Phytochemical Properties, Extraction, and Pharmacological Benefits of Naringin: History
Please note this is an old version of this entry, which may differ significantly from the current revision.

Naringin is a nutritional flavanone glycoside that has been shown to be effective in the treatment of a few chronic disorders associated with ageing. Citrus fruits contain a common flavone glycoside that has specific pharmacological and biological properties. Naringin, a flavone glycoside with a range of intriguing characteristics, is abundant in citrus fruits. Naringin has been shown to have a variety of biological, medicinal, and pharmacological effects. Naringin is hydrolyzed into rhamnose and prunin by the naringinase, which also possesses l-rhamnosidase activity. D-glucosidase subsequently catalyzes the hydrolysis of prunin into glucose and naringenin. Naringin is known for having anti-inflammatory, antioxidant, and tumor-fighting effects. Numerous test animals and cell lines have been used to correlate naringin exposure to asthma, hyperlipidemia, diabetes, cancer, hyperthyroidism, and osteoporosis.

  • naringin
  • flavonoid
  • extraction
  • bioactive potential
  • pharmaceutical

1. Introduction

Numerous phytochemicals, such as flavonoids (such as hesperidin and naringin), limonoids (such as limonin and nomilin), carotenoids (such as beta-carotene and lutein), and vitamin C are abundant in citrus fruits. Citrus fruits’ vivid colors, distinctive flavors, and distinctive scents are all influenced by these phytochemicals. Citrus fruits include a variety of phytochemicals that have many health advantages [1]. They have antioxidant capabilities that assist the body in fighting off dangerous free radicals and guarding against oxidative stress and cellular damage. Citrus phytochemicals have also been associated with anti-inflammatory effects, which can help reduce the risk of chronic diseases like cardiovascular disease and certain types of cancer. Citrus fruit polyphenols have also been linked to stronger immune systems, better cardiovascular health, and potential anti-diabetic effects. According to some research, these substances may assist with healthy weight management, lowering cholesterol levels, and lowering blood pressure. Citrus fruits include a wide variety of phytochemicals that are essential for overall health and wellbeing, thus including them in the diet is crucial. Beyond what can be achieved by a single vitamin, these chemicals act synergistically to promote health. Regular citrus fruit consumption can support optimal health and lower the risk of chronic diseases by promoting a balanced and nutrient-rich diet [2].

2. Chemical Composition of Naringin

The flavonoid substance naringin is mostly present in grapefruits and other citrus fruits. In chemical terms, it is a glycoside made up of the disaccharide neohesperidose and the flavone naringenin. The chemical structure of naringin consists of a flavonoid backbone, two phenolic rings, and a heterocyclic pyran ring. Its molecular weight per mole is 580.54 g and its chemical formula is C27H32O14. Pharmaceutical and nutraceutical research is interested in naringin because of its bitter taste and its variety of biological qualities, such as antioxidant, anti-inflammatory, anticancer, and cardioprotective properties [1].

2.1. Significance of Flavonoids

In plants, animals, and microbes, flavonoids have a variety of biological effects. Long known to be synthesized at specific locations in plants, flavonoids are also important for the color and scent of flowers, the ability of fruits to draw pollinators and, as a result, fruit dispersion, the germination of seeds and spores, and the development and growth of seedlings. Plants are protected from various biotic and abiotic challenges by flavonoids, which also serve as special UV filters, allopathic substances, signal molecules, phytoalexins, antimicrobial defensive components, and detoxifying agents. Flavonoids have protective effects against frost drought resistance and hardiness, and they may serve to help plants adapt to heat and tolerate freezing temperatures. There are six types of flavonoids [3]. The major classes of flavonoids, their examples, chemical structures, and main dietary sources are listed in Table 1.
Table 1. The major classes of flavonoids, examples, chemical structures, and main dietary sources.
Flavonoids Examples Chemical Structure with Molar Mass (g/mol) Food Sources Reference
Anthocyanin Cyanidin,
pelargonidin,
peonidin
Molecules 28 05623 i001
Cyanidin (287.24)
Solanum melongena, Rubus fruticosus,
Ribes nigrum, Vaccinium sect. Cyanococcus
[4][5]
Flavan-3-ol Catechin, epicatechin, epigallocatechin Molecules 28 05623 i002
Catechin (290.26)
Green tea, Chocolate, Phaseolus vulgaris L., Prunus avium [5]
Flavanones Hesperidin,
Naringin,
Eriodictyol
Molecules 28 05623 i003
Naringin (580.54)
Orange juice, grapefruit juice, lemon juice  
Flavanones Apigenin, luteolin Molecules 28 05623 i004
Apigenin (270.05)
Petroselinum crispum, Apium graveolens, Capsicum annuum [6]
Flavonols Quercetin,
kaempferol,
myricetin
Molecules 28 05623 i005
Quercetin (302.23)
Allium cepa, Malus domestica, Brassica oleracea var. sabellica, Allium porrum [7]
Isoflavones Genistein, daidzein, glycitein Molecules 28 05623 i006
Genistein (270.24)
Soyflour, soymilk, Glycine max. [8][9]

The two primary groups of phenolic chemicals present in citrus fruits are flavonoids and phenolic acids. Citrus flavonoids have been proven to have anti-cancer, anti-inflammatory, anti-aging, anti-bacterial, hepatoprotective, and cardiovascular protective effects, according to several studies. The primary class of phytochemicals found in citrus fruits, particularly in the pulp, peels, and seeds, are called flavonoids. Flavones, flavanones, and flavonols are the three main categories of citrus flavonoids. Table 2 depicts the classification of citrus flavonoids isolated in citrus species, major fruit sources, C-ring structures, and substitution patterns [10].

