You are on page 1of 33

Journal Pre-proof

A mechanistic review of the anticancer potential of hesperidin, a


natural flavonoid from citrus fruits

Pratibha Pandey , Fahad Khan

PII: S0271-5317(21)00034-8
DOI: https://doi.org/10.1016/j.nutres.2021.05.011
Reference: NTR 8241

To appear in: Nutrition Research

Received date: 2 March 2021


Revised date: 23 May 2021
Accepted date: 31 May 2021

Please cite this article as: Pratibha Pandey , Fahad Khan , A mechanistic review of the anti-
cancer potential of hesperidin, a natural flavonoid from citrus fruits, Nutrition Research (2021), doi:
https://doi.org/10.1016/j.nutres.2021.05.011

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2021 Elsevier Inc. All rights reserved.


Highlights

 Perspective of hesperidin as potential anticancer drug candidate

 Therapeutic potential of hesperidin in in vivo and in vitro models

 Major targets of hesperidin and their associated signaling pathways

 Enhanced anticancerous properties of hesperidin formulations


A mechanistic review of the anticancer potential of hesperidin, a natural flavonoid from

citrus fruits

Pratibha Pandey1, Fahad Khan1*


1
Department of Biotechnology, Noida Institute of Engineering & Technology, 19, Knowledge

Park-II, Institutional Area, Greater Noida, 201306, India

*Corresponding author:
Dr. Fahad Khan
Assistant Professor
Department of Biotechnology
Noida Institute of Engineering and Technology, Greater Noida
fahadintegralian@gmail.com
Abstract

Hesperidin, a phytoactive compound, is an abundant and economical dietary bioflavonoid

possessing numerous biological and medicinal benefits. Several studies have strongly proven the

significant chemotherapeutic potential of hesperidin. Therefore this review is an effort to bring

together the existing studies demonstrating hesperidin as a potential anticancer agent with its

mode of action reported in the therapeutic strategies for numerous cancer types. Hesperidin acts

via modulating multiple pathways involving cell cycle arrest, apoptosis, antiangiogenic,

antimetastatic and DNA repair in various cancer cells. Hesperidin has been reported to alter

several molecular targets related to carcinogenesis, such as reactive nitrogen species, cellular

kinases, transcription factors, reactive oxygen species, drug transporters, cell cycle mediators and

inflammatory cytokines. Altogether, this review provides significant insights for the potential of

hesperidin to be a strong and promising candidate for pharmaceuticals, functional foods, dietary

supplements, nutraceuticals and geared toward the better management of carcinoma.

Keywords

Anticancer potential; Phytocompound; Metabolism; Hesperidin; Citrus


1. Introduction

Bioflavonoids have been gaining wider attention in cancer management for their strong

antioxidant and bioactive potential (1). Hesperidin (3,5,7-trihydroxyflavanone-7-

rhamnoglucoside) is a natural phenolic compound with several biological effects (2). It is a much

known fact that products derived from citrus fruits have been widely used for combating several

human ailments for thousands of years (3). Flavonoids are considered to be the major category of

such dietary polyphenols (4). Amongst diverse flavonoids, hesperidin, a flavanone glycoside, is

widely found in sweet oranges, lemons, and some other vegetables and fruits (5). It is a colorless

and odorless plant chemical with the chemical formula of C28H34O15 and a molecular weight of

610.1898 g/mol (Figure 1A). Its name was derived from the word "hesperidium", specifically for

the fruit of citrus trees (6). Various studies have reported limited bioavailability and poor

absorption of hesperidin in experimental animals due to the attachment of rutinoside moiety with

the dietary flavonoid (7). However, hesperidin has been recognized as a potent candidate

phytocompound with numerous health benefits such as antimicrobial, anticancer, antioxidant,

anti-inflammatory, antidiabetic, and cardiovascular protective activities (8). Specifically,

hesperidin has been confirmed to possess both chemopreventive and chemotherapeutic potential

(9). To illustrate a better perceptive of the anticancerous potential of hesperidin, this review

focuses on major natural sources of hesperidin, its bioavailability, anticancer potential,

pharmacokinetics, and its associated molecular mechanisms. Furthermore, the anticancer

potential of hesperidin analogs, hesperidin-nanoformulations, and hesperidin metabolites are also

outlined. Several safety issues associated with the usage of hesperidin as a dietary component are

also summarized. We emphasize that hesperidin is a promising natural flavonoid with significant

anticancerous potential, and this review would be highly helpful in furthering the usage of
hesperidin as one of the natural compounds for the prevention and management of various

carcinomas.

2. Natural sources of hesperidin

Hesperidin is isolated from ordinary orange Citrus aurantium L, C. sinensis, C. unshiu, and other

species of the genus Citrus (10). It has also been reported in several other plants (apart from

Citrus) such as in Betulaceae, Fabaceae, Papilionaceae and Lamiaceae, the bark of Zanthoxylum.

cuspidatum, and Zanthoxylum avicennae. Hesperidin is found in various vegetables, fruits, and

several medicinal plants. Though hesperidin is widely distributed in many plants, Cyclopia

maculata stems have been recognized as the major source of hesperidin (11). Citrus fruits,

Cyclopia maculate tea, and oranges are also considered as rich sources of hesperidin. Table 1

outlines the approximate hesperidin content of selected plants with medicinal and nutritional

value.

3. Absorption and metabolism of hesperidin

Numerous research reports have emphasized the importance of absorption and metabolism of

phytocompound for its further utilization as the safe and effective anticancer lead candidate (19).

Dietary flavonoids usually have limited or low bioavailability. During their oral administration,

hesperidin is converted to hesperetin (a trihydroxyflavanone), an aglycone of hesperidin, by β-

glucosidase present in intestinal microflora and then absorbed from the gastrointestinal tract

(GIT) (20). Hesperetin is the predominant flavonoid with having three hydroxy groups at 3'-, 5-

and 7-positions and methoxy substituent at the 4'-position (Figure 1B). Additionally, several

reports have displayed the degradation of flavonoid into several carboxylic acid and phenolic

products after the flavonoid glycosides hydrolysis by gut bacteria (21). Thus, intestinal bacteria

metabolism might play an important role in hesperidin uptake. It is therefore conceivable that
changes in intestinal microflora might affect the absorption and subsequent pharmacokinetics of

hesperidin, and consequently lead to alterations in the biological activity exhibited by hesperidin.

The gut microbiome exhibits a significant role in hesperidin metabolism (22). Several studies

had reported that intake of citrus juice leads to the suppression of inflammatory enzymes and

pro-inflammatory cytokines expression in the colon, thereby resulting in the inhibition of colon

cancer (23), and further explained that some intestinal bacterial enzymes such as rhamnosidases

may exhibit differentiated activity between neohesperidoside moiety of naringenin and rutinoside

moiety of hesperetin (hesperidin). Hesperidin might increase the epithelial barrier role of Caco-2

cells in a concentration-dependent manner (24). Three types of metabolites of hesperidin

including sulfates, aglycons, phenolic acids were absorbed in the intestinal tissue. The liver and

intestinal tract are the key metabolic site, where both host metabolic enzymes and intestinal

bacteria take part in chemical biotransformation (25). Hesperidin is known to be mostly resistant

to enzymatic breakdown in the small intestine and stomach, thus, mainly get into the intact colon

(26). Several earlier reports have suggested the appearance of hesperidin in the small intestine,

even though the majority is transformed in the colon, some breakdown can previously occur in

the small intestine (distal part) (27). It was also found that the antiproliferative and antioxidant

activities of flavonoids were greatly improved after its deglycosylation (28-29). Thus, after

administration of hesperetin in the duodenum, cecum, and colon were valuable to several

intestinal bioactivities. Several bacteria are found in the intestinal tract and gut microbiota

contains numerous enzymes which modify food compounds (30). Microbial enzymes can

remove glucuronides, glycosides, sulfates, and produce glycons (flavonoid), which gets further

metabolized into several ring fission products, thus, a range of phenolic acid metabolites was

produced after hesperidin administration in mice, these metabolites were absorbed in the colonic
tissue (32). Previously reported studies suggested that the metabolites exhibited decreased

polarity, smaller molecular weight, and stronger retention in comparison to the prototype that

might improve absorption and distribution during in vivo (33). Both phase II enzymes including

UGT (uridine-5ʹ-diphosphate-glucuronosyl transferases) and SULT (sulfotransferases) are

associated with the production of the conjugated metabolites (34). Both enzymes are present in

various tissues such as the small intestine, liver, and colon (35). These experiments further

indicated that the sulfates of hesperidin absorbed in the ileum, after oral administration of

hesperidin, the conjugated metabolites would be absorbed in the ileum and duodenum.

