Skip to main content
Intended for healthcare professionals
Open access
Review article
First published online October 7, 2020

The emerging role of targeting cancer metabolism for cancer therapy

Abstract

Glucose, as the main consuming nutrient of the body, faces different destinies in cancer cells. Glycolysis, oxidative phosphorylation, and pentose phosphate pathways produce different glucose-derived metabolites and thus affect cells’ bioenergetics differently. Tumor cells’ dependency to aerobic glycolysis and other cancer-specific metabolism changes are known as the cancer hallmarks, distinct cancer cells from normal cells. Therefore, these tumor-specific characteristics receive the limelight as targets for cancer therapy. Glutamine, serine, and fatty acid oxidation together with 5-lipoxygenase are main pathways that have attracted lots of attention for cancer therapy. In this review, we not only discuss different tumor metabolism aspects but also discuss the metabolism roles in the promotion of cancer cells at different stages and their difference with normal cells. Besides, we dissect the inhibitors potential in blocking the main metabolic pathways to introduce the effective and non-effective inhibitors in the field.

Introduction

The increasing incidence in the number of cancer patients worldwide clearly calls for a more comprehensive investigation on this threatening dilemma.1 Accordingly, numerous studies in genetics, molecular biology, and cancer metabolism were performed. Cancer metabolism was first described by Otto Warburg in 1920s.2,3 In cancer, the main pathway for glucose consumption is the transformation to lactate, to produce adenosine triphosphate (ATP) which is a much faster process than tricarboxylic acid cycle (TCA cycle) occurring in mitochondria. However, this pathway has low output which results in more glucose consumption and several other metabolites.4,5 This cancer hallmark is an escape way for cancerous cells to bypass immune attack and thus proliferates and maintains malignancy.6,7 Furthermore, higher oxidative phosphorylation in other cancers, such as lymphoma and leukemia, generating ATP through electron transport chain (ETC) occurring in inner membrane of mitochondria, is another metabolic pathway.8
Cancerous cells use pentose phosphate pathway (PPP) according to metabolic demand providing a reduced form of nicotinamide adenine dinucleotide phosphate (NADPH), electron donor in reductive reactions, and essential components for cancer cells under limited situations.9 Glutamine, as glutathione, carbon, and nitrogen source in cancer metabolism, is of importance in this case.10 Beside aforementioned pathways, we will discuss other metabolic pathways including 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase 3 (PFKFB3), serine, fatty acid oxidation (FAO), 5-lipoxygenase (5-LOX), forkhead box O (FOXO), and pyruvate dehydrogenase kinase (PDK) metabolism in cancer metabolism. Hence, we aim to discuss the beneficial metabolic changes in cancers by focusing on a more selective approach as a valuable therapeutic target for treating cancers with minimal side effects for normal cells.

Glycolysis

Normal cells use the pyruvate derived from glucose metabolism in the TCA cycle to produce 30–32 ATP molecules per each glucose molecule. This is while cancer cells devise a faster pathway called aerobic glycolysis that occurs in the presence of plentiful oxygen molecules and converting one glucose molecule to two lactate molecules and producing only two ATP molecules. Lactate production is vital for cancer cells survival. Ascribed to low yield of ATP production in aerobic glycolysis, in a compensatory mechanism, glucose uptake in cancer cells will increase. Alteration in metabolism in cancer cell can be the result of impairment in function of mitochondria which has several mutations in related genes such as chain termination mutation in a de novo cytochrome C oxidase subunit 1 (CO1) which is under positive selection in cancer cells.11,12 Since glycolytic pathway performs as an intermediate and signaling network, its targeting can serve as a promising therapeutic candidate.13 Enzymes and transporters in glycolysis are attractive targets, which are the most important ones, as shown in Figure 1. Among the target enzymes in the glycolytic pathway, hexokinase with four isoforms (HK-I to HK-IV) could be the first choice. HK converts glucose to glucose-6-phosphate (G-6-P) (which is rate limiting in glycolysis) by phosphorylation.13 HK-II overexpression has been shown in cancer cells, in contrast to normal cells.14 2-deoxyglucose (2-DG), a glucose analog, is an important inhibitor of HK-II (Table 1). 2-deoxyglucose-6-phosphate (2-DG6P) is the result of 2-DG phosphorylation by HK without being metabolized by cells to progress the glycolysis pathway, resulting in 2-DG6P accumulation and inhibition of HK-II.5 Loss of the intracellular glucose leads to cancer cell death, since it results in ATP depletion. The inhibitory effect of 2-DG is dose-dependent as high dose may be toxic and low dose is insufficient to work.15,16 Despite 2-DG efficiency through glycolysis inhibition in cancer cells, it has toxicity side effects to normal cells, skeletal muscles, and neuronal cells, which also apply glucose for energy. Also, it is striking that 2-DG impairs T cells’ metabolism, which results in reduced cytokines secretion and diminished antitumor function of T cell that can be critical for therapeutic achievement.17,18 According to the toxicity effect of 2-DG in metabolism of normal cells,17 such as T cells,18 which is in paradox with its anticancer properties, it has been demonstrated that combinational application of 2-DG with other treatment could be more successful specially in preclinical phase.19 Also, the therapeutic effect of other inhibitors of HK-II, such as 3-bromopyruvate (3BP), which works through overcoming the chemoresistance and lonidamine that induces apoptosis should attract more attentions. In order to ascertain prognostic value of HK-II, a meta-analysis investigates the suggestion that HK-II can be considered as an independent prognostic factor for patients with solid tumors. They demonstrated that there is a significant association between HK-II expression with shorter overall survival (OS) in hepatocellular carcinoma (HCC), gastric cancer, and colorectal cancer (CRC) as well as progression-free survival (PFS) in patients with solid tumor, although the correlation changes according to cancer type.20 In addition, to assess the prognostic significance of HK-II expression in HCC patients, the survival time of a liver cancer patient cohort from Gene Expression Omnibus (GEO) database and The Cancer Genome Atlas (TCGA) database was analyzed. The results showed a significant association between HK-II expression and OS time in HCC patients. Therefore, elevated HK-II expression in HCC is relevant to its poor clinical prognosis.21
Figure 1. Glycolytic pathway and its inhibitors.
Cells uptakes glucose and makes flow of reactions to transform glucose to pyruvate which has two destinies—enter to TCA cycle or convert to lactate. Here, we can see the inhibitors of hexokinase (HK), glucose transporters (GLUT), and monocarboxylate transporter (MCT).
G-6-P: glucose-6-phosphate; F-6-P: fructose-6-phosphate; G-3-P: glyceraldehyde-3-phosphate; PEP: phosphoenolpyruvate; PFK: phosphofructokinase; 2-DG: 2-deoxyglucose; 3-BP: 3-bromopyruvate.
Table 1. Targets and inhibitors in cancer metabolism pathways.
Agent Target Activity Tumor type Stage
2-deoxyglucose (2-DG)
3-bromopyruvate (3BP)
Lonidamine
Hexokinase-II (HK-II) ATP depletion
Chemoresistance overcoming
Apoptosis inducing
Lung, breast, and prostate cancer Preclinical and clinical studies
Silibinin
Cytochalasin B
GLUTs Anticancer activity Breast and lung cancer Preclinical studies
AZD3965
AR-C155858
MCT-1 and MCT-2 Cytosol acidifying
Diminish solid tumor growth
Advanced solid tumor
Diffuse large B cell lymphomas
Phase I clinical trials
3-PO
PFK-15
PFK-158
PFKFB3 Apoptosis induction
Growth inhibition
Cell cycle arrest
Breast cancer
HCC
AML
Colon and pancreatic adenocarcinomas
Preclinical and clinical studies
Metformin
Phenformin
Complex I (OXPHOS) Anticancer activity
Tumor growth inhibition
Hypoxia reduction
Pancreatic, endometrial, and colon cancers
CRC and NSCLC
Preclinical and clinical studies
Lonidamine Complex II (OXPHOS) Tumor growth delay Breast cancer Phase III
Atovaquone Complex III (OXPHOS) Oxygen consumption rate alleviation
Improves IR response
NSCLC Phase 0
Arsenic trioxide
Hydrocortisone
Nitric oxide
Complex IV (OXPHOS) Reducing tumor hypoxia
Improving radiation sensitivity in tumors
APL and NSCLC Preclinical and clinical studies
Dichloroacetate
Mitaplatin and cisplatin (DCA derivatives)
PDK Cellular metabolism remodeling
Mitochondrial function reactivates
Mitochondrial defects in cancer cells
Breast and NSCLC Phase I
Dehydroepiandrosterone (DHEA) G6PD In vivo G6PD inhibitor    
Polydatin G6PD ROS accumulation apoptosis
Limiting tumor growth
Oral cancer Phase II clinical trial
6-aminonicotinamide (6-AN) 6PGD Limiting ATP production
Multidrug resistance inhibition in doxorubicin-resistant increased radiosensitivity
Human colon cancer HT29-DX cell line (in combination with DHEA)
Human gliomas and squamous carcinoma cell lines (in combination with 2-DG)
 
