Svoboda | Graniru | BBC Russia | Golosameriki | Facebook
Next Article in Journal
Distinct Cerebrospinal Fluid Lipid Signature in Patients with Subarachnoid Hemorrhage-Induced Hydrocephalus
Next Article in Special Issue
Mometasone Furoate Inhibits the Progression of Head and Neck Squamous Cell Carcinoma via Regulating Protein Tyrosine Phosphatase Non-Receptor Type 11
Previous Article in Journal
The Role of miRNA in Renal Fibrosis Leading to Chronic Kidney Disease
Previous Article in Special Issue
Development of a Novel Anti-CD44 Variant 7/8 Monoclonal Antibody, C44Mab-34, for Multiple Applications against Oral Carcinomas
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Comprehensive Review of Natural Products as Therapeutic or Chemopreventive Agents against Head and Neck Squamous Cell Carcinoma Cells Using Preclinical Models

by
Yoon Xuan Liew
1,
Lee Peng Karen-Ng
1,* and
Vui King Vincent-Chong
2,*
1
Oral Cancer Research & Coordinating Centre (OCRCC), Faculty of Dentistry, University of Malaya, Kuala Lumpur 50603, Malaysia
2
Department of Oral Oncology, Roswell Park Comprehensive Cancer Center, Buffalo, NY 14263, USA
*
Authors to whom correspondence should be addressed.
Biomedicines 2023, 11(9), 2359; https://doi.org/10.3390/biomedicines11092359
Submission received: 16 June 2023 / Revised: 9 August 2023 / Accepted: 18 August 2023 / Published: 23 August 2023
(This article belongs to the Special Issue Oral Cancer: From Pathophysiology to Novel Therapeutic Approaches)

Abstract

:
Head and neck squamous cell carcinoma (HNSCC) is a type of cancer that arises from the epithelium lining of the oral cavity, hypopharynx, oropharynx, and larynx. Despite the advancement of current treatments, including surgery, chemotherapy, and radiotherapy, the overall survival rate of patients afflicted with HNSCC remains poor. The reasons for these poor outcomes are due to late diagnoses and patient-acquired resistance to treatment. Natural products have been extensively explored as a safer and more acceptable alternative therapy to the current treatments, with numerous studies displaying their potential against HNSCC. This review highlights preclinical studies in the past 5 years involving natural products against HNSCC and explores the signaling pathways altered by these products. This review also addresses challenges and future directions of natural products as chemotherapeutic and chemoprevention agents against HNSCC.

Graphical Abstract

1. Introduction

Head and neck cancer (HNC) represents cancers occurring in the head and neck region, which includes the lip and oral cavity, nasal cavity, larynx, and pharynx [1]. More than 90% of HNCs are head and neck squamous cell carcinomas (HNSCCs), which account for approximately 300,000 deaths and 500,000 new cases worldwide annually [2]. In the United States, HNSCC has been diagnosed as one of the top 10 leading cancers in men in 2022 [2,3,4]. The low survival rate of HNSCC has been proposed to be associated with cancer recurrence, distant metastases, the progression of second primary cancers, and resistance to chemo/radiotherapy [5,6], making it a public health issue that compromises patients’ quality of life.
Tobacco smoking, excessive alcohol drinking, betel quid chewing, and high-risk human papillomavirus (HPV) have been documented as risk factors for HNSCC [7,8,9,10]. Clinical intervention of HNSCC often takes place at advanced stages of the disease due to late diagnoses, especially among individuals with a lower socioeconomic background [11,12]. The most common clinical interventions for HNSCC are surgery, radiotherapy, chemotherapy, or combined therapy, which causes numerous side effects during the treatment of this disease [11]. However, even with a successful clinical intervention, approximately 30% of these patients treated at advanced stages of the disease develop recurrent locoregional or second primary cancers, with the onset of chemo- or radio-resistance [13,14,15]. Notably, these clinical interventions are only effective on a limited subgroup of HNSCC patients and often result in additional morbidities.
Cetuximab, an epidermal growth factor receptor (EGFR)-targeting monoclonal antibody, was the first molecular-targeted drug approved by the U.S. Food and Drug Administration (FDA) as a chemotherapeutic agent for HNSCC in 2006 [5,16]. A retrospective study in Japan reported a total effective rate of 57.1%, a median progression-free survival (PFS) of 5.5 months, and an overall survival (OS) of 8.0 months with cetuximab in locally advanced HNSCC, while a total effective rate of 60.0%, a PFS of 3.8 months, and an OS of 5.8 months were reported in distantly metastatic HNSCC [17]. However, the chemoresistance ability of certain mutated cancer cell types, such as EGFRvIII, has shown resistance toward cetuximab [18]. In recent years, both pembrolizumab and nivolumab, which act as anti-programmed cell death receptor 1 (PD-1) immunotherapy drugs, were approved by the FDA for recurrent or metastatic HNSCC treatment [19,20]. However, most patient previously exposed to anti-PD-1 monoclonal antibody develop acquired resistance to immunotherapeutic drugs, making it difficult to treat recurrent or metastatic cancers [21]. There is certainly an urgent need for alternative therapeutic agents to overcome the acquired resistance of HNSCC to the standard of care. Because of the challenges faced using immunotherapies, there is a great need for therapeutic agents that can work effectively as chemoprevention or to enhance the effectiveness of chemotherapeutic agents when used in combination to kill cancer cells.
Natural products are compounds naturally found in natural resources such as plants, which possess biological activities [22]. In recent years, natural products have been widely reported for their chemotherapeutic and chemoprevention properties against HNSCC due to their low cytotoxicity, efficacy against cancers, availability, and low cost [12,22]. According to U.S. National Cancer Institute, chemoprevention is defined as the use of certain drugs or other substances to help lower a person’s risk of developing cancer or keep it from coming back [23]. The beneficial properties of natural products are consistent and could overcome the challenges with current treatment such as acquired chemoresistance, cytotoxicity against normal cells, and expensive therapy. Numerous studies have been carried out with various natural products on preclinical models of HNSCC, including various HNSCC cell lines and xenograft or carcinogen-induced tumor animal models (Figure 1). For example, psorachromene, a flavonoid found in Psoralea corylifolia, which has been used in traditional Chinese medicine (TCM) and Ayurveda, had shown therapeutic effects against HNSCC via regulation of the EGFR signaling pathways and other carcinogenesis-related signaling pathway, making it a strong candidate to act as a chemotherapeutic agent [24]. Other natural products, such as calcitriol, have been reported to possess chemoprevention properties against carcinogen-induced HNSCC in animal models [25]. Therefore, natural products have shown potential as candidates for further exploration as adjuvant, neoadjuvant chemotherapy, and chemoprevention agents in HNSCC. However, we acknowledge that, in many cases, these compounds are studied only at the beginning state in animal models. More studies on safety and efficacy are needed to improve the therapeutic potential in patients, which include clinical trials.
The activation of complex signaling pathways, such as phosphoinositide 3-kinase/protein kinase B/mammalian target of rapamycin (PI3K/Akt/mTOR) and mitogen-activated protein kinase/extracellular signal-regulated protein kinase (MAPK/ERK), is important for tumor development, cell survival, and angiogenesis, while epithelial to mesenchymal transition (EMT) signaling leads to tumor invasion and migration [26,27]. Thus, the inhibition of these signaling pathways could lead to cell death, inhibiting tumor development and metastasis [27]. Notwithstanding the extensive knowledge at the molecular level of tumor-associated signaling pathways, the chemotherapeutic agents targeting the oncogenic pathways are limited [28]. To date, Crooker et al. and Rahman et al. have described the use of natural products as chemoprevention agents against HNSCC [29,30]. Several natural products, such as vitamin A, green tea extract, and curcumin, which have shown promising results in preclinical studies, were reported to show toxicity with limited bioavailability in clinical studies [31,32,33,34]. Vitamin A has been shown by Shin et al. [31] and Papadimitrakopoulou et al. [33] to induce toxicity via oral mucosa and lip inflammation, conjunctivitis, skin reactions, fatigue, and joint and muscle pain. On the other hand, poor oral absorption of green tea extract and curcumin has been observed, which limits the bioavailability of both natural products [33,34]. Other emerging chemotherapeutic phytochemicals or herbal derivatives against HNC have recently been described by Aggarwal et al. [12]. However, several new phytochemicals against HNSCC, such as actein, calcitriol, and psorachromene, were not fully addressed. Therefore, in this review, the comprehensive mechanisms of various natural products showing significant preclinical results in HNSCC models for the past 5 years are discussed.

2. Chemotherapeutic Properties of Natural Products on Essential Pathways for HNSCC

2.1. PI3K/Akt/mTOR Pathway

It was previously reported that the activation of the PI3K/Akt/mTOR pathway has been observed in approximately 90% of HNSCC cases, making it a prominent target for treatments of HNSCC [35,36]. The activation of PI3K/Akt/mTOR signaling also plays an important role in HNSCC chemotherapy and radiotherapy resistances, as the inhibition of the signaling pathway has shown positive effects on tumor proliferation and radiotherapy sensitization in preclinical studies [37,38]. The PI3K/Akt/mTOR pathway is activated when a ligand-like growth factor binds with the receptor tyrosine kinase (RTK), leading to the activation of PI3K, which, in turn, partially activates Akt [39]. Then, mTORC2 is required to completely activate Akt via phosphorylation, leading to the activation of multiple proteins involved in cell proliferation and motility [40]. Inhibition of Akt formation will therefore limit the expression of oncoprotein. Various natural products, namely actein, salicylate, tanshinone IIA, xanthohumol, fucoidan, honokiol, ilimaquinone, nimbolide, and cinnamaldehyde, have led to the downregulation of Akt expression and phosphorylation of Akt (p-Akt) using HNSCC preclinical models [41,42,43,44,45,46,47,48,49]. The downregulation of mTOR expression in HNSCC preclinical models has been shown by salicylate, honokiol, cinnamaldehyde, and Seco-A-ring oleanane, with PI3K also limiting the phosphorylation of Akt [42,45,48,50]. The expression of glycogen synthase kinase-3β (GSK-3β) plays an important role in regulating growth, cell cycle progression, apoptosis, and cancer cell invasion in HNSCC [51,52]. The inactivation of GSK-3β via phosphorylation has previously shown significant inhibition in cancer cell growth and migration in HNSCC [51]. Preclinical studies with nimbolide have also shown the upregulation of phosphorylated GSK-3β, inactivating GSK-3β and thus inhibiting cell proliferation [47]. The expression of c-Myc has been associated with the upregulation of Akt-related pathways, leading to cancer cell proliferation and tumorigenesis [53]. The downregulation of c-Myc using HNSCC preclinical models has been observed with tanshinone IIA, leading to the inhibition of tumorigenesis [43].

2.2. MAPK/ERK Pathway

Activation of the MAPK/ERK signaling pathway, which includes signaling molecules such as interleukin-8 (IL-8), vascular endothelial growth factor (VEGF), mitogen-activated extracellular protein kinase (MEK), and extracellular signal-regulated protein kinase (ERK), is strongly associated with the expression of oncoproteins leading to cell proliferation and angiogenesis [54]. Honokiol, sodium Danshensu, cinnamaldehyde, and protocatechuic acid have been reported to inhibit MAPK/ERK signaling in HNSCC preclinical models, thus inhibiting cell proliferation [45,48,55,56]. Protocatechuic acid inhibits MAPK/ERK signaling via activating the c-Jun N-terminal kinase/p38 (JNK/p38) signaling pathway [56]. However, the expression of JNK plays a dual role in HNSCC, both tumor-suppressive and -progressive, due to the complex crosstalk between multiple signaling molecules and pathways [57]. Therefore, understanding of the role of JNK and p38 in HNSCC is required to ensure relevant clinical research in future. The overexpression of matrix metalloproteinases (MMPs), such as MMP-2 and MMP-9, has been associated with the progression and metastasis of HNSCC [58]. Fucoidan, sodium Danshensu, [(3β)-3-hydroxy-lup-20(29)-en-28-oic acid], and Seco-A-ring oleanane have been reported to downregulate MMP-2 [49,50,55]. Meanwhile, fucoidan and trichodermin downregulate MMP-9 in HNSCC preclinical models [49,56], inhibiting cancer cell metastasis. The expression of VEGF plays an important role in blood vessel formation, known as angiogenesis, which, in turn, promotes cancer cell growth due to the supply of nutrients via newly formed blood vessels [59,60]. It was also noticed that VEGF could be upregulated by human growth factor (HGF) through the PI3K/Akt/mTOR and MAPK signaling pathways [61]. Preclinical studies of HNSCC have indicated that cinnamaldehyde, [(3β)-3-hydroxy-lup-20(29)-en-28-oic acid], and Seco-A-ring oleanane could downregulate VEGF, thus inhibiting angiogenesis [48,50].