Table 2. Flavonoids isolated in citrus sp. their structure, molecular weight, fruit sources, C-ring structure and substitution pattern (FLA: flavanone FLO: flavone FOL: flavonol).
Flavonoid Molecular Weight C-Ring Structure Fruit Sources Substitution Pattern Reference
Naringin 580.541 g/mol FLA
FLA
Citrus paradisi
Citrus aurantium
5,4′-OH
7-O-Neo
[11][12]
Neoeriocitrin 596.5 g/mol FLA Citrus aurantium 5,3′,4′-OH
7-O-Neo
[7][11]
Diosmin 608.54 g/mol FLO Citrus sinensis
Citrus limonia
5,3′-OH
4′-OMe
7-O-Rut
[10]
Hesperidin 610.1898 g/mol FLA Citrus sinensis 5,3′-OH,
4′-OMe
7-O-Rut
[13]
Rutin 610.517 g/mol FOL Citrus limonia 5,7,3′,4′-OH
3-O-Rut
[13][14]
Naringenin 272.257 g/mol FLA Citrus paradisi 5,7,4′-OH [15][16]
Hesperetin 302.27 g/mol FLA Citrus sinensis 5,7,3′-OH
4′-OMe
[12][17]
Kaempferol 286.23 g/mol FOL Citrus paradisi 5,7,3,4′-OH [11]
Quercetin 302.236 g/mol FOL Citrus limonia 5,7,3,3′,4′-OH [13]
Tangeretin 372.37 g/mol FLO Citrus aurantium
Citrus paradisi
Citrus limonia
5,6,7,8,4′-OMe [18]
Luteolin 286.24 g/mol FLO Citrus limonia
Citrus aurantium
5,7,3′,4′-OH [19]

2.2. Structure of Naringin

Asahina and Inubuse identified and characterized the chemical structure and molecular formula of naringin in 1928. A 2-O-(alpha-L-rhamnopyranosyl)-beta-D-glucopyranosyl moiety is substituted at position 7 of the disaccharide derivative naringin by an alpha-L-rhamnopyranosyl group via a glycosidic bond. The melting point of naringin is 83 °C at a solubility of 1 mg/mL at 40 °C and the molecular weight of naringin is 580.5 g/mol [20]. The molecular structure of naringin is shown in Figure 1. With a rise in temperature, naringin and naringenin, its aglycon equivalent, become more soluble in various solvents. In the order of methanol, ethyl acetate, n-butanol, isopropanol, petroleum ether, and hexane, naringin was soluble in the six solvents [15]. Naringin complexes are 15 times more soluble in water at 37 ± 0.1 °C than free naringin. It starts to degrade at temperatures above 100 °C or when light is present [16]. The presence of a carboxylic group is suggested by the wide, strong -OH stretching absorption from 3300 per cm to 2500 per cm. Alcohols and phenols are represented by the strong and wide hydrogen-bonded O-H stretching bands centered at 3300 cm−1 and 3400 cm−1 [6]. The C=C stretching bands for aromatic rings typically appear outside the typical region where C=C emerges for alkenes (1650 cm−1) between 1600 and 1450 cm−1. These peaks only occur with naringin. When researching flavonoid-cyclodextrin inclusion complexes, it was discovered that the characteristic peaks for aromatic rings and phenols in naringin at 1519 cm−1 and 1361 cm−1 had disappeared. The amount of naringin measured and the correlation coefficient (r) for sensory bitterness was 0.97IBU [21].
Figure 1. Molecular structure of naringin.