BPTES
CB-839
Glutaminase (GLS1) TCA anaplerosis inhibition
Anticancer activity
Survival GEMM liver cancer, kidney, hematologic tumors, and solid tumors Preclinical and clinical studies
Methotrexate Dihydrofolate reductase Anticancer activity Lymphomas and leukemias, bladder cancer, and gestational trophoblastic tumors Clinical practice
5-FU Thymidylate synthase Anticancer activity Gastrointestinal malignancies Clinical practice
Etomoxir CPT1 Increasing ROS and decreasing NADPH production
Cancer cell growth, survival, and tumorigenicity inhibition
Leukemia, breast, and prostate cancer cell lines Retired from phase II clinical trial for diabetes and heart failures
Perhexiline CPT1 Approved for use as an antiangina therapy Prostate cancer cell lines
Primary CLL cells
Experimental
ST1326 (Teglicar) CPT1 Growth inhibition and cytotoxicity in cell lines Leukemia and lymphoma cell line Experimental
NDGA 5-LOX Apoptosis induction
Cell proliferating inhibition
Prostate cancer, brain tumor, CNS tumors, and cervical intraepithelial neoplasia Phase II
Zileuton 5-LOX and LTB4 synthesis Remedying bronchial asthma
Reduce esophageal adenocarcinogenesis
Lung cancer and head-and-neck cancer Phase II
Docebenone (AA861) 5-LOX Blocking proliferation
Diminishing metastasis
Inducing apoptosis
Reducing cancer growth
DMBA-induced skin tumor
Human leukemia cell lines
Gastric and esophageal bladder cancers
Discontinued phase II (asthma, allergy)
Phase III (rheumatoid arthritis)
ATP: adenosine triphosphate; GLUT: glucose transporter; MCT: monocarboxylate transporter; PFKFB3: 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase 3; 3-PO: 3-(3-pyridinyl)-1-(4-pyridinyl)-2-propen-1-one; HCC: hepatocellular carcinoma; AML: acute myeloid leukemia; OXPHOS: oxidative phosphorylation; NSCLC: non-small cell lung cancer; CRC: colorectal cancer; IR: ionizing radiation; APL: acute promyelocytic leukemia; PDK: pyruvate dehydrogenase kinase; G6PD: glucose-6-phosphate dehydrogenase; ROS: reactive oxygen species; 6PGD: 6-phosphogluconate dehydrogenase; TCA: tricarboxylic acid; GEMM: granulocyte, erythrocyte, monocyte, and megakaryocyte; 5-FU: fluorouracil; CLL: chronic lymphocytic leukemia; CPT: carnitine palmitoyl transferase; NDGA: nordihydroguaiaretic acid; 5-LOX: 5-lipoxygenase; CNS: central nervous system; LTB4: leukotriene B4; DMBA: 7,12-dimethylbenz[a]anthracene.
In addition to the enzymes, transporters can be the other suitable potential targets for cancer therapy. Glucose transporters (GLUTs) as it is evident from their names are vital for cancer cells, thus interruption in their normal biological activity can decrease the source of nutrients for the cells. It has been shown that GLUT-1 inhibition induces apoptosis in breast and lung cancers22 and diminishes tumor growth. Also, observed GLUT-3 overexpression in some cancers makes these transporters as important targets for cancer therapy. Silibinin and Cytochalasin B are two effective GLUT inhibitors with less side effects on normal cells.22,23
Evaluating the correlation between transcription rates of GLUT-1, GLUT-3, and GLUT-4 and prognostic clinical data in breast cancer patients from the ONCOMINE database demonstrated the elevated transcription of GLUT-1 and lower rate transcription of GLUT-3 and GLUT-4 in breast cancer patient. Besides, the transcription levels of GLUT-1 and GLUT-3 were associated with clinical cancer stage. While GLUT-3 overexpression, introduced as a potential marker of poor prognosis in oral squamous cell carcinoma and non-small cell lung cancer (NSCLC) alleviated transcription level of GLUT-3 in breast cancer associated with a worse OS and clinical cancer stage. Hence, elevated transcription of GLUT-3/4 can suggest as a potential prognostic marker (better prognosis) for breast cancer patients.24 In addition, meta-analysis illustrates the association between GLUT-1 overexpression and poor prognosis in breast cancer, notably these results are more reliable for Asian patients who were mostly included in the studies.25 As such, GLUT-1 expression is associated with poor prognosis as well as being metabolic biomarker for the Lauren classification in locally advanced gastric cancer.26 Monocarboxylate transporters (MCTs) including MCT-1, 2, and 4, facilitate the export of lactate, the substrate for oxygenated cancer cells, across the membrane.14,27 Increased MCT-4 expression associated with high risk of mortality as well as advanced stage and histological grade in patients with breast cancer. As such, elevated expression of MCT-4 is correlated to short OS, which may suggest poor prognosis in breast cancer consistent with data in head-and-neck squamous cell carcinoma.28 AR-C155858 and AZD3965 can be introduced as the lactate efflux inhibitors which make cytosol to be acidified and non-tolerable environment for proliferation.29 AZD3965 blocks MCT-1 and MCT-2 which inhibits solid tumor growth and large B cell lymphomas; however, its side effect on normal T cell has been shown.30 Lenalidomide is known as MCT inhibitor that has a role in impairing the CD147–MCT-1 ligation. Also, it has been shown that it enhances secretion of interleukin-2 (IL-2) and interferon gamma (IFN-γ) in T cells, which could repress tumor cell proliferation while supporting T cells activation.18

PFKFB3 in glycolysis

A critical enzyme in the glycolysis pathway is phosphofructokinase (PFK) that makes irreversible phosphorylation in fructose-6-phosphate (F-6-P). This enzyme is divided to two types such as PFK-1—its overexpression has been found in transformed cells and different primary cancers31 and PFK-2—expressed in solid tumor such as breast and colon cancer.32 PFKFB family with four isoforms (PFKFB 1–4) have crucial role in glycolysis regulation as they have two distinct functions of kinase-phosphatase activities. These bifunctional enzymes synthesize fructose-2,6-bisphosphate (F2,6BP) by kinase activity of 6-phosphofructo-2-kinase in N-terminal region and hydrolytically degrade the F2,6BP by the activity of fructose-2,6-bisphosphatase in C-terminal.33 Different properties, such as kinase/phosphatase ratio, have been demonstrated among four isoforms in spite of their 85% sequence homology.34 PFK-1 can draw back the deterrence effect of intracellular inhibitor such as citric acid and being activated by F2,6BP activity which itself is under control of PFKFB3 with high kinase activity causing the increase in PFK-1 activity, F1,6BP synthesizing, and finally increasing the rate of glycolysis.35 PFKFB3 gene is located on chromosome 10 at position 10p15.1, with 19 exons contains constant and variable regions finally producing six isoforms of PFKFB3.34 PFKFB3 has two distinct roles, such as tumor cell proliferation by strong kinase activity increasing the rate of glycolysis and cell cycle transition by phosphorylation of p27, G1/S suppressor, due to F2,6BP activity and decline in p27 level leading to cell cycle progression.36,37 PFKFB3 gene expression induction happens by some stimulants such as progestin, estrogen, hypoxia, and stress, based on their interaction by receptors which consensus respond elements exist in pfkfb3 promoter.35 Proliferation, survival, and migration effect of increasing activity of PFKFB3 has been shown in different cancer cell such as breast, colorectal, gastric, and leukemia.38,39 Critical roles of PFKFB3 in tumorigenesis make it a potential target in cancer therapy, as inhibitory effect on PFKFB3 in gynecologic cancer cells induces the apoptosis behavior.40 Analyzing PFKFB3 expression level demonstrates more frequent high expression of PFKFB3 in distant metastasis. Consistently, according to Kaplan–Meier analysis, elevated PFKFB3 expression correlated with poor OS in breast cancer patients. Therefore, since PFKFB3 showed determined characteristics of tumor cell metabolism as well as involving in poor prognosis in breast cancer patients, it can serve as a potential target in tumor therapies.41
3-(3-pyridinyl)-1-(4-pyridinyl)-2-propen-1-one (3-PO) introduced as the PFKFB3 inhibitor via diminishing F2,6BP and binding to the function site of PFKFB3.42 Breast cancer,43 HCC, and several adenocarcinomas42 took the advantage of inhibitory effects such as apoptosis induction and growth inhibition of 3-PO. However, due to low water solubility of this compound, the clinical application is not possible.34 PFK-15 derived from 3-PO has a potential role in cell cycle arrest, apoptosis induction, and a selective activity against PFKFB3.44 PFK-15 and PFK-158 are considered as two applicable compounds for cancer therapy in clinical phase.34 PFK-158 as a new inhibitor of PFKFB3 has decreasing effect on glucose uptake and energy production, and more effectiveness activity of this inhibitor was shown while accompanied with standard chemotherapies.40 Also, 5-triazolo-2-arylpyridazinone,45 1-(3-pyridinyl)-3-(2-quinolinyl)-2-propen-1-one (PQP),46 5,6,7,8 tetrahydroxy-2-(4-hydroxyphenyl)chrome-4-one (N4A), and 7,8-dihydroxy-3-(4 hydroxyphenyl) chromen-4-one (YN1)47 can be introduced as the other PFKFB3 inhibitors.

Oxidative phosphorylation

Oxidative phosphorylation (OXPHOS) is a process generally observed in a downregulated manner in cancers, such as breast and gastric cancers, in which glycolysis beats the first word. Alleviation in OXPHOS activities can be ascribed to mutations in mitochondrial DNA and its DNA content reduction.48 However, recently studies showed the OXPHOS upregulation in some cancers such as lymphoma, leukemia, endometrial carcinoma, pancreatic ductal adenocarcinoma, and a subtype of melanoma.49,50 Thus, OXPHOS targeting seems to be attractive for cancer therapy. OXPHOS produces ATP through electron transport among protein complexes (I–IV) in the inner membrane of mitochondria called ETC where the oxygen is the final acceptor51,52 Moreover, inhibition of proteins in ETC complexes, especially complex I, is considered as attractive therapeutic targets. Inhibitors that generate high amounts of reactive oxygen species (ROS) will disrupt ETC complexes.53 Metformin is one of the important tumor growth inhibitors in various types of tumors54,55 with radiation response improvement effects and hypoxia reduction.56 Metformin primarily has been reported as a selectively inhibitor of the mitochondrial respiratory-chain complex 1 which causes reducing nicotinamide adenine dinucleotide-reduced form (NADH) oxidation, diminishing proton gradient across the inner mitochondrial membrane, and decreasing oxygen consumption rate. Considering, metformin has a crucial role in antitumorigenic situations. Indeed, the complex 1 inhibition was reported in several cancer cells that may result in mitochondrial OXPHOS reduction and ATP depletion, eventually lead to adenosine monophosphate–activated protein kinase (AMPK)-mediated activation of catabolic pathways and anabolic proceeding inhibition via its regulation of mechanistic target of rapamycin complex 1 (mTORC1). As a result of co-existing of some AMPK- and mTORC1-independent mechanisms, this metabolic reprogramming reduce growth and proliferation of cancer cells, promotion of cell cycle arrest, and apoptosis in cells that cannot stand the energetic stress.57 Metformin may express antitumor effect via alleviation of insulin-like receptor tyrosine kinases activity.58 Metformin in suitable concentration can arrest complex I in cancer cells.59 Low concentration of metformin in cancer cell will be inefficient; hence, phenformin with low dose can compensate the inefficient metformin.60,61 Besides, the outstanding result of combinational therapeutic effect of metformin and 2-DG in pancreatic cancer by depleting ATP suggests that the blocking of both glycolysis and OXPHOS pathways presents a promising cancer therapy approach.62 It is noteworthy that cancer cell’s mitochondria membranes are more negatively charged compared to the normal cells; therefore, cationic drugs showed to be more selective in targeting tumor cells. As an example, mitochondria-targeted metformin (Mito-metformin), a product of metformin and triphenyl phosphonium (TPP+) group conjunction, has been more effective than untargeted counterparts.63 Oxymatrine, myricetin, capsaicin, and casticin have antitumor role by generating high level of ROS damaging mitochondrial transmembrane in cancer cells.64 Also, tamoxifen, α-tocopheryl succinate (α-TOS), and 3BP target the complexes of ETC in terms of destroying cancer cell via ROS inducing.53 Carboxyamidotriazole (CAI), ME344, and fenofibrate are the other inhibitors targeting complex I.65,66 Complex II disruption by lonidamine effect leading to delaying in growth in many tumors.67 Atovaquone acts as a complex III inhibitor in cancer cells, causing alleviation in oxygen consumption rate.68 Herein, complex IV can be the goal of action of arsenic trioxide, hydrocortisone, and nitric oxide via reducing tumor hypoxia and improving radiation sensitivity in tumors.69,70

