Open access peer-reviewed chapter

Investigating the Antioxidant Properties of Quercetin

Written By

Kate Nyarko

Submitted: 13 February 2024 Reviewed: 14 February 2024 Published: 04 September 2024

DOI: 10.5772/intechopen.1004648

From the Edited Volume

Quercetin - Effects on Human Health

Joško Osredkar

Chapter metrics overview

19 Chapter Downloads

View Full Metrics

Abstract

The antioxidant properties of quercetin stem from its ability to neutralize reactive oxygen species (ROS) and counteract oxidative stress, a key contributor to various chronic diseases. Numerous in vitro studies have demonstrated quercetin’s effectiveness in scavenging free radicals and protecting cellular structures from oxidative damage. Beyond its direct antioxidant effects, quercetin also interacts with cellular signaling pathways, influencing gene expression and modulating enzymatic activities associated with oxidative stress. In vivo studies, both in animals and human trials, have provided insights into the bioavailability and physiological impact of quercetin, yet its significance remains underappreciated. This chapter will focus on the mechanisms by which quercetin enters circulation, its distribution in tissues, and the subsequent effects on markers of oxidative stress. Additionally, we will highlight findings from previous epidemiological studies linking quercetin-rich diets to reduced risk of chronic diseases, emphasizing the potential translational significance of these antioxidant properties in real-world health outcomes. In conclusion, this chapter will provide an overview of quercetin’s antioxidant properties and its potential for therapeutic interventions associated with chronic diseases.

Keywords

  • antioxidant
  • quercetin
  • flavonoid
  • oxidative stress
  • chronic diseases

1. Introduction

Quercetin’s chemical structure, characterized by the presence of multiple hydroxyl groups, provides it with potent free radical scavenging abilities, making it a key player in cellular defense against oxidative stress [1]. While there have been various reports indicating the potential therapeutic benefits of quercetin, its mechanism of action remains elusive and there is a dearth of clinical data to support its potential antioxidant properties in humans [2]. Quercetin, belonging to the flavonol subclass is widely distributed in nature and can be found in various food sources, including fruits vegetables and beverages. The availability of quercetin in commonly consumed dietary items has contributed to its accessibility and the growing interest in harnessing its potential health benefits. The structural arrangement of quercetin includes several hydroxyls (∙OH) groups attached to different positions on the benzene rings, specifically at C3, C5, C7, C3′, C4′, and C5′ [3]. These hydroxyl groups are responsible for the antioxidant properties of quercetin, making it a potent scavenger of free radicals in the human body [4]. The conjugation of double bonds in the aromatic rings and the presence of hydroxyl groups further create a complex structure that plays a crucial role in quercetin’s bioactivity and physiological effects. Additionally, the presence of a glycosyl group linked to the quercetin aglycone can affect its absorption, solubility, and in vivo effects [5]. Most studies focused on the biological effects of quercetin are based on in vitro research with quercetin aglycone, which is rarely found in human plasma [6, 7]. The findings from these studies have yielded inconclusive results regarding the exact mechanism of action of quercetin aglycone and its derived metabolites in humans [8]. Given the insufficient data in this research area, several authors have emphasized the importance of studying quercetin and their derived metabolites, considering that these metabolites may be the bioactive forms in the body, requiring further research to understand their mechanism of action and antioxidant activities [9, 10, 11]. In this study, our aim is to conduct a thorough investigation of the antioxidant capabilities of quercetin and its derivatives, and their impact on human health. Specifically, we will focus on their cellular signaling pathways, bioavailability, tissue distribution, and their association with reduced risk of chronic diseases.

Advertisement

2. Antioxidant role of quercetin in scavenging free radicals

Free radicals, including reactive oxygen species and reactive nitrogen species (RNS), are highly reactive molecules produced naturally during metabolic processes or by external factors like UV radiation, pollution, and certain chemicals. When present in excess, these free radicals can induce oxidative stress to cellular components such as proteins and lipids, leading to various chronic health conditions including neurodegenerative and cardiovascular diseases. Quercetin exerts its antioxidative effects through multiple mechanisms. Firstly, it acts as a direct scavenger of free radicals, neutralizing them and preventing them from causing harm to cellular structures as shown in Figure 1. Additionally, quercetin enhances the activity of endogenous antioxidant enzymes such as superoxide dismutase (SOD), catalase, and glutathione peroxidase, which further contribute to the cellular defense against oxidative stress.

Figure 1.

The antioxidant activity of quercetin in scavenging free radicals.

Numerous in vitro and in vivo studies have highlighted the ability of quercetin to neutralize oxidative stress, making it a promising candidate for mitigating various chronic diseases associated with oxidative damage [12, 13, 14, 15]. Various studies have evaluated the antioxidant efficacy of quercetin and its derivatives using different assays, including 2,2-diphenyl-1-picrylhydrazyl (DPPH), Ferric Reducing Antioxidant Power (FRAP), Fe2+ chelation, and 2,2′-azino-bis (3-ethylbenzothiazoline-6-sulfonic acid) (ABTS) assays [16, 17]. The antioxidant activity of quercetin derivatives can be influenced by factors such as the acyl group chain length and the molar mass of the polymer. While the acylation of quercetin increases its lipophilicity, this can potentially influence its antioxidant efficacy. According to Oh et al. [18], the esterification of quercetin with various fatty acyl chlorides resulted in the formation of different derivatives, including Q-3′-O-monoester, Q-3-O-monoester, Q-7,3′-O-diester, Q-3′,4’-O- diester, Q-3,3′-O-diester, and Q-3,4′-O diester. The lipophilicity of these derivatives following the esterification process increased as expected. The results showed that quercetin had the highest activity in getting rid of harmful radicals compared to its derivatives. Interestingly, superior antioxidant activity was observed with the short-chain fatty acids derivatives in the ABTS assay, while those with C3:0, C4:0, C6:0, and C8:0 chains showed better performance in the DPPH radical scavenging assay.

