Next Article in Journal
Combination of Fetal Fraction Estimators Based on Fragment Lengths and Fragment Counts in Non-Invasive Prenatal Testing
Next Article in Special Issue
Overexpression of Platelet-Derived Growth Factor and Its Receptor Are Correlated with Oral Tumorigenesis and Poor Prognosis in Oral Squamous Cell Carcinoma
Previous Article in Journal
Chlorin e6-Loaded PEG-PCL Nanoemulsion for Photodynamic Therapy and In Vivo Drug Delivery
Previous Article in Special Issue
Oral Submucous Fibrosis: A Review on Etiopathogenesis, Diagnosis, and Therapy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Targeting Cellular Metabolism Modulates Head and Neck Oncogenesis

1
Institute of Oral Biology, School of Dentistry, National Yang-Ming University, Taipei 11221, Taiwan
2
Department of Dentistry, School of Dentistry, National Yang-Ming University, Taipei 11221, Taiwan
3
Department of Stomatology, Taipei Veterans General Hospital, Taipei 11217, Taiwan
4
Cancer Progression Research Center, National Yang-Ming University, Taipei 11221, Taiwan
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2019, 20(16), 3960; https://doi.org/10.3390/ijms20163960
Submission received: 1 August 2019 / Revised: 12 August 2019 / Accepted: 13 August 2019 / Published: 14 August 2019
(This article belongs to the Special Issue Oral Fibrosis and Oral Cancer: From Molecular Targets to Therapeutics)

Abstract

:
Considering the great energy and biomass demand for cell survival, cancer cells exhibit unique metabolic signatures compared to normal cells. Head and neck squamous cell carcinoma (HNSCC) is one of the most prevalent neoplasms worldwide. Recent findings have shown that environmental challenges, as well as intrinsic metabolic manipulations, could modulate HNSCC experimentally and serve as clinic prognostic indicators, suggesting that a better understanding of dynamic metabolic changes during HNSCC development could be of great benefit for developing adjuvant anti-cancer schemes other than conventional therapies. However, the following questions are still poorly understood: (i) how does metabolic reprogramming occur during HNSCC development? (ii) how does the tumorous milieu contribute to HNSCC tumourigenesis? and (iii) at the molecular level, how do various metabolic cues interact with each other to control the oncogenicity and therapeutic sensitivity of HNSCC? In this review article, the regulatory roles of different metabolic pathways in HNSCC and its microenvironment in controlling the malignancy are therefore discussed in the hope of providing a systemic overview regarding what we knew and how cancer metabolism could be translated for the development of anti-cancer therapeutic reagents.

1. Introduction

Malignancies of the head and neck influence a variety of anatomic sites, including the oral cavity, oropharynx, nasopharynx, hypopharynx, larynx, and salivary glands [1]. The oncogenic stimuli of Head and Neck Squamous Cell Carcinomas (HNSCC), including smoking, alcohol consumption, viral infection and an imbalanced metabolism, could lead to genetic mutations and epigenetic modulations that serve as potential triggers for head and neck tumorigenesis [2,3]. Clinical therapeutic regimens for HNSCC patients have been widely discussed; single or combinational treatments of surgery, chemotherapy and radiotherapy are common choices for HNSCC, depending on the tumor sizes, locations, histological subtypes and clinical stages [4,5]. Nevertheless, 5-year survival rates for HNSCC patients are still below 50% and have not changed much over the past 50 years. The poor survival rates could be attributable to the late diagnosis of the disease, lack of better prognostic tools or development of resistance to conventional therapies [6,7,8,9,10].
Although cancer is generally considered to be a genetic disease [11], inconsistencies regarding the somatic nuclear gene theory based on nuclear/cytoplasmic transfer experiments between tumorigenic and non-tumorigenic cells show that tumorigenicity could originate from disrupted metabolic homeostasis [12,13,14]. To meet great demands for cell growth, neoplastic cells require large quantities of energy and macromolecules from an extracellular milieu; the extrinsic signals could then be transduced into cells and co-opt the numbers of core metabolic pathways, including glycolysis, mitochondrial metabolism, and lipid and amino acid anabolism/catabolism to support cell survival [15,16,17]. On a physiological level, oxygen availability is important for cancer cells to determine their metabolic identities, as cells in tumor tissues expose to various oxygen levels with respect to their distance from the closest blood vessels [18] (Figure 1A). While cancer metabolism is receiving increasing attention [19,20], most studies were conducted to target a single metabolic enzyme or metabolite in controlling tumorigenesis, without analyzing global metabolic alterations. In this way, to escape death, cancer cells could possibly evolve and develop alternative compensatory metabolic changes [21]. In light of this, systemic manipulations to direct the tumor cell metabolic status “back” to the normal cell status, thereby lessening the cancer malignancy, is desired (Figure 1B). To achieve this aim, the identification of reagent(s) that could reduce preferential metabolic effectors in tumors as well as trigger unfavorable carcinogenic metabolic cues could be molecules of interest for suppressing malignancy in cancers. The goal of the review is to provide a systemic overview regarding the current understanding of cancer metabolism and its clinical potential, with an emphasis on HNSCCs.

2. Identification for HNSCC-Specific Metabolic Profile

In early years, HNSCC-specific metabolic fingerprints were defined mainly by immunohistochemistry staining analysis and serological examination. For example, cellular retinoic acid binding protein (CRABP) expression was enriched in tumor tissues compared with its adjacent normal tissues [22], while additional studies confirmed that external retinoic acid administration could modulate the Epidermal Growth Factor Receptor (EGFR) activity, a key predisposition of HNSCC development [23]. Furthermore, a higher glutathione (GSH) concentration was detected in metastatic tumors, compared with those concentrations derived from the corresponding primary lesions, suggesting a possible impact of the GSH metabolism on the formation of metastases in HNSCCs [24,25]. Indeed, more recent studies determined prognostic roles of glutathione metabolic enzymes such as glutathione-S-transferases in controlling HNSCC oncogenicity [26]. Other investigations focusing on the association of the ornithine decarboxylase activity with cellular DNA distributions [27], the influence of the intracellular cAMP:cGMP ratio on protein kinase activity and cell growth [28] and the impact of decreased polyglutamylation for methotrexate resistance [29], were also described. On the other hand, the results showing that the serological metabolite orosomucoid:prealbumin ratio predicted a clinical prognosis [30] and total parenteral nutrition enhanced cell susceptibility to the cell cycle inhibitor [31] further confirmed that circulating metabolic molecules could affect the HNSCC malignancy. Interestingly, a study led by Byerley et al. showed that, using a computerized euglycemic clamp, a significantly lower plasma insulin concentration and elevated whole body glucose appearance were detected in cancer patients than in control subjects, indicating that tumors could act as a glucose drain, aiding against local cachexia [32].

3. Association of Systemic Disease and HNSCCs

From a broad point of view, the correlation between systemic diseases and HNSCC cancer is an interesting topic worthy of further investigations. Taking Diabetes mellitus (DM) as an example, although for more than fifty years clinicians have reported that DM and cancer are diagnosed within the same individual more frequently, both diseases are complex with multiple subtypes, making it difficult to conclude the association between them [33]. Previous epidemiological reports aiming to clarify the association between DM and oral pre-cancerous malignancy or cancer are sparse and highly debated as the conclusions are rather inconsistent. The relationship between DM and pre-cancerous lesions was more emphasized. A very early review examining 1600 DM patients for the occurrence of oral leukoplakia (OL) and oral lichen planus (OLP) found that the prevalence of OL and OLP in DM patients was higher compared to healthy controls [34]. It was also reported that DM is a significant and independent predictor of OL [35], while the results from the population-based Study of Health in Pomerania (SHIP) showed that subjects with OL exhibited higher levels of DM-related metabolites, suggesting that DM is associated with the risk of OL [36]. A recent cohort study with a samples size of more than 4.5 million US veterans reported a significantly low risk of Oral Squamous Cell Carcinomas (OSCCs) in DM patients [37]. However, this study recruited only male patients, and the age of the groups are rather old (> 50 years old), limiting further interpretation of the results. Local studies from Taiwanese hospitals brought about different conclusions regarding the role of DM in correlating with oral carcinogenesis or prognosis. It was reported that DM patients tend to have a lower overall survival, recurrence-free survival and cancer-specific survival compared to non DM subjects, even in less aggressive tumor stages (stages I and II) [38]. Nevertheless, a recent epidemiological study showed that the DM status and duration were strongly associated with an OSCC prevalence in unadjusted models but that none was significant after a multivariable adjustment [39]. At the molecular level, using an animal model, it was demonstrated that insulin receptor substrat-1 (IRS-1) and focal adhesion kinase might be potential molecules leading to an increased DM-mediated risk of oral cancer development [40]. Our group recently demonstrated that prolonged diabetic/hyperglycemic incubation is significantly correlated with phenotypic aggressiveness in HNSCC cells in vitro, in vivo and in clinic, further implying that DM is likely a positive regulator for HNSCC/OSCC development [41]. In brief, external metabolic challenges could modulate HNSCC oncogenicity; however, it is still less feasible to define the cause-and-effect relationship between systemic diseases such as DM and HNSCC/OSCC development, particularly in a clinical setting, since both diseases are age-related.

4. Development of Anti-Cancer Scheme by Targeting Distinct Metabolic Pathways in HNSCCs

As discussed in recent studies, by using NMR spectroscopy, HNSCC-specific metabolic features correlated with an acquired EGFR-TKI resistance [42]; the intracellular metabolism therefore becomes one of the potential targets to develop non-canonical anti-HNSCC drugs to better improve therapeutic efficiency. Below, we therefore focus in detail on studies seeking the development of metabolism-mediated anti-HNSCC scheme(s)/molecule(s) by targeting different pathways.

4.1. Targeting Aerobic Glycolysis in HNSCCs

Hyperactive glycolytic metabolism is the most promising metabolic change in highly proliferating cells [43]. Despite a lesser amount of energy production, many rapidly growing cells rely primarily on glucose fermentation during proliferation, regardless of oxygen availability, a phenomenon known as aerobic glycolysis or the “Warburg effect” [44]. Aerobic glycolysis is a metabolic hallmark uniquely observed in many cancers and had therefore been considered a well-studied target for developing cancer sensors or inhibitors [45]. For example, 2-deoxy-D-glucose (2-DG), a glucose analogue, has been used as medical detector in Positron Emission Tomography (PET) scanning to detect tumors in clinic [46,47], including HNSCCs [48]. Other studies also found that the combinational treatment of 2-DG with cisplatin (CDDP) and metformin in HNSCC cells exhibited an improved therapeutic efficacy compared with a single treatment, through the enhancement of oxidative stress, both in vitro and in vivo [49,50,51]. Moreover, the suppression of the Glucose transporter SLC2A (so-called GLUT-1) activity by the selective inhibitors Fasentin and WZB-117 decreased the HNSCC malignancy [52,53]. Other anti-cancer glycolytic enzyme inhibitors in HNSCC cells were widely studied. For example, the HK II inhibitor Fenofibrate, acting through the suppression of the binding of HK II and the mitochondrial transporter VDAC, modulates the AKT/mTOR signaling pathway and was found to repress tongue tumor development [54]. The PFKFB3 inhibitor PFK15 could induce cancer cell apoptosis, via the reduction of glucose uptake and lactate production [55]. The LDHA inhibitor Oxamate, acting via direct targeting of LDHA activity by suppressing a conversion of pyruvate to lactate, was reported to suppress cell growth in cetuximab (EGFR inhibitor) resistance HNSCC cells in vitro [56] or to inhibit tumor growth in vivo [57]. In agreement with previous findings, our group also recently defined LDHA as an oncogenic cue in regulating HNSCC malignancy in vitro, in vivo and in clinic [21]. At the molecular level, another interesting finding showed that the inhibition of the membrane bidirectional transporter MCT1, which converts monocarboxylates such as lactate into pyruvate for more energy supply (in HNSCC derived xerographic tumors, via AZD3965 (in combination with a simvastatin treatment)), could delay tumor growth [58,59]. Taken together, this confirms that targeting glycolytic molecules could efficiently suppress tumor growth — even though to date no glycolytic inhibitors are used as anti-HNSCC agents in clinic. One possibility for this slow progress might be due to the fact that it is still difficult to clearly define the differential dependency of glycolysis in HNSCC tissues compared with non-cancerous tissues in vivo. On a cellular level, how HNSCC cells stimulate alternative compensatory non-glycolytic metabolic changes to evade cell death in response to the treatment of glycolytic inhibitors remains to be determined.

4.2. Targeting Mitochondrial Related Metabolism in HNSCCs

The progression of biogenetic changes during cancer development, based on Warburg’s theory, is: insufficient respiration acts early during malignant transformation, followed by a compensatory energy production through glycolysis to continuously ferment lactate in the presence of oxygen, leading to deficient oxidative phosphorylation (OxPhos) that is irreversible [14,44]. Indeed, a number of mitochondrial transfer experiments support this statement by demonstrating that various tumorigenic phenotypes could be suppressed when mitochondria from normal cells are transferred to a tumor cell, whereas the tumorigenicity is enhanced when tumor mitochondria are transferred into a normal cell [60,61]. Furthermore, in contrast to normal mitochondria, which contain numerous cristae, structures with respiratory proteins that play essential structural roles in facilitating energy production through OxPhos, mitochondria from tumor tissues showed swelling cristae with partial or total cristolysis, implying that a respiratory defect is indeed one of the key hallmarks of cancer cells [62]. The numbers of metabolic pathways related to ATP production, including the TCA cycle and Electron Transport Chain (ETC) reactions, occur within mitochondria. According to Warburg’s observation, tumor cells avidly used aerobic glycolysis, as mitochondrial respiration exhibits an irreversible defect, suggesting that mitochondrial dysfunction might also play a key role in controlling malignant transformation. Thus, a further emphasis on how mitochondrial metabolic cues, including the ROS-mediated apoptotic machinery, TCA cycle and ETC, regulate the HNSCC malignancy and are, furthermore, targets for the development of an anti-HNSCC therapeutic method is discussed [63,64,65].

4.2.1. Reactive Oxygen Species (ROS) and Apoptotic Pathway

ROS, a by-product of ETC, are molecules that could capture a singlet oxygen atom in mitochondria under a normal physiological condition. Under a certain pathological status, an accumulating ROS generation results in an imbalanced respiration chain, thereby leading to oxidative stress, a well-known toxic physiological stimulus for various diseases including cancer formation [66,67]. Providing the importance of ROS in tumorigenesis, a number of studies have been reported to target the ROS level with some natural compound like Flavonoid or Triterpene, which was extracted from a plant or fruit, aiming to establish alternative approaches to lessen the HNSCC malignancy. In general, these agents could induce a decrease of the mitochondrial activity, thereby reducing the HNSCC cell viability — although different mechanisms were proposed. For example, Hederagenin inhibits the NRF2/ARE mediated antioxidant pathway, thereby increasing the ROS production in OSCC cells [68]. CHW09 and CYT-Rx20, two novel chemical compounds, also trigger the ROS levels and mitochondrial superoxide (MitoSox), followed by Glutathione suppression [69,70]. Carbolitin is a less toxic analog of CDDP, which when combined with Thioridazine has a dramatic decrease on cell proliferation due to the upregulation of proteasome subunit alpha 5 (PSMA5) via ROS production [71]. Interestingly, the treatment of Bufalin triggers cell growth inhibition via an ROS-independent mechanism based on the detection of an increased nitric oxide level and calcium production, as well as decreased ROS levels and Mitochondrial Membrane Potential (MMP) in HNSCCs [72]. In addition to the ROS-mediated cell death pathway, other studies also suggested that mitochondrial dysfunction could also modulate HNSCC growth. For example, Fisetin could induce cytochrome C, an apoptosis-inducing factor, as well as Endonuclease G, released from mitochondria into the cytosol and could trigger an apoptotic pathway [73]. Casticin treatment led to cell death via a G2/M cell cycle arrest [74]. Moreover, HL156A, a derivative of metformin, could inhibit the oncogenic Akt-mTOR-ERK1/2 pathway, thereby downregulating cancerous phenotypes in OSCCs [75]. In brief, chemical compounds that could facilitate mitochondrial-mediated cell death pathways could be ideal targets for the development of anti-HNSCC drugs, although the effects on normal cells need to be evaluated before clinic use.