PDK

PDKs are important enzymes in oxidative phosphorylation of mitochondria acting as the inhibitor of pyruvate dehydrogenase complex (PDC). PDC is the complex settled in the inner membrane of mitochondria with the interconnection ability of glycolysis and OXPHOS by oxidative decarboxylation of pyruvate to acetyl-CoA.1,71 Decreasing in the amount of acetyl-CoA, the prominent substrate of Krebs cycle, is the result of the inhibitory effect of these kinases which cause halting mitochondrial respiration; hence, the crucial role of PDK in cancer metabolism is obvious (Figure 2).72 PDC has two major parts such as the enzymatic part—consists of four subunits such as pyruvate dehydrogenases (E1), dihydrolipoamide transacetylase (E2), dihydrolipoamide dehydrogenase (E3), E3 binding protein (E3BP) altering pyruvate to acetyl-CoA, CO2, and NADH73 and the regulatory part—consists of subunits such as PDKs and pyruvate dehydrogenase phosphatases (PDPs). PDK has four isoforms (PDK1–PDK4) that almost have tissue-specific functions.74 Three serine residues of E1 α 1 in pyruvate dehydrogenase have phosphorylated by the effect of PDKs in different order depending on tissue. PDK activities have been affected by acetyl-CoA, NADH, and ATP in mitochondria. Different expressions of PDK in various tissues and overexpression of PDK1 in cancer cells have been reported,75 accordingly this kinase could be an attractive target in cancer therapy. Dichloroacetate (DCA), known as the primary inhibitor of PDKs, besides its toxic effect in high dosage, has the anticancer characteristic like remodeling cellular metabolism and reactivating mitochondrial function as well.76 In addition, some agents, such as mitaplatin and cisplatin with various anticancer potentials, as the components derived from DCA, come to work.77,78
Figure 2. Inhibitory effect of dicholoroacetate (DCA) and its derivatives on pyruvate dehydrogenase kinase (PDK).
PDC: pyruvate dehydrogenase complex.

PPP

The PPP is another way for glucose consumption through G-6-P, the result of glucose phosphorylation by HK. This pathway is one of the major sources of NADPH, the molecule which is necessary to neutralize produced ROS in the ATP production process in the cells.79 The other product of PPP is ribulose-5-phosphate (Ru-5-P) which is generated in oxidative phase and can change into ribose-5-phosphate (R-5-P) as a nucleotide precursor. F-6-P and glyceraldehyde-3-phosphate (G-3-P), as two products resulting from non-oxidative branch of PPP, can be used in the glycolysis (Figure 3).80,81 Glucose-6-phosphate dehydrogenase (G6PD) is the first and key enzyme in oxidative phase of PPP that has more expression in cancer cells compared to normal cells.9 Overexpression of G6PD has been observed in human cervical and esophageal carcinomas, hepatomas, and lung and colon adenocarcinomas.82 There are a growing body of evidences showing that inhibition of G6PD reduces proliferation and increases susceptibility to chemotherapy in cancer cells; hence, it can be a powerful target in cancer therapy.83,84 Although this pathway is a promising therapeutic target, its effective inhibitors are limited.54 Assessment of G6PD levels in nasopharyngeal cancer male patients showed that low rates of G6PD activity have significant correlation with tumor recurrence, although it has no association with tumor stage or lymph node or distant metastasis. According to these findings, there is an association between low activity of G6PD with poor prognosis in nasopharyngeal cancer patients.85 Dehydroepiandrosterone (DHEA) was introduced as the in vivo G6PD inhibitor; however, its effect is unclear because it is transformed to steroid hormones immediately.86 Polydatin (3,4′,5-trihydroxystilbene-3-β-d-glucoside) is a natural molecule that has recently been used as the G6PD inhibitor which induces accumulation of ROS and apoptosis in carcinoma cells and limiting tumor growth.87,88 The 6-aminonicotineamide (6-AN) is the other PPP inhibitor at 6PGD level by limiting ATP production.89 Antitumorigenic effect of this inhibitor was reported in some experimental tumors.54 Also, multidrug resistance inhibition in doxorubicin-resistant human colon cancer HT29-DX cell line was reported via DHEA and 6-AN.90 However, combination of 2-DG and 6-AN increased radiosensitivity in human gliomas and squamous carcinoma cell lines, showing the efficiency of combinational targeting of glycolysis and PPP.91 Interestingly, G6PD deficiency causes disorders such as mild anemia; hence, it brings this question to our mind if G6PD deficiency help to overcome cancer development.92
Figure 3. Schematic representation of pentose phosphate pathway (PPP). PPP branches from G-6-P, the first metabolite of glycolysis, which finally terminate to produce nucleotide, NADPH, F-6-P, and G-3-P. Here, it shows the inhibitory effect of Polydatin on G6PD and 6-AN on 6PGD.
G-6-P: glucose-6-phosphate; F-6-P: fructose-6-phosphate; G-3-P: glyceraldehyde-3-phosphate; G6PD: glucose-6-phosphate dehydrogenase; 6PGD: 6-phosphogluconate dehydrogenase; R-5-P: ribose-5-phosphate.

Glutamine

Glutamine is the second nutrition source for growth and proliferation of cancer cells. Glutamine is the source of nitrogen and carbon used for building purine and pyrimidine nucleotides, nonessential amino acids, and glucosamine-6-phosphate.93,94 In addition, glutamine is considered as a fuel for TCA cycle via α-ketoglutarate (α-KG) resulting in ATP production.95 Besides, glutamine reduces oxidative stress by generating NADPH and biosynthesis of glutathione,96 and it has a key role in cellular antioxidative mechanisms.97 Glutamine can also regulate energy generation, redox homeostasis, and intracellular signaling.98 The glutamine entered into the cell has different fates as follows: (1) It can be used for nucleotide synthesis by the effect of CAD (carbamoyl-phosphate synthetase 2, aspartate transcarbamylase, and dihydroorotase) or (2) it can be exported outside to import leucine into the cell to be used as coactivator of glutamate dehydrogenase or (3) it can be converted to glutamate and ammonium by glutaminase (GLS) effect. Then, this glutamate changes to α-KG, the TCA cycle intermediate, via glutamate dehydrogenase.99 Glutamine transporters, which are known as SLC1A5/ASCT2 and are involved in glutamine supplying, are upregulated in breast cancer.100 Indeed, “glutamine addicted” is the characteristic of some cancers that they cannot sustain in the absence of glutamine source.101 Therefore, glutamine and the enzymes involved in its pathway can be targeted in cancer therapies. GLS, the first enzyme in glutaminolysis, has three isoforms including GLS1, the splice variant of GLS1 and GLS2.102 GLS1 has a critical role in the various types of tumor cells;103 therefore, it is the attractive part of targeting, which has an inhibitor named Bis-2-(5-phenylacetamido-1,3,4-thiadiazol-2-yl) ethyl sulfide (BPTES) reducing cancer cell proliferation in vitro.104,105 In fact, BPTES can inhibit enzymatic activity of GLS by binding it.106 CB-839 is the other robust GLS1 inhibitor that is more effective than BPTES107 and nowadays applied in clinical trials in cancer patients.96 Acivicin, azaserine,108 and 6-diazo-5-oxo-l-norleucine (l-DON) are glutamine mimetic with antitumor activity through inhibitory effect on GLS and biosynthesis of purine; however, their toxic side effect, such as neurotoxicity, was observed in clinical trial (Figure 4).106
Figure 4. Different fates of entered glutamine in cancer cells and inhibitors of glutaminase (GLS).
CAD: carbamoyl-phosphate synthetase 2, aspartate transcarbamylase, and dihydroorotase; BPTES: bis-2-(5-phenylacetamido-1,3,4-thiadiazol-2-yl) ethyl sulfide; l-DONL: 6-diazo-5-oxo-l-norleucine.
Evaluating prognostic value of GLSs by systematic multiomic analysis demonstrates the association between overexpression of kidney-type glutaminase (GLS) in breast, esophagus, head-and-neck, and blood cancers with poor prognosis, as such there was a correlation between liver-type glutaminase (GLS2) overexpression with poor OS in colon, blood, ovarian, and thymoma cancers.109 It is worth to mention that depletion of asparagine and glutamine by L-asparaginase from cancer cell in acute lymphoblastic leukemia (ALL) patients shows promising result in cancer treatment.110 Also, combination of L-asparaginase by other glutamine metabolism inhibitors can increase the effect of therapeutic outcome. Overall, combinational therapy using inhibitors to target glutamine metabolism and other pathways will be more effective, such as combined inhibition of GLS and mTOR kinase causes tumor cell death in tumor-bearing mice.111

Serine

Serine, a nonessential amino acid, is another most used compound by the cancer cell,112,113 which can be provided by intracellular synthesizing from glucose as the major source in cancer cells and also can be imported from extracellular environment.114 Serine, by providing carbon and nitrogen, has an important and direct or indirect role in biosynthesis of molecules, which are crucial for cell building such as sphingolipids, phospholipids, purines, and thymidine.115 Also, it maintains redox homeostasis via the folate cycle.116,117 3-phosphoglycerate (3-PG), one of the glycolytic pathway metabolite, is the precursor of serine which is affected by the phosphoglycerate dehydrogenase (PHGDH). Increasing the role of this key enzyme, either via gene amplification or oncogenic signaling, supports the growing production of serine observed in breast cancer, lung adenocarcinoma, and melanoma.118,119 Accordingly, we can consider the inhibition of PHGDH as a strategy in cancer therapy. NCT-503 and CBR-5884 are considered as two PHGDH inhibitors.120 Serine hydroxymethyltransferases (SHMTs) are the enzymes that control converting serine to glycine and one carbon unit. Adding one carbon to tetrahydrofolate (THF) produces N5,N10-methylenetetrahydrofolate (THF-CH2) that itself has also the potential of being a target in cancer therapy121 and can participate in three pathways such as (1) purine metabolisms by producing formyltetrahydrofolate (CHO-THF), (2) thymidine metabolism in cytosol via donating methyl group, and (3) S-adenosylmethionine (SAM) cycle to produce methionine.120 Methotrexate and fluorouracil (5-FU) can be introduced as the inhibitors of thymidylate synthase and dihydrofolate reductase which are involved in thymidine synthesis.122

