Next Article in Journal
Therapeutic Applications of Essential Oils from Native and Cultivated Ecuadorian Plants: Cutaneous Candidiasis and Dermal Anti-Inflammatory Activity
Next Article in Special Issue
Synthetic Studies toward 5,6,7,3′,4′-Monomethoxytetrahydroxyflavones: Synthesis of Pedalitin
Previous Article in Journal
Phytochemistry and Biological Activities of Murraya Species
Previous Article in Special Issue
Preparation, Characterization and Evaluation of Flavonolignan Silymarin Effervescent Floating Matrix Tablets for Enhanced Oral Bioavailability
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Tectorigenin: A Review of Its Sources, Pharmacology, Toxicity, and Pharmacokinetics

1
West China Centre of Excellence for Pancreatitis, Institute of Integrated Traditional Chinese and Western Medicine, Sichuan Provincial Pancreatitis Centre and West China-Liverpool Biomedical Research Centre, West China Hospital, Sichuan University, Chengdu 610041, China
2
Advanced Mass Spectrometry Center, Research Core Facility, Frontiers Science Center for Disease-Related Molecular Network, West China Hospital, Sichuan University, Chengdu 610041, China
3
Proteomics-Metabolomics Platform, Research Core Facility, West China-Washington Mitochondria and Metabolism Centre, Institutes for Systems Genetics, West China Hospital, Sichuan University, Chengdu 610041, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2023, 28(15), 5904; https://doi.org/10.3390/molecules28155904
Submission received: 9 July 2023 / Revised: 26 July 2023 / Accepted: 30 July 2023 / Published: 5 August 2023

Abstract

:
Tectorigenin is a well-known natural flavonoid aglycone and an active component that exists in numerous plants. Growing evidence suggests that tectorigenin has multiple pharmacological effects, such as anticancer, antidiabetic, hepatoprotective, anti-inflammatory, antioxidative, antimicrobial, cardioprotective, and neuroprotective. These pharmacological properties provide the basis for the treatment of many kinds of illnesses, including several types of cancer, diabetes, hepatic fibrosis, osteoarthritis, Alzheimer’s disease, etc. The purpose of this paper is to provide a comprehensive summary and review of the sources, extraction and synthesis, pharmacological effects, toxicity, pharmacokinetics, and delivery strategy aspects of tectorigenin. Tectorigenin may exert certain cytotoxicity, which is related to the administration time and concentration. Pharmacokinetic studies have demonstrated that the main metabolic pathways in rats for tectorigenin are glucuronidation, sulfation, demethylation and methoxylation, but that it exhibits poor bioavailability. From our perspective, further research on tectorigenin should cover: exploring the pharmacological targets and mechanisms of action; finding an appropriate concentration to balance pharmacological effects and toxicity; attempting diversified delivery strategies to improve the bioavailability; and structural modification to obtain tectorigenin derivatives with higher pharmacological activity.

Graphical Abstract

1. Introduction

It is well known that natural products or botanicals are valuable resources in the treatment of dozens of diseases. Up until now, natural products remain an important basis for the development of emerging therapeutic agents [1]. Newman and Cragg (2020) showed that between 1981 and 2019, natural products, natural product derivatives, and botanical drugs accounted for 33.5% of all newly approved drugs [2]. Flavonoids are secondary metabolites of plants that function as signaling molecules, antioxidants, and detoxifying agents, and protect plants against various biotic and abiotic threats [3]. Flavonoids belong to polyphenol groups, and have been found to exert anti-inflammatory [4], antioxidative [5], antibacterial [6], antidiabetic [7], neuroprotective [8], anticancer, and other activities [9,10,11,12]. Therefore, they have the potential to be used in several fields, including medicine [13], nutraceuticals as dietary supplements [14], as well as the food and cosmetic industries as biopreservatives for their antibacterial activity together with antioxidant potential [15,16].
Tectorigenin is an isoflavone that exists in numerous plant resources, and it is especially abundant in Belamcandae Rhizoma and Puerariae flos [17,18], which both have the effects of clearing away heat, removing toxic substances, relieving sore throat, and reducing swelling [19,20]. As one of the most important active ingredients, the pharmacological effects of tectorigenin have been demonstrated by an increasing number of investigators [21,22,23,24]. Due to its extensive pharmacological activities, including anticancer, antioxidation, hepatoprotection, anti-inflammation, etc., tectorigenin has received a large amount of attention [21,22,23,24]. This paper mainly reviews the research on the sources, the extraction and synthesis, the pharmacology, the toxicity, the pharmacokinetics, and the delivery strategy of tectorigenin in order to assist researchers to better understand and make use of it.

2. Sources of Tectorigenin

Tectorigenin, a methoxy isoflavone with three hydroxyl groups, is chemically noted as 5,7-dihydroxy-3-(4-hydroxyphenyl)-6-methoxy-4H-chromen-4-one (Figure 1). As a bioactive compound, tectorigenin is found in several plant species, most notably in the rhizomes of Belamcanda chinensis (L.) DC., a traditional Chinese medicine (TCM) named Belamcandae Rhizoma [25,26]. The Iris family is the major source of tectorigenin, and many Iridaceous plants, such as Iris spuria L. (Calizona) [27], Iris tectorum Maxim [28], Iris japonica Thunb. [29,30], Iris dichotoma Pall. [31], Iris germanica L. [32], Iris unguicularis Poiret [33], Iris loczyi Kan. [33], Iris kashmiriana Baker [34], Iris crocea Jacq. ex R. C. Foster [35,36], Iris ensata Thunb. [36], Iris germanica L. [37], Iris hungarica Waldst. et Kit. [38], Iris confusa Sealy [37], and Iris pseudacorus L. [37], contain abundant tectorigenin, so tectorigenin is also known as iris flavone [39]. Tectorigenin also exists in some leguminous plants, such as the flowers of Pueraria lobata (Willd.) Ohwi [40,41], Pueraria thomsonii Benth. [18,40], and Pueraria thunbergiana Benth. [42], which are also TCM called Puerariae Flos. Tectorigenin was isolated from the leaves of Dalbergia odorifera T. Chen [43], the heartwood of Dalbergia parviflora Roxb. [44], and the roots of Euchresta formosana (Hayata) Ohwi as well [45]. Moreover, some plants in other families also contain tectorigenin, including Codonopsis pilosula (Franch.) Nannf. (Campanulaceae) [46], Morus alba L. (Moraceae) [46], Viola hondoensis W. Becker et H. Boissieu. (Violaceae) [47], and Eleocharis dulcis (Burm. f.) Trin. ex Hensch. (Cyperaceae) [48]. The plant sources of tectorigenin are shown in Table 1.

3. Extraction, Isolation, and Synthesis of Tectorigenin

The extraction and isolation of tectorigenin from plants with a high tectorigenin content have been widely performed in practical industrial production. For instance, tectorigenin can be prepared on a large scale from Belamcandae Rhizoma and Puerariae Flos [17,40,49]. The extraction methods of tectorigenin from dried slices or powders of raw materials include impregnation [50], ultrasonic extraction [17], reflux extraction [51], etc. Methanol, ethanol, and their aqueous solution are the most commonly used extraction solvents [40,52,53]. In addition, supercritical fluid extraction (with liquid CO2) was also used in some studies [54]. The crude extracts can be obtained by partitioning with non-polar organic solvents, such as ethyl acetate [55], chloroform [17], methylene chloride [49], etc. Moreover, repeated thin-layer chromatography and column chromatography were often performed for further isolation and purification [51]. In most cases, silica gel [56], and octadecyl silane were used as stationary phases [57]. Joung et al. [17] used the method of 70% ethanol ultrasound, partition with chloroform, and purification by silica gel column chromatography, and the yield of tectorigenin was 0.387 mg/g from 1 kg Belamcandae Rhizoma. To improve extraction efficiency, Li et al. [57] first developed an ionic-liquid-based ultrasound-assisted extraction method applied to the isolation of active compounds from Pueraria thomsonii, and an extraction yield of about 0.15 mg/g tectorigenin was obtained from 500 g raw material by countercurrent chromatography and semipreparative liquid chromatography. High levels of glycoside forms of tectorigenin, such as tectoridin, also exist in Belamcandae Rhizoma and Puerariae Flos, so the yield of tectorigenin can be increased by acid hydrolysis during extraction [58]. Kim et al. acidized the extract of the flower of Pueraria thunbergiana with H2SO4 to a final concentration of 1% (w/v), and the tectorigenin content increased from 17.10 µmol/g before acid hydrolysis to 49.58 µmol/g [59].
In addition to being isolated from plants, there are two approaches to obtaining synthetic tectorigenin (Figure 2): (i) It can be produced by hydrolyzing tectoridin with 3% HCl in MeOH/H2O (1:1, v/v) at 85 ℃ under reflux for 4 h in MeOH/H2O with a yield of around 37.5% [60,61], and (ii) the de novo synthesis of tectorigenin for the first time, which was reported by Kagal et al. (1956) [62], although the yield was very low. Subsequent attempts to synthesize tectorigenin failed because of the generation of a mixture of tectorigenin and ψ-tectorigenin [63,64]. In 1965, Varady [65] pointed out that ψ-tectorigenin could be catalyzed into tectorigenin through the isomerisation of electrons rearranged under alkaline conditions with alcoholic potash and alcoholic potassium ethylate by reflux. In this context, Xiao et al. [66,67] proposed a route to synthesize tectorigenin from 3-methoxy-methyl gallate.

4. Pharmacological Insights of Tectorigenin

As a ubiquitous plant isoflavone, tectorigenin has been widely reported to be effective in many areas, such as anticancer, anti-diabetes, hepatoprotection, anti-inflammation, antioxidation, antimicrobial, etc. (Figure 3 and Table 2). It seems to exert its pharmacological effects via modulating the signaling pathways, including PPARγ/NF-κB, PI3K/AKT, TLR4/NF-κB, IKKβ/NF-κB/JNK, ERK/JNK, MAPK/JNK/AP-1, AKT/MAPK, and TGF-β1/Smad.

4.1. Anticancer Effects

4.1.1. Prostatic Cancer

Several epidemiological studies support the role of phytoestrogens, nonsteroidal plant-derived compounds with estrogenic activity, in reducing cancer risk [68]. In the prostate, phytoestrogens could bind to estrogen receptor β (ERβ), which might be closely related to the prevention of prostate cancer progression [68,69]. Isoflavones were reported to exhibit estrogenic effects [70,71,72]. A study revealed that tectorigenin, as a phytoestrogen, could reverse the abnormal expression of some key factors that lead to prostate cancer malignancy [73]. Actually, tectorigenin has been proven to be a selective estrogen-receptor modulator that could bind to estrogen receptors ERα and Erβ, with a higher affinity for Erβ, and prevent the proliferation of some hormone-dependent prostate cancer cells [73,74]. Morrissey et al. (2004) [75] reported the positive effects of tectorigenin alone or combined with bicalutamide on a range of prostate epithelial cells in vitro, showing that tectorigenin caused cell cycle G1 arrest. In another study, tectorigenin extracted from the flowers of Puerariae thomsonii was found to possess the highest anti-proliferation activity against prostate cancer cells (IC50 = 0.08 μM) [53]. Stettner et al. [69] also reported that tectorigenin treatment with LNCaP prostate cancer cells up-regulated ERβ, resulting in antiproliferative effects.

4.1.2. Breast Cancer

Breast cancer is a major life-threatening malignancy that affects women all around the world [76]. MCF-7, MDA-MB-231, and T-47D are three commonly used breast cancer cells as models for breast tumors [77]. Zeng et al. (2018) [78] found that tectorigenin suppressed MCF-7 and MDA-MB-231 cell proliferation both in a dose- and time-dependent manner. The mechanisms of tectorigenin on human breast cancer cell apoptosis and metastasis might owe to the downregulation of protein kinase B (AKT)/mitogen-activated protein kinase (MAPK) signaling and upregulation of the expression of the caspase family. However, two earlier reports showed an opposite phenomenon: tectorigenin could stimulate the growth of MCF-7 and T-47D [44,55]. One explanation for this discrepancy is the difference in the number of cells and the concentration of tectorigenin [78]. Therefore, future studies in more diverse cell lines and animals are needed to confirm the protective effect against breast cancer of this molecule as well as its underlying mechanisms.

4.1.3. Ovarian Cancer

Ovarian cancer is another common gynecologic malignancy with a high mortality rate [79,80]. New therapeutic agents for ovarian cancer are urgently needed. It was reported that many plant-derived drugs and their derivatives induce apoptosis in ovarian cancer cell lines [81]. In paclitaxel-resistant ovarian cancer cells, tectorigenin heightened the growth-inhibitory activity of paclitaxel through downregulating the AKT/IκB kinase (IKK)/inhibitor of NF-κB (IκB)/noncanonical nuclear factor-κB (NF-κB) signaling pathway [82]. To be specific, tectorigenin combined with paclitaxel inhibited the NF-κB nuclear translocation and phosphorylation of IκB and IKK by activating caspases-3/8/9 and AKT, and downregulated the expression of NF-κB-dependent genes, thereby promoting synergistic apoptosis [82]. Another study investigated the anti-ovarian cancer effect of the methanol extract of Puerariae Flos and found that the extract presented a good anti-proliferation effect against the human ovarian cancer cell line A2780. Among the active compounds, the IC50 of tectorigenin against A2780 cells was 48.67 ± 0.31 μM [41].

4.1.4. Lung Cancer

There are also several studies on the effect of tectorigenin on lung cancer. The ethyl acetate extract of Chinese water chestnut peel, which contains tectorigenin at levels of 12.41 mg/g, showed good inhibitory activities on human alveolar adenocarcinoma cell line A549 (IC50 = 776.12 μg/mL) [48]. Tectorigenin (30 mg/kg) isolated from Belamcandae Rhizoma was administered subcutaneously to mice transplanted with Lewis lung carcinoma, and the inhibition ratio on tumor volume reached 30.8% [83]. In general, tumors try to escape from the host immune system and contrive to benefit from infiltrating immune cells by altering immune cell function, thereby creating a pro-inflammatory microenvironment that is conducive to tumor progression and metastasis [34]. It was reported by Amin et al. [34] that tectorigenin could restrain the pro-inflammatory response of monocytes induced by lung cancer cells, repressing the secretion of pro-inflammatory cytokines, TNF-α and IL-6. Nevertheless, the mechanism needs to be further elucidated.

4.1.5. Other Cancer

It has been shown by Jiang et al. [84] that tectorigenin could reduce the vitality of HepG2 (a human hepatocellular carcinoma cell line) in a time- and concentration-dependent manner via a mitochondrial-mediated pathway to induce apoptosis. Moreover, tectorigenin restrained the proliferation of Saos2 and U2OS, which are two osteosarcoma cell lines, and it dramatically inhibited the migration and invasion of osteosarcoma cells [21]. A similar study showed that tectorigenin inhibited tumor necrosis factor-α (TNF-α) via NF-κB inhibition, thereby decreasing CXCL10 overproduction to hinder the invasion of Caco-2, which is a human colon cancer cell line [85]. In another experiment, tectorigenin could not only induce human promyelocytic leukemia HL-60 cells differentiation into granulocytes and monocytes/macrophages but also cause intracellular apoptotic variations of DNA [86]. A point of view was put forward that the 5-hydroxyl group of tectorigenin was valuable for its cytotoxic activities, which may be instructive to us [86]. Moreover, tectorigenin lowered the expression of the stem cell factors POU5F1 and NANOG, and inhibited the proliferation of malignant testicular germ cell tumor (TGCT) cells [87]. A genomic hybridization microarrays analysis displayed that over 20% of the microarray genes, including telomeres, microdeletions, oncogenes and tumor suppressor genes, were aberrated in endometrial cancer cells treated with tectorigenin [88].

4.2. Antidiabetic and Anti-Obesity Effects

Diabetes is a metabolic disorder characterized by hyperglycemia that has become a global epidemic [89]. Tectorigenin was reported to effectively decrease the serum glucose level of rats in streptozotocin-induced or high-fat and high-sucrose diet models [42,90,91,92], and it was reported as a potential antidiabetic agent exhibiting potent inhibitory activity on aldose reductase in rat lenses [47,93]. In vitro, tectorigenin inhibited glucotoxicity- and lipotoxicity-induced oxidative stress, apoptosis, and endoplasmic reticulum stress in islet cells. The mechanism was that tectorigenin regulated the expression of pancreas/duodenum homeobox protein 1 (PDX1) and extracellular signal-regulated kinase (ERK) [92]. Endothelial dysfunction is frequently seen in diabetic patients [94,95]. Tectorigenin was demonstrated to alleviate diabetic nephropathy, which was attributed to its protective effect on injured endothelial cells by inhibiting inflammation and lipotoxicity, and by restoring insulin sensitivity [96]. Mechanistically, the pharmacological properties of tectorigenin were related to the adiponectin receptor 1/2 (AdipoR1/2)-mediated activation of adenosine monophosphate-activated protein kinase (AMPK) and peroxisome proliferator-activated receptor (PPAR) pathways [96]. Similarly, tectorigenin exerted positive regulation of insulin action in palmitate-injured human umbilical vein endothelial cells (HUVECs) via modulating reactive oxygen species (ROS)-related inflammation and insulin receptor substrate-1 (IRS-1) signaling [97]. Glucose transporter protein 4 (GLUT4) is the major glucose transporter, whose decrease is one of the important molecular bases of insulin resistance [98]. Recently, it was proven that tectorigenin targeted protein kinase A catalytic subunit α (PKACα) to promote the PKA/AMPK/myocyte enhancer factor 2 (MEF2) pathway, subsequently enhancing GLUT4 expression, and thus slowing and stopping insulin resistance progression for the intervention and treatment of glucose metabolism syndrome [99].
Diabetes is closely linked to the epidemic of obesity [100]. The anti-obesity effects of tectorigenin were also investigated, as evidenced by the decreased body weight, triglycerides, total cholesterol, and low-density lipoprotein cholesterol (LDL-C) in several animal models [42,90,92]. According to the findings of Li et al. [101], tectorigenin restrained 3T3-L1 adipogenesis and reversed TNF-α-induced changes of IL-6, monocyte chemoattractant protein-1 (MCP-1), and adiponectin. Further investigation identified tectorigenin as a functional peroxisome proliferator-activated receptors γ (PPARγ) partial agonist (IC50 = 13.3 mM) and suggested that tectorigenin may ameliorate hyperglycemia through the inhibition of preadipocyte differentiation and adipocytokine secretion [101]. The studies presented above highlight the potential therapeutic value of tectorigenin in diabetes and obesity.

4.3. Hepatoprotective Effects

The liver plays a pivotal role in bile synthesis, metabolic function, and the degradation of toxins in the body [102]. Several factors, including oxidative stress, lipid peroxidation, and proinflammatory mediators (chemokines and cytokines), are involved in hepatic diseases [103]. Tectorigenin is a potential hepatoprotective agent that has been evaluated by many researchers. It has been demonstrated that tectorigenin exerted protective functions against carbon tetrachloride (CCl4) and tert-butyl hyperoxide-induced liver damage [56,104,105,106]. Tectorigenin significantly inhibited the activities of aspartate aminotransferase (AST) and alanine aminotransferase (ALT), and the protective activity of tectorigenin was higher than that of dimethyl diphenyl bicarboxylate and silybin [104,106]. Several studies showed a regulation in the activities of antioxidative enzymes, lipid peroxides, and cytokines in the tectorigenin-treated group, as evidenced by the superoxide dismutase (SOD), ROS, malondialdehyde (MDA), glutathione (GSH), TNF-α, interleukin (IL)-1β levels, etc., which demonstrated that the hepatoprotective mechanisms of tectorigenin might be related to its antioxidant and anti-inflammatory actions [56,105,107]. Liver damage can lead to hepatic fibrosis, cirrhosis, liver failure, and even liver cancer [108]. It has been revealed that tectorigenin exhibited anti-proliferative and pro-apoptotic activities on hepatic stellate cells and human hepatocellular carcinoma HepG2 cells, and might possess anti-fibrotic and anti-hepatoma potential [84,109]. An in vivo study suggested that tectorigenin significantly prevented fat accumulation, promoted bile acid circulation, and exerted beneficial effects on mice with nonalcoholic fatty liver disease through anti-inflammation and improvement of gut microbial dysbiosis [110]. Another study specifically reported the effects of tectorigenin on cholestatic liver disease, and the results proved that tectorigenin alleviated intrahepatic cholestasis via the activation of peroxisome proliferator-activated receptor gamma (PPARγ) and subsequent NF-κB inhibition and bile salt export pump (BSEP) activation [23]. Additionally, tectorigenin mitigated experimental fulminant hepatic failure via modulating the toll-like receptor 4 (TLR4)/MAPK and TLR4/NF-κB pathways and autophagy [107].

4.4. Anti-Inflammatory Effects

Several flavonoids, including tectorigenin, have been reported to inhibit nitric oxide (NO) production, one of the inflammatory mediators [111]. Tectorigenin was also found in multiple studies to inhibit the induction of cyclooxygenase-2 (COX-2) in a dose-dependent manner to restrain the production of prostaglandin E2 in inflammatory cells. This may be one of the mechanisms of action by which Belamcandae Rhizoma exert an anti-inflammatory effect [112,113,114,115]. Lipopolysaccharide (LPS), an endotoxin, could upregulate inflammatory cytokines, such as IL-1/6/12, TNF-α, p-MAPK, and p-NF-κB. Those changes above were attenuated by tectorigenin in PC12 cells with spinal cord injury [116], LPS-stimulated BV-2 Microglia [115], and mice with acute lung injury [117]. In the A549: THP-1 co-culture model, A549 cells activated and induced THP-1 cells to secrete IL-6, TNF-α, and pro-inflammatory cytokines. Co-incubation of tectorigenin with A549 cells prevented this induction behavior [34]. An in vivo model of acute inflammation study showed that tectorigenin (60 mg/kg) significantly alleviated carrageenan-induced edema in an inflammatory rat model [118]. Wang et al. [24] (2020) also investigated the anti-inflammatory ability of tectorigenin and explored the underlying mechanism in treating the allergic asthma model of guinea pigs. In that study, tectorigenin (25 mg/kg) efficiently reduced the frequency of cough, inflammatory cell numbers, and the content of pro-inflammatory factors. Further research found that tectorigenin might mitigate pulmonary fibrosis and airway inflammation through the TLR4/NF-κB and transforming growth factor β1 (TGF-β1)/Smad signaling pathways, which was consistent with the conclusion of another study that tectorigenin was suggested to play an anti-inflammatory role by antagonizing the activation of the NF-κB signaling pathway [101]. Tectorigenin protected HaCaT keratinocytes against M5 cytokine-induced abnormal proliferation and inflammatory response via suppression of the TLR4/NF-κB pathway [119]. In addition, tectorigenin also protected against multiple organ damage by alleviating inflammation-related pathways or oxidative stress [107,117,120].

4.5. Antioxidant Effects

A loss of balance in biological systems between the production of ROS and antioxidant defense levels can induce oxidative stress, which may result in extensive intracellular damage [121,122]. The harmful effects brought by oxidative stress are mainly on account of the overproduction of oxidants or depleting antioxidant potential [123], which causes cellular damage through the oxidation of protein, lipids, and DNA [124,125]. Oxidative stress participates in the nosogenesis of many illnesses, including several cancers, Alzheimer’s disease, and atherosclerosis [125]. Flavonoids are reported as natural scavengers of free radicals, and this may be due to the relatively strong reduction capacity of their phenol groups, which could form resonance-stabilized anion radicals [126]. Both tectorigenin and tectoridin have been reported to show antioxidant ability in vivo and in vitro [42,56]. In an early study, the evaluation of the antioxidant ability of Belamcandae Rhizoma extract containing tectorigenin showed the results of reducing free-radical 1,1-Diphenyl-2-picrylhydrazyl (DPPH) and transition-metal ions and decreasing linoleic acid peroxidation [51]. Another antioxidant capacity assay of tectorigenin displayed that tectorigenin at 10 μg/mL exerted significant intracellular ROS scavenging activity (63.2 ± 2.3%) and DPPH radical scavenging activity (54.3 ± 2.3%), which was superior to tectoridin [127]. In addition, tectorigenin increased the activities and protein expression of the antioxidants, including catalase (CAT), glutathione peroxidase (GPx), and SOD, which might be related to the restraint of the overproduction of ROS and lipid peroxidation. Therefore, the antioxidant capacity of tectorigenin might originate from two aspects: direct scavenging of oxygen free radicals and an indirect effect on the induction of antioxidative enzymes. In vivo experiments also revealed the antioxidative potentiality of tectorigenin, which showed that tectorigenin restrained CCl4 and bromobenzene-induced malondialdehyde formation in the rats [56,128]. Tectorigenin sodium sulfonate, reported by Han et al. [60], maintained and elevated the antioxidant activity of tectorigenin through Fe3+/ferricyanide reduction, hydroxyl radical and superoxide anion radical scavenging, DPPH, and lipid peroxidation assays in vitro. Another study evaluated the effects of Belamcandae Rhizoma extract containing tectorigenin on collagen degradation and apoptosis in UV-B-induced HaCaT cells. The extract exhibited good free radical scavenging capability and cytoprotective effects, and mitigated cell apoptosis through diminishing caspase-3 levels and increasing the B-cell lymphoma-2 (Bcl-2)/Bcl-2 associated X (Bax) ratio [26]. Taken together, these findings suggest that tectorigenin has a certain antioxidant capacity.

4.6. Antimicrobial Effects

The increasing incidence of microorganisms’ drug resistance makes the search for new antimicrobial agents urgent. In this connection, medicinal plants are promising resources [129]. In 2001, two reports revealed the antimicrobial potential of tectorigenin. Its minimum inhibitory concentration (MIC) against four species of dermatophytes of the genera Trichophyton ranged from 3.12–6.25 μg/mL [50], and also inhibited the growth of six Helicobacter pylori strains (MIC: 50–100 μg/mL) [130]. In another similar study, tectorigenin from the Belamcandae Rhizoma exerted antibacterial activity against S. aureus ATCC 33591, S. aureus ATCC 25923, which were strains of methicillin-resistant Staphylococcus aureus (MRSA), and three clinical isolates of MRSA with MIC values of 125 μg/mL [17]. Further examination confirmed that the anti-MRSA effect of tectorigenin was achieved by suppressing the adenosine triphosphatase (ATPase) and enhancing the permeability of the cytoplasmic membrane. In a cell experiment for in vitro treatment of Clostridium perfringens infection, the biofilm formation rates in the presence of 4, 8, 16, and 32 μg/mL tectorigenin were well below the control in a dose-dependent manner [131]. In addition, tectorigenin noteworthily suppressed the gliding movement, biofilm formation, and adhesion of Caco-2 cells by targeting type IV pilus (TFP) and suppressing TFP-associated genes, although it exhibited little antibacterial activity directly against Clostridium perfringens. However, more research is still needed to further verify the antibacterial activity of tectorigenin in cells and animals [131].

4.7. Bone-Protective Effects

Common skeletal diseases, such as osteoarthritis, osteoporosis, fractures, etc., greatly affect the quality of life of older adults [132]. The homeostasis of bone metabolism is critical for bone health [133]. Bone homeostasis is sustained by the balance between osteoblast-mediated bone formation and osteoclast-mediated bone resorption [134]. Many factors, including aging, postmenopausal estrogen deficiency, and prolonged immobilization, can lead to the disturbance of bone metabolism and result in bone-related diseases [135]. Some phytoestrogens, isoflavonoids, and plant-derived nonsteroidal compounds possessing estrogen-like activity could function as inhibitors of osteoporosis [136]. The healthy functional food containing tectorigenin effectively reduced bone resorption and promoted bone formation in ovariectomized-induced osteoporosis mice [137]. Ma et al. [138] investigated the therapeutic effects of tectorigenin on osteoporosis, and the results showed that tectorigenin reduced mRNA levels of osteoclast-specific genes, including nuclear factor of activated T cell cytoplasmic 1 (NFATc1), tartrate-resistant acid phosphatase (TRAP), matrix metalloproteinase (MMP)-9 and cathepsin K in RAW264.7 cells, and bone marrow mononuclear cells, and mitigated the bone loss in osteoporosis mice [138]. Another study also showed that tectorigenin regulated bone homeostasis by stimulating osteogenic differentiation and inhibiting osteoclast differentiation through bone morphogenetic protein (BMP) and MAPK pathways [120]. Tectorigenin is also attractive as an anti-osteoarthritis drug due to its anti-inflammatory activities. Tectorigenin observably suppressed the NF-κB P65 pathway and inhibited articular cartilage degeneration and chondrocyte apoptosis [139]. Tendinopathy, as a painful overuse musculoskeletal injury, is extremely common in athletes and middle-aged overweight patients [140]. As an inhibitor of MAPK and NF-κB pathways, tectorigenin has the capability to attenuate the inflammation, apoptosis, and ossification of tendon-derived stem cells in vitro and in vivo, thus effectively improving tendinopathy [141].

4.8. Anti-Skin-Damage and Antiallergic Effects

Ultraviolet radiation (UV) can induce skin photoaging and inflammation, and also lead to free radical and ROS accumulation, collagen loss and degradation, epidermal pigmentation, and abnormal cell death [142]. Several reports have uncovered the skin-protective effect potential of tectorigenin. Belamcandae Rhizoma extract containing tectorigenin exhibited good free radical scavenging and cytoprotective activities by increasing antioxidant enzyme expression and improving UV-B-induced collagen degradation and apoptosis [26]. Noh et al. [22] also reported that in human keratinocytes, tectorigenin (1 and 10 µM) exerted anti-skin-damage effects through mitigating UV-B-induced hyperoxidation, collagen degradation, and apoptosis. Recently, Dai et al. [143] designed a study to pick out a new retinoic acid γ receptor (RAR-γ)-selective agonist. Computational screening suggested that tectorigenin was a matching selective RAR-γ agonist. In this study, tectorigenin was validated to be able to inhibit UV-induced release of inflammatory factors, oxidative damage, and MMP production, and also to reverse collagen loss. Further research suggested that tectorigenin exerted its effects mainly via regulating the MAPK/c-Jun N-terminal kinase (JNK)/activating protein-1 (AP-1) pathway. Tretinoin can attenuate skin photoaging and inflammation, but its use is restricted because of the strong skin irritation and teratogenic effect [144]. Therefore, it is necessary to find a new agent with protective effects on the skin. These evidences above suggest that tectorigenin has the potential to be an effective strategy for the treatment of UV-induced skin damage. Moreover, Tamura et al. [145] disclosed tectorigenin and another compound from the extract of Puerariae Flos as inhibitors against the expression of immunoglobulin E (IgE) receptor, the key molecule triggering the allergic reactions, via diminishing the generation of γ-chain subunit. Another study demonstrated that tectorigenin effectively attenuated passive skin allergic reaction and restrained IgE-induced release of β-hexosaminidase from RBL-2H3 cells [146].

4.9. Cardioprotective and Cerebroprotective Effects

Cardiovascular and cerebrovascular diseases, including angor pectoris, myocardial infarction, stroke, etc., are the leading causes of death globally [147]. Atherosclerosis is the main cause of cardiovascular and cerebrovascular disease development [148]. Blood vessels severely affected by atherosclerosis are prone to enhance platelet aggregation, resulting in vessel occlusion and ischemia [149]. Tectorigenin was an effective antiplatelet compound with a much better effect than acetylsalicylic acid [150]. Tectorigenin was also reported as a lactate dehydrogenase (LDH) inhibitor from several Chinese medicinal herbs by UF-HPLC-DAD-MS [151]. The protective mechanism of tectorigenin was related to oxidative stress and inflammation in which phosphoinositide 3-kinase (PI3K)/AKT, TLR4/NF-κB, PPARγ/NF-κB pathways were involved [39,152,153]. Chen et al. [39] revealed that tectorigenin effectively prevented HUVECs from H2O2-induced oxidative stress injury by upregulating the PI3K/AKT pathway. In addition, tectorigenin could alleviate cognitive impairment, hippocampal tissue and myelin damage, and inflammation of chronic cerebral ischemia mice [153]. In vitro, tectorigenin benefited HT-22 cell survival and protected against oxygen-glucose deprivation/reoxygenation (OGD/R) damage. However, overexpressing TLR4, or using PI3K/AKT inhibitor LY294002, PPARγ inhibitor GW9662, or NF-κB activator LPS reversed the protection of tectorigenin, suggesting that PI3K/AKT, TLR4, PPARγ, and NF-κB pathways were crucial in the cardioprotective and cerebroprotective effects of tectorigenin [152,153].

4.10. Protective Effects on the Respiratory System

Tectorigenin was regarded as an interventional strategy by inhibiting respiratory dysfunction and death in respiratory diseases. Several studies have illustrated the protective effect of tectorigenin in respiratory disorders, including asthma, airway epithelial injury, and pulmonary fibrosis [24,154,155]. Asthma is a common chronic disease characterized by variable respiratory symptoms and airflow limitation [156]. Tectorigenin significantly diminished the frequency of coughs, the number of inflammatory cells, and the levels of inflammation-related factors in the allergic asthma guinea pigs model [24]. For the clinical treatment of asthma, glucocorticoids are currently the first-line drugs, but can also cause airway epithelial injury [24]. It was demonstrated that tectorigenin regulated migration, invasion, and apoptosis of human airway epithelial cells in the treatment of glucocorticoids. Mechanistically, tectorigenin enhanced miR-222-3p expression and inhibited the MAPK pathway, thus protecting the airway epithelium [154]. Idiopathic pulmonary fibrosis is also a chronic, progressive pulmonary disease characterized by the anomalous accumulation of fibrotic tissue in the lung’s parenchyma [157]. In vitro, tectorigenin prevented the proliferation of pulmonary fibroblasts in rats treated with bleomycin, indicating that tectorigenin has the potential to ameliorate pulmonary fibrosis. Further results revealed that its mechanism was related to the regulation of miR-338* (miR-338-5p) expression [155].

4.11. Neuroprotective Effects

In neurodegenerative diseases caused by multiple pathological procedures, neuroprotection is an interventional strategy to delay and even stop neuronal dysfunction and abnormal death [158]. Microglia are the primary immune cells in the central nervous system, and their activation is associated with neurodegeneration and alcoholic toxication [159]. Yuan et al. [40] found that isoflavonoids from Puerariae Flos, which included tectorigenin, had obvious inhibitory effects on the release of NO from microglia activated by LPS (IC50 values were 1.3–2.3 μM). From the structure-activity relationships of a number of isoflavonoids with inhibitory activity against microglial activation, the methoxyl group at the 6-position of tectorigenin enhanced the activity. Glioblastoma cell viability was dose-dependently decreased after 24 h exposure to tectorigenin, and 200 μM and 300 μM could block the cell cycle arrest at G0/G1 phase [160]. Another study also confirmed the anti-neuroinflammatory effects of tectorigenin in both LPS-treated BV-2 microglial and mouse models [115]. As a glycoprotein hormone, Erythropoietin (EPO) has neuroprotective function [161,162], but EPO circulating in the blood cannot cross the blood-brain barrier [163]. Tectorigenin was proved to promote the stacking of hypoxia-inducible factor (HIF)-1α to induce EPO gene expression in rat cortical neurons and neuron-like NT2/D1 cells, raising endogenous cerebral EPO levels [164]. Monoamine oxidase B (MAO-B) can produce ROS to directly damage neuronal cells, which is a possible target for the treatment of Alzheimer’s disease [57]. Li et al. [57] isolated tectorigenin from Pueraria thomsonii, and the IC50 value of tectorigenin against MAO-B was 54.36 μg/mL. In addition, pretreatment with tectorigenin exhibited a protective effect against neuronal damage in PC12 cells. Gong et al. (2017) [165] investigated whether tectorigenin could prevent the neurotoxicity of SH-SY5Y cells and illuminated the potential protection mechanism. In that study, tectorigenin (0.1, 1 and 10 μM) exhibited neuroprotective effect against cytotoxicity and apoptosis induced by MPP+ (1-methyl-4-phenylpyridinium), which may be related to the reduction of oxidative stress and the enhancement of the antioxidant defense activity of tectorigenin [165].

4.12. Other Effects

Except for the pharmacological activities mentioned above, tectorigenin also has other biological activities. For instance, tectorigenin was verified as a therapeutic strategy for the treatment of obstructive nephropathy. In mice suffering unilateral ureteral obstruction, tectorigenin significantly decreased the levels of the kidney injury index, including creatinine, blood urea nitrogen, and kidney injury molecule-1 (KIM-1), and alleviated pathological damage and renal interstitial fibrosis [166]. In vitro, tectorigenin treatment inhibited Smad3-mediated ferroptosis and fibrosis [166]. It was also revealed that tectorigenin could induce the release of growth hormone in rat pituitary cells [167].
Table 2. Biological and pharmacological effects of tectorigenin.
Table 2. Biological and pharmacological effects of tectorigenin.
Pharmacological EffectsAnimal/Cell LineInducerRouteDosageDurationMechanism of ActionResultsRef.
Anticancer effects
(Prostate cancer)
PC-3 and LNCaP cells/In vitro10, 50, 100 μM24 hUp-regulated ERβ, decreased the tumor cell proliferation-
related gene expression
ERβ, PDEF expression↑; PSA, prostate cancer-specific indicator gene DD3 (PCA3), hTERT, IGF-I receptor expression↓[69]
Anticancer effects
(Prostate cancer)
LNCaP cells/In vitro100 μM24 hRegulated the aberrant expression of genes relevant in proliferation, invasion, immortalization, and apoptosisProstate-derived Ets factor (PDEF), prostate
-specific antigen (PSA), human telomerase reverse transcriptase (hTERT), insulin-like growth factor 1 (IGF-1) receptor expression↓; Telomerase activity↓; inhibited MMP and induced apoptosis in LNCaP cells
[73]
Anticancer effects
(Prostate cancer)
RWPE-1, LNCaP and PC-3 cells/In vitro50 μM24, 72 hRegulated cell cycleInduced cycle arrest at G1 phase and p21 (WAF1) or p27kip1 protein expression[75]
Anticancer effects
(Prostate cancer)
RM-1, H22 and MFC cells/In vitro/24, 48 h/Exhibited anti-prostate cancer activity with IC50 value of 0.08 μM[53]
Anticancer effects (Breast cancer)MCF-7 and T-47D cells/In vitro/96 h/Stimulated MCF-7 and T-47D cell proliferation[55]
Anticancer effects (Breast cancer)MDA-MB-231, MCF-7 and hMSC cells/In vitro50, 100, 200 μM24, 48, 72, 96 hDownregulated matrix metalloproteinases
and AKT/MAPK signaling; upregulated caspase family
Induced apoptosis, G0/G1-phase arrest, migration and invasion; MMP-2, MMP-9, Bcl-2, p-AKT, and MAPK expression↓; Bax, cleaved poly [ADP-ribose] polymerase and cleaved caspase-3 expression↑[78]
Anticancer effects (Ovarian cancer)MPSC1TR, A2780TR and SKOV3TR cells/In vitro25, 100 μM48 hInactivated AKT/IKK/IκB/NF-κB pathwayActivated caspases-3, -8 and -9; Nuclear translocation of NF-κB and FLICE inhibitory protein (FLIP), X-linked inhibitor of apoptosis protein (XIAP), Bcl-2, Bcl-xL and COX-2 expression↓; phosphorylation of IκB and IKK and the activation of AKT↓[82]
Anticancer effects (Ovarian cancer)A2780 and IOSE80PC cells/In vitro1.56–100 μM48 h/IC50 against A2780 cells is 48.67 ± 0.31 μM[41]
Anticancer effects (Lung cancer)A549 cells/In vitro15, 30, 60, 120, 240 μg/mL24 h/Inhibited A549 cells growth with IC50 221.52 μg/mL[48]
Anticancer effects (Lung cancer)LLC, S180 and CPAE cells; C57BL/6 mice Injection of LLCs (in vivo)In vitro and in vivo 1, 10, 100 μM (in vitro); 30 mg/kg (in vivo)3 days (in vitro); once a day for 10 days (in vivo)/Inhibited tumor growth[83]
Anticancer effects (Lung cancer)A549 and THP-1 cells/In vitro10, 25, 50 μM24 hSuppressed lung cancer-induced pro-inflammatory responseTNF-α and IL-6 secretion↓; snail expression↓; E-cadherin↑[34]
Anticancer effects (Hepatocellular cancer)HepG2 cells/In vitro2.5, 5, 10, 15, 20, 30, 40 μg/mL12, 24, 48 hInduced apoptosis via mitochondrial-mediated pathwayHepG2 cell viability↓; induced the condensation of chromatin and fragmentation of nuclei; ROS, intracellular [Ca2+], caspase-9 and -3↑; mitochondrial membrane potential↓ [84]
Anticancer effects (Osteosarcoma)Saos2 and U2OS OS cells/In vitro100, 200, 400 µM24, 48, 72 h,Inhibited the proliferation, migration and invasionInhibited migration and invasion; cleaved caspase3↑, MMP-1, MMP-2, and MMP-9↓[21]
Anticancer effects (Colon cancer)Caco-2 cells/In vitro20 µM24 hNF-κB pathway suppressionArrested invasion; CXCL-10 expression, p-IκB, p-RelA↓; CXCL-10 promoter activity↓[85]
Anticancer effects (Leukemia)HL-60, U-937, HepG2 and SNU-C5 cells/In vitro20, 50, 75 μM1–4 daysInduced differentiation and apoptosisInhibited HL-60, U-937, HepG2 and SNU-C cell growth with IC50 values of 22.3, 28, 84, and 62.7 μM, respectively; induced HL-60 cells differentiation, caused apoptotic changes of DNA[86]
Anticancer effects (Testicular cancer)TGCT, TCam-2
and NTera-2 cells
/In vitro100, 250, 500 μM48, 72 hDownregulated stem cell factorsInhibited proliferation, the stem cell factors NANOG and POU5F1↓[87]
Anticancer effects (Endometrial cancer)Ishikawa cells/In vitro0.1 μM72 hGenomic aberrationsInduced array genes aberrated[88]
Anticancer effectsCOR-L23, C32, MCF-7, HepG2/In vitro50, 100, 200, 400 µM48 hInhibited S phase and G2/M phaseIC50 against COR-L23, C32, MCF-7, HepG2 were 189, 207, 149, 105 µM, respectively; 400 µM increased significantly G2/M phase cells[54]
Anticancer effectsP388, L1210, SNU C4, A549 and MA104 cells/In vitro///IC50 against P388, L1210, SNU C4, A549 and MA104 were 0.2, 0.04, 0.03, 0.29, >1 mM, respectively[90]
Antidiabetic effectsMale SD ratsStreptozotocinIn vivo5, 10 mg/kgA weekAntioxidant activitySerum cholesterol, triglyceride, LDL- and VLDL-cholesterol↓; High-density lipoprotein (HDL)-cholesterol↑; DPPH radical, xanthine oxidase, superoxide anion radical, lipid peroxidation↓[42]
Antidiabetic effectsMale SD ratsStreptozotocinIn vivo 10 mg/kgOnce per day for 3 days/Body weight↑; Serum glucose, cholesterol↓[90]
Antidiabetic effectsMale SD ratsStreptozotocinIn vivo 100 mg/kg10 consecutive days/Exhibited high aldose reductase inhibitory potency with IC50 1.12 μM; inhibited the sorbitol accumulation in red blood cells, sciatic nerves, and lenses by 87.2%, 75.5%, and 50.5%, respectively[91]
Antidiabetic effectsINS-1, RIN-
m5F, and HEK293T cells; Male C57BL/6J mice
Glucose, palmitic acid (in vitro); high-fat +sucrose diet (in vivo)In vivo and in vitro40 μg/mL (in vitro); 10, 20, 40 mg/kg (in vivo)3 h, 9 h, 24 h (in vitro); once every two days for one month (in vivo)Enhanced PDX1 expression and protected pancreatic β-cells by activating ERK and
reducing ER stress
Improved insulin secretion; weight gain, ROS↓; ameliorated hyperglycemia and glucose intolerance and lipotoxicity and apoptosis[92]
Antidiabetic effectsRat lenses/In vitro///Aldose reductase inhibition IC50 = 1.12 ± 0.08 μM[47]
Antidiabetic effectsRat lens/In vitro///Aldose reductase inhibition IC50 = 6.43 μM[93]
Antidiabetic effectsHRGECs; Male type 2 diabetic (BKS.C g–m +/+ Lepr db/J, db/db) mice, a model homozygous for the diabetes spontaneous mutationHigh-glucose (in vitro)In vitro and in vivo5, 10 μM (in vitro); 75 mg/kg (in vivo)24 h (in vitro); 12 weeks (in vivo)Restored the reduction of AdipoR1/2, pi-LKB1, pi-AMPKα, PPARα, decreased lipotoxicity, reduced macrophage infiltration and macrophage polarizationAttenuated metabolic disorders and exerted renoprotective effects; improved intrarenal lipid metabolism, endothelial functions, and renal insulin sensitivity[96]
Antidiabetic effectsHUVECsPalmitic acid In vitro0.1, 1, 10 μM30 minInhibited IKKβ/NF-κB/JNK pathwayROS, MMP, TNF-α, IL-6↓; protected endothelium-dependent relaxation[97]
Antidiabetic effectsC2C12 myotubes; Male C57BL/6J micePalmitic acid (in vitro); high fat (in vivo) In vitro and in vivo10 µg/mL (in vitro); 10, 20, 40 mg/kg or 50, 100, 200 mg/kg (in vivo)24 h (in vitro); every other day for 30 days (in vivo)PKACα/AMPK/MEF2 pathwayImproved GLUT4, glucose uptake, and insulin sensitivity; ameliorated insulin resistance and hyperglycemia[99]
Anti-obesity effects3T3-L1 cells; Male SD ratsDexamethasone (in vitro); high-fat diet (in vivo) In vitro and in vivo 10, 25, 50, 75 μM (in vitro); 50, 100 mg/kg (in vivo)24 h (in vitro); once daily for 2 weeks (in vivo)PPARγ and IKK/NF-κB pathwayInhibited adipocyte differentiation; Triglyceride, glycerol-3-phosphate dehydrogenase, adipogenesis-related genes expression↓; IL-6, MCP-1↓; adiponectin secretion↑; increased glucose uptake and insulin sensitivity[101]
Hepatoprotective effectsmale ICR miceCCl4In vivo50, 100 mg/kgOnceInhibited the β-glucuronidase activityInhibited the levels of serum ALT, AST and LDH by 22.4%, 44.4%, and 58.7%, respectively; MDA, Ca2+↓, GSH, GST↑[104]
Hepatoprotective effectsMale SD ratsCCl4, high-fat + cholesterol diet, alcoholIn vivo 7.5, 15, 30 mg/kgOnce per day for 6 weeksAntioxidant activitySerum ALT, AST, hyaluronate, laminin, procollagen III N-terminal peptide↓; collagen in the livers↓, serum albumin concentration and ratio of albumin to globulin↑, liver lipid peroxidation↓, liver SOD and GPx↑[105]
Hepatoprotective effectsHepG2 cells; Male ICR micet-BHP (in vitro and in vivo)In vitro and in vivo 0.01, 0.1, 1, 10 μM (in vitro); 25, 50 mg/kg (in vivo)2 h (in vitro); once (in vivo)Inhibited the hepatotoxicityInhibited the levels of plasma ALT and AST by 39% and 41%, respectively[106]
Hepatoprotective effectsRAW 264.7 cells; Male C57BL/6J miceLPS and D-GalNIn vitro and in vivo1, 10, 100 μM (in vitro); 12.5, 25, 50 mg/kg24 h (in vitro); once (in vivo)TLR4/MAPK, TLR4/NF-κB, decreased inflammatory cytokine levels, promoted autophagySerum ALT, AST↓; Ameliorated the histological injury, apoptosis, and the mortality[107]
Hepatoprotective effectsHSC-T6 cells/In vitro10, 20, 40, 60, 80, 100 μg/mL12, 24, 48 hInhibited proliferation and induced apoptosisCell viability↓; ROS, intracellular [Ca2+]↑; mitochondrial membrane potential↓; translocation of cytochrome c, caspase-3/9↑[109]
Hepatoprotective effectsMale C57BL/6N miceHigh-fat dietIn vivo 25, 50 mg/kgOnce per day for 6 weeksLPS/TLR-4/NF-κB/TNF-α pathway, modulated gut microbiotaAmeliorated the obese characteristics; total cholesterol, total triglyceride, HDL-C, LDL-C, LPS, total bile acid, ALT, AST↓; fecal total bile acid↑; ameliorated the histological injury[110]
Hepatoprotective effectsPrimary mouse Kupffer cells; Male C57BL/6J miceANIT, DDC dietIn vitro and in vivo 10 μM (in vitro); 75 mg/kg24 h (in vitro); once per day for 5 days (in vivo)PPARγ/NF-κB, PPARγ/BSEP, inhibited macrophage activationSerum AST, ALT, γ-glutamyltransferase, and AP↓; ameliorated the histological injury, cell apoptosis, inflammation, and bile metabolic dysfunction[23]
Anti-inflammatory effectsRaw 264.7 cellsIFN-γ/LPSIn vitro50, 100, 200 μM24 hBlocking of NF-κB activationNitric oxide synthase (iNOS)↓, NO↓, IL-1β↓, COX-2↓, Prostaglandin E2 (PGE2)↓ [114]
Anti-inflammatory effectsPC12 cellsLPSIn vitro25, 50, 100, 200 μM24 hAlleviated apoptosis, inflammation, and activation of NF-κB signaling in SCI cell models via inhibiting IGFBP6Cell viability↑, cell apoptosis↓, caspase-3/8/9, cleaved caspase-3/8/9; IL-1β, IL-6, TNF-α, IGFBP6, TLR4↓; inactivated IκBα and p65[116]
Anti-inflammatory effectsFemale BALB/c miceLPSIn vivo5, 10 mg/kg6 hNF-κB P65 pathwayInflammatory cell numbers↓, lung NF-κB p65 mRNA and protein level↓, myeloperoxidase↓, SOD↑, inhibited LPS-induced neutrophils in the lung [117]
Anti-inflammatory effectsSwiss mice; Wistar ratsAcetic acid or carrageenanIn vivo50, 100 mg/kg; 10, 60 mg/kg//In mice, LD50 was 1.78 g/kg, had an analgesic effect; in inflammatory rats, reduced carrageenan-induced edema[118]
Anti-inflammatory effectsHaCaT cellsM5 cytokinesIn vitro2.5, 5, 10, 20, 40 μM/TLR4/NF-κB pathway; promoted autophagyPromoted autophagy: LC3-II/LC3-I, beclin-1, LC3↑; P62↓; suppressed inflammation: IL-6, IL-1β, TNF-α↓; NOD-like receptor family pyrin domain containing 3 (NLRP3), apoptosis-associated speck-like protein (ASC) and caspase-1↓[119]
Antioxidant effectsMale SD ratsCCl4In vivo100 mg/kgOnce a day for 7 daysPrevented lipid peroxidation, antioxidationIC50 of free radical scavenging potency was 275 μM; Liver MDA, SOD, CAT, GPx↓ by 75.6%, 63.8%, and 70.4%, respectively; serum AST, ALT↓ by 47.4% and 39.8%, respectively[56]
Antioxidant effectsV79-4 cellsH2O2In vitro0.1, 1, 10 μg/mL25 hActivated ERK pathway; enhanced antioxidative levels, apoptotic and cycle arrest Prevented lipid peroxidation: intracellular ROS, DPPH↓; apoptotic cells, cell cycle arrest at G2/M↓; Cellular SOD, GPx, CAT↑; cell viability↑[127]
Antioxidant effectsMale SD ratsBromobenzeneIn vivo10 mg/kgOnce a day for 7 days/Inhibited the AFB1-induced mutagenicity by 90% and MNNG-induced one by 76%; prevented the MDA formation[128]
Antioxidant effects//In vitro50, 100, 150, 200 μM//Exhibited reductive capability and DPPH scavenging activity, IC50 of OH, O2•– scavenging activity and inhibition of lipid peroxidation were 87, 46.62, and 23 μg/mL, respectively[60]
Anti-microbial effects (fungal)7 strains of fungi and 6 strains of bacteria/In vitro/72 h/Inhibited dermatophytes of the genera Trichophyton (MIC: 3.12–6.25 μg/mL) and such yeast-like fungi as Can dida and Saccharomyces (25–50 μg/mL); inhibited the growth of P. aeruginosa, P. vulgαris, M. luteus and S. aureus (50–100 μg/mL)[50]
Anti-microbial effects (HP)HP ATCC43504, NCTC11637, NCTC11638, HP82516, HP82548 and HP4/In vitro/3 days/Inhibited HP growth (MIC: 50–100 μg/mL)[130]
Anti-microbial effects (MRSA)Methicillin-Resistant Staphylococcus aureus/In vitro/18 hAnti-MRSA action is related to cytoplasmic membrane permeability and ABC transporterInhibited all tested strains (MIC: 125 μg/mL)[17]
Anti-microbial effects (Clostridium perfringens)Caco-2/In vitro4, 8, 16, 32 μg/mL96 hTargeting type IV pilus (TFP) systemInhibited gliding motility, biofilm formation, and adherence to Caco-2 cells; TFP-encoding genes↓[131]
Bone-protective effectsBone marrow mononu clear (BMM) cells, RAW264.7 cells; Female C57BL/6 mice RANKL (in vitro); Ovariectomized model (in vivo)In vitro and in vivo10, 40, 80, 160 μM (in vitro); 1, 10 mg/kg (in vivo)7 days (in vitro); every 3 days for 6 weeks (in vivo)NF-κB pathwayReduced osteoclast differentiation, TRAP; NFATc1, cathepsin K, MMP-9↓; reduced the bone loss of trabecular bone; BMD, BV/TV, Tb.N↑, Tb.Sp↓; improved trabecular numbers, decreased osteoclasts numbers [138]
Bone-protective effectsPDLCs and BMMs; Female ICR miceOsteogenic medium (in vitro); RANKL, LPS (in vivo)In vitro and in vivo10, 50, 100 μM (in vitro); 3 mg/kg (in vivo)14 days (in vitro); dosing on days 0 and 3 (in vivo)BMP and MAPK pathwaysALP, OPN, OCN, Runx2, Osterix, BMP-2, BMP-4, Smad-4↑; TRAP, cathepsin-K, MMP-9↓; stimulated osteogenic differentiation, increased bone regeneration, inhibited osteoclast differentiation, and suppressed inflammatory bone loss[120]
Bone-protective effectsPrimary chondrocytes;
Male SD rats
Medial collateral ligament transection and
medial meniscal tear on the knee joints
In vitro and in vivo25, 50, 100, 200, 400 μM (in vitro); 0.75, 1.5 μg/kg (in vivo) 24 h (in vitro); every 5 days for 8 weeks (in vivo) Prevented articular cartilage degeneration and
chondrocyte apoptosis via the NF-κB P65 and Bax/Bcl-2/caspase-3 pathway
Type X collagen, cyclooxigenase-2, MMP-3, and MMP-13 expression↓; Runx1, type II collagen, and aggrecan↑; inhibited apoptosis: p-Bad, caspase-3↑, Bax/Bcl-2 ratio↓, improved osteoarthritic injury[139]
Bone-protective effectsTendon-derived stem cells; Male SD ratsTNF-α (in vitro); Achilles tenotomy (in vivo)In vitro and in vivo50, 100 µM (in vitro); 100 µM (in vivo)1 h (in vitro); once a week for 8 weeks (in vivo) NF-κB and MAPK pathwaysMMP-3, MMP-9, MMP-13, iNOS, COX-2, IL-6, IL-10, and collagen I, RUNX-2↓; inhibited apoptosis, senescence, and ossification, and alleviated calcification of tendon[141]
Anti-skin-damage effectsHaCaT cellsUltraviolet-B lightIn vitro0.1, 1, 10 µM24 hAntioxidant, apoptosis, and collagen degradationIntracellular ROS↓, GSH level↑, GPx, CAT expression↑; caspase-3, Bcl-2/Bax ratio↓; attenuated collagen degradation[22]
Anti-skin-damage effectsHaCaT, B16-F10, 293T cells; Kunming mice and New Zealand white rabbitsUltravioletIn vitro and in vivo10, 20, 40, 80, 160, 320 µM24, 48 hApoptosis and pyroptosis pathways; MAPK/JNK/AP-1 pathwayTargeting binding RAR-γ; inhibited UV-induced oxidative damage, inflammatory factor releases and MMP production, reversed the loss of collagen[143]
Antiallergic effectsHMC-1/In vitro10, 25, 50 µM72 hReduced generation of c-chain subunitInhibited the expression of IgE receptor[145]
Antiallergic effectsRBL-2H3 and RAW 264.7 cells/In vitro//Inhibited release of b-hexosaminidase induced by IgEInhibited the passive cutaneous anaphylaxis reaction and inhibited the release of β-hexosaminidase from RBL-2H3 cells induced by IgE[146]
Cardioprotective effectsHuman bloodArachidonic acidIn vitro//Competitive antagonism at thromboxane receptorsInhibited whole blood platelet aggregation[150]
Cardioprotective and cerebroprotective effectsHUVECsH2O2In vitro0.1, 0.2, 0.5, 1, 10 μM24 hPI3K/AKTCell viability↑, LDH↓, SOD↑, GSH-Px↑, MDA↓; Attenuated apoptosis[39]
Cerebroprotective effectsHT-22 cellsOGD/R injuryIn vitro1, 5, 10, 20 mM1 hPI3K/AKT and PPARγ/NF-κB pathwaysCell viability↑; LDH, IL-1β, IL-6, TNF-α, ROS↓; alleviated apoptosis[152]
Cerebroprotective effectsHT-22 cells; Male C57BL/6 miceOGD/R injury (in vitro); Bilateral carotid artery stenosis (BCAS) (in vivo)In vitro and in vivo1, 10, 50, 100, 200 μM (in vitro); 12.5, 25, 50 mg/kg (in vivo)24 h (in vitro); once daily for 15 days (in vivo)Inhibited the TLR4/NF-κB pathwayAlleviated cognitive impairment, hippocampal tissue and myelin damage, and inflammation; Cell viability↑; alleviated apoptosis[153]
Protective effects on the respiratory systemMale guinea pigs OvalbuminIn vivo10, 25 mg/kg14 daysTGF-β1/Smad and TLR4/NF-κB pathwaysInhibited pulmonary fibrosis and airway inflammation: the number of coughs, inflammatory cells, TGF-β1, p-Smad2/3/4, VEGFA, TNF-α, TLR4, MyD88, NF-κB, p-IKKβ↓; Smad7↑; inhibited pulmonary fibrosis[24]
Protective effects on the respiratory system9HTE cellsDexamethasoneIn vitro0.1, 0.2,0.5, 1, 10 μM24 hEnhanced miR-222-3p expression, and inhibited the MAPK pathwayImproved the cell viability, migration, and invasion, and alleviated apoptosis[154]
Protective effects on the respiratory systemPulmonary fibroblasts of male SD ratsBleomycinIn vitro10, 100, 500 μM3 daysEnhanced miR-338* (miR-338-5p) expression, inhibited LPA1 expressionCell viability↓[155]
Neuroprotective effectsPrimary rat microgliaLPSIn vitro///Showed potency of inhibiting NO release from LPS-activated microglia with IC50 values of 9.3 µM[40]
Neuroprotective effectsGBM-8401 and GBM-8901 cells/In vitro25, 50, 100, 200, 300 µM24 hCell cycle arrestGlioblastoma cell viability↓; cell cycle arrest at G0/G1 phase; p-retinoblastoma protein, cyclin-dependent kinase 4 (CDK4)↓; p21 expression↑[160]
Neuroprotective effectsBV-2 cells; Male ICR mice (8 weeks)LPSIn vitro and in vivo25, 50, 100 μM (in vitro); 5 or 10 mg/kg (in vivo)24 h (in vitro); once per day for 5 days (in vivo)NF-κB/ERK/JNK pathwayIntracellular ROS, NO, PGE2, TNF-α, IL-6↓; Extracellular ERK, JNK, iNOS, COX-2↓; Hippocampus MDA, iNOS↓; Serum TNF-α, IL-6 levels↓; TLR4, MyD88 protein↓[115]
Neuroprotective effectsNT2/D1 cells/In vitro10, 20, 30 μM24 hUpregulation of erythropoietin in neuronsInduced transcription of HIF-1α, reduced degradation of HIF-1α-OH[164]
Neuroprotective effectsPC12 cellsβ-Amyloid protein In vitro25, 50, 100 μg/mL12, 24, 48 h/IC50 against MAO-B was 54.36 μg/mL; had a protective effect against Aβ(25–35)-induced cell damage[57]
Neuroprotective effectsSH-SY5YMPP+In vitro0.1, 1, 10 μM24 hOxidative stress, apoptosisCell viability↑, caspase-3 activity and cytochrome c expression, Bax and Bcl-2 levels↓; ROS and NADPH oxidase↓; SOD, CAT, GPx↑[165]
Protective effects on obstructive nephropathyPrimary renal tubular epithelial cells; Male C57BL/6 miceTGF-β1, erastin/RSL3 (in vitro); Unilateral ureteral obstruction (in vivo)In vitro and in vivo20, 40, 60 μM (in vitro); 20 mg/kg (in vivo)24 h (in vitro); daily for 7 consecutive days (in vivo)Inhibited Smad3-mediated ferroptosis and fibrosisBlood urea nitrogen and creatinine, KIM-1↓, alleviated pathological damage and renal interstitial fibrosis[166]
Effects on growth hormone releasePituitary cells/In vitro2.5–20 μg/mL15 min/Promoted the release of growth hormone, twice as effective as in the control group[167]
“↑” means up-regulation; “↓” means down-regulation.

5. Toxicity of Tectorigenin

Evidence shows that tectorigenin affected Ca2+ homeostasis in HepG2 cells [84], HSC-T6 cells [109], isolated liver parenchymal cells [168], and Madin-Darby Canine Kidney cells [169]. The change in Ca2+ concentration may trigger some cellular responses, so caution is needed when using tectorigenin. The results of Cheng et al. [169] showed that Ca2+ influx might be evoked by protein kinase C (PKC)-insensitive store-operated Ca2+ entry and Ca2+ release from the endoplasmic reticulum via phospholipase C (PLC)-associated pathways. Tectorigenin at concentrations between 10–60 μM exerted renal cytotoxicity, and the IC50 was approximately 45.5 μM. However, using a cytosolic Ca2+ chelating agent did not affect the cytotoxicity of tectorigenin, suggesting that tectorigenin caused cell death in a Ca2+-independent manner [169]. In terms of other cell models, tectorigenin with a concentration below 80 μmol/L did not cause significant cytotoxicity on BMM cells or RAW267.4 after administration for 24 h or 48 h, but 160 μmol/L tectorigenin significantly decreased the cell viability of BMM cells and RAW267.4 within 48 h [138]. In another study, 200 μM tectorigenin exhibited cytotoxicity in tendon-derived stem cells within 24 h [141]. In addition to the cytotoxicity studies in vitro, in vivo toxicological studies of tectorigenin were also carried out in some experimental animals by a variety of scholars. An acute toxicity study revealed that the LD50 of tectorigenin by intragastric administration was 1.78 g/kg in swiss mice, and no significant differences in body weights, food consumption, or hematological or biochemical parameters were observed at tectorigenin’s doses not exceeding 300 mg/kg during a consecutive 28-day treatment [118]. In another acute toxicity test, tectorigenin at a dose of 5 g/kg/day was orally administrated to mice for 14 days, and it did not exhibit acute toxicity [105]. Similarly, tectorigenin (75 mg/kg/day) orally administrated to C57BL/6J mice for 14 days caused no significant renal or splenic toxicity [23]. In streptozotocin and high-fat diet-induced diabetes and obesity models, SD rats gavaged with 50 and 100 mg/kg/day tectorigenin for 14 days showed no significant side effects, including body weight gain, fluid retention, or cardiac hypertrophy [101]. These studies show that tectorigenin may be toxic above a certain concentration or duration, whereas in most cases, tectorigenin did not exert toxicity during the therapeutic range [120,143]. However, the implications of the toxic effects of tectorigenin for clinical use are unclear, so it is necessary to perform a risk assessment of human exposure to tectorigenin.

6. Pharmacokinetics of Tectorigenin

The pharmacokinetic properties, including absorption, distribution, metabolism, and excretion, are critical for determining the optimal dose of compounds and the frequency of administration as well as for ensuring drug adherence, and all of these play an important role in clinical application [170]. After rat oral administration, tectorigenin can be directly absorbed from the intestinal tract by passive diffusion, and it then undergoes glucuronidation and/or sulfation metabolic pathways by UDP-glucuronosyltransferases (UGT) and sulfotransferases to form corresponding metabolites [171]. UGT1A1 and UGT1A9 were identified as the primary enzymes catalyzing the glucuronidation of tectorigenin in human liver microsomes [172]. Several investigators have conducted HPLC-MS analyses of metabolites in plasma, urine, and bile after oral tectorigenin consumption in rats. The phase II metabolites of tectorigenin in rat bile and urine were identified by HPLC combined with ion trap tandem and time-of-flight mass spectrometry, and they were either glucuronide and/or sulfate conjugated compounds [173,174]. Shi et al. [175] detected 26 metabolites in rat urine after oral administration of tectorigenin (65 and 130 mg/kg). The results demonstrated that the potential biochemical transformations of tectorigenin included mono-glucuronidation, mono-sulfation, bis-glucuronidation-sulfation, bis-sulfation, bis-glucuronidation, glucuronidation after hydroxylation, methylation, methoxylation, demethylation, etc. [175]. In another study, nine tectorigenin metabolites, including mono-glucuronide and mono-sulfate conjugates, glucuronide-sulfate bis-conjugate, bis-sulfate conjugate, demethylation glucuronide conjugate, and methoxylation glucuronide conjugate, were identified in rats plasma after oral dosing at 65 and 130 mg/kg, and the results of the pharmacokinetic study on six major metabolites showed that the plasma level of mono-glucuronide conjugate tectorigenin-7-O-glucuronide was much higher than tectorigenin and other metabolites [171]. Overall, the main metabolic pathways of tectorigenin in rats include glucuronidation, sulfation, demethylation, and methoxylation, which are displayed in Figure 4.
Nevertheless, due to its low water solubility and low permeability, tectorigenin showed poor absolute bioavailability after oral administration in rats. It was reported that oral tectorigenin (130 mg/kg) administration resulted in a peak concentration (Cmax) of 12.0 ± 0.63 μmol/L, an area under the curve (AUC0–t) of 84.2 ± 8.15 μmol/L × h, whereas the long half-life period (t½) 11.7 ± 5.74 h showed that tectorigenin had a long action time [171]. Another study indicated that oral administration of tectorigenin (80 mg/kg) resulted in a Cmax of 1.46 ± 0.30 μg/mL and AUC(0–t) of 7.31 ± 1.20 μg/mL × h [176]. Additionally, sublingual intravenous administration of tectorigenin (5 mg/kg) resulted in an AUC(0–t) of 219 ± 94 ng/mL × h [177]. Pharmacokinetic studies of some TCMs, such as Shejin-liyan Granule, extracts of Belamcandae Rhizoma and rhizome of Iris tectorum, also described the pharmacokinetic characteristics of tectorigenin as the main component [178,179,180]. Tectorigenin was found to exhibit the highest exposure among eight constituents after oral administration of Shejin-liyan Granule [178]. The pharmacokinetic parameters of tectorigenin from multiple studies were summarized in Table 3. Yang et al. [180] found an appearance of a double-peak on the plasma concentration-time curves of tectorigenin after oral administration of Iris tectorum extract, and the same phenomenon was observed in some other studies as well [178,179], which suggested that tectorigenin was undergoing enterohepatic circulation, thus resulting in a long excretion time [180]. Moreover, tectorigenin is a metabolite of some isoflavone glycosides and aglycones, including tectoridin [127], kakkalide [181], and irisolidone [182]. Tectorigenin can be quickly generated through the transformation of tectoridin by human intestinal microflora [28,183] and eliminated through urine and feces [184]. However, tectorigenin exerted greater potential than that of tectoridin in many respects [91,185,186]. Therefore, tectoridin may be metabolized into tectorigenin to play a pharmacological role [127]. Although the above research has elucidated some pharmacokinetic characteristics of tectorigenin, more detailed metabolic mechanisms need to be elucidated by conducting more laboratory animal experiments, and due to its low oral bioavailability, more studies should also be carried out to improve the bioavailability.

7. Delivery Strategy of Tectorigenin

As described above, the low bioavailability of tectorigenin has severely restricted its clinical application. Generally, many natural flavonoids have the disadvantages of poor water solubility, susceptibility to oxidation, and low bioavailability [187,188,189,190]. These problems can be solved by some methods, such as prodrug design [191], formulation optimization [192], and nano-drug delivery systems, including liposomes [193,194], nanoparticles [195,196], phytosomes [197], dendrimers [198], polymer-drug conjugates [199], and microparticles [200]. Similarly, numerous attempts have been made to increase tectorigenin’s water solubility, stability, and bioavailability by employing diverse delivery strategies. An attempt was made for the first time in 2011 to improve the solubility of tectorigenin by sulfonation, a chemical modification [60]. After sulfonation, the solubility was improved 9-fold and the antioxidant activity increased as well. Solid dispersion is an applicable strategy for the structural transformation of insoluble compounds that have been widely studied to enhance solubility and bioavailability [201]. Shuai et al. [202] prepared the solid dispersions of tectorigenin and the % release was increased 4.35-fold, exceeding pure tectorigenin, and an in vivo experiment proved that the oral bioavailability in rats was also increased. Then, the above team reported that the tectorigenin-loaded self-microemulsifying drug delivery system (TG-SMEDDS) improved physical stability could be kept stable for at least 3 months at room temperature and exhibited good dissolution behavior [203]. An in vivo study proved that TG-SMEDDS had better oral bioavailability both in bile duct ligation and bile duct non-ligation rats. Dai et al. [143] displayed a new drug delivery system that self-assembled to form hyaluronic acid-polyethyleneimine-loaded tectorigenin-nano. When tectorigenin-nano was presented and delivered in the form of nanoparticles, it was better absorbed by skin tissue, improving the bioavailability and exhibiting a better effect than tectorigenin itself [143]. Controlled-release formulations are dosage forms that can release drugs at a regular, quantitative, and uniform rate, and can keep the blood drug concentration for a long time, thereby prolonging the duration of action time and reducing the total dose required for potency [204]. Tectorigenin intragastric floating sustained-release tablets were reported by Wang et al. [205], which could float on the gastric fluid after oral administration, prolong the retention time in the stomach, gradually dissolve, slowly release the drug, and then increase absorption and improve bioavailability.

8. Conclusions and Future Perspectives

Tectorigenin, an effective isoflavone, is abundant in a variety of medicinal plants, such as Belamcandae Rhizoma, Puerariae flos, and Chinese water chestnut peel, which are TCMs often used for clearing heat, detoxifying, and diminishing inflammation in many diseases [19,20,48]. Tectorigenin has attracted wide attention as one of the main functional activity ingredients of these TCMs in numerous studies [20,48,114]. This paper reviews comprehensive sources, extraction and synthesis, pharmacology, toxicity, pharmacokinetics, and delivery strategy aspects of tectorigenin in order to provide an important source of information for improving the potential of feasible treatment for a variety of disorders.
The extraction and isolation methods for tectorigenin are relatively simple and mature. Conventional extraction reagents such as methanol, ethanol, and ethyl acetate combined with chromatographic methods can realize the extraction and purification of tectorigenin [40,52,55]. In comparison, the de novo synthesis method of tectorigenin is more complicated. Therefore, extraction from plants with a high tectorigenin content is still the mainstream method to obtain tectorigenin at present. Tectorigenin has multifarious promising medicinal values in anticancer, antioxidant, antimicrobial, anti-inflammatory, and neuroprotective treatments. Through the interpretation and summary of the literature related to tectorigenin, we found that tectorigenin, as well as Chinese herbal extracts containing tectorigenin, exert extensive pharmacological effects by regulating several key enzymes function and the expression of related genes and proteins, including up-regulating antioxidant enzymes such as SOD, GPx and catalase, down-regulating ERβ, IL-6, IL-1β, TNF-α, COX-2, iNOS, and PG-E2, caspase-3/8/9, MMP, p53, and modulating PPARγ/NF-κB, PI3K/AKT, TLR4/NF-κB, IKKβ/NF-κB/JNK, ERK/JNK, MAPK/JNK/AP-1, AKT/MAPK, and TGF-β1/Smad. Existing evidence also suggests that tectorigenin could regulate diverse events involved in cell invasion, immortalization, cell cycle arrest, and cell apoptosis, which obviously supports its traditional application in the treatment of cancer and other diseases. The group and the location of tectorigenin substituents are the key factors for pharmacological activity. Structure-activity relationship analysis of flavonoids suggested that the 5-hydroxyl group and the 6-methoxyl group of the tectorigenin structure are crucial for the inhibitory activity of tumor cells. However, adverse effects of tectorigenin have also been reported, including renal cytotoxicity and hepatotoxicity [84,109,168,169]. The cytotoxicity of tectorigenin is mainly related to the time and concentration of the drug’s administration, whereas in most pharmacodynamic studies, tectorigenin did not exert toxicity during the therapeutic range [23,101,105,120,143]. To ensure the safety of tectorigenin’s clinical application, it is necessary to conduct a risk assessment of human exposure to tectorigenin, carry out more in-depth research on the mechanism of toxicity, and explore ways to reduce its toxicity in the future. In terms of the current pharmacokinetic study, the main metabolic pathways of tectorigenin in rats are glucuronidation, sulfation, demethylation, and methoxylation [171]. However, tectorigenin exhibited poor absolute bioavailability, mainly because of its water-insoluble chemical structure [202]. This may be one of the reasons why fewer experiments are performed on animal models. Meanwhile, due to some differences in pharmacokinetics between the model and normal animals, pharmacokinetic studies should also be carried out on the corresponding model animals.
To advance tectorigenin into a viable clinical therapeutic drug, we summarize several key directions for future research on tectorigenin to conclude this study. (1) Although various biological activities of tectorigenin are confirmed in vitro and in vivo, the molecular mechanisms of action are still unclear, especially the binding sites of definitive target proteins, which remain undetermined. Hence, it is of great significance to further explore the mechanism of its biological activity at the molecular level. (2) The safety of a new drug is especially important before it is used in clinical practice, and more systematic studies of dosage should be conducted to balance both its pharmacological effect and its toxicity. It is also necessary to carry out toxicity evaluations of multiple organs. (3) At present, there are few studies on pharmacokinetics, which should be conducted on more animals in the future, such as dogs and monkeys. Different methods of administration are also worth trying. (4) Pro-drug design, formulation optimization, and novel drug delivery systems, such as nanotechnology, including liposomes, nanoparticles, phytosomes, dendrimers, polymer-drug conjugates, and microparticles, could be attempted to increase tectorigenin’s water solubility, stability, and bioavailability. (5) Structural modification is still an efficient and promising way to obtain tectorigenin derivatives with higher pharmacological activity and relatively better bioavailability, while some key active substituent groups need to be retained.

Author Contributions

J.R. and F.F. designed and completed the original manuscript; D.D. and Q.X. provided guidance and modified the language in the paper; C.H. and Y.W. lent a hand for collecting the literature. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Program of Science and Technology Department of Sichuan Province (D.D., 2023NSFSC1755) and the 1.3.5 Project for Disciplines of Excellence, West China Hospital, Sichuan University (X.Q., ZYJC18005).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data are contained within the manuscript.

Acknowledgments

We thank Figdraw (www.figdraw.com) for the assistance in creating the graphical abstract.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not applicable.

References

  1. Sin, V.J.; Anand, G.S.; Koh, H.L. Botanical medicine and natural products used for erectile dysfunction. Sex. Med. Rev. 2021, 9, 568–592. [Google Scholar] [CrossRef] [PubMed]
  2. Newman, D.J.; Cragg, G.M. Natural products as sources of new drugs over the nearly four decades from 01/1981 to 09/2019. J. Nat. Prod. 2020, 83, 770–803. [Google Scholar] [CrossRef] [Green Version]
  3. Mutha, R.E.; Tatiya, A.U.; Surana, S.J. Flavonoids as natural phenolic compounds and their role in therapeutics: An overview. Futur. J. Pharm. Sci. 2021, 7, 25. [Google Scholar] [CrossRef] [PubMed]
  4. Maleki, S.J.; Crespo, J.F.; Cabanillas, B. Anti-inflammatory effects of flavonoids. Food Chem. 2019, 299, 125124. [Google Scholar] [CrossRef]
  5. Hernandez, I.; Alegre, L.; Van Breusegem, F.; Munne-Bosch, S. How relevant are flavonoids as antioxidants in plants? Trends Plant Sci. 2009, 14, 125–132. [Google Scholar] [CrossRef]
  6. Cushnie, T.P.; Lamb, A.J. Antimicrobial activity of flavonoids. Int. J. Antimicrob. Agents. 2005, 26, 343–356. [Google Scholar] [CrossRef] [PubMed]
  7. Proenca, C.; Ribeiro, D.; Freitas, M.; Carvalho, F.; Fernandes, E. A comprehensive review on the antidiabetic activity of flavonoids targeting PTP1B and DPP-4: A structure-activity relationship analysis. Crit. Rev. Food Sci. Nutr. 2022, 62, 4095–4151. [Google Scholar] [CrossRef]
  8. Calderaro, A.; Patane, G.T.; Tellone, E.; Barreca, D.; Ficarra, S.; Misiti, F.; Lagana, G. The neuroprotective potentiality of flavonoids on Alzheimer’s disease. Int. J. Mol. Sci. 2022, 23, 14835. [Google Scholar] [CrossRef]
  9. Sun, Q.; Liu, Q.; Zhou, X.; Wang, X.; Li, H.; Zhang, W.; Yuan, H.; Sun, C. Flavonoids regulate tumor-associated macrophages—From structure-activity relationship to clinical potential (Review). Pharmacol. Res. 2022, 184, 106419. [Google Scholar] [CrossRef]
  10. Williamson, G.; Kay, C.D.; Crozier, A. The bioavailability, transport, and bioactivity of dietary flavonoids: A review from a historical perspective. Compr. Rev. Food Sci. Food Saf. 2018, 17, 1054–1112. [Google Scholar] [CrossRef] [Green Version]
  11. Terahara, N. Flavonoids in foods: A review. Nat. Prod. Commun. 2015, 10, 521–528. [Google Scholar] [CrossRef] [Green Version]
  12. Hodgson, J.M.; Croft, K.D. Tea flavonoids and cardiovascular health. Mol. Aspects Med. 2010, 31, 495–502. [Google Scholar] [CrossRef] [PubMed]
  13. Kafi, M.K.; Bolvari, N.E.; Pour, S.M.; Moghadam, S.K.; Shafaei, N.; Karimi, E.; Oskoueian, E. Encapsulated phenolic compounds from Ferula gummosa leaf: A potential phytobiotic against Campylobacter jejuni infection. J. Food Process. Pres. 2022, 46, e16802. [Google Scholar]
  14. Poorbagher, M.R.M.; Karimi, E.; Oskoueian, E. Hepatoprotective effect of nanoniosome loaded Myristica fragrans phenolic compounds in mice-induced hepatotoxicity. J. Cell. Mol. Med. 2022, 26, 5517–5527. [Google Scholar] [CrossRef] [PubMed]
  15. Karimi, E.; Oskoueian, E.; Karimi, A.; Noura, R.; Ebrahimi, M. Borago officinalis L. flower: A comprehensive study on bioactive compounds and its health-promoting properties. J. Food Meas. Charact. 2018, 12, 826–838. [Google Scholar] [CrossRef]
  16. Oskoueian, E.; Karimi, E.; Noura, R.; Ebrahimi, M.; Shafaei, N.; Karimi, E. Nanoliposomes encapsulation of enriched phenolic fraction from pistachio hulls and its antioxidant, anti-inflammatory, and anti-melanogenic activities. J. Microencapsul. 2020, 37, 1–13. [Google Scholar] [CrossRef]
  17. Joung, D.K.; Mun, S.H.; Lee, K.S.; Kang, O.H.; Choi, J.G.; Kim, S.B.; Gong, R.; Chong, M.S.; Kim, Y.C.; Lee, D.S.; et al. The antibacterial assay of tectorigenin with detergents or ATPase inhibitors against methicillin-resistant Staphylococcus aureus. Evid. Based Complement. Alternat. Med. 2014, 2014, 716509. [Google Scholar] [CrossRef] [Green Version]
  18. Han, J.; Xu, K.; Yan, Q.; Sui, W.; Zhang, H.; Wang, S.; Zhang, Z.; Wei, Z.; Han, F. Qualitative and quantitative evaluation of Flos Puerariae by using chemical fingerprint in combination with chemometrics method. J. Pharm. Anal. 2022, 12, 489–499. [Google Scholar] [CrossRef]
  19. Wozniak, D.; Matkowski, A. Belamcandae chinensis rhizome—A review of phytochemistry and bioactivity. Fitoterapia 2015, 107, 1–14. [Google Scholar] [CrossRef]
  20. Chen, C.; Li, X.; Kano, Y.; Yuan, D.; Qu, J. Oriental traditional herbal Medicine--Puerariae Flos: A systematic review. J. Ethnopharmacol. 2023, 306, 116089. [Google Scholar] [CrossRef]
  21. Guo, Y.; Chen, Y.H.; Cheng, Z.H.; Ou-Yang, H.N.; Luo, C.; Guo, Z.L. Tectorigenin inhibits osteosarcoma cell migration through downregulation of matrix metalloproteinases in vitro. Anticancer Drugs 2016, 27, 540–546. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Noh, D.; Choi, J.G.; Huh, E.; Oh, M.S. Tectorigenin, a flavonoid-based compound of leopard lily rhizome, attenuates UV-B-induced apoptosis and collagen degradation by inhibiting oxidative stress in human keratinocytes. Nutrients 2018, 10, 1998. [Google Scholar] [CrossRef] [Green Version]
  23. Xiang, J.; Yang, G.; Ma, C.; Wei, L.; Wu, H.; Zhang, W.; Tao, X.; Jiang, L.; Liang, Z.; Kang, L.; et al. Tectorigenin alleviates intrahepatic cholestasis by inhibiting hepatic inflammation and bile accumulation via activation of PPARγ. Br. J. Pharmacol. 2021, 178, 2443–2460. [Google Scholar] [CrossRef] [PubMed]
  24. Wang, Y.; Jing, W.; Qu, W.; Liu, Z.; Zhang, D.; Qi, X.; Liu, L. Tectorigenin inhibits inflammation and pulmonary fibrosis in allergic asthma model of ovalbumin-sensitized guinea pigs. J. Pharm. Pharmacol. 2020, 72, 956–968. [Google Scholar] [CrossRef]
  25. Zhang, Y.Y.; Wang, Q.; Qi, L.W.; Qin, X.Y.; Qin, M.J. Characterization and determination of the major constituents in Belamcandae Rhizoma by HPLC-DAD-ESI-MS(n). J. Pharm. Biomed. Anal. 2011, 56, 304–314. [Google Scholar] [CrossRef] [PubMed]
  26. Noh, D.; Choi, J.G.; Lee, Y.B.; Jang, Y.P.; Oh, M.S. Protective effects of Belamcandae Rhizoma against skin damage by ameliorating ultraviolet-B-induced apoptosis and collagen degradation in keratinocytes. Environ. Toxicol. 2019, 34, 1354–1362. [Google Scholar] [CrossRef]
  27. Farag, S.F.; Kimura, Y.; Ito, H.; Takayasu, J.; Tokuda, H.; Hatano, T. New isoflavone glycosides from Iris spuria L. (Calizona) cultivated in Egypt. J. Nat. Med. 2009, 63, 91–95. [Google Scholar] [CrossRef] [PubMed]
  28. Sun, Y.; Liu, Z.; Wang, J. Ultrasound-assisted extraction of five isoflavones from Iris tectorum Maxim. Sep. Purif. Technol. 2011, 78, 49–54. [Google Scholar] [CrossRef]
  29. Yang, Q.; Zheng, X.; Wei, J.; Zhang, C. Determination of two isoflavones and their antitumor activity in Iris japonica Thunb. J. Sichuan Norm. Univ. (Nat. Sci.) 2020, 43, 5. [Google Scholar]
  30. Shi, G.R. Study on Active Substances and Functions of Leaf of Rehmannia glutinosa Libosch and Whole Plants of Iris japonica Thunb; Chinese Academy of Medical Sciences: Beijing, China, 2017. [Google Scholar]
  31. Di, Z.Z.; Zhang, Y.; Wang, G.H.; Zou, G.X.; Li, G.X. Research of quality control of Iris dichotoma based on medicinal material base of Belamcanda chinensis. Liaoning J. Tradit. Chin. Med. 2017, 44, 4. [Google Scholar]
  32. Zhao, Y.; Li, G.X. Determination of effective active ingredients of three plants in Iridaceae plants at different havest periods by HPLC. J. Liaoning Univ. TCM 2012, 14, 39–41. [Google Scholar]
  33. Mosihuzzman, M.; Naheed, S.; Hareem, S.; Talib, S.; Abbas, G.; Khan, S.N.; Choudhary, M.I.; Sener, B.; Tareen, R.B.; Israr, M. Studies on α-glucosidase inhibition and anti-glycation potential of Iris loczyi and Iris unguicularis. Life Sci. 2013, 92, 187–192. [Google Scholar] [CrossRef]
  34. Amin, A.; Mokhdomi, T.A.; Bukhari, S.; Wani, S.H.; Wafai, A.H.; Lone, G.N.; Qadri, A.; Qadri, R.A. Tectorigenin ablates the inflammation-induced epithelial-mesenchymal transition in a co-culture model of human lung carcinoma. Pharmacol. Rep. 2015, 67, 382–387. [Google Scholar] [CrossRef] [PubMed]
  35. Bhat, G.A.; Mir, F.; Shawl, A.S.; Ganai, B.A.; Kamili, A.N.; Masood, A.; Tantry, M.A. Crocetenone, a new rotenoid with an unusual trans-fused ring system from Iris crocea. Nat. Prod. Commun. 2015, 10, 503–504. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Wani, S.H.; Amin, A.; Sofi, S.N.; Mokhdomi, T.A.; Bukhari, S.; Hassan, Q.P.; Qadri, R.A. RP- HPLC facilitated quantitative analysis of tectorigenin in the different species of iris plant and evaluation of its invitro anticancer potential. Int. J. Curr. Res. 2013, 5, 206–211. [Google Scholar]
  37. Okba, M.M.; Abdel Baki, P.M.; Abu-Elghait, M.; Shehabeldine, A.M.; El-Sherei, M.M.; Khaleel, A.E.; Salem, M.A. UPLC-ESI-MS/MS profiling of the underground parts of common Iris species in relation to their anti-virulence activities against Staphylococcus aureus. J. Ethnopharmacol. 2022, 282, 114658. [Google Scholar] [CrossRef]
  38. Mykhailenko, O.; Kovalyov, V.; Kovalyov, S.; Krechun, A. Isoflavonoids from the rhizomes of Iris hungarica and antibacterial activity of the dry rhizomes extract. Ars Pharm. 2018, 58, 39–45. [Google Scholar] [CrossRef]
  39. Chen, X.; Zhang, W.; Sun, L.; Lian, Y. Tectorigenin protect HUVECs from H2O2 -induced oxidative stress injury by regulating PI3K/Akt pathway. Tissue. Cell. 2021, 68, 101475. [Google Scholar] [CrossRef]
  40. Yuan, D.; Xie, Y.Y.; Bai, X.; Wu, X.; Yang, J.Y.; Wu, C.F. Inhibitory activity of isoflavones of Pueraria flowers on nitric oxide production from lipopolysaccharide-activated primary rat microglia. J. Asian Nat. Prod. Res. 2009, 11, 471–481. [Google Scholar] [CrossRef]
  41. Kim, Y.; Kim, J.; Son, S.R.; Kim, J.Y.; Choi, J.H.; Jang, D.S. Chemical constituents of the flowers of Pueraria lobata and their cytotoxic properties. Plants 2022, 11, 1651. [Google Scholar] [CrossRef]
  42. Lee, K.T.; Sohn, I.C.; Kim, D.H.; Choi, J.W.; Kwon, S.H.; Park, H.J. Hypoglycemic and hypolipidemic effects of tectorigenin and kaikasaponin III in the streptozotocin-induced diabetic rat and their antioxidant activity in vitro. Arch. Pharm. Res. 2000, 23, 461–466. [Google Scholar] [CrossRef] [PubMed]
  43. Li, L.; Liu, J.Z.; Luo, M.; Wang, W.; Huang, Y.Y.; Efferth, T.; Wang, H.M.; Fu, Y.J. Efficient extraction and preparative separation of four main isoflavonoids from Dalbergia odorifera T. Chen leaves by deep eutectic solvents-based negative pressure cavitation extraction followed by macroporous resin column chromatography. J. Chromatogr. B 2016, 1033–1034, 40–48. [Google Scholar] [CrossRef]
  44. Umehara, K.; Nemoto, K.; Matsushita, A.; Terada, E.; Monthakantirat, O.; De-Eknamkul, W.; Miyase, T.; Warashina, T.; Degawa, M.; Noguchi, H. Flavonoids from the heartwood of the thai medicinal plant Dalbergia parviflora and their effects on estrogenic-responsive human breast cancer cells. J. Nat. Prod. 2009, 72, 2163–2168. [Google Scholar] [CrossRef] [PubMed]
  45. Lo, W.L.; Wu, C.C.; Chang, F.R.; Wang, W.Y.; Khalil, A.T.; Lee, K.H.; Wu, Y.C. Antiplatelet and anti-HIV constituents from Euchresta formosana. Nat. Prod. Res. 2003, 17, 91–97. [Google Scholar] [CrossRef]
  46. Ru, J.; Li, P.; Wang, J.; Zhou, W.; Li, B.; Huang, C.; Li, P.; Guo, Z.; Tao, W.; Yang, Y.; et al. TCMSP: A database of systems pharmacology for drug discovery from herbal medicines. J. Cheminform. 2014, 6, 13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Moon, H.I.; Jung, J.C.; Lee, J. Aldose reductase inhibitory effect by tectorigenin derivatives from Viola hondoensis. Bioorg. Med. Chem. 2006, 14, 7592–7594. [Google Scholar] [CrossRef]
  48. Zhan, G.; Pan, L.; Tu, K.; Jiao, S. Antitumor, antioxidant, and nitrite scavenging effects of Chinese water chestnut (Eleocharis dulcis) peel flavonoids. J. Food Sci. 2016, 81, 2578–2586. [Google Scholar] [CrossRef]
  49. Shin, K.H.; Kim, Y.P.; Lim, S.S.; Lee, S.; Ryu, N.; Yamada, M.; Ohuchi, K. Inhibition of prostaglandin E2 production by the isoflavones tectorigenin and tectoridin isolated from the rhizomes of Belamcanda chinensis. Planta Med. 1999, 65, 776–777. [Google Scholar] [CrossRef]
  50. Oh, K.B.; Kang, H.; Matsuoka, H. Detection of antifungal activity in Belamcanda chinensis by a single-cell bioassay method and isolation of its active compound, tectorigenin. Biosci. Biotechnol. Biochem. 2001, 65, 939–942. [Google Scholar] [CrossRef]
  51. Wozniak, D.; Janda, B.; Kapusta, I.; Oleszek, W.; Matkowski, A. Antimutagenic and anti-oxidant activities of isoflavonoids from Belamcanda chinensis (L.) DC. Mutat. Res. 2010, 696, 148–153. [Google Scholar] [CrossRef]
  52. Miyazawa, M.; Sakano, K.; Nakamura, S.; Kosaka, H. Antimutagenic activity of isoflavone from Pueraria lobata. J. Agric. Food Chem. 2001, 49, 336–341. [Google Scholar] [CrossRef] [PubMed]
  53. Wang, Q.; Cheng, X.L.; Li, H.; Qin, X.Y.; Ge, C.Y.; Liu, R.; Qi, L.W.; Qin, M.J. Application of an efficient strategy for discovery and purification of bioactive compounds from chinese herbal medicines, a case study on the Puerariae thomsonii flos. J. Pharm. Biomed. Anal. 2013, 75, 25–32. [Google Scholar] [CrossRef] [PubMed]
  54. Fang, R.; Houghton, P.J.; Hylands, P.J. Cytotoxic effects of compounds from Iris tectorum on human cancer cell lines. J. Ethnopharmacol. 2008, 118, 257–263. [Google Scholar] [CrossRef] [PubMed]
  55. Monthakantirat, O.; De-Eknamkul, W.; Umehara, K.; Yoshinaga, Y.; Miyase, T.; Warashina, T.; Noguchi, H. Phenolic constituents of the rhizomes of the thai medicinal plant Belamcanda chinensis with proliferative activity for two breast cancer cell lines. J. Nat. Prod. 2005, 68, 361–364. [Google Scholar] [CrossRef]
  56. Jung, S.H.; Lee, Y.S.; Lim, S.S.; Lee, S.; Shin, K.H.; Kim, Y.S. Antioxidant activities of isoflavones from the rhizomes of Belamcanda chinensis on carbon tetrachloride-induced hepatic injury in rats. Arch. Pharm. Res. 2004, 27, 184–188. [Google Scholar] [CrossRef]
  57. Li, S.; Li, Y.; Wang, Y.; Li, R.; Niu, H.; Liu, C.; Zhang, Y. Ionic-liquid-based ultrasound-assisted extraction combined with countercurrent chromatography and semipreparative LC for the preparation of monoamine oxidase B inhibitors from Pueraria thomsonii. J. Sep. Sci. 2022, 45, 1116–1127. [Google Scholar] [CrossRef]
  58. Guo, C.L. Extraction, Purification and Microbial Biotransformation of Tectorigenin Isolated from both Belamcanda Chinensis and the Flowers of Pueraria thunbergiana; Agricultural University of Hebei: Baoding, China, 2013. [Google Scholar]
  59. Kim, M.Y.; Yoo, Y.M.; Chung, I.M.; Nam, J.H.; Park, H.J. Quantitative analysis of the isoflavone content in the flower and the root of Pueraria thunbergiana before and after acid hydrolysis. Nat. Prod. Sci. 2008, 14, 187–191. [Google Scholar]
  60. Han, T.; Cheng, G.; Liu, Y.; Yang, H.; Hu, Y.T.; Huang, W. In vitro evaluation of tectoridin, tectorigenin and tectorigenin sodium sulfonate on antioxidant properties. Food Chem. Toxicol. 2012, 50, 409–414. [Google Scholar] [CrossRef]
  61. Li, Y.; Wang, Q.; He, J.; Yan, J.; Li, H. Fluorescence spectroscopy and docking study in two flavonoids, isolated tectoridin and its aglycone tectorigenin, interacting with human serum albumin: A comparison study. Luminescence 2016, 31, 38–46. [Google Scholar] [CrossRef] [PubMed]
  62. Kagal, S.; Karmarkar, S.; Venkataraman, K. Synthetical experiments in the chromone group: Part XXXII. A synthesis of tectorigenin. In Proceedings of the Indian Academy of Sciences-Section A; Springer: Berlin/Heidelberg, Germany, 1956. [Google Scholar]
  63. Baker, W.; Downing, D.F.; Floyd, A.J.; Gilbert, B.; Ollis, W.D. Synthesis of isoflavones. V. Irigenin and tectorigenin. J. Chem. Soc. Perkin. 1 1970, 9, 1219–1223. [Google Scholar] [CrossRef]
  64. Baker, W.; Downing, D.F.; Floyd, A.J.; Gilbert, B.; Ollis, W.D.; Russell, R.C. Synthesis of irigenin and of tectorigenin, derivatives of 5,7-dihydroxy-6-methoxyisoflavone. Tetrahedron Lett. 1960, 1, 6–10. [Google Scholar] [CrossRef]
  65. Varady, J. A new method for the ring isomerization of isoflavones. Direct synthesis of tectorigenin, 4′-methyl-tectorigenin, caviunin and other isoflavones. Tetrahedron Lett. 1965, 6, 4273–4275. [Google Scholar] [CrossRef]
  66. Xiao, Z.P.; Fang, R.Q.; Shi, L.; Ding, H.; Zhu, H.L. Synthesis, crystal structure, and growth inhibition of human hepatoma cell (HepG2) of polyphenolic compounds based on gallates. Canad. J. Chem. 2011, 85, 951–957. [Google Scholar] [CrossRef]
  67. Xiao, Z.P.; Li, H.Q.; Xue, J.Y.; Shi, L.; Zhu, H.L. Efficient method for the synthesis of tectorigenin. Synth. Commun. 2008, 38, 525–529. [Google Scholar] [CrossRef]
  68. Strom, S.S.; Yamamura, Y.; Duphorne, C.M.; Spitz, M.R.; Babaian, R.J.; Pillow, P.C.; Hursting, S.D. Phytoestrogen intake and prostate cancer: A case-control study using a new database. Nutr. Cancer 1999, 33, 20–25. [Google Scholar] [CrossRef]
  69. Stettner, M.; Kaulfuss, S.; Burfeind, P.; Schweyer, S.; Strauss, A.; Ringert, R.H.; Thelen, P. The relevance of estrogen receptor-beta expression to the antiproliferative effects observed with histone deacetylase inhibitors and phytoestrogens in prostate cancer treatment. Mol. Cancer Ther. 2007, 6, 2626–2633. [Google Scholar] [CrossRef] [Green Version]
  70. Branca, F. Dietary phyto-oestrogens and bone health. Proc. Nutr. Soc. 2003, 62, 877–887. [Google Scholar] [CrossRef] [PubMed]
  71. Lehmann, L.; Esch, H.L.; Wagner, J.; Rohnstock, L.; Metzler, M. Estrogenic and genotoxic potential of equol and two hydroxylated metabolites of daidzein in cultured human ishikawa cells. Toxicol. Lett. 2005, 158, 72–86. [Google Scholar] [CrossRef]
  72. Beck, V.; Rohr, U.; Jungbauer, A. Phytoestrogens derived from red clover: An alternative to estrogen replacement therapy? J. Steroid Biochem. Mol. Biol. 2005, 94, 499–518. [Google Scholar] [CrossRef] [PubMed]
  73. Thelen, P.; Scharf, J.G.; Burfeind, P.; Hemmerlein, B.; Wuttke, W.; Spengler, B.; Christoffel, V.; Ringert, R.H.; Seidlova-Wuttke, D. Tectorigenin and other phytochemicals extracted from leopard lily Belamcanda chinensis affect new and established targets for therapies in prostate cancer. Carcinogenesis 2005, 26, 1360–1367. [Google Scholar] [CrossRef]
  74. Seidlova-Wuttke, D.; Hesse, O.; Jarry, H.; Rimoldi, G.; Thelen, P.; Christoffel, V.; Wuttke, W. Belamcanda chinensis and the thereof purified tectorigenin have selective estrogen receptor modulator activities. Phytomedicine 2004, 11, 392–403. [Google Scholar] [CrossRef] [PubMed]
  75. Morrissey, C.; Bektic, J.; Spengler, B.; Galvin, D.; Christoffel, V.; Klocker, H.; Fitzpatrick, J.M.; Watson, R.W. Phytoestrogens derived from Belamcanda chinensis have an antiproliferative effect on prostate cancer cells in vitro. J. Urol. 2004, 172, 2426–2433. [Google Scholar] [CrossRef] [PubMed]
  76. Barzaman, K.; Karami, J.; Zarei, Z.; Hosseinzadeh, A.; Kazemi, M.H.; Moradi-Kalbolandi, S.; Safari, E.; Farahmand, L. Breast cancer: Biology, biomarkers, and treatments. Int. Immunopharmacol. 2020, 84, 106535. [Google Scholar] [CrossRef] [PubMed]
  77. Lacroix, M.; Leclercq, G. Relevance of breast cancer cell lines as models for breast tumours: An update. Breast Cancer Res. Treat 2004, 83, 249–289. [Google Scholar] [CrossRef] [PubMed]
  78. Zeng, L.; Yuan, S.; Shen, J.; Wu, M.; Pan, L.; Kong, X. Suppression of human breast cancer cells by tectorigenin through downregulation of matrix metalloproteinases and MAPK signaling in vitro. Mol. Med. Rep. 2018, 17, 3935–3943. [Google Scholar] [CrossRef] [Green Version]
  79. Reid, B.M.; Permuth, J.B.; Sellers, T.A. Epidemiology of ovarian cancer: A review. Cancer Biol. Med. 2017, 14, 9–32. [Google Scholar]
  80. Jayson, G.C.; Kohn, E.C.; Kitchener, H.C.; Ledermann, J.A. Ovarian cancer. Lancet 2014, 384, 1376–1388. [Google Scholar] [CrossRef]
  81. Pistollato, F.; Calderon Iglesias, R.; Ruiz, R.; Aparicio, S.; Crespo, J.; Dzul Lopez, L.; Giampieri, F.; Battino, M. The use of natural compounds for the targeting and chemoprevention of ovarian cancer. Cancer Lett. 2017, 411, 191–200. [Google Scholar] [CrossRef]
  82. Yang, Y.I.; Lee, K.T.; Park, H.J.; Kim, T.J.; Choi, Y.S.; Shih Ie, M.; Choi, J.H. Tectorigenin sensitizes paclitaxel-resistant human ovarian cancer cells through downregulation of the Akt and NFκB pathway. Carcinogenesis 2012, 33, 2488–2498. [Google Scholar] [CrossRef] [Green Version]
  83. Jung, S.H.; Lee, Y.S.; Lee, S.; Lim, S.S.; Kim, Y.S.; Ohuchi, K.; Shin, K.H. Anti-angiogenic and anti-tumor activities of isoflavonoids from the rhizomes of Belamcanda chinensis. Planta Med. 2003, 69, 617–622. [Google Scholar]
  84. Jiang, C.P.; Ding, H.; Shi, D.H.; Wang, Y.R.; Li, E.G.; Wu, J.H. Pro-apoptotic effects of tectorigenin on human hepatocellular carcinoma HepG2 cells. World. J. Gastroenterol. 2012, 18, 1753–1764. [Google Scholar] [CrossRef] [PubMed]
  85. Dai, X.Y.; Yu, Y.M.; Kong, Z.F.; Cao, Y.S.; Zhang, S. Tectorigenin inhibits Caco-2 human colon cells via NF-κB pathway suppression. Bangl. J. Pharmacol. 2015, 10, 948. [Google Scholar] [CrossRef]
  86. Lee, K.T.; Sohn, I.C.; Kim, Y.K.; Choi, J.H.; Choi, J.W.; Park, H.J.; Itoh, Y.; Miyamoto, K. Tectorigenin, an isoflavone of Pueraria thunbergiana Benth., induces differentiation and apoptosis in human promyelocytic leukemia HL-60 cells. Biol. Pharm. Bull. 2001, 24, 1117–1121. [Google Scholar] [CrossRef] [Green Version]
  87. Hasibeder, A.; Venkataramani, V.; Thelen, P.; Radzun, H.J.; Schweyer, S. Phytoestrogens regulate the proliferation and expression of stem cell factors in cell lines of malignant testicular germ cell tumors. Int. J. Oncol. 2013, 43, 1385–1394. [Google Scholar] [CrossRef] [Green Version]
  88. O’Toole, S.A.; Sheppard, B.L.; Sheils, O.; O’Leary, J.J.; Spengler, B.; Christoffel, V. Analysis of DNA in endometrial cancer cells treated with phyto-estrogenic compounds using comparative genomic hybridisation microarrays. Planta Med. 2005, 71, 435–439. [Google Scholar] [CrossRef]
  89. Eid, S.; Sas, K.M.; Abcouwer, S.F.; Feldman, E.L.; Gardner, T.W.; Pennathur, S.; Fort, P.E. New insights into the mechanisms of diabetic complications: Role of lipids and lipid metabolism. Diabetologia 2019, 62, 1539–1549. [Google Scholar] [CrossRef] [Green Version]
  90. Bae, E.A.; Han, M.J.; Lee, K.T.; Choi, J.W.; Park, H.J.; Kim, D.H. Metabolism of 6″-O-xylosyltectoridin and tectoridin by human intestinal bacteria and their hypoglycemic and in vitro cytotoxic activities. Biol. Pharm. Bull. 1999, 22, 1314–1318. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Jung, S.H.; Lee, Y.S.; Lee, S.; Lim, S.S.; Kim, Y.S.; Shin, K.H. Isoflavonoids from the rhizomes of Belamcanda chinensis and their effects on aldose reductase and sorbitol accumulation in streptozotocin induced diabetic rat tissues. Arch. Pharm. Res. 2002, 25, 306–312. [Google Scholar] [CrossRef]
  92. Yao, X.; Li, K.; Liang, C.; Zhou, Z.; Wang, J.; Wang, S.; Liu, L.; Yu, C.L.; Song, Z.B.; Bao, Y.L.; et al. Tectorigenin enhances PDX1 expression and protects pancreatic β-cells by activating ERK and reducing ER stress. J. Biol. Chem. 2020, 295, 12975–12992. [Google Scholar] [CrossRef]
  93. Qu, J.; Wu, Z.; Gao, J.; Wen, H.; Wang, T.; Yuan, D. Excretion of tectoridin metabolites in rat urine and bile orally administrated at different dosages and their inhibitory activity against aldose reductase. Fitoterapia 2014, 99, 99–108. [Google Scholar] [CrossRef]
  94. Munoz, M.; Lopez-Oliva, M.E.; Rodriguez, C.; Martínez, M.P.; Saenz-Medina, J.; Sanchez, A.; Climent, B.; Benedito, S.; Garcia-Sacristan, A.; Rivera, L.; et al. Differential contribution of Nox1, Nox2 and Nox4 to kidney vascular oxidative stress and endothelial dysfunction in obesity. Redox Biol. 2020, 28, 101330. [Google Scholar] [CrossRef] [PubMed]
  95. Ghosh, A.; Gao, L.; Thakur, A.; Siu, P.M.; Lai, C.W.K. Role of free fatty acids in endothelial dysfunction. J. Biomed. Sci. 2017, 24, 50. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Yang, S.; Ma, C.; Wu, H.; Zhang, H.; Yuan, F.; Yang, G.; Yang, Q.; Jia, L.; Liang, Z.; Kang, L. Tectorigenin attenuates diabetic nephropathy by improving vascular endothelium dysfunction through activating AdipoR1/2 pathway. Pharmacol. Res. 2020, 153, 104678. [Google Scholar] [CrossRef] [PubMed]
  97. Wang, Q.; Cheng, X.L.; Zhang, D.Y.; Gao, X.J.; Zhou, L.; Qin, X.Y.; Xie, G.Y.; Liu, K.; Qin, Y.; Liu, B.L.; et al. Tectorigenin attenuates palmitate-induced endothelial insulin resistance via targeting ROS-associated inflammation and IRS-1 pathway. PLoS ONE 2013, 8, e66417. [Google Scholar] [CrossRef] [Green Version]
  98. Zanquetta, M.M.; Alves-Wagner, A.B.; Mori, R.C.; Campello, R.S.; Machado, U.F. Recovery of insulin sensitivity and Slc2a4 mRNA expression depend on T3 hormone during refeeding. Metabolism 2014, 63, 328–334. [Google Scholar] [CrossRef]
  99. Yao, X.; Liu, L.; Shao, W.; Bai, M.; Ding, X.; Wang, G.; Wang, S.; Zheng, L.; Sun, Y.; Wang, G.; et al. Tectorigenin targets PKACα to promote GLUT4 expression in skeletal muscle and improve insulin resistance in vitro and in vivo. Int. J. Biol. Sci. 2023, 19, 1579–1596. [Google Scholar] [CrossRef]
  100. Pappachan, J.M.; Fernandez, C.J.; Chacko, E.C. Diabesity and antidiabetic drugs. Mol. Aspects Med. 2019, 66, 3–12. [Google Scholar] [CrossRef]
  101. Li, Q.Y.; Chen, L.; Yan, M.M.; Shi, X.J.; Zhong, M.K. Tectorigenin regulates adipogenic differentiation and adipocytokines secretion via PPARγ and IKK/NF-κB signaling. Pharm. Biol. 2015, 53, 1567–1575. [Google Scholar] [CrossRef]
  102. Zhou, M.; Deng, Y.; Liu, M.; Liao, L.; Dai, X.; Guo, C.; Zhao, X.; He, L.; Peng, C.; Li, Y. The pharmacological activity of berberine, a review for liver protection. Eur. J. Pharmacol. 2021, 890, 173655. [Google Scholar] [CrossRef]
  103. Elufioye, T.O.; Habtemariam, S. Hepatoprotective effects of rosmarinic acid: Insight into its mechanisms of action. Biomed. Pharmacother. 2019, 112, 108600. [Google Scholar] [CrossRef]
  104. Lee, H.W.; Choo, M.K.; Bae, E.A.; Kim, D.H. Beta-glucuronidase inhibitor tectorigenin isolated from the flower of Pueraria thunbergiana protects carbon tetrachloride-induced liver injury. Liver Int. 2003, 23, 221–226. [Google Scholar] [CrossRef] [PubMed]
  105. Gao, X.X.; Shi, D.H.; Chen, Y.X.; Cui, J.T.; Wang, Y.R.; Jiang, C.P.; Wu, J.H. The therapeutic effects of tectorigenin on chemically induced liver fibrosis in rats and an associated metabonomic investigation. Arch. Pharm. Res. 2012, 35, 1479–1493. [Google Scholar] [CrossRef] [PubMed]
  106. Lee, H.U.; Bae, E.A.; Kim, D.H. Hepatoprotective effect of tectoridin and tectorigenin on tert-butyl hyperoxide-induced liver injury. J. Pharmacol. Sci. 2005, 97, 541–544. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Zhang, L.; Zhao, Y.; Fan, L.; Xu, K.; Ji, F.; Xie, Z.; Ouyang, X.; Wu, D.; Li, L. Tectorigenin protects against experimental fulminant hepatic failure by regulating the TLR4/mitogen-activated protein kinase and TLR4/nuclear factor-κB pathways and autophagy. Phytother. Res. 2019, 33, 1055–1064. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Asrani, S.K.; Devarbhavi, H.; Eaton, J.; Kamath, P.S. Burden of liver diseases in the world. J. Hepatol. 2019, 70, 151–171. [Google Scholar] [CrossRef]
  109. Wu, J.H.; Wang, Y.R.; Huang, W.Y.; Tan, R.X. Anti-proliferative and pro-apoptotic effects of tectorigenin on hepatic stellate cells. World J. Gastroenterol. 2010, 16, 3911–3918. [Google Scholar] [CrossRef]
  110. Duan, R.; Huang, K.; Guan, X.; Li, S.; Xia, J.; Shen, M.; Sun, Z.; Yu, Z. Tectorigenin ameliorated high-fat diet-induced nonalcoholic fatty liver disease through anti-inflammation and modulating gut microbiota in mice. Food Chem. Toxicol. 2022, 164, 112948. [Google Scholar] [CrossRef]
  111. Kim, H.K.; Cheon, B.S.; Kim, Y.H.; Kim, S.Y.; Kim, H.P. Effects of naturally occurring flavonoids on nitric oxide production in the macrophage cell line RAW 264.7 and their structure-activity relationships. Biochem. Pharmacol. 1999, 58, 759–765. [Google Scholar] [CrossRef]
  112. Kim, Y.P.; Yamada, M.; Lim, S.S.; Lee, S.H.; Ryu, N.; Shin, K.H.; Ohuchi, K. Inhibition by tectorigenin and tectoridin of prostaglandin E2 production and cyclooxygenase-2 induction in rat peritoneal macrophages. Biochim. Biophys. Acta 1999, 1438, 399–407. [Google Scholar] [CrossRef]
  113. Hong, J.; Shin, K.H.; Lim, S.S.; Kwak, J.H.; Zee, O.; Ishihara, K.; Hirasawa, N.; Seyama, T.; Ohuchi, K. Lead compounds for anti-inflammatory drugs isolated from the plants of the traditional oriental medicine in Korea. Inflamm. Allergy Drug Targets 2008, 7, 195–202. [Google Scholar] [CrossRef]
  114. Pan, C.H.; Kim, E.S.; Jung, S.H.; Nho, C.W.; Lee, J.K. Tectorigenin inhibits IFN-gamma/LPS-induced inflammatory responses in murine macrophage RAW 264.7 cells. Arch. Pharm. Res. 2008, 31, 1447–1456. [Google Scholar] [CrossRef] [PubMed]
  115. Lim, H.S.; Kim, Y.J.; Kim, B.Y.; Park, G.; Jeong, S.J. The anti-neuroinflammatory activity of tectorigenin pretreatment via downregulated NF-κB and ERK/JNK pathways in BV-2 microglial and microglia inactivation in mice with lipopolysaccharide. Front. Pharmacol. 2018, 9, 462. [Google Scholar] [CrossRef] [PubMed]
  116. Zhou, L.; Yan, K.; Xing, S.; Cheng, J. Tectorigenin alleviates the apoptosis and inflammation in spinal cord injury cell model through inhibiting insulin-like growth factor-binding protein 6. Open Med. 2023, 18, 20230680. [Google Scholar] [CrossRef]
  117. Ma, C.H.; Liu, J.P.; Qu, R.; Ma, S.P. Tectorigenin inhibits the inflammation of LPS-induced acute lung injury in mice. Chin. J. Nat. Med. 2014, 12, 841–846. [Google Scholar] [CrossRef]
  118. Ha, L.M.; Que, D.T.N.; Huyen, D.T.T.; Long, P.Q.; Dat, N.T. Toxicity, analgesic and anti-inflammatory activities of tectorigenin. Immunopharmacol. Immunotoxicol. 2013, 35, 336–340. [Google Scholar] [CrossRef]
  119. Li, J.; Yan, W.; Ren, F.; Sang, H. Tectorigenin inhibits inflammation in keratinocytes by inhibition of NLRP3 inflammasome regulated by the TLR4/NF-κB pathway. Allergol. Immunopathol. 2023, 51, 82–89. [Google Scholar] [CrossRef]
  120. Lee, S.Y.; Kim, G.T.; Yun, H.M.; Kim, Y.C.; Kwon, I.K.; Kim, E.C. Tectorigenin promotes osteoblast differentiation and in vivo bone healing, but suppresses osteoclast differentiation and in vivo bone resorption. Mol. Cells 2018, 41, 476–485. [Google Scholar]
  121. Forman, H.J.; Zhang, H. Targeting oxidative stress in disease: Promise and limitations of antioxidant therapy. Nat. Rev. Drug Discov. 2021, 20, 689–709. [Google Scholar] [CrossRef]
  122. Tayab, M.A.; Chowdhury, K.A.A.; Jabed, M.; Mohammed Tareq, S.; Kamal, A.; Islam, M.N.; Uddin, A.M.K.; Hossain, M.A.; Emran, T.B.; Simal-Gandara, J. Antioxidant-rich woodfordia fruticosa leaf extract alleviates depressive-like behaviors and impede hyperglycemia. Plants 2021, 10, 287. [Google Scholar] [CrossRef] [PubMed]
  123. Akhter, S.; Arman, M.S.I.; Tayab, M.A.; Islam, M.N.; Xiao, J. Recent advances in the biosynthesis, bioavailability, toxicology, pharmacology, and controlled release of citrus neohesperidin. Crit. Rev. Food Sci. Nutr. 2022; Online ahead of print. [Google Scholar] [CrossRef]
  124. Chen, W.; Zhuang, J.; Li, Y.; Shen, Y.; Zheng, X. Myricitrin protects against peroxynitrite-mediated DNA damage and cytotoxicity in astrocytes. Food Chem. 2013, 141, 927–933. [Google Scholar] [CrossRef]
  125. Valko, M.; Leibfritz, D.; Moncol, J.; Cronin, M.T.; Mazur, M.; Telser, J. Free radicals and antioxidants in normal physiological functions and human disease. Int. J. Biochem. Cell Biol. 2007, 39, 44–84. [Google Scholar] [CrossRef] [PubMed]
  126. Arct, J.; Pytkowska, K. Flavonoids as components of biologically active cosmeceuticals. Clin. Dermatol. 2008, 26, 347–357. [Google Scholar] [CrossRef]
  127. Kang, K.A.; Lee, K.H.; Chae, S.; Zhang, R.; Jung, M.S.; Kim, S.Y.; Kim, H.S.; Kim, D.H.; Hyun, J.W. Cytoprotective effect of tectorigenin, a metabolite formed by transformation of tectoridin by intestinal microflora, on oxidative stress induced by hydrogen peroxide. Eur. J. Pharmacol. 2005, 519, 16–23. [Google Scholar] [CrossRef] [PubMed]
  128. Park, K.Y.; Jung, G.O.; Choi, J.; Lee, K.T.; Park, H.J. Potent antimutagenic and their anti-lipid peroxidative effect of kaikasaponin III and tectorigenin from the flower of Pueraria thunbergiana. Arch. Pharm. Res. 2002, 25, 320–324. [Google Scholar] [CrossRef] [PubMed]
  129. Vaou, N.; Stavropoulou, E.; Voidarou, C.; Tsigalou, C.; Bezirtzoglou, E. Towards advances in medicinal plant antimicrobial activity: A review study on challenges and future perspectives. Microorganisms 2021, 9, 2041. [Google Scholar] [CrossRef]
  130. Bae, E.A.; Han, M.J.; Kim, D.H. In vitro anti-Helicobacter pylori activity of irisolidone isolated from the flowers and rhizomes of Pueraria thunbergiana. Planta Med. 2001, 67, 161–163. [Google Scholar] [CrossRef] [PubMed]
  131. Liu, S.; Wang, J.; Zhang, J.; Wang, T.; Zhou, Y.; Lv, Q.; Hu, N.; Shen, X.; Deng, X. Tectorigenin reduces type IV pilus-dependent cell adherence in Clostridium perfringens. FEMS Microbiol. Lett. 2019, 366, 112. [Google Scholar] [CrossRef]
  132. Guan, Z.; Luo, L.; Liu, S.; Guan, Z.; Zhang, Q.; Li, X.; Tao, K. The role of depletion of gut microbiota in osteoporosis and osteoarthritis: A narrative review. Front. Endocrinol. 2022, 13, 847401. [Google Scholar] [CrossRef]
  133. Shan, S.K.; Lin, X.; Li, F.; Xu, F.; Zhong, J.Y.; Guo, B.; Wang, Y.; Zheng, M.H.; Wu, F.; Yuan, L.Q. Exosomes and bone disease. Curr. Pharm. Des. 2019, 25, 4536–4549. [Google Scholar] [CrossRef]
  134. Wang, S.; Deng, Z.; Ma, Y.; Jin, J.; Qi, F.; Li, S.; Liu, C.; Lyu, F.J.; Zheng, Q. The role of autophagy and mitophagy in bone metabolic disorders. Int. J. Biol. Sci. 2020, 16, 2675–2691. [Google Scholar] [CrossRef] [PubMed]
  135. Zhang, L.; Zheng, Y.L.; Wang, R.; Wang, X.Q.; Zhang, H. Exercise for osteoporosis: A literature review of pathology and mechanism. Front. Immunol. 2022, 13, 1005665. [Google Scholar] [CrossRef] [PubMed]
  136. Riggs, B.L.; Hartmann, L.C. Selective estrogen-receptor modulators—Mechanisms of action and application to clinical practice. N. Engl. J. Med. 2003, 348, 618–629. [Google Scholar] [CrossRef] [PubMed]
  137. Jeong, H.J.; Kim, M.H.; Kim, H.; Kim, H.Y.; Nam, S.Y.; Han, N.R.; Lee, B.; Cho, H.; Moon, P.D.; Kim, H.M. PCE17 and its active compounds exert an anti-osteoporotic effect through the regulation of receptor activator of nuclear factor-κB ligand in ovariectomized mice. J. Food Biochem. 2018, 42, e12561. [Google Scholar] [CrossRef]
  138. Ma, C.; Xu, K.; Meng, J.; Ran, J.; Adel Abdo Moqbel, S.; Liu, A.; Yan, S.; Wu, L. Tectorigenin inhibits RANKL-induced osteoclastogenesis via suppression of NF-κB signalling and decreases bone loss in ovariectomized C57BL/6. J. Cell. Mol. Med. 2018, 22, 5121–5131. [Google Scholar] [CrossRef]
  139. Wang, C.L.; Li, D.; Wang, C.D.; Xiao, F.; Zhu, J.F.; Shen, C.; Zuo, B.; Cui, Y.M.; Wang, H.; Gao, Y.; et al. Anti-inflammatory and anti-osteoarthritis effects of tectorigenin. Biol. Open 2017, 6, 1130–1136. [Google Scholar] [CrossRef] [Green Version]
  140. Aicale, R.; Oliviero, A.; Maffulli, N. Management of achilles and patellar tendinopathy: What we know, what we can do. J. Foot Ankle. Res. 2020, 13, 59. [Google Scholar] [CrossRef]
  141. Moqbel, S.A.A.; Xu, K.; Chen, Z.; Xu, L.; He, Y.; Wu, Z.; Ma, C.; Ran, J.; Wu, L.; Xiong, Y. Tectorigenin alleviates inflammation, apoptosis, and ossification in rat tendon-derived stem cells via modulating NF-κB and MAPK pathways. Front. Cell Dev. Biol. 2020, 8, 568894. [Google Scholar] [CrossRef]
  142. Kammeyer, A.; Luiten, R.M. Oxidation events and skin aging. Ageing Res. Rev. 2015, 21, 16–29. [Google Scholar] [CrossRef]
  143. Dai, X.; Jin, J.; Jia, Y.; Yang, K.; Han, J.; Zhang, Z.; Ding, X.; Yao, C.; Sun, T.; Zhu, C.; et al. A non-retinol retinoic acid receptor-γ (RAR-γ/NR1B3) selective agonist, tectorigenin, can effectively inhibit the ultraviolet a-induced skin damage. Br. J. Pharmacol. 2022, 179, 4722–4737. [Google Scholar] [CrossRef]
  144. Khalil, S.; Bardawil, T.; Stephan, C.; Darwiche, N.; Abbas, O.; Kibbi, A.G.; Nemer, G.; Kurban, M. Retinoids: A journey from the molecular structures and mechanisms of action to clinical uses in dermatology and adverse effects. J. Dermatolog. Treat 2017, 28, 684–696. [Google Scholar] [CrossRef] [PubMed]
  145. Tamura, S.; Yoshihira, K.; Tokumaru, M.; Zisheng, X.; Murakami, N. Inhibitors for expression of IgE receptor on human mast cell from Puerariae Flos. Bioorg. Med. Chem. Lett. 2010, 20, 3872–3875. [Google Scholar] [CrossRef] [PubMed]
  146. Park, E.K.; Shin, Y.W.; Lee, H.U.; Lee, C.S.; Kim, D.H. Passive cutaneous anaphylaxis-inhibitory action of tectorigenin, a metabolite of tectoridin by intestinal microflora. Biol. Pharm. Bull. 2004, 27, 1099–1102. [Google Scholar] [CrossRef] [Green Version]
  147. Pan, Y.; Wu, W.; Jiang, X.; Liu, Y. Mesenchymal stem cell-derived exosomes in cardiovascular and cerebrovascular diseases: From mechanisms to therapy. Biomed. Pharmacother. 2023, 163, 114817. [Google Scholar] [CrossRef] [PubMed]
  148. Liberale, L.; Ministrini, S.; Carbone, F.; Camici, G.G.; Montecucco, F. Cytokines as therapeutic targets for cardio- and cerebrovascular diseases. Basic. Res. Cardiol. 2021, 116, 23. [Google Scholar] [CrossRef]
  149. Chaturvedi, S.; Yadav, J.S. The role of antiplatelet therapy in carotid stenting for ischemic stroke prevention. Stroke 2006, 37, 1572–1577. [Google Scholar] [CrossRef]
  150. Applova, L.; Karlíckova, J.; Ríha, M.; Filipsky, T.; Macakova, K.; Spilkova, J.; Mladenka, P. The isoflavonoid tectorigenin has better antiplatelet potential than acetylsalicylic acid. Phytomedicine 2017, 35, 11–17. [Google Scholar] [CrossRef]
  151. Tang, Y.; Li, S.; Li, S.; Yang, X.; Qin, Y.; Zhang, Y.; Liu, C. Screening and isolation of potential lactate dehydrogenase inhibitors from five chinese medicinal herbs: Soybean, Radix pueraria, Flos pueraria, Rhizoma belamcandae, and Radix astragali. J. Sep. Sci. 2016, 39, 2043–2049. [Google Scholar] [CrossRef]
  152. Yao, L.; Yang, M.; Zhang, J.; Wang, F.; Liu, Q.; Xie, X.; Liu, Z.; Guo, Q.; Su, H.; Zhai, J.; et al. Tectorigenin attenuates the OGD/R-induced HT-22 cell damage through regulation of the PI3K/AKT and the PPARγ/NF-κB pathways. Hum. Exp. Toxicol. 2021, 40, 1320–1331. [Google Scholar] [CrossRef]
  153. Feng, W. Tectorigenin attenuates cognitive impairments in mice with chronic cerebral ischemia by inhibiting the TLR4/NF-κB signaling pathway. Biosci. Biotechnol. Biochem. 2021, 85, 1665–1674. [Google Scholar] [CrossRef]
  154. Qian, X.; Xiao, Q.; Li, Z. Tectorigenin regulates migration, invasion, and apoptosis in dexamethasone-induced human airway epithelial cells through up-regulating miR-222-3p. Drug Dev. Res. 2021, 82, 959–968. [Google Scholar] [CrossRef]
  155. Zhang, H.; Liu, X.; Chen, S.; Wu, J.; Ye, X.; Xu, L.; Chen, H.; Zhang, D.; Tan, R.; Wang, Y. Tectorigenin inhibits the in vitro proliferation and enhances miR-338* expression of pulmonary fibroblasts in rats with idiopathic pulmonary fibrosis. J. Ethnopharmacol. 2010, 131, 165–173. [Google Scholar] [CrossRef] [PubMed]
  156. Papi, A.; Brightling, C.; Pedersen, S.E.; Reddel, H.K. Asthma. Lancet 2018, 391, 783–800. [Google Scholar] [CrossRef] [PubMed]
  157. Richeldi, L.; Collard, H.R.; Jones, M.G. Idiopathic pulmonary fibrosis. Lancet 2017, 389, 1941–1952. [Google Scholar] [CrossRef] [PubMed]
  158. Cummings, J. Disease modification and Neuroprotection in neurodegenerative disorders. Transl. Neurodegener. 2017, 6, 25. [Google Scholar] [CrossRef] [Green Version]
  159. Crews, F.T.; Bechara, R.; Brown, L.A.; Guidot, D.M.; Mandrekar, P.; Oak, S.; Qin, L.; Szabo, G.; Wheeler, M.; Zou, J. Cytokines and alcohol. Alcohol. Clin. Exp. Res. 2006, 30, 720–730. [Google Scholar] [CrossRef]
  160. Yeh, L.T.; Hsu, L.S.; Chung, Y.H.; Chen, C.J. Tectorigenin inhibits glioblastoma proliferation by G0/G1 cell cycle arrest. Medicina 2020, 56, 681. [Google Scholar] [CrossRef]
  161. Marti, H.H.; Wenger, R.H.; Rivas, L.A.; Straumann, U.; Digicaylioglu, M.; Henn, V.; Yonekawa, Y.; Bauer, C.; Gassmann, M. Erythropoietin gene expression in human, monkey and murine brain. Eur. J. Neurosci. 1996, 8, 666–676. [Google Scholar] [CrossRef]
  162. Sun, Y.Y.; Lin, S.H.; Lin, H.C.; Hung, C.C.; Wang, C.Y.; Lin, Y.C.; Hung, K.S.; Lien, C.C.; Kuan, C.Y.; Lee, Y.H. Cell type-specific dependency on the PI3K/Akt signaling pathway for the endogenous Epo and VEGF induction by baicalein in neurons versus astrocytes. PLoS ONE 2013, 8, e69019. [Google Scholar] [CrossRef] [Green Version]
  163. John, M.J.; Jaison, V.; Jain, K.; Kakkar, N.; Jacob, J.J. Erythropoietin use and abuse. Indian J. Endocrinol. Metab. 2012, 16, 220–227. [Google Scholar]
  164. Liu, E.Y.; Zheng, Z.X.; Zheng, B.Z.; Xia, Y.; Guo, M.S.; Dong, T.T.; Tsim, K.W.K. Tectorigenin, an isoflavone aglycone from the rhizome of Belamcanda chinensis, induces neuronal expression of erythropoietin via accumulation of hypoxia-inducible factor-1α. Phytother. Res. 2020, 34, 1329–1337. [Google Scholar] [CrossRef]
  165. Gong, P.; Deng, F.; Zhang, W.; Ji, J.; Liu, J.; Sun, Y.; Hu, J. Tectorigenin attenuates the MPP(+)-induced SH-SY5Y cell damage, indicating a potential beneficial role in Parkinson’s disease by oxidative stress inhibition. Exp. Ther. Med. 2017, 14, 4431–4437. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Li, J.; Yang, J.; Zhu, B.; Fan, J.; Hu, Q.; Wang, L. Tectorigenin protects against unilateral ureteral obstruction by inhibiting Smad3-mediated ferroptosis and fibrosis. Phytother. Res. 2022, 36, 475–487. [Google Scholar] [CrossRef]
  167. Jung, D.Y.; Ha, H.; Kim, C. Induction of growth hormone release by Pueraria thunbergiana BENTH. Horm. Metab. Res. 2004, 36, 86–91. [Google Scholar] [PubMed]
  168. Tomonaga, T.; Mine, T.; Kojima, I.; Taira, M.; Hayashi, H.; Isono, K. Isoflavonoids, genistein, psi-tectorigenin, and orobol, increase cytoplasmic free calcium in isolated rat hepatocytes. Biochem. Biophys. Res. Commun. 1992, 182, 894–899. [Google Scholar] [CrossRef]
  169. Cheng, H.H.; Liang, W.Z.; Liao, W.C.; Kuo, C.C.; Hao, L.J.; Chou, C.T.; Jan, C.R. Investigation of effect of tectorigenin (O-methylated isoflavone) on Ca2+ signal transduction and cytotoxic responses in canine renal tubular cells. Chin. J. Physiol. 2020, 63, 60–67. [Google Scholar]
  170. Madden, J.C.; Thompson, C.V. Pharmacokinetic tools and applications. Methods Mol. Biol. 2022, 2425, 57–83. [Google Scholar] [PubMed]
  171. Wang, S.; Gong, T.; Lu, J.; Kano, Y.; Yuan, D. Simultaneous determination of tectorigenin and its metabolites in rat plasma by ultra performance liquid chromatography/quadrupole time-of-flight mass spectrometry. J. Chromatogr. B 2013, 933, 50–58. [Google Scholar] [CrossRef] [PubMed]
  172. Li, J.; Xu, Z.; Gu, J. UGT1A1 and UGT1A9 are responsible for phase ii metabolism of tectorigenin and irigenin in vitro. Molecules 2022, 27, 4104. [Google Scholar] [CrossRef] [PubMed]
  173. Zhang, W.D.; Qi, L.W.; Yang, X.L.; Huang, W.Z.; Li, P.; Yang, Z.L. Identification of the major metabolites of tectorigenin in rat bile by liquid chromatography combined with time-of-flight and ion trap tandem mass spectrometry. Rapid. Commun. Mass. Spectrom. 2008, 22, 2677–2684. [Google Scholar] [CrossRef]
  174. Zhang, W.D.; Yang, X.L.; Cao, J.; Li, P.; Yang, Z.L. Identification of key metabolites of tectorigenin in rat urine by HPLC-MS(n). Biomed. Chromatogr. 2009, 23, 219–221. [Google Scholar] [CrossRef] [PubMed]
  175. Shi, Z.; Zhang, G.; Zhao, L.; Wang, S.; Kano, Y.; Yuan, D. Excretion of tectorigenin in rat urine orally administrated at different dosages by ultra-performance liquid chromatography coupled to quadrupole time-of-flight mass spectrometry. Eur. J. Drug. Metab. Pharmacokinet 2015, 40, 255–266. [Google Scholar] [CrossRef] [PubMed]
  176. Zhang, W.D.; Qi, L.W.; Yang, X.L.; Lu, Y.W.; Li, P.; Yang, Z.L. Determination of tectorigenin in rat plasma: Application to a pharmacokinetic study after oral administration of tectorigenin or its prodrug tectoridin. Chromatographia 2008, 68, 1021. [Google Scholar] [CrossRef]
  177. Li, J.; Yao, Y.; Zhou, M.; Yu, Z.; Wang, X. Pharmacokinetics of tectorigenin, tectoridi, irigenin, and iridin in mouse blood after intravenous administration by UPLC-MS/MS. Acta Chromatogr. 2021, 34, 246–252. [Google Scholar] [CrossRef]
  178. Li, L.; Zhang, X.; Bu, F.; Chen, N.; Zhang, H.; Gu, J. Simultaneous determination of eight constituents in rat plasma by HPLC-MS/MS and its application to a pharmacokinetic study after oral administration of Shejin-liyan Granule. Biomed. Chromatogr. 2019, 33, e4648. [Google Scholar] [CrossRef]
  179. Zhang, W.D.; Yang, W.J.; Wang, X.J.; Gu, Y.; Wang, R. Simultaneous determination of tectorigenin, irigenin and irisflorentin in rat plasma and urine by UHPLC-MS/MS: Application to pharmacokinetics. J. Chromatogr. B 2011, 879, 3735–3741. [Google Scholar] [CrossRef]
  180. Yang, M.; Yang, X.; An, J.; Xiao, W.; Wang, Z.; Huang, W.; Yang, Z.; Li, F. Comparative pharmacokinetic profiles of tectorigenin in rat plasma by UPLC-MS/MS after oral administration of Iris tectorum Maxim extract and pure tectoridin. J. Pharm. Biomed. Anal. 2015, 114, 34–41. [Google Scholar] [CrossRef]
  181. Bai, X.; Xie, Y.; Liu, J.; Qu, J.; Kano, Y.; Yuan, D. Isolation and identification of urinary metabolites of kakkalide in rats. Drug Metab. Dispos. 2010, 38, 281–286. [Google Scholar] [CrossRef] [Green Version]
  182. Zhang, G.; Qi, W.; Xu, L.; Kano, Y.; Yuan, D. Pharmacokinetics of irisolidone and its main metabolites in rat plasma determined by ultra performance liquid chromatography/quadrupole time-of-flight mass spectrometry. J. Chromatogr. B 2015, 1005, 23–29. [Google Scholar] [CrossRef]
  183. Qu, J.; Gao, J.; Sun, J.; Zhang, L.; Makino, T.; Yuan, D. Pharmacokinetics of conjugated metabolites in rat plasma after oral administration of tectoridin. J. Chromatogr. B 2012, 902, 61–69. [Google Scholar] [CrossRef]
  184. Chen, Y.; Song, W.; Peng, Z.H.; Ge, B.Y.; Han, F.M. Identification of metabolites of tectoridin in-vivo and in-vitro by liquid chromatography-tandem mass spectrometry. J. Pharm. Pharmacol. 2008, 60, 709–716. [Google Scholar] [CrossRef]
  185. Thelen, P.; Seseke, F.; Ringert, R.H.; Wuttke, W.; Seidlova-Wuttke, D. Pharmacological potential of phytoestrogens in the treatment of prostate cancer. Urologe. A 2006, 45, 195–201. [Google Scholar] [CrossRef]
  186. Shin, J.E.; Bae, E.A.; Lee, Y.C.; Ma, J.Y.; Kim, D.H. Estrogenic effect of main components kakkalide and tectoridin of Puerariae Flos and their metabolites. Biol. Pharm. Bull. 2006, 29, 1202–1206. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Zheng, Y.; Zhang, R.; Shi, W.; Li, L.; Liu, H.; Chen, Z.; Wu, L. Metabolism and pharmacological activities of the natural health-benefiting compound diosmin. Food Funct. 2020, 11, 8472–8492. [Google Scholar] [CrossRef] [PubMed]
  188. Khan, H.; Ullah, H.; Martorell, M.; Valdes, S.E.; Belwal, T.; Tejada, S.; Sureda, A.; Kamal, M.A. Flavonoids nanoparticles in cancer: Treatment, prevention and clinical prospects. Semin. Cancer Biol. 2021, 69, 200–211. [Google Scholar] [CrossRef] [PubMed]
  189. Bangar, S.P.; Chaudhary, V.; Sharma, N.; Bansal, V.; Ozogul, F.; Lorenzo, J.M. Kaempferol: A flavonoid with wider biological activities and its applications. Crit. Rev. Food Sci. Nutr. 2022; Online ahead of print. [Google Scholar] [CrossRef] [PubMed]
  190. Ravetti, S.; Garro, A.G.; Gaitan, A.; Murature, M.; Galiano, M.; Brignone, S.G.; Palma, S.D. Naringin: Nanotechnological strategies for potential pharmaceutical applications. Pharmaceutics 2023, 15, 863. [Google Scholar] [CrossRef]
  191. Biasutto, L.; Zoratti, M. Prodrugs of quercetin and resveratrol: A strategy under development. Curr. Drug Metab. 2014, 15, 77–95. [Google Scholar] [CrossRef]
  192. Zhang, R.; Shi, H.; Li, S.; Zhang, H.; Zhang, D.; Wu, A.; Zhang, C.; Li, C.; Fu, X.; Chen, S.; et al. A double-layered gastric floating tablet for zero-order controlled release of dihydromyricetin: Design, development, and in vitro/in vivo evaluation. Int. J. Pharm. 2023, 638, 122929. [Google Scholar] [CrossRef]
  193. Taghavinia, F.; Teymouri, F.; Farokhrouz, F.; Bagherabad, E.H.; Farjami, S.; Karimi, E.; Oskoueian, E.; Le, H.H.; Shakeri, M. Nanoliposome-loaded phenolics from Nasturtium officinale improves health parameters in a colorectal cancer mouse model. Animals 2022, 12, 3492. [Google Scholar] [CrossRef]
  194. Beyrami, M.; Karimi, E.; Oskoueian, E. Synthesized chrysin-loaded nanoliposomes improves cadmium-induced toxicity in mice. Environ. Sci. Pollut. Res. Int. 2020, 27, 40643–40651. [Google Scholar] [CrossRef] [PubMed]
  195. Zarei, M.; Karimi, E.; Oskoueian, E.; Es-Haghi, A.; Yazdi, M.E.T. Comparative study on the biological effects of sodium citrate-based and apigenin-based synthesized silver nanoparticles. Nutr. Cancer 2021, 73, 1511–1519. [Google Scholar] [CrossRef]
  196. Tabatabaeain, S.F.; Karimi, E.; Hashemi, M. Satureja khuzistanica essential oil-loaded solid lipid nanoparticles modified with chitosan-folate: Evaluation of encapsulation efficiency, cytotoxic and pro-apoptotic properties. Front. Chem. 2022, 10, 904973. [Google Scholar] [CrossRef] [PubMed]
  197. Moeini, S.; Karimi, E.; Oskoueian, E. Antiproliferation effects of nanophytosome-loaded phenolic compounds from fruit of Juniperus polycarpos against breast cancer in mice model: Synthesis, characterization and therapeutic effects. Cancer Nanotechnol. 2022, 13, 20. [Google Scholar] [CrossRef]
  198. Chauhan, A.S. Dendrimers for Drug Delivery. Molecules 2018, 23, 938. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Javia, A.; Vanza, J.; Bardoliwala, D.; Ghosh, S.; Misra, L.A.; Patel, M.; Thakkar, H. Polymer-drug conjugates: Design principles, emerging synthetic strategies and clinical overview. Int. J. Pharm. 2022, 623, 121863. [Google Scholar]
  200. Vlachopoulos, A.; Karlioti, G.; Balla, E.; Daniilidis, V.; Kalamas, T.; Stefanidou, M.; Bikiaris, N.D.; Christodoulou, E.; Koumentakou, I.; Karavas, E.; et al. Poly(lactic acid)-based microparticles for drug delivery applications: An overview of recent advances. Pharmaceutics 2022, 14, 359. [Google Scholar] [CrossRef]
  201. Bhatnagar, P.; Dhote, V.; Mahajan, S.C.; Mishra, P.K.; Mishra, D.K. Solid dispersion in pharmaceutical drug development: From basics to clinical applications. Curr. Drug Deliv. 2014, 11, 155–171. [Google Scholar] [CrossRef]
  202. Shuai, S.; Yue, S.; Huang, Q.; Wang, W.; Yang, J.; Lan, K.; Ye, L. Preparation, characterization and in vitro/vivo evaluation of tectorigenin solid dispersion with improved dissolution and bioavailability. Eur. J. Drug Metab. Pharmacokinet. 2016, 41, 413–422. [Google Scholar] [CrossRef]
  203. Zhang, Y.; He, L.; Yue, S.; Huang, Q.; Zhang, Y.; Yang, J. Characterization and evaluation of a self-microemulsifying drug delivery system containing tectorigenin, an isoflavone with low aqueous solubility and poor permeability. Drug Deliv. 2017, 24, 632–640. [Google Scholar] [CrossRef] [Green Version]
  204. Kim, Y.H.; Tabata, Y. Dual-controlled release system of drugs for bone regeneration. Adv. Drug Deliv. Rev. 2015, 94, 28–40. [Google Scholar] [CrossRef] [PubMed]
  205. Wang, J.F.; Wang, F.; Zhao, N.; Yang, C.Y.; Wang, G.Y.; Wei, Y.; Zhang, W.Y. Preparation and release mechanism of tectorigenin intragastric floating sustained-release tablets. Chin. J. Chin. Mater. Med. 2017, 42, 5. [Google Scholar]
Figure 1. Chemical structure of tectorigenin.
Figure 1. Chemical structure of tectorigenin.
Molecules 28 05904 g001
Figure 2. Synthesis of tectorigenin. (i) The synthesis of tectorigenin by hydrolyzing tectoridin; (ii) The de novo synthesis of tectorigenin from3-methoxy-methyl gallate.
Figure 2. Synthesis of tectorigenin. (i) The synthesis of tectorigenin by hydrolyzing tectoridin; (ii) The de novo synthesis of tectorigenin from3-methoxy-methyl gallate.
Molecules 28 05904 g002
Figure 3. Pharmacological activity and relative mechanism of tectorigenin. (“↑” means up-regulation; “↓” means down-regulation).
Figure 3. Pharmacological activity and relative mechanism of tectorigenin. (“↑” means up-regulation; “↓” means down-regulation).
Molecules 28 05904 g003
Figure 4. The proposed metabolic pathways of tectorigenin in rats.
Figure 4. The proposed metabolic pathways of tectorigenin in rats.
Molecules 28 05904 g004
Table 1. Plants containing tectorigenin.
Table 1. Plants containing tectorigenin.
No.PlantFamilyPartRef.
1Belamcanda chinensis (L.) DC.IridaceaeRhizomes[25,26]
2Iris spuria L. (Calizona)IridaceaeRhizomes[27]
3Iris tectorum MaximIridaceaeRoots and rhizomes[28]
4Iris japonica Thunb.IridaceaeWhole plant[29,30]
5Iris dichotoma Pall.IridaceaeRhizomes[31]
6Iris germanica L.IridaceaeRhizomes[32]
7Iris unguicularis PoiretIridaceaeRhizomes[33]
8Iris loczyi Kan.IridaceaeWhole plant[33]
9Iris kashmiriana BakerIridaceaeRhizomes[34]
10Iris crocea Jacq. ex R. C. FosterIridaceaeRoots and rhizomes[35,36]
11Iris ensata Thunb.IridaceaeRhizomes[36]
12Iris germanica L.IridaceaeUnderground parts[37]
13Iris hungarica Waldst. et Kit.IridaceaeRhizomes[38]
14Iris confusa SealyIridaceaeUnderground parts[37]
15Iris pseudacorus L.IridaceaeUnderground parts[37]
16Pueraria lobata (Willd.) OhwiLeguminosaeFlowers[40,41]
17Pueraria thomsonii Benth.LeguminosaeFlowers[18,40]
18Pueraria thunbergiana Benth.LeguminosaeFlowers[42]
19Dalbergia odorifera T. ChenLeguminosaeLeaves[43]
20Dalbergia parviflora Roxb.LeguminosaeHeartwood[44]
21Euchresta formosana (Hayata) OhwiLeguminosaeRoots[45]
22Codonopsis pilosula (Franch.) Nannf.CampanulaceaeRoots[46]
23Morus alba L.MoraceaeVelamen and leaves[46]
24Viola hondoensis W. Becker et H. Boissieu.ViolaceaeAerial parts[47]
25Eleocharis dulcis (Burm. f.) Trin. ex Hensch.CyperaceaePeel[48]
Table 3. Pharmacokinetic parameters of tectorigenin in animals.
Table 3. Pharmacokinetic parameters of tectorigenin in animals.
Inclusion of Drug
Components
Dosage and RouteAnimal ModelPharmacokinetic ParametersRef.
Tectorigenin130 mg/kg, oral administrationMale Sprague-Dawley ratsCmax = 12.0 ± 0.63 μmol/L, Tmax = 0.23 ± 0.15 h, t1/2 = 11.7 ± 5.74 h, AUC(0–t) = 84.2 ± 8.15 μmol/L × h[171]
Tectorigenin80 mg/kg, oral administrationSprague-Dawley ratsCmax = 1.46 ± 0.30 μg/mL, Tmax = 1.20 ± 0.30 h, AUC(0–t) = 7.31 ± 1.20 μg/mL × h[176]
Tectoridin130 mg/kg, oral administrationSprague-Dawley ratsCmax = 1.08 ± 0.25 μg/mL, Tmax = 7.11 ± 1.10 h, AUC(0–t) = 8.02 ± 1.10 μg/mL × h[176]
Tectorigenin5 mg/kg, sublingual intravenous administrationMale miceCmax = 349.0 ± 172.4 ng/mL, t1/2 = 2.4 ± 1.5 h, AUC(0–t) = 219 ± 94 ng/mL × h[177]
Shejin-liyan Granule2.0 g/kg, oral administrationSprague-Dawley ratsCmax 1 = 112.72 ± 60.04 ng/mL, Cmax 2 = 463.67 ± 170.46 ng/mL, Tmax 1 = 0.75 ± 0.35 h, Tmax 2 = 6.00 ± 2.19 h, t1/2 = 12.22 ± 2.42 h, AUC(0–t) = 5340.68 ± 1223.89 ng/mL × h[178]
Belamcandae Rhizoma extract50 mg/kg, oral administrationMale Sprague-Dawley ratsCmax = 1473.2 ± 156.9 ng/mL, Tmax = 0.5 ± 0.0 h, t1/2 = 3.55 ± 0.34 h, AUC(0–t) = 5800.28 ± 658.0 ng/mL × h[179]
Extract of rhizome of Iris tectorum46 mg/kg, oral administrationMale Sprague-Dawley ratsCmax = 740.3 ± 96.3 ng/mL, Tmax = 0.2 ± 0.1 h, t1/2 = 6.6 ± 2.6 h, AUC(0–t) = 4189.5 ± 60.1 ng/mL × h[180]
Tectoridin32 mg/kg, oral administrationMale Sprague-Dawley ratsCmax = 476.0 ± 57.8 ng/mL, Tmax = 0.25 ± 0.2 h, t1/2 = 3.4 ± 1.4 h, AUC(0–t) = 1760.9 ± 64.2 ng/mL × h[180]
Irisolidone100 mg/kg, oral administrationMale Sprague-Dawley ratsCmax = 0.918 ± 0.400 μmol/L, Tmax = 11.3 ± 1.03 h, t1/2 = 12.8 ± 8.28 h, AUC(0–t) = 11.8 ± 5.70 μmol/L × h[182]
Tectoridin200 mg/kg, oral administrationMale Sprague-Dawley ratsCmax = 8.67 ± 3.07 μmol/L, Tmax = 4.92 ± 2.87 h, AUC(0–t) = 72.0 ± 22.0 μmol/L × h[183]
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

Rong, J.; Fu, F.; Han, C.; Wu, Y.; Xia, Q.; Du, D. Tectorigenin: A Review of Its Sources, Pharmacology, Toxicity, and Pharmacokinetics. Molecules 2023, 28, 5904. https://doi.org/10.3390/molecules28155904

AMA Style

Rong J, Fu F, Han C, Wu Y, Xia Q, Du D. Tectorigenin: A Review of Its Sources, Pharmacology, Toxicity, and Pharmacokinetics. Molecules. 2023; 28(15):5904. https://doi.org/10.3390/molecules28155904

Chicago/Turabian Style

Rong, Juan, Fei Fu, Chenxia Han, Yaling Wu, Qing Xia, and Dan Du. 2023. "Tectorigenin: A Review of Its Sources, Pharmacology, Toxicity, and Pharmacokinetics" Molecules 28, no. 15: 5904. https://doi.org/10.3390/molecules28155904

Article Metrics

Back to TopTop