Skip to main content

ORIGINAL RESEARCH article

Front. Chem., 29 November 2023
Sec. Theoretical and Computational Chemistry

In silico studies on cytotoxicity and antitumoral activity of acetogenins from Annona muricata L

  • 1Faculty of Pharmacy, Federal University of Para, Belem, Brazil
  • 2Postgraduate Program in Pharmaceutical Sciences, Federal University of Para, Belem, Brazil
  • 3Postgraduate Program in Biodiversity and Biotechnology (BIONORTE), Federal University of Para, Belem, Brazil

As life expectancy increases, the number of people affected by cancer is increasing. The available drugs still cause several adverse reactions, and it is important to look for less toxic drugs that act on resistant cancers. The present study evaluated the antitumor potential of acetogenins. Through a literature review, 44 acetogenins isolated from Annona muricata were selected and subjected to in silico studies to predict the physicochemical properties, pharmacokinetics (Preadmet and Admet lab), toxicity (Preadmet and Protox II) and molecular docking in caspase 3 (DockThor). For muricatacin, a literature review was carried out for antitumor activity and cytotoxicity. Only muricatacin met all physicochemical criteria, while all compounds showed high cutaneous and intestinal absorption (HIA), moderate permeability in Madin-Darby canine kidney and Caco2 cells, strongly bound plasma proteins, freely crossed the blood-brain barrier, inhibited CYP2C19, CYP2C9 and CYP3A4 and have an affinity for CYP3A4, being metabolized by it, an undesirable characteristic for antitumor drugs. All compounds were toxic in at least one model, while compound 28 was not carcinogenic in rats and mice. Compounds 13, 14, 15, 16, 17 and 28 were selected for molecular docking into Caspase 3. Docking showed hydrophobic interactions, hydrogen and covalent bonds performed to maintain the stability of caspase 3, and cis-uvariamicin IV stood out more through the energies and chemical bonds of this parameter. The chloroform fraction from the methanolic extract of the seeds showed activity against triple-negative breast cancer, both in vitro and in vivo, and only muricatacin has studies in which the antitumor activity was evaluated in vitro and showed to be very promising. In summary, muricatacin and cis-uvariamicin IV appear to be very promising as antitumors, especially cis-uvariamicin IV.

1 Introduction

Cancer or malignant neoplasm is a physiological condition, characterized by the disordered growth of cells that tend to invade tissues and organs, compromising their structure and, consequently, their functioning. The causes are multifactorial, usually associated with hormones, immunological conditions, genetics, exposure to chemical substances, viruses, lifestyle habits, among others (Deminice, 2022).

It is estimated that for every three people, one will be diagnosed with cancer, mainly associated with socioeconomic conditions. In 2018, studies showed the occurrence of 18 million new cases of cancer worldwide, and 9.6 million deaths (WHO, 2009). In Brazil, 704 thousand new cases of cancer are expected for each year of the 2023–2025 triennium (INCA, 2022).

In general, cancer is treated through surgery, chemotherapy, radiotherapy, immunotherapy and hormonal therapy, with chemotherapy and radiotherapy being the most requested (Deminice, 2022). Despite advances in treatment, the mortality rate is still high and may be related to late diagnosis, as well as inadequate response to treatment (Oliveira, 2002; Mansoori et al., 2017).

Tumor cells often acquire resistance to chemotherapy drugs, which may involve different mechanisms, in which the induction of multidrug resistance (MDR) is an important one (Nantajit et al., 2015). In addition to tumor resistance, available drugs are still quite toxic and can cause low-severity adverse effects such as fatigue, generalized pain (Pearce et al., 2017), nausea, vomiting (Lorusso et al., 2017) and other gastrointestinal disorders (Pearce et al., 2017), as well as highly serious events such as neurotoxicity and cardiotoxicity (Adão et al., 2013; Simão et al., 2019).

Even with the availability of antitumor drugs, it is necessary to search for therapeutic alternatives that are active in resistant tumors and less toxic to patients. Plants are important sources of pharmaceuticals (Amanda et al., 2022), and the Annona muricata is a species that is widely exploited by the cosmetic industry in the manufacture of moisturizers, shampoo, soap, and in the food industry for the preparation of sweets, drinks, ice cream, among others (Maul et al., 2000; De and Abreu, 2019).

The A. muricata is also used for medicinal purposes. The bath with leaves is used to treat skin diseases and pain in South Pacific countries, Mauritius and New Guinea (Longuefosse and Nossin, 1996; Sreekeesoon and Mahomoodally, 2014). A decoction of the leaves for ingestion is used as an analgesic in Brazil (Ross, 2010), Martinique (Longuefosse and Nossin, 1996), Mexico and Nicaragua (Ross, 2010).

In Benin (Kossouoh et al., 2007), Caribbean (Joyeux et al., 1995), Cuba (Beyra et al., 2004) and Mexico (Haiat et al., 2009) A. muricata is used to control free radicals and strengthening the immune system cells in order to avoid possible discomfort associated with colds, flu and asthma. The natives of Malaysia used the leaves to treat external and internal skin parasites (Badrie and Schauss, 2010). There are still reports of its use to treat malaria in Cameroon, Togo and Vietnam (Nguyen-pouplin and Julie, 2007; Ross, 2010; Boyom and Fekam, 2011). Furthermore, studies have shown this plant could be used to prevent breast, colorectal, lung and prostate cancer (Martínez, 1991; Moghadamtousi et al., 2015; Alonso-Castro and Josabad, 2011; Atawodi Eneojo, 2011; Bussmann and Rainer, 2010, Monigatti et al., 2013).

Chemical studies carried out on A. muricata led to the isolation of alkaloids, acetogenins, flavonoids, phenols, cyclopeptides (Moghadamtousi et al., 2015; Coria-téllez and Ana, 2018; Ferreira, 2022), carotenoids, and amides (Vijayameena et al., 2013). The acetogenins were characterized as majority compounds on this species (Mutakin et al., 2022).

Acetogenins, or polyketides, are metabolites derived from fatty acids. They have a long aliphatic chain of 35–38 carbons linked to a g-lactone ring, terminally substituted by β-unsaturated methyl, with tetrahydrofurans (THF) located along the hydrocarbon chain (Mutakin et al., 2022), being the presence of the hydroxyl groups, THF ring and the α, β-unsaturated y-lactone subunit responsible for producing the toxic effect of these compounds (Ferreira, 2022).

Studies have demonstrated that acetogenins have antitumor activity (Syed and Syed, 2016; Machado, 2019) and the mechanisms related to this activity are: 1-inhibition of ubiquinone oxidoreductase NADH, an enzyme essential to complex I of the electron transport chain (Carmen Zafra-Polo et al., 1998; Nakanishi, 2011); 2- inhibition of NADH oxidases in the plasma membrane, an enzyme involved in the production of adenosine triphosphate (ATP) under anaerobic conditions in order to restore coenzyme nicotinamide adenine dinucleotide - NAD levels (Alali and Feras, 1998; Moraes and Vieira, 2016); 3- phosphorylation of histone 3 (H3) through deregulation of the expression of aurora B protein and mitogen-activated protein kinase (pMSK1), which is overexpressed, decreasing phosphorylation at the H3S10 and H3S28 H3 positions; causing consequent cell cycle arrest in the G1 phase; 4- apoptosis activation through the intrinsic (through the mitochondria) and extrinsic (cell death ligands) pathways, activating caspases 8 and 9 (Lee et al., 2011; Gaviria et al., 2018).

Related to toxicity studies, both the ethanolic extract and the dichloromethane fraction from A. muricata seeds showed genotoxic activity against meristematic cells of Alluim cepa onion, the ethanolic extract was the most cytotoxic; in addition, they also caused a toxic effect against microcrustaceans. (Ferreira, 2022).

Aiming to understand the antitumor potential of different acetogenins isolated from A. Muricata, an in silico study was carried out to evaluate the physicochemical, pharmacokinetic and toxicological aspects, as well as molecular docking in caspase 3. Furthermore, the most promising acetogenins were selected and investigated whether there are studies on its antitumor activity.

2 Methodology

2.1 In silico evaluation

2.1.1 Selection of molecules

The molecules were selected based on a bibliographical review, according to some eligibility criteria: having been isolated from A. muricata, being classified as acetogenin, having its 2D structure and SMILE code available in the Pubchem database, in addition to responding to at least 3 criteria of Lipinski’s rules of five where molecular mass ≥ 500 Da, LogP ≤ 5; number of hydrogen bond acceptor groups ≤ 10; number of hydrogen bond donor groups ≤ 5 and topological polar surface area (TPSA) ≤ 140 Å. Then, the selected molecules were designed using the program Marvin Js online.

2.1.2 Prediction of physicochemical properties

To predict physicochemical properties, the online tool Mcule property calculator was used, considering Lipinski’s rule of five. Calculations of aspects related to the absorption and permeability of cell membranes (partition coefficient, lipophilicity; hydrophilicity) were made using the Molinspiration Online Property Calculation Toolkit website (www.molinspiration.com; Lipinski et al., 2001). In this way, the prediction of bioavailability was carried out considering the octanol-water partition coefficient (miLog P) less than or equal to 5; number of hydrogen bond acceptor groups (nHBA) less than or equal to 10; number of hydrogen bond donor groups (nHBD) less than or equal to 5; molecular mass (MM) less than or equal to 500 Da and the topological polar surface area (TPSA) less than or equal to 140 Å (Lipinski, 2004). For analysis, molecules that did not meet these criteria had low bioavailability (Silva, 2015).

2.1.3 Pharmacokinetics and toxicity

For pharmacokinetic and toxicity predictions, the PreADMET program (version 2.0, Copyright © 2005–2017) was used, which considers pharmacokinetic properties (A–absorption; D–Distribution; M–Metabolism/Biotransformation; E–Excretion) and evaluation of toxicity parameters (T–Toxicity; Preadmet, 2020).

The pharmacokinetic properties evaluated were: permeability, intestinal absorption, binding to plasma proteins, passage through the blood-brain barrier (BBB) and phase I and II metabolism through induction and inhibition of microsomal cytochrome P-450 (CYP–P450) enzymes, being considered the following parameters: Madin-Darby canine kidney (MDCK) and human colon carcinoma cell line-(Caco-2) (low permeability: <4 nm/s; medium permeability: 4–70 nm/s; high permeability: >70 nm/s; Yazdanian et al., 1998), skin permeability (high: <0.1, low: >0.1); Human Intestinal Absorption - HIA (poorly absorbed: 0%–20%; moderately absorbed: 20%–70%; well absorbed: 70%–100%; Yee, 1997), binding to plasma proteins (weak binding <90%; strong binding >90%) and passage through the BBB (crosses freely: >2; crosses moderately: 2.0–0.1; does not cross: <0.1; Ajay et al., 1999). In the analysis of results for phase I and II metabolism, inhibitory substances for 2 or more CYP enzymes and an inducing substance for at least 1 CYP enzyme were considered.

The toxicity analysis was carried out on Algae (Costa et al., 2008), Daphnia–through the Daphnia Test (Guilhermino et al., 2000) and Medaka and Minnow fish (Zuncker, 1985) according to the toxicity criteria (Table 1).

TABLE 1
www.frontiersin.org

TABLE 1. Toxicity criteria adopted for the study.

The mutagenicity risk was assessed by the Ames test with the following strains of Salmonella Typhimurium: TA100-10RLI and TA 100-NA mutation in His G46e plasmid pKM101 without S9; TA1535- 10RLI and TA1535-NA mutation in His G46 (Ames et al., 1975). The carcinogenic potential of the compounds was evaluated in rats and mice and referred to as: (+) carcinogenic and (–) non-carcinogenic.

To predict acute oral toxicity (oral LD50), the online software PROTOX II was used, considering the classification from I to VI (Table 2), according to ABNT NBR 14725–2, 2009 and as provided in Section 1 of the Annex IV of RDC No. 294, 2019.

TABLE 2
www.frontiersin.org

TABLE 2. Parameters adopted for classification of acute oral toxicity (oral LD50).

2.2 Molecular docking

For this trial, the effector Caspase 3 (cysteine-aspartic protease) receptors were the therapeutic targets for the action of acetogenins. This caspase was chosen due to its importance in the process of apoptosis, programmed cell death. This enzyme is a key zymogen in cellular apoptosis and is not activated until it is cleaved by initiator caspases during apoptotic flux.

To perform molecular docking, the caspase 3 structure was downloaded in two dimensions (2D) from the RCSB Protein Data Bank PDB website: 1 nms. Just as the 2D structures of acetogenins were downloaded in PDB format using the chemSketch program, Docking was carried out.

Molecular docking was carried out using the DockThor program to verify whether the investigated molecules had affinity for caspase 3 receptors, possible therapeutic targets. Docking was performed out using the grid center of the irreversible inhibitor already co-crystallographed with Caspase, to identify whether acetogenins bind to the same site. (Almeida, 2011) For this, some parameters were considered: the addition of hydrogen to the ligand, grid center: X = -9.091719; Y = -4.028625; Z = 24.242812; and the grid size: X:20Å, Y:20Å and Z:20 Å. The search was done by the standard algorithm with 1,000,000 reviews, population of 750 in 24 runs. This process was repeated for each acetogenin.

The results were observed using the Biovia Discovery studio visualizer 4.5 software, a program for visualizing the structure of proteins and small molecules and analyzing data. (Biovia Systemes, 2015).

2.3 Literature review

A bibliographic survey was carried out to correlate the results obtained and information in the literature on the antitumor activity of acetogenins, as well as the cell lines involved in this activity, for that, the following databases were searched: Google Scholar, Pubmed, Science direct, CAPES journal portal and SciELO.

3 Results and discussion

Several chemical studies on A. muricata have reported the isolation of acetogenins, with more than 100 different acetogenins isolated (Moghadamtousi et al., 2015). Because not all acetogenins have a 2D structure and SMILE code available in the Pubchem database, they were excluded from the present study, wherefore only 44 acetogenins were selected, whose structures and names are shown in Figure 1:

FIGURE 1
www.frontiersin.org

FIGURE 1. Acetogenins isolated from A. muricata.

1-Muricatocin B; 2-annocatalin; 3- annomontacin; 4-annomuricin A; 5-annonacin A; 6- annonacinone; 7- annopentocin A; 8-annoreticuin-9-one; 9-cis-corossolone; 10-cis-goniothalamicin; 11-cis-reticulatacin-10-one; 12-cis-solamin; 13-cis-uvariamicin IV; 14- cohibin A; 15-cohibin B; 16- cohibin C; 17-cohibin D; 18-epomusenin-A; 19-epomusenin-B; 20-epoxymurin A; 21- epoxymurin B; 22-gigantetronenin; 23- javoricin; 24-longifolicin; 25-montecristin; 26- muricadienin; 27-muricapentocin; 28-muricatacin; 29-muricatenol; 30-muricatocin A; 31-muricatocin B; 32-muricin A; 33-muricin B; 34-muricin C; 35- muricin E; 36-muricin F; 37- muricin G; 38- muricin H; 39- muricin I; 40-muridienin 3; 41-muridienin 4; 42-murisolin; 43-sabadelin; 44- xylomaticin.

3.1 In silico studies of acetogenins

Regarding physicochemical properties, from 44 compounds, only 28 met all the criteria of Lipinski’s rule of five. The remaining compounds violated two criteria established by this rule: log p > 5 and molecular mass>500 Da (Table 3). Although these molecules violated the rule, it does not mean that their therapeutic potential should be ignored (Lipinski, 2004; Chagas and Kallahan, 2022).

TABLE 3
www.frontiersin.org

TABLE 3. Values prediction ​​of physicochemical properties of acetogenins isolated from Annona muricata using Mcule property calculator.

The 43 molecules that have a greater molecular weight MM > 500Da may have difficulty to cross the plasma membrane. Each molecule’s topological polar surface area (TPSA) was less than 140 Å, which is justified by the lower bond between the hydrogen acceptors and donors, respectively ≤10 and ≤5. Therefore, the muricatacin (Figure 1 Compound 28; Table 3) would be the most promising for physicochemical properties, being able to express biological or pharmacological activity and be used orally in humans (Lipinski, 2004).

Lipinski rule: LogP- oil-water partition coefficient ≤5; TPSA: topological polarized surface area ≤140 Å; nHBA: Hydrogen bond acceptors ≤10; nHBD-number of hydrogen bond donor groups ≤5; MM-molecular mass≤ 500 g/mol (LIPINSKI, 2004).

In the pharmacokinetic prediction, all molecules presented high skin absorption, (Table 4), acetogenins have a lipophilic character (Barata et al., 2009), and can be used in dermocosmetics. However, permeabilities in MDCK (Madin-Darby canine kidney) cells and Caco2 (human colon carcinoma epithelial cells) were moderate, except for molecule 28 (Table 4). When assessing permeability in MDCK (Madin-Darby canine kidney) cells, the elimination rate of each molecule from the body was measured (Miranda et al., 2022). Molecule 28 showed high MDCK permeability (Table 4), maybe due to its lower molecular mass (Chen and Eugene, 2018). All molecules appear to have moderate permeability for Caco2 cells (Table 4), suggesting that the rate of intestinal absorption is moderate (Miranda et al., 2022) through the passive diffusion mechanism (Marinho and Thaisa, 2017; Chen and Eugene, 2018), which may be related to its high molecular mass. However, the studied acetogenins appear to have high absorption in the small intestine by human intestinal absorption analysis (Table 4). In a chemotherapy treatment, the oral route offers the benefits of convenience, the possibility of taking the medication at home without the need to go to clinics and hospitals (Vasconcelos et al., 2022). Also, acetogenins have an acidic character (Barata et al., 2009) and must have albumin affinity (Horta and Silvania Sofia dias, 2018).

TABLE 4
www.frontiersin.org

TABLE 4. Prediction of pharmacokinetic properties of acetogenins isolated from Annona muricata using Preadmet.

Despite the high molecular mass of compounds, the results suggest they cross the blood-brain barrier (BBB) freely (Table 4), probably due to their high lipid solubility (Chagas and Kallahan, 2022) (Table 3). It is noteworthy that different types of tumors develop in the central nervous system (CNS) and their growth can be rapid, making it important to develop drugs that can be administered orally and reach therapeutic concentrations in the CNS.

Another common point among acetogenins is the inhibitory potential of CYP2C19, 2C9 and 3A4, important in drug metabolism, mainly cyp3a4, which is the most abundant and its inhibition will be unfavorable for an antitumor drug. (Sukprasong, et al., 2021). The ability to inhibit CYP may be related to physicochemical properties as the greater the number of hydrogen acceptors, the greater the inhibitory potential. Furthermore, by inhibiting CYP 3A4 and others, the compound can interfere with the metabolism of different classes of drugs, requiring dose adjustment (De Lima and Cyntia, 2018). All are weakly metabolized by CYP3A4, and compounds 6, 8, 9, 11 and 27 also undergoing phase 1 metabolism (Table 4), giving rise to more polar metabolites, which are more easily excreted via the kidneys (Dolabela and Fani, 2018).

MDCK: Madin-Darby canine kidney; Caco2: Human Colon Adenocarcinoma Cells; PP-plasma protein; BBB: blood brain barrier; CYP: citrochrome P450; HIA: human intestinal absorption, S*: strongly; F *: freely; NO: not observed; W - weakly; H - high; L: low; M-medium.

In addition to the physicochemical and pharmacokinetic aspects, the toxic potential of acetogenins was evaluated using different models (Table 5). Only acetogenins 13 to 21, 25, 26 and 29 were not toxic to algae (Table 5), this model was used to predict acute oral toxicity in relation to death (Guilhermino et al., 2000). The same non-toxic compounds for algae were not toxic for Daphnia (Table 5), which can be used to predict acute and subchronic toxicities (Venturi, 2018).

TABLE 5
www.frontiersin.org

TABLE 5. Toxicity of acetogenins isolated from Annona muricata.

All compounds were toxic to fish (Table 5), suggesting acute and subchronic toxicity, as well as changes in different organs (Bauer, 2017). Through the results obtained in Algae, Daphnia and fish, we suggest the compounds with the greatest acute and subchronic toxic potential are: 13, 14, 15, 16, 17, 18, 19, 20, 21, 25, 26 and 29. However, it is important to consider the mutagenic and carcinogenic potential. Except for compound 28, all compounds showed carcinogenic potential for mice and compounds 11, 12, 18 to 21, 25, 26, 40, 41, 43 showed carcinogenic potential in rats (Table 5).

Only compounds 28 and 39 showed mutagenic potential for Salmonella typhimurium (TA 100-10RLI; Table 5), this strain is sensitive for the detection of mutagenicity and contain a pKM101 plasmid and mutation in His G46 (Dayse et al., 2020). Analyzing the toxicity results (Table 5), we suggest the compounds with the highest toxic potential are: 18, 21, 25 and 26.

In acute oral toxicity, only the molecule 28 (muricatacin) presented LD50 > 5,000, being classified in category VI, it appears to cause no harm and does not present immunotoxicity or cytotoxic effects. Molecules 9,11,12,13,24,26,38,40 and 41 appear unlikely to cause harm, but all have immunotoxic or genotoxic potential. On the other hand, molecules 7, 22, 36 and 37 appear to be extremely toxic, presenting cytotoxic and immunotoxic potential (Table 6). Based on these results, we can suggest that the least promising molecules are: 7, 18, 21, 22, 25, 26, 36 and 37 (Tables 5, Table 6).

TABLE 6
www.frontiersin.org

TABLE 6. Acute oral toxicity values ​​(oral LD50) of acetogenins isolated from Annona muricata using PROTOX II online software.

3.2 Molecular docking simulation

To select the molecules for docking, physicochemical and pharmacokinetic aspects were considered, in these criteria the most promising was molecule 28 (Tables 3, Table 4). Furthermore, the results in toxicity prediction studies were evaluated, in which some molecules were not toxic to Algae, Daphnia, nor mutagenic nor carcinogenic to rats (13, 14, 15, 16 and 17). Regarding acute oral toxicity, these compounds appear to be low toxic (14, 15, 16 and 17) or unlikely to cause harm (13). In summary, the most promising molecule is 28, but the other molecules were included for the next stage of the study (13–17).

For validating the docking protocol applied in the next stage, redocking was carried out, it showed that the best fit of the co-crystallized ligand would be at the value of Root Mean Square deviation - RMSD = 2.09Å (Figure 2), and according to the literature, the prediction of the binding mode using docking must present an RMSD value < 2.0 Å, when superimposed on the crystallographic pose of the ligand (Ramos and Ryan, 2019; Araújo and Pedro, 2020; Mascarenhas and Mercia, 2021). However, the reference values accepted in the literature are RMSD up to 2 Å for rigid structures and up to 2.5 Å for flexible structures (Xiao et al., 2018; Santos et al., 2020; Dardenne, 2021; Altê Alves, 2022). Therefore, this result shows the model was suitable for predicting the docking profile between the ligand molecule and the active site of the selected target.

FIGURE 2
www.frontiersin.org

FIGURE 2. Structure obtained by redocking for the co-crystallographed ligand (irreversible inhibitor) with 1mns target (caspase 3).

For the formation of the enzyme complex with each acetogenin, cis-uvariamicin IV; cohibin A; cohibin B; cohibin C; cohibin D; muricatacin, the binding free energies were respectively -8.436; -9.151; -9.339; -8.715; -9.275 and -7.632 kcal/mol. The results demonstrated ligands are capable of interacting with the enzyme in a favorable way and cohibin B would have the highest affinity with a minimum score of -9.339 (kcal/mol). As for total energy, it was higher with compound 13 (-48,007 kcal/mol) and lower for 28 (-6,123 kcal/mol). Van der Waals energy was higher for molecule 14 (-30,613 kcal/mol) and lower for 28 (-15,609 kcal/mol). The interaction and electrostatistic energies were respectively higher for the numbers 17 (-15.830 kcal/mol) and 13 (-23.064 kcal/mol) and lower for the compounds 28 (-27,981 kcal/mol) and 14 (-7,904 kcal/mol). Compound 13 cstood out most with its total and electrostatistic energies, important energy for the formation of chemical bonds (Table 7). Although compound 14 has stronger bonds, such as the pi-alkyl type, they require more energy to form bonds, which justifies their lower value of electrostatic energy, due to the different nature of the bond.

TABLE 7
www.frontiersin.org

TABLE 7. Binding affinity between acetogenins isolated from Annona muricata and caspase 3 obtained from molecular docking.

The results suggest that Cis-uvariamicin IV performs hydrophobic pi-alkyl interaction with HIS 88 and hydrogen bonds on THR 29 and SER 172. Cohibin compounds A, B, C and D perform hydrophobic pi-alkyl bonds with TRP 169, HIS 88, PHE 219, MET 28, PHE 213 and PHE 215; conventional hydrogen bonds with SER172, TYR 167, PHE 213, THR 29, ARG 170 and PHE 215; covalent bonds with PHE 213. Muricatacin makes hydrophobic alkyl bonds in ME28 and pi-alkyl in HIS 88, and TRP 169; conventional hydrogen bonds in ARG 170 and TRP 177 Even though both bonds are hydrophobic, the alkyl-type bond tends to carry out a nucleophilic interaction that favors the interaction between the ligand and the receptor (Figure 3). The distances were relatively smaller between hydrogen bonds (between 1 and 2.5 Å) and covalent bonds (between 2 and 3.5 Å) and larger for hydrophobic bonds (4–5.5 Å).

FIGURE 3
www.frontiersin.org

FIGURE 3. (Continued) Representation of 2D interactions of cis-uvariamicin IV (13), cohibin A (14), cohibin B (15), cohibin C (16), cohibin D (17) and muricatacin (28) in the active site of the caspase 3 enzyme. Image generated with Discovery Studio 4.5 Visualizer.

The approach of hydrophobic surfaces contributes to the destruction of the organized structure of water, with interactions between ligands and receptors due to the increase in entropy associated with the disorganization of the system. Hydrogen bonds are important interactions for maintaining a protein’s structure, and drugs that interact through covalent bonds inactivate the receptor site and consequently the action of enzymes (Barreiro and Fraga, 2015). Table 8 also shows compound 13 establishes four hydrogen bonds, with relatively smaller distances with the amino acids and consequently stronger bonds.

TABLE 8
www.frontiersin.org

TABLE 8. Interactions between ligands and amino acid residues.

3.3 Antitumor and cytotoxicity activity of muricatacin on literature

The last stage of this study was to correlate our results with studies evaluating the antitumor activity of the selected acetogenins (13–17 and 28), however, we observed the scarcity of in vitro and in vivo studies. One factor that may limit these studies is the chemical aspects, because the plant may contain several acetogenins and some may be isomers (Sun, et al., 2014), which makes it difficult to isolate compounds.

Another limiting factor may be the low concentrations of acetogenins in the extracts, requiring a large amount of plant material to isolate enough for compounds identification and tests. A very positive point is the storage location of acetogenins in A. muricata, the majority are stored in the seeds (3, 5, 6, 8, 10, 17, 22, 23, 24, 28, 29, 32, 33, 34, 35, 36, 37, 38, 39, 42 and 44) (Moghadamtousi, et al., 2015), as the fruit pulp is widely used in the food industry and the seeds discarded, these seeds can be reused and not discarded in the environment.

Despite being very promising molecules, there is still a lack of studies about them, and the available ones were carried out with fractions from extracts containing acetogenins. A study carried out with the chloroform fraction of the methanolic extract of Annona muricata seeds (CMAM) showed activity against triple-negative breast cancer (TNBC), both in vitro and in vivo (Kariyil and Bibu, 2021). We emphasize that TNBC treatment is still a challenge, as it usually does not respond satisfactorily to available therapy (Garrido Castro et al., 2019; So, 2022).

The anticancer effect of CMAM may be related to ROS dependent caspase activated mitochondria-mediated apoptosis and S-phase arrest in TNBCs. Furthermore, the results indicated the presence of annonaceous acetogenins, especially muricatacin, that may be contributing to the anticancer effect of CMAM (Kariyil and Bibu, 2021).

In one study, the MTT (3-(4,5-dimethylthiazol-2yl)-2,5-diphenyl tetrazoline bromide) cell viability assay was applied to evaluate the cytotoxic potential of CMAM in TNBC cells (MDA- MB-231, BT-549 and 4T1) and the normal breast epithelial cells, MCF10A. The results showed the plant fraction was cytotoxic to TNBC cells and the level of damage was concentration dependent. The IC50 values were 2.5 ± 0.14, 4.8 ± 0.3 and 4.5 ± 0.16 μg/mL, respectively, 4TI, BT-549 and MDA-MD-231. However, for normal mammary epithelial cell line, MCF10A, CMAM showed lower cytotoxicity with IC50 value of 81.1 ± 2.28 μg/mL, demonstrating its differential cytotoxicity towards cancerous and normal cells (Kariyil and Bibu, 2021).

Aiming to evaluate the selectivity of CMAM antitumor activity, the selectivity index (IC50 4/IC50 1, 2 or 3) was established, and the MDA-MB-231 lineage was the most selective with SI = 32 (Table 9).

TABLE 9
www.frontiersin.org

TABLE 9. Cytotoxicity and selectivity of the chloroform fraction of the methanolic extract of Annona muricata seeds in TNBC cell lines.

Despite their high selectivity for cancer cells, some acetogenins have neurotoxic potential, leading to neurodegenerative diseases such as Alzheimer’s and Parkinson’s (Lannuzel et al., 2007; Chuzeville, 2015; Das and Nascimento, 2022). Acetogenins can promote the intracellular aggregation of a tau protein known as tauopathy, leading to the redistribution and accumulation of tau in somatodendrites (Rottscholl, et al., 2016).

In vitro and in vivo studies indicated anonacin can cause the development of signs associated with tauopathies, such as tau hyperphosphorylation, retrograde mitochondrial transport, tau redistribution from the axon to the neuronal body, and cell death (Lannuzel et al., 2007; Guillopé, et al., 2011; Bruch, et al., 2014; Rottscholl, et al., 2016). Different studies suggest that anonacin (mono-tetrahydrofuran) causes a reduction in ATP in neurons, through the inhibition of mitochondrial complex I (Lannuzel et al., 2003; Champy et al., 2004; Escobar-khondiker et al., 2007; Potts, et al., 2012; Rottscholl et al., 2016).

4 Conclusion

The in silico study allowed to observe that almost all acetogenins are fat-soluble compounds, cross the blood-brain barrier and inhibit CYP2C19, CYP2C9 and CYP3A4; in addition to having acute and subchronic toxic potential. Therefore, muricatacin presented different characteristics with greater bioavailability and less acute toxicity. Docking showed acetogenins perform hydrophobic interactions, hydrogen bonds and covalent bonds to maintain both affinity and stability between caspase 3 and the selected ligands, and the cis-uvariamicin IV was the most promising compound for this parameter. Furthermore, the literature has shown the cytotoxic activity of acetogenins. The chloroform fraction from methanolic extract of Annona muricata seeds presented activity against triple-negative breast cancer, both in vitro and in vivo, which would be associated with the presence of acetogenins, mainly muricatacin. However, muricatacin and cis-uvariamycin IV would probably be the most promising molecules in the present study, with cis-uvariamycin IV being safer as a possible adjuvant for antitumor drugs.

Data availability statement

The raw data supporting the conclusion of this article will be made available by the authors, without undue reservation.

Author contributions

HF: Formal Analysis, Investigation, Methodology, Software, Writing–original draft, Writing–review and editing. CS: Data curation, Investigation, Writing–review and editing. GG: Data curation, Investigation, Writing–review and editing. MC: Supervision, Writing–review and editing. MD: Supervision, Writing–review and editing.

Funding

The author(s) declare financial support was received for the research, authorship, and/or publication of this article. Dean of Research and Postgraduate Studies at the State University of Pará, for payment for the processing and publication of this article - Notice 02/2023 - PAPQ/PROPESP.

Conflict of interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

References

Adão, R., de Keulenaer, G., Leite-Moreira, A., and Brás-Silva, C. (2013). Cardiotoxicity associated with cancer therapy: pathophysiology and prevention strategies. Rev. Port. Cardiol. 32 (5), 395–409. doi:10.1016/j.repc.2012.11.002

PubMed Abstract | CrossRef Full Text | Google Scholar

Ajay, A., Bermis, G. W., and Murkco, M. A. (1999). Designing libraries with CNS activity. J. Med. Chem. 42 (24), 4942–4951. doi:10.1021/jm990017w

PubMed Abstract | CrossRef Full Text | Google Scholar

ALALI Feras, Q., Zhang, Y., and McLaughlin, J. L. (1998). Unusual bioactive annonaceous acetogenins from Goniothalamus giganteus. Tetrahedron 54 (22), 5833–5844. doi:10.1016/s0040-4020(98)00286-5

CrossRef Full Text | Google Scholar

Almeida, (2011). Diogo Marinho. Dockthor: implementação, aprimoramento e validação de um programa de atracamento receptor-ligante. Dissertation (Doctorate degree). National Scientific Computing Laboratory.

Google Scholar

Alonso-Castro, , Josabad, A., Salazar-Olivo, L. A., Gomez-Sanchez, M., Dominguez, F., and Garcia-Carranca, A. (2011). Mexican medicinal plants used for cancer treatment: pharmacological, phytochemical and ethnobotanical studies. J. Ethnopharmacol. 133 (3), 945–972. doi:10.1016/j.jep.2010.11.055

PubMed Abstract | CrossRef Full Text | Google Scholar

Altê Alves, G. (2022). Planejamento computacional de inibidores da glutationa peroxidase 4 - GPX4. Dissertation (Master's degree) - federal University of Santa Catarina, Center for Biological Sciences, Postgraduate Program in Biochemistry, Florianópolis. Avaliable in: https://repositorio.ufsc.br/handle/123456789/236107.

Google Scholar

Amanda, THAIS, da Silva Neves, N. C., and Brito da Silva, I. (2022). Uso popular de plantas medicinais no estado de Pernambuco: uma revisão integrativa. Amaz. Sci. Health 10 (2), 2–20. doi:10.18606/2318-1419/amazonia.sci.health.v10n2p2-20

CrossRef Full Text | Google Scholar

Ames, B. N., Mccann, J., and Yamasaki, E. (1975). Methods for detecting carcinogens and mutagens with the Salmonella/mammalian-microsome mutagenicity test. Mutat. Research/environmental Mutagen. Relat. Subj. 31 (6), 347–363. doi:10.1016/0165-1161(75)90046-1

PubMed Abstract | CrossRef Full Text | Google Scholar

ARAÚJO Pedro, H. F., da Cruz, J. N., Silva, S. G., Ferreira, E. F. B., de Lima, L. R., et al. (2020). Identification of potential COX-2 inhibitors for the treatment of inflammatory diseases using molecular modeling approaches. Molecules 25 (18), 4183. doi:10.3390/molecules25184183

PubMed Abstract | CrossRef Full Text | Google Scholar

Atawodi Eneojo, S. (2011). “Nigerian foodstuffs with prostate cancer chemopreventive polyphenols,” in Infectious agents and cancer (London, United Kingdom: BioMed Central), 1–4.

PubMed Abstract | CrossRef Full Text | Google Scholar

Badrie, N., and Schauss, A. G. (2010). “Soursop (Annona muricata L.): composition, nutritional value, medicinal uses, and toxicology,” in Bioactive foods in promoting health (Cambridge, Massachusetts, United States: Academic Press), 621–643.

CrossRef Full Text | Google Scholar

Barata, L. E. S., Alencar, A. A. J., Tascone, M., and Tamashiro, J. (2009). Plantas medicinais brasileiras. IV. Annona muricata L. (Graviola). Rev. Fitos 4, 132–138. 01. doi:10.32712/2446-4775.2009.94

CrossRef Full Text | Google Scholar

Barreiro, E. J., and Fraga, C. A. M. (2015). Química Medicinal: as Bases Moleculares da Ação dos Fármacos. 3a Edição. Porto Alegre: ArtMed Editora Ltda.

Google Scholar

Bauer, L. H. (2017). Sensibilidade do método de ensaios de toxicidade aguda com embriões de peixes na avaliação de efluente de indústria metalmecânica. Dissertation (Bachelor's degree). Porto Alegre: Federal University of Rio Grande do Sul. Institute of Biosciences. Biological Sciences Course.

Google Scholar

Beyra, Á., et al. (2004). “Estudios etnobotánicos sobre plantas medicinales en la provincia de Camagüey (Cuba),” in Anales del jardín botánico de Madrid, 185–204.

Google Scholar

Biovia Systèmes, D. (2015). Discovery studio visualizer, release 4.5. San Diego, CA, USA: Dassault Systèmes.

Google Scholar

Boyom, , Fekam, F., Yamthe, L. R. T., Mfopa, A. N., Kemgne, E. M., Mbacham, W. F., et al. (2011). Potent antiplasmodial extracts from Cameroonian Annonaceae. J. Ethnopharmacol. 134 (3), 717–724. doi:10.1016/j.jep.2011.01.020

PubMed Abstract | CrossRef Full Text | Google Scholar

Bruch, J., Xu, H., De Andrade, A., and Höglinger, G. (2014). Mitochondrial complex 1 inhibition increases 4-repeat isoform tau by SRSF2 upregulation. PloS one 9 (11), e113070. doi:10.1371/journal.pone.0113070

PubMed Abstract | CrossRef Full Text | Google Scholar

Bussmann, , Rainer, W., Glenn, A., Sharon, D., Chait, G., Díaz, D., et al. (2010). Minimum inhibitory concentrations of medicinal plants used in Northern Peru as antibacterial remedies. J. Ethnopharmacol. 132 (1), 101–108. doi:10.1016/j.jep.2010.07.048

PubMed Abstract | CrossRef Full Text | Google Scholar

Carmen Zafra-Polo, M., Figadere, B., Gallardo, T., Tormo, J., and Cortes, D. (1998). Natural acetogenins from annonaceae, synthesis and mechanisms of action. Phytochemistry 48, 1087–1117. doi:10.1016/s0031-9422(97)00917-5

CrossRef Full Text | Google Scholar

Chagas, C., Kallahan, S., Martins, H. P. d. S., and Dolabela, M. F. (2022). Estudo in silico de compostos fenólicos isolados de Inga laurina. Res. Soc. Dev. 11 (2), e24511225592. doi:10.33448/rsd-v11i2.25592

CrossRef Full Text | Google Scholar

Champy, P., Höglinger, G. U., Féger, J., Gleye, C., Hocquemiller, R., Laurens, A., et al. (2004). Annonacin, a lipophilic inhibitor of mitochondrial complex I, induces nigral and striatal neurodegeneration in rats: possible relevance for atypical parkinsonism in Guadeloupe. J. Neurochem. 88 (1), 63–69. doi:10.1046/j.1471-4159.2003.02138.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Chen, , Plise, E., Chen, B., Liu, L., Cheong, J., et al. (2018). Evaluating the utility of canine Mdr1 knockout madin-darby canine kidney I cells in permeability screening and efflux substrate determination. Mol. Pharm. 15 (11), 5103–5113. doi:10.1021/acs.molpharmaceut.8b00688

PubMed Abstract | CrossRef Full Text | Google Scholar

Chuzeville, M. (2015). Caroleé. Identificación de la modulación de la proteína tau en el tratamiento in vitro con acetogeninas de guanábana (Annona muricata). Dissertation (Master's degree). Xalapa, México: Universidad Veracruzana. Avaliable in: https://cdigital.uv.mx/handle/123456789/42611?show=full.

Google Scholar

Coria-Téllez, , Yahia, E. M., Obledo-Vázquez, E. N., et al. (2018). Annona muricata: a comprehensive review on its traditional medicinal uses, phytochemicals, pharmacological activities, mechanisms of action and toxicity. Arabian J. Chem. 11 (5), 662–691. doi:10.1016/j.arabjc.2016.01.004

CrossRef Full Text | Google Scholar

Costa, C. R., Olivi, P., Botta, C. M. R., and Espindola, E. L. G. (2008). A toxicidade em ambientes aquáticos: discussão e métodos de avaliação. Quim Nova 31 (7), 1820–1830. doi:10.1590/s0100-40422008000700038

CrossRef Full Text | Google Scholar

Dardenne, L. E. (2021). Practical short course on ReceptorLigante docking and virtual screening methods from the X school of molecular modeling in biological systems. LNCC/MCTI.

Google Scholar

Das, CHAGAS LIMA, Nascimento, N., Allahdadi, K. J., França, L. S. d. A., and Pinto, L. C. (2022). Acetogenins from Annonaceae plants: potent antitumor and neurotoxic compounds. PharmaNutrition 20, 100295. doi:10.1016/j.phanu.2022.100295

CrossRef Full Text | Google Scholar

Dayse, L. do N. B., Martins, H. P. d. S., Tomaz, J. M. O., Ferreira, G. G., Ramos, H. O., Percário, S., et al. (2020). Estudo in silico de diterpenos isolados de Portulaca pilosa L e sua correlação com estudos etnobotânicos. Res. Soc. Dev. 9 (11), e99491110635. doi:10.33448/rsd-v9i11.10635

CrossRef Full Text | Google Scholar

De, Andrade, and Abreu, A. (2019). Guia prático para produção de óleo de graviolaa partir de rejeitos do beneficiamento da fruta [Practical guide to soursop oil production from fruit processing waste. Avaliable in: https://educapes.capes.gov.br/bitstream/capes/560743/2/Comunicado_T%C3%A9cnico_N_1.pdf.

Google Scholar

De Lima Braz, C., de Figueiredo, T. P., Barroso, S. C. C., and Reis, A. M. M. (2018). Medicamentos com atividade sobre o citocromo P450 utilizados por idosos em domicílio. Rev. Médica Minas Gerais. doi:10.5935/2238-3182.20180069

CrossRef Full Text | Google Scholar

Deminice, R. (2022). Exercício físico para o tratamento do câncer: evidências científicas eo contexto brasileiro. J. Phys. Educ. 33 (1). doi:10.4025/jphyseduc.v33i1.3301

CrossRef Full Text | Google Scholar

DOLABELA Fâni, M., Ohashi, L. H., Bastos, M. L. C., Da Silva, M. C. M., and Vale, V. V. (2018). Estudo in silico das atividades de triterpenos e iridoides isolados de Himatanthus articulatus (Vahl) Woodson. Rio J. 12 (3), 227–242. doi:10.17648/2446-4775.2018.602

CrossRef Full Text | Google Scholar

Escobar-Khondiker, M., Muriel, M. P., Champy, P., Bach, A., Depienne, C., et al. (2007). Annonacin, a natural mitochondrial complex I inhibitor, causes tau pathology in cultured neurons. J. Neurosci. 27 (29), 7827–7837. doi:10.1523/jneurosci.1644-07.2007

PubMed Abstract | CrossRef Full Text | Google Scholar

Ferreira, G., Quaresma, A. C. S., Brandão, D. L. d. N., Marinho, A. M. d. R., Siqueira, J. E. d. S., Correa, K. L., et al. (2022). Evaluation of genotoxicity and toxicity of Annona muricata L. Seeds and in silico. Molecules 28 (1), 231. doi:10.3390/molecules28010231

PubMed Abstract | CrossRef Full Text | Google Scholar

Garrido-Castro, A. C., Lin, N. U., and Polyak, K. (2019). Insights into molecular classifications of triple-negative breast cancer: improving patient selection for treatment. Cancer Discov. 9 (2), 176–198. doi:10.1158/2159-8290.cd-18-1177

PubMed Abstract | CrossRef Full Text | Google Scholar

Gaviria, CALLE, Marcela, M., Posada, ARIAS, Mira, HERNÁNDEZ, and Juliana, (2018). Acetogeninas, alternativa en el tratamiento de cáncer en caninos. CES Med. Veterinaria Zootec. 13 (2), 157–172. doi:10.21615/cesmvz.13.2.5

CrossRef Full Text | Google Scholar

Guilhermino, L., Diamantino, T., Silva, M. C., and Soares, A. M. M. (2000). Acute toxicity test with Daphnia magna: Na alternative to mammals in the Prescreening of Chemical Toxicity. Ecotoxicol. Environ. Saf. 46 (3), 57–362. doi:10.1006/eesa.2000.1916

PubMed Abstract | CrossRef Full Text | Google Scholar

Guillopé, R., Escobar-Khondiker, M., Guérineau, V., Laprévote, O., Höglinger, G. U., and Champy, P. (2011). Kaurenoic acid from pulp of Annona cherimolia in regard to annonaceae-induced parkinsonism. Phytotherapy Res. 25 (12), 1861–1864. doi:10.1002/ptr.3508

CrossRef Full Text | Google Scholar

Haiat, , Waizel, S., and Bucay, J. W. (2009). “Algunas plantas utilizadas en México para el tratamiento del asma,” in Anales de Otorrinolaringología mexicana, 145–171.

Google Scholar

Horta, S. S. D. (2018). Interações medicamento-alimento: caracterização do conhecimento dos profissionais de saúde. Dissertation (Master’s degree). Lisbon: Polytechnic Institute of Lisbon, Lisbon School of Health Technology. Available at: https://repositorio.ipl.pt/handle/10400.21/9338.

Google Scholar

Icb, (2019). Produtos químicos - informações sobre segurança, saúde e meio ambiente Parte 2: sistema de classificação de perigo. https://ww3.icb.usp.br/wp-content/uploads/2019/11/Parte2_NBR_14725-2-2009.pdf.

Google Scholar

INTITUTO NACIONAL DE CÂNCER (INCA) (2022). Estima 704 mil casos de câncer por ano no Brasil até 2025. Avalieble in: https://www.gov.br/inca/pt-br/assuntos/noticias/2022/inca-estima-704-mil-casos-de-cancer-por-ano-no-brasil-ate-2025.2022.

Google Scholar

Jardim, A. K. R., and Parente, C. A. (2021). Uso da graviola como coadjuvante no tratamento do câncer de próstata. Dissertation (Bachelor's degree). São Paulo: Ânima.

Google Scholar

Joyeux, M., Mortier, F., and Fleurentin, J. (1995). Screening of antiradical, antilipoperoxidant and hepatoprotective effects of nine plant extracts used in Caribbean folk medicine. Phytotherapy Res. 9 (3), 228–230. doi:10.1002/ptr.2650090316

CrossRef Full Text | Google Scholar

KARIYIL Bibu, J., Gopalakrishnan, A., and George, A. J. (2021). Chloroform fraction of methanolic extract of seeds of Annona muricata induce S phase arrest and ROS dependent caspase activated mitochondria-mediated apoptosis in triple-negative breast cancer. Anti-Cancer Agents Med. Chem. Former. Curr. Med. Chemistry-Anti-Cancer Agents) 21 (10), 1250–1265. doi:10.2174/1871520620666200918101448

CrossRef Full Text | Google Scholar

Kossouoh, C., Moudachirou, M., Adjakidje, V., Chalchat, J. C., and Figuérédo, G. (2007). Essential oil chemical composition ofAnnona muricataL. Leaves from Benin. J. Essent. Oil Res. 19 (4), 307–309. doi:10.1080/10412905.2007.9699288

CrossRef Full Text | Google Scholar

Lannuzel, A., Hoglinger, G. U., Verhaeghe, S., Gire, L., Belson, S., Escobar-Khondiker, M., et al. (2007). Atypical parkinsonism in Guadeloupe: a common risk factor for two closely related phenotypes? Brain 130 (3), 816–827. doi:10.1093/brain/awl347

PubMed Abstract | CrossRef Full Text | Google Scholar

Lannuzel, A., Michel, P., Höglinger, G., Champy, P., Jousset, A., Medja, F., et al. (2003). The mitochondrial complex I inhibitor annonacin is toxic to mesencephalic dopaminergic neurons by impairment of energy metabolism. Neuroscience 121 (2), 287–296. doi:10.1016/s0306-4522(03)00441-x

PubMed Abstract | CrossRef Full Text | Google Scholar

Lee, C.-C., Lin, Y. H., Chang, W. H., Lin, P. C., Wu, Y. C., and Chang, J. G. (2011). Squamocin modulates histone H3 phosphorylation levels and induces G1 phase arrest and apoptosis in cancer cells. BMC cancer 11 (1), 58–59. doi:10.1186/1471-2407-11-58

PubMed Abstract | CrossRef Full Text | Google Scholar

Lipinski, C. A. (2004). Lead-and drug-like compounds: the rule-of-five revolutio. Drug DiscoToday Technol. 1, 337–342. doi:10.1016/j.ddtec.2004.11.007

PubMed Abstract | CrossRef Full Text | Google Scholar

Lipinski, C. A., Dominy, F. B. W., and Feeney, P. J. (2001). Experimental and computational approaches to estimate solubility and permeability in drug discovery and development settings 1PII of original article: S0169-409X(96)00423-1. The article was originally published in Advanced Drug Delivery Reviews 23 (1997) 3–25. 1. Advenced Drug Deliv. Rev. 46 (1-3), 3–26. doi:10.1016/s0169-409x(00)00129-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Longuefosse, J.-L., and Nossin, E. (1996). Medical ethnobotany survey in Martinique. J. Ethnopharmacol. 53 (3), 117–142. doi:10.1016/0378-8741(96)01425-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Lorusso, D., Bria, E., Costantini, A., Di Maio, M., Rosti, G., and Mancuso, A. (2017). Patients’ perception of chemotherapy side effects: expectations, doctor–patient communication and impact on quality of life–An Italian survey. Eur. J. cancer care 26 (2), e12618. doi:10.1111/ecc.12618

PubMed Abstract | CrossRef Full Text | Google Scholar

Machado, G. C. (2019). As acetogeninas da Anonna muricata e suas propriedades anticancerigenas. Dissertation (Bachelor's degree) - faculty of Nutrition. Available in: https://bdm.ufpa.br:8443/jspui/handle/prefix/5143.

Google Scholar

Mansoori, B., Mohammadi, A., Davudian, S., Shirjang, S., and Baradaran, B. (2017). The different mechanisms of cancer drug resistance: a brief review. Adv. Pharm. Bull. 7 (3), 339–348. doi:10.15171/apb.2017.041

PubMed Abstract | CrossRef Full Text | Google Scholar

Marinho, D., de Silva, M. M. C., and dos Reis Serra, C. H. (2017). In situ intestinal perfusion in rodents: future perspectives for application on absorption studies and classification of drugs. Mini Rev. Med. Chem. 17 (9), 746–757. doi:10.2174/1389557516666160921145613

PubMed Abstract | CrossRef Full Text | Google Scholar

Mascarenhas, A., Mércia, S., de Araujo Neto, M. F., Mendes, G. O., da Cruz, J. N., dos Santos, C. B. R., et al. (2021). Pharmacophore-based virtual screening and molecular docking to identify promising dual inhibitors of human acetylcholinesterase and butyrylcholinesterase. J. Biomol. Struct. Dyn. 39 (16), 6021–6030. doi:10.1080/07391102.2020.1796791

PubMed Abstract | CrossRef Full Text | Google Scholar

Maul, A. A., Coutinho, C. B. M., Brazzach, M. L., and Silva, C. M. (2000). Extração supercrítica das sementes de graviola. XVI Simpósio de Plantas Medicinais do Brasil. UFPE, Recife, An. do Congr., 174.

Google Scholar

MCULE (2020). Prosperty calculador. Recuperado de. https://mcule.com/apps/property-calculator/.

Google Scholar

Miranda, D. A. SILVA, Cristina, C., Victor Alexandre Cardoso, SALAZAR, and Gomes, A. (2022). Predição in silico da atividade antiviral e avaliação de características farmacocinéticas e toxicológicas de compostos presentes no óleo essencial de Petiveria alliacea L. Rev. Casos Consult. 13 (1), e30705.

Google Scholar

Moghadamtousi, S. Z., Fadaeinasab, M., Nikzad, S., Mohan, G., Ali, H., and Kadir, H. (2015). Annona muricata (Annonaceae): a review of its traditional uses, isolated acetogenins and biological activities. Int. J. Mol. Sci. 16 (7), 15625–15658. doi:10.3390/ijms160715625

PubMed Abstract | CrossRef Full Text | Google Scholar

Monigatti, M., Bussmann, R. W., and Weckerle, C. S. (2013). Medicinal plant use in two Andean communities located at different altitudes in the Bolívar Province, Peru. J. Ethnopharmacol. 145 (2), 450–464. doi:10.1016/j.jep.2012.10.066

PubMed Abstract | CrossRef Full Text | Google Scholar

Moraes, I., Vieira, M. de, Schmidt, F. L., Canuto, K. M., Zocolo, G. J., de Brito, E. S., et al. (2016). UPLC–QTOF–MS and NMR analyses of graviola (Annona muricata) leaves. Rev. Bras. Farmacogn. 26, 174–179. doi:10.1016/j.bjp.2015.12.001

CrossRef Full Text | Google Scholar

Mutakin, M., Fauziati, R., Fadhilah, F. N., Zuhrotun, A., Amalia, R., and Hadisaputri, Y. E. (2022). Pharmacological activities of soursop (Annona muricata Lin.). Molecules 27 (4), 1201. doi:10.3390/molecules27041201

PubMed Abstract | CrossRef Full Text | Google Scholar

Nakanishi, S., Abe, M., Yamamoto, S., Murai, M., and Miyoshi, H. (2011). Bis-THF motif of acetogenin binds to the third matrix-side loop of ND1 subunit in mitochondrial NADH-ubiquinone oxidoreductase. Biochimica Biophysica Acta (BBA)-Bioenergetics 1807 (9), 1170–1176. doi:10.1016/j.bbabio.2011.05.012

PubMed Abstract | CrossRef Full Text | Google Scholar

Nantajit, D., Lin, D., and Li, J. J. (2015). The network of epithelial–mesenchymal transition: potential new targets for tumor resistance. J. cancer Res. Clin. Oncol. 141, 1697–1713. doi:10.1007/s00432-014-1840-y

PubMed Abstract | CrossRef Full Text | Google Scholar

Nguyen-Pouplin, , Julie, N., Tran, H., Phan, T. A., Dolecek, C., Farrar, J., et al. (2007). Antimalarial and cytotoxic activities of ethnopharmacologically selected medicinal plants from South Vietnam. J. Ethnopharmacol. 109 (3), 417–427. doi:10.1016/j.jep.2006.08.011

PubMed Abstract | CrossRef Full Text | Google Scholar

Oliveira, R. B. de, and Alves, R. J. (2002). Agentes antineoplásicos biorredutíveis: uma nova alternativa para o tratamento de tumores sólidos. Quimica Nova 25, 976–984. doi:10.1590/s0100-40422002000600015

CrossRef Full Text | Google Scholar

Pearce, A., Haas, M., Viney, R., Pearson, S. A., Haywood, P., Brown, C., et al. (2017). Incidence and severity of self-reported chemotherapy side effects in routine care: a prospective cohort study. PloS one 12 (10), e0184360. doi:10.1371/journal.pone.0184360

PubMed Abstract | CrossRef Full Text | Google Scholar

Potts, L. F., Luzzio, F. A., Smith, S. C., Hetman, M., Champy, P., and Litvan, I. (2012). Annonacin in Asimina triloba fruit: implication for neurotoxicity. Neurotoxicology 33 (1), 53–58. doi:10.1016/j.neuro.2011.10.009

PubMed Abstract | CrossRef Full Text | Google Scholar

Preadmet, (2020). ADME TOX calculatioRecuperado de. http://preadmet.bmdrc.kr.Recuperadode.