FAO and CPT1A roles

Sustainable cell energy homeostasis is somehow depending on the energy yielding and energy consuming processes named as fatty acid metabolism. Fatty acid synthesis (FAS) leads to the main processes including cell membrane integrity, energy saving, and interceding signaling.123 In return, FAO or β-oxidation process has a crucial role in producing energy by cyclical reaction in shortening of fatty acid by two carbons in each cycle until producing two acetyl-CoA molecules.124 Lipid accumulation, which is considered as a determined sign of cancer, occurs as a result of FAS and FAO equilibrium disturbance.125 Major part of FAO occurs in mitochondria, and the generated ATP and NADPH molecules are the result of entering the NADH and FADH2, produced in each turn of FAO, to the ETC.126 NADPH promotes cell proliferation by producing FA and cholesterol.127 FAO commences by generating long-chain acyl-CoA via the enzymatic effect of long-chain acyl-CoA synthetase (LCACS).128 Mitochondrial inner membrane impermeability to long-chain acyl-CoAs makes development of carnitine palmitoyl transferase (CPT) system by three members: CPT1, the carnitine acylcarnitine translocase (CACT), and CPT2.123 Altering acyl-CoAs to acylcarnitine, rate limiting stage of FAO, is catalyzed by CPT1 which has three tissue-specific isoforms (CPT1A, CPT1B, and CPT1C).129,130 Swapping acylcarnitine and carnitine between the mitochondrial membranes is done by CACT activity. CPT2, the matrix located member, is used for changing acylcarnitine to acyl-CoAs for oxidation.126 FAO rate is regulated by multiple mechanisms. Malonyl-CoA, a product of the first FAS step, is considered as a physiological inhibitor of CPT1A (liver form) and CPT1B (muscle form). Despite the defined similarity between these two isoforms, their resistance behavior against malonyl-CoA is different. CPT1B has been known as the more sensitive isoform; hence, CPT1A activity can be shown in high level.131 ATP-citrate lyase (ACLY) with high expression level reported in breast cancer, ovarian cancer, and CRC is the starter enzyme of FAS and has the ability to convert citrate to oxaloacetate and cytosolic acetyl-CoA, the building block for FAS.132 Acetyl-CoA carboxylase (ACC) is the other prominent lipogenic enzyme, which has two isoforms such as ACC1 located in cytoplasm producing malonyl-CoA and ACC2 in mitochondrial membrane which hinders Acyl-CoA importing to mitochondria via CPT1.133 ACC overexpression has been reported in gastric, lung, and breast cancers.134 In metabolic stress, produced ATP during FAO acts as the tumor nutrition and promotes cancer cell proliferation and metastasis.135 Upregulation of lipogenic enzymes, such as fatty acid synthetase, has been observed in breast, ovarian, endometrial, and prostate cancers.136,137 Therefore, targeting these enzymes plays an important role in cancer therapy. Inhibition of CPT1A activity has been reported in several cancers. Etomoxir (ETO), pharmacological inhibitor of CPT1A, will sensitize the leukemic cells to the chemotherapeutic drug, cytarabine.138 Also, enzalutamide can be more effective on depleted CPT1A prostate cancer.139 Abnormal expression of CPT1B has been seen in CRCs and bladder cancers140,141 and causes chemoresistance in breast cancer cell.142 While effectiveness of knock down and depletion of CPT2 were observed in triple-negative breast cancer (TNBC),143 high expression of CPT2 had better effect in CRC patient.144 CPT1C upregulation not only is observed in breast, lung, and brain cancers’ survival145 but also is caused by BCR-ABL positive leukemia cells resistance to imatinib or rapamycin.146 Resistance to cetuximab in head-and-neck cancer and metastatic state in breast and lung cancers have been reported as a result of ACC1 phosphorylation at serine 80.147,148 FAO provides the ATP fuel for growth, survival, and proliferation of cancer cells such as prostate cancer,149 diffuse large B-cell lymphoma, Burkitt’s lymphoma,150 and human glioblastoma.151 Inhibition of FAO can reduce the chemoresistance property of cancer cell which can explain the overcoming of lung adenocarcinoma to its resistance to paclitaxel by FAO inhibition.152 Notable point is the tumor neovascularization that is promoted by the properties of CPT1 in tumor microenvironment.126 ETO, CPT1 inhibitor, leads to human glioblastoma cell death by increasing ROS and decreasing NADPH production.153 However, ST1326 (Teglicar), as the other CPT1 inhibitor derived from amino carnitine, has a defined role in preventing cancer cell growth126 and treatment of leukemia and lymphoma.154 Perhexiline, less selective inhibitor of CPT1, makes metabolic disturbance; however, it has toxic effect in long-term therapy.155 Trimetazidine or ranolazine is another example inhibiting 3-KAT of the trifunctional protein (TFP) (hydroxyacyl-CoA dehydrogenase/enoyl-CoA hydratase/3-KAT).135 Avocatin B, a natural inhibitor extracted from avocado, affects acute myeloid leukemia (AML) cell growth and survival via FAO inhibition.156 In general, as FAS has an important function in cancer cells relative to normal tissues, de novo targeting of this pathway is considered.157 Furthermore, taking up FAs through proteins in membrane can cause cell demand, thus targeting of this pathway seems to have the potential therapeutic effects in cancer therapy.158

5-LOX

High-fat diets will result in overweight and potentially might result in cancers including breast and colon.159 Arachidonic acid, an essential C-20 fatty acid, is the indispensable element of dietary.160 Arachidonic acid generates bioactive lipid signaling molecules named eicosanoids through three major pathways such as cyclooxygenase (COX), LOX, and cytochrome P450.161 The effect of high level of eicosanoids has been reported in different phases of cancer such as promotion, progression, metastasis, and angiogenesis.162 Among the three mentioned enzymes, involvement of LOXs family in cancer has been reported both clinically and experimentally.161 This family even acts as a tumor microenvironment regulator.163 Importantly, there are two opposite activities for 15-LOX-1; promoting activity, reported in prostate cancer, and suppressing activity, reported in CRC and chronic myeloid leukemia (CML).164166 These two edge situations have been reported in breast cancer as well.167 There are enough evidences showing high overexpression of 12-LOX in prostate and breast cancers.168,169 5-LOX activity, which generates 5-hydroxyeicosatetraenoic acid (5-HETE) from arachidonic acid, was reported for the first time in 1976 in rabbit peritoneal neutrophils.170 5-LOX gene containing 14 exons and 13 introns is located on chromosome 10q12.2.171 Generated products of 5-LOX activities are 5-hydroperoxyeicosatetraenoic acid (5-HPETE) which is then reduced to 5-HETE via glutathione peroxidase (GPx) and leukotriene A4 (LTA4), which then converted to leukotriene E4 (LTE4) under sequential reactions by participating enzymes such as hydrolase, synthase, G-glutamyl transferase, and bound membrane dipeptidase.172 Peroxidase also has a crucial effect on 5-LOX product generation.173 5-LOX affects cancer growth and proliferation as its overexpression has been reported in several cancers such as breast, prostate, renal, and colorectal162 and even in atherosclerotic plaque.174 Oxidative stress and ROS have critical effect in expression and function of 5-LOX.175 However, catalytic activity of 5-LOX is owing to active site iron, which gets lost at the presence of oxygen.176 IL-4 leads to downregulation of 5-LOX in dendritic cells.177 Hence, 5-LOX expression is affected by different factors. The produced metabolites of 5-LOX pathway have major role in cell proliferation, viability, and chemoresistance.178 Epidermal growth factor (EGF) and neurotensin result in progression in prostate cancer through 5-LOX.179 Angiogenesis induction was observed in colon cancer by vascular endothelial growth factor (VEGF) expression via 5-HETE.163 In addition, LOX overexpression is related to metastasis180 and invasion in several cancers.181 High levels of 5-LOX and leukotriene B4 (LTB4) have been observed in Kaposi sarcoma lesions and latent endothelial cells.182 VEGF overexpression occurs as a result of 5-LOX, 5-HETE, and LTA4 activation in human malignant mesothelioma in vitro.183 COX-2, the other principal eicosanoid with tumorigenesis effect, along with 5-LOX results in expression and release VEGF.184 Shunting mechanism of these two elements showed that inhibition of COX-2 makes 5-LOX overexpression but vice versa, 5-LOX inhibition has no redundancy effect in COX-2 expression.185 Hence, 5-LOX inhibition is more effective and can be a potential target in cancer therapy. According to Nottingham prognostic index, a tool for speculating the prognosis of the patients, elevated rates of 5-LOX-activating protein transcription have a significant correlation with poor prognosis, which explain the stimulator role of downstream metabolites of 5-LOX in the invasion and angiogenesis in cancer.186 A63162, a selective inhibitor for 5-LOX, limits DNA synthesis and PC3 prostate cancer cell growth.187 Apoptosis induction in prostate cancer cell and cell proliferating inhibition in lung cancer cell are caused by the effect of AA861, MK886, MK591, and nordihydroguaiaretic acid (NDGA) inhibitors.175 Apoptosis and growth limitation have been reported as the inhibitory effects of boswellic acids (BAs) on 5-LOX in brain tumor cell lines. BAs have been determined as active principles of the Boswellia serrata resin extracts, which has inhibition effect on leukotriene biosynthesis in intact polymorphonuclear leukocytes. In vitro, BAs (belonging to the ursane type pentacyclic triterpene saponins) selectively blocked the leukotriene production without modifying the product formation by other dioxygenases. As such, a derivative of BAs, acetyl-11-keto-β-BA (AKBA), showed inhibition effect on 5-LOX product formation. Considering, pentacyclic triterpenes without the 11-keto-function did not express any 5-LOX inhibition, which directly states the role of AKBA on a selective site for pentacyclic triterpenes on the 5-LOX enzyme. In addition, the pentacyclic triterpene ring system is important for binding to the selective effector site, whereas functional groups are necessary for the 5-LOX inhibitory effect of BAs.188 Zileuton, as the other important 5-LOX inhibitor, is approved by United States Food and Drug Administration (US FDA) for remedying bronchial asthma and is able to reduce esophageal adenocarcinogenesis.189 Inhibitory effect of zileuton is caused by chelating iron from active site of 5-LOX. A79175 (ABT-175) and A85761 (ABT-761) have been introduced as two modified compounds of zileuton with more lasting effect in vivo as a result of the loss of glucuronidation rate. In addition, NDGA, Vigna furan, and caffeic acid have been introduced as the natural, specific, and effective inhibitors of 5-LOX.190 There are several interesting studies reporting the effect of Docebenone (AA861) to be useful in cancer therapy. The 5-LOX inhibitor with powerful effect in 7,12-dimethylbenz[a]anthracene (DMBA)-induced skin tumor causes blocking proliferation in human leukemia cell lines,175 diminishing metastasis, and inducing apoptosis in gastric cancer,191 also reducing cancer growth in esophageal and bladder cancers.192,193 BWA4C, another inhibitor for 5-LOX, increases apoptosis in lymphoma and reduces the growth and tumor size in pancreatic cancer cell lines.163 Meanwhile, esculetin194 (Esc), a natural compound 6,7-dihydroxycoumarin, has inhibitory effect on 5-LOX, 12-LOX, and LT synthesis195 in leukemia.196

Forkhead box O1 (FOXO1)