In a previous study where the impact of nicotine and quercetin treatments on the viability and antioxidant defense system of HepG2 cell line cultures was investigated, quercetin showed an antioxidant protective effect by restoring the balance between free radical production and antioxidant defense systems [19]. The in vitro experimental results showed that quercetin increased the superoxide dismutase activity in both the control and nicotine treated group. The results were comparable to those from previous studies [20, 21]. Some studies have shown the ability of quercetin to combat oxidative stress by modulating glutathione (GSH) levels. For instance, the supplementation of quercetin in a Western diet increased the GSH/GSSG ratio in mice liver and mitigated obesity and metabolic syndrome [15]. The concentration of quercetin did not significantly influence body weight, fat storage, or blood composition in healthy mice following a standard diet. However, it did decrease oxidative stress levels, specifically in the liver. A higher quercetin concentration was observed to further decrease lipid peroxidation markers in various tissues. Additionally, quercetin’s capacity to neutralize free radicals can be associated with a decrease in ischemia-reperfusion injury, achieved through alterations in the endothelial nitric oxide system [22]. This highlights the potential of quercetin to mitigate oxidative stress and protect against damage caused by ischemia-reperfusion processes. As an extension to this, recent research has emphasized quercetin’s role in enhancing mitochondrial function [23, 24], further contributing to its overall protective effects on chronic and cardiovascular health diseases.

2.1 Absorption, distribution and metabolism of quercetin in tissues

The absorption of quercetin is a complex process influenced by several factors. Dietary intake is the primary source, and its absorption occurs in the small intestine. Quercetin, in its glycosidic form, undergoes hydrolysis by enzymes in the small intestine, releasing aglycones that are absorbed through passive diffusion [25]. Glucose transporters (GLUTs) and multidrug resistance-associated proteins (MRPs) are involved in the active transport of quercetin [26]. The release of aglycones from quercetin glycosides can be mediated by the lactase phlorizin hydrolase (LPH) in the small intestines [25]. Various factors impact the absorption of quercetin, including the food matrix, co-ingestion with other nutrients, and the gut microbiota. The presence of fats and certain nutrients enhances absorption, while high-fiber diets may reduce bioavailability. The bioavailability of quercetin can be improved when consumed alongside fatty acids in higher doses. Humans can absorb significant amounts from food or supplements, with a reported half-life of 11–28 hours. While daily quercetin intakes range from 3 to 40 mg in Western diets, consumers of fruits and vegetables may ingest about 250 mg per day. The recommended daily dose of quercetin supplements in diets should range from 500 to 1000 mg [27]. Additionally, the gut microbiota plays a crucial role in quercetin metabolism, influencing its absorption and bioactivity [28].

Once absorbed, quercetin is distributed to various tissues through the bloodstream. The distribution is influenced by its lipophilic nature, allowing it to penetrate cell membranes. On the other hand, quercetin exhibits a wide tissue distribution, with higher concentrations found in organs like the lungs, kidneys, and liver. The ability of quercetin to cross the blood-brain barrier has also raised interest in its potential neuroprotective effects [29, 30]. In the bloodstream and tissues, it undergoes extensive metabolism and biotransformation with phase II enzymes, such as glucuronosyltransferases and sulfotransferases, facilitating its excretion [31]. Special transporters also play a crucial role in facilitating the absorption and transport of quercetin and its metabolites. The lipid bilayers of cellular membranes, characterized by high polarity, limits the passage of quercetin conjugates where membrane-related transporters come to play to overcome this barrier. In the intestinal epithelial cells, numerous transporters are expressed to aid the passage of substrates from the gastrointestinal tract into the circulatory system [32]. Among these transporters, sodium-dependent glucose co-transporters (SGLTs) located on the apical membrane of intestinal epithelial cells play a significant role in the absorption of quercetin and its glycosides [33]. Specifically, quercetin 4′-β-glucoside and quercetin-3-glucoside are efficiently absorbed through the involvement of SGLT-1. There is evidence that the type of sugar linked to quercetin can influence the absorption ratio of quercetin glycosides in the small intestine [34]. Quercetin circulates in the bloodstream bound to plasma proteins, such as albumin [35]. The presence of binding proteins can influence its bioavailability and may impact its physiological effects. Understanding the binding characteristics of quercetin to plasma proteins is important for optimizing its delivery and enhancing its therapeutic potential. Further research is also needed to elucidate the specific mechanisms by which binding proteins interact with quercetin and how the interaction influences its bioavailability and physiological effects [36]. Furthermore, exploring strategies to enhance the bioavailability of quercetin, such as encapsulation in delivery systems like Quercetin LipoMicel, may improve its absorption and increase its circulating levels in the body [37].

2.2 Cellular signaling pathways of quercetin: modulation of enzymatic activities in oxidative stress

The pathway associated with the nuclear factor erythroid 2-related factor 2 (Nrf2) is vital for protecting cells from oxidative stress. Quercetin has been shown to activate the Nrf2 pathway, leading to the upregulation of antioxidant response element (ARE)-driven genes. This activation enhances the cellular antioxidant defense system, including the expression of enzymes like heme oxygenase-1 (HO-1) and superoxide dismutase. Quercetin and its derivative dihydroquercetin have demonstrated the ability to mitigate cellular damage induced by oxidative stress by activating the Nrf2-ARE pathway, with the involvement of stress-responsive Extracellular Signal-Regulated Kinase (ERK) and c-Jun N-terminal Kinase (JNK) signaling pathways, suggesting a potential neurohormetic role for quercetin as a phytochemical [14, 38, 39].