2.3. NF-κB and STAT3 Transcription Factors

Nuclear factor-κappa B (NF-κB), a protein complex consisting of five transcription factors, namely RelA, RelB, c-Rel, NF-κB1, and NF-κB2, has been reported to play an important role in cell proliferation and survival in HNSCC [62,63]. Preclinical studies using HNSCC models have demonstrated the inhibition of NF-κB, leading to apoptosis with neferine, trichodermin, cinnamaldehyde, and Seco-A-ring oleanane [48,50,64,65].
The expression of the signal transducer and activator of transcription-3 (STAT3), a transcription factor from the STAT family, plays an important role in cell proliferation, survival, and metastasis in HNSCC [66,67,68]. Phosphorylation of STAT3 (p-STAT3) has been proposed as a crucial reaction for the complete activation of STAT3, which is regulated by p38, EGFR, and janus kinases (JAKs) [69,70]. The activation of STAT3 would also lead to various STAT3-dependent pathways, such as the interleukin-8/STAT3 (IL-8/STAT3) and EGFR/STAT3/SRY-box transcription factor 2 (EGFR/STAT3/SOX2) pathways, which have been shown to play important roles in cancer stemness [71,72]. Preclinical studies in HNSCC models have indicated the downregulation of p-STAT3 with trichodermin and Seco-A-ring oleanane [50,65]. Meanwhile, psorachromene caused the downregulation of EGFR, leading to apoptosis and inhibiting metastasis [24].
Actein, a biologically active compound found in the rhizome of Cinicifuga foetida, has been discovered to inhibit forkhead box O1 (FoxO1) by upregulating the phosphorylation of FoxO1 (p-FoxO1), leading to the inhibition of cell proliferation via the Akt/FoxO1 signaling pathway [41]. Knockdown of FoxO1 would reverse the antiproliferation properties of actein, indicating that the Akt/FoxO1 pathway plays an important role in actein-induced effects toward HNSCC [41]. FoxO1 is one of many transcription factors among the FoxO family which regulates various biological activities in HNSCC, such as cancer cell invasion and proliferation [73].
Activated p53 via phosphorylation acts as a tumor suppressive molecule, which has been associated with induced apoptosis in HNSCC [74,75]. About half of HNSCC cases have shown a loss-of-function p53 gene mutation [76], making p53 an interesting target in improving the efficiency of HNSCC therapy through the reactivation of p53. In HPV-positive HNSCC, p53 was downregulated by the E6 oncoprotein, leading to tumorigenesis [76]. Notably, neferine and ilimaquinone have been shown to upregulate or activate p53 in preclinical HNSCC studies, leading to apoptosis [46,64].

2.4. LC3-Dependent Autophagy

Autophagy is a major process involving protein degradation followed by the turnover of cellular components, which helps maintain the intercellular homeostasis [77]. Light chain 3 (LC3) has been strongly associated with the regulation of autophagosome, a double-membrane vesicle formed during autophagy [77,78]. During the formation of an autophagosome, cytosolic LC3 (LC3-I) is converted into the activated form of intra-autophagosomal LC3 (LC3-II) via a uibiquitylation-like reaction catalyzed by Atg7 and Atg3, by conjugating with phosphatidylethanolamine [77,78,79]. Therefore, the increased conversion of LC3-I to LC3-II indicates higher levels of autophagosome and thus the activation of autophagy, leading to tumor suppression [80]. The accumulation of p62 was also found to be associated with the induction of autophagy in HNSCC [81]. Honokiol, ilimaquinone, and nimbolide have been reported to increase LC3-II/LC3-I ratio in HNSCC preclinical models [42,45,47].
Paradoxically, autophagy may potentially lead to tumor survival when certain conditions are met [82,83,84,85]. Hypoxia, a condition where the oxygen level is below the physiological level, is a common feature in tumor progression when the oxygen supply does not meet the demand due to the exponential growth of tumor [83,86]. Hypoxia-induced autophagy in tumors has been observed to induce tumor survival in vitro [83]. Meanwhile, the knockdown of essential autophagy proteins with a xenograft model was found to promote tumor suppression [84]. The turnover of cellular components by autophagy may essentially help with hypoxia and nutrient stress faced by exponentially growing tumors, thus enhancing tumor survival [83,84].

2.5. Bcl-2/Bax Signaling

Bcl-2 and Bax are proteins found in the mitochondrial membrane, nuclear envelop, and endoplasmic reticulum [87,88]. Bcl-2 inhibits the release of cytochrome c, which impedes tointrinsic apoptosis, while Bax reverses the reaction, leading to the induction of intrinsic apoptosis in HNSCC [89]. Cytochrome c, when released followed by a cascade of reaction involving apoptotic protease activating factor 1 (Apaf-1), Caspase-9, Caspase-7, and Caspase-3, causes apoptosis [89,90]. A preclinical study with fisetin showed that the upregulation of cytochrome c leads to apoptosis [91]. Preclinical studies in HNSCC with actein, xanthohumol, fucoidan, ilimaquinone, nimbolide, 4-O-methylhonokiol, cinnamaldehyde, kaempferol, and fisetin have indicated higher a Bax/Bcl-2 ratio (upregulating Bax and/or downregulating Bcl-2) on the mitochondrial membrane, supporting the induction of intrinsic apoptosis of HNSCC [41,44,46,47,48,49,91,92]. The activation of Caspase-8 has been associated with the activation of Caspase-7 and Caspase-3 followed by apoptosis induction, while Survivin acts as an anti-apoptotic protein [75,93,94]. In preclinical studies of HNSCC, Ilimaquinone was reported to upregulate Caspase-8, while the downregulation of Survivin has been observed with actein, xanthohumol, ilimaquinone, and seco-A-ring oleanane [41,44,46,50].

2.6. Cell Cycle Arrest by Cyclin and the CDK Signaling Pathway

The cell cycle acts as a fundamental process for cancer progression in HNSCC, including cell proliferation and differentiation [95]. It is well established that cyclins and cyclin-dependent kinases (Cdks) play an important role in the regulation and transition of cell cycles [96,97,98]. G0/G1 transitioning is strongly dependent on cyclin-D/Cdk4/6 complexes; S phase entry is dependent on the cyclin E/Cdk2 complex; S/G2 transitioning is dependent on the cyclin A/Cdk2 complex, followed by mitotic phase entry, which is dependent on the cyclin A/Cdk1 complex, and finally, M/G0 transitioning, which is dependent on the cyclin B/Cdk1 complex [99,100,101,102,103]. The inhibition of the cell cycle via the formation of cyclin-Cdk complexes would lead to cell cycle arrest and, eventually, cell death in HNSCC [104]. Anti-mitogenic signals such as p16, p21, and p53 could inhibit the cell cycle transition effectively by inhibiting the formation of cyclin-Cdk complexes, thus acting as an important target for cell cycle arrest induction [99,105,106]. Xanthohumol, neferine, fucoidan, hydroxygenkwanin, ilimaquinone, honokiol, trichodermin, cinnamaldehyde, resveratrol, and curcumin have been discovered to induce cell cycle arrest via cyclin-dependent signaling with HNSCC preclinical models [44,45,46,48,49,56,64,107,108].

2.7. Potential Natural Products as Therapeutic Agent for HNSCC

Actein, salicylate, honokiol, trichodermin, psorachromene, protocatechuic acid, fucoidan, hydroxygenkwanin, nimbolide, 4-O-methylhonokiol, [(3β)-3-hydroxy-lup-20(29)-en-28-oic acid], seco-A-ring oleanane, cinnamaldehyde, kaempferol, resveratrol, and curcumin induced cell cycle arrest via the regulation of cyclins and Cdks, which play an important role in coordinating the cell cycle progression [24,41,42,45,47,48,49,50,56,65,91,92,107,108]. The expressions of p21, p27, p16, and p53, acting as cyclin-Cdk complex inhibitors, have also been upregulated by actein, hydroxygenkwanin, honokiol, ilimaquinone, tanshinone IIA, resveratrol, and curcumin, which leads to induced cell cycle arrest [41,43,45,46,107,108].
Cell migration and angiogenesis have been widely studied and strongly associated with regulations of MMPs. Fucoidan, trichodermin, [(3β)-3-hydroxy-lup-20(29)-en-28-oic acid], seco-A-ring oleanane, and sodium danshensu have shown to downregulate MMPs, which inhibit cell migration and angiogenesis in HNSCC [49,50,55,65]. The VEGF and VEGF receptors play an important role in coordinating angiogenesis, while ilimaquinone, cinnamaldehyde, [(3β)-3-hydroxy-lup-20(29)-en-28-oic acid], and seco-A-ring oleanane have shown inhibition of the VEGF signaling in vitro [46,48,50].
The PI3K/Akt/mTOR and MAPK/ERK signaling pathways have been well studied and have shown a strong association with the cell proliferation and survival of HNSCC. In vitro studies with actein, salicylate, tanshinone IIA, fucoidan, honokiol, xanthohumol, ilimaquinone, cinnamaldehyde, nimbolide, and seco-A-ring oleanane have disrupted the PI3K/Akt/mTOR signaling pathway, leading to the inhibition of cell proliferation and survival [41,42,43,44,45,46,47,48,49,50]. In contrary, in vitro studies with honokiol, sodium danshensu, protocatechuic acid, and cinnamaldehyde have shown inhibition of the MAPK/ERK signaling pathway, leading to the inhibition of cell proliferation and survival [45,48,55,56].
Cell apoptosis via the mitochondrial pathway is regulated by caspases, where the proteases are activated by the Bax and Bcl-2 proteins present in the mitochondrial membrane. In vitro studies have shown that the downregulation of Bcl-2 and upregulation of Bax by fucoidan, ilimaquinone, nimbolide, 4-O-methylhonokiol, actein, and cinnamaldehyde leads to a cascade reaction of apoptosis induction [41,46,47,48,49,92]. The upregulation of caspases leading to apoptosis have also been shown by in vitro studies with hydroxygenkwanin, ilimaquinone, nimbolide, psorachromene, actein, conocarpan, trichodermin, protocatechuic acid, kaempferol, and fisetin [24,41,46,47,56,65,91,107,109]. Table 1 summarizes in vitro preclinical studies using natural products, while Table 2 summarizes in vitro and in vivo preclinical studies of natural products.

3. Chemoprevention Properties of Natural Products against HNSCC Oral Carcinogenesis Mechanism

Oral carcinogenesis often involves the formation of abnormalities in the oral tissue, known as oral potentially malignant disorder (OPMD), before proceeding to oral squamous cell carcinoma (OSCC), a major type of HNSCC [111,112]. Previous studies have found the malignant transformation (MT) rate of OPMD to be 7.9%, while high-risk OPMD, such as erythroplakia, has shown an average MT rate of 33.1% [113]. To date, no preventive strategies including the use of drugs and/or surgical procedures has been considered the standard of care for OPMDs, thus indicating the need to investigate the chemoprevention properties of various natural products against oral carcinogenesis [114]. Several drugs, such as celecoxib, erlotinib, and metformin, have been investigated for their HNSCC prevention properties via clinical trials on oral premalignant lesions [115,116,117]. However, the use of erlotinib and celecoxib has shown no significant result in reducing the oral cancer-free survival rate while possessing higher toxicity [115,116]. The use of metformin has shown a low clinical response rate (17%) in terms of reduction in lesion size [117].
In vivo studies involving the use of genetically altered rodents or rodents treated with chemical carcinogens could lead to site-specific carcinogenesis, mimicking carcinogenesis in humans [118]. Moreover, 4-nitroquinoline 1-oxide (4NQO) acts as a tobacco-mimicking carcinogen, which has been widely used in carcinogen-induced HNSCC animal models, mainly due to the similarity in terms of genetic alteration and expression between 4NQO-induced mouse models and human oral carcinogenesis [119,120]. Other carcinogens, including 7,12-dimethylbenz(a)anthracene (DMBA) and dibenzo[a,l]pyrene (DBP), have also been widely used as HNSCC-inducing agents in animal models [64,121].
Several review studies have introduced various natural products as chemoprevention agents against HNSCC, where vitamin A, green tea extracts, and curcumin have shown promising results in preclinical and clinical trials [29,30]. However, vitamin A has shown toxicity [31,32], while green tea extracts and curcumin [33,34] have both shown limitations in bioavailability. In a randomized chemoprevention trial reported by Papadimitrakopoulou et al. [32], low-dose 13-cis retinoic acid (a derivative of vitamin A) could induce grade 1 (45%), 2 (37%), 3 (15%), and 4 (1%) toxicity, including cheilitis, conjunctivitis, and skin reactions. Similarly, a phase II chemoprevention trial by Shin et al. [31] with combinations of interferon-alpha, 13-cis retinoic acid, and alpha-tocopherol induced mild to moderate non-hematologic toxicity. One patient was reported with a severe throat infection due to beta-hemolytic streptococci which required an emergency tracheostomy. The patient was still able to complete the planned treatment after fully recovering from the infection [31]. Green tea extract was reported to induce adverse effects such as insomnia, nausea, nervousness, and headache, which is most likely due to the presence of caffeine in green tea extract [33]. The poor oral absorption of epigallocatechin-3-gallate, the most abundant polyphenol in green tea extract, was reported by Tsao et al. [33], leading to variability in plasma epigallocatechin-3-gallate concentrations. Similarly, a phase I chemoprevention trial by Cheng et al. [34] with curcumin also indicated the poor gastrointestinal absorption of curcumin, as the peak serum curcumin concentration was recorded at 1.77 µM with 8000 mg daily dosage. The poor bioavailability of both green tea extract and curcumin have introduced difficulties in dosage estimation as absorption varies among patients, which may lead to ineffective treatment. Therefore, in this review, we seek to provide a greater variety of promising natural products with chemoprevention properties using 4NQO, DMBA, or DBP-induced carcinogenesis animal models [25,64,121]. Table 3 summarizes preclinical studies with natural products involved in the chemoprevention of HNSCC investigated in the past 5 years.
Preclinical in vivo studies involving neferine and nimbolide have indicated an inhibition in carcinogenesis and a reduction in the tumor growth of DMBA-induced HNSCC, making both possible chemoprevention agents [41]. Similarly, calcitriol successfully inhibited 4NQO-induced HNSCC carcinogenesis [25]. Interestingly, a study by Vincent-Chong et al. [25] with calcitriol on a 4NQO-induced animal model showed the influence of treatment stage intervention and duration of exposure to treatment in carcinogenesis. Understanding the pathways involved during carcinogenesis will allow effective treatment intervention in HNSCC patients.
Calcitriol, nimbolide, and neferine are being investigated in preclinical studies for chemoprevention, and these products have shown promising results by reducing or inhibiting carcinogenesis-related molecular mechanisms [25,47,64].