Hesperidin metabolite production is commonly contributed by gut microbiota, liver, and

enterocyte (36). Altogether, hesperidin and its metabolites have been broadly absorbed in various

intestinal tissues. After oral consumption of hesperidin, conjugates of glucuronide, and sulfate

(hesperetin) were found in plasma demonstrating absorption in the small intestine. The

microbiota could produce rhamnosidases that perform absorption by cleaving the rhamnose

moiety from hesperidin.

4. Anticancerous effects of hesperidin

4.1 In vitro anticancerous efficacy

Hesperidin has been reported to limit the proliferation of several cancer cells in vitro, such as

glioma, prostate, pancreas, ovarian, skin, liver, colon, lung, cervical, and breast cancer cells (37).

Table 2 summarizes the recent reports on the anticancerous efficacy of hesperidin and its

associated molecular mechanism behind their mode of action in various cancer cells. Hesperidin

concentration varies with the origin of different cancer cell types and showed significant

anticancerous potential at relatively high doses in in vitro studies (38-39). Hesperidin contributes

multifaceted roles in anticancerous potential, such as inhibition of the invasion, proliferation,


metastasis, migration, and angiogenesis of cancer cells, along with apoptotic and autophagic

induction. Anticancerous efficacy of hesperidin is mediated via the regulation of numerous

molecular signaling pathways and molecular targets involved in carcinogenesis. Altogether,

hesperidin is a promising anticancer phytocompound against several cancer cells (in vitro) (40).

Phytocompounds can increase the cytotoxicity of various chemotherapeutic drugs and irradiation

by decreasing the drug-induced toxicity as well as increasing the sensitivity of several cancer

cells. Hesperidin either alone or in combination with chemotherapeutic drugs, including

paclitaxel, tamoxifen, cyclophosphamide, temozolomide, pirarubicin, methotrexate, oxaliplatin,

and doxorubicin has been reported with significant tumor growth inhibition (Table 3). A

combination of hesperidin and naringenin, a bioactive flavonoid exhibited synergistic

anticancerous effects in pancreatic cancer cells (41). Synergistic anticancerous effects of

hesperidin with chemotherapeutic drugs or irradiation are mediated via apoptotic induction, drug

efflux pumps inhibition, modulation of several cell signaling pathways reported in cancer

progression and, cell cycle protein regulation (Table 3). Altogether, these in vitro studies suggest

that hesperidin could be a potential anticancerous adjuvant agent.

4.2 In vivo anticancerous efficacy

In vivo anticancerous efficacy of hesperidin has been explained in xenograft mice models (58).

Hesperidin has exhibited significant anticancerous potential against numerous cancer types in

vivo, including prostate, breast, colon, and lymphoma. Swiss albino mice, Balb/c nude mice, and

Wistar rat models are commonly used to investigate the chemotherapeutic potential of hesperidin

(59). Cancer models are established by implantation of xenografts or murine cancer cells via

intraperitoneal or subcutaneous injection. Several studies have reported the hesperidin

administration either by oral or intraperitoneal administration at varying doses and different


treatment duration depending on the experimental setup. The mechanism behind the tumor

growth inhibitory potential of hesperidin in xenograft has been mainly associated with apoptotic

induction, angiogenesis inhibition, metastasis suppression, and regulation of associated cell

signaling pathways (60). Table 4 summarizes the in vivo anticancerous potential of hesperidin.

5. Anticancer mechanism of hesperidin

Several research studies have reported the role of hesperidin in cell cycle progression arrest in

various cancer cells (67). Hesperidin can arrest cell cycle progression at different checkpoints

including G1, G2/M, or S phase (68). Hesperidin-mediated cell cycle arrest may be associated

with the modulation of cell cycle regulatory proteins including CDK inhibitors, cyclin-dependant

kinases, and cyclins (Figure 2). Hesperidin was reported to enhance wild-type p53 levels in lung

cancer, breast cancer, leukemia cell lines, and colon cancer (69). Xia et al., 2018 reported that

hesperidin inhibited the G1/S transition in the A549 lung adenocarcinoma cell line via

upregulated p21 and downregulated cyclin D1 expression levels. Altogether, hesperidin can

target cell cycle signaling pathways in the cancer therapeutic approach. WNT/β-catenin signaling

pathway contributes a major role in cell proliferation, growth, and survival (70). Aberrant

WNT/β-catenin signaling pathway has been associated with numerous carcinomas such as

prostate cancer, colorectal cancer, and breast cancer (71). Thus, β-catenin signaling has been

considered as one of the attractive targets in cancer therapeutics (72). Hesperidin has been

reported to induce apoptosis in lung cancer cells via downregulation of PCNA protein, β-catenin,

and c-myc, expressions. Wei Hong et al., 2020 reported the therapeutic potential of hesperidin in

promoting the differentiation of human alveolar osteoblasts through the Wnt/β-Catenin signaling

pathway activation. Hesperidin has also presented significant hepatoprotective efficacy in the
hepatocellular carcinoma group via increased β-catenin, Wnt3a, Wnt5a, Cyclin D1 gene, and

protein expressions (73).

Hesperidin has depicted significant anticancerous potential via targeting apoptosis-mediated cell

death pathways. Recent studies have reported apoptosis as one of the major phenomena by which

various anticancerous agents remove neoplastic or preneoplastic cells (74, 75). Through

literature findings, it was postulated that hesperidin induces apoptosis via several mechanisms

such as increasing DNA fragmentation and nuclear condensation. These apoptosis-inducing

effects of hesperidin are associated with caspase-9 and -3 activation, inhibition of cell cycle

progression, regulation of Bcl-2 (B-cell lymphoma-2) family proteins, decreased levels of NF-κB

(nuclear factor-κB), and elevation of reactive oxygen species levels. Hesperidin was also

reported with inhibition of MMP-9, COX-2 (cyclooxygenase-2), MMP-2 (matrix

metalloproteinase-2), and regulation of mitogen-activated protein kinases (c-Jun N-terminal

kinases and extracellular signal-regulated kinases) phosphorylation (76). Hesperidin induced

apoptosis through triggering of mitochondrial apoptotic pathway and inducing G0/G1 arrest in

A549 (human non-small cell lung cancer) cells (77). Hesperidin also promoted programmed cell

death via downregulation of estrogen receptor signaling (non-genomic) pathway in endometrial

cancer cells (78). Besides, hesperidin activated both intrinsic and extrinsic pathways of apoptosis

in several cancer cells via downregulation of Bcl-2 and upregulation of caspases (caspase-3, 8,9),

and Bax (79). This research strongly supported the fact that hesperidin can induce both intrinsic

(mitochondria-mediated) and extrinsic (death receptor-mediated) apoptotic pathways in various

cancer cells. p53 (tumor suppressor gene) activation is correlated with apoptotic induction via

negative and positive regulation of Bcl-2 and Bax (80). The role of hesperidin in apoptotic

induction has also been confirmed via p53 activation (81). Hesperidin inhibited
hepatocarcinogenesis and suppressed cell proliferation, inflammation, collagen deposition,

oxidative stress via activation of PPARγ and Nrf2/ HO-1/ARE pathways (82).