For instance, Jia et al. [40] examined the antioxidant effects of quercetin on diquat-induced oxidative stress in porcine enterocytes. Pretreatment with quercetin in intestinal porcine epithelial cell line 1 (IPEC-1) cells demonstrated protective effects by mitigating apoptosis through a caspase-3-dependent mechanism, reducing ROS production, maintaining mitochondrial function, and preserving tight junction proteins. The authors proposed that quercetin’s protective role may be linked to the upregulation of Nrf2 protein levels and the modulation of intracellular glutathione content, highlighting its potential in regulating redox homeostasis. In a similar scenario, the activity of quercetin on p38-MAPK and its regulatory effects on the nuclear transcription factor erythroid-2p45-Nrf2 and the GSH antioxidant defense system in HepG2 cells was studied. The study showed a concentration-dependent modulation of p38 and Nrf2 over varying incubation periods, with 50 μM quercetin concentration [14]. In 2014, Saw and colleagues explored the cancer preventive properties of three flavonoids—quercetin, kaempferol, and pterostilbene—found in berries. The research focused on their ability to counteract free radicals using the DPPH and 2′,7′-Dichlorofluorescin diacetate (DCFH-DA) assays. The combination of all three flavonoids showed a synergistic effect in reducing intracellular ROS levels by stimulating the activation of ARE and the levels of mRNA and proteins associated with genes regulated by Nrf2 [38].

There is also evidence that quercetin can modulate enzymatic activities associated with oxidative stress [41, 42, 43]. One example is the inhibition of enzymes involved in ROS generation, such as NADPH oxidase [44, 45, 46]. By downregulating these enzymes, quercetin contributes to the reduction of intracellular ROS levels, thereby mitigating oxidative stress. In addition to that, quercetin interacts with key cellular signaling pathways, including mitogen-activated protein kinase (MAPK) and nuclear factor-kappa B (NF-κB) [47, 48, 49]. By inhibiting these pathways, quercetin plays a role in attenuating the expression of pro-inflammatory and pro-oxidative genes, contributing to its overall protective effects against oxidative stress-related damage. Furthermore, it modulates enzymatic activities involved in inflammatory pathways, including the inhibition of cyclooxygenase-2 (COX-2) and inducible nitric oxide synthase (iNOS) [50, 51, 52]. Quercetin was shown to markedly reduce the mRNA expression of inflammatory markers such as iNOS, COX-2, and C-reactive protein (CRP) in Chang Liver cells, demonstrating its anti-inflammatory effects [50].

2.3 Protective role of quercetin on chronic disease risk infection

Quercetin exerts its protective effects through diverse mechanisms. Its ability to neutralize ROS and free radicals arises from the presence of hydroxyl (∙OH) groups and a catechol-type B-ring in its molecular structure. Additionally, quercetin activates the Nrf2 pathway, promoting the expression of antioxidant enzymes and cellular defense mechanisms. The interplay of these mechanism positions quercetin as a potent defender against oxidative stress-induced damage. Quercetin, known for its potent antioxidant properties, exhibits cardiovascular protection by reducing oxidative damage to endothelial cells, lowering blood pressure, and improving lipid profiles. Oxidative stress arises from an imbalance between ROS production and the body’s antioxidant defense mechanisms [53, 54]. This imbalance is associated with the onset and progression of several chronic conditions such as myocardial infarction, chronic inflammation, aging, and neurodegenerative disorders [55, 56]. Numerous studies have demonstrated that quercetin exhibits neuroprotective effects and counteracts oxidative stress in vivo. For instance, its neuroprotective effect was shown in a study by Denny et al. [57] who found that the oral administration of quercetin and fish oil supplement enhanced neuroprotection in rats subjected to chronic exposure to the insecticide rotenone, which serves as an animal model for Parkinson’s disease. Since the brain is particularly susceptible to oxidative stress, and a target for the damaging effects of ROS, the antioxidant effects of quercetin against oxidative stress pathways and its potential in preventing brain health diseases such as Alzheimer’s disease (AD) through various pathways, including Nrf2, Paraoxonase-2, JNK, Protein kinase C, MAPK was discussed by Saikia et al. [58]. The study discussed that quercetin activates the Nrf2 pathway by increasing the production of protective proteins in AD brain cells, acting as a natural defense system against oxidative stress. Consequently, the neuroprotective potential of quercetin against oxidative stress responses was demonstrated by its interaction with paraoxonase-2-enzyme, JNK pathways and protein kinase C pathways which plays a role in cell signaling and function. Persistent activation of glutamate receptors, leading to excitotoxicity, contributes to various neurological disorders including Alzheimer’s disease, hypoxic-ischemic brain injury, multiple sclerosis etc. [59]. However, quercetin extracts protect neuronal cells from excitotoxicity-induced damage through mechanisms involving reduced ROS production, preservation of mitochondrial membrane potential, and modulation of multiple biochemical markers associated with cell death and autophagy [60]. Quercetin has emerged as a significant component demonstrating acetylcholinesterase (AChE) inhibitory activity and exerts a positive influence on the expression of nicotinic receptors, thereby augmenting cognitive memory function in Alzheimer’s patients [61, 62, 63].

Chronic inflammation is a prevalent contributing factor in various diseases, and quercetin’s anti-inflammatory properties have been studied in the context of conditions such as rheumatoid arthritis and inflammatory bowel diseases [64]. Previous studies have suggested that quercetin-rich diets may be associated with a decreased risk of developing these inflammatory conditions, providing insights into potential dietary strategies for therapeutic interventions [65, 66, 67]. Its anti-inflammatory properties have been linked to potential cardiovascular benefits. While many studies suggest potential cardiovascular benefits of quercetin, it is important to note that research in this area is ongoing, and not all studies have produced consistent results. Factors such as dosage, bioavailability, and individual variability may influence quercetin’s effects on cardiovascular health.