4. Limitation and Future Direction

As multiple signaling and crosstalk between pathways occur during carcinogenesis, with most recurrent or metastatic HNSCC failing the primary standard of care, an effective treatment for cancer may require a combined therapeutic approach such as the use of multiple signaling inhibitors combined with DNA-damaging drugs for the most efficient outcome [26]. Therefore, the synergic or antagonistic effects of natural products with standards of care such as chemotherapy (cisplatin, cetuximab, pembrolizumab and nivolumab) and radiotherapy should be analyzed using preclinical models. However, fewer than 10% of the reviewed studies reported the combination effects of natural products with the standard of care. For instance, xanthohumol, psorachromene, honokiol, calcitriol, and salicylate were shown to provide synergistic effects with standard-of-care treatment such as chemotherapy and radiotherapy using HNSCC preclinical models [24,25,42,44,45].
The major reason for the low survival rate of HNSCC is due to late diagnoses and risk factors associated with HNSCC progression, which lead to the risk of recurrent or metastatic SCC [5,12]. Chemoprevention therapy could potentially act as an important barrier to lower the risk of recurrent or metastatic SCC and the malignant transformation of OPMD; therefore, chemoprevention should be widely studied in the future. The prevention of HNSCC involving single-agent chemotherapy, such as retinoids and isotretinoin, possesses high toxicity and low efficacy, indicating the need for the development of new chemoprevention agents either as alternative or adjunctive agents for HNSCC prevention. In the current review, only 5 out of 37 studies explored the potential usage of natural products as chemoprevention therapy using 4NQO/DMBA/DBP-induced oral carcinogenesis [25,47,64,121]. Furthermore, the stage and duration of natural products’ intervention on oral carcinogenesis should also be extensively explored given the [25] strong association with the progression of carcinogenesis. All six natural products (Table 3) are strongly encouraged to proceed with clinical trials for high-risk OPMD patients, as previous trials with celecoxib, erlotinib, and metformin on mild to advanced OPMD showed no significant clinical improvements [115,116,117].
A preclinical study reported by Dai et al. [122] using various cancer lines (melanoma, breast, colon, and liver cancer cell lines (H1299, BT549, MDA-MB-231, MDA-MB-468, SW620, MHCC97H, and B16F10)) indicated the crucial role of immunomodulatory roles of natural products in anti-cancer treatments, including the involvement of CD3+ CD8+ T lymphocytes in a co-culture system together with cancer cell lines. Similarly, Cattanaeo et al. [123] and Neal et al. [124] have proposed the use of co-culture organoid-tumor reactive T lymphocyte system to investigate the role of anti-PD-1 drugs on T lymphocyte activity. Other studies have also used similar systems to screen natural product-derived drugs and epigenetic inhibitors for the non-cytotoxic T lymphocyte immunomodulating effects [122,125]. Apart from the in vitro co-culture system proposed, Verma et al. [126] implemented an in vivo RPMOC1 synergic HNSCC animal model to investigate the effects of a non-oncological drug, calcitriol, on the immunomodulation of T lymphocytes, which was the only preclinical study involving synergic HNSCC in an in vivo model present at the time of preparing this review. The initiation of this review was prompted by a recent publication by Crooker et al. [29] which provided informaion on the use of natural products in HNSCC preclinical models in 2018. Over the past five years, there have been limited efforts to publish reviews related to natural products, with one notable publication by Aggarwal et al. [12] focusing solely on HPV-related HNSCC. To address the paucity of information, we explored the possibilities of using the latest research of natural products with in vitro and in vivo studies, listed in Table 1, Table 2 and Table 3, that has not been discussed in previous review papers, particularly targeting HNSCC, as some signaling pathways are common among different subtypes. The review also highlights various challenges in the field that hinder the clinical translation of natural products.
Among the reviewed natural products, actein, salicylate, tanshinone IIA, xanthohumol, honokiol, trichodermin, psorachromene, and protocatechuic acid have been investigated as chemotherapeutic agents with HNSCC cell lines and xenografted animal models, and these products have shown promising targeted molecular mechanisms against HNSCC, making them ideal candidates for further clinical trials for safety and efficacy analyses [24,41,42,43,44,45,56,65]. Furthermore, salicylate and xanthohumol have been investigated to provide synergistic effect with cisplatin and radiotherapy, respectively, making them the strongest candidates for future clinical trials for HNSCC or OPMD patients [42,44]. Finally, calcitriol, nimbolide, and neferine were the major natural products investigated in this review; these products showed promising chemoprevention properties against induced-carcinogenesis animal models. We encourage further investigation into the safety and efficacy of these products with human trials, as well as in high-risk OPMD such as erythroplakia and leukoplakia patients [25,47,64]. Nevertheless, it is crucial that the results of preclinical studies presented in this manuscript serve only as a principle for further investigations and should not be directly extrapolated into clinical practices without additional detailed evaluation. Preclinical studies, either in vitro or in vivo, provide valuable insights on the potential of natural products as therapeutic agents; however, their efficacy, toxicity profile, and applicability in human patients should be widely explored. Finally, caution must be exercised when interpreting the findings of preclinical studies and their potential implication for clinical trials.

5. Conclusions

Both chemotherapeutic and chemoprevention approaches showed promising anti-cancer effects through preclinical studies involving various HNSCC cell lines and animal models via various pathways (Figure 2). However, the lack of extensive molecular mechanisms of natural products and lack of combination of natural products with the current standard of care limits the use of natural products as new chemotherapeutic drugs. The only study that used calcitriol to determine the immunomodulatory effect of natural product in the context of HNSCC highlighted the effort to investigate the role of these natural products since the immune checkpoint inhibitor has been recognized as one of the standards of care for HNSCC patients. The low intrinsic toxicity of natural products in normal cells and significant therapeutic effects toward cancer cells have sparked an interest in oncology studies in recent years. The synergistic effect of natural products with standard of cares, including chemotherapy, radiotherapy, and immunotherapy, has showed promising properties as both alternative and adjunctive chemotherapeutic or chemoprevention agents in cancer treatment or prevention, respectively.
As a conclusion, we attempted to provide a comprehensive database of natural products used for in vivo and in vitro preclinical studies involving HNSCC in the recent years, which would facilitate the identification of effective natural products that show promising chemotherapeutic and chemoprevention properties against HNSCC while possessing low toxicity. Natural products have shown to be promising molecular targets for therapy and prevention against HNSCC, making them great alternatives and adjunctive agents in cancer treatment. However, further toxicity profiles should be analyzed for natural products in future clinical trials to ensure the safety and efficacy of natural products against HNSCC.