The FOXOs are transcription factors,197 which have the critical regulatory functions in most important cellular processes and metabolism such as cell cycle, proliferation, development, differentiation, apoptosis, autophagy, and tumor suppressing. FOXO belongs to FOX family of transcription factors in which their forkhead domain has been conserved during evolution.198 There are four members of FOXO in vertebrates including FOXO1, FOXO3, FOXO4, and FOXO6, unlike invertebrates such as Drosophila melanogaster which has only one ortholog named dFOXO.199 Structure of FOXO1 or forkhead rhabdomyosarcoma transcription factor (FKHR), the key member of FOXO family, has four parts such as a forkhead DNA-binding domain (FHD) in amino terminal, nuclear localization signal (NLS), nuclear export sequence (NES), and a transactivation domain (TAD) in C-terminal.200 Phosphorylation, ubiquitination, acetylation, and deacetylation are posttranslational modifications in this structure affecting FOXO1 activities.201 Moreover, absence or presence of growth factor can regulate FOXO1. In this case, on one hand, FOXO1 activation is caused by insulin binding to its receptor which in turn activates phosphoinositide kinase (PI3K), Akt, and serum glucocorticoid–inducible kinase (PI3K-AKT-SGK); on the other hand, absence of insulin leads to involvement of FOXO1 in cell death.198 Transcriptional activity of FOXO is lost via its phosphorylation in some residues induced by AKT and SGK.202 It should be mentioned that nuclear export of FOXO1 and its phosphorylation manner is performed via insulin-like growth factor-1 (IGF-1)/Akt signaling and 14-3-3 proteins.203 FOXO1, which is mostly expressed in adipose tissue,200 acts as a tumor suppressor that its deletion, inactivation, or downregulation has been observed in several cancers.204 There are evidences demonstrating the roles of FOXO1 in the inhibition of proliferation, differentiation, angiogenesis, and enhancing apoptosis.198 Anticarcinogenesis characteristic of FOXO1 is amplified by PI3K, mitogen-activated protein kinase (MAPK), and IκB kinase (IKK) signaling pathways205 and stress signaling such as oxidative stress.206 Limited proliferation can be achieved by FOXO1 activities; for example, AQP9 overexpression in HCC restricts cell proliferation, which takes place via the inhibition of PI3K/Akt pathway and increase in FOXO1.207 FOXO1 suppression induce overexpression of hypoxia-inducible factor-1-α (HIF-1-α) and angiogenesis in gastric cancer,208 and also endothelial cell migration is limited following the overexpression of FOXO1 and FOXO3.209 Inhibition of Cyclin D1 and D2 by FOXO leads to cell cycle arrest.203 Restraining of PI3K/Akt pathway, which promotes FOXO1, induces apoptosis in pancreatic cancer in response to curcumin application.210,211 Also, apoptosis is induced by ROS-mediated FOXO1 activation.212 Extrinsic apoptosis pathway is activated by FOXO via expression of three ligands including Fas ligand, tumor necrosis factor (TNF)-related apoptosis inducing ligand, and TNF receptor–associated death domain. However, expression of Bim, BNIP3, and Puma, belonging to Bcl-2 family, induces intrinsic pathway through FOXO.203 It is well documented that FOXOs can inhibit glycolysis and suppress the Warburg effect via FOXO-mediated antagonism with Myc (one of the major oncogenes promoting the Warburg effect). As an example, FOXOs inhibit the expression of glycolysis genes in renal cancer cells through Myc inhibition. In addition, inhibiting FOXOs by acetylation via mTORC2 in glioblastoma leads to activation of Myc and Warburg effect upregulation.213215 Also, it is demonstrated that glutamine synthetase expression can be regulated by FOXO3 and FOXO4 in the PI3K-AKT-FOXO pathway. Besides, FOXO activation leads to mTOR inhibition through glutamine synthetase as well as promoting autophagy. In addition, amino acids involve in regulating FOXO including inactivation of FOXO through mTORC2 and AKT activation in mammalian cells. It is worth to mention that, despite the unclear role of FOXO in lipid metabolism, it can cause downregulation of SREBP1c (sterol regulatory element binding proteins) through occupying the SREBP1c promoter.202 Notable point is the conflicting activity of FOXO that may be observed in some ways. In this regard, sustaining cancer stem cells especially in CML and AML has been reported,216 which arising from the increase in FOXO3, and also the expression of MDR1 protein by FOXO1 leads to drug resistance to doxorubicin and adriamycin.203
Nevertheless, according to the functions reported for FOXO in cellular and physiological behaviors, it can be considered as a target in cancers. Paclitaxel, an anticancer drug applied for breast, ovarian, pancreatic, and other cancers, allows FOXO3a to be accumulated. FOXO3a nuclear translocation occurs by the effect of doxorubicin used as a drug in cancer therapy. Also, cisplatin and epirubicin are other drugs that can target this axis.217

Conclusion

Common therapies, such as chemotherapy and hormone therapy, are widely used for cancer treatment; however, side effects of these approaches show the demand for alternative targeted routes. Targeting cancer metabolism with various conversion pathways, enzymes, and metabolites seems to be the way to proceed. Altered metabolic pathways in cancer cells drive cells to enhance proliferation, the event that rarely occurs in differentiated normal cells. Hence, targeting these statues can destroy tumor cells. Alteration in glycolysis, that is, producing lactate instead of delivery of pyruvate to TCA cycle, encourages the tumor cells to majorly rely on aerobic glycolysis. This is a huge finding in cancer cells relative to normal counterparts, although an increase in OXPHOS pathway has been reported in certain cancers. Key role of some enzymes in these pathways has opened a novel window to combat uncontrolled cell proliferation. So, the inhibitory effect of some drugs was examined in the preclinical or clinical phases and promising results have shown the right direction for more progress. However, finding and focusing on the approaches to improve more selective effects are still challenging. Especially focusing on handling immune cells according to their pro- or antitumor capacities makes the story more complex.218 It is worth mentioning that cancer cells have shown varieties of altered metabolism pathways which provide sufficient fuel for their viability. Sieging multilateral using combinational blocking of these pathways is suggested as the more useful approach. Besides, when cancer cells cannot stand the inhibition effect of drugs and become sensitive, normal cells can overcome this situation using flexible characteristics.

Acknowledgments

The authors would like to especially thank Dr Ali Mohsen Rastegar-Pouyani and Dr Hamid Reza Mohammadi Motlagh for their valuable discussions and help with the preparation of this article. The authors also thank the Research Council of Kermanshah University of Medical Sciences (KUMS) for supporting this investigation.

Declaration of conflicting interests

The author(s) declared no potential conflicts of interest with respect to the research, authorship, and/or publication of this article.

Funding

The author(s) received no financial support for the research, authorship, and/or publication of this article.