Excessive generation of reactive oxygen species resulting in oxidative stress has been recognized as a significant factor contributing to endothelial dysfunction-induced hypertension and various cardiovascular diseases [68]. Consequently, mitigating oxidative stress is considered a practical approach for comprehensive management of hypertension and other cardiovascular conditions. Several in vivo studies in animal models that explored the cardioprotective effects of quercetin and demonstrated its ability to lower blood pressure including those with hypertension, high-fat high-sucrose diet-induced conditions, nitric oxide deficiency, angiotensin infusion, and aortic constriction [69, 70, 71, 72]. In a 4-week trial involving eighteen Dahl salt-sensitive rats, the antihypertensive effects of captopril (CAP) and quercetin (QUE) on the renin-angiotensin-aldosterone system were explored with a specific focus on renal effects. While the results from this study indicated that quercetin does not exhibit a significant difference in aldosterone levels, it effectively lowered blood pressure in the hypertensive rat model, suggesting a potential modulation of renal function as the underlying mechanism [72]. Hackl et al. [73] investigated the impact of quercetin on angiotensin-converting enzyme (ACE) activity by assessing cardiovascular responses to bradykinin and angiotensin I. Quercetin pretreatment, administered orally (88.7 umol/kg, 45 min) and intravenously (14.7 umol/kg, 5 min), significantly enhanced the hypotensive effect of bradykinin (10 nmol/kg, i.v.). This study indicated that quercetin possess an ACE inhibitory activity in vitro similar to captopril, demonstrating its potential antihypertensive properties.

In contrast to findings in animal studies, human research trials have not conclusively demonstrated consistent results regarding the antioxidant potential of quercetin, even at elevated dosages. Earlier studies by Egert et al. [8] and others [74, 75, 76] demonstrated no impact on plasma oxidized low-density lipoprotein in healthy individuals or those with pre-hypertension and stage 1 hypertension after quercetin supplementation. The discrepancy between the antioxidant effects observed in hypertensive animal models and the equivocal findings in humans may be attributed to variations in quercetin doses. However, Hertog and his group [77] provided promising results regarding the epidemiological and in vitro/in vivo antioxidant effects of flavonoids, supporting their cardioprotective function. Their study provided evidence supporting the strong cardioprotective effects of various flavonoids, including quercetin. The study showed that men who consumed over 29 mg of flavonols per day experienced a substantial 68% reduction in the risk of coronary death compared to those consuming less than 10 mg per day. Although the study did not specifically explore the link between quercetin intake and blood pressure, the authors noted an inverse relationship between high-quercetin foods and blood pressure, and flavonol intake, including quercetin. While these studies conducted in animal models have indicated a decrease in blood pressure and enhanced antioxidant status with quercetin administration, the evidence from human trials has not been particularly compelling. Human studies examining the effects of quercetin on blood pressure and antioxidant status have not yielded convincing results, indicating the need for additional investigation and exploration of potential factors that influence the outcomes in human subjects. The translation of promising animal findings to human contexts requires careful consideration of various variables that may contribute to the differences observed in the effects of quercetin on blood pressure and antioxidant status.

Quercetin’s ability to scavenge free radicals and modulate inflammatory pathways suggests that it may contribute to reducing oxidative stress and inflammation associated with diabetes, thereby potentially alleviating the risk of diabetic complications. Quercetin’s potential protective effects on pancreatic beta-cells, responsible for insulin production, have been extensively investigated [78]. Various studies have shown that quercetin may protect beta-cells from oxidative stress and apoptosis, thereby preserving their function. Furthermore, quercetin may stimulate insulin secretion, contributing to better glycemic control. Additionally, it may inhibit key enzymes involved in carbohydrate digestion, leading to a more gradual release of glucose into the bloodstream. Various in vitro studies have focused on the antidiabetic effects of quercetin in cellular and animal models. For example, the impact of quercetin on glucose or glibenclamide-induced insulin secretion and its ability to protect against hydrogen peroxide-induced beta-cell dysfunctions, using the INS-1 beta-cell line was examined. The researchers observed that quercetin enhances insulin secretion, protects beta-cells from oxidative damage, and implicates the extracellular signal-regulated kinase (ERK)1/2 pathway [79]. Additionally, quercetin exhibited concentration-dependent inhibition of alpha-amylase and alpha-glucosidase activities. These findings suggest that quercetin’s ability to regulate blood glucose levels may be attributed to its disruption of enzymatic processes and protection of pancreatic tissues from oxidative damage, as evidenced by the inhibition of carbohydrate-metabolizing enzymes and prevention of pancreatic lipid peroxidation [80]. Moreover, potential interactions with medications and variations in individual responses should be considered when assessing the overall safety profile of quercetin for individuals with diabetes.

Advertisement

3. Conclusions

In conclusion, quercetin, with its powerful antioxidant properties, is crucial for defending cells against oxidative stress. While much research has focused on its potential therapeutic benefits, clinical evidence supporting its antioxidant effects in humans is limited. Understanding the mechanisms of action of quercetin and its derivatives is crucial, especially considering that these metabolites may be the active forms in the body. Despite challenges in absorption and bioavailability, quercetin’s ability to mitigate oxidative damage holds promise for combating various chronic diseases. Additionally, its activation of the Nrf2 pathway and interaction with signaling pathways further underscores its potential as a neuroprotective agent. Further research is needed to fully elucidate the mechanisms underlying quercetin’s antioxidant activities including clinical trials to fully elucidate the therapeutic potential and optimal dosage of quercetin for specific health conditions.

References

  1. 1. Srimathi Priyanga K, Vijayalakshmi K. Investigation of antioxidant potential of quercetin and hesperidin: An in vitro approach. Asian Journal of Pharmaceutical and Clinical Research. 2017;10:83-86. DOI: 10.22159/AJPCR.2017.V10I11.20260
  2. 2. Hagde P, Pingle P, Mourya A, Katta CB, Srivastava S, Sharma R, et al. Therapeutic potential of quercetin in diabetic foot ulcer: Mechanistic insight, challenges, nanotechnology driven strategies and future prospects. Journal of Drug Delivery Science and Technology. 2022;74:103575. DOI: 10.1016/J.JDDST.2022.103575
  3. 3. Li Y, Yao J, Han C, Yang J, Chaudhry MT, Wang S, et al. Quercetin, inflammation and immunity. Nutrients. 2016;8:167. DOI: 10.3390/NU8030167
  4. 4. Ozgen S, Kilinc OK, Selamoğlu Z. Antioxidant activity of quercetin: A mechanistic review. Turkish Journal of Agriculture-Food Science and Technology. 2016;4:1134-1138
  5. 5. Wang W, Sun C, Mao L, Ma P, Liu F, Yang J, et al. The biological activities, chemical stability, metabolism and delivery systems of quercetin: A review. Trends in Food Science and Technology. 2016;56:21-38. DOI: 10.1016/J.TIFS.2016.07.004
  6. 6. Kroon PA, Clifford MN, Crozier A, Day AJ, Donovan JL, Manach C, et al. How should we assess the effects of exposure to dietary polyphenols in vitro? The American Journal of Clinical Nutrition. 2004;80:15-21. DOI: 10.1093/AJCN/80.1.15
  7. 7. Mullen W, Edwards CA, Crozier A. Absorption, excretion and metabolite profiling of methyl-, glucuronyl-, glucosyl- and sulpho-conjugates of quercetin in human plasma and urine after ingestion of onions. British Journal of Nutrition. 2006;96:107. DOI: 10.1079/BJN20061809
  8. 8. Egert S, Wolffram S, Bosy-Westphal A, Boesch-Saadatmandi C, Wagner AE, Frank J, et al. Daily quercetin supplementation dose-dependently increases plasma quercetin concentrations in healthy humans. The Journal of Nutrition. 2008;138:1615-1621. DOI: 10.1093/JN/138.9.1615
  9. 9. Wiczkowski W, Szawara-Nowak D, Topolska J, Olejarz K, Zieliński H, Piskuła MK. Metabolites of dietary quercetin: Profile, isolation, identification, and antioxidant capacity. Journal of Functional Foods. 2014;11:121-129. DOI: 10.1016/J.JFF.2014.09.013
  10. 10. Santos MR, Rodríguez-Gómez MJ, Justino GC, Charro N, Florencio MH, Mira L. Influence of the metabolic profile on the in vivo antioxidant activity of quercetin under a low dosage oral regimen in rats. British Journal of Pharmacology. 2008;153:1750. DOI: 10.1038/BJP.2008.46
  11. 11. Cho JM, Chang SY, Kim DB, Needs PW, Jo YH, Kim MJ. Effects of physiological quercetin metabolites on interleukin-1β-induced inducible NOS expression. The Journal of Nutritional Biochemistry. 2012;23:1394-1402. DOI: 10.1016/J.JNUTBIO.2011.08.007
  12. 12. Ademosun AO, Oboh G, Bello F, Ayeni PO. Antioxidative properties and effect of quercetin and its glycosylated form (Rutin) on acetylcholinesterase and butyrylcholinesterase activities. Evidence-based Complementary and Alternative Medicine. 2016;21:NP11-NP17. DOI: 10.1177/2156587215610032/ASSET/IMAGES/LARGE/10.1177_2156587215610032-FIG7.JPEG
  13. 13. Braun KF, Ehnert S, Freude T, Egaña JT, Schenck TL, Buchholz A, et al. Quercetin protects primary human osteoblasts exposed to cigarette smoke through activation of the antioxidative enzymes HO-1 and SOD-1. Research article. The Scientific World Journal. 2011;11:2348-2357. DOI: 10.1100/2011/471426
  14. 14. Granado-Serrano AB, Martín MA, Bravo L, Goya L, Ramos S. Quercetin modulates Nrf2 and glutathione-related defenses in HepG2 cells: Involvement of P38. Chemico-Biological Interactions. 2012;195:154-164. DOI: 10.1016/J.CBI.2011.12.005
  15. 15. Kobori M, Takahashi Y, Akimoto Y, Sakurai M, Matsunaga I, Nishimuro H, et al. Chronic high intake of quercetin reduces oxidative stress and induces expression of the antioxidant enzymes in the liver and visceral adipose tissues in mice. Journal of Functional Foods. 2015;15:551-560. DOI: 10.1016/J.JFF.2015.04.006
  16. 16. Murugesan N, Damodaran C, Krishnamoorthy S, Raja M. In-vitro evaluation of synergism in antioxidant efficiency of quercetin and resveratrol. Chemical Biology Letters. 2023;10:534-534
  17. 17. Zizkova P, Stefek M, Rackova L, Prnova M, Horakova L. Novel quercetin derivatives: From redox properties to promising treatment of oxidative stress related diseases. Chemico-Biological Interactions. 2017;265:36-46. DOI: 10.1016/J.CBI.2017.01.019
  18. 18. Oh WY, Ambigaipalan P, Shahidi F. Preparation of quercetin esters and their antioxidant activity. Journal of Agricultural and Food Chemistry. 2019;67:10653-10659. DOI: 10.1021/ACS.JAFC.9B04154
  19. 19. Yarahmadi A, Zal F, Bolouki A. Protective effects of quercetin on nicotine induced oxidative stress in ‘HepG2 cells’. Toxicology Mechanisms and Methods. 2017;27:609-614. DOI: 10.1080/15376516.2017.1344338
  20. 20. Lee YJ, Beak SY, Choi I, Sung JS. Quercetin and its metabolites protect hepatocytes against ethanol-induced oxidative stress by activation of Nrf2 and AP-1. Food Science and Biotechnology. 2018;27:809. DOI: 10.1007/S10068-017-0287-8