ROS (reactive oxygen species) formations during metabolism have been involved in various

physiological functions (83). The balance between ROS production and its elimination via

antioxidant defense system has been properly maintained in normal cells whereas ROS

homeostasis is deregulated in cancer cells, leading to higher ROS generation (84). Nonetheless,

if ROS generation reaches beyond its threshold level, it could trigger apoptotic induction in

cancer cells, thereby limiting cancer progression, explaining that ROS can be identified as one of

the potent therapeutic tools in cancer management (85). Several studies have reported that

flavonoids induce anticancer potential via increased ROS generation (86). For instance, the

anticancerous potential of hesperidin is regulated by ROS-dependent apoptotic pathways in

several cancer cells such as primary gall bladder cancer cells, HeLa, HepG2, and prostate cancer

cells (87). Also, hesperidin can be a good ROS scavenger and could act as a potent antioxidant

(non-enzymatic) defense system. Hesperidin has also been reported to promote the loss of MMP

(mitochondrial membrane potential), ROS formation, intracellular Ca2+ mobilization, caspase-3

activation, enhanced release of apoptosis-inducing factor, and cytochrome c from mitochondria

(87). It also mediated cell cycle growth arrest in the G0/G1 phase in HeLa cells via

downregulation of cyclinE1, cyclinD1, and CDKs (cyclin-dependent kinase 2) protein expression

level. In this perspective, hesperidin, induced apoptosis in human HepG2 cells via death receptor

pathway and mitochondrial pathway (88). Hesperidin administration has also shown a significant

reduction in the level of serum AFP level, liver function enzymes, and oxidative stress marker.

Various cellular kinases are recognized as the molecular targets of hesperidin (89). EGFR

(receptor tyrosine kinase), is also activated by its ligands that further promote the
phosphorylation of PI3K (phosphoinositide 3-kinase) and MAPK signaling, contributing a

prominent role in cell survival (90). Furthermore, hesperidin treatment suppressed DEN-induced

upregulation of CDK‐2, PI3K, Akt protein expression, and maintained the integrity of liver

tissues from hepatocellular carcinoma formation (91-92). Altogether, the hepatoprotective

potential of hesperidin is mediated through downregulation of the PI3K/Akt pathway.

Inflammation has been considered as one of the crucial contributing factors in cancer

progression. Pro-inflammatory molecules, including interferon c, interleukins, and TNFα, have

also been implicated in cancer progression and development. During inflammation, MAPK

pathway modulation results in activation of AP-1 and NF- ⱪB which is further associated with

increased COX-2 and iNOS gene expression (93-94). Hesperidin can stimulate TNFα secretion

in numerous cancer cells, including A549 and LAN-5 cells (95). Overall, hesperidin has

exhibited significant anticancerous potential through the suppression of the production of various

inflammatory molecules involved in cancer development and progression (Figure 2).

6. Anticancer efficacy of hesperidin analogs, nanoformulations, and metabolites

The anticancerous properties of bioactive phytocompounds can be significantly improved by

various nanoparticle formulations as well as enzymatic and chemical modifications (96). Due to

its poor bioavailability and low solubility, access of hesperidin into cells for its pharmacological

properties remains limited. Numerous drug delivery approaches, including microencapsulation,

enzymatic modification, and complexation have been recommended for enhancing its targeted

delivery, solubility and bioavailability thereby improving its anticancer, and biological activity

(97). According to Ali et al., 2019 (98), modified nanohesperidin exhibited higher growth

inhibitory potential and better cytocompatibility than the native hesperidin in breast cancer cells

by modulating the apoptotic pathway of Caspase-3 and p53. Research conducted by Byun et al.,
2019 explained that hesperidin irradiated with 150 kGy gamma has also resulted in higher

growth inhibition of B16BL6 melanoma cells and lung metastasis in C57BL/6 mice in

comparison to non-irradiated hesperidin (99). Balakrishnan et al., 2020 (100) reported that

formulated hesperidin nanoparticles were more competent in comparison to native hesperidin in

Hep-2 cells. Altogether, hesperidin could be considered a safe and economical drug candidate for

the better management of cancer treatments.

7. Conclusion

Increasing cancer incidence, costlier anticancer drugs with their severe side effects, multidrug

resistance are some of the major challenges in cancer therapeutics. Furthermore, various

chemotherapeutic drugs fail to manage numerous carcinomas. There has been mounting demand

for phytocompounds as potent anticancer drugs because they are easily available with low or

minimal toxicity. It is evident from our review that hesperidin could be a promising anti-cancer

agent. Combining chemotherapeutic drugs with phytocompounds would be a better approach for

the treatment of cancer. In combination, hesperidin usually does not affect the activity of chemo

drugs, but somewhat increases the drug-induced cytotoxicity in organ tissues. Hesperidin has

exhibited both chemopreventive and chemotherapeutic effects against numerous carcinomas.

Hesperidin could affect several cell signaling pathways associated with cancer progression.

Several approaches including the combination of hesperidin with other phytochemicals,

chemotherapeutic drugs, irradiation, and nanoformulations could be utilized further to improve

its bioavailability and efficacy as an anticancer agent. Nevertheless, the antitumor efficacy of

hesperidin has been elucidated only in preclinical studies including in vivo and in vitro cancer

models. However respect to limited clinical studies related to pharmacological properties of

hesperidin, it is quite difficult to draw a clear picture about the effective dose consideration for
cancer management in the human body. Therefore, further studies are still needed to elucidate

the best effective doses for future clinical trials in cancer patients to validate the candidature of

hesperidin as one of the promising and effective therapeutic strategies for cancer management.

Sources of Support

This research received no external funding

Declaration of competing interest

All authors declare no conflict of interest.

Acknowledgements

The authors thank the management of Noida Institute of Engineering & Technology for

providing an opportunity to complete the review.

References

[1] Li C, Schluesener H. Health-promoting effects of the citrus flavanone hesperidin. Crit Rev
Food Sci Nutr 2017;57:613-631. doi: 10.1080/10408398.2014.906382.

[2] Ferreira de Oliveira JMP, Santos C, Fernandes E. Therapeutic potential of hesperidin and its
aglycone hesperetin: Cell cycle regulation and apoptosis induction in cancer models.
Phytomedicine 2020;73:152887. doi: 10.1016/j.phymed.2019.152887.

[3] Zaidun NH, Thent ZC, Latiff AA. Combating oxidative stress disorders with citrus flavonoid:
naringenin. Life Sci 2018;208:111-122. doi: 10.1016/j.lfs.2018.07.017.

[4] Grosso G, Godos J, Lamuela-Raventos R, Ray S, Micek A, Pajak A, Sciacca S, D'Orazio N,


Del Rio D, Galvano F. A comprehensive meta-analysis on dietary flavonoid and lignan intake
and cancer risk: Level of evidence and limitations. Mol Nutr Food Res. 2017;61. doi:
10.1002/mnfr.201600930.

[5] Tejada S, Pinya S, Martorell M, Capó X, Tur JA, Pons A, et al. Potential anti-inflammatory
effects of hesperidin from the genus Citrus. Curr Med Chem 2018;25:4929-4945. doi:
10.2174/0929867324666170718104412.