4.2.2. Tricarboxylic Acid Cycle (TCA Cycle) & Electron Transport Chain (ETC)

It is widely known that the TCA cycle acts as an essential metabolic pathway for the cell energy supply. The numbers of compounds were therefore developed to modulate the TCA cycle aiming to regulate cellular bioenergetics-mediated phenotypes in cancers. For HNSCCs, it was reported that Lanthanide could inhibit HNSCC growth in glucose free conditions by suppressing cytosolic Malic enzyme 1 (ME1) that converts malate into pyruvate, leading to a decreased pyruvate level and consequent lactate production, accompanied by increased mitochondrial areas. This study suggested a metabolic shift away from aerobic glycolysis toward oxidative phosphorylation in response to lanthanide treatment [76]. A more recent study further demonstrated that ME2 depletion potentiated ionizing radiation therapy induced cellular senescence driven by ROS and linked to better clinical outcomes, revealing a potential therapeutic benefit of targeting ME2 [77]. Moreover, rotenone, a specific inhibitor for ETC complex I, has been shown to induce apoptosis in HNSCC via the activation of the caspase 8/9 mediated pathway [78]. It was recently found that a common anti-diabetic drug, metformin, could reduce the mitochondrial respiration activity and increase the sensitivity to external-beam radiation (XRT) in HNSCC carrying the wild-type TP53 protein, under a combinational treatment of 2-DG [49,79]. Another interesting finding was to discover the unexpected mitochondrial regulatory impact of “old” drugs, showing that atovaquone, an anti-malarial drug, inhibited complex III activity and caused an OCR reduction, leading to a reduced tumor growth in HNSCCs [80].
At a genetic basis, mitochondria possess their own transcriptional codes. Mitochondrial DNA (mtDNA) is located in the mitochondrial matrix and encodes 37 subunits of the tETC complex involved in the OxPhos system. Based on previous findings, HNSCC cells utilize aerobic glycolysis preferentially over the mitochondrial metabolism, suggesting that a decreased OxPhos activity is a potential predisposition for HNSCC development [21,81]. Indeed, it was found that a significantly lower mtDNA copy number was detected in the peripheral blood leukocytes of oral premalignant lesion patients when compared to those from a healthy group [82]. In HNSCC patients, lower COX II mRNA as well as D-loop abundance was detected in cancerous tissues when compared with their normal counterparts [83]. Consistent with these results, by using the TCGA database, less mtDNA content is found in HNSCC tumors in contrast to adjacent normal tissues [84]. Nevertheless, another study found that COX I mRNA expression was upregulated during HNSCC development from a benign lesion to a premalignant one, and ultimately to the tumorous stage [85], implying that the regulatory role of the mtDNA number is still unclear in HNSCCs. In brief, a number of questions remain to be elucidated; they include: (1) What are the upstream signaling pathway(s) for controlling the mtDNA expression? (2) Could a nuclear DNA-encoded subunit compensate for an mtDNA deficiency? (3) What are the associations between epidemiological risk factors including racism, smoking, alcohol consumption, areca nut chewing and HPV infection, and the mtDNA content in HNSCC patients? and (4) Is mtDNA modulation or OxPhos activity playing a more important role in regulating HNSCC oncogenicity?

4.3. Targeting Lipid Metabolism in HNSCCs

Cancer cell proliferation also requires the duplication of cellular macromolecular components during each cell division, making lipid metabolism a principal physiological pathway for maintaining cell architecture, growth and physiological homeostasis [86,87,88,89]. The link between glycolysis and the lipid metabolism could be evident by the fact that a portion of acetyl-CoA could be carboxylated into malonyl-CoA by acetyl-CoA carboxylase (ACACA), the primary rate-limiting enzyme of the de novo fatty acid biosynthesis pathway, followed by the condensation of acetyl-CoA and malonyl-CoA by fatty acid synthase (FASN) to produce saturated fatty acids (FAs). Saturated long-chain FAs can be further modified by elongases or desaturases to form more complex FAs, which are used for the synthesis of various cellular lipids such as phospholipids, triglycerides and cholesterol esters [90,91]. While FA building blocks come from either exogenous sources or from de novo biosynthesis, most normal cells prefer an external uptake, as tumors synthesize FA mainly through intrinsic de novo lipogenesis [92,93]. Based on previous studies, aberrant lipid metabolism is now recognized as one of the key features of cancer cells because cell proliferation requires an increased lipid biosynthesis in order to produce bioactive molecules that act as signal molecules regulating cancer progression [94]. Indeed, earlier studies showed that elevated activities of citrate synthase (CS), FASN and ATP citrate lyase (ACLY) were observed in different malignant cells [95,96]. The blockage of key lipid metabolic enzymes such as ACLY, FASN, ACACA, and stearoyl CoA desaturase (SCD), as well as the upstream regulator sterol regulatory element-binding proteins (SREBPs) in various neoplastic cells could suppress tumor cell malignancy both in vitro and in vivo [92,97,98,99]. We and other groups also found that either the imbalanced circulating cholesterol or upregulated expression of FASN, ACACA and SREBPs correlated with a poorer diagnosis in different cancers, implying a significance of the lipid metabolism during the cancer development in clinic [100,101,102].
As for HNSCC, the significance of the lipid metabolism in controlling the pathological transformation, in comparison with glycolysis- and mitochondria-related cues, is much less emphasized. It was firstly studied from the findings that cholesterol-lowering drugs, including lovastatin and simvastatin, exhibited a great anti-HNSCC effect by facilitating cell apoptosis and attenuating 1 integrin mediated cell migration, implying that circulating lipids might play a role in regulating HNSCC malignancy [103,104]. Further investigations revealed that elevated lipid peroxidation products such as lipid-derived ROS and Reactive Nitrogen Species (RNS) could be crucial in the HNSCC development [105,106,107], given that the antioxidant compound apigenin suppressed CDDP/5-fluorouracil (5-Fu) induced cytotoxicity via the Tumor Necrosis Factor (TNF)/Bcl-2 pathway [108]. The link between the lipid metabolism associated molecules and neoplastic phenotypic changes in HNSCC was initially evident from a finding that caspase-triggered SREBP activation is a key molecular mechanism for HNSCC cell death under CDDP treatment [109]. A number of studies were later conducted to delineate the roles of different lipogenic proteins, including ACCs [110,111,112], FASN [113] and sphinogosine-1-phosphate metabolic enzymes [114], as well as a lipid catabolic factor lipolysis-stimulated lipoprotein receptor (LSR) [115] in controlling the HNSCC malignancy. In general, the molecules involved in lipogenesis are upregulated during HNSCC development and serve as an oncogenic indicator in clinic [110,114,116]; in contrast, lipid catabolic cues display a tumor-suppressive effect in HNSCCs.

4.4. Targeting Amino Acid Metabolism in HNSCCs

Like glucose, there are major differences in the utilization and metabolism of amino acids in tumors when compared to normal cells. All twenty amino acids could serve differently as important regulatory cues to many of the physiological processes, including the biosynthesis of macromolecules, cellular redox homeostasis, post-translational and epigenetic modifications and anaplerostic reactions [117,118,119]. Previous studies have found that several amino acids, including phenylalanine, valine, threonine, tyrosine, glycine, proline, histidine, aspartic acid, glutamic acid, asparagine, lysine, serine, methionine, and alanine, were significantly increased in tumor tissues compared with adjacent normal tissues from HNSCC patients [120,121]. By using L-[3-18F]-α–methyltyrosine (18F-FAMT) as a PET imaging tracer, it was also discovered that primary tumor cells, as well as cervical lymph node tissues, exhibit a greater amino acid uptake rate in HNSCC patients, suggesting that the amino acid metabolism might serve a critical role in HNSCC oncogenicity [4].
In addition to intrinsic amino acids, cells could obtain amino acids externally via different transporters, including the L-type amino acid transporter 1 (LAT1), which is specific for the branched-chain and aromatic amino acid intake; the system ASC amino-acid transporter-2 (ASCT2), which is responsible for the neutral amino acid uptake; and SLC7A11, also named xCT, an bidirectional exchanger for cysteine-glutamate. Previous studies have found that these transporters are enriched in many cancers and that their expression is correlated with a worse survival rate [117,122,123,124,125,126,127,128]. In HNSCC tissues, a greater ALT expression positively correlated with a higher percentage of Ki67+ cells [129]. Further studies confirmed the role of LAT in controlling HNSCC tumorigenicity via the detection of the anti-proliferative ability in HNSCC cells, both in vitro and in vivo, in response to a combinational treatment of the LAT-specific inhibitor JPH203 and metformin [130]. The synergetic anti-cancer effect of LAT inhibition was also reported in a study using gefitinib, a clinically targeted therapeutic EGFR inhibitor, to enhance cytotoxicity under stress (e.g., amino acid starvation culture condition) in HNSCC cells [131]. Several studies have also documented that the other two amino acid transporters ASCT2 and xCT mediate chemotherapy and EGFR-targeted therapy resistance in HNSCC cells after treatments of CDDP, cetuximab and AG1478 [132,133,134], implying that amino acid transporters could be a therapeutic target in patients with HNSCCs. On a molecular basis, a recent study demonstrated that xCT-mediated therapeutic resistance occurs mainly through the regulation of ferroptosis [135].
Including different amino acids in the discussion, individual amino acids play differential roles in regulating tumor cell malignancy. For example, methionine-derivative S-adenosyl methionine (SAM) is a cofactor acting mainly as a methyl donor to cytosine in DNA during epigenetic modification [136]. A recent study found a reduced oral mucositis after an oral D-methionine treatment in patients with grade 3/4 HNSCCs [137]; an in vitro study showed that the administration of methionine could also downregulate the proliferative capacity in HNSCC cells; it was also reported that there was a positive correlation of high methionine serum levels with greater overall and relapse-free survival rates [121,138]. These findings suggest that methionine could be tumor-suppressive and that the methionine content might serve as a prognostic factor for HNSCC patients, although one must still mechanistically clarify the way in which methionine regulates epigenetic modification in HNSCC development. Another important amino acid for cell homeostasis is arginine. Arginine is a conditional essential amino acid that could either be obtained from a diet or from a biosynthesis pathway by converting glutamine into citrulline. Several studies showed that cell proliferation is abolished when HNSCC cells were cultured in an arginine-free medium or a medium containing an arginine depleting agent, Arginine deiminase (ADI) [139,140]. In addition, a disease-free survival and late-stage overall survival rates seem significantly correlated with the expression of argininosuccinate synthetase (ASS), a rate-limiting enzyme of the urea cycle that involves an arginine de novo synthesis by converting citrulline into argininosuccinate and arginine [139]. These findings revealed the oncogenic impact of the arginine-related metabolism and provide a possibility for targeting the arginine metabolic pathway for HNSCC therapy. A very recent investigation found that cysteine deficiency led to a decrease of the glutathione level, resulting in lipid ROS-mediated cell death in CDDP resistant HNSCC cells compared with their parental counterparts [135].
Glutaminolysis, a metabolic pathway for converting glutamine into glutamate via glutaminase, is widely documented as playing an essential role in regulating neoplastic phenotypes. Immunohistochemistry staining and Western blot analysis both showed a higher glutaminase expression in metastatic HNSCC tissues than in normal tissues. Furthermore, shRNA-mediated silencing for glutaminase abrogated the HNSCC proliferation and downregulated the cellular glutamate level. At the cellular level, a positive correlation between glutaminase and aldehyde dehydrogenase (ALDH) was reported, while the administration of glutamine could enhance the ALDH expression and restored the tumor growth in ALDHloCD44lo cells, suggesting that glutaminolysis-related manipulations could be a potential method for the development of a HNSCC stem cell based therapeutic regimen [141]. Another glutamine-related enzyme, glutaminase 1 (GLS1), is also highly expressed in tumorous tissues, and the GLS1 expression is correlated with a poorer survival rate in HNSCC patients. Under the treatment of the GLS1 selective inhibitor bis–2–(5–phenylacetamido–1,3,4–thiadiazol–2–yl)ethyl sulfide (BPTES), it was shown that a reduced glutamine consumption and HNSCC growth was detected, highlighting the significance of glutaminolysis in controlling the HNSCC malignancy [142].
Taken together, amino acid metabolic cues are important for HNSCCs, although much effort should be put in better identifying the significance of different amino acids in controlling the HNSCC physiology under the same experimental settings. In addition, based on the fact that different amino acid metabolisms crosstalk to each other, it becomes essential to apply isotope labeling methods to trace amino acid related metabolites, and their associations with phenotypic alterations and metabolic shifts. In this way, the metabolic plasticity of HNSCC cells could be more clearly illustrated, and key amino acid metabolic pathways could be uncovered.

5. Molecular Basis of Metabolic Regulations in HNSCCs

5.1. Non-Coding RNA (ncRNA) Mediated Control for HNSCC Metabolic Cues

Although the majority of cancer research focuses on investigating the roles of the protein-coding genes in controlling oncogenicity, the exon-coding strands only consist of 3% of the whole human genome [143]. The regulations of non-coding RNAs (ncRNAs) in physical and pathological conditions have remained largely unknown until recent years. Aberrant ncRNAs expression was found in a number of pathological conditions, such as Alzheimer’s disease, Down’s syndrome, Parkinson’s disease, neurodevelopmental disease, diabetes, leukemia, and various type of solid tumors [144]. As for tumourigenesis, ncRNAs regulate pathways for cancer initiation and progression in a tissue-specific manner [145]. Long non-coding RNA (LncRNA) and microRNA (miRNA), based on their length of nucleotides, were the most studied ncRNAs. LncRNAs are defined as ncRNAs that are more than 200 nucleotides long and associated with a range of biological conditions, including cell metabolism and immune response, as well as cancer development and progression [146]. In contrast, miRNAs are 19–22 nucleotides in length and involved in the regulation of gene expression by targeting mRNAs in order to cause a transcriptional repression or mRNA degradation [147,148].
Most ncRNA-mediated metabolic controls for tumorigenesis rely on regulatory cues derived by the surrounding environment. Cancer progression is highly dependent on the interactions between the tumor cell and the tumor microenvironment (TME); TME is comprised of various cell types, such as endothelial cells, cancer-associated fibroblasts (CAF), and immune cells, which may cross-talk with cancer cells via cytokines, growth factors, proteases or hormones [149]. It therefore may not be surprising that metabolic modulations in the tumor niche have considerable impacts in regulating oncogenicity. For example, immune checkpoint blockade therapies against CTLA-4, Program Death-1 (PD-1), and PD-L1 could restore glucose in the tumor microenvironment, resulting in an enhanced glycolytic activity in tumor infiltrating T lymphocytes (TILs) and IFN–production [150]. It was found that LncUCA1 promotes aerobic glycolysis by upregulating hexokinase 2 (HK2) and the mTOR-STAT3 pathway in bladder cancer [151]. LncHULC is highly expressed in hepatocellular carcinoma (HCC) patient tissues and positively correlated with the lipogenic gene ACSL1, which in turn controls intracellular triglycerides and cholesterol levels by increasing the miR-9-PPARA-ACSL1 axis [152]. Moreover, increased LncRNA LINC01234 promotes colon cancer cell growth and is associated with poorer survival via a regulation of the serine/glycine metabolism by serving as a competing endogenous RNA (ceRNA) for miR-642a [153]. The potential regulatory roles of ncRNAs in cancers such as lung, liver, gastric, colorectal, ovarian and prostate cancer have already been described elsewhere [146,154].
In HNSCC, the upregulation of LncCAF (FLJ22447) in cancer-associated fibroblasts (CAFs) upregulates cytokine IL-33, supports tumor growth and leads to a poor prognosis [155]. Using the TCGA database, it was found that LncZFAS1 serves as a stimulator in HNSCC and can be utilized as a potential tumor marker in clinic. On the molecular level, LncZFAS1 is associated with transcripts related to cell adhesion, cell differentiation, cell death, angiogenesis, oxidative stress response, as well as endothelial signal regulation. More recently, high levels of EGFR and PD-L1 were found in LncZFAS1-low patients, suggesting that a low expression of LncZFAS1 might be more responsive to anti-EGFR and anti-PDL1 therapies [156]. Furthermore, the upregulation of LncHIFCAR (long noncoding HIF-1αco-activating RNA, MIR31HG) in OSCC patients can also be a prognostic indicator via the direct interaction with Hypoxia inducible factor-1α (HIF-1α under hypoxia condition [157]). Interestingly, LncHIFCAR could induce pseudohypoxia by interacting with HIF-1 and a modulated hypoxia-induced glucose uptake and lactate production. In addition to LncZFAS1 and LncHIFCAR, hypoxia could also enhance the expression of another non-coding RNA, HAS2-AS1, in an HIF-1α dependent manner, and the increase of HAS2-AS1 contributes to hypoxia-regulated EMT and invasiveness in OSCC cells [158].
In addition to LncRNA, several studies supported the roles of miRNAs in orchestrating cross-talks between tumor cells and tumor stroma [159]. In HNSCC, miR-34a plays an inhibitory role by shaping a microenvironment favorable for tumor growth; a recent study demonstrated that miR-34a differentially modulate endothelial cell growth, migration and tube formation through a vascular endothelial growth factor (VEGF) mediated machinery in the microenvironments of pre-cancerous lesions and cancers [160]. Interestingly, CAF and CAF-derived exosomes contain lower miR-34a levels than normal fibroblasts do, as a lower miR-34a expression enhances the AKT, GSK3β, β-catenin, and snail signaling activity by targeting AXL [161]. Whilst there was still no evidence supporting the role of miR-34a in controlling the immune system in HNSCCs, nevertheless, in acute myeloid leukemia (AML), miR-34a could target PD-L1 to lessen PD-L1-mediated T cell apoptosis [162]. In contrast, miR-34a was reported as a potential invasive biomarker for the diagnosis of progressive pancreatic ductal adenocarcinoma (PDAC) [163], implying that miR-34-related regulatory mechanisms could vary in different cancers. Under hypoxia conditions, miR-210 can be induced in many cancer cells (breast, pancreatic, head and neck, lung, colon, and renal carcinoma) [164,165,166,167]. The HIF-1α-driven overexpression of glycolytic enzymes and miR-210 is coupled to the downregulation of its target iron-sulfur cluster assembly enzyme (ISCU) in oropharyngeal squamous cell carcinoma (OPSCC) [168]. Moreover, decreased miR-340 levels in OSCC cells led to an increased cell growth via targeting GLUT1 expression, which consequently resulted in an enhanced lactate secretion and glucose uptake [169]. As ncRNAs could potentially target multiple downstream molecular cues, a cooperative regulation between LncRNAs and miRNAs was also reported. It was shown that LncPCAT19 promotes the proliferation and poor prognosis of larynx carcinoma, and that the LncPCAT19-mediated regulation in laryngeal cancer cells occurred through the miR-182-mediated facilitation of the mitochondrial PDK4 expression [156].
Aside from the LncRNAs and miRNAs detected in the tumor or its microenvironment, the analysis for serological and salivary LncRNAs and miRNAs as potential diagnostic tools has recently been carried out [163,170,171,172]. It was found that LncRNAs HOXA11-AS, LINC00964 and MALAT-1 were upregulated in the plasma of HNSCC patients when compared to healthy controls, while the levels of miR-21 [173] and miR-146a [174] were significantly increased in plasma samples obtained from HNSCC patients. On the other hand, the numbers of LncRNAs, including MALAT-1, HOTAIR, NEAT-1, HULC, MEG-3 and UCA1, were abundant in both HNSCC tumor tissues and patients’ saliva [175] as two salivary miRNAs, miR-125a and miR-200a, were significantly decreased in OSCC patients compared to healthy subjects [176]. Interestingly, miR-21 is not only a plasma biomarker for HNSCC but also a salivary marker for the detection of early esophageal cancer [177,178]. Based on our findings, one of the oncogenic miRNAs, miR-31, was increased in both the plasma and saliva of the OSCC patients at all clinical stages when compared with normal subjects; the miR-31 level is abundantly detected in the saliva when compared to the plasma, suggesting that salivary miR-31 is a more sensitive marker for oral malignant transformation [179,180]. Additionally, in a 4NQO-induced oral cancer mouse model, the plasma and salivary miR-21, 31, 146a, and 184 expressions were all increased in progressive tongue cancers [181]. Taken together, these findings demonstrated that the LncRNA and/or miRNA levels in saliva and plasma could be used as an early, noninvasive, and rapid tool for the diagnosis of oral cancer. In order to gain a more comprehensive view of metabolic networks regulated in HNSCC development, using different databases such as TCGA or GEO datasets to define potential therapeutic ncRNAs could be an alternative for exploring new HNSCC-associated ncRNAs [182,183]. The significance of ncRNAs in controlling the HNSCC malignancy is summarized in Table 1.

5.2. Crosstalk of Tumors and Microenvironment in HNSCCs

Tumor cells rely on external nutrients for proliferation and survival, and it is therefore not surprising that the cancer milieu plays a crucial role in controlling tumourigenesis [184,185,186,187,188]. By using an immunohistochemistry analysis, an early focus in investigating the role of TME during cancer development was initiated by defining stromal-enriched proteins. For example, a number of studies reported that VEGF [189], Receptor for Advanced Glycation End products (RAGE) [190], Platelet-derived Growth Factor Receptor PDGFR) [191], Hepatocyte growth factor/c-met (HGF/HGFR) signaling [192,193,194], Tumor Growth Factor 1(TGF-1)-Throbospondin-1(THBS-1) axis [195], Notch signaling [196] and Wnt signaling [197] molecules serve as tumor stroma specific factors and have a great impact in regulating the HNSCC malignancy. Regarding the methodology, in order to better define the importance of stromal cells in modulating the HNSCC cell malignancy, a tumor-stroma co-culture system was developed [198]. Much effort was also spent to elucidate the significance of different cell populations detected in tumor stroma — mainly CAFs and tumor-associated immune cells, including tumor-associated macrophages (TAMs) and tumor-infiltrating lymphocytes (TILs) — in modulating HNSCC oncogenicity (Figure 2). The importance of another important environmental cue, hypoxia, in influencing the tumorous identification of HNSCC cells was also a research field of great interest. Last, very recent studies showed that tumor-secreted lactate and monocarboxylate transporters (MCTs) detected in tumor and stromal cells also play important roles in modulating HNSCC progression [199,200,201,202], suggesting that a better understanding of the metabolic competition between tumor cells and surrounding immune cells is essential for development of anti-HNSCC therapy [203].

5.2.1. CAFs in HNSCCs

The first observation for determining the importance of CAFs in HNSCC was described by Rosenthal et al., showing that elevated TGF 1 was detected in the stromal counterpart of HNSCCs compared to normal mucosa [204]; the stroma-derived TGF 1 is important to the migration of Disseminated Tumor Cells (DTCs) [205]. Moreover, a co-culture of HNSCC/OSCC cells with CAFs could trigger an HNSCC cell proliferation, invasion and metastatic activity [206], as well as metabolic alteration [81]. The large-scale screening of secretome by CAFs was also carried out to more systemically identify stroma-specific molecules, and the results found that the epithelial-mesenchymal-transition (EMT) mediator S100 calcium-binding protein A4 (S100A4) serves as a potential therapeutic target for the CAFs-based treatment in HNSCCs [207]. Interestingly, a very recent study demonstrated that the crosstalk between OSCC cells and CAFs via the tumor microvesicles mediated “Reverse Warburg effect” could promote tumor cell motility and facilitate glycometabolic activity [208].

5.2.2. Immune Cells in HNSCCs

In theory, the immune system could protect cells from attacks by microorganisms and pathological cells including tumor cells, in order to maintain homeostasis in the body. Although TILs could be found in clinic HNSCC tumor tissues, the elevated expression of the death-associated domains Fas and Fas ligand (FasL) was also detected [209], indicating that immune cells in TME might be suppressive and functionally exhausted. Different immune cell populations were defined in HNSCC TME, and they execute immune suppression via various molecules, mainly cytokines, to promote tumor growth. For instance, CD4+/CD25high/Foxp3+ Treg cells secreted IL-10/TGFβ1 to inhibit TME [210]; Neutropolin-1-/- Treg cells produced Interferon-γ triggering Neutropolin-1+/+ Treg cell fragility to modulate TME immune activity [211]; HNSCC cells target Th17+ cells for an increased Cox2 expression, resulting in greater tumor cell growth [212]; TAMs enhanced HNSCC invasiveness and lymph node metastasis via IL-33 and the CCR4/CCL22 axis, respectively [213,214]; and immunosuppressive adenosine (ADO) is reduced in HNSCC patients compared with healthy control subjects [215]. With great attention paid to immune checkpoint molecules such as PD-1 and its ligand PDL1 in tumor immunology [216], several studies were conducted to test whether a blockade of immunosuppressive TME could rescue anti-tumor activity in HNSCC patients. The blockade of PD-1/SHP2 could restore a robust Th1 immunity and thereby reverse immunosuppression in TME [217]. Stromal IL-33 is important for the suppression of immune activity and is correlated with a poorer prognosis in HNSCC patients [218]. It is worth noting that scores of different immune cell subpopulations (dendritic cells (DCs), CD4+/CD8+ T cells, IL-12Rβ2+ TILs [219,220,221] and CD47+ macrophages [222]) predict differential clinical outcomes in various studies, even though a very recent meta-analysis showed that a high PDL1 expression did not correlate with a poorer prognosis for OSCC patients [223]. The discrepant findings indicated that a more complex molecular regulation for modulating the immune activity remains to be explored.

5.2.3. Hypoxia/Vascular Signals in HNSCCs

Hypoxia stimulates a complex oncogenic alteration in tumors [224]. Oxygen deprivation in cancerous tissues often leads to an advanced but dysfunctional vascularization to support tumor growth and invasion [225]. In HNSCCs, it was long found that a very low oxygen pressure (< 10 mmHg) was detected, compared to normal tissues (~43 mmHg) [226]. At the molecular level, the HIF-1α/MIF and NF-κB/IL-6 axes contributed to the recruitment of CD11b+Gr-1+ immunosuppressive myeloid cells and promoted angiogenesis and tumor invasion [227]. Hypoxic HNSCC tumors could secret miR-21 containing exosomes into normoxic cells to enhance metastatic activity [228]. Under hypoxic conditions, vascular distant HNSCC cells displayed a lower expression of the Epidermal Growth Factor Receptor (EGFRs), which may underlie anti-EGFR therapy resistance [229]. A very recent study demonstrated that primary hypoxic TME gives rise to a population of dormant DTCs that evade therapy and may contribute to disease relapse and poor prognosis [230]. On a molecular level, it was shown that the oncogenic p38-MAPK signaling pathway could facilitate angiogenesis and lymphangiogenesis in HNSCCs [231].

6. Conclusions and Future Perspectives

Cellular metabolism is highly dynamic, and the regulation between environmental and intracellular metabolic cues is closely intertwined. In order to translate the current knowledge about cancer metabolism for the development of clinical therapeutic schemes, several points should be addressed. First, the effects of using combinational therapy of conventional chemo/radiotherapy plus metabolic modulators in treating HNSCCs were rarely examined. While working mechanisms vary for conventional therapies (mainly targeting cell growth) and metabolism-mediated anti-cancer agents, it would be interesting to determine whether combinational therapies could be superior to single treatments. Second, in order to manipulate tumorous metabolic cues in a broader manner, an ideal platform for screening potential reagents that suppress cancerous metabolic molecules and simultaneously upregulate less active metabolic factors in cancer cells would be required. Last, it is likely that HNSCC tumors and stromal cells exhibit reciprocal metabolic cues that could be important in helping cancer cells escape from immune attacks and, at the same time, in exhausting immune cells in TME. It therefore becomes essential to define metabolic molecules that display opposite expression pattern in tumors/CAFs and tumor-associated immune cells on a single-cell population level. In this way, the candidate compound(s) could be more efficiently suppressing the malignancy and stimulating immune cell activity (Figure 3).

Funding

This work is supported by grants from Ministry of Science and Technology, Taiwan (Most-103-2314-B-010-024-MY3 and 106-2314-B-010-005-MY3); Yen Tjing Ling Medical Foundation (CI-105-8 & CI-106-13) and a grant from Ministry of Education, Aiming for the Top University Plan as well as the Higher Education Sprout Project by the Ministry of Education (MOE) in Taiwan.

Acknowledgments

We thank Jude Clapper (Taipei American School) for critical review and English corrections for the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cohen, N.; Fedewa, S.; Chen, A.Y. Epidemiology and demographics of the head and neck cancer population. Oral Maxillofac Surg Clin. North. Am. 2018, 30, 381–395. [Google Scholar] [CrossRef] [PubMed]
  2. Leemans, C.R.; Braakhuis, B.J.; Brakenhoff, R.H. The molecular biology of head and neck cancer. Nat. Rev. Cancer 2011, 11, 9–22. [Google Scholar] [CrossRef] [PubMed]
  3. Meurman, J.H. Infectious and dietary risk factors of oral cancer. Oral Oncol. 2010, 46, 411–413. [Google Scholar] [CrossRef]
  4. Cognetti, D.M.; Weber, R.S.; Lai, S.Y. Head and neck cancer: An evolving treatment paradigm. Cancer 2008, 113, 1911–1932. [Google Scholar] [CrossRef] [PubMed]
  5. Du, E.; Mazul, A.L.; Farquhar, D.; Brennan, P.; Anantharaman, D.; Abedi-Ardekani, B.; Weissler, M.C.; Hayes, D.N.; Olshan, A.F.; Zevallos, J.P. Long-term survival in head and neck cancer: Impact of site, stage, smoking, and human papillomavirus status. Laryngoscope 2019. [Google Scholar] [CrossRef] [PubMed]
  6. Bray, F.; Ferlay, J.; Soerjomataram, I.; Siegel, R.L.; Torre, L.A.; Jemal, A. Global cancer statistics 2018: Globocan estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J. Clin. 2018, 68, 394–424. [Google Scholar] [CrossRef] [PubMed]
  7. Santuray, R.T.; Johnson, D.E.; Grandis, J.R. New therapies in head and neck cancer. Trends Cancer 2018, 4, 385–396. [Google Scholar] [CrossRef] [PubMed]
  8. Siegel, R.L.; Miller, K.D.; Jemal, A. Cancer statistics, 2019. CA Cancer J. Clin. 2019, 69, 7–34. [Google Scholar] [CrossRef] [PubMed]
  9. Leemans, C.R.; Snijders, P.J.F.; Brakenhoff, R.H. The molecular landscape of head and neck cancer. Nat. Rev. Cancer 2018, 18, 269–282. [Google Scholar] [CrossRef]
  10. Mascitti, M.; Rubini, C.; De Michele, F.; Balercia, P.; Girotto, R.; Troiano, G.; Lo Muzio, L.; Santarelli, A. American joint committee on cancer staging system 7th edition versus 8th edition: Any improvement for patients with squamous cell carcinoma of the tongue? Oral Surg. Oral Med. Oral Pathol Oral Radiol. 2018, 126, 415–423. [Google Scholar] [CrossRef]
  11. Vogelstein, B.; Papadopoulos, N.; Velculescu, V.E.; Zhou, S.; Diaz, L.A., Jr.; Kinzler, K.W. Cancer genome landscapes. Science 2013, 339, 1546–1558. [Google Scholar] [CrossRef] [PubMed]
  12. Seyfried, T.N.; Flores, R.; Poff, A.M.; D’Agostino, D.P.; Mukherjee, P. Metabolic therapy: A new paradigm for managing malignant brain cancer. Cancer Lett. 2015, 356, 289–300. [Google Scholar] [CrossRef] [PubMed]
  13. Seyfried, T.N.; Flores, R.E.; Poff, A.M.; D’Agostino, D.P. Cancer as a metabolic disease: Implications for novel therapeutics. Carcinogenesis 2014, 35, 515–527. [Google Scholar] [CrossRef] [PubMed]
  14. Seyfried, T.N.; Shelton, L.M. Cancer as a metabolic disease. Nutr. Metab 2010, 7, 7. [Google Scholar] [CrossRef] [PubMed]
  15. Vander Heiden, M.G.; DeBerardinis, R.J. Understanding the intersections between metabolism and cancer biology. Cell 2017, 168, 657–669. [Google Scholar] [CrossRef] [PubMed]
  16. Lee, N.; Kim, D. Cancer metabolism: Fueling more than just growth. Mol. Cells 2016, 39, 847–854. [Google Scholar] [CrossRef] [PubMed]
  17. Li, Z.; Zhang, H. Reprogramming of glucose, fatty acid and amino acid metabolism for cancer progression. Cell Mol. Life Sci. 2016, 73, 377–392. [Google Scholar] [CrossRef] [PubMed]
  18. Racker, E. History of the pasteur effect and its pathobiology. Mol. Cell Biochem. 1974, 5, 17–23. [Google Scholar] [CrossRef]
  19. Pavlova, N.N.; Thompson, C.B. The emerging hallmarks of cancer metabolism. Cell Metab. 2016, 23, 27–47. [Google Scholar] [CrossRef]
  20. Hirschey, M.D.; DeBerardinis, R.J.; Diehl, A.M.E.; Drew, J.E.; Frezza, C.; Green, M.F.; Jones, L.W.; Ko, Y.H.; Le, A.; Lea, M.A.; et al. Dysregulated metabolism contributes to oncogenesis. Semin. Cancer Biol. 2015, 35, 129–150. [Google Scholar] [CrossRef]
  21. Chen, T.Y.; Hsieh, Y.T.; Huang, J.M.; Liu, C.J.; Chuang, L.T.; Huang, P.C.; Kuo, T.Y.; Chia, H.Y.; Chou, C.Y.; Chang, C.W.; et al. Determination of pyruvate metabolic fates modulates head and neck tumorigenesis. Neoplasia 2019, 21, 641–652. [Google Scholar] [CrossRef] [PubMed]
  22. Gates, R.E.; Rees, R.S. Altered vitamin a-binding proteins in carcinoma of the head and neck. Cancer 1985, 56, 2598–2604. [Google Scholar] [CrossRef]
  23. Kim, J.S.; Steck, P.A.; Gallick, G.E.; Lee, J.S.; Blick, M.; Hong, W.K.; Lotan, R. Suppression by retinoic acid of epidermal growth factor receptor autophosphorylation and glycosylation in cultured human head and neck squamous carcinoma cells. J. Natl. Cancer Inst. Monogr 1992, 101–110. [Google Scholar]
  24. Bier, H.; Bergler, W.; Mende, S.; Ganzer, U. Glutathione content and gamma-glutamyltranspeptidase activity in squamous cell head and neck cancer xenografts. Arch. Otorhinolaryngol 1988, 245, 166–169. [Google Scholar] [CrossRef] [PubMed]
  25. Janot, F.; Massaad, L.; Ribrag, V.; de Waziers, I.; Beaune, P.H.; Luboinski, B.; Parise, O., Jr.; Gouyette, A.; Chabot, G.G. Principal xenobiotic-metabolizing enzyme systems in human head and neck squamous cell carcinoma. Carcinogenesis 1993, 14, 1279–1283. [Google Scholar] [CrossRef] [PubMed]
  26. Tripathy, C.B.; Roy, N. Meta-analysis of glutathione s-transferase m1 genotype and risk toward head and neck cancer. Head Neck 2006, 28, 217–224. [Google Scholar] [CrossRef] [PubMed]
  27. Edstrom, S.; Westin, T.; Delle, U.; Lundholm, K. Cell cycle distribution and ornithine decarboxylase activity in head and neck cancer in response to enteral nutrition. Eur. J. Cancer Clin. Oncol. 1989, 25, 227–232. [Google Scholar] [CrossRef]
  28. Rydell, E.L.; Axelsson, K.L.; Olofsson, J.; Hellem, S. Protein kinase activities in neoplastic squamous epithelia and normal epithelia from the upper aero-digestive tract. Cancer Biochem. Biophys 1990, 11, 187–194. [Google Scholar]
  29. Pizzorno, G.; Chang, Y.M.; McGuire, J.J.; Bertino, J.R. Inherent resistance of human squamous carcinoma cell lines to methotrexate as a result of decreased polyglutamylation of this drug. Cancer Res. 1989, 49, 5275–5280. [Google Scholar]
  30. Caldani, C.; Thyss, A.; Schneider, M.; Milano, G.; Buray, L.; Demard, F. Orosomucoid:Prealbumin ratio--a marker of the host-tumor relationship in head and neck cancer. Eur. J. Cancer Clin. Oncol. 1988, 24, 653–657. [Google Scholar] [CrossRef]
  31. Frank, J.L.; Lawrence, W., Jr.; Banks, W.L., Jr.; McKinnon, J.G.; Chan, W.M.; Collins, J.M. Modulation of cell cycle kinetics in human cancer with total parenteral nutrition. Cancer 1992, 69, 1858–1864. [Google Scholar] [CrossRef]
  32. Byerley, L.O.; Heber, D.; Bergman, R.N.; Dubria, M.; Chi, J. Insulin action and metabolism in patients with head and neck cancer. Cancer 1991, 67, 2900–2906. [Google Scholar] [CrossRef]
  33. Joslin, E.P.; Lombard, H.L.; Burrows, R.E.; Manning, M.D. Diabetes and cancer. N Engl. J. Med. 1959, 260, 486–488. [Google Scholar] [CrossRef] [PubMed]
  34. Albrecht, M.; Banoczy, J.; Dinya, E.; Tamas, G., Jr. Occurrence of oral leukoplakia and lichen planus in diabetes mellitus. J. Oral Pathol. Med. 1992, 21, 364–366. [Google Scholar] [CrossRef] [PubMed]
  35. Dietrich, T.; Reichart, P.A.; Scheifele, C. Clinical risk factors of oral leukoplakia in a representative sample of the us population. Oral Oncol. 2004, 40, 158–163. [Google Scholar] [CrossRef]
  36. Meisel, P.; Dau, M.; Sumnig, W.; Holtfreter, B.; Houshmand, M.; Nauck, M.; Kocher, T. Association between glycemia, serum lipoproteins, and the risk of oral leukoplakia: The population-based study of health in pomerania (ship). Diabetes Care 2010, 33, 1230–1232. [Google Scholar] [CrossRef] [PubMed]
  37. Atchison, E.A.; Gridley, G.; Carreon, J.D.; Leitzmann, M.F.; McGlynn, K.A. Risk of cancer in a large cohort of u.S. Veterans with diabetes. Int. J. Cancer 2011, 128, 635–643. [Google Scholar] [CrossRef]
  38. Wu, C.H.; Wu, T.Y.; Li, C.C.; Lui, M.T.; Chang, K.W.; Kao, S.Y. Impact of diabetes mellitus on the prognosis of patients with oral squamous cell carcinoma: A retrospective cohort study. Ann. Surg Oncol. 2010, 17, 2175–2183. [Google Scholar] [CrossRef]
  39. Tseng, C.H. Oral cancer in taiwan: Is diabetes a risk factor? Clin. Oral Investig. 2013, 17, 1357–1364. [Google Scholar] [CrossRef]
  40. Goutzanis, L.; Vairaktaris, E.; Yapijakis, C.; Kavantzas, N.; Nkenke, E.; Derka, S.; Vassiliou, S.; Acil, Y.; Kessler, P.; Stavrianeas, N.; et al. Diabetes may increase risk for oral cancer through the insulin receptor substrate-1 and focal adhesion kinase pathway. Oral Oncol. 2007, 43, 165–173. [Google Scholar] [CrossRef]
  41. Liu, C.J.; Chang, W.J.; Chen, C.Y.; Sun, F.J.; Cheng, H.W.; Chen, T.Y.; Lin, S.C.; Li, W.C. Dynamic cellular and molecular modulations of diabetes mediated head and neck carcinogenesis. Oncotarget 2015, 6, 29268–29284. [Google Scholar] [PubMed] [Green Version]
  42. Beloueche-Babari, M.; Box, C.; Arunan, V.; Parkes, H.G.; Valenti, M.; De Haven Brandon, A.; Jackson, L.E.; Eccles, S.A.; Leach, M.O. Acquired resistance to egfr tyrosine kinase inhibitors alters the metabolism of human head and neck squamous carcinoma cells and xenograft tumours. Br. J. Cancer 2015, 112, 1206–1214. [Google Scholar] [CrossRef] [PubMed]
  43. Liberti, M.V.; Locasale, J.W. The warburg effect: How does it benefit cancer cells? Trends Biochem. Sci. 2016, 41, 211–218. [Google Scholar] [CrossRef] [PubMed]
  44. Warburg, O. On the origin of cancer cells. Science 1956, 123, 309–314. [Google Scholar] [CrossRef] [PubMed]
  45. Deberardinis, R.J. Q&A: Targeting metabolism to diagnose and treat cancer. Cancer Metab. 2014, 2, 5. [Google Scholar] [PubMed]
  46. Miele, E.; Spinelli, G.P.; Tomao, F.; Zullo, A.; De Marinis, F.; Pasciuti, G.; Rossi, L.; Zoratto, F.; Tomao, S. Positron emission tomography (pet) radiotracers in oncology--utility of 18f-fluoro-deoxy-glucose (fdg)-pet in the management of patients with non-small-cell lung cancer (nsclc). J. Exp. Clin. Cancer Res. 2008, 27, 52. [Google Scholar] [CrossRef] [PubMed]
  47. Wick, A.N.; Drury, D.R.; Nakada, H.I.; Wolfe, J.B. Localization of the primary metabolic block produced by 2-deoxyglucose. J. Biol. Chem. 1957, 224, 963–969. [Google Scholar] [PubMed]
  48. Park, G.C.; Kim, J.S.; Roh, J.L.; Choi, S.H.; Nam, S.Y.; Kim, S.Y. Prognostic value of metabolic tumor volume measured by 18f-fdg pet/ct in advanced-stage squamous cell carcinoma of the larynx and hypopharynx. Ann. Oncol 2013, 24, 208–214. [Google Scholar] [CrossRef] [PubMed]
  49. Sandulache, V.C.; Ow, T.J.; Pickering, C.R.; Frederick, M.J.; Zhou, G.; Fokt, I.; Davis-Malesevich, M.; Priebe, W.; Myers, J.N. Glucose, not glutamine, is the dominant energy source required for proliferation and survival of head and neck squamous carcinoma cells. Cancer 2011, 117, 2926–2938. [Google Scholar] [CrossRef] [PubMed]
  50. Simons, A.L.; Ahmad, I.M.; Mattson, D.M.; Dornfeld, K.J.; Spitz, D.R. 2-deoxy-d-glucose combined with cisplatin enhances cytotoxicity via metabolic oxidative stress in human head and neck cancer cells. Cancer Res. 2007, 67, 3364–3370. [Google Scholar] [CrossRef]
  51. Simons, A.L.; Fath, M.A.; Mattson, D.M.; Smith, B.J.; Walsh, S.A.; Graham, M.M.; Hichwa, R.D.; Buatti, J.M.; Dornfeld, K.; Spitz, D.R. Enhanced response of human head and neck cancer xenograft tumors to cisplatin combined with 2-deoxy-d-glucose correlates with increased 18f-fdg uptake as determined by pet imaging. Int. J. Radiat Oncol. Biol. Phys. 2007, 69, 1222–1230. [Google Scholar] [CrossRef] [PubMed]
  52. Celentano, A.; McCullough, M.; Cirillo, N. Glucocorticoids reduce chemotherapeutic effectiveness on oscc cells via glucose-dependent mechanisms. J. Cell Physiol. 2019, 234, 2013–2020. [Google Scholar] [CrossRef] [PubMed]
  53. Kraus, D.; Reckenbeil, J.; Veit, N.; Kuerpig, S.; Meisenheimer, M.; Beier, I.; Stark, H.; Winter, J.; Probstmeier, R. Targeting glucose transport and the nad pathway in tumor cells with stf-31: A re-evaluation. Cell Oncol. 2018, 41, 485–494. [Google Scholar] [CrossRef] [PubMed]
  54. Jan, C.I.; Tsai, M.H.; Chiu, C.F.; Huang, Y.P.; Liu, C.J.; Chang, N.W. Fenofibrate suppresses oral tumorigenesis via reprogramming metabolic processes: Potential drug repurposing for oral cancer. Int. J. Biol. Sci. 2016, 12, 786–798. [Google Scholar] [CrossRef] [PubMed]
  55. Li, H.M.; Yang, J.G.; Liu, Z.J.; Wang, W.M.; Yu, Z.L.; Ren, J.G.; Chen, G.; Zhang, W.; Jia, J. Blockage of glycolysis by targeting pfkfb3 suppresses tumor growth and metastasis in head and neck squamous cell carcinoma. J. Exp. Clin. Cancer Res. 2017, 36, 7. [Google Scholar] [CrossRef] [PubMed]
  56. Lu, H.; Li, X.; Luo, Z.; Liu, J.; Fan, Z. Cetuximab reverses the warburg effect by inhibiting hif-1-regulated ldh-a. Mol. Cancer 2013, 12, 2187–2199. [Google Scholar] [CrossRef] [PubMed]
  57. Zhai, X.; Yang, Y.; Wan, J.; Zhu, R.; Wu, Y. Inhibition of ldh-a by oxamate induces g2/m arrest, apoptosis and increases radiosensitivity in nasopharyngeal carcinoma cells. Oncol. Rep. 2013, 30, 2983–2991. [Google Scholar] [CrossRef] [PubMed]
  58. Guan, X.; Rodriguez-Cruz, V.; Morris, M.E. Cellular uptake of mct1 inhibitors ar-c155858 and azd3965 and their effects on mct-mediated transport of l-lactate in murine 4t1 breast tumor cancer cells. AAPS J. 2019, 21, 13. [Google Scholar] [CrossRef]
  59. Mehibel, M.; Ortiz-Martinez, F.; Voelxen, N.; Boyers, A.; Chadwick, A.; Telfer, B.A.; Mueller-Klieser, W.; West, C.M.; Critchlow, S.E.; Williams, K.J.; et al. Statin-induced metabolic reprogramming in head and neck cancer: A biomarker for targeting monocarboxylate transporters. Sci. Rep. 2018, 8, 16804. [Google Scholar] [CrossRef]
  60. Elliott, R.L.; Jiang, X.P.; Head, J.F. Mitochondria organelle transplantation: Introduction of normal epithelial mitochondria into human cancer cells inhibits proliferation and increases drug sensitivity. Breast Cancer Res. Treat. 2012, 136, 347–354. [Google Scholar] [CrossRef]
  61. Kaipparettu, B.A.; Ma, Y.; Park, J.H.; Lee, T.L.; Zhang, Y.; Yotnda, P.; Creighton, C.J.; Chan, W.Y.; Wong, L.J. Crosstalk from non-cancerous mitochondria can inhibit tumor properties of metastatic cells by suppressing oncogenic pathways. PLoS ONE 2013, 8, e61747. [Google Scholar] [CrossRef] [PubMed]
  62. Arismendi-Morillo, G. Electron microscopy morphology of the mitochondrial network in human cancer. Int. J. Biochem. Cell Biol. 2009, 41, 2062–2068. [Google Scholar] [CrossRef] [PubMed]
  63. Porporato, P.E.; Filigheddu, N.; Pedro, J.M.B.; Kroemer, G.; Galluzzi, L. Mitochondrial metabolism and cancer. Cell Res. 2018, 28, 265–280. [Google Scholar] [CrossRef] [PubMed]
  64. Potter, M.; Newport, E.; Morten, K.J. The warburg effect: 80 years on. Biochem Soc. Trans. 2016, 44, 1499–1505. [Google Scholar] [CrossRef] [PubMed]
  65. Senyilmaz, D.; Teleman, A.A. Chicken or the egg: Warburg effect and mitochondrial dysfunction. F1000Prime Rep. 2015, 7, 41. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Sabharwal, S.S.; Schumacker, P.T. Mitochondrial ros in cancer: Initiators, amplifiers or an achilles’ heel? Nat. Rev. Cancer 2014, 14, 709–721. [Google Scholar] [CrossRef] [PubMed]
  67. Galadari, S.; Rahman, A.; Pallichankandy, S.; Thayyullathil, F. Reactive oxygen species and cancer paradox: To promote or to suppress? Free Radic Biol. Med. 2017, 104, 144–164. [Google Scholar] [CrossRef] [PubMed]
  68. Kim, E.H.; Baek, S.; Shin, D.; Lee, J.; Roh, J.L. Hederagenin induces apoptosis in cisplatin-resistant head and neck cancer cells by inhibiting the nrf2-are antioxidant pathway. Oxid Med. Cell Longev. 2017, 2017, 5498908. [Google Scholar] [CrossRef] [PubMed]
  69. Tang, J.Y.; Wu, C.Y.; Shu, C.W.; Wang, S.C.; Chang, M.Y.; Chang, H.W. A novel sulfonyl chromen-4-ones (chw09) preferentially kills oral cancer cells showing apoptosis, oxidative stress, and DNA damage. Env. Toxicol. 2018, 33, 1195–1203. [Google Scholar] [CrossRef] [PubMed]
  70. Wang, Y.Y.; Chen, Y.K.; Hsu, Y.L.; Chiu, W.C.; Tsai, C.H.; Hu, S.C.; Hsieh, P.W.; Yuan, S.F. Synthetic beta-nitrostyrene derivative cyt-rx20 as inhibitor of oral cancer cell proliferation and tumor growth through glutathione suppression and reactive oxygen species induction. Head Neck 2017, 39, 1055–1064. [Google Scholar] [CrossRef] [PubMed]
  71. Seo, S.U.; Cho, H.K.; Min, K.J.; Woo, S.M.; Kim, S.; Park, J.W.; Kim, S.H.; Choi, Y.H.; Keum, Y.S.; Hyun, J.W.; et al. Thioridazine enhances sensitivity to carboplatin in human head and neck cancer cells through downregulation of c-flip and mcl-1 expression. Cell Death Dis 2017, 8, e2599. [Google Scholar] [CrossRef] [PubMed]
  72. Chou, H.Y.; Chueh, F.S.; Ma, Y.S.; Wu, R.S.; Liao, C.L.; Chu, Y.L.; Fan, M.J.; Huang, W.W.; Chung, J.G. Bufalin induced apoptosis in scc4 human tongue cancer cells by decreasing bcl2 and increasing bax expression via the mitochondriadependent pathway. Mol. Med. Rep. 2017, 16, 7959–7966. [Google Scholar] [CrossRef] [PubMed]
  73. Shih, Y.L.; Hung, F.M.; Lee, C.H.; Yeh, M.Y.; Lee, M.H.; Lu, H.F.; Chen, Y.L.; Liu, J.Y.; Chung, J.G. Fisetin induces apoptosis of hsc3 human oral cancer cells through endoplasmic reticulum stress and dysfunction of mitochondria-mediated signaling pathways. In Vivo 2017, 31, 1103–1114. [Google Scholar] [PubMed]
  74. Chou, G.L.; Peng, S.F.; Liao, C.L.; Ho, H.C.; Lu, K.W.; Lien, J.C.; Fan, M.J.; La, K.C.; Chung, J.G. Casticin impairs cell growth and induces cell apoptosis via cell cycle arrest in human oral cancer scc-4 cells. Env. Toxicol. 2018, 33, 127–141. [Google Scholar] [CrossRef] [PubMed]
  75. Lam, T.G.; Jeong, Y.S.; Kim, S.A.; Ahn, S.G. New metformin derivative hl156a prevents oral cancer progression by inhibiting the insulin-like growth factor/akt/mammalian target of rapamycin pathways. Cancer Sci. 2018, 109, 699–709. [Google Scholar] [CrossRef] [PubMed]
  76. Nakashima, C.; Yamamoto, K.; Fujiwara-Tani, R.; Luo, Y.; Matsushima, S.; Fujii, K.; Ohmori, H.; Sasahira, T.; Sasaki, T.; Kitadai, Y.; et al. Expression of cytosolic malic enzyme (me1) is associated with disease progression in human oral squamous cell carcinoma. Cancer Sci. 2018, 109, 2036–2045. [Google Scholar] [CrossRef] [PubMed]
  77. Woo, S.H.; Yang, L.P.; Chuang, H.C.; Fitzgerald, A.; Lee, H.Y.; Pickering, C.; Myers, J.N.; Skinner, H.D. Down-regulation of malic enzyme 1 and 2: Sensitizing head and neck squamous cell carcinoma cells to therapy-induced senescence. Head Neck 2016, 38, 934–940. [Google Scholar] [CrossRef]
  78. Lee, J.; Huang, M.S.; Yang, I.C.; Lai, T.C.; Wang, J.L.; Pang, V.F.; Hsiao, M.; Kuo, M.Y. Essential roles of caspases and their upstream regulators in rotenone-induced apoptosis. Biochem. Biophys. Res. Commun. 2008, 371, 33–38. [Google Scholar] [CrossRef]
  79. Sandulache, V.C.; Skinner, H.D.; Ow, T.J.; Zhang, A.; Xia, X.; Luchak, J.M.; Wong, L.J.; Pickering, C.R.; Zhou, G.; Myers, J.N. Individualizing antimetabolic treatment strategies for head and neck squamous cell carcinoma based on tp53 mutational status. Cancer 2012, 118, 711–721. [Google Scholar] [CrossRef]
  80. Ashton, T.M.; Fokas, E.; Kunz-Schughart, L.A.; Folkes, L.K.; Anbalagan, S.; Huether, M.; Kelly, C.J.; Pirovano, G.; Buffa, F.M.; Hammond, E.M.; et al. The anti-malarial atovaquone increases radiosensitivity by alleviating tumour hypoxia. Nat. Commun. 2016, 7, 12308. [Google Scholar] [CrossRef] [Green Version]
  81. Zhang, Z.; Gao, Z.; Rajthala, S.; Sapkota, D.; Dongre, H.; Parajuli, H.; Suliman, S.; Das, R.; Li, L.; Bindoff, L.A.; et al. Metabolic reprogramming of normal oral fibroblasts correlated with increased glycolytic metabolism of oral squamous cell carcinoma and precedes their activation into carcinoma associated fibroblasts. Cell Mol. Life Sci. 2019, 1–19. [Google Scholar] [CrossRef] [PubMed]
  82. He, Y.; Gong, Y.; Gu, J.; Lee, J.J.; Lippman, S.M.; Wu, X. Increased leukocyte mitochondrial DNA copy number is associated with oral premalignant lesions: An epidemiology study. Carcinogenesis 2014, 35, 1760–1764. [Google Scholar] [CrossRef] [PubMed]
  83. Takeda, D.; Hasegawa, T.; Ueha, T.; Sakakibara, A.; Kawamoto, T.; Minamikawa, T.; Sakai, Y.; Komori, T. Decreased mitochondrial copy numbers in oral squamous cell carcinoma. Head Neck 2016, 38, 1170–1175. [Google Scholar] [CrossRef] [PubMed]
  84. Reznik, E.; Miller, M.L.; Senbabaoglu, Y.; Riaz, N.; Sarungbam, J.; Tickoo, S.K.; Al-Ahmadie, H.A.; Lee, W.; Seshan, V.E.; Hakimi, A.A.; et al. Mitochondrial DNA copy number variation across human cancers. Elife 2016, 5, e10769. [Google Scholar] [CrossRef]
  85. Kim, M.M.; Clinger, J.D.; Masayesva, B.G.; Ha, P.K.; Zahurak, M.L.; Westra, W.H.; Califano, J.A. Mitochondrial DNA quantity increases with histopathologic grade in premalignant and malignant head and neck lesions. Clin. Cancer Res. 2004, 10, 8512–8515. [Google Scholar] [CrossRef] [PubMed]
  86. Nakagawa, H.; Hayata, Y.; Kawamura, S.; Yamada, T.; Fujiwara, N.; Koike, K. Lipid metabolic reprogramming in hepatocellular carcinoma. Cancers 2018, 10, 447. [Google Scholar] [CrossRef] [PubMed]
  87. Dalal, S. Lipid metabolism in cancer cachexia. Ann. Palliat. Med. 2019, 8, 13–23. [Google Scholar] [CrossRef] [PubMed]
  88. Munir, R.; Lisec, J.; Swinnen, J.V.; Zaidi, N. Lipid metabolism in cancer cells under metabolic stress. Br. J. Cancer 2019, 120, 1090–1098. [Google Scholar] [CrossRef]
  89. Yu, X.H.; Ren, X.H.; Liang, X.H.; Tang, Y.L. Roles of fatty acid metabolism in tumourigenesis: Beyond providing nutrition (review). Mol. Med. Rep. 2018, 18, 5307–5316. [Google Scholar] [CrossRef]
  90. Fahy, E.; Subramaniam, S.; Murphy, R.C.; Nishijima, M.; Raetz, C.R.; Shimizu, T.; Spener, F.; van Meer, G.; Wakelam, M.J.; Dennis, E.A. Update of the lipid maps comprehensive classification system for lipids. J. Lipid Res. 2009, 50, 9–14. [Google Scholar] [CrossRef]
  91. Huang, C.; Freter, C. Lipid metabolism, apoptosis and cancer therapy. Int. J. Mol. Sci. 2015, 16, 924–949. [Google Scholar] [CrossRef] [PubMed]
  92. Currie, E.; Schulze, A.; Zechner, R.; Walther, T.C.; Farese, R.V., Jr. Cellular fatty acid metabolism and cancer. Cell Metab 2013, 18, 153–161. [Google Scholar] [CrossRef] [PubMed]
  93. Medes, G.; Thomas, A.; Weinhouse, S. Metabolism of neoplastic tissue. Iv. A study of lipid synthesis in neoplastic tissue slices in vitro. Cancer Res. 1953, 13, 27–29. [Google Scholar] [PubMed]
  94. Santos, C.R.; Schulze, A. Lipid metabolism in cancer. FEBS J. 2012, 279, 2610–2623. [Google Scholar] [CrossRef] [PubMed]
  95. Schlichtholz, B.; Turyn, J.; Goyke, E.; Biernacki, M.; Jaskiewicz, K.; Sledzinski, Z.; Swierczynski, J. Enhanced citrate synthase activity in human pancreatic cancer. Pancreas 2005, 30, 99–104. [Google Scholar] [CrossRef] [PubMed]
  96. Vazquez-Martin, A.; Colomer, R.; Brunet, J.; Lupu, R.; Menendez, J.A. Overexpression of fatty acid synthase gene activates her1/her2 tyrosine kinase receptors in human breast epithelial cells. Cell Prolif. 2008, 41, 59–85. [Google Scholar] [CrossRef] [PubMed]
  97. Ricoult, S.J.; Yecies, J.L.; Ben-Sahra, I.; Manning, B.D. Oncogenic pi3k and k-ras stimulate de novo lipid synthesis through mtorc1 and srebp. Oncogene 2016, 35, 1250–1260. [Google Scholar] [CrossRef] [PubMed]
  98. Wang, C.; Ma, J.; Zhang, N.; Yang, Q.; Jin, Y.; Wang, Y. The acetyl-coa carboxylase enzyme: A target for cancer therapy? Expert Rev. Anticancer 2015, 15, 667–676. [Google Scholar] [CrossRef]
  99. Igal, R.A. Stearoyl-coa desaturase-1: A novel key player in the mechanisms of cell proliferation, programmed cell death and transformation to cancer. Carcinogenesis 2010, 31, 1509–1515. [Google Scholar] [CrossRef]
  100. Bao, J.; Zhu, L.; Zhu, Q.; Su, J.; Liu, M.; Huang, W. Srebp-1 is an independent prognostic marker and promotes invasion and migration in breast cancer. Oncol. Lett. 2016, 12, 2409–2416. [Google Scholar] [CrossRef]
  101. Duan, J.; Sun, L.; Huang, H.; Wu, Z.; Wang, L.; Liao, W. Overexpression of fatty acid synthase predicts a poor prognosis for human gastric cancer. Mol. Med. Rep. 2016, 13, 3027–3035. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Lee, Y.L.; Li, W.C.; Tsai, T.H.; Chiang, H.Y.; Ting, C.T. Body mass index and cholesterol level predict surgical outcome in patients with hepatocellular carcinoma in taiwan - a cohort study. Oncotarget 2016, 7, 22948–22959. [Google Scholar] [CrossRef] [PubMed]
  103. Knox, J.J.; Siu, L.L.; Chen, E.; Dimitroulakos, J.; Kamel-Reid, S.; Moore, M.J.; Chin, S.; Irish, J.; LaFramboise, S.; Oza, A.M. A phase i trial of prolonged administration of lovastatin in patients with recurrent or metastatic squamous cell carcinoma of the head and neck or of the cervix. Eur. J. Cancer 2005, 41, 523–530. [Google Scholar] [CrossRef] [PubMed]
  104. Takeda, I.; Maruya, S.; Shirasaki, T.; Mizukami, H.; Takahata, T.; Myers, J.N.; Kakehata, S.; Yagihashi, S.; Shinkawa, H. Simvastatin inactivates beta1-integrin and extracellular signal-related kinase signaling and inhibits cell proliferation in head and neck squamous cell carcinoma cells. Cancer Sci. 2007, 98, 890–899. [Google Scholar] [CrossRef] [PubMed]
  105. Aksoy, A.; Kurnaz, S.C. An investigation of oxidative stress and coenzyme q10 levels in patients with head and neck squamous cell carcinomas. Eur. Arch. Otorhinolaryngol. 2019, 276, 1197–1204. [Google Scholar] [CrossRef] [PubMed]
  106. Gupta, A.; Bhatt, M.L.; Misra, M.K. Lipid peroxidation and antioxidant status in head and neck squamous cell carcinoma patients. Oxid Med. Cell Longev. 2009, 2, 68–72. [Google Scholar] [CrossRef]
  107. Rasheed, M.H.; Beevi, S.S.; Geetha, A. Enhanced lipid peroxidation and nitric oxide products with deranged antioxidant status in patients with head and neck squamous cell carcinoma. Oral Oncol. 2007, 43, 333–338. [Google Scholar] [CrossRef]
  108. Chan, L.P.; Chou, T.H.; Ding, H.Y.; Chen, P.R.; Chiang, F.Y.; Kuo, P.L.; Liang, C.H. Apigenin induces apoptosis via tumor necrosis factor receptor- and bcl-2-mediated pathway and enhances susceptibility of head and neck squamous cell carcinoma to 5-fluorouracil and cisplatin. Biochim. Biophys Acta. 2012, 1820, 1081–1091. [Google Scholar] [CrossRef]
  109. Huang, Y.; Bell, L.N.; Okamura, J.; Kim, M.S.; Mohney, R.P.; Guerrero-Preston, R.; Ratovitski, E.A. Phospho-deltanp63alpha/srebf1 protein interactions: Bridging cell metabolism and cisplatin chemoresistance. Cell Cycle 2012, 11, 3810–3827. [Google Scholar] [CrossRef]
  110. Su, Y.W.; Lin, Y.H.; Pai, M.H.; Lo, A.C.; Lee, Y.C.; Fang, I.C.; Lin, J.; Hsieh, R.K.; Chang, Y.F.; Chen, C.L. Association between phosphorylated amp-activated protein kinase and acetyl-coa carboxylase expression and outcome in patients with squamous cell carcinoma of the head and neck. PLoS ONE 2014, 9, e96183. [Google Scholar] [CrossRef]
  111. Luo, J.; Hong, Y.; Lu, Y.; Qiu, S.; Chaganty, B.K.; Zhang, L.; Wang, X.; Li, Q.; Fan, Z. Acetyl-coa carboxylase rewires cancer metabolism to allow cancer cells to survive inhibition of the warburg effect by cetuximab. Cancer Lett. 2017, 384, 39–49. [Google Scholar] [CrossRef] [PubMed]
  112. Li, K.; Zhang, C.; Chen, L.; Wang, P.; Fang, Y.; Zhu, J.; Chen, S.; Du, J.; Shen, B.; Wu, K.; et al. The role of acetyl-coa carboxylase2 in head and neck squamous cell carcinoma. Peer J. 2019, 7, e7037. [Google Scholar] [CrossRef] [PubMed]
  113. Mims, J.; Bansal, N.; Bharadwaj, M.S.; Chen, X.; Molina, A.J.; Tsang, A.W.; Furdui, C.M. Energy metabolism in a matched model of radiation resistance for head and neck squamous cell cancer. Radiat Res. 2015, 183, 291–304. [Google Scholar] [CrossRef] [PubMed]
  114. Vishwakarma, S.; Agarwal, R.; Goel, S.K.; Panday, R.K.; Singh, R.; Sukumaran, R.; Khare, S.; Kumar, A. Altered expression of sphingosine-1-phosphate metabolizing enzymes in oral cancer correlate with clinicopathological attributes. Cancer Invest. 2017, 35, 139–141. [Google Scholar] [CrossRef] [PubMed]
  115. Takano, K.; Kakuki, T.; Obata, K.; Nomura, K.; Miyata, R.; Kondo, A.; Kurose, M.; Kakiuchi, A.; Kaneko, Y.; Kohno, T.; et al. The behavior and role of lipolysis-stimulated lipoprotein receptor, a component of tricellular tight junctions, in head and neck squamous cell carcinomas. Anticancer Res. 2016, 36, 5895–5904. [Google Scholar] [CrossRef] [PubMed]
  116. Hu, Q.; Peng, J.; Chen, X.; Li, H.; Song, M.; Cheng, B.; Wu, T. Obesity and genes related to lipid metabolism predict poor survival in oral squamous cell carcinoma. Oral Oncol 2019, 89, 14–22. [Google Scholar] [CrossRef] [PubMed]
  117. Lukey, M.J.; Katt, W.P.; Cerione, R.A. Targeting amino acid metabolism for cancer therapy. Drug Discov. Today 2017, 22, 796–804. [Google Scholar] [CrossRef] [PubMed]
  118. Choi, B.H.; Coloff, J.L. The diverse functions of non-essential amino acids in cancer. Cancers 2019, 11, 675. [Google Scholar] [CrossRef]
  119. Reid, M.A.; Dai, Z.; Locasale, J.W. The impact of cellular metabolism on chromatin dynamics and epigenetics. Nat. Cell Biol. 2017, 19, 1298–1306. [Google Scholar] [CrossRef]
  120. Ide, A.L.; Portari, G.V.; Padovan, G.J.; Rosa, F.T.; Mello-Filho, F.V.; Marchini, J.S. Amino acids in squamous cell carcinomas and adjacent normal tissues from patients with larynx and oral cavity lesions. Clin. (Sao Paulo) 2012, 67, 1225–1227. [Google Scholar]
  121. Yonezawa, K.; Nishiumi, S.; Kitamoto-Matsuda, J.; Fujita, T.; Morimoto, K.; Yamashita, D.; Saito, M.; Otsuki, N.; Irino, Y.; Shinohara, M.; et al. Serum and tissue metabolomics of head and neck cancer. Cancer Genom. Proteom. 2013, 10, 233–238. [Google Scholar]
  122. Bothwell, P.J.; Kron, C.D.; Wittke, E.F.; Czerniak, B.N.; Bode, B.P. Targeted suppression and knockout of asct2 or lat1 in epithelial and mesenchymal human liver cancer cells fail to inhibit growth. Int. J. Mol. Sci. 2018, 19, 93. [Google Scholar] [CrossRef] [PubMed]
  123. Hafliger, P.; Charles, R.P. The l-type amino acid transporter lat1-an emerging target in cancer. Int. J. Mol. Sci. 2019, 20, 2093. [Google Scholar] [CrossRef] [PubMed]
  124. Kaira, K.; Oriuchi, N.; Imai, H.; Shimizu, K.; Yanagitani, N.; Sunaga, N.; Hisada, T.; Tanaka, S.; Ishizuka, T.; Kanai, Y.; et al. Prognostic significance of l-type amino acid transporter 1 expression in resectable stage i-iii nonsmall cell lung cancer. Br. J. Cancer 2008, 98, 742–748. [Google Scholar] [CrossRef] [PubMed]
  125. Kaira, K.; Sunose, Y.; Arakawa, K.; Ogawa, T.; Sunaga, N.; Shimizu, K.; Tominaga, H.; Oriuchi, N.; Itoh, H.; Nagamori, S.; et al. Prognostic significance of l-type amino-acid transporter 1 expression in surgically resected pancreatic cancer. Br. J. Cancer 2012, 107, 632–638. [Google Scholar] [CrossRef] [PubMed]
  126. Sugano, K.; Maeda, K.; Ohtani, H.; Nagahara, H.; Shibutani, M.; Hirakawa, K. Expression of xct as a predictor of disease recurrence in patients with colorectal cancer. Anticancer Res. 2015, 35, 677–682. [Google Scholar] [PubMed]
  127. Wang, J.; Papanicolau-Sengos, A.; Chintala, S.; Wei, L.; Liu, B.; Hu, Q.; Miles, K.M.; Conroy, J.M.; Glenn, S.T.; Costantini, M.; et al. Collecting duct carcinoma of the kidney is associated with cdkn2a deletion and slc family gene up-regulation. Oncotarget 2016, 7, 29901–29915. [Google Scholar] [CrossRef] [PubMed]
  128. Wu, Y.; Sun, X.; Song, B.; Qiu, X.; Zhao, J. Mir-375/slc7a11 axis regulates oral squamous cell carcinoma proliferation and invasion. Cancer Med. 2017, 6, 1686–1697. [Google Scholar] [CrossRef] [PubMed]
  129. Nobusawa, A.; Kim, M.; Kaira, K.; Miyashita, G.; Negishi, A.; Oriuchi, N.; Higuchi, T.; Tsushima, Y.; Kanai, Y.; Yokoo, S.; et al. Diagnostic usefulness of (1)(8)f-famt pet and l-type amino acid transporter 1 (lat1) expression in oral squamous cell carcinoma. Eur. J. Nucl Med. Mol. Imaging 2013, 40, 1692–1700. [Google Scholar] [CrossRef] [PubMed]
  130. Ueno, S.; Kimura, T.; Yamaga, T.; Kawada, A.; Ochiai, T.; Endou, H.; Sakurai, H. Metformin enhances anti-tumor effect of l-type amino acid transporter 1 (lat1) inhibitor. J. Pharm. Sci. 2016, 131, 110–117. [Google Scholar] [CrossRef] [PubMed]
  131. Saito, Y.; Moriya, S.; Kazama, H.; Hirasawa, K.; Miyahara, K.; Kokuba, H.; Hino, H.; Kikuchi, H.; Takano, N.; Hiramoto, M.; et al. Amino acid starvation culture condition sensitizes egfr-expressing cancer cell lines to gefitinib-mediated cytotoxicity by inducing atypical necroptosis. Int. J. Oncol. 2018, 52, 1165–1177. [Google Scholar] [CrossRef] [PubMed]
  132. Tao, X.; Lu, Y.; Qiu, S.; Wang, Y.; Qin, J.; Fan, Z. Ap1g1 is involved in cetuximab-mediated downregulation of asct2-egfr complex and sensitization of human head and neck squamous cell carcinoma cells to ros-induced apoptosis. Cancer Lett. 2017, 408, 33–42. [Google Scholar] [CrossRef] [PubMed]
  133. Yoshikawa, M.; Tsuchihashi, K.; Ishimoto, T.; Yae, T.; Motohara, T.; Sugihara, E.; Onishi, N.; Masuko, T.; Yoshizawa, K.; Kawashiri, S.; et al. Xct inhibition depletes cd44v-expressing tumor cells that are resistant to egfr-targeted therapy in head and neck squamous cell carcinoma. Cancer Res. 2013, 73, 1855–1866. [Google Scholar] [CrossRef] [PubMed]
  134. Zhang, P.; Wang, W.; Wei, Z.; Xu, L.I.; Yang, X.; Du, Y. Xct expression modulates cisplatin resistance in tca8113 tongue carcinoma cells. Oncol Lett. 2016, 12, 307–314. [Google Scholar] [CrossRef] [PubMed]
  135. Roh, J.L.; Kim, E.H.; Jang, H.J.; Park, J.Y.; Shin, D. Induction of ferroptotic cell death for overcoming cisplatin resistance of head and neck cancer. Cancer Lett. 2016, 381, 96–103. [Google Scholar] [CrossRef] [PubMed]
  136. Zhang, N. Role of methionine on epigenetic modification of DNA methylation and gene expression in animals. Anim Nutr. 2018, 4, 11–16. [Google Scholar] [CrossRef] [PubMed]
  137. Hamstra, D.A.; Lee, K.C.; Eisbruch, A.; Sunkara, P.; Borgonha, S.; Phillip, B.; Campbell, K.C.M.; Ross, B.D.; Rehemtulla, A. Double-blind placebo-controlled multicenter phase ii trial to evaluate d-methionine in preventing/reducing oral mucositis induced by radiation and chemotherapy for head and neck cancer. Head Neck 2018, 40, 1375–1388. [Google Scholar] [CrossRef] [PubMed]
  138. Vsiansky, V.; Svobodova, M.; Gumulec, J.; Cernei, N.; Sterbova, D.; Zitka, O.; Kostrica, R.; Smilek, P.; Plzak, J.; Betka, J.; et al. Prognostic significance of serum free amino acids in head and neck cancers. Cells 2019, 8, 428. [Google Scholar] [CrossRef]
  139. Huang, C.C.; Tsai, S.T.; Kuo, C.C.; Chang, J.S.; Jin, Y.T.; Chang, J.Y.; Hsiao, J.R. Arginine deprivation as a new treatment strategy for head and neck cancer. Oral Oncol. 2012, 48, 1227–1235. [Google Scholar] [CrossRef]
  140. Tomlinson, B.K.; Thomson, J.A.; Bomalaski, J.S.; Diaz, M.; Akande, T.; Mahaffey, N.; Li, T.; Dutia, M.P.; Kelly, K.; Gong, I.Y.; et al. Phase i trial of arginine deprivation therapy with adi-peg 20 plus docetaxel in patients with advanced malignant solid tumors. Clin. Cancer Res. 2015, 21, 2480–2486. [Google Scholar] [CrossRef]
  141. Kamarajan, P.; Rajendiran, T.M.; Kinchen, J.; Bermudez, M.; Danciu, T.; Kapila, Y.L. Head and neck squamous cell carcinoma metabolism draws on glutaminolysis, and stemness is specifically regulated by glutaminolysis via aldehyde dehydrogenase. J. Proteome Res. 2017, 16, 1315–1326. [Google Scholar] [CrossRef] [PubMed]
  142. Yang, J.; Guo, Y.; Seo, W.; Zhang, R.; Lu, C.; Wang, Y.; Luo, L.; Paul, B.; Yan, W.; Saxena, D.; et al. Targeting cellular metabolism to reduce head and neck cancer growth. Sci. Rep. 2019, 9, 4995. [Google Scholar] [CrossRef] [PubMed]
  143. Carninci, P.; Kasukawa, T.; Katayama, S.; Gough, J.; Frith, M.C.; Maeda, N.; Oyama, R.; Ravasi, T.; Lenhard, B.; Wells, C.; et al. The transcriptional landscape of the mammalian genome. Science 2005, 309, 1559–1563. [Google Scholar] [PubMed]
  144. Esteller, M. Non-coding rnas in human disease. Nat. Rev. Genet. 2011, 12, 861–874. [Google Scholar] [CrossRef] [PubMed]
  145. Denaro, N.; Merlano, M.C.; Russi, E.G.; Lo Nigro, C. Non coding rnas in head and neck squamous cell carcinoma (hnscc): A clinical perspective. Anticancer Res. 2014, 34, 6887–6896. [Google Scholar] [PubMed]
  146. Guglas, K.; Bogaczynska, M.; Kolenda, T.; Rys, M.; Teresiak, A.; Blizniak, R.; Lasinska, I.; Mackiewicz, J.; Lamperska, K. Lncrna in hnscc: Challenges and potential. Contemp Oncol. (Pozn) 2017, 21, 259–266. [Google Scholar] [CrossRef] [PubMed]
  147. Chen, Y.F.; Wei, Y.Y.; Yang, C.C.; Liu, C.J.; Yeh, L.Y.; Chou, C.H.; Chang, K.W.; Lin, S.C. Mir-125b suppresses oral oncogenicity by targeting the anti-oxidative gene prxl2a. Redox Biol. 2019, 22, 101140. [Google Scholar] [CrossRef] [PubMed]
  148. Chen, Y.F.; Yang, C.C.; Kao, S.Y.; Liu, C.J.; Lin, S.C.; Chang, K.W. Microrna-211 enhances the oncogenicity of carcinogen-induced oral carcinoma by repressing tcf12 and increasing antioxidant activity. Cancer Res. 2016, 76, 4872–4886. [Google Scholar] [CrossRef] [PubMed]
  149. Sounni, N.E.; Noel, A. Targeting the tumor microenvironment for cancer therapy. Clin. Chem. 2013, 59, 85–93. [Google Scholar] [CrossRef] [PubMed]
  150. Chang, C.H.; Qiu, J.; O’Sullivan, D.; Buck, M.D.; Noguchi, T.; Curtis, J.D.; Chen, Q.; Gindin, M.; Gubin, M.M.; van der Windt, G.J.; et al. Metabolic competition in the tumor microenvironment is a driver of cancer progression. Cell 2015, 162, 1229–1241. [Google Scholar] [CrossRef] [PubMed]
  151. Li, Z.; Li, X.; Wu, S.; Xue, M.; Chen, W. Long non-coding rna uca1 promotes glycolysis by upregulating hexokinase 2 through the mtor-stat3/microrna143 pathway. Cancer Sci. 2014, 105, 951–955. [Google Scholar] [CrossRef] [PubMed]
  152. Cui, M.; Xiao, Z.; Wang, Y.; Zheng, M.; Song, T.; Cai, X.; Sun, B.; Ye, L.; Zhang, X. Long noncoding rna hulc modulates abnormal lipid metabolism in hepatoma cells through an mir-9-mediated rxra signaling pathway. Cancer Res. 2015, 75, 846–857. [Google Scholar] [CrossRef] [PubMed]
  153. Lin, C.; Zhang, Y.; Chen, Y.; Bai, Y.; Zhang, Y. Long noncoding rna linc01234 promotes serine hydroxymethyltransferase 2 expression and proliferation by competitively binding mir-642a-5p in colon cancer. Cell Death Dis. 2019, 10, 137. [Google Scholar] [CrossRef] [PubMed]
  154. Peng, W.X.; Koirala, P.; Mo, Y.Y. Lncrna-mediated regulation of cell signaling in cancer. Oncogene 2017, 36, 5661–5667. [Google Scholar] [CrossRef] [PubMed]
  155. Ding, L.; Ren, J.; Zhang, D.; Li, Y.; Huang, X.; Hu, Q.; Wang, H.; Song, Y.; Ni, Y.; Hou, Y. A novel stromal lncrna signature reprograms fibroblasts to promote the growth of oral squamous cell carcinoma via lncrna-caf/interleukin-33. Carcinogenesis 2018, 39, 397–406. [Google Scholar] [CrossRef]
  156. Xu, S.; Guo, J.; Zhang, W. Lncrna pcat19 promotes the proliferation of laryngocarcinoma cells via modulation of the mir-182/pdk4 axis. J. Cell Biochem. 2019, 120, 12810–12821. [Google Scholar] [CrossRef]
  157. Shih, J.W.; Chiang, W.F.; Wu, A.T.H.; Wu, M.H.; Wang, L.Y.; Yu, Y.L.; Hung, Y.W.; Wang, W.C.; Chu, C.Y.; Hung, C.L.; et al. Long noncoding rna lnchifcar/mir31hg is a hif-1alpha co-activator driving oral cancer progression. Nat. Commun. 2017, 8, 15874. [Google Scholar] [CrossRef]
  158. Zhu, G.; Wang, S.; Chen, J.; Wang, Z.; Liang, X.; Wang, X.; Jiang, J.; Lang, J.; Li, L. Long noncoding rna has2-as1 mediates hypoxia-induced invasiveness of oral squamous cell carcinoma. Mol. Carcinog. 2017, 56, 2210–2222. [Google Scholar] [CrossRef]
  159. Rupaimoole, R.; Calin, G.A.; Lopez-Berestein, G.; Sood, A.K. Mirna deregulation in cancer cells and the tumor microenvironment. Cancer Discov. 2016, 6, 235–246. [Google Scholar] [CrossRef]
  160. Kumar, B.; Yadav, A.; Lang, J.; Teknos, T.N.; Kumar, P. Dysregulation of microrna-34a expression in head and neck squamous cell carcinoma promotes tumor growth and tumor angiogenesis. PLoS ONE 2012, 7, e37601. [Google Scholar] [CrossRef]
  161. Li, Y.Y.; Tao, Y.W.; Gao, S.; Li, P.; Zheng, J.M.; Zhang, S.E.; Liang, J.; Zhang, Y. Cancer-associated fibroblasts contribute to oral cancer cells proliferation and metastasis via exosome-mediated paracrine mir-34a-5p. EBioMedicine 2018, 36, 209–220. [Google Scholar] [CrossRef] [PubMed]
  162. Wang, X.; Li, J.; Dong, K.; Lin, F.; Long, M.; Ouyang, Y.; Wei, J.; Chen, X.; Weng, Y.; He, T.; et al. Tumor suppressor mir-34a targets pd-l1 and functions as a potential immunotherapeutic target in acute myeloid leukemia. Cell Signal. 2015, 27, 443–452. [Google Scholar] [CrossRef] [PubMed]
  163. Rapado-Gonzalez, O.; Majem, B.; Muinelo-Romay, L.; Alvarez-Castro, A.; Santamaria, A.; Gil-Moreno, A.; Lopez-Lopez, R.; Suarez-Cunqueiro, M.M. Human salivary micrornas in cancer. J. Cancer 2018, 9, 638–649. [Google Scholar] [CrossRef] [PubMed]
  164. Camps, C.; Buffa, F.M.; Colella, S.; Moore, J.; Sotiriou, C.; Sheldon, H.; Harris, A.L.; Gleadle, J.M.; Ragoussis, J. Hsa-mir-210 is induced by hypoxia and is an independent prognostic factor in breast cancer. Clin. Cancer Res. 2008, 14, 1340–1348. [Google Scholar] [CrossRef] [PubMed]
  165. Ho, A.S.; Huang, X.; Cao, H.; Christman-Skieller, C.; Bennewith, K.; Le, Q.T.; Koong, A.C. Circulating mir-210 as a novel hypoxia marker in pancreatic cancer. Transl. Oncol. 2010, 3, 109–113. [Google Scholar] [CrossRef] [PubMed]
  166. Huang, X.; Ding, L.; Bennewith, K.L.; Tong, R.T.; Welford, S.M.; Ang, K.K.; Story, M.; Le, Q.T.; Giaccia, A.J. Hypoxia-inducible mir-210 regulates normoxic gene expression involved in tumor initiation. Mol. Cell 2009, 35, 856–867. [Google Scholar] [CrossRef] [PubMed]
  167. Favaro, E.; Ramachandran, A.; McCormick, R.; Gee, H.; Blancher, C.; Crosby, M.; Devlin, C.; Blick, C.; Buffa, F.; Li, J.L.; et al. Microrna-210 regulates mitochondrial free radical response to hypoxia and krebs cycle in cancer cells by targeting iron sulfur cluster protein iscu. PLoS ONE 2010, 5, e10345. [Google Scholar] [CrossRef]
  168. Saenz-de-Santa-Maria, I.; Bernardo-Castineira, C.; Secades, P.; Bernaldo-de-Quiros, S.; Rodrigo, J.P.; Astudillo, A.; Chiara, M.D. Clinically relevant hif-1alpha-dependent metabolic reprogramming in oropharyngeal squamous cell carcinomas includes coordinated activation of caix and the mir-210/iscu signaling axis, but not mct1 and mct4 upregulation. Oncotarget 2017, 8, 13730–13746. [Google Scholar] [CrossRef]
  169. Xu, P.; Li, Y.; Zhang, H.; Li, M.; Zhu, H. Microrna-340 mediates metabolic shift in oral squamous cell carcinoma by targeting glucose transporter-1. J. Oral Maxillofac Surg. 2016, 74, 844–850. [Google Scholar] [CrossRef]
  170. Yoshizawa, J.M.; Wong, D.T. Salivary micrornas and oral cancer detection. Methods Mol. Biol. 2013, 936, 313–324. [Google Scholar]
  171. Schneider, A.; Victoria, B.; Lopez, Y.N.; Suchorska, W.; Barczak, W.; Sobecka, A.; Golusinski, W.; Masternak, M.M.; Golusinski, P. Tissue and serum microrna profile of oral squamous cell carcinoma patients. Sci. Rep. 2018, 8, 675. [Google Scholar] [CrossRef] [PubMed]
  172. Chandra Gupta, S.; Nandan Tripathi, Y. Potential of long non-coding rnas in cancer patients: From biomarkers to therapeutic targets. Int. J. Cancer 2017, 140, 1955–1967. [Google Scholar] [CrossRef] [PubMed]
  173. Ren, W.; Qiang, C.; Gao, L.; Li, S.M.; Zhang, L.M.; Wang, X.L.; Dong, J.W.; Chen, C.; Liu, C.Y.; Zhi, K.Q. Circulating microrna-21 (mir-21) and phosphatase and tensin homolog (pten) are promising novel biomarkers for detection of oral squamous cell carcinoma. Biomarkers 2014, 19, 590–596. [Google Scholar] [CrossRef] [PubMed]
  174. Hung, P.S.; Liu, C.J.; Chou, C.S.; Kao, S.Y.; Yang, C.C.; Chang, K.W.; Chiu, T.H.; Lin, S.C. Mir-146a enhances the oncogenicity of oral carcinoma by concomitant targeting of the irak1, traf6 and numb genes. PLoS ONE 2013, 8, e79926. [Google Scholar] [CrossRef] [PubMed]
  175. Tang, H.; Wu, Z.; Zhang, J.; Su, B. Salivary lncrna as a potential marker for oral squamous cell carcinoma diagnosis. Mol. Med. Rep. 2013, 7, 761–766. [Google Scholar] [CrossRef] [PubMed]
  176. Park, N.J.; Zhou, H.; Elashoff, D.; Henson, B.S.; Kastratovic, D.A.; Abemayor, E.; Wong, D.T. Salivary microrna: Discovery, characterization, and clinical utility for oral cancer detection. Clin. Cancer Res. 2009, 15, 5473–5477. [Google Scholar] [CrossRef] [PubMed]
  177. Ye, M.; Ye, P.; Zhang, W.; Rao, J.; Xie, Z. diagnostic values of salivary versus and plasma microrna-21 for early esophageal cancer. J. Southern Med. University 2014, 34, 885–889. [Google Scholar]
  178. Xie, Z.J.; Chen, G.; Zhang, X.C.; Li, D.F.; Huang, J.; Li, Z.J. Saliva supernatant mir-21: A novel potential biomarker for esophageal cancer detection. Asian Pac. J. Cancer Prev. 2012, 13, 6145–6149. [Google Scholar] [CrossRef]
  179. Liu, C.J.; Kao, S.Y.; Tu, H.F.; Tsai, M.M.; Chang, K.W.; Lin, S.C. Increase of microrna mir-31 level in plasma could be a potential marker of oral cancer. Oral Dis. 2010, 16, 360–364. [Google Scholar] [CrossRef]
  180. Liu, C.J.; Lin, S.C.; Yang, C.C.; Cheng, H.W.; Chang, K.W. Exploiting salivary mir-31 as a clinical biomarker of oral squamous cell carcinoma. Head Neck 2012, 34, 219–224. [Google Scholar] [CrossRef]
  181. Kao, Y.Y.; Tu, H.F.; Kao, S.Y.; Chang, K.W.; Lin, S.C. The increase of oncogenic mirna expression in tongue carcinogenesis of a mouse model. Oral Oncol. 2015, 51, 1103–1112. [Google Scholar] [CrossRef] [PubMed]
  182. Li, S.; Chen, X.; Liu, X.; Yu, Y.; Pan, H.; Haak, R.; Schmidt, J.; Ziebolz, D.; Schmalz, G. Complex integrated analysis of lncrnas-mirnas-mrnas in oral squamous cell carcinoma. Oral Oncol. 2017, 73, 1–9. [Google Scholar] [CrossRef] [PubMed]
  183. Qiu, Y.L.; Liu, Y.H.; Ban, J.D.; Wang, W.J.; Han, M.; Kong, P.; Li, B.H. Pathway analysis of a genomewide association study on a long noncoding rna expression profile in oral squamous cell carcinoma. Oncol. Rep. 2019, 41, 895–907. [Google Scholar] [PubMed]
  184. Curry, J.M.; Sprandio, J.; Cognetti, D.; Luginbuhl, A.; Bar-ad, V.; Pribitkin, E.; Tuluc, M. Tumor microenvironment in head and neck squamous cell carcinoma. Semin. Oncol. 2014, 41, 217–234. [Google Scholar] [CrossRef] [PubMed]
  185. Freiser, M.E.; Serafini, P.; Weed, D.T. The immune system and head and neck squamous cell carcinoma: From carcinogenesis to new therapeutic opportunities. Immunol. Res. 2013, 57, 52–69. [Google Scholar] [CrossRef] [PubMed]
  186. Lim, Y.Z.; South, A.P. Tumour-stroma crosstalk in the development of squamous cell carcinoma. Int. J. Biochem. Cell Biol. 2014, 53, 450–458. [Google Scholar] [CrossRef]
  187. Schmitz, S.; Machiels, J.P. Targeting the tumor environment in squamous cell carcinoma of the head and neck. Curr. Treat. Options Oncol. 2016, 17, 37. [Google Scholar] [CrossRef]
  188. Peltanova, B.; Raudenska, M.; Masarik, M. Effect of tumor microenvironment on pathogenesis of the head and neck squamous cell carcinoma: A systematic review. Mol. Cancer 2019, 18, 63. [Google Scholar] [CrossRef]
  189. Eisma, R.J.; Spiro, J.D.; Kreutzer, D.L. Vascular endothelial growth factor expression in head and neck squamous cell carcinoma. Am. J. Surg. 1997, 174, 513–517. [Google Scholar] [CrossRef]
  190. Mark, R.; Bermejo, J.L.; Bierhaus, A.; Plinkert, P.K.; Angel, P.; Hess, J. The receptor for advanced glycation end products is dispensable in a mouse model of oral and esophageal carcinogenesis. Histol. Histopathol. 2013, 28, 1585–1594. [Google Scholar]
  191. Kartha, V.K.; Stawski, L.; Han, R.; Haines, P.; Gallagher, G.; Noonan, V.; Kukuruzinska, M.; Monti, S.; Trojanowska, M. Pdgfrbeta is a novel marker of stromal activation in oral squamous cell carcinomas. PLoS ONE 2016, 11, e0154645. [Google Scholar] [CrossRef] [PubMed]
  192. Hartmann, S.; Bhola, N.E.; Grandis, J.R. Hgf/met signaling in head and neck cancer: Impact on the tumor microenvironment. Clin. Cancer Res. 2016, 22, 4005–4013. [Google Scholar] [CrossRef] [PubMed]
  193. Szturz, P.; Budikova, M.; Vermorken, J.B.; Horova, I.; Gal, B.; Raymond, E.; de Gramont, A.; Faivre, S. Prognostic value of c-met in head and neck cancer: A systematic review and meta-analysis of aggregate data. Oral Oncol. 2017, 74, 68–76. [Google Scholar] [CrossRef] [PubMed]
  194. Szturz, P.; Raymond, E.; Abitbol, C.; Albert, S.; de Gramont, A.; Faivre, S. Understanding c-met signalling in squamous cell carcinoma of the head & neck. Crit. Rev. Oncol. Hematol. 2017, 111, 39–51. [Google Scholar] [PubMed]
  195. Pal, S.K.; Nguyen, C.T.; Morita, K.I.; Miki, Y.; Kayamori, K.; Yamaguchi, A.; Sakamoto, K. Thbs1 is induced by tgfb1 in the cancer stroma and promotes invasion of oral squamous cell carcinoma. J. Oral Pathol. Med. 2016, 45, 730–739. [Google Scholar] [CrossRef] [PubMed]
  196. Mao, L.; Zhao, Z.L.; Yu, G.T.; Wu, L.; Deng, W.W.; Li, Y.C.; Liu, J.F.; Bu, L.L.; Liu, B.; Kulkarni, A.B.; et al. Gamma-secretase inhibitor reduces immunosuppressive cells and enhances tumour immunity in head and neck squamous cell carcinoma. Int. J. Cancer 2018, 142, 999–1009. [Google Scholar] [CrossRef]
  197. Le, P.N.; Keysar, S.B.; Miller, B.; Eagles, J.R.; Chimed, T.S.; Reisinger, J.; Gomez, K.E.; Nieto, C.; Jackson, B.C.; Somerset, H.L.; et al. Wnt signaling dynamics in head and neck squamous cell cancer tumor-stroma interactions. Mol. Carcinog. 2019, 58, 398–410. [Google Scholar] [CrossRef]
  198. Chen, Y.C.; Cheng, Y.H.; Kim, H.S.; Ingram, P.N.; Nor, J.E.; Yoon, E. Paired single cell co-culture microenvironments isolated by two-phase flow with continuous nutrient renewal. Lab. Chip. 2014, 14, 2941–2947. [Google Scholar] [CrossRef] [Green Version]
  199. Bisetto, S.; Whitaker-Menezes, D.; Wilski, N.A.; Tuluc, M.; Curry, J.; Zhan, T.; Snyder, C.M.; Martinez-Outschoorn, U.E.; Philp, N.J. Monocarboxylate transporter 4 (mct4) knockout mice have attenuated 4nqo induced carcinogenesis; a role for mct4 in driving oral squamous cell cancer. Front. Oncol. 2018, 8, 324. [Google Scholar] [CrossRef]
  200. Ohashi, T.; Aoki, M.; Tomita, H.; Akazawa, T.; Sato, K.; Kuze, B.; Mizuta, K.; Hara, A.; Nagaoka, H.; Inoue, N.; et al. M2-like macrophage polarization in high lactic acid-producing head and neck cancer. Cancer Sci. 2017, 108, 1128–1134. [Google Scholar] [CrossRef]
  201. Curry, J.; Tassone, P.; Gill, K.; Tuluc, M.; BarAd, V.; Mollaee, M.; Whitaker-Menezes, D.; Rodeck, U.; Luginbuhl, A.; Cognetti, D.; et al. Tumor metabolism in the microenvironment of nodal metastasis in oral squamous cell carcinoma. Otolaryngol. Head Neck Surg. 2017, 157, 798–807. [Google Scholar] [CrossRef] [PubMed]
  202. Krupar, R.; Robold, K.; Gaag, D.; Spanier, G.; Kreutz, M.; Renner, K.; Hellerbrand, C.; Hofstaedter, F.; Bosserhoff, A.K. Immunologic and metabolic characteristics of hpv-negative and hpv-positive head and neck squamous cell carcinomas are strikingly different. Virchows Arch. 2014, 465, 299–312. [Google Scholar] [CrossRef] [PubMed]
  203. Na, K.J.; Choi, H. Tumor metabolic features identified by (18)f-fdg pet correlate with gene networks of immune cell microenvironment in head and neck cancer. J. Nucl. Med. 2018, 59, 31–37. [Google Scholar] [CrossRef] [PubMed]
  204. Rosenthal, E.; McCrory, A.; Talbert, M.; Young, G.; Murphy-Ullrich, J.; Gladson, C. Elevated expression of tgf-beta1 in head and neck cancer-associated fibroblasts. Mol. Carcinog. 2004, 40, 116–121. [Google Scholar] [CrossRef] [PubMed]
  205. Metzler, V.M.; Pritz, C.; Riml, A.; Romani, A.; Tuertscher, R.; Steinbichler, T.; Dejaco, D.; Riechelmann, H.; Dudas, J. Separation of cell survival, growth, migration, and mesenchymal transdifferentiation effects of fibroblast secretome on tumor cells of head and neck squamous cell carcinoma. Tumour Biol. 2017, 39, 1010428317705507. [Google Scholar] [CrossRef] [PubMed]
  206. Wheeler, S.E.; Shi, H.; Lin, F.; Dasari, S.; Bednash, J.; Thorne, S.; Watkins, S.; Joshi, R.; Thomas, S.M. Enhancement of head and neck squamous cell carcinoma proliferation, invasion, and metastasis by tumor-associated fibroblasts in preclinical models. Head Neck 2014, 36, 385–392. [Google Scholar] [CrossRef]
  207. Rasanen, K.; Sriswasdi, S.; Valiga, A.; Tang, H.Y.; Zhang, G.; Perego, M.; Somasundaram, R.; Li, L.; Speicher, K.; Klein-Szanto, A.J.; et al. Comparative secretome analysis of epithelial and mesenchymal subpopulations of head and neck squamous cell carcinoma identifies s100a4 as a potential therapeutic target. Mol. Cell Proteom. 2013, 12, 3778–3792. [Google Scholar] [CrossRef]
  208. Jiang, E.; Xu, Z.; Wang, M.; Yan, T.; Huang, C.; Zhou, X.; Liu, Q.; Wang, L.; Chen, Y.; Wang, H.; et al. Tumoral microvesicle-activated glycometabolic reprogramming in fibroblasts promotes the progression of oral squamous cell carcinoma. FASEB J. 2019, 33, 5690–5703. [Google Scholar] [CrossRef]
  209. Reichert, T.E.; Strauss, L.; Wagner, E.M.; Gooding, W.; Whiteside, T.L. Signaling abnormalities, apoptosis, and reduced proliferation of circulating and tumor-infiltrating lymphocytes in patients with oral carcinoma. Clin. Cancer Res. 2002, 8, 3137–3145. [Google Scholar]
  210. Strauss, L.; Bergmann, C.; Gooding, W.; Johnson, J.T.; Whiteside, T.L. The frequency and suppressor function of cd4+cd25highfoxp3+ t cells in the circulation of patients with squamous cell carcinoma of the head and neck. Clin. Cancer Res. 2007, 13, 6301–6311. [Google Scholar] [CrossRef]
  211. Overacre-Delgoffe, A.E.; Chikina, M.; Dadey, R.E.; Yano, H.; Brunazzi, E.A.; Shayan, G.; Horne, W.; Moskovitz, J.M.; Kolls, J.K.; Sander, C.; et al. Interferon-gamma drives treg fragility to promote anti-tumor immunity. Cell 2017, 169, 1130–1141. [Google Scholar] [CrossRef] [PubMed]
  212. Kesselring, R.; Thiel, A.; Pries, R.; Trenkle, T.; Wollenberg, B. Human th17 cells can be induced through head and neck cancer and have a functional impact on hnscc development. Br. J. Cancer 2010, 103, 1245–1254. [Google Scholar] [CrossRef] [PubMed]
  213. Chen, S.F.; Nieh, S.; Jao, S.W.; Wu, M.Z.; Liu, C.L.; Chang, Y.C.; Lin, Y.S. The paracrine effect of cancer-associated fibroblast-induced interleukin-33 regulates the invasiveness of head and neck squamous cell carcinoma. J. Pathol. 2013, 231, 180–189. [Google Scholar] [CrossRef] [PubMed]
  214. Tsujikawa, T.; Yaguchi, T.; Ohmura, G.; Ohta, S.; Kobayashi, A.; Kawamura, N.; Fujita, T.; Nakano, H.; Shimada, T.; Takahashi, T.; et al. Autocrine and paracrine loops between cancer cells and macrophages promote lymph node metastasis via ccr4/ccl22 in head and neck squamous cell carcinoma. Int. J. Cancer 2013, 132, 2755–2766. [Google Scholar] [CrossRef] [PubMed]
  215. Schuler, P.J.; Westerkamp, A.M.; Kansy, B.A.; Bruderek, K.; Dissmann, P.A.; Dumitru, C.A.; Lang, S.; Jackson, E.K.; Brandau, S. Adenosine metabolism of human mesenchymal stromal cells isolated from patients with head and neck squamous cell carcinoma. Immunobiology 2017, 222, 66–74. [Google Scholar] [CrossRef] [PubMed]
  216. Havel, J.J.; Chowell, D.; Chan, T.A. The evolving landscape of biomarkers for checkpoint inhibitor immunotherapy. Nat. Rev. Cancer 2019, 19, 133–150. [Google Scholar] [CrossRef] [PubMed]
  217. Li, J.; Jie, H.B.; Lei, Y.; Gildener-Leapman, N.; Trivedi, S.; Green, T.; Kane, L.P.; Ferris, R.L. Pd-1/shp-2 inhibits tc1/th1 phenotypic responses and the activation of t cells in the tumor microenvironment. Cancer Res. 2015, 75, 508–518. [Google Scholar] [CrossRef]
  218. Wen, Y.H.; Lin, H.Q.; Li, H.; Zhao, Y.; Lui, V.W.Y.; Chen, L.; Wu, X.M.; Sun, W.; Wen, W.P. Stromal interleukin-33 promotes regulatory t cell-mediated immunosuppression in head and neck squamous cell carcinoma and correlates with poor prognosis. Cancer Immunol. Immunother 2019, 68, 221–232. [Google Scholar] [CrossRef]
  219. Nguyen, N.; Bellile, E.; Thomas, D.; McHugh, J.; Rozek, L.; Virani, S.; Peterson, L.; Carey, T.E.; Walline, H.; Moyer, J.; et al. Tumor infiltrating lymphocytes and survival in patients with head and neck squamous cell carcinoma. Head Neck 2016, 38, 1074–1084. [Google Scholar] [CrossRef] [Green Version]
  220. Tao, Y.; Gross, N.; Liu, Y.; Zhang, L.; Li, G.; Huang, Z.; Yang, J. A high ratio of il-12rbeta2-positive tumor-infiltrating lymphocytes indicates favorable prognosis in laryngeal cancer. Oral Oncol 2017, 74, 148–156. [Google Scholar] [CrossRef]
  221. Wallis, S.P.; Stafford, N.D.; Greenman, J. Clinical relevance of immune parameters in the tumor microenvironment of head and neck cancers. Head Neck 2015, 37, 449–459. [Google Scholar] [CrossRef] [PubMed]
  222. Sakakura, K.; Takahashi, H.; Kaira, K.; Toyoda, M.; Murata, T.; Ohnishi, H.; Oyama, T.; Chikamatsu, K. Relationship between tumor-associated macrophage subsets and cd47 expression in squamous cell carcinoma of the head and neck in the tumor microenvironment. Lab. Invest. 2016, 96, 994–1003. [Google Scholar] [CrossRef] [PubMed]
  223. Troiano, G.; Caponio, V.C.A.; Zhurakivska, K.; Arena, C.; Pannone, G.; Mascitti, M.; Santarelli, A.; Lo Muzio, L. High pd-l1 expression in the tumour cells did not correlate with poor prognosis of patients suffering for oral squamous cells carcinoma: A meta-analysis of the literature. Cell Prolif 2019, 52, e12537. [Google Scholar] [CrossRef] [PubMed]
  224. Petrova, V.; Annicchiarico-Petruzzelli, M.; Melino, G.; Amelio, I. The hypoxic tumour microenvironment. Oncogenesis 2018, 7, 10. [Google Scholar] [CrossRef] [PubMed]
  225. Muz, B.; de la Puente, P.; Azab, F.; Azab, A.K. The role of hypoxia in cancer progression, angiogenesis, metastasis, and resistance to therapy. Hypoxia (Auckl) 2015, 3, 83–92. [Google Scholar] [CrossRef] [PubMed]
  226. Lartigau, E.; Le Ridant, A.M.; Lambin, P.; Weeger, P.; Martin, L.; Sigal, R.; Lusinchi, A.; Luboinski, B.; Eschwege, F.; Guichard, M. Oxygenation of head and neck tumors. Cancer 1993, 71, 2319–2325. [Google Scholar] [CrossRef]
  227. Zhu, G.; Tang, Y.; Geng, N.; Zheng, M.; Jiang, J.; Li, L.; Li, K.; Lei, Z.; Chen, W.; Fan, Y.; et al. Hif-alpha/mif and nf-kappab/il-6 axes contribute to the recruitment of cd11b+gr-1+ myeloid cells in hypoxic microenvironment of hnscc. Neoplasia 2014, 16, 168–179. [Google Scholar] [CrossRef]
  228. Li, L.; Li, C.; Wang, S.; Wang, Z.; Jiang, J.; Wang, W.; Li, X.; Chen, J.; Liu, K.; Li, C.; et al. Exosomes derived from hypoxic oral squamous cell carcinoma cells deliver mir-21 to normoxic cells to elicit a prometastatic phenotype. Cancer Res. 2016, 76, 1770–1780. [Google Scholar] [CrossRef]
  229. Mayer, A.; Zahnreich, S.; Brieger, J.; Vaupel, P.; Schmidberger, H. Downregulation of egfr in hypoxic, diffusion-limited areas of squamous cell carcinomas of the head and neck. Br. J. Cancer 2016, 115, 1351–1358. [Google Scholar] [CrossRef]
  230. Fluegen, G.; Avivar-Valderas, A.; Wang, Y.; Padgen, M.R.; Williams, J.K.; Nobre, A.R.; Calvo, V.; Cheung, J.F.; Bravo-Cordero, J.J.; Entenberg, D.; et al. Phenotypic heterogeneity of disseminated tumour cells is preset by primary tumour hypoxic microenvironments. Nat. Cell Biol. 2017, 19, 120–132. [Google Scholar] [CrossRef] [Green Version]
  231. Leelahavanichkul, K.; Amornphimoltham, P.; Molinolo, A.A.; Basile, J.R.; Koontongkaew, S.; Gutkind, J.S. A role for p38 mapk in head and neck cancer cell growth and tumor-induced angiogenesis and lymphangiogenesis. Mol. Oncol. 2014, 8, 105–118. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Metabolic reprogramming in HNSCCs. (A) Diagraphic illustration of the metabolic shift during the oncogenic transformation in HNSCC cells. The pathways presented in red indicate the pathways upregulated in HNSCC cells compared to normal cells, while the metabolic pathways shown in green are pathways less active in tumors; (B) the nutritional balance (e.g., maintenance of normoglycemia in DM patients) and reverse for intrinsic metabolic cues by inhibitors could be a potential method to suppress cancerous identity in HNSCCs. E1-E11: Enzymes in glycolytic pathway.
Figure 1. Metabolic reprogramming in HNSCCs. (A) Diagraphic illustration of the metabolic shift during the oncogenic transformation in HNSCC cells. The pathways presented in red indicate the pathways upregulated in HNSCC cells compared to normal cells, while the metabolic pathways shown in green are pathways less active in tumors; (B) the nutritional balance (e.g., maintenance of normoglycemia in DM patients) and reverse for intrinsic metabolic cues by inhibitors could be a potential method to suppress cancerous identity in HNSCCs. E1-E11: Enzymes in glycolytic pathway.
Ijms 20 03960 g001
Figure 2. Exploring TME-associated tumor activators/inhibitors in HNSCCs. A comparison between the metabolic profiles of sorted tumor cells, CAFs and tumor-associated immune cells could define oncogenic or tumor suppressive metabolic molecules. The reciprocal expression between HNSCCs/CAFs and tumor-associated immune cells could be molecules of interest. CAFs: Cancer associated fibroblasts; TILs: Tumor infiltrating lymphocytes; and TAMs: Tumor associated macrophages.
Figure 2. Exploring TME-associated tumor activators/inhibitors in HNSCCs. A comparison between the metabolic profiles of sorted tumor cells, CAFs and tumor-associated immune cells could define oncogenic or tumor suppressive metabolic molecules. The reciprocal expression between HNSCCs/CAFs and tumor-associated immune cells could be molecules of interest. CAFs: Cancer associated fibroblasts; TILs: Tumor infiltrating lymphocytes; and TAMs: Tumor associated macrophages.
Ijms 20 03960 g002
Figure 3. The hypothetic experimental design for novel metabolism-mediated anti-cancer drug screening. HNSCC and CAF cells containing reporters driven under a cancer favorable gene promoter (A promoter) and cancer unfavorable gene promoter (B promoter) could be established, while the same plasmid cassettes could also be introduced into tumor-associated immune cells. The system could be then utilized to screen potential cancer inhibitors via the detection of reporter alterations. CAFs: Cancer associated fibroblasts; TILs: Tumor infiltrating lymphocytes; and TAMs: Tumor associated macrophages.
Figure 3. The hypothetic experimental design for novel metabolism-mediated anti-cancer drug screening. HNSCC and CAF cells containing reporters driven under a cancer favorable gene promoter (A promoter) and cancer unfavorable gene promoter (B promoter) could be established, while the same plasmid cassettes could also be introduced into tumor-associated immune cells. The system could be then utilized to screen potential cancer inhibitors via the detection of reporter alterations. CAFs: Cancer associated fibroblasts; TILs: Tumor infiltrating lymphocytes; and TAMs: Tumor associated macrophages.
Ijms 20 03960 g003
Table 1. Summary of the ncRNA-mediated regulations for the HNSCC metabolism.
Table 1. Summary of the ncRNA-mediated regulations for the HNSCC metabolism.
Cancer TypencRNARegulating Target/RoleMolecular AlterationsReference
Bladder cancerLncUCA1glucose metabolismhexokinase 2 (HK2)↑, mTOR-STAT3↑[151]
Hepatocellular carcinomaLncHULClipid metabolismACSL1↑[152]
Colon cancerLncRNA LINC01234amino acid metabolismmiR-642a↓[153]
Head and neck cancerLncCAF (FLJ22447)oncogenicIL-33↑[155]
Head and neck cancerLncZFAS1tumor suppressiveEGFR↓, PD-L1↓[156]
Head and neck cancerLncHIFCAR (MIR31HG)glucose metabolismHIF-1↑[157]
Head and neck cancerncHAS2-AS1hypoxia [158]
Larynx carcinomaLncPCAT19mitochondriamiR-182↑[156]
Head and neck cancermiR-34atumor suppressiveVEGF↓, AXL↓[160,161]
Acute myeloid leukemiamiR-34a PD-L1↓[162]
Pancreatic ductal adenocarcinomamiR-34a [163]
Oropharyngeal squamous cell carcinomamiR-210 ISCU↓[168]
Head and neck cancermiR-340glucose metabolismGLUT1↑[169]
Head and neck cancermiR-21, LncRNAs HOXA11-AS, LINC00964 and MALAT-1plasma biomarker [173]
Head and neck cancermiR-146aplasma biomarker [174]
Head and neck cancerMALAT-1, HOTAIR, NEAT-1, HULC, MEG-3 and UCA1oncogenic [175]
Head and neck cancermiR-125a and miR-200asalivary marker [176]
Esophageal cancermiR-21salivary marker [177,178]
Head and neck cancermiR-31plasma and salivary biomarker [179,180]

Share and Cite

MDPI and ACS Style

Hsieh, Y.-T.; Chen, Y.-F.; Lin, S.-C.; Chang, K.-W.; Li, W.-C. Targeting Cellular Metabolism Modulates Head and Neck Oncogenesis. Int. J. Mol. Sci. 2019, 20, 3960. https://doi.org/10.3390/ijms20163960

AMA Style

Hsieh Y-T, Chen Y-F, Lin S-C, Chang K-W, Li W-C. Targeting Cellular Metabolism Modulates Head and Neck Oncogenesis. International Journal of Molecular Sciences. 2019; 20(16):3960. https://doi.org/10.3390/ijms20163960

Chicago/Turabian Style

Hsieh, Yi-Ta, Yi-Fen Chen, Shu-Chun Lin, Kuo-Wei Chang, and Wan-Chun Li. 2019. "Targeting Cellular Metabolism Modulates Head and Neck Oncogenesis" International Journal of Molecular Sciences 20, no. 16: 3960. https://doi.org/10.3390/ijms20163960

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