  21. 21. Muthukumaran S, Sudheer AR, Menon VP, Nalini N. Protective effect of quercetin on nicotine-induced prooxidant and antioxidant imbalance and DNA damage in Wistar rats. Toxicology. 2008;243:207-215. DOI: 10.1016/J.TOX.2007.10.006
  22. 22. Shoskes D, Lapierre C, Cruz-Corerra M, Muruve N, Rosario R, Fromkin B, et al. Beneficial effects of the bioflavonoids curcumin and quercetin on early function in cadaveric renal transplantation: A randomized placebo controlled trial. Transplantation. 2005;80:1556-1559. DOI: 10.1097/01.TP.0000183290.64309.21
  23. 23. Qiu L, Luo Y, Chen X. Quercetin attenuates mitochondrial dysfunction and biogenesis via upregulated AMPK/SIRT1 signaling pathway in OA rats. Biomedicine & Pharmacotherapy. 2018;103:1585-1591. DOI: 10.1016/J.BIOPHA.2018.05.003
  24. 24. Rayamajhi N, Kim SK, Go H, Joe Y, Callaway Z, Kang JG, et al. Quercetin induces mitochondrial biogenesis through activation of HO-1 in HepG2 cells. Oxidative Medicine and Cellular Longevity. 2013;2013:2-10. DOI: 10.1155/2013/154279
  25. 25. Chen L, Cao H, Huang Q , Xiao J, Teng H. Absorption, metabolism and bioavailability of flavonoids: A review. Critical Reviews in Food Science and Nutrition. 2022;62:7730-7742. DOI: 10.1080/10408398.2021.1917508
  26. 26. Wenzel U. Flavonoids as drugs at the small intestinal level. Current Opinion in Pharmacology. 2013;13:864-868. DOI: 10.1016/J.COPH.2013.08.015
  27. 27. Andres S, Pevny S, Ziegenhagen R, Bakhiya N, Schäfer B, Hirsch-Ernst KI, et al. Safety aspects of the use of quercetin as a dietary supplement. Molecular Nutrition & Food Research. 2018;62:2-17. DOI: 10.1002/MNFR.201700447
  28. 28. Murota K, Nakamura Y, Uehara M. Flavonoid metabolism: The interaction of metabolites and gut microbiota. Bioscience, Biotechnology, and Biochemistry. 2018;82:600-610. DOI: 10.1080/09168451.2018.1444467
  29. 29. Grabowska D, Galanty K, Sobolewska A, Podolak D, Saldanha L, Fernanda De Campos Fraga-Silva T, et al. A flavonoid on the brain: Quercetin as a potential therapeutic agent in central nervous system disorders. Life. 2022;12:591. DOI: 10.3390/LIFE12040591.
  30. 30. Grewal AK, Singh TG, Sharma D, Sharma V, Singh M, Rahman MH, et al. Mechanistic insights and perspectives involved in neuroprotective action of quercetin. Biomedicine & Pharmacotherapy. 2021;140:111729. DOI: 10.1016/J.BIOPHA.2021.111729
  31. 31. Jiang W, Hu M. Mutual interactions between flavonoids and enzymatic and transporter elements responsible for flavonoid disposition via phase II metabolic pathways. RSC Advances. 2012;2:7948-7963. DOI: 10.1039/C2RA01369J
  32. 32. Han F, Yang P, Wang H, Fernandes I, Mateus N, Liu Y. Digestion and absorption of red grape and wine anthocyanins through the gastrointestinal tract. Trends in Food Science and Technology. 2019;83:211-224. DOI: 10.1016/J.TIFS.2018.11.025
  33. 33. Röder PV, Geillinger KE, Zietek TS, Thorens B, Koepsell H, Daniel H. The role of SGLT1 and GLUT2 in intestinal glucose transport and sensing. PLoS One. 2014;9:2-10. DOI: 10.1371/JOURNAL.PONE.0089977
  34. 34. Arts ICW, Sesink ALA, Faassen-Peters M, Hollman PCH. The type of sugar moiety is a major determinant of the small intestinal uptake and subsequent biliary excretion of dietary quercetin glycosides. The British Journal of Nutrition. 2004;91:841-847. DOI: 10.1079/BJN20041123
  35. 35. Aghababaei F, Hadidi M. Recent advances in potential health benefits of quercetin. Pharmaceuticals. 2023;16:2-31. DOI: 10.3390/PH16071020
  36. 36. Papakyriakopoulou P, Velidakis N, Khattab E, Valsami G, Korakianitis I, Kadoglou NPE. Potential pharmaceutical applications of quercetin in cardiovascular diseases. Pharmaceuticals. 2022;15:1019-1019. DOI: 10.3390/PH15081019
  37. 37. Solnier J, Chang C, Roh K, Du M, Kuo YC, Hardy M, et al. Quercetin LipoMicel—A novel delivery system to enhance bioavailability of quercetin. Journal of Natural Health Product Research. 2021;3:1-8. DOI: 10.33211/JNHPR.17
  38. 38. Saw CLL, Guo Y, Yang AY, Paredes-Gonzalez X, Ramirez C, Pung D, et al. The berry constituents quercetin, kaempferol, and pterostilbene synergistically attenuate reactive oxygen species: Involvement of the Nrf2-ARE signaling pathway. Food and Chemical Toxicology. 2014;72:303-311. DOI: 10.1016/J.FCT.2014.07.038
  39. 39. Liang L, Gao C, Luo M, Wang W, Zhao C, Zu Y, et al. Dihydroquercetin (DHQ) induced HO-1 and NQO1 expression against oxidative stress through the Nrf2-dependent antioxidant pathway. Journal of Agricultural and Food Chemistry. 2013;61:2755-2761. DOI: 10.1021/JF304768P/ASSET/IMAGES/LARGE/JF-2012-04768P_0006.JPEG
  40. 40. Jia H, Zhang Y, Si X, Jin Y, Jiang D, Dai Z, et al. Quercetin alleviates oxidative damage by activating nuclear factor erythroid 2-related factor 2 signaling in porcine enterocytes. Nutrients. 2021;13:1-15. DOI: 10.3390/NU13020375