Author Contributions

Conceptualization, V.K.V.-C.; Investigation, Y.X.L., L.P.K.-N. and V.K.V.-C.; Visualization, Y.X.L.; Writing—Original Draft Preparation, Y.X.L.; Writing—Review and Editing, L.P.K.-N. and V.K.V.-C. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Johnson, D.E.; Burtness, B.; Leemans, C.R.; Lui, V.W.Y.; Bauman, J.E.; Grandis, J.R. Head and Neck Squamous Cell Carcinoma. Nat. Rev. Dis. Prim. 2020, 6, 92. [Google Scholar] [CrossRef] [PubMed]
  2. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global Cancer Statistics 2020: GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef] [PubMed]
  3. Magnes, T.; Wagner, S.M.; Melchardt, T.; Weiss, L.; Rinnerthaler, G.; Huemer, F.; Kopp, M.; Gampenrieder, S.P.; Mayrbäurl, B.; Füreder, T.; et al. Postoperative Chemoradiotherapy with Cisplatin Is Superior to Radioimmunotherapy with Cetuximab and Radiotherapy Alone: Analysis of the Austrian Head and Neck Cancer Registry of the AGMT. Wien. Klin. Wochenschr. 2021, 133, 1131–1136. [Google Scholar] [CrossRef]
  4. Siegel, R.L.; Miller, K.D.; Fuchs, H.E.; Jemal, A. Cancer Statistics, 2022. CA Cancer J. Clin. 2022, 72, 7–33. [Google Scholar] [CrossRef] [PubMed]
  5. Bonner, J.A.; Harari, P.M.; Giralt, J.; Cohen, R.B.; Jones, C.U.; Sur, R.K.; Raben, D.; Baselga, J.; Spencer, S.A.; Zhu, J.; et al. Radiotherapy plus Cetuximab for Locoregionally Advanced Head and Neck Cancer: 5-Year Survival Data from a Phase 3 Randomised Trial, and Relation between Cetuximab-Induced Rash and Survival. Lancet Oncol. 2010, 11, 21–28. [Google Scholar] [CrossRef]
  6. Ferris, R.L.; Blumenschein, G.; Fayette, J.; Guigay, J.; Colevas, A.D.; Licitra, L.; Harrington, K.; Kasper, S.; Vokes, E.E.; Even, C.; et al. Nivolumab for Recurrent Squamous-Cell Carcinoma of the Head and Neck. N. Engl. J. Med. 2016, 375, 1856–1867. [Google Scholar] [CrossRef]
  7. Talamini, R.; Bosetti, C.; La Vecchia, C.; Dal Maso, L.; Levi, F.; Bidoli, E.; Negri, E.; Pasche, C.; Vaccarella, S.; Barzan, L.; et al. Combined Effect of Tobacco and Alcohol on Laryngeal Cancer Risk: A Case-Control Study. Cancer Causes Control 2002, 13, 957–964. [Google Scholar] [CrossRef]
  8. Lewin, F.; Norell, S.E.; Johansson, H.; Gustavsson, P.; Wennerberg, J.; Biörklund, A.; Rutqvist, L.E. Smoking Tobacco, Oral Snuff, and Alcohol in the Etiology of Squamous Cell Carcinoma of the Head and Neck A Population-Based Case-Referent Study in Sweden. Cancer 1998, 82, 1367–1375. [Google Scholar] [CrossRef]
  9. Hansson, B.G.; Rosenquist, K.; Antonsson, A.; Wennerberg, J.; Schildt, E.B.; Bladström, A.; Andersson, G. Strong Association between Infection with Human Papillomavirus and Oral and Oropharyngeal Squamous Cell Carcinoma: A Population-Based Case-Control Study in Southern Sweden. Acta Otolaryngol. 2005, 125, 1337–1344. [Google Scholar] [CrossRef]
  10. Van Wyk, C.W.; Stander, I.; Padayachee, A.; Grobler-Rabie, A.F. The Areca Nut Chewing Habit and Oral Squamous Cell Carcinoma in South African Indians. A Retrospective Study. S. Afr. Med. J. 1993, 83, 425–429. [Google Scholar]
  11. Haddad, R.I.; Shin, D.M. Recent Advances in Head and Neck Cancer Reconstruction. N. Engl. J. Med. 2008, 359, 1143–1154. [Google Scholar] [CrossRef] [PubMed]
  12. Aggarwal, N.; Yadav, J.; Chhakara, S.; Janjua, D.; Tripathi, T.; Chaudhary, A.; Chhokar, A.; Thakur, K.; Singh, T.; Bharti, A.C. Phytochemicals as Potential Chemopreventive and Chemotherapeutic Agents for Emerging Human Papillomavirus–Driven Head and Neck Cancer: Current Evidence and Future Prospects. Front. Pharmacol. 2021, 12, 699044. [Google Scholar] [CrossRef] [PubMed]
  13. Posner, M.R.; Hershock, D.M.; Blajman, C.R.; Mickiewicz, E.; Winquist, E.; Gorbounova, V.; Tjulandin, S.; Shin, D.M.; Cullen, K.; Ervin, T.J.; et al. Cisplatin and Fluorouracil Alone or with Docetaxel in Head and Neck Cancer Marshall. N. Engl. J. Med. 2007, 357, 1705–1715. [Google Scholar] [CrossRef] [PubMed]
  14. Hashim, D.; Genden, E.; Posner, M.; Hashibe, M.; Boffetta, P. Head and Neck Cancer Prevention: From Primary Prevention to Impact of Clinicians on Reducing Burden. Ann. Oncol. 2019, 30, 744–756. [Google Scholar] [CrossRef]
  15. Brockstein, B.; Haraf, D.J.; Rademaker, A.W.; Kies, M.S.; Stenson, K.M.; Rosen, F.; Mittal, B.B.; Pelzer, H.; Fung, B.B.; Witt, M.E.; et al. Patterns of Failure, Prognostic Factors and Survival in Locoregionally Advanced Head and Neck Cancer Treated with Concomitant Chemoradiotherapy: A 9-Year, 337-Patient, Multi-Institutional Experience. Ann. Oncol. 2004, 15, 1179–1186. [Google Scholar] [CrossRef]
  16. Bonner, J.A.; Harari, P.M.; Giralt, J.; Azarnia, N.; Shin, D.M.; Cohen, R.B.; Jones, C.U.; Sur, R.; Raben, D.; Jassem, J.; et al. Results of Radiotherapy plus Cetuximab for Squamous Cell Carcinoma of the Head and Neck. N. Engl. J. Med. 2006, 354, 567–578. [Google Scholar] [CrossRef]
  17. Naruse, T.; Yanamoto, S.; Matsushita, Y.; Sakamoto, Y.; Morishita, K.; Ohba, S.; Shiraishi, T.; Yamada, S.-I.; Asahina, I.; Umeda, M. Cetuximab for the Treatment of Locally Advanced and Recurrent/Metastatic Oral Cancer: An Investigation of Distant Metastasis. Mol. Clin. Oncol. 2016, 5, 246–252. [Google Scholar] [CrossRef]
  18. Sok, J.C.; Coppelli, F.M.; Thomas, S.M.; Lango, M.N.; Xi, S.; Hunt, J.L.; Freilino, M.L.; Graner, M.W.; Wikstrand, C.J.; Bigner, D.D.; et al. Mutant Epidermal Growth Factor Receptor (EGFRvIII) Contributes to Head and Neck Cancer Growth and Resistance to EGFR Targeting. Clin. Cancer Res. 2006, 12, 5064–5073. [Google Scholar] [CrossRef]
  19. US Food and Drug Administration. FDA Approves Pembrolizumab for First-Line Treatment of Head and Neck Squamous Cell Carcinoma. Available online: https://www.fda.gov/drugs/resources-information-approved-drugs/fda-approves-pembrolizumab-first-line-treatment-head-and-neck-squamous-cell-carcinoma (accessed on 1 February 2023).
  20. US Food and Drug Administration. Nivolumab for SCCHN. Available online: https://www.fda.gov/drugs/resources-information-approved-drugs/nivolumab-scchn (accessed on 1 February 2023).
  21. dos Santos, L.V.; Abrahão, C.M.; William, W.N. Overcoming Resistance to Immune Checkpoint Inhibitors in Head and Neck Squamous Cell Carcinomas. Front. Oncol. 2021, 11, 596290. [Google Scholar] [CrossRef]
  22. Newman, D.J.; Cragg, G.M. Natural Products as Sources of New Drugs over the Nearly Four Decades from 01/1981 to 09/2019. J. Nat. Prod. 2020, 83, 770–803. [Google Scholar] [CrossRef]
  23. National Cancer Institute. Dictionary of Cancer Terms. Available online: https://www.cancer.gov/publications/dictionaries/cancer-terms (accessed on 26 September 2022).
  24. Wang, T.H.; Leu, Y.L.; Chen, C.C.; Shieh, T.M.; Lian, J.H.; Chen, C.Y. Psorachromene Suppresses Oral Squamous Cell Carcinoma Progression by Inhibiting Long Non-Coding RNA GAS5 Mediated Epithelial-Mesenchymal Transition. Front. Oncol. 2019, 9, 1168. [Google Scholar] [CrossRef]
  25. Vincent-Chong, V.K.; DeJong, H.; Attwood, K.; Hershberger, P.A.; Seshadri, M. Preclinical Prevention Trial of Calcitriol: Impact of Stage of Intervention and Duration of Treatment on Oral Carcinogenesis. Neoplasia 2019, 21, 376–388. [Google Scholar] [CrossRef] [PubMed]
  26. Hashem, S.; Ali, T.A.; Akhtar, S.; Nisar, S.; Sageena, G.; Ali, S.; Al-Mannai, S.; Therachiyil, L.; Mir, R.; Elfaki, I.; et al. Targeting Cancer Signaling Pathways by Natural Products: Exploring Promising Anti-Cancer Agents. Biomed. Pharmacother. 2022, 150, 113054. [Google Scholar] [CrossRef]
  27. Sever, R.; Brugge, J.S. Signal Transduction in Cancer. Cold Spring Harb. Perspect. Med. 2015, 5, a006098. [Google Scholar] [CrossRef]
  28. Sanchez-Vega, F.; Mina, M.; Armenia, J.; Chatila, W.K.; Luna, A.; La, K.C.; Dimitriadoy, S.; Liu, D.L.; Kantheti, H.S.; Saghafinia, S.; et al. Oncogenic Signaling Pathways in The Cancer Genome Atlas. Cell 2019, 173, 321–337. [Google Scholar] [CrossRef] [PubMed]
  29. Crooker, K.; Aliani, R.; Ananth, M.; Arnold, L.; Anant, S.; Thomas, S.M. A Review of Promising Natural Chemopreventive Agents for Head and Neck Cancer. Cancer Prev. Res. 2018, 11, 441–450. [Google Scholar] [CrossRef]
  30. Rahman, M.A.; Amin, A.R.M.R.; Shin, D.M. Chemopreventive Potential of Natural Compounds in Head and Neck Cancer. Nutr. Cancer 2010, 62, 973–987. [Google Scholar] [CrossRef] [PubMed]
  31. Shin, D.M.; Khuri, F.R.; Murphy, B.; Garden, A.S.; Clayman, G.; Francisco, M.; Liu, D.; Glisson, B.S.; Ginsberg, L.; Papadimitrakopoulou, V.; et al. Combined Interferon-Alfa, 13-Cis-Retinoic Acid, and Alpha-Tocopherol in Locally Advanced Head and Neck Squamous Cell Carcinoma: Novel Bioadjuvant Phase II Trial. J. Clin. Oncol. 2001, 19, 3010–3017. [Google Scholar] [CrossRef]
  32. Papadimitrakopoulou, V.A.; Lee, J.J.; William, W.N.; Martin, J.W.; Thomas, M.; Kim, E.S.; Khuri, F.R.; Shin, D.M.; Feng, L.; Waun, K.H.; et al. Randomized Trial of 13-Cis Retinoic Acid Compared with Retinyl Palmitate with or without Beta-Carotene in Oral Premalignancy. J. Clin. Oncol. 2009, 27, 599–604. [Google Scholar] [CrossRef]
  33. Tsao, A.S.; Liu, D.; Martin, J.; Tang, X.M.; Lee, J.J.; El-Naggar, A.K.; Wistuba, I.; Culotta, K.S.; Mao, L.; Gillenwater, A.; et al. Phase II Randomized, Placebo-Controlled Trial of Green Tea Extract in Patients with High-Risk Oral Premalignant Lesions. Cancer Prev. Res. 2009, 2, 931–941. [Google Scholar] [CrossRef]
  34. Cheng, A.L.; Hsu, C.H.; Lin, J.K.; Hse, M.W.; Ho, Y.F.; Shen, T.S.; Ko, J.Y.; Lin, J.T.; Lin, B.R.; Wu, M.S.; et al. Phase I Clinical Trial of Curcumin, a Chemopreventive Agent, in Patients with High-Risk or Pre-Malignant Lesions. Anticancer Res. 2001, 21, 2895–2900. [Google Scholar] [PubMed]
  35. Molinolo, A.A.; Hewitt, S.M.; Amornphimoltham, P.; Keelawat, S.; Rangdaeng, S.; García, A.M.; Raimondi, A.R.; Jufe, R.; Itoiz, M.; Gao, Y.; et al. Dissecting the Akt/Mammalian Target of Rapamycin Signaling Network: Emerging Results from the Head and Neck Cancer Tissue Array Initiative. Clin. Cancer Res. 2007, 13, 4964–4973. [Google Scholar] [CrossRef] [PubMed]
  36. Freudlsperger, C.; Horn, D.; Weißfuß, S.; Weichert, W.; Weber, K.J.; Saure, D.; Sharma, S.; Dyckhoff, G.; Grabe, N.; Plinkert, P.; et al. Phosphorylation of AKT(Ser473) Serves as an Independent Prognostic Marker for Radiosensitivity in Advanced Head and Neck Squamous Cell Carcinoma. Int. J. Cancer 2015, 136, 2775–2785. [Google Scholar] [CrossRef] [PubMed]
