc2)O)C(C(C1O)O)
OC1OC(C)C(C(C1O)O)O
C27H32O14580.53
MalvidinCOc1cc(cc(c1O)OC)c1[o+]c2cc(O)cc(c2cc1O)OC17H15O7+331.3
QuercetinOc1cc(O)c2c(c1)oc(c(c2=O)O)c1ccc(c(c1)O)OC15H10O7302.24
NaringeninOc1ccc(cc1)C1CC(=O)c2c(O1)cc(cc2O)OC15H12O5272.25
GenisteinOc1ccc(cc1)c1coc2c(c1=O)c(O)cc(c2)OC15H10O5270.24
Syringic acidCOc1cc(cc(c1O)OC)C(=O)OC9H10O5198.17
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

Mareş, C.; Udrea, A.-M.; Şuţan, N.A.; Avram, S. Bioinformatics Tools for the Analysis of Active Compounds Identified in Ranunculaceae Species. Pharmaceuticals 2023, 16, 842. https://doi.org/10.3390/ph16060842

AMA Style

Mareş C, Udrea A-M, Şuţan NA, Avram S. Bioinformatics Tools for the Analysis of Active Compounds Identified in Ranunculaceae Species. Pharmaceuticals. 2023; 16(6):842. https://doi.org/10.3390/ph16060842

Chicago/Turabian Style

Mareş, Cătălina, Ana-Maria Udrea, Nicoleta Anca Şuţan, and Speranţa Avram. 2023. "Bioinformatics Tools for the Analysis of Active Compounds Identified in Ranunculaceae Species" Pharmaceuticals 16, no. 6: 842. https://doi.org/10.3390/ph16060842

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