  41. 41. Costa LG, Garrick JM, Roquè PJ, Pellacani C. Mechanisms of neuroprotection by quercetin: Counteracting oxidative stress and more. Oxidative Medicine and Cellular Longevity. 2016;2016:1-10. DOI: 10.1155/2016/2986796
  42. 42. Ramyaa P, Krishnaswamy R, Padma VV. Quercetin modulates OTA-induced oxidative stress and redox signalling in HepG2 cells — Up regulation of Nrf2 expression and down regulation of NF-ΚB and COX-2. Biochimica et Biophysica Acta (BBA) - General Subjects. 2014;1840:681-692. DOI: 10.1016/J.BBAGEN.2013.10.024
  43. 43. Alía M, Ramos S, Mateos R, Granado-Serrano AB, Bravo L, Goya L. Quercetin protects human hepatoma HepG2 against oxidative stress induced by tert-butyl hydroperoxide. Toxicology and Applied Pharmacology. 2006;212:110-118. DOI: 10.1016/J.TAAP.2005.07.014
  44. 44. Baba RA, Mir HA, Mokhdomi TA, Bhat HF, Ahmad A, Khanday FA. Quercetin suppresses ROS production and migration by specifically targeting Rac1 activation in gliomas. Frontiers in Pharmacology. 2024;15:1318797. DOI: 10.3389/FPHAR.2024.1318797
  45. 45. Sul OJ, Ra SW. Quercetin prevents LPS-induced oxidative stress and inflammation by modulating NOX2/ROS/NF-KB in lung epithelial cells. Molecules. 2021;22:1-11. DOI: 10.3390/MOLECULES26226949
  46. 46. Luo M, Tian R, Yang Z, Peng YY, Lu N. Quercetin suppressed NADPH oxidase-derived oxidative stress via heme oxygenase-1 induction in macrophages. Archives of Biochemistry and Biophysics. 2019;671:69-76. DOI: 10.1016/J.ABB.2019.06.007
  47. 47. Devi KP, Kiruthiga P, Pandian SK. Emerging role of flavonoids in inhibition of NF-B-mediated signaling pathway: A review. International Journal of Biomedical and Pharmaceutical Sciences. 2009;3:31-45
  48. 48. Endale M, Park SC, Kim S, Kim SH, Yang Y, Cho JY, et al. Quercetin disrupts tyrosine-phosphorylated phosphatidylinositol 3-kinase and myeloid differentiation factor-88 association, and inhibits MAPK/AP-1 and IKK/NF-ΚB-induced inflammatory mediators production in RAW 264.7 cells. Immunobiology. 2013;218:1452-1467. DOI: 10.1016/J.IMBIO.2013.04.019
  49. 49. Ruiz PA, Braune A, Hölzlwimmer G, Quintanilla-Fend L, Haller D. Quercetin inhibits TNF-induced NF-KappaB transcription factor recruitment to proinflammatory gene promoters in murine intestinal epithelial cells. The Journal of Nutrition. 2007;137:1208-1215. DOI: 10.1093/JN/137.5.1208
  50. 50. García-Mediavilla V, Crespo I, Collado PS, Esteller A, Sánchez-Campos S, Tuñón MJ, et al. The anti-inflammatory flavones quercetin and kaempferol cause inhibition of inducible nitric oxide synthase, cyclooxygenase-2 and reactive C-protein, and down-regulation of the nuclear factor KappaB pathway in Chang liver cells. European Journal of Pharmacology. 2007;557:221-229. DOI: 10.1016/J.EJPHAR.2006.11.014
  51. 51. Chen YC, Shen SC, Lee WR, Hou WC, Yang LL, Lee TJF. Inhibition of nitric oxide synthase inhibitors and lipopolysaccharide induced inducible NOS and cyclooxygenase-2 gene expressions by Rutin, quercetin, and quercetin pentaacetate in RAW 264.7 macrophages. Journal of Cellular Biochemistry. 2001;82:537-548. DOI: 10.1002/JCB.1184
  52. 52. Raso GM, Meli R, Di Carlo G, Pacilio M, Di Carlo R. Inhibition of inducible nitric oxide synthase and cyclooxygenase-2 expression by flavonoids in macrophage J774A.1. Life Sciences. 2001;68:921-931. DOI: 10.1016/S0024-3205(00)00999-1
  53. 53. Dong YS, Wang JL, Feng DY, Qin HZ, Wen H, Yin ZM, et al. Protective effect of quercetin against oxidative stress and brain edema in an experimental rat model of subarachnoid hemorrhage. International Journal of Medical Sciences. 2014;11:282. DOI: 10.7150/IJMS.7634
  54. 54. Betteridge DJ. What is oxidative stress? Metabolism. 2000;49:3-8. DOI: 10.1016/S0026-0495(00)80077-3
  55. 55. Uttara B, Singh AV, Zamboni P, Mahajan RT. Oxidative stress and neurodegenerative diseases: A review of upstream and downstream antioxidant therapeutic options. Current Neuropharmacology. 2009;7:65. DOI: 10.2174/157015909787602823
  56. 56. Fridovich I. Fundamental aspects of reactive oxygen species, or what’s the matter with oxygen? Annals of the New York Academy of Sciences. 1999;893:13-18. DOI: 10.1111/J.1749-6632. 1999.TB07814.X
  57. 57. Denny Joseph KM, Muralidhara. Combined oral supplementation of fish oil and quercetin enhances neuroprotection in a chronic rotenone rat model: Relevance to Parkinson’s disease. Neurochemical Research. 2015;40:894-905. DOI: 10.1007/S11064-015-1542-0
  58. 58. Saikia S. Ascertaining the antioxidant properties of quercetin against oxidative stress in combating Alzheimer’s disease: A review. Uttar Pradesh Journal of Zoology. 2022;43:680-694. DOI: 10.56557/UPJOZ/2022/V43I243385
  59. 59. Akyuz E, Paudel YN, Polat AK, Dundar HE, Angelopoulou E. Enlightening the neuroprotective effect of quercetin in epilepsy: From mechanism to therapeutic opportunities. Epilepsy & Behavior. 2021;115:1525-5050. DOI: 10.1016/J.YEBEH.2020.107701
  60. 60. Silva B, Oliveira PJ, Dias A, Malva JO. Quercetin, kaempferol and Biapigenin from hypericum Perforatum are neuroprotective against excitotoxic insults. Neurotoxicity Research. 2008;13:265-279. DOI: 10.1007/BF03033510
  61. 61. Parsons CG, Danysz W, Dekundy A, Pulte I. Memantine and cholinesterase inhibitors: Complementary mechanisms in the treatment of Alzheimer’s disease. Neurotoxicity Research. 2013;24:358. DOI: 10.1007/S12640-013-9398-Z
  62. 62. Islam MR, Zaman A, Jahan I, Chakravorty R, Chakraborty S. In silico QSAR analysis of quercetin reveals its potential as therapeutic drug for Alzheimer’s disease. Journal of Young Pharmacists. 2013;5:173. DOI: 10.1016/J.JYP.2013.11.005
  63. 63. Jung M, Park M. Acetylcholinesterase inhibition by flavonoids from Agrimonia Pilosa. Molecules: A Journal of Synthetic Chemistry and Natural Product Chemistry. 2007;12:2130. DOI: 10.3390/12092130
  64. 64. El-Said KS, Atta A, Mobasher MA, Germoush MO, Mohamed TM, Salem MM. Quercetin mitigates rheumatoid arthritis by inhibiting adenosine deaminase in rats. Molecular Medicine. 2022;28:1-13. DOI: 10.1186/S10020-022-00432-5/FIGURES/11
  65. 65. Habtemariam S, Belai A. Natural therapies of the inflammatory bowel disease: The case of Rutin and its aglycone, quercetin. Mini-reviews in Medicinal Chemistry. 2017;18(3):234-243. DOI: 10.2174/1389557517666170120152417
  66. 66. Comalada M, Camuesco D, Sierra S, Ballester I, Xaus J, Gálvez J, et al. In vivo quercitrin anti-inflammatory effect involves release of quercetin, which inhibits inflammation through down-regulation of the NF-ΚB pathway. European Journal of Immunology. 2005;35:584-592. DOI: 10.1002/EJI.200425778
  67. 67. Kleemann R, Verschuren L, Morrison M, Zadelaar S, van Erk MJ, Wielinga PY, et al. Anti-inflammatory, anti-proliferative and anti-atherosclerotic effects of quercetin in human in vitro and in vivo models. Atherosclerosis. 2011;218:44-52. DOI: 10.1016/J.ATHEROSCLEROSIS.2011.04.023
  68. 68. Schiffrin EL. Antioxidants in hypertension and cardiovascular disease. Molecular Interventions. 2010;10:354-362. DOI: 10.1124/MI.10.6.4
  69. 69. Yamamoto Y, Oue E. Antihypertensive effect of quercetin in rats fed with a high-fat high-sucrose diet. Bioscience, Biotechnology, and Biochemistry. 2006;70:933-939. DOI: 10.1271/BBB.70.933
  70. 70. Duarte J, Jiménez R, O’Valle F, Galisteo M, Pérez-Palencia R, Vargas F, et al. Protective effects of the flavonoid quercetin in chronic nitric oxide deficient rats. Journal of Hypertension. 2002;20:1843-1854. DOI: 10.1097/00004872-200209000-00031
  71. 71. Jalili T, Carlstrom J, Kim S, Freeman D, Jin H, Wu TC, et al. Quercetin-supplemented diets lower blood pressure and attenuate cardiac hypertrophy in rats with aortic constriction. Journal of Cardiovascular Pharmacology. 2006;47:531-541. DOI: 10.1097/01.FJC.0000211746.78454.50
  72. 72. Mackraj I, Govender T, Ramesar S. The antihypertensive effects of quercetin in a salt-sensitive model of hypertension. Journal of Cardiovascular Pharmacology. 2008;51:239-245. DOI: 10.1097/FJC.0B013E318162011F
  73. 73. Häckl LPN, Cuttle G, Sanches Dovichi S, Lima-Landman MT, Nicolau M. Inhibition of angiotensin-converting enzyme by quercetin alters the vascular response to bradykinin and angiotensin I. Pharmacology. 2002;65:182-186. DOI: 10.1159/000064341
  74. 74. Larson AJ, Symons JD, Jalili T. Quercetin: A treatment for hypertension?—A review of efficacy and mechanisms. Pharmaceuticals. 2010;3:237. DOI: 10.3390/PH3010237
  75. 75. Conquer JA, Maiani G, Azzini E, Raguzzini A, Holub BJ. Supplementation with quercetin markedly increases plasma quercetin concentration without effect on selected risk factors for heart disease in healthy subjects. The Journal of Nutrition. 1998;128:593-597. DOI: 10.1093/JN/128.3.593
  76. 76. Edwards RL, Lyon T, Litwin SE, Rabovsky A, Symons JD, Jalili T. Quercetin reduces blood pressure in hypertensive subjects. The Journal of Nutrition. 2007;137:2405-2411. DOI: 10.1093/JN/137.11.2405
  77. 77. Hertog MGL, Feskens EJM, Kromhout D, Hertog MGL, Hollman PCH, Hertog MGL, et al. Dietary antioxidant flavonoids and risk of coronary heart disease: The Zutphen elderly study. The Lancet. 1993;342:1007-1011. DOI: 10.1016/0140-6736(93)92876-U
  78. 78. Wang JY, Nie YX, Dong BZ, Cai ZC, Zeng XK, Du L, et al. Quercetin protects islet β-cells from oxidation-induced apoptosis via Sirt3 in T2DM. Iranian Journal of Basic Medical Sciences. 2021;24:629. DOI: 10.22038/IJBMS.2021.52005.11792
  79. 79. Youl E, Bardy G, Magous R, Cros G, Sejalon F, Virsolvy A, et al. Quercetin potentiates insulin secretion and protects INS-1 pancreatic b-cells against oxidative damage via the ERK1/2 pathway. British Journal of Pharmacology. 2010;161(4):799-814. DOI: 10.1111/j.1476-5381.2010.00910.x
  80. 80. Oboh G, Ademosun AO, Ayeni PO, Omojokun OS, Bello F. Comparative effect of quercetin and Rutin on α-amylase, α-glucosidase, and some pro-oxidant-induced lipid peroxidation in rat pancreas. Comparative Clinical Pathology. 2015;24:1103-1110. DOI: 10.1007/S00580-014-2040-5

Written By

Kate Nyarko

Submitted: 13 February 2024 Reviewed: 14 February 2024 Published: 04 September 2024