[6] Stanisic D, Liu LHB, Dos Santos RV, Costa AF, Durán N, Tasic L. New sustainable process
for hesperidin isolation and anti-ageing effects of hesperidin nanocrystals. Molecules
2020;25:4534. doi: 10.3390/molecules25194534.
[7] Guo X, Li K, Guo A, Li E. Intestinal absorption and distribution of naringin, hesperidin, and
their metabolites in mice. J Funct Foods 2020; 74:104158. doi:10.1016/j.jff.2020.104158.

[8] Ali AM, Gabbar MA, Abdel-Twab SM, Fahmy EM, Ebaid H, Alhazza IM, et al. Antidiabetic
potency, antioxidant effects, and mode of actions of Citrus reticulata fruit peel hydroethanolic
extract, hesperidin, and quercetin in nicotinamide/streptozotocin-induced wistar diabetic rats.
Oxid Med Cell Longev 2020;2020:1730492. doi: 10.1155/2020/1730492.

[9] Chikara S, Nagaprashantha LD, Singhal J, Horne D, Awasthi S, Singhal SS. Oxidative stress
and dietary phytochemicals: role in cancer chemoprevention and treatment. Cancer Lett
2018;413:122-134. doi: 10.1016/j.canlet.2017.11.002.

[10] Garg A, Garg S, Zaneveld LJ, Singla AK. Chemistry and pharmacology of the Citrus
bioflavonoid hesperidin. Phytother Res 2001;15:655-69. doi: 10.1002/ptr.1074.

[11] du Preez BV, de Beer D, Joubert E. By-product of honeybush (Cyclopia aculata) tea
processing as source of hesperidin-enriched nutraceutical extract. Ind Crops Prod 2016; 87:132-
141. doi:10.1016/j.indcrop.2016.04.012

[12] Soares MS, da Silva DF, Forim MR, da Silva MF, Fernandes JB, Vieira PC, et al.
Quantification and localization of hesperidin and rutin in Citrus sinensis grafted on C. limonia
after Xylella fastidiosa infection by HPLC-UV and MALDI imaging mass spectrometry.
Phytochemistry 2015;115:161-70. doi: 10.1016/j.phytochem.2015.02.011.

[13] Alam P, Alam A, Anwer MK, Alqasoumi SI. Quantitative estimation of hesperidin by
HPTLC in different varieties of citrus peels. Asian Pac J Trop Biomed 2014;4:262-266. doi:
10.12980/APJTB.4.2014C1007.

[14] Jokić S, Šafranko S, Jakovljević M, Cikoš AM, Kajić N, Kolarević F, et al. Sustainable
green procedure for extraction of hesperidin from selected croatian mandarin peels. Processes
2019;7:469. doi:10.3390/pr7070469.

[15] Najafian S, Moradi M, Sepehrimanesh M. Polyphenolic contents and antioxidant activities


of two medicinal plant species, Mentha piperita and Stevia rebaudiana, cultivated in Iran. Comp
Clin Path 2016;25:743-747. doi:10.1007/s00580-016-2258-5.

[16] Gu H, Chen F, Zhang Q, Zang J. Application of ionic liquids in vacuum microwave-assisted


extraction followed by macroporous resin isolation of three flavonoids rutin, hyperoside and
hesperidin from Sorbus tianschanica leaves. J Chromatogr B Analyt Technol Biomed Life Sci
2016;1014:45-55. doi: 10.1016/j.jchromb.2016.01.045.

[17] Hamdan DI, Mahmoud MF, Wink M, El-Shazly AM. Effect of hesperidin and
neohesperidin from bittersweet orange (Citrus aurantium var. bigaradia) peel on indomethacin-
induced peptic ulcers in rats. Environ Toxicol Pharmacol 2014;37:907-15. doi:
10.1016/j.etap.2014.03.006.

[18] Damián-Reyna AA, González-Hernández JC, Maya-Yescas R, de Jesús Cortés-Penagos C,


Del Carmen Chávez-Parga M. Polyphenolic content and bactericidal effect of Mexican Citrus
limetta and Citrus reticulata. J Food Sci Technol 2017;54:531-537. doi: 10.1007/s13197-017-
2498-7.

[19] Man MQ, Yang B, Elias PM. Benefits of hesperidin for cutaneous functions. Evid Based
Complement Alternat Med 2019;2019:2676307. doi: 10.1155/2019/2676307.

[20] Nectoux AM, Abe C, Huang SW, Ohno N, Tabata J, Miyata Y, et al. Absorption and
metabolic behavior of hesperidin (rutinosylated hesperetin) after single oral administration to
Sprague-Dawley rats. J Agric Food Chem 2019;67:9812-9819. doi: 10.1021/acs.jafc.9b03594.

[21] Kawabata K, Yoshioka Y, Terao J. Role of intestinal microbiota in the bioavailability and
physiological functions of dietary polyphenols. Molecules 2019;24:370. doi:
10.3390/molecules24020370.

[22] Bagwe-Parab S, Kaur G, Buttar HS, Tuli HS. Absorption, metabolism, and disposition of
flavonoids and their role in the prevention of distinctive cancer types. In: Singh Tuli H. (eds)
Current aspects of flavonoids: their role in cancer treatment. Springer, Singapore. 2019, pp 125-
137. doi:10.1007/978-981-13-5874-6_6

[23] Jiang J, Yan L, Shi Z, Wang L, Shan L, Efferth T. Hepatoprotective and anti-inflammatory
effects of total flavonoids of Qu Zhi Ke (peel of Citrus changshan-huyou) on non-alcoholic fatty
liver disease in rats via modulation of NF-κB and MAPKs. Phytomedicine 2019;64:153082. doi:
10.1016/j.phymed.2019.153082

[24] Kobayashi S, Tanabe S, Sugiyama M, Konishi Y. Transepithelial transport of hesperetin and


hesperidin in intestinal Caco-2 cell monolayers. Biochim Biophys Acta 2008;1778:33-41. doi:
10.1016/j.bbamem.2007.08.020.

[25] Nagula RL, Wairkar S. Recent advances in topical delivery of flavonoids: a review. J
Control Release 2019;296:190-201. doi: 10.1016/j.jconrel.2019.01.029.

[26] Stevens Y, Rymenant EV, Grootaert C, Camp JV, Possemiers S, Masclee A, et al. The
intestinal fate of citrus flavanones and their effects on gastrointestinal health. Nutrients
2019;11:1464. doi: 10.3390/nu11071464.

[27] Borges G, Lean ME, Roberts SA, Crozier A. Bioavailability of dietary (poly) phenols: a
study with ileostomists to discriminate between absorption in small and large intestine. Food
Funct 2013;4:754-762. doi: 10.1039/c3fo60024f.
[28] de Araújo ME, Franco YE, Alberto TG, Sobreiro MA, Conrado MA, Priolli DG, et al.
Enzymatic de-glycosylation of rutin improves its antioxidant and antiproliferative activities.
Food Chem 2013;141:266-73. doi: 10.1016/j.foodchem.2013.02.127.

[29] De Souza VT, De Franco ÉP, De Araújo ME, Messias MC, Priviero FB, Frankland Sawaya
AC, et al. Characterization of the antioxidant activity of aglycone and glycosylated derivatives of
hesperetin: an in vitro and in vivo study. J Mol Recognit 2016;29:80-87. doi: 10.1002/jmr.2509.

[30] Murota K, Nakamura Y, Uehara M. Flavonoid metabolism: the interaction of metabolites


and gut microbiota. Biosci Biotechnol Biochem 2018;82:600-610. doi:
10.1080/09168451.2018.1444467.

[31] Wu L, Kang A, Shan C, Chai C, Zhou Z, Lin Y, et al. LC-Q-TOF/MS-oriented systemic


metabolism study of pedunculoside with in vitro and in vivo biotransformation. J Pharm Biomed
Anal. 2019;175:112762. doi: 10.1016/j.jpba.2019.07.010.