ORCID iD

References

1. Zhang W, Zhang S-L, Hu X, et al. Targeting tumor metabolism for cancer treatment: is pyruvate dehydrogenase kinases (PDKs) a viable anticancer target. Int J Biol Sci 2015; 11(12): 1390–1400.
2. Hanahan D, Weinberg RA. Hallmarks of cancer: the next generation. Cell 2011; 144: 646–674.
3. DeBerardinis RJ, Chandel NS. Fundamentals of cancer metabolism. Sci Adv 2016; 2(5): e1600200.
4. Chae YC, Kim JH. Cancer stem cell metabolism: target for cancer therapy. BMB Rep 2018; 51: 319.
5. Martinez-Outschoorn UE, Peiris-Pages M, Pestell RG, et al. Cancer metabolism: a therapeutic perspective. Nat Rev Clin Oncol 2017; 14: 11.
6. Deshmukh A, Deshpande K, Arfuso F, et al. Cancer stem cell metabolism: a potential target for cancer therapy. Mol Cancer 2016; 15: 69.
7. Villalba M, Rathore MG, Lopez-Royuela N, et al. From tumor cell metabolism to tumor immune escape. Int J Biochem Cell Biol 2013; 45(1): 106–113.
8. Ashton TM, McKenna WG, Kunz-Schughart LA, et al. Oxidative phosphorylation as an emerging target in cancer therapy. Clin Cancer Res 2018; 24: 2482–2490.
9. Cho ES, Cha YH, Kim HS, et al. The pentose phosphate pathway as a potential target for cancer therapy. Biomol Ther 2018; 26: 29.
10. Nguyen T-L, Durán RV. Glutamine metabolism in cancer therapy. Cancer Drug Resist 2018; 1: 126–138.
11. Wallace DC. Mitochondria and cancer. Nat Rev Cancer 2012; 12: 685–698.
12. de Souza AC, Justo GZ, de Araújo DR, et al. Defining the molecular basis of tumor metabolism: a continuing challenge since Warburg’s discovery. Cell Physiol Biochem 2011; 28(5): 771–792.
13. Gill KS, Fernandes P, O’Donovan TR, et al. Glycolysis inhibition as a cancer treatment and its role in an anti-tumour immune response. Biochim Biophys Acta 2016; 1866(1): 87–105.
14. Wolf A, Agnihotri S, Micallef J, et al. Hexokinase 2 is a key mediator of aerobic glycolysis and promotes tumor growth in human glioblastoma multiforme. J Exp Med 2011; 208: 313–326.
15. Raez LE, Papadopoulos K, Ricart AD, et al. A phase I dose-escalation trial of 2-deoxy-D-glucose alone or combined with docetaxel in patients with advanced solid tumors. Cancer Chemoth Pharm 2013; 71: 523–530.
16. Stein M, Lin H, Jeyamohan C, et al. Targeting tumor metabolism with 2-deoxyglucose in patients with castrate-resistant prostate cancer and advanced malignancies. Prostate 2010; 70: 1388–1394.
17. Kalyanaraman B. Teaching the basics of cancer metabolism: developing antitumor strategies by exploiting the differences between normal and cancer cell metabolism. Redox Biol 2017; 12: 833–842.
18. Kouidhi S, Ben Ayed F, Benammar Elgaaied A. Targeting tumor metabolism: a new challenge to improve immunotherapy. Front Immunol 2018; 9: 353.
19. Vander Heiden MG, DeBerardinis RJ. Understanding the intersections between metabolism and cancer biology. Cell 2017; 168: 657–669.
20. Liu Y, Wu K, Shi Xiang F, et al. Prognostic significance of the metabolic marker hexokinase-2 in various solid tumors: a meta-analysis. PLoS ONE 2016; 11(11): e0166230.
21. Wang L, Yang Q, Peng S, et al. The combination of the glycolysis inhibitor 2-DG and sorafenib can be effective against sorafenib-tolerant persister cancer cells. Onco Targets Ther 2019; 12: 5359–5373.
22. Rastogi S, Banerjee S, Chellappan S, et al. Glut-1 antibodies induce growth arrest and apoptosis in human cancer cell lines. Cancer Lett 2007; 257: 244–251.
23. Kapoor K, Finer-Moore JS, Pedersen BP, et al. Mechanism of inhibition of human glucose transporter GLUT1 is conserved between cytochalasin B and phenylalanine amides. Proc Natl Acad Sci USA 2016; 113: 4711–4716.
24. Zeng K, Ju G, Wang H, et al. GLUT1/3/4 as novel biomarkers for the prognosis of human breast cancer. Transl Cancer Res 2020; 9: 2363–2377.
25. Deng Y, Zou J, Deng T, et al. Clinicopathological and prognostic significance of GLUT1 in breast cancer: a meta-analysis. Medicine 2018; 97(48): e12961.
26. Yin C, Gao B, Yang J, et al. Glucose transporter-1 (GLUT-1) expression is associated with tumor size and poor prognosis in locally advanced gastric cancer. Med Sci Monit 2020; 26: 920778–920771.
27. Cairns R, Harris I, McCracken S, et al. Cancer cell metabolism. Cold Spring Harbor Symp Quant Biol 2011; 76: 299-311.
28. Xiao S, Zhu H, Shi Y, et al. Prognostic and predictive value of monocarboxylate transporter 4 in patients with breast cancer. Oncol Lett 2020; 20(3): 2143–2152.
29. Payen VL, Mina E, Van Hée VF, et al. Monocarboxylate transporters in cancer. Mol Metab 2020; 33: 48.
30. Bola BM, Chadwick AL, Michopoulos F, et al. Inhibition of monocarboxylate transporter-1 (MCT1) by AZD3965 enhances radiosensitivity by reducing lactate transport. Mol Cancer Ther 2014; 13(12): 2805–2816.
31. Vora S, Halper JP, Knowles DM. Alterations in the activity and isozymic profile of human phosphofructokinase during malignant transformation in vivo and in vitro: transformation-and progression-linked discriminants of malignancy. Cancer Res 1985; 45(7): 2993–3001.
32. Atsumi T, Chesney J, Metz C, et al. High expression of inducible 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase (iPFK-2; PFKFB3) in human cancers. Cancer Res 2002; 62: 5881–5887.
33. Yalcin A, Telang S, Clem B, et al. Regulation of glucose metabolism by 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatases in cancer. Exp Mol Pathol 2009; 86(3): 174–179.
34. Shi L, Pan H, Liu Z, et al. Roles of PFKFB3 in cancer. Signal Transduct Target Ther 2017; 2: 17044.
35. Lu L, Chen Y, Zhu Y. The molecular basis of targeting PFKFB3 as a therapeutic strategy against cancer. Oncotarget 2017; 8: 62793.
36. Wang X, Gorospe M, Huang Y, et al. P27 Kip1 overexpression causes apoptotic death of mammalian cells. Oncogene 1997; 15: 2991–2997.
37. Yalcin A, Clem B, Imbert-Fernandez Y, et al. 6-phosphofructo-2-kinase (PFKFB3) promotes cell cycle progression and suppresses apoptosis via Cdk1-mediated phosphorylation of p27. Cell Death Dis 2014; 5: e1337.
38. Minchenko OH, Tsuchihara K, Minchenko DO, et al. Mechanisms of regulation of PFKFB expression in pancreatic and gastric cancer cells. World J Gastroenterol 2014; 20: 13705.
39. Yamamoto T, Takano N, Ishiwata K, et al. Reduced methylation of PFKFB3 in cancer cells shunts glucose towards the pentose phosphate pathway. Nat Commun 2014; 5: 1–16.
40. Mondal S, Roy D, Sarkar Bhattacharya S, et al. Therapeutic targeting of PFKFB3 with a novel glycolytic inhibitor PFK158 promotes lipophagy and chemosensitivity in gynecologic cancers. Int J Cancer 2019; 144: 178–189.
41. Peng F, Li Q, Sun JY, et al. PFKFB3 is involved in breast cancer proliferation, migration, invasion and angiogenesis. Int J Oncol 2018; 52(3): 945–954.
42. Clem B, Telang S, Clem A, et al. Small-molecule inhibition of 6-phosphofructo-2-kinase activity suppresses glycolytic flux and tumor growth. Mol Cancer Ther 2008; 7(1): 110–120.
43. Pisarsky L, Bill R, Fagiani E, et al. Targeting metabolic symbiosis to overcome resistance to anti-angiogenic therapy. Cell Rep 2016; 15: 1161–1174.
44. Zhu W, Ye L, Zhang J, et al. PFK15, a small molecule inhibitor of PFKFB3, induces cell cycle arrest, apoptosis and inhibits invasion in gastric cancer. PLoS ONE 2016; 11(9): e0163768.
45. Brooke DG, van Dam EM, Watts CK, et al. Targeting the Warburg effect in cancer; relationships for 2-arylpyridazinones as inhibitors of the key glycolytic enzyme 6-phosphofructo-2-kinase/2,6-bisphosphatase 3 (PFKFB3). Bioorg Med Chem 2014; 22: 1029–1039.
46. Lea MA, Guzman Y, Desbordes C. Inhibition of growth by combined treatment with inhibitors of lactate dehydrogenase and either phenformin or inhibitors of 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase 3. Anticancer Res 2016; 36(4): 1479–1488.
47. Seo M, Kim J-D, Neau D, et al. Structure-based development of small molecule PFKFB3 inhibitors: a framework for potential cancer therapeutic agents targeting the Warburg effect. PLoS ONE 2011; 6(9): e24179.
48. Yu M. Generation, function and diagnostic value of mitochondrial DNA copy number alterations in human cancers. Life Sci 2011; 89: 65–71.
49. Weinberg SE, Chandel NS. Targeting mitochondria metabolism for cancer therapy. Nat Chem Biol 2015; 11(1): 9–15.
50. Moreno Sánchez R, Rodríguez Enríquez S, Marín Hernández A, et al. Energy metabolism in tumor cells. FEBS J 2007; 274: 1393–1418.
51. Zong W-X, Rabinowitz JD, White E. Mitochondria and cancer. Mol Cell 2016; 61: 667–676.
52. Viale A, Corti D, Draetta GF. Tumors and mitochondrial respiration: a neglected connection. Cancer Res 2015; 75: 3687–3691.
53. Dong L, Neuzil J. Targeting mitochondria as an anticancer strategy. Cancer Commun 2019; 39: 63.
54. Kowalik MA, Columbano A, Perra A. Emerging role of the pentose phosphate pathway in hepatocellular carcinoma. Front Oncol 2017; 7: 87.
55. Lonardo E, Cioffi M, Sancho P, et al. Metformin targets the metabolic achilles heel of human pancreatic cancer stem cells. PLoS ONE 2013; 8(10): e76518.
56. Zannella VE, Dal Pra A, Muaddi H, et al. Reprogramming metabolism with metformin improves tumor oxygenation and radiotherapy response. Clin Cancer Res 2013; 19: 6741–6750.
57. Vial G, Detaille D, Guigas B. Role of mitochondria in the mechanism (s) of action of metformin. Front Endocrinol 2019; 10: 294.
58. Pernicova I, Korbonits M. Metformin—mode of action and clinical implications for diabetes and cancer. Nat Rev Endocrinol 2014; 10(3): 143–156.
59. Wheaton WW, Weinberg SE, Hamanaka RB, et al. Metformin inhibits mitochondrial complex I of cancer cells to reduce tumorigenesis. Elife 2014; 3: e02242.
60. Cheng G, Zielonka J, Ouari O, et al. Mitochondria-targeted analogues of metformin exhibit enhanced antiproliferative and radiosensitizing effects in pancreatic cancer cells. Cancer Res 2016; 76: 3904–3915.
61. Yuan P, Ito K, Perez-Lorenzo R, et al. Phenformin enhances the therapeutic benefit of BRAFV600E inhibition in melanoma. Proc Natl Acad Sci USA 2013; 110: 18226–18231.
62. Kalyanaraman B, Cheng G, Hardy M, et al. A review of the basics of mitochondrial bioenergetics, metabolism, and related signaling pathways in cancer cells: therapeutic targeting of tumor mitochondria with lipophilic cationic compounds. Redox Biol 2018; 14: 316–327.
63. Dickey JS, Gonzalez Y, Aryal B, et al. Mito-tempol and dexrazoxane exhibit cardioprotective and chemotherapeutic effects through specific protein oxidation and autophagy in a syngeneic breast tumor preclinical model. PLoS ONE 2013; 8(8): e70575.