  37. Zumsteg, Z.S.; Morse, N.; Krigsfeld, G.; Gupta, G.; Higginson, D.S.; Lee, N.Y.; Morris, L.; Ganly, I.; Shiao, S.L.; Powell, S.N.; et al. Taselisib (GDC-0032), a Potent β-Sparing Small Molecule Inhibitor of PI3K, Radiosensitizes Head and Neck Squamous Carcinomas Containing Activating PIK3CA Alterations. Clin. Cancer Res. 2016, 22, 2009–2019. [Google Scholar] [CrossRef] [PubMed]
  38. Iwase, M.; Yoshiba, S.; Uchid, M.; Takaoka, S.; Kurihara, Y.; Ito, D.; Hatori, M.; Shintani, S. Enhanced Susceptibility to Apoptosis of Oral Squamous Cell Carcinoma Cells Subjected to Combined Treatment with Anticancer Drugs and Phosphatidylinositol 3-Kinase Inhibitors. Int. J. Oncol. 2007, 31, 1141–1147. [Google Scholar] [PubMed]
  39. Cantley, L.C. The Phosphoinositide 3-Kinase Pathway. Science 2002, 296, 1655–1657. [Google Scholar] [CrossRef]
  40. Sarbassov, D.D.; Guertin, D.A.; Ali, S.M.; Sabatini, D.M. Phosphorylation and Regulation of Akt/PKB by the Rictor-MTOR Complex. Science 2005, 307, 1098–1101. [Google Scholar] [CrossRef]
  41. Zhao, C.; Zhang, Z.; Dai, X.; Wang, J.; Liu, H.; Ma, H. Actein Antagonizes Oral Squamous Cell Carcinoma Proliferation through Activating FoxO1. Pharmacology 2021, 106, 551–563. [Google Scholar] [CrossRef]
  42. Zhang, X.; Wang, F.; Zeng, Y.; Zhu, X.; Peng, L.; Zhang, L.; Gu, J.; Han, H.; Yi, X.; Shi, J. Salicylate Sensitizes Oral Squamous Cell Carcinoma to Chemotherapy through Targeting MTOR Pathway. Oral Dis. 2020, 26, 1131–1140. [Google Scholar] [CrossRef]
  43. Li, M.; Gao, F.; Zhao, Q.; Zuo, H.; Liu, W.; Li, W. Tanshinone IIA Inhibits Oral Squamous Cell Carcinoma via Reducing Akt-c-Myc Signaling-Mediated Aerobic Glycolysis. Cell Death Dis. 2020, 11, 381. [Google Scholar] [CrossRef]
  44. Li, M.; Gao, F.; Yu, X.; Zhao, Q.; Zhou, L.; Liu, W.; Li, W. Promotion of Ubiquitination-Dependent Survivin Destruction Contributes to Xanthohumol-Mediated Tumor Suppression and Overcomes Radioresistance in Human Oral Squamous Cell Carcinoma. J. Exp. Clin. Cancer Res. 2020, 39, 88. [Google Scholar] [CrossRef]
  45. Huang, K.J.; Kuo, C.H.; Chen, S.H.; Lin, C.Y.; Lee, Y.R. Honokiol Inhibits in Vitro and in Vivo Growth of Oral Squamous Cell Carcinoma through Induction of Apoptosis, Cell Cycle Arrest and Autophagy. J. Cell. Mol. Med. 2018, 22, 1894–1908. [Google Scholar] [CrossRef]
  46. Lin, C.W.; Bai, L.Y.; Su, J.H.; Chiu, C.F.; Lin, W.Y.; Huang, W.T.; Shih, M.C.; Huang, Y.T.; Hu, J.L.; Weng, J.R. Ilimaquinone Induces Apoptosis and Autophagy in Human Oral Squamous Cell Carcinoma Cells. Biomedicines 2020, 8, 296. [Google Scholar] [CrossRef]
  47. Sophia, J.; Kowshik, J.; Dwivedi, A.; Bhutia, S.K.; Manavathi, B.; Mishra, R.; Nagini, S. Nimbolide, a Neem Limonoid Inhibits Cytoprotective Autophagy to Activate Apoptosis via Modulation of the PI3K/Akt/GSK-3β Signalling Pathway in Oral Cancer. Cell Death Dis. 2018, 9, 1087. [Google Scholar] [CrossRef] [PubMed]
  48. Aggarwal, S.; Bhadana, K.; Singh, B.; Rawat, M.; Mohammad, T.; Al-Keridis, L.A.; Alshammari, N.; Hassan, M.I.; Das, S.N. Cinnamomum Zeylanicum Extract and Its Bioactive Component Cinnamaldehyde Show Anti-Tumor Effects via Inhibition of Multiple Cellular Pathways. Front. Pharmacol. 2022, 13, 918479. [Google Scholar] [CrossRef] [PubMed]
  49. Zhang, N.; Gao, L.; Ren, W.; Li, S.; Zhang, D.; Song, X.; Zhao, C.; Zhi, K. Fucoidan Affects Oral Squamous Cell Carcinoma Cell Functions in Vitro by Regulating FLNA-Derived Circular RNA. Ann. N. Y. Acad. Sci. 2020, 1462, 65–78. [Google Scholar] [CrossRef] [PubMed]
  50. Aswathy, M.; Banik, K.; Parama, D.; Sasikumar, P.; Harsha, C.; Joseph, A.G.; Sherin, D.R.; Thanathu, M.K.; Kunnumakkara, A.B.; Vasu, R.K. Exploring the Cytotoxic Effects of the Extracts and Bioactive Triterpenoids from Dillenia Indica against Oral Squamous Cell Carcinoma: A Scientific Interpretation and Validation of Indigenous Knowledge. ACS Pharmacol. Transl. Sci. 2021, 4, 834–847. [Google Scholar] [CrossRef]
  51. Schulz, L.; Pries, R.; Lanka, A.S.; Drenckhan, M.; Rades, D.; Wollenberg, B. Inhibition of GSK3α/β Impairs the Progression of HNSCC. Oncotarget 2018, 9, 27630–27644. [Google Scholar] [CrossRef]
  52. Ugolkov, A.V.; Matsangou, M.; Taxter, T.J.; O’halloran, T.V.; Cryns, V.L.; Giles, F.J.; Mazar, A.P. Aberrant Expression of Glycogen Synthase Kinase-3β in Human Breast and Head and Neck Cancer. Oncol. Lett. 2018, 16, 6437–6444. [Google Scholar] [CrossRef]
  53. Marconi, G.D.; Della Rocca, Y.; Fonticoli, L.; Melfi, F.; Rajan, T.S.; Carradori, S.; Pizzicannella, J.; Trubiani, O.; Diomede, F. C-Myc Expression in Oral Squamous Cell Carcinoma: Molecular Mechanisms in Cell Survival and Cancer Progression. Pharmaceuticals 2022, 15, 890. [Google Scholar] [CrossRef]
  54. Bancroft, C.C.; Chen, Z.; Dong, G.; Sunwoo, J.B.; Yeh, N.; Park, C.; Van Waes, C. Coexpression of Proangiogenic Factors IL-8 and VEGF by Human Head and Neck Squamous Cell Carcinoma Involves Coactivation by MEK-MAPK and IKK-NF-ΚB Signal Pathways. Clin. Cancer Res. 2001, 7, 435–442. [Google Scholar] [PubMed]
  55. Kumar, V.B.; Lin, S.H.; Mahalakshmi, B.; Lo, Y.S.; Lin, C.C.; Chuang, Y.C.; Hsieh, M.J.; Chen, M.K. Sodium Danshensu Inhibits Oral Cancer Cell Migration and Invasion by Modulating P38 Signaling Pathway. Front. Endocrinol. 2020, 11, 568436. [Google Scholar] [CrossRef] [PubMed]
  56. Li, Z.; Cao, L.; Yang, C.; Liu, T.; Zhao, H.; Luo, X.; Chen, Q. Protocatechuic Acid-Based Supramolecular Hydrogel Targets SerpinB9 to Achieve Local Chemotherapy for OSCC. ACS Appl. Mater. Interfaces 2022, 14, 32. [Google Scholar] [CrossRef] [PubMed]
  57. Gkouveris, I.; Nikitakis, N.; Karanikou, M.; Rassidakis, G.; Sklavounou, A. JNK1/2 Expression and Modulation of STAT3 Signaling in Oral Cancer. Oncol. Lett. 2016, 12, 699–706. [Google Scholar] [CrossRef]
  58. charoenrat, P.; Rhys-Evans, P.H.; Eccles, S.A. Expression of Matrix Metalloproteinases and Their Inhibitors Correlates with Invasion and Metastasis in Squamous Cell Carcinoma of the Head and Neck. Arch. Otolaryngol. Head Neck Surg. 2001, 127, 813–820. [Google Scholar]
  59. Carmeliet, P. VEGF as a Key Mediator of Angiogenesis in Cancer. Oncology 2005, 69, 4–10. [Google Scholar] [CrossRef] [PubMed]
  60. Pepper, M.S.; Ferrara, N.; Orci, L.; Montesano, R. Potent Synergism between Vascular Endothelial Growth Factor and Basic Fibroblast Growth Factor in the Induction of Angiogenesis in Vitro. Biochem. Biophys. Res. Commun. 1992, 189, 824–831. [Google Scholar] [CrossRef]
  61. Zhang, Y.H.; Wei, W.; Xu, H.; Wang, Y.Y.; Wu, W.X. Inducing Effects of Hepatocyte Growth Factor on the Expression of Vascular Endothelial Growth Factor in Human Colorectal Carcinoma Cells through MEK and PI3K Signaling Pathways. Chin. Med. J. 2007, 120, 743–748. [Google Scholar] [CrossRef]
  62. Lun, M.; Zhang, P.L.; Pellitteri, P.K.; Law, A.; Kennedy, T.L.; Brown, R.E. Nuclear Factor-KappaB Pathway as a Therapeutic Target in Head and Neck Squamous Cell Carcinoma: Pharmaceutical and Molecular Validation in Human Cell Lines Using Velcade and SiRNA/NF-ΚB. Ann. Clin. Lab. Sci. 2005, 35, 251–258. [Google Scholar]
  63. Oeckinghaus, A.; Ghosh, S. The NF-ΚB Family of Transcription Factors and Its Regulation. Cold Spring Harb. Perspect. Biol. 2009, 1, a000034. [Google Scholar] [CrossRef]
  64. Wang, J.; Hu, Y.; Yuan, J.; Zhang, Y.; Wang, Y.; Yang, Y.; Alahmadi, T.A.; Ali Alharbi, S.; Zhuang, Z.; Wu, F. Chemomodulatory Effect of Neferine on DMBA-Induced Squamous Cell Carcinogenesis: Biochemical and Molecular Approach. Environ. Toxicol. 2021, 36, 460–471. [Google Scholar] [CrossRef] [PubMed]
  65. Chen, H.; Lo, Y.; Lin, C.; Lee, T.; Leung, W.; Wang, S.; Lin, I.; Lin, M.; Lee, C. Biomedicine & Pharmacotherapy Trichodermin Inhibits the Growth of Oral Cancer through Apoptosis-Induced Mitochondrial Dysfunction and HDAC-2-Mediated Signaling. Biomed. Pharmacother. 2022, 153, 113351. [Google Scholar] [CrossRef] [PubMed]
  66. Ihle, J.N. The Stat Family in Cytokine Signaling. Curr. Opin. Cell Biol. 2001, 13, 211–217. [Google Scholar] [CrossRef] [PubMed]
  67. Sun, S.S.; Zhou, X.; Huang, Y.Y.; Kong, L.P.; Mei, M.; Guo, W.Y.; Zhao, M.H.; Ren, Y.; Shen, Q.; Zhang, L. Targeting STAT3/MiR-21 Axis Inhibits Epithelial-Mesenchymal Transition via Regulating CDK5 in Head and Neck Squamous Cell Carcinoma. Mol. Cancer 2015, 14, 213. [Google Scholar] [CrossRef]
  68. Zhou, X.; Ren, Y.; Liu, A.; Han, L.; Zhang, K.; Li, S.; Li, P.; Li, P.; Kang, C.; Wang, X.; et al. STAT3 Inhibitor WP1066 Attenuates MiRNA-21 to Suppress Human Oral Squamous Cell Carcinoma Growth in Vitro and in Vivo. Oncol. Rep. 2014, 31, 2173–2180. [Google Scholar] [CrossRef]
  69. Riebe, C.; Pries, R.; Schroeder, K.N.; Wollenberg, B. Phosphorylation of STAT3 in Head and Neck Cancer Requires P38 MAPKinase, Whereas Phosphorylation of STAT1 Occurs via a Different Signaling Pathway. Anticancer Res. 2011, 31, 3819–3825. [Google Scholar]
  70. Zhao, C.; Yang, L.; Zhou, F.; Yu, Y.; Du, X.; Xiang, Y.; Li, C.; Huang, X.; Xie, C.; Liu, Z.; et al. Feedback Activation of EGFR Is the Main Cause for STAT3 Inhibition-Irresponsiveness in Pancreatic Cancer Cells. Oncogene 2020, 39, 3997–4013. [Google Scholar] [CrossRef] [PubMed]
  71. Ning, Y.; Cui, Y.; Li, X.; Cao, X.; Chen, A.; Xu, C.; Cao, J.; Luo, X. Co-Culture of Ovarian Cancer Stem-like Cells with Macrophages Induced SKOV3 Cells Stemness via IL-8/STAT3 Signaling. Biomed. Pharmacother. 2018, 103, 262–271. [Google Scholar] [CrossRef]
  72. Yang, J.; Liao, D.; Chen, C.; Liu, Y.; Chuang, T.H.; Xiang, R.; Markowitz, D.; Reisfeld, R.A.; Luo, Y. Tumor-Associated Macrophages Regulate Murine Breast Cancer Stem Cells through a Novel Paracrine Egfr/Stat3/Sox-2 Signaling Pathway. Stem Cells 2013, 31, 248–258. [Google Scholar] [CrossRef]
  73. Chan, C.Y.; Huang, S.Y.; Sheu, J.J.C.; Roth, M.M.; Chou, I.T.; Lien, C.H.; Lee, M.F.; Huang, C.Y. Transcription Factor HBP1 Is a Direct Anti-Cancer Target of Transcription Factor FOXO1 in Invasive Oral Cancer. Oncotarget 2017, 8, 14537–14548. [Google Scholar] [CrossRef]
  74. Roh, J.L.; Kang, S.K.; Minn, I.; Califano, J.A.; Sidransky, D.; Koch, W.M. P53-Reactivating Small Molecules Induce Apoptosis and Enhance Chemotherapeutic Cytotoxicity in Head and Neck Squamous Cell Carcinoma. Oral Oncol. 2011, 47, 8–15. [Google Scholar] [CrossRef] [PubMed]
  75. Khan, Z.; Tiwari, R.P.; Mulherker, R.; Sah, N.K.; Prasad, G.B.; Shrivastava, B.R.; Bisen, P.S. Detection of Survivin and P53 in Human Oral Cancer: Correlation with Clinicopathological Findings. Head Neck 2009, 31, 1039–1048. [Google Scholar] [CrossRef] [PubMed]