[32] Roohbakhsh A, Parhiz H, Soltani F, Rezaee R, Iranshahi M. Neuropharmacological


properties and pharmacokinetics of the citrus flavonoids hesperidin and hesperetin—a mini-
review. Life Sci 2014;113:1-6. doi: 10.1016/j.lfs.2014.07.029.

[33] Wu Z, Shen C, Van Den Hengel A. Wider or deeper: revisiting the resnet model for visual
recognition. Pattern Recognit 2019;90:119-133. doi:10.1016/j.patcog.2019.01.006.

[34] van der Woude H, Boersma MG, Vervoort J, Rietjens IM. Identification of 14 quercetin
phase II mono-and mixed conjugates and their formation by rat and human phase II in vitro
model systems. Chem Res Toxicol 2004;17:1520-1530. doi: 10.1021/tx049826v.

[35] Bredsdorff L, Nielsen IL, Rasmussen SE, Cornett C, Barron D, Bouisset F, et al.
Absorption, conjugation and excretion of the flavanones, naringenin and hesperetin from α-
rhamnosidase-treated orange juice in human subjects. Br J Nutr 2010;103:1602-1609. doi:
10.1017/S0007114509993679.

[36] Cuevas-Sierra A, Ramos-Lopez O, Riezu-Boj JI, Milagro FI, Martinez JA. Diet, gut
microbiota, and obesity: links with host genetics and epigenetics and potential applications. Adv
Nutr 2019;10:S17-S30. doi: 10.1093/advances/nmy078.

[37] Vijayvergia U, Bandyopadhayaya S, Mandal CC. Biphasic effects of phytochemicals and


their relevance to cancer therapeutics. In: Kumar M., Sharma A., Kumar P. (eds)
Pharmacotherapeutic botanicals for cancer chemoprevention. Springer, Singapore. 2020. pp 197-
219. doi:10.1007/978-981-15-5999-0_9

[38] Naz H, Tarique M, Ahamad S, Alajmi MF, Hussain A, Rehman MT, et al. Hesperidin‐
CAMKIV interaction and its impact on cell proliferation and apoptosis in the human hepatic
carcinoma and neuroblastoma cells. J Cell Biochem 2019;120:15119-15130. doi:
10.1002/jcb.28774.

[39] Dhanya R, Jayamurthy P. In vitro evaluation of antidiabetic potential of hesperidin and its
aglycone hesperetin under oxidative stress in skeletal muscle cell line. Cell Biochem Funct
2020;38:419-427. doi: 10.1002/cbf.3478.

[40] Magura J, Moodley R, Maduray K, Mackraj I. Phytochemical constituents and in vitro


anticancer screening of isolated compounds from Eriocephalus africanus‡. Nat Prod Res 2020:1-
4. doi: 10.1080/14786419.2020.1744138.

[41] Ansari IA, Akhtar MS. Recent Insights on the Anticancer properties of
flavonoids:prospective candidates for cancer chemoprevention and therapy. In Natural Bio-active
Compounds. Springer, Singapore 2019. pp. 425-448. doi:10.1007/978-981-13-7154-7_13.

[42] Lee J, Kim DH, Kim JH. Combined administration of naringenin and hesperetin with
optimal ratio maximizes the anti-cancer effect in human pancreatic cancer via down regulation of
FAK and p38 signaling pathway. Phytomedicine 2019;58:152762. doi:
10.1016/j.phymed.2018.11.022.

[43] Magura J, Moodley R, Mackraj I. The effect of hesperidin and luteolin isolated
fromEriocephalus africanus on apoptosis, cell cycle and miRNA expression in MCF-7. J Biomol
Struct Dyn 2020;1-10. doi: 10.1080/07391102.2020.1833757.

[44] Kongtawelert P, Wudtiwai B, Shwe TH, Pothacharoen P, Phitak T. Inhibitory effect of


hesperidin on the expression of programmed death ligand (PD-L1) in breast Cancer. Molecules
2020;25:252. doi: 10.3390/molecules25020252.

[45] Tan S, Dai L, Tan P, Liu W, Mu Y, Wang J, et al. Hesperidin administration suppresses the
proliferation of lung cancer cells by promoting apoptosis via targeting the miR-132/ZEB2
signalling pathway. Int J Mol Med 2020;46:2069-2077. doi: 10.3892/ijmm.2020.4756.

[46] Yang Z, Yang H, Dong X, Pu M, Ji F. Hesperidin loaded Zn2+@ SA/PCT nanocomposites


inhibit the proliferation and induces the apoptosis in colon cancer cells (HCT116) through the
enhancement of pro-apoptotic protein expressions. J Photochem Photobiol B 2020;204:111767.
doi: 10.1016/j.jphotobiol.2019.111767

[47] Pandey P, Sayyed U, Tiwari RK, Siddiqui MH, Pathak N, Bajpai P. Hesperidin induces
ROS-mediated apoptosis along with cell cycle arrest at G2/M phase in human gall bladder
carcinoma. Nutr Cancer 2019;71:676-687. doi: 10.1080/01635581.2018.1508732.

[48] Fontana F, Raimondi M, Marzagalli M, Di Domizio A, Limonta P. The emerging role of


paraptosis in tumor cell biology: perspectives for cancer prevention and therapy with natural
compounds. Biochim Biophys Acta Rev Cancer 2020;1873:188338. doi:
10.1016/j.bbcan.2020.188338.

[49] Zhang F, Zhang YY, Sun YS, Ma RH, Thakur K, Zhang JG, et al. Asparanin a from
Asparagus officinalis L. induces G0/G1 cell cycle arrest and apoptosis in human endometrial
carcinoma ishikawa cells via mitochondrial and PI3K/AKT signaling pathways. J Agric Food
Chem 2020;68:213-224. doi: 10.1021/acs.jafc.9b07103.

[50] Kamaraj S, Anandakumar P, Jagan S, Ramakrishnan G, Periyasamy P, Asokkumar S, et al.


Hesperidin inhibits cell proliferation and induces mitochondrial-mediated apoptosis in human
lung cancer cells through down regulation of β-catenin/c-myc. Biocatal Agric Biotechnol
2019;18:101065. doi:10.1016/j.bcab.2019.101065

[51] Kumar V, Chauhan D. Discovery of hesperidin based novel AMPK/mTOR kinase inhibitor
against colorectal cancer cells. Gut Liver 2019;13.

[52] Yu W, Xie X, Yu Z, Jin Q, Wu H.. Mechanism of hesperidin-induced apoptosis in human


gastric cancer AGS cells. Trop J Pharm Res 2019;18: 2363-2369.
http://dx.doi.org/10.4314/tjpr.v18i11.19.

[53] Febriansah R, Putri DD, Sarmoko, Nurulita NA, Meiyanto E, Nugroho AE. Hesperidin as a
preventive resistance agent in MCF-7 breast cancer cells line resistance to doxorubicin. Asian
Pac J Trop Biomed 2014;4:228-33. doi: 10.1016/S2221-1691(14)60236-7.

[54] Wang Y, Liu S, Dong W, Qu X, Huang C, Yan T, et al. Combination of hesperetin and
platinum enhances anticancer effect on lung adenocarcinoma. Biomed Pharmacother
2019;113:108779. doi: 10.1016/j.biopha.2019.108779.

[55] Yunita E, Muflikhasari HA, Ilmawati GP, Meiyanto E, Hermawan A. Hesperetin alleviates
doxorubicin-induced migration in 4T1 breast cancer cells. Future J Pharm
Sci 2020;6:1-9. doi:10.1186/s43094-020-00036-y.

[56] Khamis AAA, Ali EMM, El-Moneim MAA, Abd-Alhaseeb MM, El-Magd MA, Salim EI.
Hesperidin, piperine and bee venom synergistically potentiate the anticancer effect of tamoxifen
against breast cancer cells. Biomed Pharmacother 2018;105:1335-1343. doi:
10.1016/j.biopha.2018.06.105.

[57] Aboismaiel MG, El-Mesery M, El-Karef A, El-Shishtawy MM. Hesperetin upregulates


Fas/FasL expression and potentiates the antitumor effect of 5-fluorouracil in rat model of
hepatocellular carcinoma. Egypt J Basic Appl Sci 2020;7:20-34.
doi:10.1080/2314808X.2019.1707627.

[58] Memariani Z, Abbas SQ, Ul Hassan SS, Ahmadi A, Chabra A. Naringin and naringeninin as
anticancer agents and adjuvants in cancer combination therapy; efficacy and molecular
mechanisms of action, a comprehensive narrative review. Pharmacol Res 2020:105264. doi:
10.1016/j.phrs.2020.105264.

[59] Afshari K, Haddadi NS, Haj-Mirzaian A, Farzaei MH, Rohani MM, Akramian F, et al.
Natural flavonoids for the prevention of colon cancer: a comprehensive review of preclinical and
clinical studies. J Cell Physiol 2019;234:21519-21546. doi: 10.1002/jcp.28777.

[60] Birsu Cincin Z, Unlu M, Kiran B, Sinem Bireller E, Baran Y, Cakmakoglu B. Anti-
proliferative, apoptotic and signal transduction effects of hesperidin in non-small cell lung cancer
cells. Cell Oncol (Dordr) 2015;38:195-204. doi: 10.1007/s13402-015-0222-z.

[61] Donia TI, Gerges MN, Mohamed TM. Amelioration effect of Egyptian sweet orange
hesperidin on ehrlich ascites carcinoma (EAC) bearing mice. Chem Biol Interact 2018;285:76-
84. doi: 10.1016/j.cbi.2018.02.029.

[62] Du GY, He SW, Zhang L, Sun CX, Mi LD, Sun ZG. Hesperidin exhibits in vitro and in vivo
antitumor effects in human osteosarcoma MG-63 cells and xenograft mice models via inhibition
of cell migration and invasion, cell cycle arrest and induction of mitochondrial-mediated
apoptosis. Oncol Lett 2018;16:6299-6306. doi: 10.3892/ol.2018.9439.

[63] Wu D, Zhang J, Wang J, Li J, Liao F, Dong W. Hesperetin induces apoptosis of esophageal


cancer cells via mitochondrial pathway mediated by the increased intracellular reactive oxygen
species. Tumour Biol 2016;37:3451-3459. doi:10.1007/s13277-015-4176-6.

[64] Siddiqi A, Nafees S, Rashid S, Sultana S, Saidullah B. Hesperidin ameliorates


trichloroethylene-induced nephrotoxicity by abrogation of oxidative stress and apoptosis in
wistar rats. Mol Cell Biochem 2015;406:9-20. doi: 10.1007/s11010-015-2400-8.

[65] Saiprasad G, Chitra P, Manikandan R, Sudhandiran G. Hesperidin induces apoptosis and


triggers autophagic markers through inhibition of aurora-A mediated phosphoinositide-3-
kinase/Akt/mammalian target of rapamycin and glycogen synthase kinase-3 beta signalling
cascades in experimental colon carcinogenesis. Eur J Cancer 2014;50:2489-507. doi:
10.1016/j.ejca.2014.06.013.

[66] de Oliveira JR, Camargo SEA, de Oliveira LD. Rosmarinus officinalis L. (rosemary) as
therapeutic and prophylactic agent. J Biomed Sci 2019;26:5. doi: 10.1186/s12929-019-0499-8.

[67] Kabała-Dzik A, Rzepecka-Stojko A, Kubina R, Iriti M, Wojtyczka RD, Buszman E, et al.


Flavonoids, bioactive components of propolis, exhibit cytotoxic activity and induce cell cycle
arrest and apoptosis in human breast cancer cells MDA-MB-231 and MCF-7–a comparative
study. Cell Mol Biol (Noisy-le-grand) 2018;64:1-10. doi:10.14715/cmb/2018.64.8.1.
[68] Lee DH, Park KI, Park HS, Kang SR, Nagappan A, Kim JA, et al. Flavonoids isolated from
Korea Citrus aurantium L. Induce G2/M phase arrest and apoptosis in human gastric cancer AGS
cells. Evid Based Complement Alternat Med 2012;2012:515901. doi: 10.1155/2012/515901.

[69] Devi KP, Rajavel T, Nabavi SF, Setzer WN, Ahmadi A, Mansouri K, et al. Hesperidin: a
promising anticancer agent from nature. Ind Crops Prod 2015;76: 582-589.
doi:10.1016/j.indcrop.2015.07.051

[70] ang JN, Li L, Li LY, Yan Q, Li J, Xu T. Emerging role and therapeutic implication of Wnt
signaling pathways in liver fibrosis. Gene 2018;674:57-69. doi:10.1016/j.gene.2018.06.053.

[71] Cheng X, Xu X, Chen D, Zhao F, Wang W. Therapeutic potential of targeting the Wnt/β-
catenin signaling pathway in colorectal cancer. Biomed Pharmacother 2019;110:473-481. doi:
10.1016/j.biopha.2018.11.082.

[72] Zhang X, Wang L, Qu Y. Targeting the β-catenin signaling for cancer therapy. Pharmacol
Res 2020;160:104794. doi: 10.1016/j.phrs.2020.104794.

[73] Zaghloul RA, Elsherbiny NM, Kenawy HI, El-Karef A, Eissa LA, El-Shishtawy MM.
Hepatoprotective effect of hesperidin in hepatocellular carcinoma: involvement of Wnt signaling
pathways. Life Sci 2017;185:114-125. doi: 10.1016/j.lfs.2017.07.026.

[74] Pfeffer CM, Singh ATK. Apoptosis: a target for anticancer therapy. Int J Mol Sci
2018;19:448. doi: 10.3390/ijms19020448.

[75] Kottaiswamy A, Kizhakeyil A, Padmanaban AM, Mirza FB, Vijay VR, Lee PS, et al. The
citrus flavanone hesperetin induces apoptosis in CTCL cells via STAT3/Notch1/NFκB-mediated
signaling axis. Anticancer Agents Med Chem 2020;20:1459-1468. doi:
10.2174/1871521409666200324110031.

[76] Ahmad ST, Arjumand W, Nafees S, Seth A, Ali N, Rashid S, et al. Hesperidin alleviates
acetaminophen induced toxicity in Wistar rats by abrogation of oxidative stress, apoptosis and
inflammation. Toxicol Lett 2012;208:149-161. doi: 10.1016/j.toxlet.2011.10.023

[77] Xia R, Xu G, Huang Y, Sheng X, Xu X, Lu H. Hesperidin suppresses the migration and


invasion of non-small cell lung cancer cells by inhibiting the SDF-1/CXCR-4 pathway. Life
Sci 2018;201:111-120. doi: 10.1016/j.lfs.2018.03.046.

[78] Cincin ZB, Kiran B, Baran Y, Cakmakoglu B. Hesperidin promotes programmed cell death
by downregulation of nongenomic estrogen receptor signalling pathway in endometrial cancer
cells. Biomed Pharmacother 2018;103:336-345. doi: 10.1016/j.biopha.2018.04.020.

[79] Meng C, Guo Z, Li D, Li H, He J, Wen D, et al. Preventive effect of hesperidin modulates


inflammatory responses and antioxidant status following acute myocardial infarction through the
expression of PPAR‐γ and Bcl‐2 in model mice. Mol Med Rep 2018;17:1261-1268. doi:
10.3892/mmr.2017.7981.

[80] Aubrey BJ, Kelly GL, Janic A, Herold MJ, Strasser A. How does p53 induce apoptosis and
how does this relate to p53-mediated tumour suppression? Cell Death Differ 2018;25:104-113.
doi: 10.1038/cdd.2017.169.

[81] Ren F, Zhang G, Li C, Li G, Cao Y, Sun F. Hesperidin induces mitochondria mediated


intrinsic apoptosis in HPV-positive cervical cancer cells via regulation of E6/p53 expression
2020. doi:10.21203/rs.3.rs-105340/v1

[82] Heo SD, Kim J, Choi Y, Ekanayake P, Ahn M, Shin T. Hesperidin improves motor
disability in rat spinal cord injury through anti-inflammatory and antioxidant mechanism via Nrf-
2/HO-1 pathway. Neurosci Lett 2020;715:134619. doi: 10.1016/j.neulet.2019.134619.

[83] Moloney JN, Cotter TG. ROS signalling in the biology of cancer. Semin Cell Dev Biol
2018;80:50-64. doi: 10.1016/j.semcdb.2017.05.023.

[84] Kim J, Kim J, Bae JS. ROS homeostasis and metabolism: a critical liaison for cancer
therapy. Exp Mol Med 2016;48:e269. doi: 10.1038/emm.2016.119.

[85] Kohan R, Collin A, Guizzardi S, de Talamoni NT, Picotto G. Reactive oxygen species in
cancer: a paradox between pro-and anti-tumour activities. Cancer Chemother Pharmacol
2020;86:1-13. doi: 10.1007/s00280-020-04103-2.

[86] Tavsan Z, Kayali HA. Flavonoids showed anticancer effects on the ovarian cancer cells:
involvement of reactive oxygen species, apoptosis, cell cycle and invasion. Biomed
Pharmacother 2019;116;109004. doi: 10.1016/j.biopha.2019.109004.

[87] Antunes MS, Ladd FV, Ladd AA, Moreira AL, Boeira SP, Souza LC. Hesperidin protects
against behavioral alterations and loss of dopaminergic neurons in 6-OHDA-lesioned mice: the
role of mitochondrial dysfunction and apoptosis. Metab Brain Dis 2020;36:1-15. doi:
10.1007/s11011-020-00618-y.

[88] Aggarwal V, Tuli HS, Thakral F, Singhal P, Aggarwal D, Srivastava S, et al. Molecular
mechanisms of action of hesperidin in cancer: recent trends and advancements. Exp Biol Med
(Maywood) 2020;245:486-497. doi: 10.1177/1535370220903671.

[89] Parhiz H, Roohbakhsh A, Soltani F, Rezaee R, Iranshahi M. Antioxidant and anti-


inflammatory properties of the citrus flavonoids hesperidin and hesperetin: an updated review of
their molecular mechanisms and experimental models. Phytother Res 2015;29:323-31. doi:
10.1002/ptr.5256.

[90] Sooro MA, Zhang N, Zhang P. Targeting EGFR-mediated autophagy as a potential strategy
for cancer therapy. Int J Cancer 2018;143:2116-2125. doi: 10.1002/ijc.31398.
[91] Li S, Tan HY, Wang N, Cheung F, Hong M, Feng Y. The potential and action mechanism
of polyphenols in the treatment of liver diseases. Oxid Med Cell 2018; 2018:8394818 doi:
10.1155/2018/8394818.

[92] Mo’men YS, Hussein RM, Kandeil MA. Involvement of PI3K/Akt pathway in the
protective effect of hesperidin against a chemically induced liver cancer in rats. J Biochem Mol
Toxicol 2019;33:e22305. doi: 10.1002/jbt.22305.

[93] Saltarella I, Frassanito MA, Lamanuzzi A, Brevi A, Leone P, Desantis V, et al. "Homotypic
and heterotypic activation of the notch pathway in multiple myeloma–enhanced angiogenesis: a
novel therapeutic target?." Neoplasia. 2019;21:93-105. doi: 10.1016/j.neo.2018.10.011.

[94] Birch AM, Cheung J, Oluwadare C, Burton J, Huang W. Alterations in the expression of
transcription factors PPARγ and NFκB in the brain of models of chronic pain. Biochem
Pharmacol 2016;5:177-0501. doi: 10.4172/2177-0501.1000217.

[95] Fantini M, Benvenuto M, Masuelli L, Frajese GV, Tresoldi I, Modesti A, et al. In vitro and
in vivo antitumoral effects of combinations of polyphenols, or polyphenols and anticancer drugs:
perspectives on cancer treatment. Int J Mol Sci 2015;16:9236-82. doi: 10.3390/ijms16059236.

[96] Bonferoni MC, Rossi S, Sandri G, Ferrari F. Nanoparticle formulations to enhance tumor
targeting of poorly soluble polyphenols with potential anticancer properties. Semin Cancer Biol.
2017;46:205-214. doi: 10.1016/j.semcancer.2017.06.010.

[97] Roberts TC, Langer R, Wood MJ. Advances in oligonucleotide drug delivery. Nat Rev Drug
Discov 2020;19:1-22. doi: 10.1038/s41573-020-0075-7.

[98] Ali SH, Sulaiman GM, Al-Halbosiy MM, Jabir MS, Hameed AH. Fabrication of hesperidin
nanoparticles loaded by poly lactic co-glycolic acid for improved therapeutic efficiency and
cytotoxicity. Artif Cells Nanomed Biotechnol 2019;47:378-394. doi:
10.1080/21691401.2018.1559175

[99] Byun EB, Kim HM, Song HY, Kim WS. Hesperidin structurally modified by gamma
irradiation induces apoptosis in murine melanoma B16BL6 cells and inhibits both subcutaneous
tumor growth and metastasis in C57BL/6 mice. Food Chem Toxicol 2019;127:19-30. doi:
10.1016/j.fct.2019.02.042.

[100] Balakrishnan K, Casimeer SC, Ghidan AY, Ghethan FY, Venkatachalam K, Singaravelu
A. Bioformulated hesperidin-loaded PLGA nanoparticles counteract the mitochondrial-mediated
intrinsic apoptotic pathway in cancer cells. J Inorg Organomet Polym Mater 2021;31:331-343.
doi:10.1007/s10904-020-01746-9.
Figure 1: Chemical structure of (A) Hesperidin and (B) Hesperitin. PubChem database has been

used to retrieve the chemical structure of these two potential phytocompounds.

Figure 2: Mechanism associated with the anticancer potential of hesperidin and its associated

targets
Table 1: Occurrence of hesperidin in plants

Plant species Hesperidin content References


Citrus limonia 15.4 g kg-1 (12)
Grape fruits (Citrus paradisi) 11.15 % w/w
Mosambi (Citrus limetta) 3.79% w/w
(13)
Lemon (Citrus lemon) 1.72% w/w
Orange (Citrus sinensis) 1.29% w/w
Croatian mandarin peels 31.42 mg/g (14)
Mentha piperita L. 504.2 mg/L (15)
Stevia rebaudiana 493.4 mg/L
Sorbus tianschanica leaves 0.48 mg/g (16)
Citrus aurantium var. bigaradia 2.66 (w/w) % (17)
0.62 mg/g (bagasse)
Citrus limetta 27.3 mg/L (juice) (18)
0.04 mg/g (seed)
0.98 mg/g (bagasse)
Citrus reticulate 21.7 mg/L (juice) (18)
0.09 mg/g (seed)
Citrus sinensis L 23.0 µg/mL (19)
Table 2: In vitro studies of anticancerous effects of hesperidin

Hesperidin
Cancer cell lines Anticancerous potential References
concentration
Maximizes the anticancerous effects in
human pancreatic cancer via down
Panc-1 cells 1-20µM (42)
regulation of p38 and FAK signaling
pathways
Significant apoptotic cells accumulation in
G0/G1 cell cycle phases, apoptotic
MCF-7 human
induction via extrinsic and intrinsic
breast cancer cell 20-40µg/mL (43)
pathways, down regulation of anti
line
apoptotic gene (Bcl-2), and upregulation of
pro-apoptotic gene (Bax).
Decreased PD-L1 expression via inhibition
MDA-MB231
of NF-κB and PI3K/Akt pathway.
triple negative
Hesperidin prevented metastasis
breast 10 to 50 µM (44)
phenotypes via reduction in PD-L1
adenocarcinoma
(immune checkpoint inhibitor) expression
cancer cell line
PD-L1
Hesperidin administration resulted in
alleviation of non small cell lung cancer
A549 and H460
1-2.5 µM by antiproliferative effect and apoptotic (45)
cells
induction in NSCLC cells via the miR-
132/ZEB2 pathway.
Hesperidin loaded nanocomposites
administration induced apoptosis via
HCT116 cells 5-15 µg/mL enhanced ROS generation leading to (46)
excessive mitochondrial membrane
permeability in HCT116 cancer cells.
Apoptosis inducing potential via enhanced
reactive oxygen species generation, loss of
MMP (mitochondrial membrane potential)
Human gall
25-200 µM cell cycle arrest, and caspase-3 activation (47)
bladder carcinoma
in the primary cells generated from
surgically removed cancerous gall bladder
tissues.
Hesperidin inhibits proliferation through
apoptosis mediated by endoplasmic
HeLa cell 20-100 µM (47)
reticulum stress pathways and cell cycle
arrest
Hesperidin treatment induces paraptosis
HepG2 Cells 0.1-2 mM like cell death via phosphorylation of (48)
ERK1/2
Reduces cell viability, and induces
apoptosis via activated protein expression
A2780 cells 0.1-10 µM (49)
of caspase-3 through cytochrome c
signaling pathways
Inhibition of cell proliferation via down
regulation of β-catenin, PCNA, and c-myc
protein expression. apoptotic induction via
A549 cells 6.25-100 µM (50)
enhanced levels of APAF-1, cytochrome c,
DNA fragmentation, and caspase-3 in
hesperidin treated cells
Exhibited cytotoxic, anti-proliferative and
HT-29 and HCT-
58.45 μmol/L pro-apoptotic effect via AMPK/mTOR (51)
116
signaling pathway
Induces mitochondrial dependent apoptosis
Gastric cancer
1-100 μM by increasing ROS levels and MAPK (51)
(GC) AGS cells
signaling pathway regulation in AGS cells
Abbreviations: FAK, focal adhesion kinase; PD-L1, programmed death-ligand 1; NF-κB,
nuclear factor kappa light chain enhancer of activated B cells; PI3K, phosphatidylinositol 3-
kinase; Akt, protein kinase B; NSCLC, non-small-cell lung carcinoma; ZEB2, zinc finger E-box
Binding homeobox 2; ROS, reactive oxygen species; MMP, mitochondrial membrane potential;
ERK1, extracellular-signal-regulated kinase PCNA, proliferating cell nuclear antigen; c-myc,
cellular myelocytomatosis oncogene; AMPK, AMP-activated protein kinase; mTOR, mammalian
target of rapamycin complex 1; AGS, adenocarcinoma gastric cell line

Table 3: Studies on anticancerous potential of combinations of hesperidin and

chemotherapeutic drug/irradiation

Cancer
Combinatorial treatment Synergistic effects References
cells/Model
Hesperidin in combination with
gemicitabine synergistically
Gall bladder
Hesperidin+Gemicitabine reduced the proliferation of GBC (47, 52)
carcinoma
cells in a more significant way in
comparison to hesperidin alone
Combinatorial treatment of
MCF–7 breast
hesperidin and doxorubicin
cancer cells Hesperidin+Doxorubicin (53)
resulted in the inhibition of Pgp
line
expression in MCF-7/Dox cells
Combined treatment of
hesperetin and platinum
exhibited more significant
A549 cells Hesperetin + Platinum (54)
inhibitory effect on lung cancer
progression in compared with
single drug treatment.
Combined treatment induced
4T1 murine apoptosis and cell cycle arrest in
metastatic G2/M phase. Combined
Hesperitin+Doxorubicin (55)
breast cancer treatment inhibited MMP-9
cells expression and migration and in
4T1 cells
Synergistically enhance the
anticancerous effect of
MCF7 and Tamoxifen+Hesperidin+Pip
tamoxifen and can be further (56)
T47D cells erine
utilized as safe adjuvant or
vehicle to tamoxifen
Rat model
Significant increase in apoptotic
(Sprague-
effects induced by 5-FU via
Dawley) of Hesperetin+5-fluorouracil
remarkable elevation in gene (57)
HCC (5-FU)
expression of FasL, Fas,
(hepatocellular
caspase-3 and caspase-8
carcinoma)
Abbreviations: GBC, gall bladder cancer; MCF-7, michigan cancer foundation-7; Dox,
Doxorubicin; FasL, fas-ligand; MMP-9, matrix metallopeptidase 9; T47D, Human Breast Cancer
Cells; 4T1, breast cancer cell line derived from the mammary gland tissue of a mouse BALB/c
strain; Pgp, P-glycoprotein
Table 4: In vivo anticancerous studies of Hesperidin

Animal cancer Hesperidin Outcome and possible mode of


References
BALB/c mice model treatment action
Apoptosis induction in tumor cells
Ehrlich ascites
100 mg/kg via DNA fragmentation, down-
carcinoma (EAC) (60)
regulation of antiapoptotic gene
bearing Balb/C mice
Bcl2 and Caspase 3 stimulation
Displayed in vivo and in vitro
apoptotic and antitumor effects in
Bone cancer
MG-63 cells via mediating cell
MG-63 xenograft/ 80 mg/kg (61)
cycle arrest, cell migration
BALB/c mice
inhibition and mitochondrial
membrane mediated apoptosis.
Inhibited esophageal cancer cell
Esophageal cancer
growth and invasion by inhibiting
Eca109 xenograft/ 90 mg/kg (62)
or downregulating the PI3K/AKT
BALB/c-nu/nu mice
signaling pathway.
Cell apoptosis induction in
Esophageal cancer
esophageal cancer cells via
Eca109 xenograft/ 60 mg/kg (63)
mitochondrial mediated intrinsic
BALB/c-nu/nu mice
pathway by ROS accumulation.
hesperidin to be a potent
chemopreventive agent against
DEN-induced Kidney renal carcinogenesis possibly by
200 mg/kg (64)
Cancer/ Wistar rats virtue of its antioxidant properties
and by modulation of multiple
molecular pathways.
Hesperidin treatments combat with
AOM-induced the inflammation via
Colorectal cancer/ 25 mg/kg downregulation of NF-κB and its (65)
Swiss albino mice related target molecules COX-2 and
iNOS.
Cyclophosphamide
Hesperidin treatment resulted in an
induced Premature
80 mg/kg effective treatment for premature (66)
ovarian failure (POF)/
ovarian failure in rats.
rat model
Abbreviations: Bcl-2, B-cell lymphoma 2; BALB, bagg albino; COX-2, cyclooxygenase-2;
iNOS, inducible nitric oxide synthase; DEN, diethylnitrosamine; AOM, azoxymethane; NF-κB,
nuclear factor kappa light chain enhancer of activated B cells; ROS, reactive oxygen species
Figure 1.
Figure 2.
Graphical abstract

You might also like