Google Scholar

Ramos, D. A. SILVA, da Costa, G., Rodrigues, A., Rabelo, É., et al. (2019). Identification of potential inhibitors from pyriproxyfen with insecticidal activity by virtual screening. Pharmaceuticals 12 (1), 20. doi:10.3390/ph12010020

PubMed Abstract | CrossRef Full Text | Google Scholar

Ross, I. A. (2010). “Medicinal plants of the world,” in Chemical constituents, traditional and modern medicinal (Totowa: Humana Press), 133–142.

Google Scholar

Rottscholl, R., Haegele, M., Jainsch, B., Xu, H., Respondek, G., Höllerhage, M., et al. (2016). Chronic consumption of Annona muricata juice triggers and aggravates cerebral tau phosphorylation in wild-type and <scp>MAPT</scp> transgenic mice. J. Neurochem. 139 (4), 624–639. doi:10.1111/jnc.13835

PubMed Abstract | CrossRef Full Text | Google Scholar

Santos, K. B., Guedes, I. A., Karl, A. L. M., and Dardenne, L. E. (2020). Highly flexible ligand docking: benchmarking of the DockThor program on the LEADS-PEP protein–peptide data set. J. Chem. Inf. Model. 60 (2), 667–683. doi:10.1021/acs.jcim.9b00905

PubMed Abstract | CrossRef Full Text | Google Scholar

Silva, D. A. da (2015). “Screening de produtos naturais com potencialidade para tratamento de doenças ocasionadas por helicobacter pylori: um estudo in vitro e in silico (Curitiba: Federal Technological University of Paraná). Available at: https://educapes.capes.gov.br/handle/capes/656986.

Google Scholar

Simão, D. A. D. A. S., Abreu, M. N. S., Gomez, R. S., Oliveira, L. D. d., Souza, R. S., Silva, T. M. R. d., et al. (2019). Evolução Clínica e Preditores da Neuropatia Periférica Induzida por Quimioterapia. Rev. Bras. Cancerol. 65 (2). doi:10.32635/2176-9745.rbc.2019v65n2.392

CrossRef Full Text | Google Scholar

So, J. Y., Ohm, J., Lipkowitz, S., and Yang, L. (2022). Triple negative breast cancer (TNBC): non-genetic tumor heterogeneity and immune microenvironment: emerging treatment options. Pharmacol. Ther. 237, 108253. doi:10.1016/j.pharmthera.2022.108253

PubMed Abstract | CrossRef Full Text | Google Scholar

Sreekeesoon, D. P., and Mahomoodally, M. (2014). Ethnopharmacological analysis of medicinal plants and animals used in the treatment and management of pain in Mauritius. J. Ethnopharmacol. 157, 181–200. doi:10.1016/j.jep.2014.09.030

PubMed Abstract | CrossRef Full Text | Google Scholar

Sukprasong, R. (2021). Allele frequencies of single nucleotide polymorphisms of clinically important drug-metabolizing enzymes CYP2C9, CYP2C19, and CYP3A4 in a Thai populatio. Sci. Rep. 11 (1), 1–8. doi:10.1038/s41598-021-90969-y

PubMed Abstract | CrossRef Full Text | Google Scholar

Sun, S., Liu, J., Kadouh, H., Sun, X., and Zhou, K. (2014). Three new anti-proliferative Annonaceous acetogenins with mono-tetrahydrofuran ring from graviola fruit (Annona muricata). Bioorg. Med. Chem. Lett. 24 (12), 2773–2776. doi:10.1016/j.bmcl.2014.03.099

PubMed Abstract | CrossRef Full Text | Google Scholar

Syed, NAJMUDDIN, and Syed, Umar Faruq (2016). Anti-cancer effect of Annona muricata linn leaves crude extract (AMCE) on breast cancer cell line. BMC complementary Altern. Med. 16, 1–20. doi:10.1186/s12906-016-1290-y

CrossRef Full Text | Google Scholar

Vasconcelos, G., Caetano, T., and Oliveira, T. (2022). A importância do farmacêutico na farmacoterapia antineoplásica oral. Dissertation (bachelor's degree). Avaliable in: https://repositorio.animaeducacao.com.br/handle/ANIMA/24800.2022.

Google Scholar

Venturi, I. (2018). Estudo Fitoquímico e Avaliação Biológica de Flores e Frutos de Solanum dipolconos (Mart) Bohs. Dissertation (Doctorate) - Vale do Rio. Itajai, Santa Catarina: Doce University.

Google Scholar

Vijayameena, C., Subhashini, G., Loganayagi, M., and Ramesh, B. (2013). Phytochemical screening and assessment of antibacterial activity for the bioactive compounds in Annona muricata. Int. J. Curr. Microbiol. Appl. Sci. 2 (1), 1–8.

Google Scholar

WORLD HEALTH ORGANIZATIONet al. (2009). Medicinal plants in Papua New Guinea: information on 126 commonly used medicinal plants in Papua New Guinea. Geneva, Switzerland: WHO Regional Office for the Western Pacific.

Google Scholar

Xiao, W., Wang, D., Shen, Z., Li, S., and Li, H. (2018). Multi-body interactions in molecular docking program devised with key water molecules in protein binding sites. Molecules 23 (9), 2321. doi:10.3390/molecules23092321

PubMed Abstract | CrossRef Full Text | Google Scholar

Xiong, G., Wu, Z., Yi, J., Fu, L., Yang, Z., Hsieh, C., et al. (2021). ADMETlab 2.0: an integrated online platform for accurate and comprehensive predictions of ADMET properties. Nucleic Acids Res. 49, W5–W14. doi:10.1093/nar/gkab255

PubMed Abstract | CrossRef Full Text | Google Scholar

Yazdanian, M., Glynn, S. L., Wright, J. L., and Hawi, A. (1998). Correlating partitioning and caco-2 cell permeability of structurally diverse small molecular weight compounds. Pharm. Res. 15 (9), 1490–1494. doi:10.1023/a:1011930411574

PubMed Abstract | CrossRef Full Text | Google Scholar

Yee, S. (1997). In vitro permeability across Caco-2 cells (colonic) can predict in vivo (small intestinal) absorption in man – fact or myth. Pharm. Res. 14 (6), 763–766. doi:10.1023/a:1012102522787

PubMed Abstract | CrossRef Full Text | Google Scholar

Zucker, E. H. (1985). Hazard Evaluation Division—standard evaluation procedure—acute test for freshwater fish. Washington, D.C., United States: U.S. EPA Publication.

Google Scholar

Keywords: cancer, pharmacokinetics, caspase 3, toxicology, plants, medicinal

Citation: Fallon Adido HE, Silva Chagas CK, Ferreira GG, Carmo Bastos ML and Dolabela MF (2023) In silico studies on cytotoxicity and antitumoral activity of acetogenins from Annona muricata L. Front. Chem. 11:1316779. doi: 10.3389/fchem.2023.1316779

Received: 10 October 2023; Accepted: 08 November 2023;
Published: 29 November 2023.

Edited by:

Snehasis Daschakraborty, Indian Institute of Technology Patna, India

Reviewed by:

Suman Das, University of Toronto, Canada
Niladri Patra, Indian Institute of Technology Dhanbad, India

Copyright © 2023 Fallon Adido, Silva Chagas, Ferreira, Carmo Bastos and Dolabela. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Maria Fâni Dolabela, fanidolabela03@gmail.com

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.