  76. The Cancer Genome Atlas Network. Comprehensive Genomic Characterization of Head and Neck Squamous Cell Carcinomas. Nature 2015, 517, 576–582. [Google Scholar] [CrossRef]
  77. Tanida, I.; Ueno, T.; Kominami, E. LC3 and Autophagy. Methods Mol. Biol. 2008, 445, 77–88. [Google Scholar] [CrossRef] [PubMed]
  78. Kabeya, Y.; Mizushima, N.; Ueno, T.; Yamamoto, A.; Kirisako, T.; Noda, T.; Kominami, E.; Ohsumi, Y.; Yoshimori, T. LC3, a Mammalian Homologue of Yeast Apg8p, Is Localized in Autophagosome Membranes after Processing. EMBO J. 2000, 19, 5720–5728. [Google Scholar] [CrossRef] [PubMed]
  79. Yun, C.W.; Lee, S.H. The Roles of Autophagy in Cancer. Int. J. Mol. Sci. 2018, 19, 3466. [Google Scholar] [CrossRef]
  80. Khan, T.; Relitti, N.; Brindisi, M.; Magnano, S.; Zisterer, D.; Gemma, S.; Butini, S.; Campiani, G. Autophagy Modulators for the Treatment of Oral and Esophageal Squamous Cell Carcinomas. Med. Res. Rev. 2020, 40, 1002–1060. [Google Scholar] [CrossRef]
  81. Lee, M.; Nam, H.Y.; Kang, H.B.; Lee, W.H.; Lee, G.H.; Sung, G.J.; Han, M.W.; Cho, K.J.; Chang, E.J.; Choi, K.C.; et al. Epigenetic Regulation of P62/SQSTM1 Overcomes the Radioresistance of Head and Neck Cancer Cells via Autophagy-Dependent Senescence Induction. Cell Death Dis. 2021, 12, 250. [Google Scholar] [CrossRef]
  82. Eisenberg-Lerner, A.; Kimchi, A. The Paradox of Autophagy and Its Implication in Cancer Etiology and Therapy. Apoptosis 2009, 14, 376–391. [Google Scholar] [CrossRef]
  83. Garcia-Medina, R.; Gounon, P.; Chiche, J.; Pouysse, J.; Mazure, N.M. Hypoxia-Induced Autophagy Is Mediated through Hypoxia-Inducible Factor Induction of BNIP3 and BNIP3L via Their BH3 Domains. Mol. Cell. Biol. 2009, 29, 2570–2581. [Google Scholar] [CrossRef]
  84. Wei, H.; Wei, S.; Gan, B.; Peng, X.; Zou, W.; Guan, J. Suppression of Autophagy by FIP200 Deletion Inhibits Mammary Tumorigenesis. Genes Dev. 2011, 37, 1510–1527. [Google Scholar] [CrossRef] [PubMed]
  85. Singh, S.S.; Vats, S.; Chia, A.Y.; Zea, T.; Shuo, T.; Mei, D.; Ong, S.; Arfuso, F.; Yap, C.T.; Cher, B.; et al. Dual Role of Autophagy in Hallmarks of Cancer. Oncogene 2018, 37, 1142–1158. [Google Scholar] [CrossRef] [PubMed]
  86. Muz, B.; Azab, A.K. The Role of Hypoxia in Cancer Progression, Angiogenesis, Metastasis, and Resistance to Therapy. Hypoxia 2015, 3, 83–92. [Google Scholar] [CrossRef] [PubMed]
  87. Oltval, Z.N.; Milliman, C.L.; Korsmeyer, S.J. Bcl-2 Heterodimerizes in Vivo with a Conserved Homolog, Bax, That Accelerates Programed Cell Death. Cell 1993, 74, 609–619. [Google Scholar] [CrossRef]
  88. Hockenbery, D.; Nuñez, G.; Milliman, C.; Schreiber, R.D.; Korsmeyer, S.J. Bcl-2 Is an Inner Mitochondrial Membrane Protein That Blocks Programmed Cell Death. Nature 1990, 348, 334–336. [Google Scholar] [CrossRef]
  89. Loro, L.L.; Vintermyr, O.K.; Liavaag, P.G.; Jonsson, R.; Johannessen, A.C. Oral Squamous Cell Carcinoma Is Associated with Decreased Bcl-2/Bax Expression Ratio and Increased Apoptosis. Hum. Pathol. 1999, 30, 1097–1105. [Google Scholar] [CrossRef]
  90. Zou, H.; Henzel, W.J.; Liu, X. Apaf-1, a Human Protein Homologous to C. Elegans CED-4, Participates in cytochrome c-dependent activation of caspase-3. Cell 1997, 90, 405–413. [Google Scholar] [CrossRef]
  91. Kubina, R.; Krzykawski, K.; Dziedzic, A. Kaempferol and Fisetin-Related Signaling Pathways Induce Apoptosis in Head and Neck Cancer Cells. Cells 2023, 12, 1568. [Google Scholar] [CrossRef]
  92. Xiao, S.; Chen, F.; Gao, C. Antitumor Activity of 4-O-Methylhonokiol in Human Oral Cancer Cells Is Mediated via ROS Generation, Disruption of Mitochondrial Potential, Cell Cycle Arrest and Modulation of Bcl-2/Bax Proteins. J. BUON 2017, 22, 1577–1581. [Google Scholar]
  93. Liu, J.; Uematsu, H.; Tsuchida, N.; Ikeda, M.A. Essential Role of Caspase-8 in P53/P73-Dependent Apoptosis Induced by Etoposide in Head and Neck Carcinoma Cells. Mol. Cancer 2011, 10, 95. [Google Scholar] [CrossRef]
  94. Liu, J.; Uematsu, H.; Tsuchida, N.; Ikeda, M.A. Association of Caspase-8 Mutation with Chemoresistance to Cisplatin in HOC313 Head and Neck Squamous Cell Carcinoma Cells. Biochem. Biophys. Res. Commun. 2009, 390, 989–994. [Google Scholar] [CrossRef] [PubMed]
  95. Coleman, S.C.; Stewart, Z.A.; Day, T.A.; Netterville, J.L.; Burkey, B.B.; Pietenpol, J.A. Analysis of Cell-Cycle Checkpoint Pathways in Head and Neck Cancer Cell Lines: Implications for Therapeutic Strategies. Arch. Otolaryngol.-Head Neck Surg. 2002, 128, 167–176. [Google Scholar] [CrossRef] [PubMed]
  96. Norbury, C.; Nurse, P. Animal Cell Cycles and Their Control. Annu. Rev. Biochem. 1992, 61, 441–470. [Google Scholar] [CrossRef] [PubMed]
  97. Xiong, Y.; Connolly, T.; Futcher, B.; Beach, D. Human D-Type Cyclin. Cell 1991, 65, 691–699. [Google Scholar] [CrossRef]
  98. Matsushime, H.; Ewen, M.E.; Strom, D.K.; Kato, J.Y.; Hanks, S.K.; Roussel, M.F.; Sherr, C.J. Identification and Properties of an Atypical Catalytic Subunit (P34PSK-J3/Cdk4) for Mammalian D Type G1 Cyclins. Cell 1992, 71, 323–334. [Google Scholar] [CrossRef]
  99. Lea, N.C.; Orr, S.J.; Stoeber, K.; Williams, G.H.; Lam, E.W.-F.; Ibrahim, M.A.A.; Mufti, G.J.; Thomas, N.S.B. Commitment Point during G0→G1 That Controls Entry into the Cell Cycle. Mol. Cell. Biol. 2003, 23, 2351–2361. [Google Scholar] [CrossRef]
  100. Duronio, R.J.; Brook, A.; Dyson, N.; O’Farrell, P.H. E2F-Induced S Phase Requires Cyclin E. Genes Dev. 1996, 10, 2505–2513. [Google Scholar] [CrossRef]
  101. Oakes, V.; Wang, W.; Harrington, B.; Lee, W.J.; Beamish, H.; Chia, K.M.; Pinder, A.; Goto, H.; Inagaki, M.; Pavey, S.; et al. Cyclin A/Cdk2 Regulates Cdh1 and Claspin during Late S/G2 Phase of the Cell Cycle. Cell Cycle 2014, 13, 3302–3311. [Google Scholar] [CrossRef]
  102. Vigneron, S.; Sundermann, L.; Labbé, J.C.; Pintard, L.; Radulescu, O.; Castro, A.; Lorca, T. Cyclin A-Cdk1-Dependent Phosphorylation of Bora Is the Triggering Factor Promoting Mitotic Entry. Dev. Cell 2018, 45, 637–650.e7. [Google Scholar] [CrossRef]
  103. Gavet, O.; Pines, J. Activation of Cyclin B1-Cdk1 Synchronizes Events in the Nucleus and the Cytoplasm at Mitosis. J. Cell Biol. 2010, 189, 247–259. [Google Scholar] [CrossRef]
  104. Li, B.; Zhou, P.; Xu, K.; Chen, T.; Jiao, J.; Wei, H.; Yang, X.; Xu, W.; Wan, W.; Xiao, J. Metformin Induces Cell Cycle Arrest, Apoptosis and Autophagy through ROS/JNK Signaling Pathway in Human Osteosarcoma. Int. J. Biol. Sci. 2020, 16, 74–84. [Google Scholar] [CrossRef]
  105. McConnell, B.B.; Gregory, F.J.; Stott, F.J.; Hara, E.; Peters, G. Induced Expression of P16 INK4a Inhibits Both CDK4- and CDK2-Associated Kinase Activity by Reassortment of Cyclin-CDK-Inhibitor Complexes. Mol. Cell. Biol. 1999, 19, 1981–1989. [Google Scholar] [CrossRef] [PubMed]
  106. Qin, X.Q.; Livingston, D.M.; Kaelin, W.G.; Adams, P.D. Deregulated Transcription Factor E2F-1 Expression Leads to S-Phase Entry and P53-Mediated Apoptosis. Proc. Natl. Acad. Sci. USA 1994, 91, 10918–10922. [Google Scholar] [CrossRef] [PubMed]
  107. Huang, Y.C.; Lee, P.C.; Wang, J.J.; Hsu, Y.C. Anticancer Effect and Mechanism of Hydroxygenkwanin in Oral Squamous Cell Carcinoma. Front. Oncol. 2019, 9, 911. [Google Scholar] [CrossRef] [PubMed]
  108. Bostan, M.; Petrică-Matei, G.G.; Radu, N.; Hainarosie, R.; Stefanescu, C.D.; Diaconu, C.C.; Roman, V. The Effect of Resveratrol or Curcumin on Head and Neck Cancer Cells Sensitivity to the Cytotoxic Effects of Cisplatin. Nutrients 2020, 12, 2596. [Google Scholar] [CrossRef]
  109. Da Fonseca, A.C.C.; de Queiroz, L.N.; Sales Felisberto, J.; Jessé Ramos, Y.; Mesquita Marques, A.; Wermelinger, G.F.; Pontes, B.; de Lima Moreira, D.; Robbs, B.K. Cytotoxic Effect of Pure Compounds from Piper Rivinoides Kunth against Oral Squamous Cell Carcinoma. Nat. Prod. Res. 2021, 35, 6163–6167. [Google Scholar] [CrossRef]
  110. De Campos, P.S.; Matte, B.F.; Diel, L.F.; Jesus, L.H.; Bernardi, L.; Alves, A.M.; Rados, P.V.; Lamers, M.L. Low Doses of Curcuma Longa Modulates Cell Migration and Cell–Cell Adhesion. Phyther. Res. 2017, 31, 1433–1440. [Google Scholar] [CrossRef]
  111. Chourasia, N.R.; Borle, R.M.; Vastani, A. Concomitant Association of Oral Submucous Fibrosis and Oral Squamous Cell Carcinoma and Incidence of Malignant Transformation of Oral Submucous Fibrosis in a Population of Central India: A Retrospective Study. J. Maxillofac. Oral Surg. 2015, 14, 902–906. [Google Scholar] [CrossRef]
  112. Evren, I.; Brouns, E.R.; Wils, L.J.; Poell, J.B.; Peeters, C.F.W.; Brakenhoff, R.H.; Bloemena, E.; de Visscher, J.G.A.M. Annual Malignant Transformation Rate of Oral Leukoplakia Remains Consistent: A Long-Term Follow-up Study. Oral Oncol. 2020, 110, 105014. [Google Scholar] [CrossRef]
  113. Iocca, O.; Sollecito, T.P.; Alawi, F.; Weinstein, G.S.; Newman, J.G.; De Virgilio, A.; Di Maio, P.; Spriano, G.; Pardiñas López, S.; Shanti, R.M. Potentially Malignant Disorders of the Oral Cavity and Oral Dysplasia: A Systematic Review and Meta-Analysis of Malignant Transformation Rate by Subtype. Head Neck 2020, 42, 539–555. [Google Scholar] [CrossRef]
  114. Lodi, G.; Franchini, R.; Warnakulasuriya, S.; Varoni, E.M.; Sardella, A.; Kerr, A.R.; Carrassi, A.; MacDonald, L.C.; Worthington, H.V.; Mauleffinch, L.F. Interventions for Treating Oral Leukoplakia to Prevent Oral Cancer (Review). Cochrane Database Syst. Rev. 2016, 7, CD001829. [Google Scholar] [CrossRef] [PubMed]
  115. Papadimitrakopoulou, V.A.; William, W.N.; Dannenberg, A.J.; Lippman, S.M.; Lee, J.J.; Ondrey, F.G.; Peterson, D.E.; Feng, L.; Atwell, A.; El-Naggar, A.K.; et al. Pilot Randomized Phase II Study of Celecoxib in Oral Premalignant Lesions. Clin. Cancer Res. 2008, 14, 2095–2101. [Google Scholar] [CrossRef] [PubMed]
  116. William, W.N.; Papadimitrakopoulou, V.; Lee, J.J.; Mao, L.; Cohen, E.E.W.; Lin, H.Y.; Gillenwater, A.M.; Martin, J.W.; Lingen, M.W.; Boyle, J.O.; et al. Erlotinib and the Risk of Oral Cancer the Erlotinib Prevention of Oral Cancer (EPOC) Randomized Clinical Trial. JAMA Oncol. 2016, 2, 209–216. [Google Scholar] [CrossRef]
  117. Gutkind, J.S.; Molinolo, A.A.; Wu, X.; Wang, Z.; Nachmanson, D.; Harismendy, O.; Alexandrov, L.B.; Wuertz, B.R.; Ondrey, F.G.; Laronde, D.; et al. Inhibition of MTOR Signaling and Clinical Activity of Metformin in Oral Premalignant Lesions. JCI Insight 2021, 6, 147096. [Google Scholar] [CrossRef] [PubMed]