3. Sources of Naringin

Plants contain a variety of flavonoids, which are widely dispersed and have significant biological functions. Since the quantity of naringin is comparatively higher at the immature stage, citrus fruits are typically used in studies to determine the amount of naringin in fruits [22]. Citrus fruits provide a large number of flavonoids in the diet. Naringin is mostly found in the peel of grapefruit, lime, and their variations; it has several biological functions and is frequently used in food, cosmetics, and medicine. Naringin is a glycoside flavanone seen in grapes and citrus fruits. Naringin was first discovered by DeVry in 1857 [23]. It has been reported that the pith contains a higher quantity of naringin in grapefruit, followed by the peel with the membrane, the seeds, and the juice [24]. The amount of naringin in the seeds of grape fruit is 200 μg/mL and 2300 μg/mL is found in the peel of grape fruit [25]. Pummelo has plenty of naringin in it. Compared to the juice, the quantity of naringin was higher in the peel; the naringin content of the juice of pummelo is 220 μg/mL and in the peel it is 3910 μg/mL. The amount of naringin in lime is very low when compared with pummelo. In both species, a high amount of naringin content is present in the skin of the fruits. The amount of naringin found in skin, juice, and seed is 517.2 μg/mL, 98 μg/mL, and 29.2 μg/mL, respectively [26]. The distribution of naringin based on the calculations of various studies in Citrus aurantiifolia is shown in Figure 2. Naringin content in sour orange is 47.1 μg/mL. In sour orange flower, the amount of naringin in the receptacle, ovary, and stigma is 1.3444 μg/mL, 9.036 μg/mL, and 2.554 μg/mL, respectively [27]. Phenol is a chemical compound with a hydroxyl group attached to an aromatic ring. Tannin is a type of phenol compound found in plants, known for its astringent properties. Naringin is a flavonoid compound found in citrus fruits that exhibits antioxidant and anti-inflammatory effects.
Figure 2. The distribution of naringin in Citrus aurantiifolia.

4. Extraction of Naringin

The citrus peel contains significant levels of the flavanones neohesperidin, hesperidin, naringin, and narirutin as well as polymethoxylated flavones tangeretin, sinensetin, and nobiletin. Flavonols, glycosylated flavones, and hydrocinnamic acid are present in very minor amounts [28]. There are three main steps for naringin isolation from fruits including extraction, separation, and purification [10]. Only after utilizing the proper extraction process can flavonoids be isolated, recognized, and classified. The amount of naringin in fruit is determined by numerous factors which include the harvesting time of the fruit, the section of fruit utilized, and whether the peel is a source of naringin. The most prevalent method of extraction is conventional solvent extraction [29]. The various non-traditional techniques include high hydrostatic pressure extraction [30], ultrasound assisted extraction [31], microwave assisted extraction [32], and subcritical [33] and supercritical [34] extraction. The first step in the process involves pre-treating or preparing the sample, during which centrifugation, filtration, drying, and other techniques may be performed. Naringin is extracted, isolated, and purified from various plant materials in the second stage. Naringin is extracted utilizing techniques like soxhlet, maceration, water infusion, microwave extraction, ultrasound extraction, supercritical fluid extraction, auto-hydrolysis, and solid micro-phase extraction, etc. in this step, as given in Figure 3. The final phase often involves the identification, quantification, and recovery of flavonoid components using chromatography techniques on the purified and extracted extracts.
Figure 3. Extraction of naringin using different techniques.

5. Possible Health Benefit

Since ancient times, citrus fruits have been utilized as natural herbal treatments in traditional medicine. Citrus peel has been utilized in traditional Chinese medicine to enhance digestion, minimize gastric gas, bloating, and clear congestion [35]. Clinical and epidemiologic research states that eating citrus fruits lowers the risk of lifestyle-related disorders like cancer, cardiovascular disease, diabetes (type-2), and osteoporosis [36]. Naringin has been shown to have anticancer, antiapoptotic, cholesterol-lowering, antiatherogenic, and metal binding capabilities, as well as antioxidant qualities, as shown in Figure 4. Naringin is also said to enhance medication absorption and metabolism [37].
Figure 4. Potential health benefits of naringin; SOD: superoxide dismutase, ROS: Reactive oxygen species, GSH: γ-l-glutamyl-l-cysteinyl-glycine (glutathione).

5.1. Anticancer Properties of Naringin

Naringin has been reported to inhibit many malignancies through the regulation of various cellular signaling cascades, including the inhibition of malignant cell growth, the induction of apoptosis and also the arresting of the cell cycle and the regulation of oxidative stress, inflammatory processes, and angiogenesis [38]. It was discovered that naringin at concentrations of 250–2000 M promoted cell apoptosis in cervical cancer cells (SiHa) in a dose-dependent way. This impact of naringin is thought to have contributed to the suppression of cell growth as well and also increase in apoptosis [10].
Naringin in the concentrations of 1 M, 5 M, and 10 M has reduced cell mortality caused by rotenone in human neuroblastoma cells (SH-SY5Y). In 4, 6-diamidino-2-phenylindol (DAPI) staining and terminal deoxynucleotidyl transferase dUTP nick end labeling (TUNEL) tests, naringin prevents condensation of chromatin and breakage of DNA strand production by rotenone [39]. Naringin also decreases rotenone-induced phosphorylation of the mitogene-activated kinase (MAPK) family members p38 and Jun NH2-terminal protein kinase (JNK) [40]. According to one study, naringin inhibits the growth of cells, and apoptosis was induced in K562, HL-60, and Kasumi-1 human myeloid leukemia cells in a concentration- and time-dependent manner by downregulating Mcl-1 expression and activating the caspase and PARP pathways. In U937 and THP-1 human leukemia cells, naringin therapy increased cell death and lowered cell cervical proliferation and expansion [41].
Breast cancer is a term that refers to various types of cancers. A vast variety of individualized treatments for breast cancer have recently been offered, all of which have been shown to be effective [42]. Chemotherapy and cancer chemoprevention are both carried out with natural products containing bioactive chemicals. In MCF-7 cell lines, naringin treatment reduced proliferation and growth while also increasing apoptosis. In canine mammary cancer cells (CMT-U27), naringin oxime treatment decreased cell proliferation and viability [40]. Cervical cancer is the second most common cancer in women around the globe and continues to be difficult. At a dose of 750 M, naringin displayed a 50% suppression of SiHa human cervical cancer cells. 

