Skip to main content
  • Review
  • Published:

Myricetin: biological activity related to human health

Abstract

Myricetin (3,5,7,3′,4′,5′-hexahydroxyflavone) is one of the natural flavonols from fruit, vegetable, tea, and medicinal plants. A great deal of attention has been paid to this compound owing to its various health-promoting effects. In this review, we provide a comprehensive discussion of the therapeutic potential of myricetin by focusing on its biological activity and pharmacokinetic properties.

Myricetin (3,3′,4′,5,5′,7-hexahydroxyflavone, Fig. 1), together with quercetin and kaempferol, is a member of the class of flavonoids called flavonols, which are abundant in fruit, vegetables, tea, berries, and red wine. Myricetin is characterized by the pyrogallol B-ring, and the more hydroxylated structure is known to be responsible for its enhanced biological properties compared with other flavonols (Ono et al. 1990; Sato et al. 2013). Myricetin has been highlighted because of its various biological activities including antioxidant (Ong and Khoo 1997), anticancer (Holder et al. 2007; Phillips et al. 2011), anti-inflammatory (Tsai et al. 1999; Wang et al. 2010b), anti-amyloidogenic (Sato et al. 2013), antibacterial (Xu et al. 2001), antiviral (Pasetto et al. 2014), and antidiabetic effects (Li and Ding 2012). Nevertheless, the therapeutic potential of myricetin for human diseases has not been well explored and documented. In this review, we summarize the biological activities and pharmacokinetic properties of myricetin.

Fig. 1
figure 1

Structures of myricetin, quercetin, and kaempferol

Natural sources

Generally, flavonoid content of plant materials is determined by extraction with organic solvents followed by high-performance liquid chromatography analysis. Myricetin mainly occurs in nature in the form of glycosides and can be detected in numerous nutritious foods such as blueberry leaves, rose petals (Rosa damascena), oranges, sea buckthorn, chia seeds, carob extract, pistachio extract (Pistacia lentiscus), grape seed extract, cruciferous vegetables, peppers, garlic, and black tea (Camellia sinensis). In 2008, Sultana and Anwar (Sultana and Anwar 2008) demonstrated the flavonol contents of 22 fruits, vegetables, and medicinal plants. In that study, myricetin was the second most abundant flavonol, and the total myricetin content in the plant species ranged from 146.2 to 1660.9 mg/kg of dry matter. Among the plants evaluated, spinach (1660.9 mg/kg), cauliflower (1586.9 mg/kg), carrot (525.3 mg/kg), turnip (457.0 mg/kg), and peas (146.2 mg/kg) were shown to have high content of myricetin while cabbage, onion, and garlic-lacked myricetin. A more extensive list of myricetin aglycone content in foods, converted from glycoside values using conversion factors based on the molecular weight of the specific compounds, was composed by the United States Department of Agriculture (Table 1) (US Department of Agriculture 2003).

Table 1 Myricetin content in selected foods (USDA 2003)

Physicochemical properties

It is only recently that the physicochemical properties of myricetin were reported (Yao et al. 2014) (Table 2). The average acid dissociation constant of myricetin (pK a = 6.63 ± 0.09) indicated that myricetin would be significantly ionized at physiological pH (7.4), but myricetin was only sparingly soluble in water (16.60 ± 0.92 μg/mL). In addition, in buffered solutions, myricetin showed higher aqueous solubility at acidic pH (776.74, 576.07, and 149.75 μg/mL in acetate buffer solutions of pHs 1.0, 2.0, and 3.0, respectively). In terms of solubility in organic solvents, myricetin was readily soluble in polar organic solvents (317.23 ± 21.41 mg/mL in DMF) but strikingly less soluble in nonpolar ones (0.74 ± 0.30 μg/mL in chloroform). Typically, low solubility parallels low dissolution rates, and myricetin showed a low intrinsic dissolution rate (IDR = 11.66 ± 0.82 g/min/cm2) at 37 °C. Consistent with the low aqueous solubility, myricetin exhibited hydrophobicity (apparent partition coefficient, log P = 2.76 ± 0.05). Log P values of myricetin at different pHs (logD) were also determined, and decreases in logD pH were observed at higher pH (logD 2.0 = 2.06 ± 0.09 versus logD 7.4 = 1.34 ± 0.01). The pH-dependent decrease in logD indicates that myricetin might be better absorbed under acidic conditions, which is consistent with a previous report by Xue et al. This report showed that in rats, myricetin is most likely to be absorbed at the duodenum, a weakly acidic region in the small intestine (Xue et al. 2011). Noteworthy is the moderate membrane permeability of myricetin (apparent permeability coefficient, P app = 1.7 × 10−6 cm/s) (Yee 1997), which, in combination with the favorable logD 7.4 value for absorption, suggests that the low solubility and slow dissolution rate of myricetin are the limiting factors in low absorption of myricetin. On the other hand, the stability of myricetin was dependent on both pH and temperature, as evidenced by facile decomposition upon increases in pH and temperature (Table 2).

Table 2 Physicochemical properties of myricetin (Yao et al. 2014)

Pharmacokinetic properties and bioavailability

It was not until recently that the pharmacokinetic properties of myricetin were evaluated as a single active principle. In 2014, Dang et al. developed a specific and sensitive ultra-performance liquid chromatography-tandem mass spectrometry method for myricetin quantification in rat plasma after oral and intravenous administrations (Dang et al. 2014). According to this study, owing to poor absorption, myricetin showed low oral bioavailability (9.62 and 9.74 % at oral doses of 50 and 100 mg/kg, respectively) (Table 3). Specifically, after oral administration, both C max and AUC of myricetin increased in a dose-dependent manner, which can be attributed to the passive diffusion of myricetin in vivo. However, the relatively long T max (6.4 h) to reach C max (1 488.75 ± 200.78 ng/mL) indicates that myricetin was only slowly absorbed, presumably owing to its poor aqueous solubility.

Table 3 Pharmacokinetic parameters of myricetin (Dang et al. 2014)

Myricetin is also known to enhance the bioavailability and pharmacokinetic properties of co-administered drugs through inhibition of drug-metabolizing cytochrome P450 s (CYP450 s) and/or drug efflux pumps such as P-glycoproteins (Pgps). Thus, Li et al. reported that myricetin inhibits CYP450 s (IC50 = 7.81 μM for CYP3A4 and 13.5 μM for CYP2C9) as well as Pgp (Li et al. 2011). As a result, upon combination with myricetin, both the AUC0-∞ and C max of orally administered tamoxifen (10 mg/kg) increased by 41.8–74.4 and 48.4–81.7 %, respectively, with concurrent enhancement of the absolute bioavailability (29.0–35.7 %) when compared with a control group administered oral tamoxifen alone (20.4 %). The effect of myricetin as a metabolism-boosting drug was also demonstrated by co-administration with losartan (Choi et al. 2010). In that study, inhibition of CYP3A4, CYP2C9, and Pgp by myricetin was shown to increase the AUC0-∞ (31.4–61.1 %) and C max (31.8–50.2 %) of losartan. The ratio of the losartan metabolite to the losartan parent also decreased by 20 %, presumably owing to inhibition of CYP450-mediated metabolism of losartan. However, other pharmacokinetic properties of losartan such as T max, V d, and t 1/2 were not affected by combination with myricetin.

Pharmacology

Antioxidant effects

In biological systems, reactive oxygen species (ROS) such as superoxide anion, hydroxyl radical, hydrogen peroxide, and the nitrogen oxide radical play essential roles in homeostasis of physiological processes. However, imbalance of ROS and antioxidant enzyme activity causes several problems including protein aggregation, DNA degradation, and oxidation of membrane lipids. For example, superoxide anion radicals have been proposed to cause damage to cells and other components in bodily tissues by generating highly reactive hydroxyl radicals in a Fenton reaction (Halliwell 1978; McCord and Day 1978):

$$\begin{gathered} {\text{O}}_{2} ^{{\bar{ \cdot }}} + {\text{Fe}}^{{3 + }} \to {\text{O}}_{2} + {\text{Fe}}^{2} \hfill \\ {\text{H}}_{2} {\text{O}}_{2} + {\text{Fe}}^{{2 + }} \to {\text{OH}} \cdot + {\text{OH}}^{ - } + {\text{Fe}}^{{3 + }} \hfill \\ \end{gathered}$$

Flavonols are well known for effective reduction of free radical damage by providing antioxidant benefits, and quercetin is the most extensively studied flavonol antioxidant (Metodiewa et al. 1999; Awad et al. 2000). The proposed antioxidant mechanism of quercetin includes single electron transfer followed by hemolytic cleavage of the phenolic OH bonds in the B-ring, resulting in successive formation of the ortho-semiquinone and para-quinone methide (Fig. 2). The potent antioxidant effect of quercetin was attributed to stabilization of these intermediates by the 3′,4′-catechol moiety (Boots et al. 2008).

Fig. 2
figure 2

Antioxidant mechanism of quercetin

According to Chobot et al., the mechanism of myricetin’s antioxidant effect also includes inactivation of superoxide anion radicals through single electron transfer to form an aryloxy radical (Fig. 3A) (Chobot and Hadacek 2011). In contrast, by tracing the oxidized products of myricetin by mass spectrometry, Ghidouche et al. identified a myricetin dimer linked through a C2′–C2′ covalent bond (Fig. 3B), and therefore proposed a different antioxidant mechanism of myricetin, wherein an alkoxy radical is initially generated at the 3-position and then rapidly isomerizes to a C2′ radical that is stabilized by three phenolic hydroxyl groups (Ghidouche et al. 2007).

Fig. 3
figure 3

Proposed antioxidant mechanisms of myricetin

It is noteworthy that stabilization of the aryloxy radical in the B-ring is anticipated to increase the antioxidant activity of the flavonols because a pyrogallol moiety with a 3′,4′,5′-trihydroxy-substituted phenyl group is a better radical scavenger than a catechol moiety (Furuno et al. 2002). Thus, myricetin has been proposed as a stronger antioxidant than quercetin (Oyama et al. 1994; Gordon and Roedig-Penmanm 1998). This has been demonstrated by myricetin’s effective radical scavenging activity in various radical generating systems. These systems include the superoxide anion (O2 −) generated by phenazine methosulfate-NADH (Robak and Gryglewski 1988); hydroxyl radical (HO·) generated either by hydrogen peroxide (Hussain et al. 1987) or tert-butyl hydroperoixde (Fraga et al. 1987), and DPPH (2,2-diphenyl-1-picrylhydrazyl) radical (Khanduja and Bhardwaj 2003).

The antioxidant effect of myricetin was demonstrated in various cell-based assays. In Chinese hamster lung fibroblast cells (V97-4), myricetin showed a protective effect through inactivation of H2O2-induced radicals as well as regulation of apoptosis (Kang et al. 2010). Similarly, myricetin protected yeast cells from H2O2-induced oxidative stress that leads to a reduction in intracellular oxidation and protein carbonylation (Mendes et al. 2015). In the case of vascular endothelial dysfunction, myricetin regulated the amount of oxidized low-density lipoprotein (oxLDL), which is crucial for disease development, to increase the viability of EA.Hy926 cells (Yi et al. 2011). Myricetin was also effective in preventing methylmercury (MeHg)-induced mitochondrial dysfunction by blocking ROS generation as well as lipid peroxidation (Franco et al. 2010).

Antioxidant effect of myricetin was also observed in animal models. Myricetin decreased the generation of nitric oxide, myeloperoxidase, and malondialdehyde (MDA) while increasing the activity of superoxide dismutase (SOD) and glutathione peroxide in a mice model of dextran sulfate sodium (DSS)-induced colitis (Zhao et al. 2013). Protective effects of myricetin on hypoxia-induced mitochondrial impairments were reported, and myricetin treatment attenuated acute hypoxia-induced mitochondrial impairment in rats (Zou et al. 2015). In 7,12-dimethyl benzanthracene (DMBA)-induced breast cancer in female Wistar rats (Jayakumar et al. 2014) and in a sickle cell anemia model, myricetin was effective in preventing oxidative damage (Henneberg et al. 2013). However, according to Duthie et al., the in vivo protective effects of myricetin are limited by poor bioavailability (Duthie and Morrice 2012).

Anti-inflammatory effects

Anti-inflammatory effects are closely linked with antioxidants effect because one of the main effects of inflammation is oxidative stress. Wang et al. evaluated the in vivo anti-inflammatory effect of myricetin in both acute and chronic models of inflammation and reported a significant inhibition of xylene-induced ear edema, carrageenan-induced hind paw edema, and acetic acid-induced capillary permeability (Wang et al. 2010a). In the carrageenan-induced paw edema model, myricetin treatment resulted in a significant decrease in MDA, increased levels of SOD, and a concurrent decrease in leukocyte count. Myricetin was also effective in preventing chronic inflammation, which inhibited formation of granuloma tissue. Taken together, myricetin showed potent anti-inflammatory activity for both acute and chronic inflammation, which might be associated with its antioxidant activity (Wang et al. 2010b).

Inhibition of specific anti-inflammatory target enzymes by myricetin was also demonstrated. Myricetin inhibited 5-lipoxygenase (5-LOX) with an IC50 of 4.02 μM. This is of importance because 5-LOX is related to the production of leukotriene B4 (LTB4), a mediator of several inflammatory diseases including atherosclerosis (Yamada et al. 1999). Others have also reported stimulation of cyclooxygenases 1 and 2 (COX-1 and COX-2) by myricetin; upon co-incubation with myricetin, the catalytic activity of COX-2 increased by 124.5 % compared with control. This result was attributed to tight binding of myricetin to the peroxidase active site of COX-2, resulting in re-activation of its catalytic activity (Wang et al. 2010b). However, in vivo exposure of Sprague–Dawley rats to myricetin showed a unique biphasic pattern in the catalytic activity of COX-2. At low myricetin doses (<0.3 mg/kg), COX-2 was stimulated to convert prostaglandin G2 (PGG2) to prostaglandin E2 (PGE2), while at high doses of myricetin (>0.3 mg/kg), stimulation of COX-2 decreased (Bai and Zhu 2010).

Myricetin also showed regulatory effects on the production of cytokines involved in inflammatory responses. In lipopolysaccharide-stimulated primary macrophages and RAW264.7 macrophages, myricetin was shown to inhibit the production of interleukin-12 (IL-12) via downregulation of the binding activity of macrophages to nuclear factor kappa-B (NF-κB) (Kang et al. 2005). Production of interleukin-1β (IL-1β) in SW982 human synovial sarcoma cells was also shown to be inhibited by myricetin (Lee and Choi 2010).

Myricetin also controlled inflammation in JB6 P+ mouse epidermal cells via signaling pathways such as TNF-α-induced VEGF (vascular endothelial growth factor) upregulation by inhibition of mitogen-activated protein kinase (MAPK)/extracellular signal-regulated kinase kinase 1 (MEK1), as well as MAPK kinase 4 (MKK4) (Kim et al. 2009). In another study in TNF-α-activated ECV304 cells, myricetin suppressed the activity of TNF-α-mediated NF-κB through downregulation of inhibitor-κB kinase (IKK) (Tsai et al. 1999). Ding et al. also reported that myricetin protects from cytokine-induced cell death. It also decreased cell apoptosis induced by the cytokine mixture of TNF-α (10 ng/mL), IL-1β (5 ng/mL), and IFN-γ (1000 IU/mL) in insulin-secreting RIN-m5f β cells (Ding et al. 2012).

Anti-diabetes effects

Zamora-Ros et al. (2014) reported the statistical relationship between dietary intake of flavonoids and type II diabetes in a European population through statistical analysis. Among various flavonoids, myricetin showed an inverse association with the risk of type II diabetes, which indicates that myricetin has potential anti-diabetes activity (Zamora-Ros et al. 2014).

Diabetes is a metabolic disorder resulting hyperglycemia and development of diabetes-specific complications such as macrovascular and microvascular diseases. Diabetes is closely related to inflammation and oxidative stress. Chronic inflammatory responses, including production of cytokines, result in insulin and β-cell dysfunction, and ultimately diabetes (Hotamisligil 2006). Therefore, balancing the generation and elimination of ROS is important for pathogenesis of diabetes (Cheng et al. 2011). As described above, myricetin has antioxidant as well as anti-inflammatory effects and as a result, myricetin can be useful in preventing diabetic complications (Pandey et al. 2009).

Additionally, myricetin directly affects glycogen metabolism and insulin resistance. In a mouse model of diabetes, treatment with myricetin for 2 days (3 mg/12 h) resulted in 50 % reduction of hyperglycemia compared with control along with normalization of hypertriglyceridemia (Ong and Khoo 2000). Myricetin was also reported to lower plasma glucose levels in streptozotocin-induced diabetes in rats (Liu et al. 2005) and in a rat model of insulin resistance (Tzeng et al. 2011). In the insulin resistant rat model in particular, myricetin reduced insulin resistance by increasing endogenous production of β-endorphins followed by activation of μ-opioid receptors, which ameliorated defective post-receptor insulin signaling (Tzeng et al. 2011). In mice that fed a high-fat and high-sucrose diet, 0.12 % myricetin supplementation significantly improved hypertriglyceridemia and hypercholesterolemia, which suggests that myricetin may have a protective effect against diet-induced obesity and insulin resistance (Choi et al. 2014a).

Myricetin was found to inhibit glucose uptake in isolated rat adipocytes through direct interaction with glucose-transporter subtype 4 (GLUT4) (Strobel et al. 2005). Myricetin was found to block the uptake of methylglucose by inhibiting GLUT4. However, phosphorylation of insulin receptor substrate 1 via insulin receptor tyrosine kinase was not affected by myricetin in insulin-stimulated rat adipocytes (Strobel et al. 2005).

Aggregation of islet amyloid polypeptide (IAPP) results in pancreatic β-cell death in type II diabetes. IAPP has thus been considered as a potential target for type II diabetes. According to Zelus et al., myricetin not only strongly inhibits the formation of IAPP, but also protects from the neurotoxic effect of IAPP in PC12 cells (Zelus et al. 2012).

Protective effects against Alzheimer’s disease

Major hallmarks of Alzheimer’s disease are amyloid beta (Aβ) fibrils and neurofibrillary tangles (NFTs) that are composed of highly phosphorylated forms of the microtubule-associated protein tau. Myricetin showed therapeutic potential against Alzheimer’s disease by inhibiting the formation of both Aβ fibrils and NFTs.

Aβ peptides result from cleavage of the amyloid precursor protein (APP) catalyzed by β-secretase (BACE) and γ-secretase, and myricetin is known to inhibit BACE-1 (Shimmyo et al. 2008a). Upon exposure to stress, Aβ peptides are aggregated to Aβ fibrils. The formation of Aβ fibrils results in neurotoxicity by producing ROS. Myricetin showed neuroprotective effects by scavenging radicals induced by Aβ fibrils (Choi et al. 2014b).

Myricetin was found to have a higher binding affinity for Aβ fibrils than for Aβ monomer, and it was claimed that oxidized myricetin shows higher binding affinity for Aβ fibrils than un-oxidized myricetin (Hirohata et al. 2007). It was also demonstrated that myricetin inhibited the formation of Aβ fibrils (Bartolini et al. 2011). Sato et al. reported the half-maximal inhibitory concentration (IC50) of myricetin against Aβ fibril formation to be 15.1 μM (Sato et al. 2013). Shimmyo et al. proposed that the underlying molecular mechanism for inhibition of Aβ fibrillogenesis by myricetin includes preventing a conformational change of Aβ to a toxic β-sheet structure (Shimmyo et al. 2008b). A molecular docking study of myricetin to Aβ fibrils was performed to show the binding mode of myricetin (Keshet et al. 2010).

Recently, Aβ oligomers have been suggested as more important in Alzheimer’s disease pathogenesis than Aβ fibrils. Interestingly, myricetin was also found to be an effective inhibitor of formation of Aβ oligomers (Ono et al. 2012). Necular et al. determined the IC50 value of myricetin (106.87 μM) against Aβ oligomerization using an oligomer-specific antibody (A11) (Necula et al. 2007). In a transgenic mouse model (Tg2576) of Alzheimer’s disease, myricetin was shown to decrease Aβ oligomers by inhibition of oligomerization (Hamaguchi et al. 2009).

Myricetin treatment also reduced abnormal fibrillation (IC50 = 1.2 μM), as well as oligomerization, of the tau protein (Taniguchi et al. 2005). This result was supported by molecular dynamics calculation of the binding energy of myricetin to the VQIVYK oligomer fragment of tau (Berhanu and Masunov 2015).

Anticancer effects

Myricetin shows anticancer activity against various kinds of cancers through different molecular mechanisms. Due to the structural similarity with ATP (Fig. 4), myricetin is able to mimic ATP. Thus, protein kinases and ATPases are anticancer targets of myricetin.

Fig. 4
figure 4

Structural comparison of adenosine and myricetin

There have been numerous reports regarding competitive or noncompetitive inhibition of protein kinases by myricetin. For example, Philips et al. suggested that myricetin would show anticancer activity in pancreatic cancer by acting as an inhibitor of phosphatidylinositol 3-kinase. This was confirmed by tumor regression and decreased metastatic spread in vivo (Phillips et al. 2011). The serine/threonine kinase PIM1, an oncogenic protein that has been implicated in the development of leukemias, lymphomas, and prostate cancer, was also shown to be inhibited by myricetin with an IC50 value of 0.78 μM (Holder et al. 2007). Zhang et al. reported that myricetin inhibits proliferation of human hepatoma HepG2 cells by arresting the cell cycle at the G2/M phase through inhibition of cyclin-dependent kinase 1 (Zhang et al. 2011). MKK4, another anti-tumor target, was shown to be inhibited by myricetin in an ATP-competitive manner (Kim et al. 2009). Receptor tyrosine kinase is another target for myricetin. The tyrosine kinase Met, a receptor of hepatocyte growth factor (HGF), has been proposed as an attractive target for prevention and treatment of human medulloblastoma. Labbé et al. observed that myricetin (5 μmol/L) inhibited HGF/Met signaling in a medulloblastoma cell line (Labbé et al. 2009).

Similar to kinases, ATPases also use ATP for their catalytic activity, and several ATPases such as Hsp70 ATPase (Koren et al. 2010) have been reported to be inhibited by myricetin.

Other molecular targets of myricetin for cancer have been reported. In colorectal carcinoma cells, myricetin showed inhibitory activity against matrix metalloproteinase-2, with an IC50 of 7.82 μmol/L (Ko et al. 2005). Lu et al. reported inhibition of thioredoxin reductase (TrxR), a potential anticancer target, by myricetin (IC50 = 0.63 μmol/L) in TrxR-overexpressed A549 cells and suggested that the oxidized form of myricetin is utilized as both inhibitor and substrate (Lu et al. 2006).

Myricetin has been shown to have maximal chemoprotective effects in a mouse model of skin tumorigenesis induced by PAHs and N-methyl-N-nitrosourea (Mukhtar et al. 1988). In this model, myricetin afforded significant protection against skin tumorgenicity by delaying the onset and the subsequent development of skin tumors (Mukhtar et al. 1988).

Several reports suggest that myricetin can also be used as a chemosensitizer for cancer. Myricetin increased the radiosensitivity of lung cancer cells (A549 and H1299 cells), resulting in increased apoptosis of cancer cells after radiotherapy. This was indicated by monitoring the expression levels of caspase-3 (Zhang et al. 2014). In esophageal cancer EC9706 cells, myricetin enhanced chemosensitivity of 5-fluorouracil, and the combination therapy of myricetin and 5-fluorouracil resulted in suppression of tumor growth in tumor-bearing nude mice (Wang et al. 2014).

Modulation of ion channels

Ca2+ channels are important for various cellular functions including contraction of vascular smooth muscle and modulation of intracellular Ca2+ concentration for pain signals. There are many reports demonstrating that myricetin has a stimulatory effect on Ca2+ channels, which might be useful for the treatment of hypertension and stroke. In rat tail artery myocytes, myricetin stimulated vascular Cav 1.2 channels in a concentration- and voltage-dependent manner with an EC50 value of 16 μM (Saponara et al. 2011). In rat aorta rings, myricetin induced contractile responses in vascular smooth muscle L-type Ca2+ channels with a pEC50 value of 4.43 ± 0.03. A functional study observed that myricetin induced vasoconstriction by activating L-type Ca2+ channels with similar efficacy but at a site of action different than that of the Ca2+ channel agonist, (S)-(−)-Bay K 8644 (Fusi et al. 2003). Interestingly, the stimulatory effect of myricetin was exerted only on the L-type, but not on the T-type, Ca2+ channel (Fusi et al. 2005). In neurons of hypothalamic paraventricular nucleus (PVN), myricetin increased Ca2+ currents and intracellular Ca2+ concentration through T- and L-type Ca2+ channels to enhance GABAA receptor activity (Zhang et al. 2012).

Behavioral studies indicated a potential analgesic effect of myricetin in a neuropathic pain model of spinal nerve ligation (SNL) in rats. Injection of myricetin (0.1–10 mg/kg i.p.) reduced SNL-induced mechanical allodynia and thermal hyperalgesia (Hagenacker et al. 2010).

Anti-infectious activity

Antiviral activity of myricetin against several viruses, including the human immunodeficiency virus (HIV) and human papillomaviruses (HPVs), has been reported. Pasetto, et al. tested the antiviral activity of myricetin against several HIV-1 strains. Myricetin showed potent inhibitory activity, as indicated by IC50 values, in several cells lines: 20.43 μM (HIV-1 BaL-infected TZM-b1 cells), 22.91 μM (HIV-1 MN-infected H9 cells), 1.76 μM (HIV-1 89.6-infected H9 cells), 4.49 μM (HIV-1 MN-infected PBMC cells), and 3.23 μM (HIV-1 89.6-infected PBMC cells) (Pasetto et al. 2014).

HPVs cause cervical as well as head and neck cancers, which might be attributed to two on coproteins (E6 and E7). In particular, binding of E6 to a Fas-associated protein containing a death domain and caspase-8 can prevent apoptosis of infected cells. However, the HPV-induced apoptotic blockade was shown to be prevented or reversed by myricetin with an IC50 of 850 nM (Yuan et al. 2012).

Myricetin showed antibacterial effects against methicillin-resistant Staphylococcus aureus, multidrug-resistant Burkholderia cepacia, vancomycin-resistant enterococci, and other medically important organisms such as Klebsiella pneumoniae and Staphylococcus epidermidis with minimum inhibitory concentrations (MIC) of 32–256 μg/mL (Xu and Lee 2001).

The bacterial replicative helicase DnaB helicase was shown to be the major target enzyme for antibacterial activity of myricetin. Myricetin noncompetitively inhibited Escherichia coli DnaB helicase (IC50 = 11.3 μM) which, along with primase, plays a central role during DNA replication and elongation (Griep et al. 2007). Myricetin regulated K. pneumoniae DnaB helicase (KpDnaB) activity to decrease its ssDNA-stimulated ATPase activity (59 %) as well as its dsDNA-unwinding activity (Lin and Huang 2012).

Myricetin has found additional usefulness in prevention of bacterial/viral infection by stimulating epithelial Cl− secretion (Sun et al. 2014). In contrast to quercetin, which stimulated epithelial Cl− secretion only under basal conditions in epithelial A6 cells, myricetin was able to stimulate epithelial Cl− secretion even under cAMP-stimulated conditions (Sun et al. 2014).

Other biological activities

Myricetin showed the anti-obesity effects in a rat model of diet-induced obesity. In this model, myricetin reduced not only body weight, visceral fat-pad weight, and plasma lipid levels but also decreased triglycerides and cholesterol. These effects were achieved by upregulation of peroxisome proliferator-activated receptor α and downregulation of sterol regulatory element binding proteins (SREBPs) expression (Chang et al. 2012).

Lifespan was prolonged (32.9 %) by myricetin in C. elegans through nuclear translocation of DAF-16. This was confirmed by a complete loss of myricetin-induced lifespan expansion in a DAF-16 loss-of-function mutant strain (CF1038). Myricetin also decreased the formation of lipofuscin, a biomarker of aging (Buchter et al. 2013).

Inhibition of soluble N-ethylmaleimide-sensitive factor attachment protein receptor (SNARE) zippering by myricetin was also reported (Yang et al. 2010). Neuronal SNARE proteins facilitate the synaptic fusion of vesicles to the presynaptic plasma membrane and mediate neurotransmitter release. Myricetin intercalated into the hydrophobic inner core near the middle of the SNARE complex to stop SNARE zippering.

Myricetin has been suggested to interfere with the intestinal folate transport system, and Furumiya et al. demonstrated that myricetin inhibited proton-coupled folate transporter in a noncompetitive manner both in Caco-2 (K m = 0.407 µM, K i = 61 µM) and MDCK cells (K m = 1.246 µM, K i = 130 µM) (Furumiya et al. 2014).

References

  • Awad HM, Boersma MG, Vervoort J, Rietjens IMCM (2000) Peroxidase-catalyzed formation of quercetin quinone methide–glutathione adducts. Arch Biochem Biophys 378:224–233

    Article  CAS  Google Scholar 

  • Bai HW, Zhu BT (2010) Myricetin and quercetin are naturally occurring co-substrates of cyclooxygenase in vivo. Prostaglandins Leukot Essent 82:45–50

    Article  CAS  Google Scholar 

  • Bartolini M, Naldi M, Fiori J, Valle F, Biscarini F, Nicolau DV, Andrisano V (2011) Kinetic characterization of amyloid-beta 1-42 aggregation with a multimethodological approach. Anal Biochem 414:215–225

    Article  CAS  Google Scholar 

  • Berhanu WM, Masunov AE (2015) Atomistic mechanism of polyphenol amyloid aggregation inhibitors: molecular dynamics study of curcumin, exifone, and myricetin interaction with the segment of tau peptide oligomer. J Biomol Struct Dyn 33:1399–1411

    Article  CAS  Google Scholar 

  • Boots AW, Haenen GR, Bast A (2008) Health effects of quercetin: from antioxidant to nutraceutical. Eur J Pharmacol 585:325–327

    Article  CAS  Google Scholar 

  • Buchter C, Ackermann D, Havermann S, Honnen S, Chovolou Y, Fritz G, Watjen W (2013) Myricetin-mediated lifespan extension in Caenorhabditis elegans is modulated by DAF-16. Int J Mol Sci 14:11895–11914

    Article  Google Scholar 

  • Chang CJ, Tzeng TF, Liou SS, Chang YS, Liu IM (2012) Myricetin increases hepatic peroxisome proliferator-activated receptor alpha protein expression and decreases plasma lipids and adiposity in rats. Evid Based Complement Alternat Med 2012:787152

    Google Scholar 

  • Cheng X, Siow RC, Mann GE (2011) Impaired redox signaling and antioxidant gene expression in endothelial cells in diabetes: a role for mitochondria and the nuclear factor-E2-related factor 2-Kelch-like ECH-associated protein 1 defense pathway. Antioxid Redox Signal 14:469–487

    Article  CAS  Google Scholar 

  • Chobot V, Hadacek F (2011) Exploration of pro-oxidant and antioxidant activities of the flavonoid myricetin. Redox Rep 16:242–247

    Article  CAS  Google Scholar 

  • Choi DH, Li C, Choi JS (2010) Effects of myricetin, an antioxidant, on the pharmacokinetics of losartan and its active metabolite, EXP-3174, in rats: possible role of cytochrome P450 3A4, cytochrome P450 2C9 and P-glycoprotein inhibition by myricetin. J Pharm Pharmacol 62:908–914

    Article  CAS  Google Scholar 

  • Choi HN, Kang MJ, Lee SJ, Kim JI (2014a) Ameliorative effect of myricetin on insulin resistance in mice fed a high-fat, high-sucrose diet. Nutr Res Pract 8:544–549

    Article  Google Scholar 

  • Choi SM, Kim BC, Cho YH, Choi KH, Chang J, Park MS, Kim JK (2014b) Effects of flavonoid compounds on β-amyloid-peptide-induced neuronal death in cultured mouse cortical neurons. Chonnam Med J 50:45–51

    Article  CAS  Google Scholar 

  • Dang Y, Lin G, Xie Y, Duan J, Ma P, Li G, Ji G (2014) Quantitative determination of myricetin in rat plasma by ultraperformance liquid chromatography tandem mass spectrometry and its absolute bioavailability. Drug Res (Stuttg) 64:516–522

    CAS  Google Scholar 

  • Ding Y, Zhang ZF, Dai XQ, Li Y (2012) Myricetin protects against cytokine-induced cell death in RIN-m5f beta cells. J Med Food 15:733–740

    Article  CAS  Google Scholar 

  • Duthie G, Morrice P (2012) Antioxidant capacity of flavonoids in hepatic microsomes is not reflected by antioxidant effects in vivo. Oxid Med Cell Longev 2012:165127

    Article  Google Scholar 

  • Fraga CG, Martino VS, Ferraro GE, Coussio JD, Boveris A (1987) Flavonoids as antioxidants evaluated by in vitro and in situ liver chemiluminescence. Biochem Pharmacol 36:717–720

    Article  CAS  Google Scholar 

  • Franco JL, Posser T, Missau F, Pizzolatti MG, Dos Santos AR, Souza DO, Farina M (2010) Structure-activity relationship of flavonoids derived from medicinal plants in preventing methylmercury-induced mitochondrial dysfunction. Environ Toxicol Pharmacol 30:272–278

    Article  CAS  Google Scholar 

  • Furumiya M, Inoue K, Nishijima C, Yamashiro T, Inaoka E, Ohta K, Yuasa H (2014) Noncompetitive inhibition of proton-coupled folate transporter by myricetin. Drug Metab Pharmacokinet 29:312–316

    Article  CAS  Google Scholar 

  • Furuno K, Akasako T, Sugihara N (2002) The contribution of the pyrogallol moiety to the superoxide radical scavenging activity of flavonoids. Biol Pharm Bull 25:19–23

    Article  CAS  Google Scholar 

  • Fusi F, Saponara S, Frosini M, Gorelli B, Sgaragli G (2003) L-type Ca2+ channels activation and contraction elicited by myricetin on vascular smooth muscles. Naunyn Schmiedebergs Arch Pharmacol 368:470–478

    Article  CAS  Google Scholar 

  • Fusi F, Sgaragli G, Saponara S (2005) Mechanism of myricetin stimulation of vascular L-type Ca2+ current. J Pharmacol Exp Ther 313:790–797

    Article  CAS  Google Scholar 

  • Ghidouche S, Es-Safi NE, Ducrot PH (2007) Trtametes versicolor laccase mediated oxidation of flavonoids. influence of the hydroxylation pattern of ring B of flavonols. Am J Food Technol 2:630–640

    Article  Google Scholar 

  • Gordon MH, Roedig-Penmanm A (1998) Antioxidant activity of quercetin and myricetin in liposomes. Chem Phys Lipids 97:79–85

    Article  CAS  Google Scholar 

  • Griep MA, Blood S, Larson MA, Koepsell SA, Hinrichs SH (2007) Myricetin inhibits Escherichia coli DnaB helicase but not primase. Bioorg Med Chem 15:7203–7208

    Article  CAS  Google Scholar 

  • Hagenacker T, Hillebrand I, Wissmann A, Busselberg D, Schafers M (2010) Anti-allodynic effect of the flavonoid myricetin in a rat model of neuropathic pain: involvement of p38 and protein kinase C mediated modulation of Ca2+ channels. Eur J Pain 14:992–998

    Article  CAS  Google Scholar 

  • Halliwell B (1978) Superoxide-dependant formation of hydroxyl radicals in the presence of iron chelates: is it a mechanism for hydroxyl radical production in biological systems? FEBS Lett 92:321–326

    Article  CAS  Google Scholar 

  • Hamaguchi T, Ono K, Murase A, Yamada M (2009) Phenolic compounds prevent Alzheimer’s pathology through different effects on the amyloid-beta aggregation pathway. Am J Pathol 175:2557–2565

    Article  CAS  Google Scholar 

  • Henneberg R, Otuki MF, Furman AE, Hermann P, do Nascimento AJ, Leonart MS (2013) Protective effect of flavonoids against reactive oxygen species production in sickle cell anemia patients treated with hydroxyurea. Rev Bras Hematol Hemoter 35:52–55

  • Hirohata M, Hasegawa K, Tsutsumi-Yasuhara S, Ohhashi Y, Ookoshi T, Ono K, Naiki H (2007) The anti-amyloidogenic effect is exerted against Alzheimer’s beta-amyloid fibrils in vitro by preferential and reversible binding of flavonoids to the amyloid fibril structure. Biochemistry 46:1888–1899

    Article  CAS  Google Scholar 

  • Holder S, Zemskova M, Zhang C, Tabrizizad M, Bremer R, Neidigh JW, Lilly MB (2007) Characterization of a potent and selective small-molecule inhibitor of the PIM1 kinase. Mol Cancer Ther 6:163–172

    Article  CAS  Google Scholar 

  • Hotamisligil GS (2006) Inflammation and metabolic disorders. Nature 444:860–867

    Article  CAS  Google Scholar 

  • Hussain SR, Cillard J, Cillard P (1987) Hydroxyl radical scavenging activity of flavonoids. Phytochemistry 26:2489–2491

    Article  Google Scholar 

  • Jayakumar JK, Nirmala P, Praveen Kumar BA, Kumar AP (2014) Evaluation of protective effect of myricetin, a bioflavonoid in dimethyl benzanthracene-induced breast cancer in female Wistar rats. South Asian J Cancer 3:107–111

    Article  CAS  Google Scholar 

  • Kang BY, Kim SH, Cho D, Kim TS (2005) Inhibition of interleukin-12 production in mouse macrophages via decreased nuclear factor-kappaB DNA binding activity by myricetin, a naturally occurring flavonoid. Arch Pharm Res 28:274–279

    Article  CAS  Google Scholar 

  • Kang KA, Wang ZH, Zhang R, Piao MJ, Kim KC, Kang SS, Hyun JW (2010) Myricetin protects cells against oxidative stress-induced apoptosis via regulation of PI3 K/Akt and MAPK signaling pathways. Int J Mol Sci 11:4348–4360

    Article  CAS  Google Scholar 

  • Keshet B, Gray JJ, Good TA (2010) Structurally distinct toxicity inhibitors bind at common loci on beta-amyloid fibril. Protein Sci 19:2291–2304

    Article  CAS  Google Scholar 

  • Khanduja KL, Bhardwaj A (2003) Stable free radical scavenging and antiperoxidative properties of resveratrol compared in vitro with some other bioflavonoids. Indian J Biochem Biophys 40:416–422

    CAS  Google Scholar 

  • Kim JE, Kwon JY, Lee DE, Kang NJ, Heo YS, Lee KW, Lee HJ (2009) MKK4 is a novel target for the inhibition of tumor necrosis factor-alpha-induced vascular endothelial growth factor expression by myricetin. Biochem Pharmacol 77:412–421

    Article  CAS  Google Scholar 

  • Ko CH, Shen SC, Lee TJ, Chen YC (2005) Myricetin inhibits matrix metalloproteinase 2 protein expression and enzyme activity in colorectal carcinoma cells. Mol Cancer Ther 4:281–290

    CAS  Google Scholar 

  • Koren J, Jinwal UK, Jin Y, O’Leary J, Jones JR, Johnson AG, Dickey CA (2010) Facilitating Akt clearance via manipulation of Hsp70 activity and levels. J Biol Chem 285:2498–2505

    Article  CAS  Google Scholar 

  • Labbé D, Provencal M, Lamy S, Boivin D, Gingras D, Beliveau R (2009) The flavonols quercetin, kaempferol, and myricetin inhibit hepatocyte growth factor-induced medulloblastoma cell migration. J Nutr 139:646–652

    Article  Google Scholar 

  • Lee YS, Choi EM (2010) Myricetin inhibits IL-1 beta-induced inflammatory mediators in SW982 human synovial sarcoma cells. Int Immunopharmacol 10:812–814

    Article  CAS  Google Scholar 

  • Li Y, Ding Y (2012) Minireview: therapeutic potential of myricetin in diabetes mellitus. Food Sci Human Well 1:19–25

    Article  Google Scholar 

  • Li C, Lim SC, Kim J, Choi JS (2011) Effects of myricetin, an anticancer compound, on the bioavailability and pharmacokinetics of tamoxifen and its main metabolite, 4-hydroxytamoxifen, in rats. Eur J Drug Metab Pharmacokinet 36:175–182

    Article  CAS  Google Scholar 

  • Lin HH, Huang CY (2012) Characterization of flavonol inhibition of DnaB helicase: real-time monitoring, structural modeling, and proposed mechanism. J Biomed Biotechnol 2012:735368

    Google Scholar 

  • Liu IM, Liou SS, Lan TW, Hsu FL, Cheng JT (2005) Myricetin as the active principle of Abelmoschus moschatus to lower plasma glucose in streptozotocin-induced diabetic rats. Planta Med 71:617–621

    Article  CAS  Google Scholar 

  • Lu J, Papp LV, Fang J, Rodriguez-Nieto S, Zhivotovsky B, Holmgren A (2006) Inhibition of mammalian thioredoxin reductase by some flavonoids: implications for myricetin and quercetin anticancer activity. Cancer Res 66:4410–4418

    Article  CAS  Google Scholar 

  • McCord JM, Day ED (1978) Superoxide-dependant production of hydroxyl radical catalyzed by iron-EDTA complex. FEBS Lett 86:139–142

    Article  CAS  Google Scholar 

  • Mendes V, Vilaca R, Freitas V, Ferreira PM, Mateus N, Costa V (2015) Effect of myricetin, pyrogallol, and phloroglucinol on yeast resistance to oxidative stress. Oxid Med Cell Longev 2015:782504

    Article  Google Scholar 

  • Metodiewa D, Jaiswal AK, Cenas N, Dickancaité E, Segura-Aguilar J (1999) Quercetin may act as a cytotoxic prooxidant after its metabolic activation to semiquinone and quinoidal product. Free Radical Biol Med 26:107–116

    Article  CAS  Google Scholar 

  • Mukhtar H, Das M, Khan WA, Wang ZY, Bik DP, Bickers DR (1988) Exceptional activity of tannic acid among naturally occurring plant phenols in protecting against 7,12-dimethylbenz(a)anthracene-, benzo(a)pyrene-, 3-methylcholanthrene-, and N-methyl-N-nitrosourea-induced skin tumorigenesis in mice. Cancer Res 48:2361–2365

    CAS  Google Scholar 

  • Necula M, Kayed R, Milton S, Glabe CG (2007) Small molecule inhibitors of aggregation indicate that amyloid beta oligomerization and fibrillization pathways are independent and distinct. J Biol Chem 282:10311–10324

    Article  CAS  Google Scholar 

  • Ong KC, Khoo HE (1997) Biological effects of myricetin. Gen Pharmacol 29:121–126

    Article  CAS  Google Scholar 

  • Ong KC, Khoo HE (2000) Effects of myricetin on glycemia and glycogen metabolism in diabetic rats. Life Sci 67:1695–1705

    Article  CAS  Google Scholar 

  • Ono K, Nakane H, Fukushima M, Chermann JC, Barre-Sinoussi F (1990) Differential inhibitory effects of various flavonoids on the activities of reverse transcriptase and cellular DNA and RNA polymerases. Eur J Biochem 190:469–476

    Article  CAS  Google Scholar 

  • Ono K, Li L, Takamura Y, Yoshiike Y, Zhu L, Han F, Yamada M (2012) Phenolic compounds prevent amyloid beta-protein oligomerization and synaptic dysfunction by site-specific binding. J Biol Chem 287:14631–14643

    Article  CAS  Google Scholar 

  • Oyama Y, Fuchs PA, Katayama N, Noda K (1994) Myricetin and quercetin, the flavonoid constituents of Ginkgo biloba extract, greatly reduce oxidative metabolism in both resting and Ca(2 +)-loaded brain neurons. Brain Res 635:125–129

    Article  CAS  Google Scholar 

  • Pandey KB, Mishra N, Rizvi SI (2009) Myricetin may provide protection against oxidative stress in type 2 diabetic erythrocytes. Z Naturforsch C 64:626–630

    Article  CAS  Google Scholar 

  • Pasetto S, Pardi V, Murata RM (2014) Anti-HIV-1 activity of flavonoid myricetin on HIV-1 infection in a dual-chamber in vitro model. Plos One 9:e115323

    Article  Google Scholar 

  • Phillips PA, Sangwan V, Borja-Cacho D, Dudeja V, Vickers SM, Saluja AK (2011) Myricetin induces pancreatic cancer cell death via the induction of apoptosis and inhibition of the phosphatidylinositol 3-kinase (PI3 K) signaling pathway. Cancer Lett 308:181–188

    Article  CAS  Google Scholar 

  • Robak J, Gryglewski RJ (1988) Flavonoids are scavengers of superoxide anions. Biochem Pharmacol 37:837–841

    Article  CAS  Google Scholar 

  • Saponara S, Carosati E, Mugnai P, Sgaragli G, Fusi F (2011) The flavonoid scaffold as a template for the design of modulators of the vascular Ca(v) 1.2 channels. Br J Pharmacol 164:1684–1697

    Article  CAS  Google Scholar 

  • Sato M, Murakami K, Uno M, Nakagawa Y, Katayama S, Akagi K, Irie K (2013) Site-specific inhibitory mechanism for amyloid β42 aggregation by catechol-type flavonoids targeting the Lys residues. J Biol Chem 288:23212–23224

    Article  CAS  Google Scholar 

  • Shimmyo Y, Kihara T, Akaike A, Niidome T, Sugimoto H (2008a) Flavonols and flavones as BACE-1 inhibitors: structure-activity relationship in cell-free, cell-based and in silico studies reveal novel pharmacophore features. Biochim Biophys Acta 1780:819–825

    Article  CAS  Google Scholar 

  • Shimmyo Y, Kihara T, Akaike A, Niidome T, Sugimoto H (2008b) Multifunction of myricetin on Aβ: neuroprotection via a conformational change of Aβ and reduction of Aβ via the interference of secretases. J Neurosci Res 86:368–377

    Article  CAS  Google Scholar 

  • Strobel P, Allard C, Perez-Acle T, Calderon R, Aldunate R, Leighton F (2005) Myricetin, quercetin and catechin-gallate inhibit glucose uptake in isolated rat adipocytes. Biochem J 386:471–478

    Article  CAS  Google Scholar 

  • Sultana B, Anwar F (2008) Flavonols (kaempeferol, quercetin, myricetin) contents of selected fruits, vegetables and medicinal plants. Food Chem 108:879–884

    Article  CAS  Google Scholar 

  • Sun H, Niisato N, Nishio K, Hamilton KL, Marunaka Y (2014) Distinct action of flavonoids, myricetin and quercetin, on epithelial Cl− secretion: useful tools as regulators of Cl− secretion. Biomed Res Int 2014:902735

    Google Scholar 

  • Taniguchi S, Suzuki N, Masuda M, Hisanaga S, Iwatsubo T, Goedert M, Hasegawa M (2005) Inhibition of heparin-induced tau filament formation by phenothiazines, polyphenols, and porphyrins. J Biol Chem 280:7614–7623

    Article  CAS  Google Scholar 

  • Tsai SH, Liang YC, Lin-Shiau SY, Lin JK (1999) Suppression of TNF alpha-mediated NF-kappaB activity by myricetin and other flavonoids through downregulating the activity of IKK in ECV304 cells. J Cell Biochem 74:606–615

    Article  CAS  Google Scholar 

  • Tzeng TF, Liou SS, Liu IM (2011) Myricetin ameliorates defective post-receptor insulin signaling via β-endorphin signaling in the skeletal muscles of fructose-fed rats. Evid Based Complement Alternat Med 2011:150752

    Article  Google Scholar 

  • US Department of Agriculture. (2003) USDA database for the flavonoid content of selected foods. Beltsville (MD): US Department of Agriculture

  • Wang P, Bai HW, Zhu BT (2010a) Structural basis for certain naturally occurring bioflavonoids to function as reducing co-substrates of cyclooxygenase I and II. Plos One 5:e12316

    Article  Google Scholar 

  • Wang SJ, Tong Y, Lu S, Yang R, Liao X, Xu YF, Li X (2010b) Anti-inflammatory activity of myricetin isolated from Myrica rubra Sieb. et Zucc. leaves. Planta Med 76:1492–1496

    Article  CAS  Google Scholar 

  • Wang L, Feng J, Chen X, Guo W, Du Y, Wang Y, Zhao G (2014) Myricetin enhance chemosensitivity of 5-fluorouracil on esophageal carcinoma in vitro and in vivo. Cancer Cell Int 14:71–78

    Article  Google Scholar 

  • Xu HX, Lee SF (2001) Activity of plant flavonoids against antibiotic-resistant bacteria. Phytother Res 15:39–43

    Article  Google Scholar 

  • Xu H, Ziegelin G, Schroder W, Frank J, Ayora S, Alonso JC, Saenger W (2001) Flavones inhibit the hexameric replicative helicase RepA. Nucleic Acids Res 29:5058–5066

    Article  CAS  Google Scholar 

  • Xue CF, Guo JM, Qian DW, Duan JA, Shu Y (2011) Absorption of flavonoids from Abelmoschus manihot extract by in situ intestinal perfusion. Yao Xue Xue Bao 46:454–459

    CAS  Google Scholar 

  • Yamada K, Shoji K, Mori M, Ueyama T, Matsuo N, Oka S, Nishiyama K, Sugano M (1999) Structure-activity relationship of polyphenols on inhibition of chemical mediator release from rat peritoneal exudate cells. In Vitro Cell Dev Biol Anim 35:169–174

    Article  CAS  Google Scholar 

  • Yang Y, Shin JY, Oh JM, Jung CH, Hwang Y, Kim S, Kweon DH (2010) Dissection of SNARE-driven membrane fusion and neuroexocytosis by wedging small hydrophobic molecules into the SNARE zipper. Proc Natl Acad Sci 107:22145–22150

    Article  CAS  Google Scholar 

  • Yao Y, Lin G, Xie Y, Ma P, Li G, Meng Q, Wu T (2014) Preformulation studies of myricetin: a natural antioxidant flavonoid. Pharmazie 69:19–26

    CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  • Yi L, Jin X, Chen CY, Fu YJ, Zhang T, Chang H, Mi MT (2011) Chemical structures of 4-oxo-flavonoids in relation to inhibition of oxidized low-density lipoprotein (LDL)-induced vascular endothelial dysfunction. Int J Mol Sci 12:5471–5489

    Article  CAS  Google Scholar 

  • Yuan CH, Filippova M, Tungteakkhun SS, Duerksen-Hughes PJ, Krstenansky JL (2012) Small molecule inhibitors of the HPV16-E6 interaction with caspase 8. Bioorg Med Chem Lett 22:2125–2129

    Article  CAS  Google Scholar 

  • Zamora-Ros R, Forouhi NG, Sharp SJ, Gonzalez CA, Buijsse B, Guevara M, Wareham NJ (2014) Dietary intakes of individual flavanols and flavonols are inversely associated with incident type 2 diabetes in European populations. J Nutr 144:335–343

    Article  CAS  Google Scholar 

  • Zelus C, Fox A, Calciano A, Faridian BS, Nogaj LA, Moffet DA (2012) Myricetin inhibits islet amyloid polypeptide (IAPP) aggregation and rescues living mammalian cells from IAPP toxicity. Open Biochem J 6:66–70

    Article  CAS  Google Scholar 

  • Zhang XH, Zou ZQ, Xu CW, Shen YZ, Li D (2011) Myricetin induces G2/M phase arrest in HepG2 cells by inhibiting the activity of the cyclin B/Cdc2 complex. Mol Med Rep 4:273–277

    CAS  Google Scholar 

  • Zhang XH, Ma ZG, Rowlands DK, Gou YL, Fok KL, Wong HY, Chan HC (2012) Flavonoid myricetin modulates GABAA receptor activity through activation of Ca2+ channels and CaMK-II pathway. Evid Based Complement Alternat Med 2012:758097

    Google Scholar 

  • Zhang S, Wang L, Liu H, Zhao G, Ming L (2014) Enhancement of recombinant myricetin on the radiosensitivity of lung cancer A549 and H1299 cells. Diagn Pathol 9:68–75

    Article  Google Scholar 

  • Zhao J, Hong T, Dong M, Meng Y, Mu J (2013) Protective effect of myricetin in dextran sulphate sodium-induced murine ulcerative colitis. Mol Med Rep 7:565–570

    Google Scholar 

  • Zou D, Liu P, Chen K, Xie Q, Liang X, Bai Q, Mi M (2015) Protective effects of myricetin on acute hypoxia-induced exercise intolerance and mitochondrial impairments in rats. Plos One 10:e0124727

    Article  Google Scholar 

Download references

Acknowledgments

This research was supported by the Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education (NRF-2013R1A1A2006455) and by a Grant from the Priority Research Centers Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education, Science, and Technology (2009-0093824).

Author information

Authors and Affiliations

Authors

Corresponding authors

Correspondence to Youhoon Chong or Mi Kyoung Kim.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Park, KS., Chong, Y. & Kim, M.K. Myricetin: biological activity related to human health. Appl Biol Chem 59, 259–269 (2016). https://doi.org/10.1007/s13765-016-0150-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s13765-016-0150-2

Keywords