  118. Tian, H.; Lyu, Y.; Yang, Y.G.; Hu, Z. Humanized Rodent Models for Cancer Research. Front. Oncol. 2020, 10, 1696. [Google Scholar] [CrossRef]
  119. Bouaoud, J.; De Souza, G.; Darido, C.; Tortereau, A.; Elkabets, M.; Bertolus, C.; Saintigny, P. The 4-NQO Mouse Model: An Update on a Well-Established in Vivo Model of Oral Carcinogenesis. Methods Cell Biol. 2021, 163, 197–229. [Google Scholar] [CrossRef]
  120. Schoop, R.A.L.; Noteborn, M.H.M.; Baatenburg De Jong, R.J. A Mouse Model for Oral Squamous Cell Carcinoma. J. Mol. Histol. 2009, 40, 177–181. [Google Scholar] [CrossRef]
  121. Wang, J.; Wang, S.; Wang, Y.; Wang, L.; Xia, Q.; Tian, Z.; Guan, X. Chemopreventive Effect of Modified Zengshengping on Oral Cancer in a Hamster Model and Assessment of Its Effect on Liver. J. Ethnopharmacol. 2020, 255, 112774. [Google Scholar] [CrossRef]
  122. Dai, Z.; Zhu, P.F.; Liu, H.; Li, X.C.; Zhu, Y.Y.; Liu, Y.Y.; Shi, X.L.; Chen, W.D.; Liu, Y.P.; Zhao, Y.L.; et al. Discovery of Potent Immune-Modulating Molecule Taccaoside A against Cancers from Structures-Active Relationships of Natural Steroidal Saponins. Phytomedicine 2022, 104, 154335. [Google Scholar] [CrossRef]
  123. Cattaneo, C.M.; Dijkstra, K.K.; Fanchi, L.F.; Kelderman, S.; Kaing, S.; van Rooij, N.; van den Brink, S.; Schumacher, T.N.; Voest, E.E. Tumor Organoid–T-Cell Coculture Systems. Nat. Protoc. 2020, 15, 15–39. [Google Scholar] [CrossRef]
  124. Neal, J.T.; Li, X.; Zhu, J.; Giangarra, V.; Grzeskowiak, C.L.; Ju, J.; Liu, I.H.; Chiou, S.H.; Salahudeen, A.A.; Smith, A.R.; et al. Organoid Modeling of the Tumor Immune Microenvironment. Cell 2018, 175, 1972–1988.e16. [Google Scholar] [CrossRef] [PubMed]
  125. Zhou, Z.; Van der Jeught, K.; Fang, Y.; Yu, T.; Li, Y.; Ao, Z.; Liu, S.; Zhang, L.; Yang, Y.; Eyvani, H.; et al. An Organoid-Based Screen for Epigenetic Inhibitors That Stimulate Antigen Presentation and Potentiate T-Cell-Mediated Cytotoxicity. Nat. Biomed. Eng. 2021, 5, 1320–1335. [Google Scholar] [CrossRef] [PubMed]
  126. Verma, A.; Vincent-chong, V.K.; Dejong, H.; Hershberger, P.A.; Seshadri, M. Impact of Dietary Vitamin D on Initiation and Progression of Oral Cancer. J. Steroid Biochem. Mol. Biol. 2020, 199, 105603. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Natural products involved in preclinical trials involving HNSCC cell lines and animal models.
Figure 1. Natural products involved in preclinical trials involving HNSCC cell lines and animal models.
Biomedicines 11 02359 g001
Figure 2. Various natural products and their respective molecular mechanisms against HNSCC: chemoprevention (natural product name in green), chemotherapeutic (natural product name in black).
Figure 2. Various natural products and their respective molecular mechanisms against HNSCC: chemoprevention (natural product name in green), chemotherapeutic (natural product name in black).
Biomedicines 11 02359 g002
Table 1. Chemotherapeutic findings from preclinical studies with natural products involving in vitro HNSCC cell lines.
Table 1. Chemotherapeutic findings from preclinical studies with natural products involving in vitro HNSCC cell lines.
NoAuthorPMIDNatural ProductPreclinical
Model (Cell Lines)
Test & DosageFindingsPathway Related
1Zhang et al.
[49]
31495936FucoidanIn vitro (SCC15, SCC25)
  • Cell Proliferation; MTT assay (IC50 value; 2 µM, 5 µM)
  • Cell Cycle Arrest; flow cytometry (dosage not available)
  • Colony Formation; clonogenic assay (dosage not available)
  • Apoptosis; Annexin V assay (dosage not available)
  • Cell Migration and Invasion; transwell assay, wound healing assay (dosage not available)
  • Reduced cell viability and inducing cell arrest (G2 Phase)
  • Inducing intrinsic apoptosis via upregulation of Bcl-2 associated X protein (Bax) expression
  • Downregulating phosphorylation of protein kinase B (p-Akt) and cyclin-dependent kinase 1 (Cdk1)
  • Inhibiting cell migration and invasion via downregulation of metalloproteinases (MMP-9) and MMP-2 expression
  • Upregulate circRNA filament A (circFLNA) expression, mediating signaling molecules
Akt
Bax/Bcl-2
2Fonseca et al.
[109]
33078660ConocarpanIn vitro (SCC9, SCC4, SCC25)
  • Cell Viability; MTT assay (1.2–300 µM)
  • Apoptosis; cell morphology observation
  • Reduced cell viability and promote cell apoptosis
  • Activating Caspase-3 expression and inducing pyknotic nuclei
Caspase
3Huang et al.
[107]
31620368HydroxygenkwaninIn vitro (SAS & OECM1)
  • Cell Proliferation; MTT assay (25, 50 & 75 µM)
  • Cell Cycle Arrest; flow cytometry (25, 50 & 75 µM)
  • Colony formation; clonogenic assay (25, 50 & 75 µM)
  • Cell Migration and Invasion; wound healing assay and invasion assay (25 & 50 µM)
  • Reduced cell growth and colony formation
  • Activated p21 and inhibited Cdk2 expression, inducing cell cycle arrest
  • Upregulating poly (ADP-ribose) polymerase (PARP) cleavage and phosphorylated X-linked inhibitor of apoptosis protein (p-H2AX)
  • Induce cell apoptosis via intrinsic pathway involving Caspase-9
  • Inhibition of cell invasion and migration via downregulation of Vimentin
Caspase
4Lin et al.
[46]
32825464IlimaquinoneIn vitro (SCC4, SCC2095)
  • Cell Viability; MTT assay (IC50 value; 7.5 µM, 8.5 µM)
  • Apoptosis; Annexin V assay (5, 10, 20 & 30 µM)
  • Autophagy; autophagic vesicle detection (2.5, 5, 10 & 20 µM)
  • Reduced cell viability
  • Upregulated caspase-3, caspase-8, caspase-9 and PARP cleavage leading to apoptosis
  • Upregulated proapoptotic protein Bax and p-p53, and downregulated anti-apoptotic protein myeloid leukemia cell differentiation protein (Mcl-1), B-cell lymphoma 2 (Bcl-2), and apoptotic inhibitor protein Survivin
  • Downregulated p-Akt and hypoxia-inducible factor 1-alpha (H1F-1α), mediating cell migration
  • Upregulated p-H2AX, as regulation due to increased reactive oxygen species (ROS) generation
  • Induced autophagy by upregulating light chain 3 (LC3-II) and autophagy related 5 (Atg5) expression
Akt
Caspase
Bax/Bcl-2
5Sophia et al.
[47]
30352996NimbolideIn vitro (SCC131, SCC4)
  • Cell Viability; MTT assay (IC50 value; 6 µM, 6.2 µM)
  • Cell Cycle Arrest; flow cytometry (6 µM, 6.2 µM)
  • Apoptosis; Annexin V, nuclear morphology, mitochondrial transmembrane potential (6 µM, 6.2 µM)
  • Autophagy; autophagic vesicle detection (6 µM, 6.2 µM)
  • Reduced cell viability
  • Upregulated Bax and downregulated Bcl-2, leading to intrinsic apoptosis
  • Upregulated cleaved Caspase-9 and Caspase-3 expression
  • Induced conversion of LC3-I to LC3-II, inducing autophagy
  • Downregulating p-Akt and upregulation of phosphorylated glycogen synthase kinase 3 beta (p-GSK-3β), inhibiting phosphoinositide 3-kinase/protein kinase B/glycogen synthase kinase 3 beta (PI3K/Akt/GSK-3β)
PI3K/Akt/GSK-3β
Caspase
Bax/Bcl-2
6Xiao et al.
[92]
293323554-O-methylhonokoilIn vitro (PE/CA-PJ41)
  • Cell Viability; MTT assay (IC50 value; 2.5 µM)
  • Cell Cycle Arrest; flow cytometry (1, 2.5 & 5 µM)
  • Induced G2/M cell cycle arrest and apoptosis
  • Upregulated formation of intracellular ROS, leading to ROS-mediated reduction in mitochondrial membrane potential, inducing intrinsic apoptosis
  • Upregulated Bax and downregulated Bcl-2 expression, leading to apoptosis
Bax/Bcl-2
7Kumar et al.
[55]
33101201Sodium DanshensuIn vitro (FaDu, CA9-22)
  • Cell Viability; MTT assay (50 µM)
  • Cell Migration and Invasion; wound healing, migration, and invasion assay (25, 50 & 100 µM)
  • Reduced cell motility, migration and invasion
  • Upregulation of E-cadherin and zonula occludens-1 (ZO-1) expression, and downregulation of MMP-2, Vimentin and N-cadherin expression, leading to anti-migratory and anti-invasive effect
  • Downregulate p38 phosphorylation leading to inhibition of mitogen-activated protein kinase (MAPK) signaling pathway and downregulation of extracellular signal-regulated protein kinase (ERK1/2)
MAPK/ERK
8Aswathy et al.
[50]
33860206[(3β)-3-hydroxy-lup-20(29)-en-28-oic acid]In vitro (SAS)
  • Cell Proliferation; MTT assay (IC50 value; 6 µM)
  • Colony Formation; colony-forming assay (10 & 15 µM)
  • Cell Cycle Arrest; flow cytometry (5, 10 & 20 µM)
  • Apoptosis; Annexin V (50 µM)
  • Cell Migration; cell migration assay (2.5 & 5 µM)
  • Reduced colony formation and migration and induced apoptosis
  • Downregulated vascular endothelial growth factor (VEGF) and MMP-2 expression, via Akt/mTOR pathway
Akt/mTOR
JAK/STAT3
VEGF
NF-κB
Seco-A-ring oleananeIn vitro (SAS)
  • Cell Proliferation; MTT assay (IC50 value; 20 µM)
  • Colony Formation; colony-forming assay (10 & 15 µM)
  • Cell Cycle Arrest; flow cytometry (1, 3 & 5 µM)
  • Apoptosis; Annexin V (50 µM)
  • Cell Migration; cell migration assay (2.5 & 5 µM)
  • Reduced colony formation and migration and induced apoptosis
  • Downregulated cyclooxygenase-2 (Cox-2), Survivin, MMP-2 and VEGF, via nuclear factor kappa light chain enhancer of activated B cells (NF-κB), mTOR and STAT3 pathway
9Aggarwal et al.
[48]
35774603CinnamaldehydeIn vitro (SCC9, SCC25)
  • Cell Viability; MTT assay (IC50 value; 40 µM, 45 µM)
  • Colony formation; clonogenic assay (40 µM, 45 µM)
  • Cell Cycle Arrest; flow cytometry (40 µM, 45 µM)
  • Apoptosis; Annexin V (40 µM, 45 µM)
  • Cell Invasion; matrigel cell invasion assay (40 µM, 45 µM)
  • Reduced cell viability and inhibited proliferation, migration and invasion
  • Induced cell cycle arrest at G2/M and S-phase
  • Induces autophagy
  • Inhibited nuclear translocation of NF-κB from cytoplasm
  • Cinnamaldehyde shows binding affinity with MAPK-p38α and dihydrofolate reductase (DHFR)
  • Downregulating expression of NF-κB/p65, Cox-2, p110a, cyclin-D1, VEGF, Akt, mTOR, p-mTOR, and Bcl-2, and upregulating beclin-1 expression
PI3K/Akt/mTOR
NF-κB
MAPK
10Kubina et al.
[91]
37371038Kaempferol & FisetinIn vitro (SCC9, SCC25)
  • Cell Proliferation; WST-1 assay (IC50 value kaempferol; 45.03 µM, 49.90 µM & fisetin; 38.85 µM, 62.34 µM)
  • Cell Cycle Arrest; flow cytometry (1/2 and ¼ IC50 value)
  • Apoptosis; Annexin V Detection Kit (1/2 and ¼ IC50 value)
  • Cell Migration; wound healing assay (1/2 and ¼ IC50 value)
  • Inhibited cell proliferation and migration
  • Induced apoptosis by activation of Caspase-3 and decreased potential of mitochondrial membrane
  • Downregulation of Bcl-2
  • Fisetin upregulate cytochrome c
  • Kaempferol induced cell cycle arrest at S phase
Bax/Bcl-2
11Bostan et al.
[108]
32859062Resveratrol & CurcuminIn vitro (PE/CA-PJ49)
  • Cell Viability; MTT assay (IC50 value resveratrol; 46.8 µM & curcumin; 16.3 µM)
