Next Article in Journal
A Natural Compound Containing a Disaccharide Structure of Glucose and Rhamnose Identified as Potential N-Glycanase 1 (NGLY1) Inhibitors
Previous Article in Journal
Traditional Uses, Pharmacological Activities, and Phytochemical Analysis of Diospyros mespiliformis Hochst. ex. A. DC (Ebenaceae): A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Friedelin: Structure, Biosynthesis, Extraction, and Its Potential Health Impact

by
Santosh Kumar Singh
1,
Shweta Shrivastava
2,
Awdhesh Kumar Mishra
3,*,
Darshan Kumar
1,
Vijay Kant Pandey
4,
Pratima Srivastava
1,
Biswaranjan Pradhan
5,
Bikash Chandra Behera
6,
Ashutosh Bahuguna
7 and
Kwang-Hyun Baek
3,*
1
Department of Biotechnology, ARKA Jain University, Jamshedpur 832108, Jharkhand, India
2
School of Pharmacy, ARKA Jain University, Jamshedpur 832108, Jharkhand, India
3
Department of Biotechnology, Yeungnam University, Gyeongsan 38541, Republic of Korea
4
Department of Agriculture, Netaji Subhas University, Jamshedpur 831012, Jharkhand, India
5
S.K. Dash Centre of Excellence of Biosciences and Engineering and Technology, Indian Institute of Technology, Bhubaneswar 752050, Odisha, India
6
School of Biological Sciences, National Institute of Science Education and Research, Bhubaneswar 752050, Odisha, India
7
Department of Food Science and Technology, Yeungnam University, Gyeongsan 38541, Republic of Korea
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(23), 7760; https://doi.org/10.3390/molecules28237760
Submission received: 25 October 2023 / Revised: 20 November 2023 / Accepted: 21 November 2023 / Published: 24 November 2023

Abstract

:
Pharmaceutical companies are investigating more source matrices for natural bioactive chemicals. Friedelin (friedelan-3-one) is a pentacyclic triterpene isolated from various plant species from different families as well as mosses and lichen. The fundamental compounds of these friedelane triterpenoids are abundantly found in cork tissues and leaf materials of diverse plant genera such as Celastraceae, Asteraceae, Fabaceae, and Myrtaceae. They possess many pharmacological effects, including anti-inflammatory, antioxidant, anticancer, and antimicrobial activities. Friedelin also has an anti-insect effect and the ability to alter the soil microbial ecology, making it vital to agriculture. Ultrasound, microwave, supercritical fluid, ionic liquid, and acid hydrolysis extract friedelin with reduced environmental impact. Recently, the high demand for friedelin has led to the development of CRISPR/Cas9 technology and gene overexpression plasmids to produce friedelin using genetically engineered yeast. Friedelin with low cytotoxicity to normal cells can be the best phytochemical for the drug of choice. The review summarizes the structural interpretation, biosynthesis, physicochemical properties, quantification, and various forms of pharmacological significance.

Graphical Abstract

1. Introduction

Plants in the form of whole plants, vegetables, fruits, whole grains, and nuts provide various phytochemicals like phenolic compounds, terpenoids, alkaloids, pigments, and other natural antioxidants [1,2]. These phytochemicals are nonnutritive substances that possess large health-protective benefits [3]. Over 80% of the world’s population relies on the traditional medical system to treat their health issues [4].
Friedelin (friedelan-3-one) is a pentacyclic triterpene first isolated from bark in 1807 using alcohol and called “cork alcohol” [5]. Later it was isolated from various plant species from different families [6,7,8] and also reported to isolate from lower plants like mosses [9], lichen [10,11], algae [12], and fungi [13]. In recent years, a substantial number of research studies have demonstrated extraordinary pharmacological actions of friedelin such as anti-inflammatory [14], anticancer [15,16,17,18], antioxidant and hepatoprotective [19,20,21], neuroprotective [22,23], antimicrobial [24,25,26], and antidiabetic effects [21,27,28]. Further, friedelin has outstanding natural anti-insect action, especially against S. litura and H. armigera [29], and could modulate the dynamics of the soil microbial community [30] and thus play an important role in agriculture. Friedelin has also been suggested for use as an herbicide due to the fact that it inhibits photosynthesis. [31,32] found that its phytotoxic activity inhibited both the root and shoot germination of wheat, rice, and pea germinating seeds. Consequently, friedelin and its derivatives offer prospective resources for the creation of novel medications or nutritional supplements.
Initially, the compound was extracted by soaking plant materials in various organic solvents. However, with the advancement of technology, modern extraction and analytical methods for the extraction of friedelin from different plant materials were used. With its great pharmaceutical usage, the amount of friedelin that can be obtained from plant sources is insufficient, whilst the amount that can be obtained through chemical processes is sometimes difficult to achieve, requires intense reaction conditions, and results in the production of unsafe compounds. In 2022, Wang et al. [33] used CRISPR/Cas9 technology and gene overexpression plasmids to produce friedelin using genetically engineered yeast.
According to the information obtained from various kinds of published research, friedelin could be a potential source in preventing various chronic diseases after the extraction and incorporation are completed at the target location. Nevertheless, the production of friedelin involves a number of significant procedures, the most important of which are the identification of the source, the determination of an appropriate quantity, and the execution of an appropriate extraction method. Therefore, this study is a complete presentation addressing the sources, qualities, and applications of friedelin, with a main emphasis on its numerous extraction procedures along with their applicability, limitations, and prospective remedies.

2. Sources of Friedelin

Due to the adverse effects of modern medications and therapies, plant products as therapeutic options are used worldwide. Friedelin has been studied for its potential biological activities and medicinal properties. One of the most significant natural sources of friedelin is found in cork. In the form of powder or granules, both natural and treated cork/bark are said to contain a large amount of friedelin. Leaves that are used with food directly are another source of friedelin. Friedelin was isolated from a wide variety of plant genera that belong to a wide range of plant families, and it was also found to be isolated from lower plants such as mosses [9], lichen [10], and algae [12]. Calophyllum pinetorum, Garcinia prainiana, Garcinia imberti, Garcinia rubroechinata, and Mammea siamensis belonging to the Clusiaceae family are major sources of friedelin. Drypetes tessmanniana, Putranjiva roxburghii, Uapaca ambanjensis, and Euphorbia tirucalli L. belonging to the family Euphorbiaceae; Salix tetrasperma and Populus davidiana belonging to the family Salicaceae; Terminalia avicennioides and Combretum duarteanum of the family Combretaceae; Luehea ochrophylla and Ancistrocarpus densispinosus of the family Tiliaceae; Prunus turfosa and Prunus lusitanica of the family Rosaceae and Maytenus ilicifolia and Maytenus aquifolium of family Celastraceae have been reported for the extraction of friedelin (Table 1).

3. Chemistry and Biosynthesis Pathway of Friedelin

Triterpenes are an essential class of chemicals that are present in both prokaryotes and eukaryotes and are involved in a wide variety of biological processes [77]. Triterpenoids and their derivatives have been shown to have a variety of functions, including those of hormones [78], lipid membrane components [79], and defense chemicals [80,81,82]. In addition, numerous triterpenoids, or their downstream products, have medical applications [83,84,85]. The biosynthesis of triterpenoids begins with the oxidation of 2-3-oxidosqualene, which is followed by protonation, cyclization, and several rearrangements [77]. This substance is capable of being transformed into a large variety of structurally distinct backbones, with over 100 of them having been found in plants alone. Members of the oxidosqualene cyclases (OSCs) and squalene-hopene cyclases (SHCs) families of enzymes commonly carry out these changes [77].
Friedelin is a pentacyclic triterpenoid that is derived from perhydropicene [45]. There is a substituent of an oxo group at position 3 of the perhydropicene molecule, as well as methyl groups at positions 4, 4a, 6b, 8a, 11, 11, 12b, and 14a [45]. It has a pentacyclic structure consisting of five rings with the molecular formula C30H50O (Figure 1). It is an extremely basic compound with a pKa value of −7.4 and showed high solubility in chloroform, sparing solubility in ethanol, and insolubility in water [40]. The molecular weight of friedelin is 426.7 g/mol with a topological surface area of 17.1 Å2. It contains one hydrogen bond acceptor, nine defined atom stereocenters, and one covalently bonded unit (details of chemical and physical properties of friedelin are shown in Table S1).
The biosynthesis of friedelin is a continuous process involving the pentacyclization of squalene oxide to the lupanyl cation and ten suprafacial 1,2-shifts of methyls and hydrogens as depicted in Figure 2. The first phase starts with the condensation of acetyl coenzyme A (CoA) units catalyzed by acetyl-CoA C-acetyltransferase (AACT) to produce acetoacetyl CoA. Subsequently, 3-hydroxy-3-methylglutaryl CoA (HMGCoA) synthase converts acetoacetyl CoA to hydroxy-3-methylglutaryl-CoA, which is regarded as a precursor to cholesterol. NADPH-dependent HMGCoA reductase converts HMGCoA into mevalonate (MVA). A series of enzymes then convert MVA into isopentenyl pyrophosphate (IPP), the five-carbon isoprenoid unit. IPP isomerase can convert IPP into dimethylallyl pyrophosphate (DMAPP) in a reversible manner. IPP and DMAPP are regarded as substrates shared by both pathways. DMAPP molecules are converted into farnesyl diphosphate by the enzyme farnesyl pyrophosphate synthase followed by squalene by the enzyme squalene synthase and finally to 2,3-Oxidosqualene by the enzyme squalene epoxidase [79].
In the second phase, oxidosqualene is protonated by oxidosqualene cyclases followed by cyclization, rearrangements, and deprotonation to synthesis friedelin [51,86]. Further, oxidosqualene cyclization begins with carbocation, which undergoes multiple rearrangements to generate various compounds. In the pathway, 2, 3-oxidosqualene converted into dammarenyl cation followed by baccharenyl cation, lupyl cation, Germanioyl cation, oleamyl cation, taraxaeyl cation, multiflorenol cation, Walsurenyl cation, companulyl cation, glutinyl cation, friedenyl cation, and finally friedelin (Figure 2).

4. Extraction Methods and Quantification

4.1. Extraction Methods and Analysis

The process of phytochemical extraction and purification from plant material is typically influenced by a number of parameters, the most important of which are time, temperature, the concentration of the solvent, and the polarity of the solvent [87]. Because it is unlikely that a single solvent could reliably extract all of the phytochemicals present in the plant material, distinct phytochemicals are extracted in solvents of varying polarity according to the chemical nature of the phytochemical.
Friedelin is a pentacyclic triterpenoid non-polar compound. Various extraction methods reported to be used for friedelin extraction are represented in Figure 3. Depending upon the plant parts used, organic solvents such as methanol, ethanol, hexane, dichloromethane, petroleum ether, and chloroform are commonly used to extract friedelin (aforementioned Table 1). Soxhlet has been the most common extraction method of friedelin for a very long time (Table 2). This assertion is supported by the fact that Soxhlet has been a standard technique for more than a century and is currently the standard against which other leaching methods are measured [88].
Supercritical fluid extraction (SFE) is an environmentally friendly and selective method to obtain plant extracts [91,92]. de Vasconcelos et al. [90] used a modified supercritical fluid extraction-CO2 method in which plant material was extracted with CO2 modified with methanol (10% v/v), ethanol (10% v/v), or pentane (10% or 20% v/v). They found that SFE CO2 modified with 10% methanol and SFE CO2 modified with 10% ethanol had the highest concentration of friedelin with 6.10 ± 0.75 mg/g plant material and 6.00 ± 0.97 mg/g plant material, respectively, as compared to SFE CO2 modified with 10% pentane, which had a friedelin yield of 3.00 ± 0.90 mg/g plant material, whereas SFE CO2 modified with 20% pentane had a friedelin yield of 3.70 ± 0.98 mg/g plant material [90]. De Melo et al. [86] used pure and ethanol-modified CO2 for the extraction of friedelin by the supercritical fluid extraction method. Different particle size (coarse particles to >80 mesh size) and ethanol (CO2 modifier) content (0–5 wt.%) in a 0.5 L capacity unit, at 300 bar, 50 ºC, and a CO2 flow rate of 11 g/min was used for the total (ηtotal) and friedelin (ηfriedelin) extraction. In this extraction method, they found that intermediate granulometries (40–60 mesh to 60–80 mesh) and a CO2: EtOH ratio 95.0: 5.0 wt% with 300 bar pressure enhance the friedelin selectivity by up to 2.6 times [86].
Vieira et al. [34] used a supercritical fluid extraction method for the extraction of friedelin from Quercus cerris bark. They analyzed two CO2:EtOH ratios, 95.0:5.0 wt% and 97.5:2.5 wt%, with QCO2 5 g/min & 8 g/min and temperature ranges 40–60 °C at a constant pressure of 300 bar. They found that the CO2:EtOH ratio of 97.5:2.5 wt%, QCO2 8 g/min, and temperature 60 °C had the highest extraction yield of 0.48 wt% with a friedelin concentration of 28 wt% [34]. Kornpointner et al. [65] used this extraction method with a condition like 2 h (1 h static/1 h dynamic) at 20 MPa and 60 °C from Cannabis sativa roots. The flow was set to 3 mL/min with 10 vol % EtOH as a co-solvent. They found that compared to conventional solvent extraction methods using n-hexane (0.698 ± 0.078 mg/g DW) and ethanol (0.0709 ± 0.036% by wt DW), the supercritical CO2 combined with ethanol method has significantly less friedelin (0.0548% by wt DW, p < 0.05) yield [65].
Microwaves are non-ionizing electromagnetic (EM) waves in the electromagnetic spectrum between radio-frequency waves and infrared. The frequency 915 MHz is best for industrial applications due to its greater penetration depth, while 2450 MHz is used in domestic microwave ovens and extraction applications with a wide range of commercial applications [93].
Alves et al. [61] used ultrasound-assisted extraction (UAE) and pressurized-liquid extraction (PLE) methods for friedelin extraction from Monteverdia aquifolia leaves extracts and compared them with the conventional Soxhlet extraction (SOX) extraction method. They found the highest yield of 8.3% in SOX with ethanol (360 min and ±78 °C) with 6.6% in UAE (30 min, 50 °C, amplitude 80%, and solvent/biomass ratio 20 mL g−1) and 5.3% in PLE (25 min and 60 °C) with the same solvent. However, UAE and PLE produced better extracts with greater TPC and AA than traditional extraction and used less solvent [61].
Mishra et al. (2020) [94] used microwave-assisted extraction (MAE) with eight vessels fitted with the exhaust in a closed extraction chamber accommodation. The friedelin concentrations of 0.010% by wt were extracted using fixed parameters like extraction vessel IR temperature limit (180 °C), pressure limit (20 psi), oscillation (ON), irradiation time (3 min), optimized microwave power (240 W), powdered drug (100 mg), and solvent (10 mL chloroform) with silicon carbide beads [94].

4.2. Quantification of Friedelin

Pharmaceutical researchers have studied friedelin extensively. Analytical quantification of friedelin is increasingly dependent on diverse methods. GC-MS and GC-FID, which enable new technology, are the main analytical methods.
Friedelin was detected in Putranjiva roxburghii Wall leaf extract (0.003% w/w), bark (0.04%), Ayurvedic formulation Femi-forte tablets (formulation 1) containing extracts of Putranjiva roxburghii (40 mg each tablet) (0.002%), and Femiplex tablets (13.05 mg each tablet) (formulation 2) (0.035%) [41]. Friedelin has an excellent linearity relationship (100–500 ng) with an r2 value of 0.9892. Friedelin had 32.15 and 97.44 ng/band detection and quantitation limits. The devised approach had interday and intraday precisions of 0.78% and 0.9%, respectively. Recovery experiments at three concentration levels showed 98.55% friedelin recovery [41].
Friedelin isolated from the cork of Quercus suber L. and cork byproduct was quantified by gas chromatography-mass spectroscopy (GC-MS) [95,96]. The chromatographic conditions were as follows: isothermal temperature 80 °C for 5 min followed by 285 °C for 15 min; 250 °C injector temperature; 285 °C transfer line temperature; 1:50 split ratio. The MS obtained data at 1 scan s−1 over m/z 33–800 in electron impact mode with 70 eV electron impact energy and the ion source were 200 °C. GC-MS quantification showed cork byproduct friedelin yields up to 1.4–5.0 g/kg [95] and Quercus suber cork yield friedelin of 2.47 g/kg dry weight [96].
Gas chromatography with flame ionization detection (GC-FID) was used for the quantification of friedelin extracted from Maytenus ilicifolia [97]. In this technique, 0.44 mg L−1 friedelin was quantified with the condition split mode (1:90) injected 1.0 L of samples in a 280 °C injector and 320 °C FID detector, with the column set at 300 °C isothermal mode, and the carrier gas was 1.5 mL/min helium [97].

5. Biological and Pharmacological Properties

Herbal therapies for the treatment of various diseases like diabetes, cancer, and liver problems are growing worldwide. Ayurvedic medicine uses herbal plant extracts to reduce the harmful side effects of medicines and improve efficacy. Numerous studies have demonstrated that Friedelin displays a diverse array of biological and pharmacological characteristics, encompassing hepatoprotective, antioxidant, anti-inflammatory, anticancer, anti-ulcerogenic, and neuroprotective properties.

5.1. Antioxidant and Hepatoprotective Activity

Under physiological conditions, the body maintains a balance between the production and elimination of free radicals. Excessive free radicals damage cellular proteins, membrane lipids, and nucleic acids, causing lipid peroxidation [98]. The antioxidant and hepatoprotective activity of friedelin is shown in Figure 4. Friedelin isolated from Azima tetracantha Lam. leaves sowed a free radical scavenging effect on 2,2-diphenyl-picrylhydrazyl (DPPH), nitric oxide, hydroxyl, and superoxide radical with IC50 values of 21.1 mM, 22.1 mM, 19.8 mM, and 21.9 mM, respectively [99]. Friedelin (25 μg/mL) isolated from Holothuria scabra showed 90.22 ± 0.15% DPPH free radical scavenging activity with an effective concentration (EC50) value that was found to be 14.63 ± 0.01μg/mL [19]. Friedelin isolated from the ethyl acetate extract of the stem of Tapinanthus bangwensis showed 73.69% free radical scavenging activities comparable to 93.96% by ascorbic acid [20].
Sunil et al. in 2021 [21] evaluated the hepatoprotective effects of friedelin in CCl4-induced oxidative stressed rats by administering a dose of 40 mg/kg for a period of 7 days. The results showed that friedelin was able to restore this hepatic enzyme to normal levels and exhibited similar hepatoprotective effects to silymarin (25 mg/kg), indicating its significant antioxidant and hepatoprotective properties [21,99]. Friedelin fulfilled Lipinski’s rule of five and showed better bioactivity than Silibinin. The findings of this investigation indicated that the bioactivity score of friedelin is superior to that of Silibinin, a potent hepatoprotective medication [100].

5.2. Anti-Ulcerogenic Activity

Due to side effects from conventional medications, herbal remedies for gastrointestinal diseases are becoming more popular globally [101]. Antonisamy et al. [14] reported the anti-ulcerogenic activity of friedelin in an ethanol-induced gastric ulcer mice model (Figure 5). They found that pretreatment of 35 mg/kg friedelin exhibited a protective effect against harmful impact in ethanol-induced gastric ulcer mice. Friedelin lowered vascular permeability, pro-inflammatory cytokines [tumor necrosis factor-α (TNF-α) and Interleukin 6 (IL-6)], iNOS, caspase-3, and apoptosis, whereas anti-inflammatory cytokines (Interleukin-10) and gastric mucus increased [14].
By comprehending network theory and systems biology, network pharmacology is the future drug discovery paradigm. Shi et al. [102] were able to identify the top 10 targets coming from the protein–protein interaction network by molecular docking of friedelin with ulcerative colitis (UC) receptors (Figure S1). Further, they demonstrated that 42 mg/kg/d intra-peritoneal friedelin (i.p) alleviated the effects of colitis by lowering inflammatory cytokines (IL-1β and IL-6), increased anti-inflammatory cytokines (IL-10), restored colon mucosa, and improved symptoms and bodily function in a mice model induced by dextran sulfate sodium [102].

5.3. Antidiabetic Activity

Different mechanisms of antidiabetic effects of friedelin are shown in Figure 6. Susanti et al. [27] investigated the antidiabetic properties of friedelin isolated from twigs of Garcinia prainiana in insulin sensitivity 3T3-L1 adipocytes. The study observed that the intracellular fat accumulation was increased by 2.02-fold in the presence of friedelin when treated with an adipogenic cocktail (0.5 mM 3-isobutyl-1-methyl-xanthine (IBMX), 0.25 mM dexamethasone, 1 µg/mL insulin) compared to the cells treated with the vehicle. Adipocyte insulin sensitivity was tested using a deoxyglucose uptake assay. Compared to insulin-treated cells, friedelin increased glucose absorption by 1.8-fold [27].
Sunil et al. [21] used STZ-induced diabetic Wistar rats to show the antidiabetic mechanism of isolated friedelin from A. tetracantha. Diabetic rats had decreased protein expression of liver PI3K, p-Akt, GLUT2 and AMPK, and skeletal muscle GLUT4. GLUT2, PI3K, AMPK, p-Akt, and GLUT4 protein expressions decreased in diabetic rats. They found that 40 mg/kg friedelin increased PI3K, p-Akt, GLUT2, and AMPK protein expression in STZ-induced diabetic rats [21].
α-glucosidase hydrolyzes polysaccharides to glucose. Diabetes medicines target α-glucosidase because only the intestinal tract is capable of absorbing monosaccharides from the digestion of the poly- and oligosaccharides. Acarbose, miglitol, and voglibose are approved α-glucosidase inhibitors. An investigation demonstrated that 100 μM friedelin isolated from Ficus drupacea leaves inhibited α-glucosidase by 20.1% [62]. Further, friedelin obtained from ethyl acetate fraction of Antidesma bunius bark exhibited greater α-glucosidase inhibitory activity with IC50 19.51 μg/mL compared to miglitol [28].
A computational approach to uncover the interaction between molecules extracted from Syzygium cumini and antidiabetic targets was done by Smruthi et al. [103]. Twenty-two phytoconstituents were docked with carbohydrate metabolism enzyme α-amylase using Autodock software (https://autodock.scripps.edu/ accessed on 19 November 2023) and the Lamarckian genetic algorithm. Friedelin had a significantly lower binding energy of −9.54 kcal/mol than the synthetic medication acarbose, which had a binding energy of −2.43 kcal/mol [103].

5.4. Anticancer Activity

Several clinical trials have shown that herbal remedies have anticancer properties [104]. Herbal medication was combined with conventional chemotherapy in anticancer therapy studies to improve therapeutic benefit, quality of life (QoL), and adverse effects [105]. Friedelin demonstrated anti-proliferative properties against various cancer cell lines [106]. The anticancer activity of friedelin is shown in Figure 7, and IC50 values are listed in Table 3.
Malignant breast cancer affects 18% of the world’s population and is the second leading cause of mortality for women [15,76]. An upregulation of estrogen receptors has been observed in several instances of breast cancer. Friedelin extracted from Hopea odorata demonstrated a −4.710 docking score with estrogen receptor alpha (ER-α) [76]. An in vitro cytotoxicity study of friedelin in human breast cancer cells (MCF-7) showed dose- as well as time-dependent inhibition of breast cancer proliferation [15]. Friedelin had an IC50 value of 0.51 μg/mL after 48 h for MCF-7 without causing any cytotoxicity in Vero and V79 cells. Friedelin caused DNA damage with significantly increased ROS levels. After 48 h of 1.2 μM friedelin treatment, Cdkn1a, pRb2, p53, Nrf2, and caspase-3 were upregulated, while Bcl-2, mdm2, and PCNA were downregulated, confirming apoptosis [15].
Prostate cancer is the second most common cancer in men and the fifth leading cause of death [109]. It has been reported that drugs targeting CYP17A1, a cytochrome P450c17 inhibitor, slow prostate cancer progression. The first approved CYP17A1 inhibitor was abiraterone acetate. However, successful drugs have side effects and therapeutic resistance in prostate cancer [110]. Friedelin derived from Cassia tora has been identified as the most optimal inhibitor of CYP17A1. Friedelin exhibits a stable binding pattern to the conserved binding pocket of CYP17A1, with a higher binding affinity compared to the control compound Orteronel. Friedelin’s IC50 was 72.025 and 81.766 μg/mL in hormone-sensitive (22Rv1-a human prostate carcinoma epithelial cell lines) as well as insensitive (DU145-a human prostate cancer cell lines) cell lines, respectively. The histopathological study confirmed that in animal trials, friedelin reduced prostate weight, blood PSA, prostate index, and testosterone [17].
Friedelin suppressed human leukemia cells (AML-196) while exhibiting minimal impact on healthy cells. The induction of apoptosis by friedelin was found to be associated with an increase in the expression of cleaved caspase-3, -8, and -9, as well as cleaved PARP. The levels of Bax protein were observed to be elevated, while those of Bcl-2 were found to be reduced. Friedelin inhibited AML-196 leukemia cell migration and invasion in transwell tests. Furthermore, Friedelin exhibited a dose-dependent inhibition of the MEK/ERK and PI3K/AT signaling pathways [16].
Using in silico molecular docking as well as molecular dynamics simulation, a total of 52 bioactive secondary metabolites from Wedelia trilobata were found to bind to the anti-apoptotic B-cell lymphoma-2 (Bcl-2) protein (PDB: 2W3L) structure. Friedelin’s binding energy against Bcl-2 protein was 10.1 kcal/mol compared to 8.4 kcal/mol for Obatoclax’s. In general, friedelin exhibits superior predicted absorption, distribution, metabolism, excretion, and toxicity (ADMET) properties compared to obatoclax. Friedelin derived from W. trilobata exhibits potential as an inhibitor of the Bcl-2 protein, which is known to be involved in cancer cell survival as well as resistance [18].
The cytotoxic effect of extracts and isolated compounds from Elaeocarpus floribundus was evaluated through a 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyl-2H-tetrazolium bromide (MTT) assay against human cervical (HeLa) as well as human T4 lymphoblastoid (CEM-SS) cancer cells. The friedelin demonstrated a strong inhibitory effect with an IC50 value of 3.54 ± 0.30 µg/mL toward HeLa cancer cells [47].
According to reports, friedelin exhibited cytotoxic properties towards the PC3 and U251 cancer cell lines. At 31 μM concentration, friedelin exhibited varying degrees of inhibition on K562, U251, and PC3 cancer cell lines of 0%, 25.8%, and 61.9%, respectively [111].
The Chinese Ministry of Health certifies bamboo shavings as useful food. Lu et al., (2010) examined the anticancer effects of triterpenoid-rich bamboo shavings extract (EBS) and friedelin. Friedelin separated from EBS was tested by MTT assay and showed strong anti-tumor activity on four cancer lines, L929 (IC50 value: 1.48 μg/mL), Hela (IC50 value: 2.59 μg/mL), A375 (IC50 value: 2.46 μg/mL), and THP-1 (IC50 value: 2.33 μg/mL), at an effected time of 48 h, compared to de-methylcantharidin [112].
Friedelin exhibited in vitro anticancer properties against glioblastoma multiforme (U87MG-GBM) cells. MTT assay indicates that FRI exhibited greater cytotoxicity towards U87MG cells in comparison to PRCC cells, as evidenced by IC50 values of 46.38 and 1271.77 µg/mL, respectively [11].

5.5. Neuroprotective Activity

Cognitive dysfunction is a major health issue in the 21st century, and many neuropsychiatric and neurodegenerative disorders, such as schizophrenia, Alzheimer’s disease dementia, seizure disorders, cerebrovascular impairment, and Parkinsonism, can severely debilitate [113]. Medicinal plant phytochemicals regulate the key inhibitory neurotransmitter receptors to maintain brain chemical homeostasis. Several plants cure cognitive problems in traditional medicine [114].
Chang et al. (2013) examined the neuroprotective effects of 176 phytochemicals on primary cortical neurons after oxygen-glucose deprivation. Friedelin showed similar cell viability after oxygen-glucose deprivation (OGD) insult, which was 0.97 ± 0.003 at 1 μM and 1.00 ± 0.009 at 10 μM, compared to the untreated control group at 1.00 [22]. After OGD insult, friedelin exhibited neuroprotective effects, as shown in Figure 8.
Oxidative stress (OS) along with c-Jun N-terminal kinase (JNK) have been significant factors involved in neuroinflammatory signaling pathways as well as their associated neurodegenerative disorders. Sandhu et al. [23] tested the friedelin neuroprotection effect. Friedelin exhibited a protective effect against scopolamine-induced oxidative stress, neuro-inflammation, glial cell activation, and p-JNK as well as NF-κB and their downstream signaling molecules. Friedelin was found to enhance neuronal synapse and improve memory deficits induced by scopolamine through the inhibition of β-secretase enzyme (BACE-1) and amyloidogenic pathways [23] as shown in Figure 8.

5.6. Antimicrobial and Antiparasitic Activity

With the rise of microbe resistance toward various antimicrobial drugs, novel antimicrobial medicines developed from natural bioactive substances were sought [115]. Friedelin has been extensively researched for its potential health advantages and various biological properties, such as its ability to inhibit the growth of microorganisms. Friedelin extracted from Pterocarpus erinaceous [25], Azima tetracantha [24], Jatropha tanjorensis [26], Calophyllum inophyllum, Maytenus undata [74], Calophyllum brasiliense [73], Garcinia smeathmannii [72], and Cola lateritia K. Schum [116] have been reported to have antimicrobial properties against bacterial and fungal pathogens, as shown in Table 4.
The antibacterial and resistance-modifying activities of friedelin isolated from the methanol extract of Paullinia pinnata L. roots were evaluated in vitro against Staphylococcus aureus strains SA1199B, RN4220, and XU212. These strains possess the Tet (K), Nor (A), and Msr (A) transporters, which confer resistance to tetracycline, norfloxacin, and macrolides, respectively [71]. Friedelin had moderate antibacterial activity against three resistant S. aureus strains with MICs between 128 and 256μg/mL. At a concentration of 10µg/mL, friedelin did not exhibit any antibacterial activity. However, when combined with tetracycline, erythromycin, and norfloxacin, friedelin demonstrated a two-fold increase in potency [71].
Friedelin isolated from Garcinia smeathmannii exhibited good antimicrobial activity against E. cloaclae, S. typhi, and S. faecalis with MIC value 0.61 µg/mL [72]. Friedelin extracted from Jatropha tanjorensis in methanol solvent showed maximum antibacterial and antifungal activity [26]. At a concentration of 2.5 mg/mL, friedelin showed 40 mm, 40 mm, and 38 mm zone of inhibition toward Gram-negative bacteria K. pneumoniae, P. mirabilis, and V. cholera, respectively, 40 mm and 37 mm toward Gram-positive bacteria B. cereus and S. epidermis, respectively, and 31 mm and 33 mm toward fungi A. fumigates and T. rubrum, respectively [26]. However, friedelin isolated from Maytenus undata (MIC value >250 µg/mL) and Calophyllum brasiliense (MIC value >1000 µg/mL) was reported to have very little microbial activity [73,74].
Recently, the SARS-CoV-2 enzyme inhibitory potential of friedelin has been analyzed using in silico computational evaluation. It was reported that friedelin had more hydrogen bonds than remdesivir after 100 ns of molecular dynamic investigations and may be useful against the SARS-CoV-2 spike protein [117]. Friedelin formed a stable interaction with inflammatory cytokines IL-6 (−10.4 ± 0.02), IL-1β (−10.8 ± 0.01), and anti-inflammatory cytokines IFN-γ (−10.1 ± 0.01), thus protecting the pathogenicity of SARS-CoV-2 [118]. Friedelin extracted from Vitex negundo bound to five SARS-CoV-2 protein targets protease, spike glycoprotein, NSP3, NSP9, and NSP15 in which NSP9 and NSP15 showed the highest binding affinity of −9.6 Kcal/mol and −8.6 Kcal/mol, respectively [70].
Parasites have been with humans since the beginning of time and cause high morbidity and mortality, especially in developing nations. Despite recent advancements, parasitic disease control remains difficult. Friedelin is reported to exhibit trypanocidal [119], leishmanicidal [119,120], and antimalarial [121] properties. Friedelin has good antiplasmodial efficacy against P. falciparum strain K1 (IC50 = 7.70 μM) [122] and the chloroquine-resistant strain (W2) P. falciparum (IC50 of 7.20 ± 0.5 µM) [123].

6. Conclusions

This work presents detailed information on the various kinds of plant species that have been explored for the isolation of friedelin. Supercritical fluid extraction (SFE), ultrasound-assisted extraction (UAE), microwave-assisted extraction (MAE), and pressurized-liquid extraction (PLE) have been used to maximize friedelin extraction from plant sources. In this review, convincing evidence has been presented for the potent antioxidant activity of friedelin in several assay systems. Friedelin also suppresses inflammation in the brain tissues and regulates different cell signaling pathways. This review summarized that friedelin possesses potential as a valuable adjunctive therapy in the prevention and management of neurodegenerative disorders, cancer, diabetes, and inflammatory disease due to its natural origin. Nevertheless, it is important to note that the majority of the cited findings in this study are derived from experiments conducted in laboratory settings and animal models, which may not accurately reflect the impact on human subjects. Therefore, further investigation is required to explore the various pharmacokinetic parameters, potentially involving human participants, in order to ascertain the suitability of this substance as a prescribed medication in the future. High demand for friedelin has led to the development of CRISPR/Cas9 technology and gene overexpression plasmids to make it in genetically altered yeast. To efficiently acquire friedelin in large quantities, different plasmids must be studied.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/molecules28237760/s1: Figure S1: Integrated targets between friedelin and ulcerative colitis (UC); Table S1: Computed chemical and physical properties of friedelin.

Author Contributions

Conceptualization, S.K.S. and S.S.; investigation, S.K.S. and V.K.P.; writing—original draft preparation, S.S., P.S. and D.K.; writing—review and editing, A.K.M., B.P. and B.C.B.; supervision, A.B. and K.-H.B.; funding acquisition, K.-H.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by Basic Science Research Program through the National Research Foundation of Korea (NRF), funded by the Ministry of Education (NRF-2021R1F1A1060297).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

There is no conflict of interest among the authors.

Abbreviations

pOSCsOxido squalene cyclases
SHCsSqualene-hopene cyclases
CoACoenzymeA
AACTAcetyl CoA Cacetyl transferase
HMGCoA3-hydroxy-3-methyl glutaryl CoA
MVAMevalonate
IPPIsopentenyl pyrophosphate
DMAPPDimethyl allyl pyrophosphate
SFESupercritical fluid extraction
EMElectromagnetic
UAEUltrasound-assisted extraction
PLEPressurized-liquid extraction
SOXSoxhlet extraction
MAEMicrowave-assisted extraction
GC-MSGas chromatography mass spectroscopy
GC-FIDGas chromatography with flame ionization detection
DPPH2,2-diphenyl-picrylhydrazyl
TNFTumor necrosis factor-α
ILInterleukin
UCUlcerative colitis
EREstrogen receptor
BclB-cell lymphoma
ADMETAbsorption, distribution, metabolism, excretion, and toxicity
MTT3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyl-2H-tetrazolium bromide
EBSBamboo shavings extract
GBMGlioblastoma Multiforme
OSOxidative stress
JNKc-JunN-terminal kinase
MICMinimum Inhibitory Concentration
SARS-CoVSevere Acute Respiratory Syndrome Corona Virus
ICInhibitory Concentration
IFN-γInterferon gamma
HSCHematopoietic stem cells
pADMSCsPorcine adipose-derived mesenchymal stem cells
MCF-7Michigan Cancer Foundation-7
CYP17A1Cytochrome P450 family 17 subfamily A member 1
HT-29Cells human colorectal adenocarcinoma cell lines
T24Urinary bladder colorectal adenocarcinoma cell lines
AML-196Acute myeloid leukemia 196
PRCCPapillary renal cell carcinoma

References

  1. Ndolo, V.; Maoni, M.; Mwamatope, B.; Tembo, D.T. Phytochemicals in commonly consumed foods in malawian diets. Funct. Foods Health Dis. 2022, 12, 564–575. [Google Scholar] [CrossRef]
  2. Anshuman, K.P. Phytochemicals: An immune booster against the pathogens. In Recent Frontiers of Phytochemicals; Pati, S., Sarkar, T., Lahiri, D., Eds.; Elsevier: Amsterdam, The Netherlands, 2023; pp. 501–509. ISBN 9780443191435. [Google Scholar]
  3. Dillard, C.J.; German, J.B. Phytochemicals: Nutraceuticals and human health. J. Sci. Food Agric. 2000, 80, 1744–1756. [Google Scholar] [CrossRef]
  4. WHO. WHO Guidelines on Safety Monitoring of Herbal Medicines in Pharmacovigilance Systems; World Health Organization: Geneva, Switzerland, 2004; ISBN 9241592214. [Google Scholar]
  5. Corey, E.J.T.; Ursprung, J.J. The Structures of the Triterpenes Friedelin and Cerin1,2. J. Am. Chem. Soc. 1956, 78, 5041–5051. [Google Scholar] [CrossRef]
  6. Caneschi, C.M.; de Souza, S.M.; Certo, T.S.; de Souza, G.H.B.; Campos, M.S.T.; Duarte, L.P.; de Silva, G.D.F.; de Gomes, M.S.; Vieira Filho, S.A. Samaras of Austroplenckia Populnea (Celastraceae): New constituents and effect of extracts and friedelin on germination of Bidens Pilosa (Asteraceae). Int. J. 2014, 6, 318–325. [Google Scholar]
  7. Alves, T.B.; Souza-Moreira, T.M.; Valentini, S.R.; Zanelli, C.F.; Furlan, M. Friedelin in Maytenus ilicifolia is produced by friedelin synthase isoforms. Molecules 2018, 23, 700. [Google Scholar] [CrossRef] [PubMed]
  8. Herrera, C.; Pérez, Y.; Morocho, V.; Armijos, C.; Malagón, O.; Brito, B.; Tacán, M.; Cartuche, L.; Gilardoni, G. Preliminary phytochemical study of the Ecuadorian plant Croton elegans Kunth (Euphorbiaceae). J. Chil. Chem. Soc. 2018, 63, 3875–3877. [Google Scholar] [CrossRef]
  9. Cao, Z.; Wang, Z.; Shang, Z.; Zhao, J. Classification and identification of Rhodobryum roseum Limpr. and its adulterants based on fourier-transform infrared spectroscopy (FTIR) and chemometrics. PLoS ONE 2017, 12, e0172359. [Google Scholar] [CrossRef] [PubMed]
  10. Solberg, Y. Chemical investigation of the lichen species Alectoria ochroleuca, Stereocaulon vesuvianum var. pulvinatum and Icmadophila ericetorum. Z. Naturforsch. C 1977, 32, 182–189. [Google Scholar] [CrossRef]
  11. Emsen, B.; Engin, T.; Turkez, H. In vitro investigation of the anticancer activity of friedelin in glioblastoma multiforme. Afyon Kocatepe Üniversitesi Mühendislik Bilim. Derg. 2018, 18, 763–773. [Google Scholar] [CrossRef]
  12. Chandler, R.F.; Hooper, S.N. Friedelin and associated triterpenoids. Phytochemistry 1979, 18, 711–724. [Google Scholar] [CrossRef]
  13. Richter, C.; Wittstein, K.; Kirk, P.M.; Stadler, M. An assessment of the taxonomy and chemotaxonomy of Ganoderma. Fungal Divers. 2015, 71, 1–15. [Google Scholar] [CrossRef]
  14. Antonisamy, P.; Duraipandiyan, V.; Aravinthan, A.; Al-Dhabi, N.A.; Ignacimuthu, S.; Choi, K.C.; Kim, J.-H. Protective effects of friedelin isolated from Azima tetracantha Lam. against ethanol-induced gastric ulcer in rats and possible underlying mechanisms. Eur. J. Pharmacol. 2015, 750, 167–175. [Google Scholar] [CrossRef] [PubMed]
  15. Subash-Babu, P.; Li, D.K.; Alshatwi, A.A. In vitro cytotoxic potential of friedelin in human MCF-7 breast cancer cell: Regulate early expression of Cdkn2a and pRb1, neutralize mdm2-p53 amalgamation and functional stabilization of p53. Exp. Toxicol. Pathol. 2017, 69, 630–636. [Google Scholar] [CrossRef]
  16. Chang, W.; Wang, J.; Xiao, Y. Friedelin inhibits the growth and metastasis of human leukemia cells via modulation of MEK/ERK and PI3K/AKT signalling pathways. J. BU. ON 2020, 25, 1594–1599. [Google Scholar]
  17. Joshi, B.P.; Bhandare, V.V.; Vankawala, M.; Patel, P.; Patel, R.; Vyas, B.; Krishnamurty, R. Friedelin, a novel inhibitor of CYP17A1 in prostate cancer from Cassia tora. J. Biomol. Struct. Dyn. 2022, 41, 9695–9720. [Google Scholar] [CrossRef] [PubMed]
  18. Gowtham, H.G.; Ahmed, F.; Anandan, S.; Shivakumara, C.S.; Bilagi, A.; Pradeep, S.; Shivamallu, C.; Shati, A.A.; Alfaifi, M.Y.; Elbehairi, S.E.I. In silico computational studies of bioactive secondary metabolites from Wedelia trilobata against anti-apoptotic B-cell lymphoma-2 (Bcl-2) protein associated with cancer cell survival and resistance. Molecules 2023, 28, 1588. [Google Scholar] [CrossRef] [PubMed]
  19. Nobsathian, S.; Tuchinda, P.; Sobhon, P.; Tinikul, Y.; Poljaroen, J.; Tinikul, R.; Sroyraya, M.; Poomton, T.; Chaichotranunt, S. An antioxidant activity of the whole body of Holothuria scabra. Chem. Biol. Technol. Agric. 2017, 4, 4. [Google Scholar] [CrossRef]
  20. Atewolara-Odule, O.C.; Aiyelaagbe, O.O.; Olubomehin, O.O.; Ogunmoye, A.O.; Feyisola, R.T.; Sanusi, A.S. Antioxidant activity of some secondary metabolites from Tapinanthus bangwensis (Engl., and K. Krause)[Loranthaceae] grown in Nigeria. Sci. Afr. 2020, 8, e00348. [Google Scholar] [CrossRef]
  21. Sunil, C.; Irudayaraj, S.S.; Duraipandiyan, V.; Alrashood, S.T.; Alharbi, S.A.; Ignacimuthu, S. Friedelin exhibits antidiabetic effect in diabetic rats via modulation of glucose metabolism in liver and muscle. J. Ethnopharmacol. 2021, 268, 113659. [Google Scholar] [CrossRef] [PubMed]
  22. Chang, H.-J.; Kim, Y.S.; Ryu, S.Y.; Chun, H.S. Screening of various sources of phytochemicals for neuroprotective activity against oxygen-glucose deprivation in vitro. J. Korean Soc. Appl. Biol. Chem. 2013, 56, 451–455. [Google Scholar] [CrossRef]
  23. Sandhu, M.; Irfan, H.M.; Shah, S.A.; Ahmed, M.; Naz, I.; Akram, M.; Fatima, H.; Farooq, A.S. Friedelin Attenuates Neuronal Dysfunction and Memory Impairment by Inhibition of the Activated JNK/NF-κB Signalling Pathway in Scopolamine-Induced Mice Model of Neurodegeneration. Molecules 2022, 27, 4513. [Google Scholar] [CrossRef] [PubMed]
  24. Duraipandiyan, V.; Gnanasekar, M.; Ignacimuthu, S. Antifungal activity of triterpenoid isolated from Azima tetracantha leaves. Folia Histochem. Cytobiol. 2010, 48, 311–313. [Google Scholar] [CrossRef] [PubMed]
  25. Ibrahim, M.; Idris, M.M.; Dauda, U.; Ali, U.; Muhammad, A. Isolation and Characterization of Friedelin and 5-DodecylResorcinol from the Stem Bark Extract of Pterocarpus erinaceus. J. Sci. Math. Lett. 2022, 10, 74–80. [Google Scholar]
  26. Viswanathan, M.B.G.; Ananthi, J.D.J.; Kumar, P.S. Antimicrobial activity of bioactive compounds and leaf extracts in Jatropha tanjorensis. Fitoterapia 2012, 83, 1153–1159. [Google Scholar] [CrossRef]
  27. Susanti, D.; Amiroudine, M.Z.A.M.; Rezali, M.F.; Taher, M. Friedelin and lanosterol from Garcinia prainiana stimulated glucose uptake and adipocytes differentiation in 3T3-L1 adipocytes. Nat. Prod. Res. 2013, 27, 417–424. [Google Scholar] [CrossRef] [PubMed]
  28. Mauldina, M.G.; Sauriasari, R.; Elya, B. α-Glucosidase inhibitory activity from ethyl acetate extract of Antidesma bunius (L.) Spreng stem bark containing triterpenoids. Pharmacogn. Mag. 2017, 13, 590–594. [Google Scholar]
  29. Baskar, K.; Duraipandiyan, V.; Ignacimuthu, S. Bioefficacy of the triterpenoid friedelin against Helicoverpa armigera (Hub.) and Spodoptera litura (Fab.) (Lepidoptera: Noctuidae). Pest Manag. Sci. 2014, 70, 1877–1883. [Google Scholar] [CrossRef]
  30. Dong, H.-Y.; Kong, C.-H.; Wang, P.; Huang, Q.-L. Temporal variation of soil friedelin and microbial community under different land uses in a long-term agroecosystem. Soil Biol. Biochem. 2014, 69, 275–281. [Google Scholar] [CrossRef]
  31. Castola, V.; Bighelli, A.; Rezzi, S.; Melloni, G.; Gladiali, S.; Desjobert, J.-M.; Casanova, J. Composition and chemical variability of the triterpene fraction of dichloromethane extracts of cork (Quercus suber L.). Ind. Crops Prod. 2002, 15, 15–22. [Google Scholar] [CrossRef]
  32. Pranab, G.; Amitava, M.; Madhumita, C.; Aniruddha, S. Triterpenoids from Quercus suber and their antimicrobial and phytotoxic activities. J. Chem. Pharm. Res. 2010, 2, 714–721. [Google Scholar]
  33. Wang, S.; Zhao, F.; Yang, M.; Lin, Y.; Han, S. Metabolic engineering of Saccharomyces cerevisiae for the synthesis of valuable chemicals. Crit. Rev. Biotechnol. 2023, 43, 1–28. [Google Scholar] [CrossRef]
  34. Vieira, P.G.; de Melo, M.M.R.; Şen, A.; Simões, M.M.Q.; Portugal, I.; Pereira, H.; Silva, C.M. Quercus cerris extracts obtained by distinct separation methods and solvents: Total and friedelin extraction yields, and chemical similarity analysis by multidimensional scaling. Sep. Purif. Technol. 2020, 232, 115924. [Google Scholar] [CrossRef]
  35. El-Shazly, A.; El-Sayed, A.; Fikrey, E. Bioactive secondary metabolites from Salix tetrasperma Roxb. Z. Naturforschung C 2012, 67, 353–359. [Google Scholar] [CrossRef]
  36. Alarcón, A.B.; Cuesta-Rubio, O.; Pérez, J.C.; Piccinelli, A.L.; Rastrelli, L. Constituents of the Cuban endemic species Calophyllum pinetorum. J. Nat. Prod. 2008, 71, 1283–1286. [Google Scholar] [CrossRef] [PubMed]
  37. Kuete, V.; Dongfack, M.D.J.; Mbaveng, A.T.; Lallemand, M.-C.; Van-Dufat, H.T.; Wansi, J.-D.; Seguin, E.; Tillequin, F.; Wandji, J. Antimicrobial activity of the methanolic extract and compounds from the stem bark of Drypetes tessmanniana. Chin. J. Integr. Med. 2010, 16, 337–343. [Google Scholar] [CrossRef] [PubMed]
  38. Sainsbury, M. Friedelin and epifriedelinol from the bark of Prunus turfosa and a review of their natural distribution. Phytochemistry 1970, 9, 2209–2215. [Google Scholar] [CrossRef]
  39. Tittikpina, N.K.; Nana, F.; Fontanay, S.; Philippot, S.; Batawila, K.; Akpagana, K.; Kirsch, G.; Chaimbault, P.; Jacob, C.; Duval, R.E. Antibacterial activity and cytotoxicity of Pterocarpus erinaceus Poir extracts, fractions and isolated compounds. J. Ethnopharmacol. 2018, 212, 200–207. [Google Scholar] [CrossRef] [PubMed]
  40. Mann, A.; Ibrahim, K.; Oyewale, A.O.; Amupitan, J.O.; Fatope, M.O.; Okogun, J.I. Antimycobacterial friedelane-terpenoid from the root bark of Terminalia avicennioides. Am. J. Chem. 2011, 1, 52–55. [Google Scholar] [CrossRef]
  41. Abhimanyu, K.K.; Ravindra, C.S.; Avanapu, R.S. A validated HPTLC method for the quantification of friedelin in Putranjiva roxburghii Wall extracts and in polyherbal formulations. Bull. Fac. Pharm. Cairo Univ. 2017, 55, 79–84. [Google Scholar] [CrossRef]
  42. Muniz, M.P.; Nunomura, S.M.; Lima, E.S.; Lima, A.S.; Almeida, P.; Nunomura, R. Quantification of bergenin, antioxidant activity and nitric oxide inhibition from bark, leaf and twig of Endopleura uchi. Quim. Nova 2020, 43, 413–418. [Google Scholar] [CrossRef]
  43. Araújo, C.R.R.; de Melo Silva, T.; Dos Santos, M.G.; Ottoni, M.H.F.; de Souza Fagundes, E.M.; de Sousa Fontoura, H.; de Melo, G.E.B.A.; de Carvalho Alcântara, A.F. Anti-inflammatory and cytotoxic activities of the extracts, fractions, and chemical constituents isolated from Luehea ochrophylla Mart. BMC Complement. Altern. Med. 2019, 19, 284. [Google Scholar] [CrossRef]
  44. Peyeino, J.H.; Djomkam, H.L.M.; Kenmogne, S.B.; Langat, M.K.; Isyaka, S.M.; Tsopgni, W.D.T.; Wansi, J.D.; Kamdem Waffo, A.F.; Sewald, N.; Vardamides, J.C. First report of compounds from an Ancistrocarpus species: Triterpenoids from A. densispinosus Oliv. (Malvaceae). Nat. Prod. Res. 2021, 36, 479–481. [Google Scholar] [CrossRef]
  45. Mujawah, A.; Rauf, A.; Bawazeer, S.; Wadood, A.; Hemeg, H.A.; Bawazeer, S. Isolation, Structural Elucidation, In Vitro Anti-α-Glucosidase, Anti-β-Secretase, and In Silico Studies of Bioactive Compound Isolated from Syzygium cumini L. Processes 2023, 11, 880. [Google Scholar] [CrossRef]
  46. Okafor, G.C.O.; Oyewale, A.O.; Habila, J.D.; Akpemi, M.A. Isolation, Characterization, and Assessment of the In Vitro Antibacterial and Antifungal Properties of Methanol Extracts and Friedelan-3-one from Uapaca ambanjensis (Leandri). J. Appl. Sci. Environ. Manag. 2022, 26, 1479–1486. [Google Scholar] [CrossRef]
  47. Utami, R.; Khalid, N.; Sukari, M.A.; Rahmani, M.; Abdul, A.B. Phenolic contents, antioxidant and cytotoxic activities of Elaeocarpus floribundus Blume. Pak. J. Pharm. Sci. 2013, 26, 245–250. [Google Scholar] [PubMed]
  48. Kumar, N.; Biswas, S.; Shrungeswara, A.H.; Mallik, S.B.; Viji, M.H.; Mathew, J.E.; Mathew, J.; Nandakumar, K.; Lobo, R. Pinocembrin enriched fraction of Elytranthe parasitica (L.) Danser induces apoptosis in HCT 116 colorectal cancer cells. J. Infect. Chemother. 2017, 23, 354–359. [Google Scholar] [CrossRef]
  49. Ndwigah, S.N.; Thoithi, G.N.; Mwangi, J.W.; Amugune, B.K.; Mugo, H.N.; Kibwage, I.O. Phytosterols from Dombeya torrida (JF Gmel.). East Cent. Afr. J. Pharm. Sci. 2013, 16, 44–48. [Google Scholar]
  50. Antonisamy, P.; Duraipandiyan, V.; Ignacimuthu, S. Anti-inflammatory, analgesic and antipyretic effects of friedelin isolated from Azima tetracantha Lam. in mouse and rat models. J. Pharm. Pharmacol. 2011, 63, 1070–1077. [Google Scholar] [CrossRef]
  51. Souza-Moreira, T.M.; Alves, T.B.; Pinheiro, K.A.; Felippe, L.G.; De Lima, G.M.A.; Watanabe, T.F.; Barbosa, C.C.; Santos, V.A.; Lopes, N.P.; Valentini, S.R. Friedelin synthase from Maytenus ilicifolia: Leucine 482 plays an essential role in the production of the most rearranged pentacyclic triterpene. Sci. Rep. 2016, 6, 36858. [Google Scholar] [CrossRef]
  52. Han, J.Y.; Ahn, C.-H.; Adhikari, P.B.; Kondeti, S.; Choi, Y.E. Functional characterization of an oxidosqualene cyclase (PdFRS) encoding a monofunctional friedelin synthase in Populus davidiana. Planta 2019, 249, 95–111. [Google Scholar] [CrossRef]
  53. Corsino, J.; de Carvalho, P.R.F.; Kato, M.J.; Latorre, L.R.; Oliveira, O.M.M.F.; Araújo, A.R.; da Bolzani, S.V.; França, S.C.; Pereira, A.M.S.; Furlan, M. Biosynthesis of friedelane and quinonemethide triterpenoids is compartmentalized in Maytenus aquifolium and Salacia campestris. Phytochemistry 2000, 55, 741–748. [Google Scholar] [CrossRef] [PubMed]
  54. Aswathy, S.V.; Joe, I.H.; Rameshkumar, K.B. Spectroscopic, quantum chemical and molecular docking studies on friedelin, the major triterpenoid isolated from Garcinia imberti. J. Mol. Struct. 2022, 1263, 133152. [Google Scholar] [CrossRef]
  55. Quintans, J.S.S.; Costa, E.V.; Tavares, J.F.; Souza, T.T.; Araújo, S.S.; Estevam, C.S.; Barison, A.; Cabral, A.G.S.; Silva, M.S.; Serafini, M.R. Phytochemical study and antinociceptive effect of the hexanic extract of leaves from Combretum duarteanum and friedelin, a triterpene isolated from the hexanic extract, in orofacial nociceptive protocols. Rev. Bras. Farmacogn. 2014, 24, 60–66. [Google Scholar] [CrossRef]
  56. Lim, W.Y.; Chan, E.W.C.; Phan, C.W.; Wong, C.W. Potent melanogenesis inhibition by friedelin isolated from Hibiscus tiliaceus leaves. Eur. J. Integr. Med. 2022, 55, 102181. [Google Scholar] [CrossRef]
  57. Vilkickyte, G.; Petrikaite, V.; Marksa, M.; Ivanauskas, L.; Jakstas, V.; Raudone, L. Fractionation and Characterization of Triterpenoids from Vaccinium vitis-idaea L. Cuticular Waxes and Their Potential as Anticancer Agents. Antioxidants 2023, 12, 465. [Google Scholar] [CrossRef] [PubMed]
  58. Mejía-Méndez, J.L.; Bach, H.; Lorenzo-Leal, A.C.; Navarro-López, D.E.; López-Mena, E.R.; Hernández, L.R.; Sánchez-Arreola, E. Biological Activities and Chemical Profiles of Kalanchoe fedtschenkoi Extracts. Plants 2023, 12, 1943. [Google Scholar] [CrossRef] [PubMed]
  59. Dharmasoth, R.D.; Rao, B.G.; Nageswara, S.; Basavaiah, K. Investigation of antibacterial activity of methanolic extract and isolation of sterol and triterpenoids from Grewia tiliaefolia vahl leaf. Int. J. Pharm. Pharm. Sci. 2022, 14, 34–43. [Google Scholar] [CrossRef]
  60. Menon, L.N.; Sindhu, G.; Raghu, K.G.; Rameshkumar, K.B. Chemical composition and cytotoxicity of Garcinia rubro-echinata, a Western Ghats endemic species. Nat. Prod. Commun. 2018, 13, 1934578X1801301121. [Google Scholar] [CrossRef]
  61. Alves, T.P.; Triques, C.C.; Palsikowski, P.A.; da Silva, C.; Fiorese, M.L.; da Silva, E.A.; Fagundes-Klen, M.R. Improved extraction of bioactive compounds from Monteverdia aquifolia leaves by pressurized-liquid and ultrasound-assisted extraction: Yield and chemical composition. J. Supercrit. Fluids 2022, 181, 105468. [Google Scholar] [CrossRef]
  62. Van Kiem, P.; Van Minh, C.; Nhiem, N.X.; Yen, P.H.; Anh, H.L.T.; Cuong, N.X.; Tai, B.H.; Quang, T.H.; Hai, T.N.; Kim, S.H. Chemical constituents of Ficus drupacea leaves and their α-glucosidase inhibitory activities. Notes 2013, 34, 263. [Google Scholar]
  63. Reyes-Chilpa, R.; Estrada-Muñiz, E.; Apan, T.R.; Amekraz, B.; Aumelas, A.; Jankowski, C.K.; Vázquez-Torres, M. Cytotoxic effects of mammea type coumarins from Calophyllum brasiliense. Life Sci. 2004, 75, 1635–1647. [Google Scholar] [CrossRef]
  64. Subhadhirasakul, S.; Pechpongs, P. A terpenoid and two steroids from the flowers of Mammea siamensis. Songklanakarin J. Sci. Technol. 2005, 27, 555–561. [Google Scholar]
  65. Kornpointner, C.; Martinez, A.S.; Marinovic, S.; Haselmair-Gosch, C.; Jamnik, P.; Schröder, K.; Löfke, C.; Halbwirth, H. Chemical composition and antioxidant potential of Cannabis sativa L. roots. Ind. Crops Prod. 2021, 165, 113422. [Google Scholar] [CrossRef]
  66. Do Céu Costa, M.; Duarte, P.; Neng, N.R.; Nogueira, J.M.F.; Costa, F.; Rosado, C. Novel insights for permeant lead structures through in vitro skin diffusion assays of Prunus lusitanica L., the Portugal Laurel. J. Mol. Struct. 2015, 1079, 327–336. [Google Scholar] [CrossRef]
  67. Tung, P.T.; Quyen, T.H.N.; Quang, T.T.; Van, N.T.T.; Duong, N.T.T.; Ngan, V.T.T.; Nga, L.T.; Phung, N.K.P.; Thu, N.T.H. Chemical constituents of the n-hexane extract of Leonotis nepetifolia (L.) R. Br (Lamiaceae). Vietnam J. Chem. 2020, 58, 719–722. [Google Scholar]
  68. Zou, D.; Li, X.; Zhou, X.; Luo, B.; Faruque, M.O.; Hu, S.; Chen, J.; Hu, X. The interaction of anti-inflammatory and anti-tumor components in the traditional Chinese medicine Solanum lyratum Thunb. Nat. Prod. Res. 2023, 37, 1–5. [Google Scholar] [CrossRef] [PubMed]
  69. De Souza, L.S.; Luz Tosta, C.; de Oliveira Borlot, J.R.P.; Varricchio, M.C.B.N.; Kitagawa, R.R.; Filgueiras, P.R.; Kuster, R.M. Chemical profile and cytotoxic evaluation of aerial parts of Euphorbia tirucalli L. on gastric adenocarcinoma (AGS cells). Nat. Prod. Res. 2023, 37, 1–7. [Google Scholar] [CrossRef] [PubMed]
  70. Nallusamy, S.; Mannu, J.; Ravikumar, C.; Angamuthu, K.; Nathan, B.; Nachimuthu, K.; Ramasamy, G.; Muthurajan, R.; Subbarayalu, M.; Neelakandan, K. Exploring phytochemicals of traditional medicinal plants exhibiting inhibitory activity against main protease, spike glycoprotein, RNA-dependent RNA polymerase and non-structural proteins of SARS-CoV-2 through virtual screening. Front. Pharmacol. 2021, 12, 667704. [Google Scholar] [CrossRef]
  71. Annan, K.; Adu, F.; Gbedema, S.Y. Friedelin: A bacterial resistance modulator from Paulinia Pinnata L. J. Sci. Technol. 2009, 29, 152–159. [Google Scholar] [CrossRef]
  72. Kuete, V.; Komguem, J.; Beng, V.P.; Meli, A.L.; Tangmouo, J.G.; Etoa, F.-X.; Lontsi, D. Antimicrobial components of the methanolic extract from the stem bark of Garcinia smeathmannii Oliver (Clusiaceae). S. Afr. J. Bot. 2007, 73, 347–354. [Google Scholar] [CrossRef]
  73. Pretto, J.B.; Cechinel-Filho, V.; Noldin, V.F.; Sartori, M.R.K.; Isaias, D.E.B.; Bella Cruz, A. Antimicrobial activity of fractions and compounds from Calophyllum brasiliense (Clusiaceae/Guttiferae). Z. Naturforsch. C 2004, 59, 657–662. [Google Scholar] [CrossRef]
  74. Mokoka, T.A.; McGaw, L.J.; Mdee, L.K.; Bagla, V.P.; Iwalewa, E.O.; Eloff, J.N. Antimicrobial activity and cytotoxicity of triterpenes isolated from leaves of Maytenus undata (Celastraceae). BMC Complement. Altern. Med. 2013, 13, 111. [Google Scholar] [CrossRef]
  75. Ali, M.S.; Mahmud, S.; Perveen, S.; Rizwani, G.H.; Ahmad, V.U. Screening for the antimicrobial properties of the leaves of Calophyllum inophyllum Linn. (Guttiferae). J. Chem. Soc. Pak. 1999, 21, 175–178. [Google Scholar]
  76. Paul, A.; Shoibe, M.; Islam, M.M.; Awal, M.A.; Hoq, M.I.; Rahman, M.H.; Mamur, A.; Hoque, M.S.U.; Chowdhury, T.A.; Kabir, M.S.H. Anticancer potential of isolated phytochemicals from Hopea odorata against breast cancer: In silico molecular docking approach. World J. Pharm. Res. 2016, 5, 1162–1169. [Google Scholar]
  77. Noushahi, H.A.; Khan, A.H.; Noushahi, U.F.; Hussain, M.; Javed, T.; Zafar, M.; Batool, M.; Ahmed, U.; Liu, K.; Harrison, M.T. Biosynthetic pathways of triterpenoids and strategies to improve their biosynthetic efficiency. Plant Growth Regul. 2022, 97, 439–454. [Google Scholar] [CrossRef]
  78. Hillier, S.G.; Lathe, R. Terpenes, hormones and life: Isoprene rule revisited. J. Endocrinol. 2019, 242, R9–R22. [Google Scholar] [CrossRef]
  79. Lu, Y.; Liu, Y.; Zhou, J.; Li, D.; Gao, W. Biosynthesis, total synthesis, structural modifications, bioactivity, and mechanism of action of the quinone-methide triterpenoid celastrol. Med. Res. Rev. 2021, 41, 1022–1060. [Google Scholar] [CrossRef] [PubMed]
  80. Wu, Z.; Wei, W.; Cheng, K.; Zheng, L.; Ma, C.; Wang, Y. Insecticidal activity of triterpenoids and volatile oil from the stems of Tetraena mongolica. Pestic. Biochem. Physiol. 2020, 166, 104551. [Google Scholar] [CrossRef] [PubMed]
  81. Li, H.; Sun, J.; Xiao, S.; Zhang, L.; Zhou, D. Triterpenoid-mediated inhibition of virus–host interaction: Is now the time for discovering viral entry/release inhibitors from nature? J. Med. Chem. 2020, 63, 15371–15388. [Google Scholar] [CrossRef] [PubMed]
  82. Yang, M.-J.; Luo, S.-H.; Guo, K.; Liu, Y.; Li, S.-H. Chemical investigation of Buddleja officinalis leaves and localization of defensive triterpenoids to its glandular trichomes. Fitoterapia 2023, 164, 105379. [Google Scholar] [CrossRef] [PubMed]
  83. Luz, D.A.; Pinheiro, A.M.; Fontes-Júnior, E.A.; Maia, C.S.F. Neuroprotective, neurogenic, and anticholinergic evidence of Ganoderma lucidum cognitive effects: Crucial knowledge is still lacking. Med. Res. Rev. 2023, 43, 1504–1536. [Google Scholar] [CrossRef]
  84. Sun, X.; Shen, B.; Yu, H.; Wu, W.; Sheng, R.; Fang, Y.; Guo, R. Therapeutic potential of demethylzeylasteral, a triterpenoid of the genus Tripterygium wilfordii. Fitoterapia 2022, 163, 105333. [Google Scholar] [CrossRef]
  85. Renda, G.; Gökkaya, İ.; Şöhretoğlu, D. Immunomodulatory properties of triterpenes. Phytochem. Rev. 2022, 21, 537–563. [Google Scholar] [CrossRef] [PubMed]
  86. De Melo, M.M.R.; Şen, A.; Silvestre, A.J.D.; Pereira, H.; Silva, C.M. Experimental and modeling study of supercritical CO2 extraction of Quercus cerris cork: Influence of ethanol and particle size on extraction kinetics and selectivity to friedelin. Sep. Purif. Technol. 2017, 187, 34–45. [Google Scholar] [CrossRef]
  87. Gil-Martín, E.; Forbes-Hernández, T.; Romero, A.; Cianciosi, D.; Giampieri, F.; Battino, M. Influence of the extraction method on the recovery of bioactive phenolic compounds from food industry by-products. Food Chem. 2022, 378, 131918. [Google Scholar] [CrossRef] [PubMed]
  88. De Castro, M.D.L.; Garcıa-Ayuso, L.E. Soxhlet extraction of solid materials: An outdated technique with a promising innovative future. Anal. Chim. Acta 1998, 369, 1–10. [Google Scholar] [CrossRef]
  89. Şen, A.; de Melo, M.M.R.; Silvestre, A.J.D.; Pereira, H.; Silva, C.M. Prospective pathway for a green and enhanced friedelin production through supercritical fluid extraction of Quercus cerris cork. J. Supercrit. Fluids 2015, 97, 247–255. [Google Scholar] [CrossRef]
  90. De Vasconcelos, E.C.; Vilegas, J.H.Y.; Lanças, F.M. Comparison of extraction and clean-up methods for the analysis of friedelan-3-ol and friedelin from leaves of Maytenus aquifolium Martius (Celastraceae). Phytochem. Anal. Int. J. Plant Chem. Biochem. Tech. 2000, 11, 247–250. [Google Scholar] [CrossRef]
  91. De Melo, M.M.R.; Silvestre, A.J.D.; Silva, C.M. Supercritical fluid extraction of vegetable matrices: Applications, trends and future perspectives of a convincing green technology. J. Supercrit. Fluids 2014, 92, 115–176. [Google Scholar] [CrossRef]
  92. Zorić, M.; Banožić, M.; Aladić, K.; Vladimir-Knežević, S.; Jokić, S. Supercritical CO2 extracts in cosmetic industry: Current status and future perspectives. Sustain. Chem. Pharm. 2022, 27, 100688. [Google Scholar] [CrossRef]
  93. Routray, W.; Orsat, V. Microwave-assisted extraction of flavonoids: A review. Food Bioprocess Technol. 2012, 5, 409–424. [Google Scholar] [CrossRef]
  94. Mishra, S.; Ramdas; Gupta, N.; Shanker, K. HPTLC method for the simultaneous determination of six bioactive terpenoids in Putranjiva roxburghii Wall. J. Planar Chromatogr. 2020, 33, 353–364. [Google Scholar] [CrossRef]
  95. Sousa, A.F.; Pinto, P.C.R.O.; Silvestre, A.J.D.; Pascoal Neto, C. Triterpenic and other lipophilic components from industrial cork byproducts. J. Agric. Food Chem. 2006, 54, 6888–6893. [Google Scholar] [CrossRef] [PubMed]
  96. Touati, R.; Santos, S.A.O.; Rocha, S.M.; Belhamel, K.; Silvestre, A.J.D. The potential of cork from Quercus suber L. grown in Algeria as a source of bioactive lipophilic and phenolic compounds. Ind. Crops Prod. 2015, 76, 936–945. [Google Scholar] [CrossRef]
  97. Vistuba, J.P.; Piovezan, M.; Pizzolatti, M.G.; Rebelo, A.M.; Azevedo, M.S.; Vitali, L.; Costa, A.C.O.; Micke, G.A. Increasing the instrumental throughput of gas chromatography method using multiple injections in a single experimental run: Application in determination of friedelan-3-ol and friedelin in Maytenus ilicifolia. J. Chromatogr. A 2013, 1274, 159–164. [Google Scholar] [CrossRef]
  98. Gao, J.; Hu, J.; Hu, D.; Yang, X. A Role of Gallic Acid in Oxidative Damage Diseases: A Comprehensive Review. Nat. Prod. Commun. 2019, 14, 1934578X19874174. [Google Scholar] [CrossRef]
  99. Sunil, C.; Duraipandiyan, V.; Ignacimuthu, S.; Al-Dhabi, N.A. Antioxidant, free radical scavenging and liver protective effects of friedelin isolated from Azima tetracantha Lam. leaves. Food Chem. 2013, 139, 860–865. [Google Scholar] [CrossRef]
  100. Patel, R.; Shukla, P.K.; Singh, M.P. Pharmacognostical, Phytochemical Evaluation and in silico lead finding of Ficus bengalensis Linn with hepatoprotective potentials. Int. J. Phytopharm. 2017, 7, 57–63. [Google Scholar]
  101. Parvez, M.K.; Rishi, V. Herb-drug interactions and hepatotoxicity. Curr. Drug Metab. 2019, 20, 275–282. [Google Scholar] [CrossRef]
  102. Shi, B.; Liu, S.; Huang, A.; Zhou, M.; Sun, B.; Cao, H.; Shan, J.; Sun, B.; Lin, J. Revealing the mechanism of friedelin in the treatment of ulcerative colitis based on network pharmacology and experimental verification. Evid. Based Complement. Altern. Med. 2021, 2021, 4451779. [Google Scholar] [CrossRef]
  103. Smruthi, G.; Mahadevan, V.; Vadivel, V.; Brindha, P. Docking studies on antidiabetic molecular targets of phytochemical compounds of Syzygium cumini (L.) skeels. Asian J. Pharm. Clin. Res. 2016, 9, 287–293. [Google Scholar]
  104. Dehelean, C.A.; Marcovici, I.; Soica, C.; Mioc, M.; Coricovac, D.; Iurciuc, S.; Cretu, O.M.; Pinzaru, I. Plant-derived anticancer compounds as new perspectives in drug discovery and alternative therapy. Molecules 2021, 26, 1109. [Google Scholar] [CrossRef]
  105. Berretta, M.; Dal Lago, L.; Tinazzi, M.; Ronchi, A.; La Rocca, G.; Montella, L.; Di Francia, R.; Facchini, B.A.; Bignucolo, A.; Montopoli, M. Evaluation of concomitant use of anticancer drugs and herbal products: From interactions to Synergic Activity. Cancers 2022, 14, 5203. [Google Scholar] [CrossRef] [PubMed]
  106. Yessoufou, K.; Elansary, H.O.; Mahmoud, E.A.; Skalicka-Woźniak, K. Antifungal, antibacterial and anticancer activities of Ficus drupacea L. stem bark extract and biologically active isolated compounds. Ind. Crops Prod. 2015, 74, 752–758. [Google Scholar] [CrossRef]
  107. Razwinani, M.; Motaung, K.S. The influence of friedelin, resinone, tingenone and betulin of compounds on chondrogenic differentiation of porcine adipose-derived mesenchymal stem cells (pADMSCs). Biochimie 2022, 196, 234–242. [Google Scholar] [CrossRef] [PubMed]
  108. Prabhu, A.; Krishnamoorthy, M.; Prasad, D.J.; Naik, P. Anticancer activity of friedelin isolated from ethanolic leaf extract of Cassia tora on HeLa and HSC-1 cell lines. Indian J. Appl. Res. 2013, 3, 1–4. [Google Scholar] [CrossRef]
  109. Bray, F.; Ferlay, J.; Soerjomataram, I.; Siegel, R.L.; Torre, L.A.; Jemal, A. Global cancer statistics 2018: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J. Clin. 2018, 68, 394–424. [Google Scholar] [CrossRef]
  110. Gomez, L.; Kovac, J.R.; Lamb, D.J. CYP17A1 inhibitors in castration-resistant prostate cancer. Steroids 2015, 95, 80–87. [Google Scholar] [CrossRef]
  111. Pillai Manoharan, K.; Yang, D.; Hsu, A.; Tan Kwong Huat, B. Evaluation of Polygonum bistorta for anticancer potential using selected cancer cell lines. Med. Chem. 2007, 3, 121–126. [Google Scholar] [CrossRef]
  112. Lu, B.; Liu, L.; Zhen, X.; Wu, X.; Zhang, Y. Anti-tumor activity of triterpenoid-rich extract from bamboo shavings (Caulis bamfusae in Taeniam). Afr. J. Biotechnol. 2010, 9, 6430–6436. [Google Scholar]
  113. Küpeli Akkol, E.; Tatlı Çankaya, I.; Şeker Karatoprak, G.; Carpar, E.; Sobarzo-Sánchez, E.; Capasso, R. Natural compounds as medical strategies in the prevention and treatment of psychiatric disorders seen in neurological diseases. Front. Pharmacol. 2021, 12, 669638. [Google Scholar] [CrossRef] [PubMed]
  114. Sharma, K.; Verma, R.; Kumar, D.; Nepovimova, E.; Kuča, K.; Kumar, A.; Raghuvanshi, D.; Dhalaria, R.; Puri, S. Ethnomedicinal plants used for the treatment of neurodegenerative diseases in Himachal Pradesh, India in Western Himalaya. J. Ethnopharmacol. 2022, 293, 115318. [Google Scholar] [CrossRef]
  115. Allemailem, K.S. Antimicrobial potential of naturally occurring bioactive secondary metabolites. J. Pharm. Bioallied Sci. 2021, 13, 155. [Google Scholar] [CrossRef]
  116. Kamdem, M.H.K.; Ojo, O.; Kemkuignou, B.M.; Talla, R.M.; Fonkui, T.Y.; Silihe, K.K.; Tata, C.M.; Fotsing, M.C.D.; Mmutlane, E.M.; Ndinteh, D.T. Pentacyclic Triterpenoids, Phytosteroids and Fatty Acid Isolated from the Stem-bark of Cola lateritia K. Schum. (Sterculiaceae) of Cameroon origin; Evaluation of Their Antibacterial Activity. Arab. J. Chem. 2022, 15, 103506. [Google Scholar] [CrossRef]
  117. Arumugam, M.; Shanmugavel, B.; Sellppan, M.; Pavadai, P. In silico evaluation of some commercially available terpenoids as spike glycoprotein of SARS-CoV-2–inhibitors using molecular dynamic approach. J. Biomol. Struct. Dyn. 2023, 41, 1–7. [Google Scholar] [CrossRef]
  118. Kar, P.; Saleh-E-In, M.M.; Jaishee, N.; Anandraj, A.; Kormuth, E.; Vellingiri, B.; Angione, C.; Rahman, P.K.S.M.; Pillay, S.; Sen, A. Computational profiling of natural compounds as promising inhibitors against the spike proteins of SARS-CoV-2 wild-type and the variants of concern, viral cell-entry process, and cytokine storm in COVID-19. J. Cell. Biochem. 2022, 123, 964–986. [Google Scholar] [CrossRef] [PubMed]
  119. Moiteiro, C.; Marcelo Curto, M.J.; Mohamed, N.; Bailén, M.; Martínez-Díaz, R.; González-Coloma, A. Biovalorization of friedelane triterpenes derived from cork processing industry byproducts. J. Agric. Food Chem. 2006, 54, 3566–3571. [Google Scholar] [CrossRef]
  120. Torres-Santos, E.C.; Lopes, D.; Oliveira, R.R.; Carauta, J.P.P.; Falcao, C.A.B.; Kaplan, M.A.C.; Rossi-Bergmann, B. Antileishmanial activity of isolated triterpenoids from Pourouma guianensis. Phytomedicine 2004, 11, 114–120. [Google Scholar] [CrossRef]
  121. Bapela, M.J.; Heyman, H.; Senejoux, F.; Meyer, J.J.M. 1H NMR-based metabolomics of antimalarial plant species traditionally used by Vha-Venda people in Limpopo Province, South Africa and isolation of antiplasmodial compounds. J. Ethnopharmacol. 2019, 228, 148–155. [Google Scholar] [CrossRef]
  122. Ngouamegne, E.T.; Fongang, R.S.; Ngouela, S.; Boyom, F.F.; Rohmer, M.; Tsamo, E.; Gut, J.; Rosenthal, P.J. Endodesmiadiol, a friedelane triterpenoid, and other antiplasmodial compounds from Endodesmia calophylloides. Chem. Pharm. Bull. 2008, 56, 374–377. [Google Scholar] [CrossRef]
  123. Lenta, B.N.; Ngouela, S.; Boyom, F.F.; Tantangmo, F.; Tchouya, G.R.F.; Tsamo, E.; Gut, J.; Rosenthal, P.J.; Connolly, J.D. Anti-plasmodial activity of some constituents of the root bark of Harungana madagascariensis L AM. (Hypericaceae). Chem. Pharm. Bull. 2007, 55, 464–467. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Structure of friedelin with UPAC name (4R,4aS,6aS,6aS,6bR,8aR,12aR,14aS,14bS)-4,4a,6a,6b,8a,11,11,14a-octamethyl-2,4,5,6,6a,7,8,9,10,12,12a,13,14,14b-tetradecahydro-1H-picen-3-one.
Figure 1. Structure of friedelin with UPAC name (4R,4aS,6aS,6aS,6bR,8aR,12aR,14aS,14bS)-4,4a,6a,6b,8a,11,11,14a-octamethyl-2,4,5,6,6a,7,8,9,10,12,12a,13,14,14b-tetradecahydro-1H-picen-3-one.
Molecules 28 07760 g001
Figure 2. Pentacyclic friedelin biosynthesis from 2, 3-oxidosqualene through oxidosqualene protonation, cyclization, several rearrangements, and deprotonating.
Figure 2. Pentacyclic friedelin biosynthesis from 2, 3-oxidosqualene through oxidosqualene protonation, cyclization, several rearrangements, and deprotonating.
Molecules 28 07760 g002
Figure 3. Extraction methods of friedelin from different plant materials.
Figure 3. Extraction methods of friedelin from different plant materials.
Molecules 28 07760 g003
Figure 4. Antioxidant and hepatoprotective effect of friedelin.
Figure 4. Antioxidant and hepatoprotective effect of friedelin.
Molecules 28 07760 g004
Figure 5. Schematic diagram of anti-ulcerogenic activity of friedelin.
Figure 5. Schematic diagram of anti-ulcerogenic activity of friedelin.
Molecules 28 07760 g005
Figure 6. Schematic diagram of antidiabetic activity of friedelin.
Figure 6. Schematic diagram of antidiabetic activity of friedelin.
Molecules 28 07760 g006
Figure 7. Schematic diagram of anticancer activity of friedelin.
Figure 7. Schematic diagram of anticancer activity of friedelin.
Molecules 28 07760 g007
Figure 8. Schematic diagram of neuroprotective activity of friedelin.
Figure 8. Schematic diagram of neuroprotective activity of friedelin.
Molecules 28 07760 g008
Table 1. Plant materials containing friedelin and solvent used for their extraction.
Table 1. Plant materials containing friedelin and solvent used for their extraction.
Plant PartPlantFamilySolvent Used for ExtractionReferences
Cork and/or stem barksQuercus cerrisFagaceaeMethanol, ethanol, dichloromethane, petroleum ether[34]
Salix tetraspermaSalicaceae80% aqueous methanol[35]
Calophyllum pinetorumClusiaceaeSequentially with n-hexane and ethyl acetate[36]
Drypetes tessmannianaEuphorbiaceaeMethanol[37]
Prunus turfosaRosaceae5% benzene in chloroform[38]
Pterocarpus erinaceusFabaceaeDichloromethane and methanol (1:1, v/v)[39]
Terminalia avicennioidesCombretaceaePetroleum ether, ethyl acetate, chloroform, and methanol [40]
Putranjiva roxburghiiEuphorbiaceaeChloroform[41]
Endopleura uchiHumiriaceaeHexane[42]
Luehea ochrophyllaTiliaceaeHexane and ethanol[43]
Ancistrocarpus densispinosus Oliv. TiliaceaeMethanol[44]
Syzygium cumini L. Myrtaceae70% methanol[45]
Garcinia prainianaClusiaceaen-hexane[27]
Uapaca ambanjensisEuphorbiaceaeSequentially with n-hexane, dichloromethane, ethyl acetate, and methanol, respectively[46]
Elaeocarpus floribundusElaeocarpaceaeSequentially with hexane, chloroform, ethyl acetate, and methanol[47]
Elytranthe parasiticaLoranthaceaeMethanol[48]
Dombeya torridaSterculiaceaeChloroform[49]
LeavesAzima tetracantha Lam.SalvadoraceaeHexane[50]
Maytenus ilicifoliaCelastraceaeHexane: Ethyl acetate (8:2, v/v)[7,51]
Populus davidianaSalicaceaeLiquid WPM medium with 1% sucrose[52]
Maytenus aquifoliumCelastraceaeEthanol[53]
Garcinia imbertiClusiaceaeHexane[54]
Combretum duarteanumCombretaceaeEthanol[55]
Hibiscus tiliaceusMalvaceaeDichloromethane[56]
Vaccinium vitisidaea L. EricaceaeChloroform[57]
Kalanchoe fedtschenkoiCrassulaceaeHexane and chloroform[58]
Grewia tiliaefoliaMalvaceaeMethanol[59]
Dombeya torridaSterculiaceaeDichloromethane: Methanol (50:50)[49]
Garcinia rubroechinataClusiaceaen-hexane followed by methanol[60]
Tapinanthus bangwensisLoranthaceaeSuccessively with n-hexane, ethyl acetate, and methanol [20]
Monteverdia aquifoliaCelastraceaeEthanol[61]
Ficus drupaceaMoraceaeNA[62]
RhizomesPolygonum bistortaPolygonaceaeChloroform[63]
FlowerMammea siamensisClusiaceaeChloroform and methanol[64]
RootCannabis sativaCannabaceaeEtOH and n-hexane[65]
Aerial partsPrunus lusitanicaRosaceaePetroleum ether[66]
Leonotis nepetifolia (L.) R. Br LamiaceaeEthanol followed by methanol[67]
LichenAlectoria ochroleucaParmeliaceaeAcetone[10]
MossRhodobryum roseumBryaceaeNA[9]
Whole plantSolanum lyratum ThunbSolanaceaeEthanol[68]
Whole plantEuphorbia tirucalliEuphorbiaceaeHexane and aqueous[69]
Vitex negundoLamiaceaeNA[70]
Paullinia pinnataSapindaceaeMethanol[71]
Garcinia smeathmanniiClusiaceaeMethanol[72]
Calophyllum brasilienseClusiaceaeMethanol[73]
Maytenus undataCelastraceaeHexane, dichloromethane, acetone, and methanol[74]
Calophyllum inophyllumClusiaceaeEthanol, butanol, chloroform[75]
Jatropha tanjorensisEuphorbiaceaeHexane, chloroform, and methanol[26]
Wedelia trilobataAsteraceaeNA[18]
Cassia toraLeguminosaeEthanol[17]
Hopea odorataDipterocarpaceaeNA[76]
Antidesma buniusEuphorbiaceaeEthyl acetate[28]
Azima tetracanthaSalvadoraceaeDistilled water, phosphate buffer K3Fe(CN)6[19]
NA—Not available.
Table 2. Extraction of friedelin using Soxhlet methods from different plant sources.
Table 2. Extraction of friedelin using Soxhlet methods from different plant sources.
PlantPlant MaterialExtraction ConditionSolventFriedelin
Concentration
References
Quercus cerrisCork120 mL solvent, 1 bar pressure, ~3 g biomass, 8 h.Methanol12.1 wt %[34]
Ethanol15.2 wt %
Dichloromethane23.7 wt %
Petroleum ether41.3 wt %
Quercus cerrisCorkNADichloromethane26.03 wt %[89]
Maytenus aquifoliumLeaves10 g plant material with successive hexane and chloroform as an extraction solvent for 20 h eachHexane0.49 wt %[90]
Dombeya torridaBark1 kg of the stem bark powder extracted with chloroform for 48 hChloroformNA[49]
Garcinia rubroechinataLeaves500 g of the dry leaf powder extracted with n-hexane followed by methanol for 24 hn-hexane followed by methanol3.0 wt %[60]
Putranjiva roxburghiileaf and bark50 g of dried powder extracted with Chloroform for 6 h.Chloroform0.003% w/w in leaf extract, 0.04% w/w in bark extract[41]
NA—Not available.
Table 3. IC50 values toward different cell lines treated with friedelin.
Table 3. IC50 values toward different cell lines treated with friedelin.
S. No.Cell LineCancer TypeBiological SourceAssay, Time of
Execution
IC50 Concentration UsedReferences
1MCF-7Breast cancerHuman breast (adenocarcinoma)MTT, 24 h,
MTT< 48 h
0.76 μg/mL and 0.51 μg/mL
22.81 ± 2.1 µg/mL
[15,106]
222Rv1Prostate cancerHuman prostateMTT, 24 h72.025 μg/mL[17]
3DU145Prostate cancerHuman prostateMTT, 24 h81.766 μg/mL[17]
4AML-196LeukemiaHuman
leukemia cells
CCK-8, 24 h34 μg/mL[16]
5pADMSCsAdipose-derived mesenchymal stem cellsPorcine mesenchymal stem cellsMTT, 48 h15 μg/mL[107]
6U87 MG-GBMBrain cancerHuman brain (glioblastoma astrocytoma)MTT, 4 h46.38 μg/mL[11]
7HeLaCervical cancerEpithelioid cervix carcinomaMTT, 72 h
MTT, 24 h
MTT, 24 h
3.54 ± 0.30 μg/mL
20.42 ± 2.3 μg/mL;
19.3 ± 1.27 μg/mL
[47,106,108]
8JurkatLeukemiaHuman blood (leukemic T-cell lymphoblast)MTT, 24 h29.15 ± 2.3 μg/mL[106]
9HT-29Colon cancerHuman colon adenocarcinomaMTT, 24 h37.21 ± 3.61 μg/mL[106]
10T24Urinary bladder cancerHuman bladder carcinomaMTT, 24 h12.81 ± 1.4 μg/mL[106]
11HSC-1Squamous carcinomaHuman squamous carcinomaMTT, 24 h28.7 ± 1.98 μg/mL[108]
Table 4. Antimicrobial activity of friedelin extracted from different plant sources.
Table 4. Antimicrobial activity of friedelin extracted from different plant sources.
Source of FriedelinMethodMicroorganismZone of Inhibition (mm)/Minimal Inhibition Concentration (µg/mL)References
Pterocarpus erinaceous1000 µg/mL
disk diffusion
Staphyloccocus aureus17[25]
Aspergillus flavus10
Azima tetracanthaMinimal inhibition concentration (MIC)T. mentagrophytes>250[24]
T. simii125
T. rubrum 57/01>250
Epidermophyton floccosum125
Scopulariopsis sp.>250
Aspergillus niger125
Curvularia lunata62.5
Magnethophora sp.125
Candida albicans>250
Jatropha tanjorensis2.5 mg/mL
Disk diffusion
Bacillus cereus 43040[26]
Staphylococcus epidermis 43537
Aeromonas hydrophila 64632
Klebsiella pneumoniae 43240
Proteus mirabilis 42540
Proteus vulgaris 42617
Salmonella paratyphi 73335
Vibrio alcaligenes 444227
Vibrio cholera 390638
Aspergillus fumigates 34331
Candida albicans 227NA
Microsporum gyseum 2819NA
Trichophyton rubrum 29633
Calophyllum inophyllumDisk diffusionStaphylococcus aureus6.6[75]
Corynebacterium dptheriae3.5
Salmonella typhi3.53
Klebsiella pneumoniae4.0
Proteus mirabilis3.11
% growth inhibition compared to standard drug miconazole and ketoconazolePseudallescheria boydii81.04
Candida albicans51.73
Aspergillus niger85.09
Trichophyton schoenleinii55.05
Maytenus undataMICStaphyloccocus aureus>250[74]
E. coli>250
Pseudomonas aeruginosa>250
Enterococcus faecalis>250
Candida albicans>250
Candida neofamans>250
Calophyllum brasiliense235 µM
MIC
Bacillus cereus>1000[73]
Staphylococcus aureus>1000
Staphylococcus saprophyticus>1000
Streptococcus agalactiae>1000
Enterobacter cloacae>1000
Escherichia coli>1000
Pseudomonas aeruginosa>1000
Proteus mirabilis>1000
Salmonella typhimurium>1000
Candida albicans>1000
Candida tropicalis>1000
Garcinia smeathmanniiMICE. cloaclae0.61[72]
P. vulgaris1.22
S. dysenteria1.22
S. flexneri1.22
S. typhi0.61
S. typhimurium1.22
B. megaterium1.22
B. stearothermophilus1.22
S. faecalis0.61
C. albicans2.44
C. krusei4.88
C. gabrata2.44
NA—Not available.
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

Singh, S.K.; Shrivastava, S.; Mishra, A.K.; Kumar, D.; Pandey, V.K.; Srivastava, P.; Pradhan, B.; Behera, B.C.; Bahuguna, A.; Baek, K.-H. Friedelin: Structure, Biosynthesis, Extraction, and Its Potential Health Impact. Molecules 2023, 28, 7760. https://doi.org/10.3390/molecules28237760

AMA Style

Singh SK, Shrivastava S, Mishra AK, Kumar D, Pandey VK, Srivastava P, Pradhan B, Behera BC, Bahuguna A, Baek K-H. Friedelin: Structure, Biosynthesis, Extraction, and Its Potential Health Impact. Molecules. 2023; 28(23):7760. https://doi.org/10.3390/molecules28237760

Chicago/Turabian Style

Singh, Santosh Kumar, Shweta Shrivastava, Awdhesh Kumar Mishra, Darshan Kumar, Vijay Kant Pandey, Pratima Srivastava, Biswaranjan Pradhan, Bikash Chandra Behera, Ashutosh Bahuguna, and Kwang-Hyun Baek. 2023. "Friedelin: Structure, Biosynthesis, Extraction, and Its Potential Health Impact" Molecules 28, no. 23: 7760. https://doi.org/10.3390/molecules28237760

Article Metrics

Back to TopTop