5.2. Antidiabetic Properties of Naringin

It has been demonstrated that naringin enhances insulin sensitivity. Naringin has shown that it can improve insulin action and cell uptake of glucose. Insulin resistance is a major contributor to the onset of type 2 diabetes. This can enhance overall glycemic control and help control blood sugar levels. Insulin moves sugars from the bloodstream into cells, where they are used or stored as energy [43]. Diabetes is when the body does not produce enough insulin or cannot utilize the insulin it makes efficiently [18]. Diabetes is divided into two types. Type 1 Diabetes is a condition that is autoimmune. Cells in the pancreas, which make insulin, are attacked and destroyed by the immune system and what generates this attack remains an enigma. Approximately 10% of diabetics have this type of diabetes [21].
Naringin is a powerful biomolecule that has the potential to help people with diabetes and its consequences [44]. Naringin restricts the secretion and sensitivity of insulin, PPAR, glucose transporters, blood lipids, hepatic glucose production, peripheral glucose uptake, intestinal glucose absorption, biosynthesis of cholesterol, oxidative stress, and inflammation [45]. Inflammatory cytokines are elevated and insulin resistance and hyperglycemia are generated by a high-fat diet. Naringin’s hypoglycemic impact has been thoroughly documented. Vitamin C (50 mg/kg) with naringin co-treatment improved insulin concentration and oxidative stress reduction in rats with streptozotocin-induced diabetes [18].

5.3. Anti-Inflammatory Properties of Naringin

The process by which the body’s white blood cells and the substances they make protect against bacterial and viral illness is known as inflammation. Flavanone-rich plants, such as naringin, hesperidin, and neohesperidin, have long been known to have anti-inflammatory properties [17]. Inflammation is divided into two types. Acute inflammation is the body’s reaction to a quick injury, such as cutting your finger. Your body sends inflammatory cells to the wound to help it heal. The healing process begins with these cells [46]. Chronic inflammation occurs when your body sends inflammatory cells even when there is no external threat.
The anti-inflammatory process controlled by nuclear factor-erythroid 2–related factor 2 (Nrf2) regulates cellular antioxidant synthesis and thus plays a very important role in preventing various degenerative illnesses [47]. In 3-nitropropionic acid-induced rats, naringin upregulates the expression of mRNA in HO-1, GST P1, NAD(P)H:quinone oxidoreductase 1, and g-glutamylcysteine ligase; this is followed by activating Nrf2 and the reduced expression of proinflammatory mediators like TNF-a, cyclooxygenase-2, and inducible NO synthase [48]. Naringin did not inhibit cell proliferation, but it did inhibit RANTES (regulated upon activation of normal T-cell expressed and secreted) production in a human epidermal keratinocytes cell line (HaCaT cells) by restricting nuclear translocation of NF-JB [49]

5.4. Hepatoprotective Properties of Naringin

The capability of a chemical compound to inhibit liver toxicity is known as hepatoprotection [50]. Naringin is suggested to enhance the functioning of the hepatic antioxidant system as well as the metabolism of hepatotoxic substances [51]. Naringin exhibits protection against naturally occurring genotoxins in food, like PhIP (2-Amino-1-methyl-6-phenylimidazo[4,5-b] pyridine) and other cooked food mutagens, by lessening PHIP induced genotoxicity in human liver segments at a concentration of 1000 M [48]. Naringin (0.05–0.125 g/L) increased ethanol and lipid metabolism in rats, alleviating the adverse effects of ethanol consumption. It also reduced necrosis, steatosis, and fibrosis in rat models of alcoholic liver disease, as demonstrated by reduced expression of PGC1α (Peroxisome proliferator-activated receptor-gamma coactivator) or Sirt1; it is an enzyme involved in regulating energy metabolism in response to calorie restrictions at a dosage of 100 mg/day [52].

5.5. Pharmacokinetics of Naringin

Studies were conducted with help of rats to understand the pharmacokinetic properties of naringin. The study of the absorption, distribution, metabolism, and excretion of drugs is known as pharmacokinetics [53]. Proton-coupled active transport and passive diffusion are used to absorb flavanone aglycones into the enterocytes. The low molecular weight, high lipophilicity, and slightly acidic character of aglycones cause passive diffusion. Once within the cells, naringin is expected to go through phase I metabolism, such as oxidation or demethylation by cytochrome P450 monooxygenases, then passing to phase II metabolism, such as sulfation, glucuronidation, or methylation, in intestinal cells or liver cells [54][55]. Naringin is rapidly absorbed in the blood, with the initial concentration peaking at 15 min and the second peaking at 3 h after naringin monomer oral administration; 480 min later, it is undetectable [56]. The affinity of these food chemicals for serum albumin, the primary transport protein, coincides with their tissue distribution and elimination.
In terms of tissue distribution, the liver had the largest quantities of flavanone conjugates after repeated or single dose flavanone treatment in rats. By partially undergoing breakage of the bacterial ring and then the three bridges of carbon to dihydrochalcone moiety, naringin is eliminated by the kidneys into the urine and by the liver into bile According to Fuhr and Kummert’s findings (1995). Urine excretion ranges from 5 to 57 percent of total intake. Sulfates were the most common naringenin type detected in the tissues of rats. Only the liver and kidney had glucuronide concentrations that could be measured [57]. The average Cmax of naringin in portal plasma was 18.83.8 min (determined by the concentration reached at tmax in portal plasma), whereas the absorption ratios of naringin in portal plasma and lymph fluid were approximately 95.9 and 4.1, respectively, after naringin administration via a duodenal cannula (600 and 1000 mg/kg). This suggests that naringin is absorbed largely by portal blood rather than mesenteric lymph fluid and that it is excreted primarily by bile, with just a tiny quantity entering systemic circulation following hepatic metabolism [58]

6. Application of Naringin

6.1. In Cosmetic Industry

The flavonoid naringin has anti-cancer, anti-oxidative, anti-aging, antibacterial, anti-inflammatory, cholesterol-lowering, and free radical scavenging properties [59]. Studies show that naringin reduces the risk of toxicity caused by other sunscreen ingredients like TiO2 when it is added to sunscreen formulations because of its antioxidant activity. It also scavenges free radicals produced by UV radiation and by the photocatalytic activity of ZnO and TiO2, which further lowers the risk of toxicity [60].

6.2. Pharmaceutical Application

The area of the wound and the length of the epithelization phase significantly decreased during treatment with naringin ointment formulation, whilst the velocity of wound contraction dramatically increased. Naringin ointment formulation modulates collagen-1 expression to promote angiogenesis, which in turn promotes wound healing. This is accomplished by down-regulating the expression of inflammatory (ILs, NF-Jb, and TNF-a), apoptotic (pol-g and Bax), and growth factor (TGF-b and VEGF) genes [61].

6.3. In Livestock Sector

Naringin and quercetin reduce protozoa and methanogen populations in the rumen and suppress methane production without negatively affecting the parameters of ruminal fermentation. Daily diets containing hesperidin and naringin have proved successful in enhancing milk’s oxidative stability while having no negative impacts on the substance’s chemical compositions, coagulation abilities, or fatty acid profile [62].

6.4. Food Industry

The use of naringin microspheres in yogurt demonstrated their ability to effectively reduce whey precipitation and to slow pH drop. According to a study, naringin-encapsulated microspheres could extend the shelf life of this bioactive product and offer a fresh concept for functional yogurt [63]. Hesperidin, naringin, and coumarins have been found to inhibit xanthine oxidase, which directly reduces cellular free radical production. When compared to dietary citrus pulp and control diets, feeding dietary citrus pulp prolonged the shelf life of beef during retail display by increasing antioxidant activity, lowering coliforms, and reducing lipid and protein oxidation [64]. Naringin’s incorporation caused significant UV blocking, plasticizing, and antioxidant and antibacterial effects. The biological oxygen demand (BOD) in saltwater was used to test the biodegradability of these films, showing excellent disintegration under these circumstances [65].

This entry is adapted from the peer-reviewed paper 10.3390/molecules28155623

References

  1. Alam, M.A.; Subhan, N.; Rahman, M.M.; Uddin, S.J.; Reza, H.M.; Sarker, S.D. Effect of citrus flavonoids, naringin and naringenin, on metabolic syndrome and their mechanisms of action. Adv. Nutr. 2014, 5, 404–417.
  2. Jakab, J.; Miškić, B.; Mikšić, Š.; Juranić, B.; Ćosić, V.; Schwarz, D.; Včev, A. Adipogenesis as a potential anti-obesity target: A review of pharmacological treatment and natural products. Diabetes Metab. Syndr. Obes. 2021, 14, 67–83.
  3. Panche, A.N.; Diwan, A.D.; Chandra, S.R. Flavonoids: An overview. J. Nutr. Sci. 2016, 5, e47.
  4. Nawaz, R.; Abbasi, N.A.; Khan, M.R.; Ali, I.; Hasan, S.Z.U.; Hayat, A. Color development in “Feutrell’s early” (Citrus reticulata Blanco) affects peel composition and juice biochemical properties. Int. J. Fruit Sci. 2020, 20, 871–890.
  5. Tan, J.; Li, Y.; Hou, D. The Effects and Mechanisms of cyanidin-3-glucoside and Its phenolic Metabolites in Maintaining intestinal Integrity. Antioxidants 2019, 8, 479.
  6. Puri, M. Updates on naringinase: Structural and biotechnological aspects. Appl. Microbiol. Biotechnol. 2012, 93, 49–60.
  7. Hollman, P.C.H.; Arts, I.C.W. Flavonols, flavones and flavanols—Nature, occurrence and dietary burden. J. Sci. Food Agric. 2000, 80, 1081–1093.
  8. Křížová, L.; Dadáková, K.; Kašparovská, J.; Kašparovský, T. Isoflavones. Molecules 2019, 24, 1076.
  9. Da Silva, F.L.; Escribano-Bailón, M.T.; Pérez Alonso, J.J.; Rivas-Gonzalo, J.C.; Santos-Buelga, C. Anthocyanin pigments in strawberry. LWT—Food Sci. Technol. 2007, 40, 374–382.
  10. Sharma, K.; Mahato, N.; Lee, Y.R. Extraction, characterization and biological activity of citrus flavonoids. Rev. Chem. Eng. 2019, 35, 265–284.
  11. Li, Q.; Zhang, N.; Sun, X.; Zhan, H.; Tian, J.; Fei, X.; Liu, X.; Chen, G.; Wang, Y. Controllable biotransformation of naringin to prunin by naringinase immobilized on functionalized silica. J. Chem. Technol. Biotechnol. 2021, 96, 1218–1227.
  12. Verma, A.K.; Pratap, R. Chemistry of biologically important flavones. Tetrahedron 2012, 68, 8523–8538.
  13. Billowria, K.; Ali, R.; Rangra, N.K.; Kumar, R.; Chawla, P.A. Bioactive Flavonoids: A Comprehensive Review on Pharmacokinetics and Analytical Aspects. Crit. Rev. Anal. Chem. 2022, 1–15.
  14. Tang, D.M.; Zhu, C.F.; Zhong, S.A.; Da Zhou, M. Extraction of naringin from pomelo peels as dihydrochalcone’s precursor. J. Sep. Sci. 2011, 34, 113–117.
  15. Zhang, L.; Song, L.; Zhang, P.; Liu, T.; Zhou, L.; Yang, G.; Lin, R.; Zhang, J. Solubilities of naringin and naringenin in Different Solvents and Dissociation Constants of naringenin. J. Chem. Eng. Data 2015, 60, 932–940.
  16. Aron, P.M.; Kennedy, J.A. Flavan-3-ols: Nature, occurrence and biological activity. Mol. Nutr. Food Res. 2008, 52, 79–104.
  17. Su, W.; Wang, Y.; Li, P.; Wu, H.; Zeng, X.; Shi, R.; Zheng, Y.; Li, P.; Peng, W. The potential application of the traditional Chinese herb Exocarpium Citri grandis in the prevention and treatment of COVID-19. Tradit. Med. Res. 2020, 5, 160–166.
  18. Ahmed, O.M.; Mahmoud, A.M.; Abdel-Moneim, A.; Ashour, M.B. Antidiabetic effects of hesperidin and naringin in type 2 diabetic rats. Diabetol. Croat. 2012, 41, 53–67.
  19. Sharma, A.; Bhardwaj, P.; Arya, S.K. Naringin: A potential natural product in the field of biomedical applications. Carbohydr. Polym. 2021, 2, 100068.
  20. Aherne, S.A.; O’Brien, N.M. Dietary flavonols: Chemistry, food content, and metabolism. Nutr. J. 2002, 18, 75–81.
  21. Ioannou, I.; Nouha, M.; Chaaban, H.; Boudhrioua, N.M.; Ghoul, M. Effect of the process, temperature, light and oxygen on naringin extraction and the evolution of its antioxidant activity. Int. J. Food Sci. Technol. 2020, 53, 2754–2760.
  22. Wang, S.; Yang, C.; Tu, H.; Zhou, J.; Liu, X.; Cheng, Y.; Luo, J.; Deng, X.; Zhang, H.; Xu, J. Characterization and metabolic diversity of flavonoids in Citrus species. Sci. Rep. 2017, 7, 10549.
  23. Hoffmann, E. Naringin (Hesperidin de Vry). Arch. Pharm. 1879, 214, 139–145.
  24. Liu, Y.; Heying, E.; Tanumihardjo, S.A. History, global distribution, and nutritional importance of citrus fruits. Compr. Rev. Food Sci. Food Saf. 2012, 11, 530–545.
  25. Pichaiyongvongdee, S.; Haruenkit, R. Comparative studies of limonin and naringin distribution in different parts of pummelo (Citrus grandis (L.) Osbeck) cultivars grown in Thailand. J. Nat. Sci. 2009, 43, 28–36.
  26. Sir, K.A.; Randa, E.; Amro, A.B. Content of phenolic compounds and vitamin C and antioxidant activity in wasted parts of Sudanese citrus fruits. Food Sci. Nutr. 2018, 6, 1214–1219.
  27. Hassan, R.A.; Hozayen, W.G.; Abo Sree, H.T.; Al-Muzafar, H.M.; Amin, K.A.; Ahmed, O.M. Naringin and hesperidin counteract diclofenac-induced hepatotoxicity in male Wistar rats via their antioxidant, anti-inflammatory, and antiapoptotic activities. Oxidative Med. Cell. Longev. 2021, 2021, 9990091.
  28. Bahorun, T.; Ramful-baboolall, D.; Neergheen-bhujun, V.; Aruoma, O.I.A.; Verma, S.; Tarnus, E.; Robert, C.; Silva, D.; Rondeau, P. Phytophenolic Nutrients in Citrus: Biochemical and Molecular Evidence. In Advances in Citrus Nutrition; Springer: Berlin/Heidelberg, Germany, 2012.
  29. Zhao, B.T.; Kim, E.J.; Son, K.H.; Son, J.K.; Min, B.S.; Woo, M.H. Quality evaluation and pattern recognition analyses of marker compounds from five medicinal drugs of Rutaceae family by HPLC/PDA. Arch. Pharm. Res. 2015, 38, 1512–1520.
  30. Grassino, A.N.; Pedisić, S.; Dragović-Uzelac, V.; Karlović, S.; Ježek, D.; Bosiljkov, T. Insight into high-hydrostatic pressure extraction of polyphenols from tomato peel waste. Plant Foods Hum. Nutr. 2020, 75, 427–433.
  31. Victor, M.M.; David, J.M.; Sakukuma, M.C.K.; França, E.L.; Nunes, A.V.J. A simple and efficient process for the extraction of naringin from grapefruit peel waste. Green Process. Synth. 2018, 7, 524–529.
  32. Chávez-González, M.L.; Sepúlveda, L.; Verma, D.K.; Luna-García, H.A.; Rodríguez-Durán, L.V.; Ilina, A.; Aguilar, C.N. Conventional and emerging extraction processes of flavonoids. Processes 2020, 8, 434.
  33. Ciğeroğlu, Z.; Bayramoğlu, M.; Kırbaşlar, Ş.İ.; Şahin, S. Comparison of microwave-assisted techniques for the extraction of antioxidants from Citrus paradisi Macf. biowastes. J. Food Sci. Technol. 2021, 58, 1190–1198.
  34. Cheigh, C.I.; Chung, E.Y.; Chung, M.S. Enhanced extraction of flavanones hesperidin and narirutin from Citrus unshiu peel using subcritical water. J. Food Eng. 2012, 110, 472–477.
  35. Bacanli, M.; Başaran, A.A.; Başaran, N. The major flavonoid of grapefruit: Naringin. In Polyphenols: Prevention and Treatment of Human Disease; Academic Press: Cambridge, MA, USA, 2018; pp. 37–44.
  36. El Kantar, S.; Rajha, H.N.; Boussetta, N.; Vorobiev, E.; Maroun, R.G.; Louka, N. Green extraction of polyphenols from grapefruit peels using high voltage electrical discharges, deep eutectic solvents and aqueous glycerol. Food Chem. 2019, 295, 165–171.
  37. Ma, G.; Zhang, L.; Sugiura, M.; Kato, M. Citrus and Health. The Genus Citrus; Elsevier: Amsterdam, The Netherlands, 2020.
  38. Yan, Y.; Zhou, H.; Wu, C.; Feng, X.; Han, C.; Chen, H.; Liu, Y.; Li, Y. Ultrasound-assisted aqueous two-phase extraction of synephrine, naringin, and neohesperidin from Citrus aurantium L. fruitlets. Prep. Biochem. Biotechnol. 2021, 51, 780–791.
  39. Gupta, A.K.; Dhua, S.; Sahu, P.P.; Abate, G.; Mishra, P.; Mastinu, A. Variation in phytochemical, antioxidant and volatile composition of pomelo fruit (Citrus grandis (L.) osbeck) during seasonal growth and development. Plants 2021, 10, 1941.
  40. Ghanbari-Movahed, M.; Jackson, G.; Farzaei, M.H.; Bishayee, A. A systematic review of the preventive and therapeutic effects of naringin against human malignancies. Front. Pharmacol. 2021, 12, 639840.
  41. Li, H.; Yang, B.; Huang, J.; Xiang, T.; Yin, X.; Wan, J.; Luo, F.; Zhang, L.; Li, H.; Ren, G. Naringin inhibits growth potential of human triple-negative breast cancer cells by targeting β-catenin signaling pathway. Toxicol. Lett. 2013, 220, 219–228.
  42. Aroui, S.; Fetoui, H.; Kenani, A. Natural dietary compound naringin inhibits glioblastoma cancer neoangiogenesis. BMC Pharmacol. Toxicol. 2020, 21, 46.
  43. Ahmed, S.; Khan, H.; Aschner, M.; Hasan, M.M.; Hassan, S.T.S. Therapeutic potential of naringin in neurological disorders. Food Chem. Toxicol. 2019, 132, 110646.
  44. Mahmoud, A.M. Anti-diabetic effect of naringin: Insights into the molecular mechanism. Int. J. Obes. 2016, 8, 324–328.
  45. Pandey, V.K.; Shams, R.; Singh, R.; Dar, A.H.; Pandiselvam, R.; Rusu, A.V.; Trif, M. A comprehensive review on clove (Caryophyllus aromaticus L.) essential oil and its significance in the formulation of edible coatings for potential food applications. Front. Nutr. 2022, 9, 987674.
  46. Chen, F.; Zhang, N.; Ma, X.; Huang, T.; Shao, Y.; Wu, C.; Wang, Q. Naringin alleviates diabetic kidney disease through inhibiting oxidative stress and inflammatory reaction. PLoS ONE. 2015, 10, e0143868.
  47. Kumar, R.; Clermont, G.; Vodovotz, Y.; Chow, C.C. The dynamics of acute inflammation. J. Theor. Biol. 2004, 230, 145–155.
  48. Lawrence, T.; Gilroy, D.W. Chronic inflammation: A failure of resolution. Int. J. Exp. Pathol. 2007, 88, 85–94.
  49. El-Desoky, A.H.; Abdel-Rahman, R.F.; Ahmed, O.K.; El-Beltagi, H.S.; Hattori, M. Anti-inflammatory and antioxidant activities of naringin isolated from Carissa carandas L.: In vitro and in vivo evidence. Phytomedicine 2018, 42, 126–134.
  50. Pari, L.; Amudha, K. Hepatoprotective role of naringin on nickel-induced toxicity in male Wistar rats. Eur. J. Clin. Pharmacol. 2011, 650, 364–370.
  51. Chen, J.C.; Li, L.J.; Wen, S.M.; He, Y.C.; Liu, H.X.; Zheng, Q.S. Quantitative analysis and simulation of anti-inflammatory effects from the active components of Paino Powder in rats. Chin. J. Integr. Med. 2011, 201203, online ahead of print.
  52. Shirani, K.; Yousefsani, B.S.; Shirani, M.; Karimi, G. Protective effects of naringin against drugs and chemical toxins induced hepatotoxicity: A review. Phytother. Res. 2020, 34, 1734–1744.
  53. Gillessen, A.; Schmidt, H.H.J. Silymarin as supportive treatment in liver diseases: A narrative review. Adv. Ther. 2020, 37, 1279–1301.
  54. Kawaguchi, K.; Maruyama, H.; Hasunuma, R.; Kumazawa, Y. Suppression of inflammatory responses after onset of collagen-induced arthritis in mice by oral administration of the Citrus flavanone naringin. Immunopharmacol. Immunotoxicol. 2011, 33, 723–729.
  55. Yang, X.; Zhao, Y.; Gu, Q.; Chen, W.; Guo, X. Effects of naringin on postharvest storage quality of bean sprouts. Foods 2022, 11, 2294.
  56. Bailey, D.G.; Dresser, G.K. Interactions between grapefruit juice and cardiovascular drugs. Am. J. Cardiol. 2004, 4, 281–297.
  57. Tsai, Y.J.; Tsai, T.H. Mesenteric lymphatic absorption and the pharmacokinetics of naringin and naringenin in the rat. J. Agric. Food Chem. 2012, 60, 12435–12442.
  58. Tesseromatis, C.; Alevizou, A. The role of the protein-binding on the mode of drug action as well the interactions with other drugs. Eur. J. Drug Metab. Pharmacokinet. 2008, 33, 225–230.
  59. Goyal, A.; Verma, A.; Dubey, N.; Raghav, J.; Agrawal, A. Naringenin: A prospec-tive therapeutic agent for Alzheimer’s and Parkinson’s disease. J. Food Biochem. 2022, 46, e14415.
  60. Turfus, S.C.; Delgoda, R.; Picking, D.; Gurley, B.J. Pharmacokinetics. In Pharmacognosy; Elsevier: Amsterdam, The Netherlands, 2017.
  61. Pandey, V.K.; Islam, R.U.; Shams, R.; Dar, A.H. A comprehensive review on the application of essential oils as bioactive compounds in nano-emulsion based edible coatings of fruits and vegetables. Appl. Food Res. 2022, 2, 100042.
  62. Oliva, J.; French, B.A.; Li, J.; Bardag-Gorce, F.; Fu, P.; French, S.W. Sirt1 is involved in energy metabolism: The role of chronic ethanol feeding and resveratrol. Exp. Mol. Pathol. 2008, 85, 155–159.
  63. Goodarzi-Boroojeni, F.; Männer, K.; Zentek, J. The impacts of Macleaya cordata extract and naringin inclusion in post-weaning piglet diets on performance, nutrient digestibility and intestinal histomorphology. Arch. Anim. Nutr. 2018, 72, 178–189.
  64. Kandhare, A.D.; Alam, J.; Patil, M.V.K.; Sinha, A.; Bodhankar, S.L. Wound healing potential of naringin ointment formulation via regulating the expression of inflammatory, apoptotic and growth mediators in experimental rats. Pharm. Biol. 2016, 54, 419–432.
  65. Wang, H.; Hu, H.; Zhang, X.; Zheng, L.; Ruan, J.; Cao, J.; Zhang, X. Preparation, physicochemical characterization, and antioxidant activity of naringin–silk fibroin–alginate microspheres and application in yogurt. Foods 2022, 11, 2147.
More
This entry is offline, you can click here to edit this entry!
ScholarVision Creations