  • Cell Proliferation; cell proliferation assay (resveratrol; 40 µM & curcumin; 15 µM)
  • Cell Cycle Arrest; flow cytometry (resveratrol; 40 µM & curcumin; 15 µM)
  • Apoptosis; Annexin V Detection Kit (resveratrol; 40 µM & curcumin; 15 µM)
  • Inhibited cell proliferation
  • Induced apoptosis
  • Amplifying effect of low concentration of cisplatin on inhibition of cell proliferation, and induction of apoptosis and cell cycle arrest
  • Upregulation of p21
Caspase
Table 2. Chemotherapeutic findings from preclinical studies with natural products involving in vitro HNSCC cell lines and in vivo xenograft models.
Table 2. Chemotherapeutic findings from preclinical studies with natural products involving in vitro HNSCC cell lines and in vivo xenograft models.
NoAuthorPMIDNatural ProductPreclinical
Model
Test & DosageFindingsPathway Related
1Zhao et al.
[41]
34175854ActeinIn vitro (CAL27, SCC9)
  • Cell Proliferation; CCK-8 Assay (IC50 value; 15 µM, 12 µM)
  • Cell Cycle Arrest; flow cytometry (7, 15 & 30 µM)
  • Apoptosis; Annexin V assay (7, 15 & 30 µM)
  • Reduced cell viability and induced cell cycle arrest
  • Induced apoptosis via upregulated Bax and downregulated Bcl-2
  • Downregulated Survivin and upregulated p21 and Bim
  • Downregulating Akt and upregulating forkhead box protein O1 (FoxO1), inhibiting Akt/FoxO1 pathway
Akt/FoxO1
Bax/Bcl-2
In vivo (CAL27)
C57BL/6 mice
  • Mammary fat pad subcutaneous inoculation
  • Intragastrically administration of 10, 20, and 50 mg/kg dosage
  • Impaired tumor growth
2Zhang et al.
[42]
32267053SalicylateIn vitro (SAS)
  • Cell Proliferation; MTS Assay (2500, 5000 and 10,000 µM)
  • Cell Cycle Arrest; flow cytometry (5000 µM)
  • Tumorsphere Formation; tumorsphere formation assay (5000 µM)
  • Combination with cisplatin enhanced the cytotoxicity and induced cell apoptosis
  • Downregulated p-Akt, mTOR, p-S6 and p-70S6, inhibiting Akt/mTOR signaling pathway
  • Downregulated Jagged-1, SRY-box transcription factor 9 (SOX9), yes-associated protein-1 (YAP-1), sonic hedgehog protein (Shh), aldehyde dehydrogenase (ALDH), oxtamer-binding transcription factor 4 (OCT4) expression
Akt/mTOR
In vivo (SAS)
Nude mice
  • Subcutaneous inoculation
  • Oral administration of 3 mg/kg dosage
  • Monotherapy impaired tumor growth
  • Combination with cisplatin enhanced tumor growth inhibition
3Li et al.
[43]
32424132Tanshinone IIAIn vitro (CAL27, SCC9, SCC15, SCC25)
  • Cell Viability; MTS assay (2 & 5 µM)
  • Reducing cell viability
  • Inducing intrinsic apoptosis via downregulation of p-Akt, c-Myc and hexokinase 2 (HK2)
  • Inhibiting glycolysis of SCC by downregulating expression of HK2
  • Promoting FBW7 E3 ligase interaction with c-Myc, shortening half-life of c-Myc
Akt/c-Myc
In vivo (CAL27, SCC15)
Athymic nude mice
  • Right flank subcutaneous inoculation
  • Intraperitoneal administration of 10 mg/kg dosage
  • Reducing population of Ki-67 positive cells
  • Downregulating p-Akt, c-Myc and HK2
  • Impaired tumor growth
4Li et al.
[44]
32410646XanthohumolIn vitro (CAL27, SCC9, SCC15, SCC25)
  • Cell Viability; MTS assay (1, 2 & 5 µM)
  • Reducing cell viability and colony formation
  • Regulating Akt/Wee1/Cdk1 signaling pathway
  • Upregulating Bax on mitochondria, PARP cleavage and Caspase-3, leading to intrinsic apoptosis
  • Promote ubiquitination and degradation of Survivin by upregulating FBXL7 E3 protein
Akt/Wee1/Cdk1
Bax/Bcl-2
Caspase
In vivo (CAL27, SCC25)
Athymic nude mice
  • Right flank subcutaneous inoculation
  • Intraperitoneal administration of 10 mg/kg dosage
  • Delayed tumor development and impaired tumor growth
  • Reducing population of Ki-67 positive cells
  • Downregulating p-Akt and Survivin
  • Sensitizing radioresistance cells to radiotherapy
5Huang et al.
[45]
29363886HonokiolIn vitro (OC2, OCSL)
  • Cell Viability; CCK-8 assay (IC50 value; 35 µM, 22 µM)
  • Cell Cycle Arrest; flow cytometry (25 & 40 µM, 15 & 30 µM)
  • Apoptosis; Annexin V assay (25 µM, 15 µM)
  • Reduced cell growth
  • Induced cell cycle arrest at G0/G1 phase via upregulating p21 and p27, accumulation of cyclin-E, and downregulating Cdk2, Cdk4 and cyclin-D1
  • Inhibiting MAPK pathway by downregulating p-Akt and p-mTOR
  • Inducing autophagy via activation of LC3-II
  • Synergic therapeutic effect with Fluorouracil
Akt/mTOR
MAPK
In vivo (SAS)
BALB/c nude mice, AnN.Cg-Foxn1nu/CrlNarl
  • Right flank subcutaneous inoculation
  • Oral administration of 5 and 15 mg/kg dosage
  • Inhibiting tumor growth and inducing apoptosis
6Chen et al.
[65]
35785707TrichoderminIn vitro (Ca922, HSC3)
  • Cell Viability; MTT assay (IC50 value; 9.65 µM, 11.49 µM)
  • Colony Formation; clonogenic assay (3 & 10 µM)
  • Cell Cycle Arrest; flow cytometry (3 & 10 µM)
  • Apoptosis; Annexin V assay; nuclear condensation observation (3 & 10 µM)
  • Cell Migration and Invasion; transwell assay (3 & 10 µM)
  • Reduced cell viability, migration and invasive
  • Downregulation of MMP-9, inhibiting cell migration and invasion
  • Downregulation of cyclin A, cyclin D1, Cdk1/Cdk2 and Cdk4 expression, leading to G2/M cell cycle arrest
  • Inducing apoptosis by upregulating Caspase-3 and cleaved PARP expression
  • Reduces mitochondrial membrane potential, basal respiration, ATP production, maximum respiration and proton leak, inducing intrinsic apoptosis
  • Downregulated expression of HDAC-2, phosphorylated STAT3 and NF-κB
HDAC-2
Caspase
In vivo (HSC3)
Zebrafish
  • Embryo tumor transplantation
  • Embryo submerged in dosage of 3 and 10 µg/mL solution
  • Inhibited tumor growth
7Wang et al.
[24]
31750253PsorachromeneIn vitro (SAS, OECM1)
  • Cell Viability; sulforhodamine B assay (25 & 50 µM)
  • Cell Cycle Arrest; flow cytometry (25, 50 & 75 µM)
  • Apoptosis; TUNEL assay (50 µM)
  • Cell Migration and Invasion; wound healing assay and invasion assay (25, 50 & 75 µM)
  • Reduced cell growth and colony formation
  • Induced cell cycle arrest in G2 phase
  • Induced apoptosis via activation of Caspase-9 and cleavage PARP
  • Inhibited cell migration and invasion, via downregulating EMT-promoting protein, Vimentin, Slug and EGFR signaling pathway
  • Synergic therapeutic effect with Cisplatin and Doxorubicin
  • Upregulating growth arrest specific 5 (GAS5), inhibiting cell growth and metastasis
EGFR
EMT-related
Caspase
SLUG
In vivo (SAS)
BALB/c nude mice
  • Right flank subcutaneous inoculation
  • Intraperitoneal administration of 100 µL dosage
  • Downregulated EMT-promoting protein and EGFR signaling pathway
  • Reduced tumor growth
8Li et al.
[56]
35904511Protocatechuic acidIn vitro (HSC3, CAL27)
  • Cell Proliferation; CCK-8 assay (250 & 500 µM)
  • Cell Cycle Arrest; flow cytometry (250 & 500 µM)
  • Apoptosis; TUNEL assay (250 & 500 µM)
  • Tumorsphere Formation; tumorsphere assay (250, 500 & 1000 µM)
  • Induced cell death by interrupting Serpinb9 and granzyme B (Sb9-GrB) complex formation
  • Induced cell apoptosis via upregulated phosphorylation of c-Jun N-terminal kinase (JNK), p38 and cleaved Caspase-3 expression
  • Upregulated superoxide dismutase and nuclear factor erythroid 2-related factor 2 (Nrf2), reducing the ROS levels
  • Inhibiting cancer stemness
JNK/p38
Caspase
In vivo (CAL27)
Nude mice
  • Right flank subcutaneous inoculation
  • Peritumoral administration of 100 µL dosage
  • Inhibiting tumor growth
  • Reducing population of Ki-67 positive cell
9de Compos et al.
[110]
28782139CurcuminIn vitro (CAL27, SCC25, HACAT, NIH-3T3)
  • Cell Proliferation; NF cell proliferation assay (2, 5, 10, 20, 30, 40 & 50 µM)
  • Cell Apoptosis; Annexin V assay (5 & 50 µM)
  • Cell Migration and Invasion; time-lapse analysis (2 & 5 µM)
  • Tumorsphere Formation; spheroid assay (10,20, 50 & 200 µM)
  • Reducing cell proliferation
  • Reduced migratory rate and impairment on tumor cell directionality
  • Reducing cell-cell adhesion, leading to less homogenous spheroid
In vivo (HNSCC Biopsy)
BALB/c nude mice
  • Right flank subcutaneous inoculation
  • Treatment with 70 mg/kg dosage
  • Inhibiting tumor growth
  • Induced a less aggressive histological phenotype
Table 3. Chemoprevention findings from preclinical studies with natural products on induced carcinogenesis.
Table 3. Chemoprevention findings from preclinical studies with natural products on induced carcinogenesis.
NoAuthorPMIDNatural ProductPreclinical
Model
Test & DosageFindingsPathway Related
1Vincent-Chong et al.
[25]
30875566CalcitriolIn vivo (4NQO-induced carcinogenesis)
C57BL/6NCr mice
  • 0.1 µg dosage, intraperitoneal administration (thrice weekly)
  • Inhibition of 4-nitroquinoline-1-oxide (4NQO)-induced carcinogenesis
  • Reduced incidence of HNSCC induced by 4NQO
  • Increased Ki-67 positive dysplastic epithelium
  • Carcinogenesis influenced by stage of intervention and duration of exposure to calcitriol
2Wang et al.
[64]
33156559NeferineIn vivo (DMBA-induced carcinogenesis)
Syrian hamster
  • 15 mg/kg dosage, intragastric administration (thrice weekly)
  • Increased body weight and suppression on formation of 7,12-Dimethylbenz[a]anthracene (DMBA) induced tumor development
  • Downregulated expression of NF-κB, mutant p53 and PCNA
NF-κB
3Sophia et al.
[47]
30352996NimbolideIn vivo (DMBA-induced carcinogenesis)
Syrian hamster
  • 100 µg/kg dosage, intragastric administration
  • Inhibited PI3K/Akt/GSK-3β signaling pathway
  • Sensitized tumor to apoptosis
PI3K/Akt/GSK-3β
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Liew, Y.X.; Karen-Ng, L.P.; Vincent-Chong, V.K. A Comprehensive Review of Natural Products as Therapeutic or Chemopreventive Agents against Head and Neck Squamous Cell Carcinoma Cells Using Preclinical Models. Biomedicines 2023, 11, 2359. https://doi.org/10.3390/biomedicines11092359

AMA Style

Liew YX, Karen-Ng LP, Vincent-Chong VK. A Comprehensive Review of Natural Products as Therapeutic or Chemopreventive Agents against Head and Neck Squamous Cell Carcinoma Cells Using Preclinical Models. Biomedicines. 2023; 11(9):2359. https://doi.org/10.3390/biomedicines11092359

Chicago/Turabian Style

Liew, Yoon Xuan, Lee Peng Karen-Ng, and Vui King Vincent-Chong. 2023. "A Comprehensive Review of Natural Products as Therapeutic or Chemopreventive Agents against Head and Neck Squamous Cell Carcinoma Cells Using Preclinical Models" Biomedicines 11, no. 9: 2359. https://doi.org/10.3390/biomedicines11092359

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop