Next Article in Journal
Translational Insights into NK Immunophenotyping: Comparative Surface Marker Analysis and Circulating Immune Cell Profiling in Cancer Immunotherapy
Previous Article in Journal
Genetic Susceptibility to Tuberculosis and the Utility of Polygenic Scores in Population Stratification
Previous Article in Special Issue
Nutritional Characteristics, Health-Related Properties, and Food Application of Teff (Eragrostis tef): An Overview
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Comprehensive Review of Robinetin: Distribution, Biological Activity and Pharmacokinetic Parameters

by
Katarzyna Jakimiuk
Department of Biology and Pharmacognosy, Faculty of Pharmacy with the Division of Laboratory Medicine, Medical University of Białystok, Mickiewicza 2a, 15-230 Białystok, Poland
Int. J. Mol. Sci. 2025, 26(19), 9546; https://doi.org/10.3390/ijms26199546
Submission received: 9 September 2025 / Revised: 22 September 2025 / Accepted: 29 September 2025 / Published: 30 September 2025
(This article belongs to the Special Issue Role of Natural Compounds in Human Health and Disease)

Abstract

Robinetin, a naturally occurring polyhydroxylated flavonol, has gained attention due to its broad spectrum of biological activities and potential therapeutic applications. This review presents a comprehensive summary of the current knowledge concerning the natural occurrence, extraction, spectroscopic characterization, and pharmacological properties of robinetin. Ethnobotanical evidence highlights its presence in various medicinal plants, particularly within the Fabaceae family, where it contributes to traditional treatments of infections, inflammation, and metabolic disorders. Robinetin exhibits diverse bioactivities, including antiviral, antibacterial, antiparasitic, antioxidant, anti-mutagenic, and enzyme-inhibitory effects. Notably, it inhibits HIV-1 integrase and acetylcholinesterase and demonstrates moderate antiproliferative activity in cancer cell lines. Despite limited water solubility, its redox behavior and metal-chelating capabilities support its antioxidant potential. Recent in vivo studies indicate its hepatoprotective and metabolic regulatory effects. Additionally, computational models reveal promising interactions with molecular targets such as CDK1. Collectively, these findings underscore the multifaceted therapeutic potential of robinetin and advocate for further pharmacokinetic and clinical investigations to validate its efficacy as a lead compound for the development of phytochemically derived pharmaceuticals.

1. Introduction

Plants are an infinite source of bioactive substances where flavonoids remain one of the most important chemicals with health-beneficial potential [1]. Flavonoids are divided into groups according to their substitution pattern and primarily are composed of phenols in their free form or glycosides [2]. Polyhydroxylated flavonoids (PHF) are a class of polyphenolic compounds characterized by multiple hydroxyl groups, which confer diverse and significant biological activities [3,4]. The biological activities of polyhydroxylated flavonoids are significantly influenced by their structural features. The number and position of hydroxyl groups, degree of glycosylation, and presence of other substituents affect their solubility, bioavailability, and interaction with biological targets [4]. Also, the diversity in physicochemical properties of this class of compounds influence their biological activity, solubility, and bioavailability which is essential to understanding their mechanisms of action and optimizing their therapeutic applications [5]. Polyhydroxylated flavonoids generally exhibit poor water solubility due to their hydrophobic aromatic ring; however, aglycone forms are more lipophilic, facilitating their penetration through lipid bilayers and membranes [6]. Furthermore, these flavonoid derivatives are effective metal chelators, particularly for transition metals (e.g., iron, copper). Chelation occurs via hydroxyl groups, especially in positions C3, C4, and C5 on the flavonoid backbone. This property is crucial for their antioxidant effects, as it prevents metal-catalyzed oxidative reactions [7,8]. Another important chemical reactivity of flavonoids with unsubstituted hydroxyl groups is redox properties. The hydroxyl groups in flavonoids undergo redox cycling, allowing them to donate electrons to neutralize free radicals. This redox capability is enhanced in flavonoids with catechol structures in their B-ring [9,10]. One of polyhydroxy flavones occurring in the plant kingdom is robinetin (norkanugin, 5-hydroxyfisetin, 5-deoxymyricetin) which belong to the flavonol subclass (Figure 1). Robinetin consists of a flavonoid skeleton with hydroxyl groups at the 3, 5, 7, 3′, and 4′ positions [11]. Available studies suggest that robinetin exhibit diverse biological activities and may be promising antiviral, anti-inflammatory, antioxidant or anticancer therapeutic agent [12,13,14].
Although various biological activities of robinetin have been established, no clearly organized review articles are available. Thus, this paper gives an outline of the findings on the techniques used for the analysis, isolation, and separation of robinetin, as well as described biological properties and therapeutic activities of this compound.

2. Methodology

A systematic literature search was conducted using a predefined query across key bibliographic platforms. The following resources were used: Scopus; (https://www.elsevier.com/ (accessed on 9 September 2025)); Google Scholar (https://scholar.google.com/ (accessed on 9 September 2025)); PubMed/MEDLINE (https://pubmed.ncbi.nlm.nih.gov/ (accessed on 9 September 2025)); Web of Science (SCI-EXPANDED) (https://www.webofscience.com/ (accessed on 9 September 2025)); Taylor & Francis Online (https://taylorandfrancis.com/online/taylor-francis-online/ (accessed on 9 September 2025)); Wiley Online Library (https://onlinelibrary.wiley.com/ (accessed on 9 September 2025)); EBSCO Discovery Service (EDS) (https://search.ebscohost.com/ (accessed on 9 September 2025)); REAXYS (https://www.reaxys.com/#/search/quick/query (accessed on 9 September 2025)); and ScienceDirect (Elsevier) (https://www.sciencedirect.com/ (accessed on 9 September 2025)). The chemical structure validation was achieved via cross-referencing entries in PubChem and REAXYS. TITLE-ABS-KEY fields in the specified databases were queried using designated keywords, either individually or in various logical combinations, in accordance with each database’s search constraints: “robinetin”, “robinetin derivatives”, “polihydroxyflavonoids”, “biological activity”, “natural compound”, “isolation and identification”, “chemical characterization”, “chemical analysis”, “metabolism”, “chemical and physical properties”. Inclusion criteria encompassed English-language studies published between 1933, marking the first documented mention of robinetin, and 2025 (September). The search process featured a two-stage screening—first assessing titles and abstracts, then evaluating full-text articles. Further relevant studies were located through review article bibliographies and cross-referencing of cited works. The chemical structure of robinetin was drawn using ACD/ChemSketch (2014.1.5 Freeware Version).

3. Natural Occurrence of Robinetin

Table 1 presents a compilation of ethnopharmacological data concerning robinetin-containing plant species, with a predominant representation from the Fabaceae family, alongside members of the Boraginaceae, Asteraceae, Loranthaceae, Nelumbonaceae, Brassicaceae, Araceae and Moringaceae families. The documented species exhibit a wide range of traditional medicinal applications, often associated with specific plant parts such as bark, leaves, seeds, and roots. For instance, Acacia mearnsii is traditionally utilized for the treatment of microbial infections, whereas Albizia lebbeck demonstrates a broad ethnomedical profile, with its leaves employed in managing ulcers, night blindness, respiratory and dermatological conditions, as well as envenomations, leprosy, gonorrhea, and pharyngeal disorders [15,16]. In Adenanthera pavonina, both seeds and leaves, is known for its application in treating boils and inflammatory states [17]. Cordia myxa, a representative of the Boraginaceae family, has traditionally been applied in the treatment of wounds, boils, tumors, gout, and ulcers, despite the unspecified plant part used [18]. Cosmos caudatus, belonging to the Asteraceae family, exhibits multifaceted therapeutic uses, with its leaves employed for antidiabetic, antihypertensive, anti-inflammatory, hepatoprotective, and antimicrobial purposes, as well as for enhancing blood circulation, strengthening bones, cooling the body, and decelerating aging processes [19]. The root of Entada africana is traditionally recognized for its role in treating inflammatory diseases, hepatitis, bronchitis, and cough, in addition to its utility in wound healing, diuresis, and as an anti-gonococcal and anti-syphilitic agent [20]. The bark of Intsia bijuga is associated with the management of rheumatism, dysentery, urinary tract infections, asthma, diabetes, ulcers, and skeletal fractures [21]. Robinia pseudoacacia bark is traditionally utilized for its laxative, antispasmodic, and diuretic properties [22]. Collectively, the data underscore the ethnomedical significance of these species and highlight their diverse pharmacological potential within traditional healing systems.
Thus, Table 1 synthesizes current knowledge regarding the botanical sources, tissue-specific localization, and geographic distribution of robinetin, highlighting its role as both a taxonomic marker and bioactive metabolite in various plant species worldwide.

4. Techniques for the Analysis of Robinetin in Plant Material

The extraction of robinetin from Robinia pseudoacacia wood was investigated using three conventional methods: Soxhlet extraction, ultrasonic extraction, and maceration with stirring. The solvents used for the extraction were acetone, methanol and ethanol, each containing 10% of water (v/v). Among these, Soxhlet extraction proved to be the most efficient, yielding the highest concentrations of robinetin as well as the related flavanonol dihydrorobinetin. In contrast, ultrasonic extraction and maceration provided lower yields overall, although ultrasonic extraction achieved relatively high recovery in a shorter time, making it advantageous in terms of reduced solvent and energy consumption. With respect to solvents, the amounts of robinetin extracted did not differ substantially between acetone, ethanol, and methanol. Overall, Soxhlet extraction with aqueous acetone was identified as the optimal laboratory-scale method for obtaining robinetin [63].
Chromatography currently represents a fundamental technique in the field of separation and analytical sciences, and its widespread adoption across research institutions and the pharmaceutical industry underscores its central role in modern analytical methodologies [64]. Column chromatography was employed to separate robinetin from the leaves of Robinia pseudoacacia. The extraction was carried out using ethanol, followed by purification on Sephadex LH-20. Further separation was performed with reversed-phase (RP) silica gel columns using a methanol–water system (9:1). Finally, a Lobar column with the same solvent system (MeOH:H2O, 9:1) was applied to achieve effective isolation of robinetin [62].
High-performance liquid chromatography (HPLC), unlike TLC, offers several critical advantages in analytical applications, including the ability to perform both qualitative and quantitative determinations with high efficiency, sensitivity, and rapid separation [65]. As a continuously evolving technique, HPLC has demonstrated broad applicability and plays a key role in the analysis of plant-derived extracts and fractions. Its utility in the detection and isolation of robinetin is well-established, with specific analytical conditions outlined in Table 2.
Sanz and co-workers using spectroscopic and spectrometric data of peaks in HPLC chromatograms seasoned and toasted Robinia pseudoacacia heartwood extracts identified robinetin with λmax 254 and 364 nm, MS/MS m/z [M-H]—301, 273, 245, 229, 135, 91 [57].

5. Biological Activities of Robinetin

Obtaining bioactive compounds from medical plants is fundamental to the production of drugs of the plant origin. Thus, Table 3 summarizes biological activities of robinetin in vitro, in vivo and in silico studies. Considering performed spectroscopic analyses, complemented by molecular modeling calculations, provide important insights into robinetin binding within a human serum albumin (HSA) matrix, highlighting its potential utility as a sensitive multiwavelength fluorescent probe. Such studies are expected to facilitate the rational screening and design of structurally optimized flavonoid-based compounds, representing a promising class of phytochemicals with significant potential as alternatives to conventional therapies [66,68]. Another study also showed that robinetin inhibits lipid peroxidation in a dose-dependent manner and prevents non-enzymatic glycosylation of hemoglobin [69].

5.1. Antiviral Activity

To study the antiviral activity, Fesen and co-workers investigated the inhibition of HIV-1 integrase by robinetin. It was revealed that robinetin inhibited 3′—processing and strand transfer with IC50 = 5.9 ± 1.9 µM and 1.6 ± 0.7 µM [70].
Moreover, robinetin has been reported to exhibit antiviral potential against SARS-CoV-2 based on in silico molecular docking studies, demonstrating a binding affinity of −8.3 kcal/mol for the main protease (Mpro) and −7.6 kcal/mol for the spike glycoprotein. These interactions suggest potential inhibition of viral replication and interference with host cell entry mechanisms. The docking results indicate that robinetin forms up to five hydrogen bonds and multiple hydrophobic contacts within the active sites of these proteins, stabilizing the ligand–protein complex [73].

5.2. Antibacterial Activity

Robinetin was also studied as an antibacterial agent against human skin bacterium bacteria Proteus vulgaris and Staphylococcus aureus, where MIC after its exposure gain 100 µg/mL in both bacterium [75].

5.3. Antiparasitic Activity

To study the antiparasitic activity of robinetin, Tasdemir et al. performed in vitro assay using Leishmania donovani strain MHOM/ET/67/L82, Trypanosoma brucei rhodesiense SSTIB 900, and Trypanosoma cruzi C2C4 with IC50 values of 5.9 µg/mL, 5.3 µg/mL and above 30 µg/mL, respectively [77].

5.4. Antioxidant Activity

Using the density functional theory (DFT) with B3LYP functional and 6–311++G (d, p) basis set, the antiradical activity of robinetin has been investigated. Calculations were carried out both in the gas phase and with consideration of solvent effects: water, dimethyl sulfoxide (DMSO), methanol, and benzene. Three mechanisms were examined—hydrogen atom transfer (HAT), single-electron transfer followed by proton transfer (SET-PT), and sequential proton loss electron transfer (SPLET)—to evaluate radical scavenging ability and identify the most likely antioxidant action pathway. Final findings indicated that the 4′-OH group is the most reactive site. Among the evaluated pathways, the HAT mechanism proved to be the most energetically favorable [78,93]. According to Huguet and co-workers, robinetin possesses superoxide scavenging activity comparable to that of previously known scavengers of superoxide anions, such as morin, rutin or luteolin [94].

5.5. Anticancer Activity

Numerous studies have explored the potential of robinetin as an anticancer agent. Unfortunately, this polyhydroxylated flavonoid exhibits limited activity against two human melanoma cell lines, C32 and A375, as the IC50 values in both MTT and NRU assays after 24, 48, and 72 h of incubation exceeded the highest tested concentration (200 µM) [81]. The effect of robinetin (6.25–200 µM) was also examined in an oral squamous carcinoma (SCC-25) cell line. In both, MTT and NRU tests IC50 was above 200 µM [80].
Robinetin was evaluated for their inhibitory efficacy against CDK1 (cyclin-dependent kinase 1) through molecular docking studies. CDK1, a key regulator of the cell cycle component central to the uncontrolled proliferation of malignant cells, has been reportedly implicated in colorectal cancer. The authors predicted that robinetin may be active in terms of carcinogenicity and mutagenicity [82].
Chang et al. applied robinetin topically at a dose of 2.5 µmol. It exhibited no intrinsic tumor-initiating activity on mouse skin. When administered 5 min prior to 200 nmol of B[a]P 7,8-diol-9,10-epoxide-2 (B[a]P), it resulted in only a 16–24% reduction in tumor incidence, which was not statistically significant. Similarly, robinetin showed little or no effect on the tumor-initiating activity of 50 nmol of B[a]P. In contrast, systemic administration of 1.4 µmol intraperitoneally to preweaning mice did not induce tumors over a 9–11 month observation period, but pretreatment before exposure to 30 nmol of diol-epoxide produced a substantial 44–75% inhibition of pulmonary tumor formation, indicating partial chemopreventive potential [83].
In another study, robinetin was tested for its effect on melanogenesis in HMV II human melanoma cells. The compound showed no melanogenesis-promoting activity under the experimental conditions [95].
Robinetin was tested for its ability to modulate multidrug resistance in Colo 320 human colon cancer cells expressing MDR1/LRP. The compound exhibited only marginal effects on Rhodamine 123 accumulation, indicating limited potential in reversing MDR (multi-drug resistance). Moreover, robinetin acted as a weak apoptosis inducer in both drug-resistant and drug-sensitive colon cancer cells, without significant differences between the two cell lines [84]. Also, in SW480 and T84 colon carcinoma cells robinetin possesses weak anticancer activity (IC50 = 100 µM) [85].

5.6. Anti-Mutagenic Activity

Anti-mutagenic activity was evaluated using the Salmonella typhimurium mutagenicity assay by assessing the ability of the test compounds to inhibit mutation rates induced by known chemical mutagens, including methyl-nitrosourea (MNU), methyl-n-nitro-N-nitrosoguanidine (MNNG), benzo-γ-pyrene (BaP), and 2-aminoanthracene (2-AA). Robinetin caused 11%, 6%, 1.2% and 87% inhibition of mutagenicity induced by MNU, MNNG, BaP and 2-AA, respectively [86].
Since aflatoxin B1 is highly toxic and mutagenic substance, Bhattacharya and Firozi evaluated the effect of robinetin and other flavonoids on microsome catalyzed reactions of aflatoxin B1, leading to activation and DNA adduct formation. According to their results, robinetin is one of the most active compounds in inhibiting microsome-mediated activation of aflatoxin B1 (11.4% of control) and subsequent DNA adduct formation (7.7% of control). Its strong activity suggests that robinetin may counteract the carcinogenic effects of aflatoxin B1 through modulation of microsomal enzyme function [87].
Robinetin was tested for mutagenicity in the Salmonella/mammalian microsome assay using multiple tester strains. It exhibited low frameshift mutagenic activity, with a revertant value of 0.06 revertants/nmol in strain TA98, indicating weak mutagenic potential compared to quercetin and kaempferol. Like other flavonols, robinetin showed significant activation by Aroclor 1254-induced rat-liver microsome preparations. These findings suggest that the mutagenic activity of robinetin is modest and dependent on microsomal enzyme-mediated activation [96]. In another study, robinetin effectively reduced mutagenicity induced by tert-butyl hydroperoxide (BHP) and cumene hydroperoxide (CHP) in Salmonella typhimurium TA102. Its antimutagenic activity was comparable to that of other polyhydroxylated flavonols such as quercetin, fisetin, and myricetin, with ID50 values ranging from 0.25 to 1.05 μmol per plate. Structural modifications, such as hydrogenation, led to a marked loss of antimutagenic potential, highlighting the importance of the double bond between carbons 2 and 3. These findings indicate that the protective effects of robinetin are largely attributable to its radical scavenging properties [97].

5.7. Anti-Necroptosis Activity

The anti-necroptotic potential of flavonoid inhibitors was assessed by measuring cell viability following TSZ-induced necroptosis. Robinetin demonstrated a significant cytoprotective effect, markedly enhancing cell viability at a concentration of 50 μM. In HT-29 cells, robinetin restored survival to levels comparable to those observed in untreated controls. Notably, robinetin maintained its anti-necroptotic efficacy across a range of TSZ-induced necroptotic conditions in both mouse embryonic fibroblasts (MEFs) and HT-29 cells, with observed survival rates ranging from 18% to 56%, corresponding to baseline cell viability. These findings indicate that robinetin confers potent protective effects against necroptotic cell death in vitro. Additionally, ADP-Glo assay revealed RIPK1 inhibition by robinetin (IC50 = 43.8 μM) [88].

5.8. Enzyme Inhibition Activity

Acetylcholinesterase inhibitory activity can be a useful tool in the treatment of Alzheimer’s disease (AD). Robinetin was evaluated for its inhibitory activity using an in vitro assay, revealing a half-maximal inhibitory concentration (IC50) of 456.48 ± 2.57 µM [89]. In another study Hanaki and co-workers identify bioactive natural products in crude drugs that inhibit Aβ42 aggregation and that could be applied to future AD therapies. They selected 6 flavonoids, including robinetin, which suppressed Aβ42 aggregation [44].
Robinetin demonstrated strong inhibitory activity against both MRP1 and MRP2 in MDCKII transfected cells. The compound exhibited IC50 values of 13.6 µM for MRP1 and 15.0 µM for MRP2, indicating comparable potency toward both transporters. Detailed kinetic analysis revealed that robinetin acts as a competitive inhibitor with respect to calcein, with apparent inhibition constants (Ki) of 5.0 µM for MRP1 and 8.5 µM for MRP2. These findings highlight robinetin as a potent and structurally significant flavonoid in modulating multidrug resistance protein activity [91]. Robinetin also possesses inhibitory potency against NADH-oxidase with IC50 value 19 nmol/mg protein [98].

5.9. Activity in the Liver Diseases

To evaluate the ameliorating effect of robinetin on the significant pathogenic features of metabolic failure in the liver and to identify the underlying molecular mechanism, Song et al. used AML-12 hepatocytes untreated and treated with OPA mixture (800 μM oleic acid and 150 μM palmitic acid) and five-week-old male C57/BL6 mice (divided into 4 groups: control, Western diet-feed group (WD), WD with 0.025% robinetin group, and WD with 0.05% robinetin. The researchers proved that robinetin caused suppression of TG accumulation in AML-12 cells, by abrogating the mRNA expression of lipogenesisrelated genes. Also, the liver tissues in the robinetin-supplemented mice exhibited reduced lipid droplet accumulation compared with that of the WD-fed (Western diet) mice. The OGTT results revealed that the blood glucose levels in robinetin-supplemented groups were lower after oral glucose loading. On the other hand, in the mice model it was established that the liver size in mice fed the robinetin-supplemented diet was similar to that in mice in the control group. Its strong activity is associated with the presence of adjacent tri-hydroxyl groups, which promote a substantial rate of auto-oxidation. During cyanide-stimulated oxidation, robinetin contributed to the non-enzymatic generation of superoxide anions, a process attenuated by superoxide dismutase [91].
Summarizing, the information about biological activity of robinetin, the available data are derived from highly heterogeneous experimental models, which complicates direct comparison of reported activities. Many studies rely on biochemical assays that yield low-micromolar IC50 values, whereas cell-based evaluations often indicate only weak effects at much higher concentrations (>200 µM), likely reflecting differences in compound solubility, stability, membrane permeability, and protein binding. Computational approaches (docking, DFT) offer mechanistic insights but require validation with orthogonal biochemical and cellular assays. Finally, very few studies have incorporated pharmacokinetic or metabolism data; yet microsome-dependent activation results suggest that metabolites may substantially alter biological outcomes. Together, these factors underscore the need for standardized protocols, careful control of redox conditions, and integrated biochemical, cellular, and in vivo studies to more reliably define the pharmacological potential of robinetin.

6. ADMET of Robinetin

Analysis of urine samples from rats administered robinetin (200 mg/rat) as well as ether extracts from incubation mixtures (10 mg/tube), demonstrated the presence of substantial amounts of unmetabolized robinetin. It is worth mentioning that in this study the metabolism of tricine and morin was also examined; in the sample with robinetin, no metabolites formed after the metabolism of this flavonoid were found. The results indicate that robinetin is resistant to bacterial catabolism [99].
Additionally, through Deep-PK Predictions analysis, the ADMET (absorption, distribution, metabolism, excretion and toxicity) properties of robinetin can be computationally estimated [100]. The selected pharmacokinetic parameters and toxicity properties of robinetin (C1=CC2=C(C=C1O)OC(=C(C2=O)O)C3=CC(=C(C(=C3)O)O)O) are given in Table 4.
The clinical use of robinetin may be significantly limited by its poor oral bioavailability and short half-life. Predictive data indicate that its oral bioavailability in humans is low, suggesting limited systemic exposure following oral administration. This reduced absorption, combined with a predicted half-life of less than 3 h, means that therapeutic plasma concentrations may not be maintained over long periods, necessitating frequent dosing. Such pharmacokinetic limitations may hinder patient compliance and reduce the overall therapeutic efficacy of the compound. To address these challenges, several strategies could be considered. Formulation approaches, such as the use of nanoemulsions, liposomes, or solid lipid nanoparticles, may enhance solubility, protect the compound from premature metabolism, and improve gastrointestinal absorption. Co-administration with permeability enhancers or inhibitors of efflux transporters could also be explored to increase oral exposure. Alternatively, chemical derivatization, including the development of prodrugs or structural analogs with improved pharmacokinetic properties, could extend systemic half-life and bioavailability. Furthermore, long-acting delivery systems (e.g., polymer-based depots or sustained-release formulations) may help overcome the rapid clearance. Overall, rational optimization through formulation science and medicinal chemistry will be essential to unlock the therapeutic potential of robinetin [101].

7. Conclusions

Robinetin, a polyhydroxylated flavonol, emerges as a bioactive compound of significant pharmacological relevance, supported by evidence from in vitro, in vivo, and in silico studies. Its wide spectrum of activities, including antiviral, antibacterial, antiparasitic, antioxidant, anti-mutagenic, anti-necroptotic, and enzyme-inhibitory effects, highlights its multifaceted therapeutic potential. Despite its relatively modest anticancer efficacy in certain cell models, robinetin demonstrates promising activity in modulating multidrug resistance proteins and protecting against liver metabolic failure, suggesting value in chemopreventive and metabolic disease contexts. Structural features, particularly hydroxyl group substitutions, play a decisive role in its redox activity, metal chelation, and radical scavenging capacity, underscoring the importance of structure–activity relationships. However, limitations such as poor water solubility and limited pharmacokinetic data constrain its direct application. Thus, future research on robinetin should prioritize comprehensive in vivo pharmacokinetic studies to validate current computational predictions and clarify its absorption, metabolism, and systemic distribution. Structural modifications or formulation strategies, such as prodrug design or nanocarrier-based delivery, are needed to improve its poor aqueous solubility and oral bioavailability. Additionally, investigations into synergistic interactions with other flavonoids could enhance its antioxidant, antiviral, or anticancer efficacy. Finally, preclinical and translational studies are essential to elucidate molecular mechanisms and assess the therapeutic potential of robinetin in clinical settings.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Acknowledgments

During the preparation of this manuscript, the author used ACD/ChemSketch for the purposes of drawing the robinetin structure. Deep-PK (Biosig Lab) was used for deep learning for robinetin pharmacokinetic and toxicity prediction. The author have reviewed and edited the output and take full responsibility for the content of this publication.

Conflicts of Interest

The author declares no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
PHFpolyhydroxylated flavonoids
HIVhuman immunodeficiency virus
TGtriglyceride
HOMA-IRhomeostatic model assessment of insulin resistance
HAThistone acetyltransferase
SET-PTsingle electron transfer followed by proton transfer
SPLETsequential proton loss electron transfer
MTT3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide
NRUneutral red uptake
DFTdensity functional theory
Mpromain protease
DMSOdimethyl sulfoxide
HSAhuman serum albumin
ADAlzheimer disease
MeOHmethanol
HPLChigh-performance liquid chromatography
TLCthin-layer chromatography
Et2Odiethyl ether
EtOAcethyl acetate
PhAphosphoric acid
ACNacetonitrile
NH4OACammonium acetate
FAaormic acid
MDRmultidrug-resistant
MNUmethylnitrosourea
MNNGmethyl-n-nltro-N-nitrosoguanidine
BaPbenzo-α-pyrene
2-AA2-aminoanthracene
TPA12-O-tetradecanoylphorbol-13-acetate
FRETfluorescence resonance energy transfer
MICminimum inhibitory concertation
CDK1cyclin-dependent kinase 1
BHPtert-butyl hydroperoxide
CHPcumene hydroperoxide
MEFmouse embryonic fibroblasts
WDwestern diet
OGTToral glucose tolerance test
ADMETabsorption, distribution, metabolism, excretion and toxicity

References

  1. Li, C.; Dai, T.; Chen, J.; Chen, M.; Liang, R.; Liu, C.; Du, L.; McClements, D.J. Modification of flavonoids: Methods and influences on biological activities. Crit. Rev. Food Sci. Nutr. 2022, 63, 10637–10658. [Google Scholar] [CrossRef] [PubMed]
  2. Ayyanna, C.; Kuppusamy, S.; Kumar, P.P. An overview of pharmacological activities and beneficial effects of 3-hydroxyflavone. Macromol. Symp. 2024, 413, 2300078. [Google Scholar] [CrossRef]
  3. Wang, T.-Y.; Li, Q.; Bi, K.-S. Bioactive flavonoids in medicinal plants: Structure, activity and biological fate. Asian J. Pharm. Sci. 2018, 13, 12–23. [Google Scholar] [CrossRef] [PubMed]
  4. Dias, M.C.; Pinto, D.C.G.A.; Silva, A.M.S. Plant flavonoids: Chemical characteristics and biological activity. Molecules 2021, 26, 5377. [Google Scholar] [CrossRef]
  5. Tsimogiannis, D.; Samiotaki, M.; Panayotou, G.; Oreopoulou, V. Characterization of flavonoid subgroups and hydroxy substitution by HPLC-MS/MS. Molecules 2007, 12, 593–606. [Google Scholar] [CrossRef]
  6. Walle, T. Absorption and metabolism of flavonoids. Free. Radic. Biol. Med. 2004, 36, 829–837. [Google Scholar] [CrossRef] [PubMed]
  7. Lomozová, Z.; Hrubša, M.; Conte, P.F.; Papastefanaki, E.; Moravcová, M.; Catapano, M.C.; Proietti Silvestri, I.; Karlíčková, J.; Kučera, R.; Macáková, K.; et al. The effect of flavonoids on the reduction of cupric ions, the copper-driven fenton reaction and copper-triggered haemolysis. Food Chem. 2022, 394, 133461. [Google Scholar] [CrossRef]
  8. Lomozová, Z.; Catapano, M.C.; Hrubša, M.; Karlíčková, J.; Macáková, K.; Kučera, R.; Mladěnka, P. Chelation of iron and copper by quercetin b-ring methyl metabolites, isorhamnetin and tamarixetin, and their effect on metal-based fenton chemistry. J. Agric. Food Chem. 2021, 69, 5926–5937. [Google Scholar] [CrossRef]
  9. Spiegel, M.; Andruniów, T.; Sroka, Z. Flavones’ and flavonols’ antiradical structure– activity relationship—A quantum chemical study. Antioxidants 2020, 9, 461. [Google Scholar] [CrossRef]
  10. Heim, K.E.; Tagliaferro, A.R.; Bobilya, D.J. Flavonoid antioxidants: Chemistry, metabolism and structure-activity relationships. J. Nutr. Biochem. 2002, 13, 572–584. [Google Scholar] [CrossRef]
  11. Guharay, J.; Sengupta, P.K. Excited-state proton-transfer and dual fluorescence of robinetin in different environments. Spectrochim. Acta A Mol. Biomol. Spectrosc. 1997, 53, 905–912. [Google Scholar] [CrossRef]
  12. Faisal Hayat, M.; Ur Rahman, A.; Tahir, A.; Batool, M.; Ahmed, Z.; Atique, U. Palliative potential of robinetin to avert polystyrene microplastics instigated pulmonary toxicity in rats. J. King Saud Univ. Sci. 2024, 36, 103348. [Google Scholar] [CrossRef]
  13. Li, H.; Sun, M.; Lei, F.; Liu, J.; Chen, X.; Li, Y.; Wang, Y.; Lu, J.; Yu, D.; Gao, Y.; et al. Methyl rosmarinate is an allosteric inhibitor of SARS-CoV-2 3 CL protease as a potential candidate against SARS-Cov-2 infection. Antivir. Res. 2024, 224, 105841. [Google Scholar] [CrossRef] [PubMed]
  14. Uzelac, M.; Sladonja, B.; Šola, I.; Dudaš, S.; Bilić, J.; Famuyide, I.M.; McGaw, L.J.; Eloff, J.N.; Mikulic-Petkovsek, M.; Poljuha, D. Invasive alien species as a potential source of phytopharmaceuticals: Phenolic composition and antimicrobial and cytotoxic activity of Robinia pseudoacacia L. leaf and flower extracts. Plants 2023, 12, 2715. [Google Scholar] [CrossRef]
  15. Ogawa, S.; Yazaki, Y. Tannins from Acacia Mearnsii De Wild. Bark: Tannin determination and biological activities. Molecules 2018, 23, 837. [Google Scholar] [CrossRef] [PubMed]
  16. Balkrishna, A.; Sakshi; Chauhan, M.; Dabas, A.; Arya, V. A Comprehensive insight into the phytochemical, pharmacological potential, and traditional medicinal uses of Albizia lebbeck (L.) Benth. eCAM 2022, 2022, 5359669. [Google Scholar] [CrossRef]
  17. Geronço, M.S.; Melo, R.C.; Mendes Barros, H.L.; Aquino, S.R.; Evangelista de Oliveira, F.d.C.; Islam, M.T.; do Ó Pessoa, C.; dos Santos Rizzo, M.; da Costa, M.P. Advances in the research of Adenanthera pavonina: From traditional use to intellectual property. J. Med. Plants Res. 2020, 14, 13–23. [Google Scholar] [CrossRef]
  18. Oza, M.J.; Kulkarni, Y.A. Traditional uses, phytochemistry and pharmacology of the medicinal species of the genus Cordia (Boraginaceae). J. Pharm. Pharmacol. 2017, 69, 755–789. [Google Scholar] [CrossRef]
  19. Ahda, M.; Jaswir, I.; Khatib, A.; Ahmed, Q.U.; Syed Mohamad, S.N.A. A review on Cosmos caudatus as a potential medicinal plant based on pharmacognosy, phytochemistry, and pharmacological activities. Int. J. Food Prop. 2023, 26, 344–358. [Google Scholar] [CrossRef]
  20. Germanò, M.P.; Certo, G.; D’Angelo, V.; Sanogo, R.; Malafronte, N.; De Tommasi, N.; Rapisarda, A. Anti-angiogenic activity of Entada africana root. Nat. Prod. Res. 2015, 29, 1551–1556. [Google Scholar] [CrossRef]
  21. Koch, M.; Kehop, D.A.; Kinminja, B.; Sabak, M.; Wavimbukie, G.; Barrows, K.M.; Matainaho, T.K.; Barrows, L.R.; Rai, P.P. An ethnobotanical survey of medicinal plants used in the east sepik province of Papua New Guinea. J. Ethnobiol. Ethnomed. 2015, 11, 79. [Google Scholar] [CrossRef]
  22. Wang, L.; Waltenberger, B.; Pferschy-Wenzig, E.M.; Blunder, M.; Liu, X.; Malainer, C.; Blazevic, T.; Schwaiger, S.; Rollinger, J.M.; Heiss, E.H.; et al. Natural product agonists of peroxisome proliferator-activated receptor gamma (PPARγ): A review. Biochem. Pharmacol. 2014, 92, 73–89. [Google Scholar] [CrossRef]
  23. Roux, D.G. Recent advances in the chemistry and chemical utilization of the natural condensed tannins. Phytochemistry 1972, 11, 1219–1230. [Google Scholar] [CrossRef]
  24. Olajuyigbe, O.O.; Afolayan, A.J. Pharmacological assessment of the medicinal potential of acacia mearnsii de wild: Antimicrobial and toxicity activities. Int. J. Mol. Sci. 2012, 13, 4255–4267. [Google Scholar] [CrossRef]
  25. Phoraksa, O.; Vongthip, W.; Juntranggoor, P.; Maiuthed, A.; Tuntipopipat, S.; Charoenkiatkul, S.; Tencomnao, T.; Muangnoi, C.; Sukprasansap, M. Bioavailable fraction from the edible leaf of Albizia lebbeck (L.) Benth. inhibits neurotoxicity in human microglial HMC3 cells and promotes lifespan in Caenorhabditis elegans. Int. J. Food Sci. Technol. 2025, 60, vvae067. [Google Scholar] [CrossRef]
  26. El-Shazly, M.; Mansour, M.; Elsayed, E.; Elgindi, M.; Singab, A.-N. Characterization of essential oils of the leaves and fruits of Adenanthera pavonina L. by GC/MS. APSU-ASU 2020, 4, 63–69. [Google Scholar] [CrossRef]
  27. Malan, J.C.S.; Young, D.A.; Steenkamp, J.A.; Ferreira, D. Oligomeric flavanoids. Part 2. The first profisetinidins with dihydroflavonol constituent units. J. Chem. Soc. 1988, 9, 2567–2572. [Google Scholar] [CrossRef]
  28. Miao, Y.; Liu, W.; Alsallameh, S.M.S.; Albekairi, N.A.; Muhseen, Z.T.; Butch, C.J. Unraveling Cordia myxa’s anti-malarial potential: Integrative insights from network pharmacology, molecular modeling, and machine learning. BMC Infect. Dis. 2024, 24, 1180. [Google Scholar] [CrossRef] [PubMed]
  29. Sumarno, L.; Hendrawati, Y.T.; Ramadhan, I.A.; Yustinah; Hasyim, H.U.; Nugrahani, A.R.; Novika, H.R.G.; Setianto, W.B.; Nasruddin; Yohanes, H. Separation and characterization of kenikir leaf extract (Cosmos caudatus Kunt) in potential as a natural antioxidant. API Conf. Proc. 2025, 3166, 020013. [Google Scholar] [CrossRef]
  30. Dewi, R.T.; Ekapratiwi, Y.; Sundowo, A.; Ariani, N.; Yolanda, T.; Filaila, E. Bioconversion of quercetin glucosides from Dendrophthoe pentandra leaf using Aspergillus acueletus LS04-3. AIP Conf. Proc. 2019, 2175, 020048. [Google Scholar] [CrossRef]
  31. Hashem, F.A. Phenolic compounds of Erucaria microcarpa Boiss. and their effect as scavengers for singlet oxygen. J. Herbs. Spices Med. Plants 2006, 12, 27–41. [Google Scholar] [CrossRef]
  32. Jurd, L.; Manners, G.D. Isoflavene, isoflavan, and flavonoid constituents of Gliricidia speium. J. Agric. Food Chem. 1977, 25, 723–726. [Google Scholar] [CrossRef]
  33. Wafaey, A.A.; El-Hawary, S.S.; Kirollos, F.N.; Abdelhameed, M.F. An overview on Gliricidia sepium in the pharmaceutical aspect: A review article. Egypt. J. Chem. 2023, 66, 479–496. [Google Scholar] [CrossRef]
  34. Astuti, S.D.; Widyobroto, B.P.; Agus, A.; Yusiati, L.M. Identification of galactogogues in Gliricidia maculata. IOP Conf. Ser. Earth Environ. Sci. 2019, 387, 012119. [Google Scholar] [CrossRef]
  35. Wahman, R.; Moser, S.; Bieber, S.; Cruzeiro, C.; Schröder, P.; Gilg, A.; Lesske, F.; Letzel, T. Untargeted analysis of Lemna minor metabolites: Workflow and prioritization strategy comparing highly confident features between different mass spectrometers. Metabolites 2021, 11, 832. [Google Scholar] [CrossRef]
  36. Sosa, D.; Alves, F.M.; Prieto, M.A.; Pedrosa, M.C.; Heleno, S.A.; Barros, L.; Feliciano, M.; Carocho, M. Lemna Minor: Unlocking the value of this duckweed for the food and feed industry. Foods 2024, 13, 1435. [Google Scholar] [CrossRef]
  37. Hillis, W.E. Formation of robinetin crystals in vessels of Intsia species. IAWA J. 1996, 17, 405–419. [Google Scholar] [CrossRef]
  38. Angio, M.H.; Renjana, E.; Firdiana, E.R. Morphology characterization and phytochemical overview of the moluccan ironwood Intsia bijuga (Colebr.) kuntze, a living collection of purwodadi botanic garden, Indonesia. J. Threat. Taxa 2022, 14, 21853–21861. [Google Scholar] [CrossRef]
  39. Norscia, I.; Borgognini-Tarli, S.M. Ethnobotanical reputation of plant species from two forests of madagascar: A preliminary investigation. South Afr. J. Bot. 2006, 72, 656–660. [Google Scholar] [CrossRef]
  40. Sari, R.K.; Prayogo, Y.H.; Sari, R.A.L.; Asidah, N.; Rafi, M.; Wientarsih, I.; Darmawan, W. Intsia bijuga heartwood extract and its phytosome as tyrosinase inhibitor, antioxidant, and sun protector. Forests 2021, 12, 1792. [Google Scholar] [CrossRef]
  41. Hawthorne, B.J.; Morgan, J.W.W. The colouring matter from Millettia stuhlamannii. Chem. Ind. 1962, 33, 1504–1505. [Google Scholar]
  42. Hong, Z.; Xie, J.; Hu, H.; Bai, Y.; Hu, X.; Li, T.; Chen, J.; Sheng, J.; Tian, Y. Hypoglycemic effect of Moringa oleifera leaf extract and its mechanism prediction based on network pharmacology. J. Futur. Foods 2023, 3, 383–391. [Google Scholar] [CrossRef]
  43. Liu, R.; Liu, J.; Huang, Q.; Liu, S.; Jiang, Y. Moringa oleifera: A systematic review of its botany, traditional uses, phytochemistry, pharmacology and toxicity. J. Pharm. Pharmacol. 2022, 74, 296–320. [Google Scholar] [CrossRef]
  44. Hanaki, M.; Murakami, K.; Gunji, H.; Irie, K. Activity-differential search for amyloid-β aggregation inhibitors using LC-MS combined with principal component analysis. Bioorg. Med. Chem. Lett. 2022, 61, 128613. [Google Scholar] [CrossRef]
  45. Tungmunnithum, D.; Pinthong, D.; Hano, C. Flavonoids from Nelumbo nucifera Gaertn., a medicinal plant: Uses in traditional medicine, phytochemistry and pharmacological activities. Medicines 2018, 5, 127. [Google Scholar] [CrossRef]
  46. Pal, I.; Dey, P. A Review on lotus (Nelumbo nucifera) seed. Intern J. Scien Res. (IJSR) 2013, 14, 2319–7064. [Google Scholar]
  47. Charlesworth, E.H.; Robinson, R. 73. Anthoxanthins. Part XIII. Synthesis of a colouring matter of Robinia pseudoacacia. J. Chem. Soc. 1933, 0, 268–270. [Google Scholar] [CrossRef]
  48. Bostyn, S.; Destandau, E.; Charpentier, J.P.; Serrano, V.; Seigneuret, J.M.; Breton, C. optimization and kinetic modelling of robinetin and dihydrorobinetin extraction from Robinia pseudoacacia wood. Ind. Crop. Prod. 2018, 126, 22–30. [Google Scholar] [CrossRef]
  49. Sanz, M.; De Simón, B.F.; Cadahía, E.; Esteruelas, E.; Muñoz, Á.M.; Hernández, M.T.; Estrella, I. Polyphenolic profile as a useful tool to identify the wood used in wine aging. Anal. Chim. Acta 2012, 732, 33–45. [Google Scholar] [CrossRef]
  50. Cerezo, A.B.; Alvarez-Fernández, M.A.; Hornedo-Ortega, R.; Troncoso, A.M.; García-Parrilla, M.C. Phenolic composition of vinegars over an accelerated aging process using different wood species (acacia, cherry, chestnut, and oak): Effect of wood toasting. J. Agric. Food Chem. 2014, 62, 4369–4376. [Google Scholar] [CrossRef] [PubMed]
  51. Sanz, M.; Fernández de Simón, B.; Esteruelas, E.; Muñoz, Á.M.; Cadahía, E.; Hernández, M.T.; Estrella, I.; Martinez, J. Polyphenols in red wine aged in acacia (Robinia pseudoacacia) and oak (Quercus petraea) wood barrels. Anal. Chim. Acta 2012, 732, 83–90. [Google Scholar] [CrossRef]
  52. Sánchez-Monedero, A.; Cañadas, R.; Martín-Sampedro, R.; Santos, J.I.; Eugenio, M.E.; Ibarra, D. A Sustainable valorization approach of kraft black liquor from Robinia pseudoacacia L.: Lignin precipitation followed by green solvents extraction of phenols from delignified liquor. J. Environ. Manag. 2025, 391, 126515. [Google Scholar] [CrossRef]
  53. Zarev, Y. Isolation and characterization of 3-O-caffeoyloleanolic acid from Robinia pseudoacacia stem bark. Pharmacia 2023, 70, 1209–1212. [Google Scholar] [CrossRef]
  54. Smith, A.L.; Campbell, C.L.; Walker, D.B.; Hanover, J.W. Extracts from black locust as wood preservatives: Extraction of decay resistance from black locust heartwood. Holzforschung 1989, 43, 293–296. [Google Scholar] [CrossRef]
  55. Tian, F.; McLaughlin, J.L. Bioactive flavonoids from the black locust tree, Robinia pseudoacacia. Pharm. Biol. 2000, 38, 229–234. [Google Scholar] [CrossRef]
  56. Destandau, E.; Charpentier, J.P.; Bostyn, S.; Zubrzycki, S.; Serrano, V.; Seigneuret, J.M.; Breton, C. Gram-scale purification of dihydrorobinetin from Robinia pseudoacacia L. wood by centrifugal partition chromatography. Separations 2016, 3, 23. [Google Scholar] [CrossRef]
  57. Sanz, M.; Fernández De Simón, B.; Esteruelas, E.; Muñoz, Á.M.; Cadahía, E.; Hernández, T.; Estrella, I.; Pinto, E. Effect of toasting intensity at cooperage on phenolic compounds in acacia (Robinia pseudoacacia) heartwood. J. Agric. Food Chem. 2011, 59, 3135–3145. [Google Scholar] [CrossRef] [PubMed]
  58. Hofmann, T.; Guran, R.; Zitka, O.; Visi-Rajczi, E.; Albert, L. Liquid chromatographic/mass spectrometric study on the role of beech (Fagus sylvatica L.) wood polyphenols in red heartwood formation. Forests 2022, 13, 10. [Google Scholar] [CrossRef]
  59. Hosseinihashemi, S.K.; Safdari, V.; Kanani, S. Comparative chemical composition of n-hexane and ethanol extractives from the heartwood of black locust. Asian J. Chem. 2013, 25, 929–933. [Google Scholar] [CrossRef]
  60. Madanat, H.M.; Al-Antary, T.M.; Abu Zarga, M. Identification and isolation of the insecticidal compounds from Robinia pseudoacacia L. (Fabaceae). Fresenius Environ. Bull 2018, 27, 1838–1849. [Google Scholar]
  61. Scheidemann, P.; Wetzel, A. Identification and characterization of flavonoids in the root exudate of Robinia pseudoacacia. Trees 1997, 11, 316–321. [Google Scholar] [CrossRef]
  62. Nasir, H.; Iqbal, Z.; Hiradate, S.; Fujii, Y. Allelopathic potential of Robinia pseudo-acacia L. J. Chem. Ecol. 2005, 31, 2179–2192. [Google Scholar] [CrossRef] [PubMed]
  63. Vek, V.; Poljanšek, I.; Oven, P. Efficiency of three conventional methods for extraction of dihydrorobinetin and robinetin from wood of black locust. Eur. J. Wood Wood Prod. 2019, 77, 891–901. [Google Scholar] [CrossRef]
  64. Yandamuri, N.; Nagabathula, S.; Kurra, S.; Batthula, S.; Allada, N.; Bandam, P. Comparative study of new trends in HPLC. Int. J. Pharm. Sci. Rev. Res. 2013, 23, 52–57. [Google Scholar]
  65. Garcia, S.; Heinzen, H.; Martinez, R.; Moyna, P. Identification of flavonoids by TLC scanning analysis. Chromatographia 1993, 35, 430–434. [Google Scholar] [CrossRef]
  66. Wang, Y.; Hu, Y.; Wu, T.; Zhou, X.; Shao, Y. Triggered excited-state intramolecular proton transfer fluorescence for selective triplex DNA recognition. Anal. Chem. 2015, 87, 11620–11624. [Google Scholar] [CrossRef]
  67. Vek, V.; Poljanšek, I.; Oven, P. Variability in content of hydrophilic extractives and individual phenolic compounds in black locust stem. Eur. J. Wood Wood Prod. 2020, 78, 501–511. [Google Scholar] [CrossRef]
  68. Pahari, B.P.; Chaudhuri, S.; Chakraborty, S.; Sengupta, P.K. Ground and excited state proton transfer of the bioactive plant flavonol robinetin in a protein environment: Spectroscopic and molecular modeling studies. J. Phys. Chem. B 2015, 119, 2533–2545. [Google Scholar] [CrossRef]
  69. Chaudhuri, S.; Pahari, B.; Sengupta, B.; Sengupta, P.K. Binding of the bioflavonoid robinetin with model membranes and hemoglobin: Inhibition of lipid peroxidation and protein glycosylation. J. Photochem. Photobiol. B 2010, 98, 12–19. [Google Scholar] [CrossRef]
  70. Fesen, M.R.; Pommier, Y.; Leteurtre, F.; Hiroguchi, S.; Yung, J.; Kohn, K.W. Inhibition of HIV-1 integrase by flavones, caffeic acid phenethyl ester (CAPE) and related compounds. Biochem. Pharmacol. 1994, 48, 595–608. [Google Scholar] [CrossRef] [PubMed]
  71. Cushnie, T.P.T.; Lamb, A.J. Antimicrobial activity of flavonoids. Int. J. Antimicrob. Agents 2005, 26, 343–356. [Google Scholar] [CrossRef]
  72. Kumar, S.; Pandey, A.K. Chemistry and biological activities of flavonoids: An overview. Sci. World J. 2013, 162750, 162750. [Google Scholar] [CrossRef]
  73. Mahmud, S.; Uddin, M.A.R.; Paul, G.K.; Shimu, M.S.S.; Islam, S.; Rahman, E.; Islam, A.; Islam, M.S.; Promi, M.M.; Emran, T.B.; et al. Virtual screening and molecular dynamics simulation study of plant-derived compounds to identify potential inhibitors of main protease from SARS-CoV-2. Brief. Bioinform. 2021, 22, 1402–1414. [Google Scholar] [CrossRef] [PubMed]
  74. Krüger, N.; Kronenberger, T.; Xie, H.; Rocha, C.; Pöhlmann, S.; Su, H.; Xu, Y.; Laufer, S.A.; Pillaiyar, T. Discovery of polyphenolic natural products as SARS-CoV-2 Mpro inhibitors for COVID-19. Pharmaceuticals 2023, 16, 190. [Google Scholar] [CrossRef]
  75. Mori, A.; Nishino, C.; Enoki, N.; Tawata, S. Antibacterial activity and mode of action of plant flavonoids against Proteus vulgaris and Staphylococcus aureus. Phytochemistry 1987, 26, 2231–2234. [Google Scholar] [CrossRef]
  76. Banerjee, A.; Basu, K.; Sengupta, P.K. Effect of β-cyclodextrin nanocavity confinement on the photophysics of robinetin. J. Photochem. Photobiol. B 2007, 89, 88–97. [Google Scholar] [CrossRef]
  77. Tasdemir, D.; Kaiser, M.; Brun, R.; Yardley, V.; Schmidt, T.J.; Tosun, F.; Rüedi, P. Antitrypanosomal and antileishmanial activities of flavonoids and their analogues: In vitro, in vivo, structure-activity relationship, and quantitative structure-activity relationship studies. Antimicrob. Agents Chemother. 2006, 50, 1352–1364. [Google Scholar] [CrossRef]
  78. Menacer, R.; Rekkab, S.; Kabouche, Z. Fisetin and robinetin antiradical activity under solvent effect: Density functional theory study. J. Mol. Model. 2022, 28, 240. [Google Scholar] [CrossRef]
  79. Butkovic, V.; Klasinc, L.; Bors, W. Kinetic study of flavonoid reactions with stable radicals. J. Agric. Food Chem. 2004, 52, 2816–2820. [Google Scholar] [CrossRef]
  80. Jakimiuk, K.; Szoka, Ł.; Surazynski, A.; Tomczyk, M. Anticancer effect of selected flavonoids on tongue squamous cell carcinoma (SCC-25): In vitro study. Acta Pol. Pharm. 2025, 82, 79–89. [Google Scholar] [CrossRef] [PubMed]
  81. Jakimiuk, K.; Szoka, Ł.; Surażyński, A.; Tomczyk, M. Using flavonoid substitution status to predict anticancer effects in human melanoma cancers: An in vitro study. Cancers 2024, 16, 487. [Google Scholar] [CrossRef] [PubMed]
  82. Ogbodo, U.C.; Enejoh, O.A.; Okonkwo, C.H.; Gnanasekar, P.; Gachanja, P.W.; Osata, S.; Atanda, H.C.; Iwuchukwu, E.A.; Achilonu, I.; Awe, O.I. Computational identification of potential inhibitors targeting cdk1 in colorectal cancer. Front. Chem. 2023, 11, 1264808. [Google Scholar] [CrossRef]
  83. Chang, R.L.; Huang, M.T.; Wood, A.W.; Wong, C.Q.; Newmark, H.L.; Yagi, H.; Sayer, J.M.; Jerina, D.M.; Conney, A.H. Effect of ellagic acid and hydroxylated flavonoids on the tumorigenicity of benzo-γ-pyrene and (±)-7β, 8α-dihydroxy-9α, 10α-epoxy-7,8,9,10-Tetrahydrobenzo-α-pyrene on mouse skin and in the newborn mouse. Carcinogenesis 1985, 6, 1127–1133. [Google Scholar] [CrossRef]
  84. Ugocsai, K.; Varga, A.; Molnár, P.; Antus, S.; Molnár, J. Effects of selected flavonoids and carotenoids on drug accumulation and apoptosis induction in multidrug-resistant colon cancer cells expressing MDR1/LRP. In Vivo 2005, 19, 433–438. [Google Scholar]
  85. Richter, M.; Ebermann, R.; Marian, B. Quercetin-induced apoptosis in colorectal tumor cells: Possible role of EGF receptor signaling. Nutr. Cancer 1999, 34, 88–99. [Google Scholar] [CrossRef]
  86. Birt, D.F.; Walker, B.; Tibbels, M.G.; Bresnick, E. Anti-mutagenesis and anti-promotion by apigenin, robinectin and indole-3-carbinol. Carcinogenesis 1986, 7, 1617–1619. [Google Scholar] [CrossRef]
  87. Bhattacharya, R.K.; Firozi, P.F. Effect of plant flavonoids on microsome catalyzed reactions of aflatoxin B1, leading to activation and DNA adduct formation. Cancer Lett. 1988, 39, 85–91. [Google Scholar] [CrossRef]
  88. Xie, H.; Li, W.; Han, X.; Li, M.; Zhao, Q.; Xu, Y.; Su, H.; Meng, W. Identification of RIPK3 as a target of flavonoids for anti-necroptosis in vitro. Bioorg. Chem. 2025, 161, 108503. [Google Scholar] [CrossRef]
  89. Jakimiuk, K.; Nazaruk, D.; Tomczyk, M. Acetylcholinesterase inhibitors: Structure-activity relationship and kinetic studies on selected flavonoids. Acta Pol. Pharm. 2022, 79, 835–840. [Google Scholar] [CrossRef]
  90. Van Zanden, J.J.; Wortelboer, H.M.; Bijlsma, S.; Punt, A.; Usta, M.; Bladeren, P.J.V.; Rietjens, I.M.C.M.; Cnubben, N.H.P. Quantitative structure activity relationship studies on the flavonoid mediated inhibition of multidrug resistance proteins 1 and 2. Biochem. Pharmacol. 2005, 69, 699–708. [Google Scholar] [CrossRef]
  91. Song, J.H.; Kim, H.J.; Lee, J.; Hong, S.P.; Chung, M.Y.; Lee, Y.G.; Park, J.H.; Choi, H.K.; Hwang, J.T. Robinetin alleviates metabolic failure in liver through suppression of P300–CD38 axis. Biomol. Ther. 2024, 32, 214–223. [Google Scholar] [CrossRef]
  92. Añón, M.T.; Ubeda, A.; Alcaraz, M.J.H. Protective effects of phenolic compounds on CcL4-induced toxicity in isolated rat hepatocytes. Z Naturforsch. C 1992, 47, 275–279. [Google Scholar] [CrossRef]
  93. Manrique-de-la-Cuba, M.F.; Gamero-Begazo, P.; Valencia, D.E.; Barazorda-Ccahuana, H.L.; Gómez, B. Theoretical study of the antioxidant capacity of the flavonoids present in the Annona muricata (Soursop) leaves. J. Mol. Model. 2019, 25, 200. [Google Scholar] [CrossRef]
  94. Huguet, A.I.; Máñez, S.; Alcaraz, M.J. Superoxide scavenging properties of flavonoids in a non-enzymic system. Z Naturforsch. C 1990, 45, 19–24. [Google Scholar] [CrossRef]
  95. Takekoshi, S.; Nagata, H.; Kitatani, K. Flavonoids enhance melanogenesis in human melanoma cells. Tokai J. Exp. Clin. Med. 2014, 39, 116–121. [Google Scholar]
  96. Brown, J.P.; Dietrich, P.S. Mutagenicity of plant flavonols in the Salmonella/Mammalian microsome test. activation of flavonol glycosides by mixed glycosidases from rat cecal bacteria and other sources. Mutat. Res. Genet. Toxicol. Environ. Mutagen 1979, 66, 223–240. [Google Scholar] [CrossRef]
  97. Edenharder, R.; Grünhage, D. Free radical scavenging abilities of flavonoids as mechanism of protection against mutagenicity induced by tert-butyl hydroperoxide or cumene hydroperoxide in Salmonella typhimurium TA102. Mutat. Res. Genet. Toxicol. Environ. Mutagen 2003, 540, 1–18. [Google Scholar] [CrossRef]
  98. Hodnick, W.F.; Duval, D.L.; Pardini, R.S. Inhibition of mitochondrial respiration and cyanide-stimulated generation of reactive oxygen species by selected flavonoids. Biochem. Pharmacol. 1994, 47, 573–580. [Google Scholar] [CrossRef] [PubMed]
  99. Griffiths, L.A.; Smith, G.E. Metabolism of myricetin and related compounds in the rat. Metabolite formation in vivo and by the intestinal microflora in vitro. Biochem. J. 1972, 130, 141–151. [Google Scholar] [CrossRef] [PubMed]
  100. Available online: https://biosig.lab.uq.edu.au/deeppk/prediction (accessed on 9 September 2025).
  101. Stielow, M.; Witczyńska, A.; Kubryń, N.; Fijałkowski, Ł.; Nowaczyk, J.; Nowaczyk, A. The Bioavailability of drugs—The current state of knowledge. Molecules 2023, 28, 8038. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Chemical structure of robinetin.
Figure 1. Chemical structure of robinetin.
Ijms 26 09546 g001
Table 1. Plants containing robinetin and its traditional use.
Table 1. Plants containing robinetin and its traditional use.
PlantFamilyPlant PartTraditional UseReferences
Acacia
mearnsii
Fabaceaebarkmicrobial infections[15,23,24]
Albizia
lebbeck
leavestreatment of ulcers, night blindness, respiratory disorders, skin disorders, snake, bite, piles, leprosy gonorrhea, scorpion bite, cough, pharyngitis[16,25]
Adenanthera
pavonina
seeds,
leaves
treatment of boils and inflammations[17,26]
Burkea
africana
barknot found[27]
Cordia
myxa
Boraginaceaenot giventreatment of wound, boils, tumors, gout and ulcer; blood purifier and febrifuge[18,28]
Cosmos
caudatus
Asteraceaeleavesanti-diabetic, anti-hypertensive, anti-inflammatory, hepatoprotective, antimicrobial, blood circulation booster, bone strengthener, body-cooling agent and anti-aging agent[19,29]
Dendrophthoe
pentandra
Loranthaceaeleavesimmunological disorders and cancers[30]
Entada
africana
Fabaceaeroottreatment of inflammatory diseases, hepatitis, bronchitis and cough and wound-healing, diuretic, anti-gonococci and anti-syphilitic agent[20]
Erucaria
microcarpa
Brassicaceaeaerial partsnot found[31]
Gliricidia
sepium
Fabaceaebarkanti-microbial, antibacterial, anti-inflammatory, thrombolytic, antisickling, wound healing agent[32,33]
Gliricidia
maculata
herbal galactogogue[32,34]
Lemna
minor
Araceaeleaveshuman food[35,36]
Intsia
bijuga
Fabaceaebarkrheumatism, dysentery, urinary tract infections, asthma, diabetes, ulcers, fractures[21,37,38,39,40]
Millettia stuhlmanniinot found[41]
Moringa
oleifera
Moringaceaeleavesparalysis, helminthiasis, sores and skin infections[42,43]
Nelumbo
nucifera
Nelumbonaceaeleaveshyperlipidemia, hematemesis, metrorrhagia, fever treatment, release skin inflammatory symptoms[44,45]
seedstissue inflammation, cancer, diuretics, skin
diseases and as poison antidote
[44,46]
rootcirculatory system disorders, diarrhea, insomnia,
fever, body heat imbalance and gastritis
[44,45]
Robinia
pseudoacacia
Fabaceaebarklaxative, antispasmodic, diuretic agent[22,47,48,49,50,51,52,53,54,55,56,57,58,59,60]
leaves[61,62]
Table 2. High-performance liquid chromatography in the identification of robinetin.
Table 2. High-performance liquid chromatography in the identification of robinetin.
ExtractColumnMobile PhaseConditionsRef.
Robinia pseudoacacia
bark
50% MeOHHypersil ODS C180.1% PhA (A)
MeOH; PhA 0.1% (B)
100–95% A: 0–50 min
95–70% A: 50–85 min
70–0% A: 85–105 min
[66]
Et2O, EtOAc
lyophilizates (in 50% MeOH)
[57]
Et2O, EtOAc100–85% A: 0–20 min
85–75% A: 20–30 min
75–50% A: 30–50 min
50–0% A: 50–70 min
90% acetoneThermo Accucore C18H2O + 0.1% FA (A); MeOH + 0.1% FA (B)5–95% of solvent B[67]
Lemna minor
leaves
50% MeOHPoroshell 120 EC-C18;
ZIC-HILIC
10 mM NH4OAC in 9:1 (v/v) H2O–ACN (A); 10 mM NH4OAC in 1:9 (v/v) H2O–ACN (B) (RPLC); ACN (C); H2O (D) (HILIC)
  • Binary pump 1:
100% A: 0–7 min
100–50% A: 7–13 min
50–0% A: 13–33 min
100% A: 33–58 min
  • Binary pump 2:
100% C: 0–7 min
100–60% A: 7–13 min
100% C: 13–53 min
[35]
Intsia bijuga
bark
EtOHAccucore C18 columnH2O + 0.1% FA (A); ACN + 0.1% FA (B)5–25% B: 0–3 min
25–55%: 3–22.5 min
55–95%: 22.5–25 min
95% B: 25–28 min
5% B: 29–30 min
[40]
MeOH—methanol; Et2O—diethyl ether; EtOAc—ethyl acetate; PhA—phosphoric acid, ACN—acetonitrile; NH4OAC—ammonium acetate; FA—formic acid.
Table 3. Biological activities of robinetin in vitro, in vivo and in silico models.
Table 3. Biological activities of robinetin in vitro, in vivo and in silico models.
Experimental ModelExposure/
Incubation
ConcentrationEfficacyRef.
ANTIVIRAL
HIV integrase
catalytic assays
1 h
incubation
4 µL of
sample
  • Cleavage IC50 = 5.9 ± 1.9 µM
  • Integration IC50 = 1.6 ± 0.7 µM
[70,71,72]
SARS-CoV-2
virtual screening
--
  • MM-GBSA ΔGBind = −47.544 kcal/mol
  • Binding affinity (Mpro) = −8.3 kcal/mol
  • Binding affinity (spike glycoprotein) = −7.6 kcal/mol
[73]
SARS-CoV-2 Mpro
inhibition (FRET)
-10 µM
  • % of inhibition at 10 µM: 5.96
  • IC50 > 10
[74]
ANTIBACTERIAL
Proteus vulgaris,
Staphylococcus aureus
1 h
incubation
not given
  • MIC (µg/mL):
  • P. vulgaris: 100
  • S. aureus: 100
[71,75]
ANTIPARASITIC
Leishmania
donovani MHOM/ET/67/L82
72 h
incubation
30 to 0.041 µg/mLIC50 = 5.9 µg/mL[76,77]
Trypanosoma brucei rhodesiense
SSTIB 900
72 h
incubation
90 to 0.123 µg/mlIC50 = 5.3 µg/mL[77]
Trypanosoma cruzi C2C496 h
incubation
not givenIC50 > 30 µg/mL
ANTIOXIDANT
PM7 semiempirical method, HAT, SET-PT, SPLET --4′-OH hydroxyl is the preferred active site, HAT mechanism is energetically the most favored pathway[78]
DPPH assay, BDPA
assay
not givennot givenKinetic data: 1.4 × 102 kF/L mol−1 × s−1 in MeOH and 1:1 H2O/2-propanol (v/v)[79]
ANTICANCER
SCC-25
cell line
24, 48 and 72 h incubation in MTT and NRU tests6.25–200 µM
  • MTT, IC50 > 200 µM
  • NRU, IC50 > 200 µM
[80]
C32
cell line
  • MTT, IC50 > 200 µM
  • NRU, IC50 > 200 µM
[81]
A375
cell line
  • MTT, IC50 = 100–200 µM
  • NRU, IC50 > 200 µM
inhibitory efficacies against CDK1 through molecular docking--stable within the binding pocket of the CDK1 protein[82]
mice with skin
tumors
20 weeks2500 nmolinhibited the number of tumors per mouse by 16–24% after 15–20 weeks of promotion with TPA[83]
CCL-220.1 and CCL-222
cell lines
(Rhodamine 123
accumulation)
20 min followed by 10 min1 mg/mL (stock
solution)
Fluorescence activity ratio (in CCL-222 MDR1/LRP-expressing cells):
  • at 40 µg/mL—0.83
  • at 4 µg/mL—1.10
Apoptosis in MDR1 (CCL-222 cells) at 10 µg/mL:
  • early—6.99%
  • apoptosis—4.25%
  • cell death—1.21%
Apoptosis in CCL-222 cells
  • early—12.99%
  • apoptosis—4.73%
  • cell death—1.95%
[84]
SW480 and T84 cell lines48 h incubationnot givenIC50 = 100 µM[85]
ANTI-MUTAGENIC
Salmonella typhimurium wit mutagenesis induced by MNU, MNNG, BaP, 2-AA48 h incubationnot given% of inhibition:
  • MNU: 11
  • MNNG: 6
  • BaP: 1.2
  • 2-AA: 87
[86]
liver microsomes from rats (inhibition of
aflatoxin B1)
1 h incubationnot givenMicrosome-mediated metabolic activation: 11.4% of control; DNA adduct formation of AFB1: 7.7% of
control.
[87]
ANTI-NECROPTOSIS
MEFs and HT-29 cells18–20 h
incubation
10 µMIC50 = 9.1 µM[88]
ENZYME INHIBITION
acetylcholinesterase15 min
incubation
50 µL of
sample
IC50 = 456.48 ± 2.57 µM[89]
MRP1 and MRP20 and 45 min
incubation
1, 10, 20, 30, 40 and 50 µM of sample
  • inhibition at 25 µM:
  • 75% for MRP1
  • 76% for MRP2
2.
IC50:
  • 13.6 ± 3.9 µM for MRP1
  • 15.0 ± 3.5 µM for MRP2
3.
Ki:
  • 5.0 mM for MRP1
  • 8.5 mM for MRP2
[90]
LIVER DISEASES
AML-12 hepatocytes, C57/BL6 mice
  • 18 h incubation with cells
  • mice: 12 weeks
Wester diet (WD) with 0.025%, 0.05% of
robinetin
  • AML-12 cells:
  • ↓ TG accumulation by downregulating lipogenesis-related transcription factors,
  • ↑CD38 expression in OPA-treated cells thought anti-HAT
2.
In mice:
  • ↓ mass liver,
  • Improved plasma insulin level and HOMA-IR values
  • ↑ CD38 expression thought anti-HAT
  • ↓ blood glucose levels
3.
Computational simulation:
  • dock into the HAT domain pocket of p300, abrogation of its catalytic activity
[91]
liver cells from male Wistar rats10 min prior to toxic challenge2.4%ALT activity: 2.16 IU/g wet cells[92]
TG—triglyceride; HOMA-IR—homeostatic model assessment of insulin resistance; HAT—histone acetyltransferase; HAT—hydrogentom transfer, SET-PT—single-electron transfer followed by proton transfer; SPLET—sequential proton loss electron transfer; MNU—methylnitrosourea; MNNG—methyl-n-nltro-N-nitrosoguanidine; BaP—benzo-α-pyrene; 2-AA—2-aminoanthracene; TPA—12-O-tetradecanoylphorbol-13-acetate; FRET—fluorescence resonance energy transfer.
Table 4. The selected predicted pharmacokinetic and toxicity properties of robinetin.
Table 4. The selected predicted pharmacokinetic and toxicity properties of robinetin.
Property NamePrediction ResultPredictive Confidence A
ABSORPTION
human oral bioavailability 20%non-bioavailablelow
human oral bioavailability 50%bioavailablelow
human intestinal absorptionabsorbedhigh
P-glycoprotein inhibitornon-inhibitorhigh
P-glycoprotein substratenon-substratemedium
DISTRIBUTION
Blood–Brain Barriernon-penetrablehigh
METABOLISM
CYP 1A2 inhibitorinhibitorhigh
CYP 1A2 substratesubstratehigh
CYP 3A4 inhibitornon-inhibitormedium
CYP 3A4 substratenon-substratehigh
OATP1B1non-inhibitormedium
EXCRETION
high-life of drug<3 hlow
TOXICITY
AMES mutagenesissafemedium
eye irritationtoxiclow
carcinogenesissafehigh
skin sensitizationtoxicmedium
A interpretation of predictive confidence available on https://biosig.lab.uq.edu.au/deeppk/theory (accessed on 9 September 2025).
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

Jakimiuk, K. A Comprehensive Review of Robinetin: Distribution, Biological Activity and Pharmacokinetic Parameters. Int. J. Mol. Sci. 2025, 26, 9546. https://doi.org/10.3390/ijms26199546

AMA Style

Jakimiuk K. A Comprehensive Review of Robinetin: Distribution, Biological Activity and Pharmacokinetic Parameters. International Journal of Molecular Sciences. 2025; 26(19):9546. https://doi.org/10.3390/ijms26199546

Chicago/Turabian Style

Jakimiuk, Katarzyna. 2025. "A Comprehensive Review of Robinetin: Distribution, Biological Activity and Pharmacokinetic Parameters" International Journal of Molecular Sciences 26, no. 19: 9546. https://doi.org/10.3390/ijms26199546

APA Style

Jakimiuk, K. (2025). A Comprehensive Review of Robinetin: Distribution, Biological Activity and Pharmacokinetic Parameters. International Journal of Molecular Sciences, 26(19), 9546. https://doi.org/10.3390/ijms26199546

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

Article Metrics

Back to TopTop