64. Cui Q, Wen S, Huang P. Targeting cancer cell mitochondria as a therapeutic approach: recent updates. Future Med Chem 2017; 9(9): 929–949.
65. Lim SC, Carey KT, McKenzie M. Anti-cancer analogues ME-143 and ME-344 exert toxicity by directly inhibiting mitochondrial NADH: ubiquinone oxidoreductase (complex I). Am J Cancer Res 2015; 5(2): 689–701.
66. Johnson EA, Marks RS, Mandrekar SJ, et al. Phase III randomized, double-blind study of maintenance CAI or placebo in patients with advanced non-small cell lung cancer (NSCLC) after completion of initial therapy (NCCTG 97-24-51). Lung Cancer 2008; 60(2): 200–207.
67. Guo L, Shestov AA, Worth AJ, et al. Inhibition of mitochondrial complex II by the anticancer agent lonidamine. J Biol Chem 2016; 291: 42–57.
68. Fiorillo M, Lamb R, Tanowitz HB, et al. Repurposing atovaquone: targeting mitochondrial complex III and OXPHOS to eradicate cancer stem cells. Oncotarget 2016; 7: 34084.
69. Jordan BF, Sonveaux P. Targeting tumor perfusion and oxygenation to improve the outcome of anticancer therapy. Front Pharmacol 2012; 3: 94.
70. Diepart C, Karroum O, Magat J, et al. Arsenic trioxide treatment decreases the oxygen consumption rate of tumor cells and radiosensitizes solid tumors. Cancer Res 2012; 72: 482–490.
71. Kaplon J, Zheng L, Meissl K, et al. A key role for mitochondrial gatekeeper pyruvate dehydrogenase in oncogene-induced senescence. Nature 2013; 498: 109–112.
72. Kim JW, Tchernyshyov I, Semenza GL, et al. HIF-1-mediated expression of pyruvate dehydrogenase kinase: a metabolic switch required for cellular adaptation to hypoxia. Cell Metab 2006; 3(3): 177–185.
73. Yu X, Hiromasa Y, Tsen H, et al. Structures of the human pyruvate dehydrogenase complex cores: a highly conserved catalytic center with flexible N-terminal domains. Structure 2008; 16(1): 104–114.
74. Rardin MJ, Wiley SE, Naviaux RK, et al. Monitoring phosphorylation of the pyruvate dehydrogenase complex. Anal Biochem 2009; 389: 157–164.
75. Houten S, Chegary M, Te Brinke H, et al. Pyruvate dehydrogenase kinase 4 expression is synergistically induced by AMP-activated protein kinase and fatty acids. Cell Mol Life Sci 2009; 66(7): 1283–1294.
76. Michelakis E, Sutendra G, Dromparis P, et al. Metabolic modulation of glioblastoma with dichloroacetate. Sci Transl Med 2010; 2: 31ra34.
77. Pathak RK, Marrache S, Harn DA, et al. Mito-DCA: a mitochondria targeted molecular scaffold for efficacious delivery of metabolic modulator dichloroacetate. ACS Chem Biol 2014; 9: 1178–1187.
78. Dhar S, Lippard SJ. Mitaplatin, a potent fusion of cisplatin and the orphan drug dichloroacetate. Proc Natl Acad Sci USA 2009; 106: 22199–22204.
79. Ghanbari Movahed Z, Rastegari-Pouyani M, Mohammadi MH, et al. Cancer cells change their glucose metabolism to overcome increased ROS: one step from cancer cell to cancer stem cell. Biomed Pharmacother 2019; 112: 108690.
80. Riganti C, Gazzano E, Polimeni M, et al. The pentose phosphate pathway: an antioxidant defense and a crossroad in tumor cell fate. Free Radic Biol Med 2012; 53: 421–436.
81. Patra KC, Hay N. The pentose phosphate pathway and cancer. Trends Biochem Sci 2014; 39: 347–354.
82. Kuo WY, Lin JY, Tang TK. Human glucose-6-phosphate dehydrogenase (G6PD) gene transforms NIH 3T3 cells and induces tumors in nude mice. Int J Cancer 2000; 85: 857–864.
83. Du W, Jiang P, Mancuso A, et al. TAp73 enhances the pentose phosphate pathway and supports cell proliferation. Nat Cell Biol 2013; 15(8): 991–1000.
84. Catanzaro D, Gaude E, Orso G, et al. Inhibition of glucose-6-phosphate dehydrogenase sensitizes cisplatin-resistant cells to death. Oncotarget 2015; 6: 30102.
85. Cheng AJ, Chiu DT, See LC, et al. Poor prognosis in nasopharyngeal cancer patients with low glucose-6-phosphate-dehydrogenase activity. Jpn J Cancer Res 2001; 92(5): 576–581.
86. Mele L, Paino F, Papaccio F, et al. A new inhibitor of glucose-6-phosphate dehydrogenase blocks pentose phosphate pathway and suppresses malignant proliferation and metastasis in vivo. Cell Death Dis 2018; 9: 1–12.
87. Cai H, Scott E, Kholghi A, et al. Cancer chemoprevention: evidence of a nonlinear dose response for the protective effects of resveratrol in humans and mice. Sci Transl Med 2015; 7: 298ra117.
88. Liu H, Zhao S, Zhang Y, et al. Reactive oxygen species-mediated endoplasmic reticulum stress and mitochondrial dysfunction contribute to polydatin-induced apoptosis in human nasopharyngeal carcinoma CNE cells. J Cell Biochem 2011; 112(12): 3695–3703.
89. Köhler E, Barrach H-J, Neubert D. Inhibition of NADP dependent oxidoreductases by the 6-aminonicotinamide analogue of NADP. FEBS Lett 1970; 6: 225–228.
90. Polimeni M, Voena C, Kopecka J, et al. Modulation of doxorubicin resistance by the glucose-6-phosphate dehydrogenase activity. Biochem J 2011; 439: 141–149.
91. Manganelli G, Masullo U, Passarelli S, et al. Glucose-6-phosphate dehydrogenase deficiency: disadvantages and possible benefits. Cardiovasc Haematol Disord Drug Targets 2013; 13: 73–82.
92. Jiang P, Du W, Wu M. Regulation of the pentose phosphate pathway in cancer. Protein Cell 2014; 5: 592–602.
93. Nicklin P, Bergman P, Zhang B, et al. Bidirectional transport of amino acids regulates mTOR and autophagy. Cell 2009; 136: 521–534.
94. Jahani M, Noroznezhad F, Mansouri K. Arginine: challenges and opportunities of this two-faced molecule in cancer therapy. Biomed Pharmacother 2018; 102: 594–601.
95. DeBerardinis RJ, Mancuso A, Daikhin E, et al. Beyond aerobic glycolysis: transformed cells can engage in glutamine metabolism that exceeds the requirement for protein and nucleotide synthesis. Proc Natl Acad Sci USA 2007; 104: 19345–19350.
96. Jiang J, Srivastava S, Zhang J. Starve cancer cells of glutamine: break the spell or make a hungry monster? Cancers 2019; 11: 804.
97. Xiang L, Xie G, Liu C, et al. Knock-down of glutaminase 2 expression decreases glutathione, NADH, and sensitizes cervical cancer to ionizing radiation. Biochim Biophys Acta 2013; 1833(12): 2996–3005.
98. Cluntun AA, Lukey MJ, Cerione RA, et al. Glutamine metabolism in cancer: understanding the heterogeneity. Trends Cancer 2017; 3(3): 169–180.
99. Icard P, Poulain L, Lincet H. Understanding the central role of citrate in the metabolism of cancer cells. Biochim Biophys Acta 2012; 1825(1): 111–116.
100. Van Geldermalsen M, Wang Q, Nagarajah R, et al. ASCT2/SLC1A5 controls glutamine uptake and tumour growth in triple-negative basal-like breast cancer. Oncogene 2016; 35: 3201–3208.
101. Wise DR, DeBerardinis RJ, Mancuso A, et al. Myc regulates a transcriptional program that stimulates mitochondrial glutaminolysis and leads to glutamine addiction. Proc Natl Acad Sci USA 2008; 105: 18782–18787.
102. Thangavelu K, Chong QY, Low BC, et al. Structural basis for the active site inhibition mechanism of human kidney-type glutaminase (KGA). Sci Rep 2014; 4: 3827.
103. Xiang Y, Stine ZE, Xia J, et al. Targeted inhibition of tumor-specific glutaminase diminishes cell-autonomous tumorigenesis. J Clin Invest 2015; 125(6): 2293–2306.
104. Robinson MM, Mcbryant SJ, Tsukamoto T, et al. Novel mechanism of inhibition of rat kidney-type glutaminase by bis-2-(5-phenylacetamido-1, 2, 4-thiadiazol-2-yl) ethyl sulfide (BPTES). Biochem J 2007; 406: 407–414.
105. Shukla K, Ferraris DV, Thomas AG, et al. Design, synthesis, and pharmacological evaluation of bis-2-(5-phenylacetamido-1, 2, 4-thiadiazol-2-yl) ethyl sulfide 3 (BPTES) analogs as glutaminase inhibitors. J Med Chem 2012; 55: 10551–10563.
106. Lukey MJ, Wilson KF, Cerione RA. Therapeutic strategies impacting cancer cell glutamine metabolism. Future Med Chem 2013; 5(14): 1685–1700.
107. Gross MI, Demo SD, Dennison JB, et al. Antitumor activity of the glutaminase inhibitor CB-839 in triple-negative breast cancer. Mol Cancer Ther 2014; 13(4): 890–901.
108. Akins NS, Nielson TC, Le HV. Inhibition of glycolysis and glutaminolysis: an emerging drug discovery approach to combat cancer. Curr Top Med Chem 2018; 18(6): 494–504.
109. Saha SK, Islam SM, Abdullah-AL-Wadud M, et al. Multiomics analysis reveals that GLS and GLS2 differentially modulate the clinical outcomes of cancer. J Clin Medicine 2019; 8: 355.
110. Grigoryan RS, Panosyan EH, Seibel NL, et al. Changes of amino acid serum levels in pediatric patients with higher-risk acute lymphoblastic leukemia (CCG-1961). In Vivo 2004; 18(2): 107–112.
111. Tanaka K, Sasayama T, Irino Y, et al. Compensatory glutamine metabolism promotes glioblastoma resistance to mTOR inhibitor treatment. J Clin Invest 2015; 125(4): 1591–1602.
112. Dolfi SC, Chan LL-Y, Qiu J, et al. The metabolic demands of cancer cells are coupled to their size and protein synthesis rates. Cancer Metab 2013; 1: 20.
113. Jain M, Nilsson R, Sharma S, et al. Metabolite profiling identifies a key role for glycine in rapid cancer cell proliferation. Science 2012; 336: 1040–1044.
114. Mattaini KR, Sullivan MR, Vander Heiden MG. The importance of serine metabolism in cancer. J Cell Biol 2016; 214: 249–257.
115. Locasale JW. Serine, glycine and one-carbon units: cancer metabolism in full circle. Nat Rev Cancer 2013; 13(8): 572–583.
116. Fan J, Ye J, Kamphorst JJ, et al. Quantitative flux analysis reveals folate-dependent NADPH production. Nature 2014; 510: 298–302.
117. Ye J, Fan J, Venneti S, et al. Serine catabolism regulates mitochondrial redox control during hypoxia. Cancer Discov 2014; 4(12): 1406–1417.
118. Possemato R, Marks KM, Shaul YD, et al. Functional genomics reveal that the serine synthesis pathway is essential in breast cancer. Nature 2011; 476: 346–350.
119. Zhang B, Zheng A, Hydbring P, et al. PHGDH defines a metabolic subtype in lung adenocarcinomas with poor prognosis. Cell Rep 2017; 19: 2289–2303.
120. Luengo A, Gui DY, Vander Heiden MG. Targeting metabolism for cancer therapy. Cell Chem Biol 2017; 24: 1161–1180.
121. Tedeschi PM, Vazquez A, Kerrigan JE, et al. Mitochondrial methylenetetrahydrofolate dehydrogenase (MTHFD2) overexpression is associated with tumor cell proliferation and is a novel target for drug development. Mol Cancer Res 2015; 13(10): 1361–1366.
122. Vazquez A, Kamphorst JJ, Markert EK, et al. Cancer metabolism at a glance. J Cell Sci 2016; 129: 3367–3373.
123. Kuo C-Y, Ann DK. When fats commit crimes: fatty acid metabolism, cancer stemness and therapeutic resistance. Cancer Commun 2018; 38: 47.
124. Carracedo A, Cantley LC, Pandolfi PP. Cancer metabolism: fatty acid oxidation in the limelight. Nat Rev Cancer 2013; 13(4): 227–232.
125. Tirinato L, Pagliari F, Limongi T, et al. An overview of lipid droplets in cancer and cancer stem cells. Stem Cells Int 2017; 2017: 1656053.
126. Qu Q, Zeng F, Liu X, et al. Fatty acid oxidation and carnitine palmitoyltransferase I: emerging therapeutic targets in cancer. Cell Death Dis 2016; 7: e2226.
127. Chiarugi A, Dölle C, Felici R, et al. The NAD metabolome—a key determinant of cancer cell biology. Nat Rev Cancer 2012; 12(11): 741–752.
128. Yan S, Yang X-F, Liu H-L, et al. Long-chain acyl-CoA synthetase in fatty acid metabolism involved in liver and other diseases: an update. World J Gastroenterol 2015; 21: 3492.
129. Schreurs M, Kuipers F, Van Der Leij F. Regulatory enzymes of mitochondrial β-oxidation as targets for treatment of the metabolic syndrome. Obes Rev 2010; 11: 380–388.
130. Rufer AC, Thoma R, Hennig M. Structural insight into function and regulation of carnitine palmitoyltransferase. Cell Mol Life Sci 2009; 66(15): 2489–2501.
131. Bonnefont J-P, Djouadi F, Prip-Buus C, et al. Carnitine palmitoyltransferases 1 and 2: biochemical, molecular and medical aspects. Mol Aspects Med 2004; 25(5–6): 495–520.
132. Qian X, Hu J, Zhao J, et al. ATP citrate lyase expression is associated with advanced stage and prognosis in gastric adenocarcinoma. Int J Clin Exp Med 2015; 8(5): 7855–7860.
133. Currie E, Schulze A, Zechner R, et al. Cellular fatty acid metabolism and cancer. Cell Metab 2013; 18: 153–161.
134. Fang W, Cui H, Yu D, et al. Increased expression of phospho-acetyl-CoA carboxylase protein is an independent prognostic factor for human gastric cancer without lymph node metastasis. Med Oncol 2014; 31(7): 15.
135. Ma Y, Temkin SM, Hawkridge AM, et al. Fatty acid oxidation: an emerging facet of metabolic transformation in cancer. Cancer Lett 2018; 435: 92–100.
136. Menendez JA, Lupu R. Fatty acid synthase and the lipogenic phenotype in cancer pathogenesis. Nat Rev Cancer 2007; 7(10): 763–777.
137. Wu X, Dong Z, Wang CJ, et al. FASN regulates cellular response to genotoxic treatments by increasing PARP-1 expression and DNA repair activity via NF-κB and SP1. Proc Natl Acad Sci USA 2016; 113: E6965–E6973.
138. Tachibana K, Yamasaki D, Ishimoto K. The role of PPARs in cancer. PPAR Res 2008; 2008: 102737.
139. Flaig TW, Salzmann-Sullivan M, Su L-J, et al. Lipid catabolism inhibition sensitizes prostate cancer cells to antiandrogen blockade. Oncotarget 2017; 8: 56051.
140. Kim WT, Yun SJ, Yan C, et al. Metabolic pathway signatures associated with urinary metabolite biomarkers differentiate bladder cancer patients from healthy controls. Yonsei Med J 2016; 57(4): 865–871.
141. Yeh C-S, Wang J-Y, Cheng T-L, et al. Fatty acid metabolism pathway play an important role in carcinogenesis of human colorectal cancers by Microarray-Bioinformatics analysis. Cancer Lett 2006; 233: 297–308.
142. Wang T, Fahrmann JF, Lee H, et al. JAK/STAT3-regulated fatty acid β-oxidation is critical for breast cancer stem cell self-renewal and chemoresistance. Cell Metab 2018; 27: 136–150.e5.
143. Park JH, Vithayathil S, Kumar S, et al. Fatty acid oxidation-driven Src links mitochondrial energy reprogramming and oncogenic properties in triple-negative breast cancer. Cell Rep 2016; 14: 2154–2165.
144. Guo H, Zeng W, Feng L, et al. Integrated transcriptomic analysis of distance-related field cancerization in rectal cancer patients. Oncotarget 2017; 8: 61107.
145. Casals N, Zammit V, Herrero L, et al. Carnitine palmitoyltransferase 1C: from cognition to cancer. Prog Lipid Res 2016; 61: 134–148.
146. Shinohara H, Kumazaki M, Minami Y, et al. Perturbation of energy metabolism by fatty-acid derivative AIC-47 and imatinib in BCR-ABL-harboring leukemic cells. Cancer Lett 2016; 371: 1–11.
147. Garcia MR, Steinbauer B, Srivastava K, et al. Acetyl-CoA carboxylase 1-dependent protein acetylation controls breast cancer metastasis and recurrence. Cell Metab 2017; 26: 842–855.e5.
148. Luo J, Hong Y, Lu Y, et al. 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.
149. Liu Y. Fatty acid oxidation is a dominant bioenergetic pathway in prostate cancer. Prostate Cancer Prostatic Dis 2006; 9(3): 230–234.
150. Pacilli A, Calienni M, Margarucci S, et al. Carnitine-acyltransferase system inhibition, cancer cell death, and prevention of myc-induced lymphomagenesis. J Natl Cancer Inst 2013; 105: 489–498.
151. Cirillo A, Di Salle A, Petillo O, et al. High grade glioblastoma is associated with aberrant expression of ZFP57, a protein involved in gene imprinting, and of CPT1A and CPT1C that regulate fatty acid metabolism. Cancer Biol Ther 2014; 15: 735–741.
152. Tung S, Shi Y, Wong K, et al. PPARα and fatty acid oxidation mediate glucocorticoid resistance in chronic lymphocytic leukemia. Blood, J Am Hematol 2013; 122: 969–980.
153. Pike LS, Smift AL, Croteau NJ, et al. Inhibition of fatty acid oxidation by etomoxir impairs NADPH production and increases reactive oxygen species resulting in ATP depletion and cell death in human glioblastoma cells. Biochim Biophys Acta 2011; 1807(6): 726–734.
154. Ricciardi MR, Mirabilii S, Allegretti M, et al. Targeting the leukemia cell metabolism by the CPT1a inhibition: functional preclinical effects in leukemias. Blood, J Am Hematol 2015; 126: 1925–1929.
155. Ashrafian H, Horowitz JD, Frenneaux MP. Perhexiline. Cardiovasc Drug Rev 2007; 25: 76–97.
156. Lee EA, Angka L, Rota S-G, et al. Targeting mitochondria with avocatin B induces selective leukemia cell death. Cancer Res 2015; 75: 2478–2488.
157. Chen M, Huang J. The expanded role of fatty acid metabolism in cancer: new aspects and targets. Precis Clin Med 2019; 2(3): 183–191.
158. Watt MJ, Clark AK, Selth LA, et al. Suppressing fatty acid uptake has therapeutic effects in preclinical models of prostate cancer. Sci Transl Med 2019; 11: eaau5758.
159. Divisi D, Di Tommaso S, Salvemini S, et al. Diet and cancer. Acta Biomed Ateneo Parmense 2006; 77: 118.
160. Steele VE, Holmes CA, Hawk ET, et al. Lipoxygenase inhibitors as potential cancer chemopreventives. Cancer Epidemiol Biomarkers Prev 1999; 8(5): 467–483.
161. Moore GY, Pidgeon GP. Cross-talk between cancer cells and the tumour microenvironment: the role of the 5-lipoxygenase pathway. Int J Mol Sci 2017; 18: 236.
162. Bishayee K, Khuda-Bukhsh AR. 5-lipoxygenase antagonist therapy: a new approach towards targeted cancer chemotherapy. Acta Biochim Biophys Sin 2013; 45(9): 709–719.
163. Wejksza K, Lee-Chang C, Bodogai M, et al. Cancer-produced metabolites of 5-lipoxygenase induce tumor-evoked regulatory B cells via peroxisome proliferator–activated receptor α. J Immunol 2013; 190: 2575–2584.
164. Kelavkar UP, Cohen C. 15-lipoxygenase-1 expression upregulates and activates insulin-like growth factor-1 receptor in prostate cancer cells. Neoplasia 2004; 6(1): 41–52.
165. Middleton MK, Zukas AM, Rubinstein T, et al. Identification of 12/15-lipoxygenase as a suppressor of myeloproliferative disease. J Exp Med 2006; 203: 2529–2540.
166. Shureiqi I, Wu Y, Chen D, et al. The critical role of 15-lipoxygenase-1 in colorectal epithelial cell terminal differentiation and tumorigenesis. Cancer Res 2005; 65: 11486–11492.
167. Jiang WG, Watkins G, Douglas-Jones A, et al. Reduction of isoforms of 15-lipoxygenase (15-LOX)-1 and 15-LOX-2 in human breast cancer. Prostaglandins Leukot Essent Fatty Acids 2006; 74(4): 235–245.
168. Gao X, Grignon DJ, Chbihi T, et al. Elevated 12-lipoxygenase mRNA expression correlates with advanced stage and poor differentiation of human prostate cancer. Urology 1995; 46(2): 227–237.
169. Mohammad A, Abdel H, Abdel W, et al. Expression of cyclooxygenase-2 and 12-lipoxygenase in human breast cancer and their relationship with HER-2/neu and hormonal receptors: impact on prognosis and therapy. Indian J Cancer 2006; 43(4): 163–168.
170. Borgeat P, Hamberg M, Samuelsson B. Transformation of arachidonic acid and homo-gamma-linolenic acid by rabbit polymorphonuclear leukocytes. Monohydroxy acids from novel lipoxygenases. J Biol Chem 1976; 251: 7816–7820.
171. Funk CD, Hoshiko S, Matsumoto T, et al. Characterization of the human 5-lipoxygenase gene. Proc Natl Acad Sci USA 1989; 86: 2587–2591.
172. Sundaram S, Ghosh J. Expression of 5-oxoETE receptor in prostate cancer cells: critical role in survival. Biochemi Biophys Res Commun 2006; 339: 93–98.
173. Weitzel F, Wendel A. Selenoenzymes regulate the activity of leukocyte 5-lipoxygenase via the peroxide tone. J Biol Chem 1993; 268: 6288–6292.
174. Ghosh J. Targeting 5-lipoxygenase for prevention and treatment of cancer. Curr Enzyme Inhib 2008; 4: 18–28.
175. Werz O, Szellas D, Steinhilber D. Reactive oxygen species released from granulocytes stimulate 5-lipoxygenase activity in a B-lymphocytic cell line. Eur J Biochem 2000; 267(5): 1263–1269.
176. Percival MD. Human 5-lipoxygenase contains an essential iron. J Biol Chem 1991; 266: 10058–10061.
177. Spanbroek R, Hildner M, Köhler A, et al. IL-4 determines eicosanoid formation in dendritic cells by down-regulation of 5-lipoxygenase and up-regulation of 15-lipoxygenase 1 expression. Proc Natl Acad Sci USA 2001; 98: 5152–5157.
178. Rioux N, Castonguay A. Inhibitors of lipoxygenase: a new class of cancer chemopreventive agents. Carcinogenesis 1998; 19(8): 1393–1400.
179. Hassan S, Carraway RE. Involvement of arachidonic acid metabolism and EGF receptor in neurotensin-induced prostate cancer PC3 cell growth. Regul Pept 2006; 133: 105–114.
180. Erler JT, Bennewith KL, Nicolau M, et al. Lysyl oxidase is essential for hypoxia-induced metastasis. Nature 2006; 440: 1222–1226.
181. Erler JT, Bennewith KL, Cox TR, et al. Hypoxia-induced lysyl oxidase is a critical mediator of bone marrow cell recruitment to form the premetastatic niche. Cancer Cell 2009; 15: 35–44.
182. Sharma-Walia N, Chandran K, Patel K, et al. The Kaposi’s sarcoma-associated herpesvirus (KSHV)-induced 5-lipoxygenase-leukotriene B4 cascade plays key roles in KSHV latency, monocyte recruitment, and lipogenesis. J Virol 2014; 88: 2131–2156.
183. Romano M, Catalano A, Nutini M, et al. 5-lipoxygenase regulates malignant mesothelial cell survival: involvement of vascular endothelial growth factor. FASEB J 2001; 15(13): 2326–2336.
184. Romano M, Claria J. Cyclooxygenase-2 and 5-lipoxygenase converging functions on cell proliferation and tumor angiogenesis: implications for cancer therapy. FASEB J 2003; 17(14): 1986–1995.
185. Park S-W, Heo D-S, Sung MW. The shunting of arachidonic acid metabolism to 5-lipoxygenase and cytochrome p450 epoxygenase antagonizes the anti-cancer effect of cyclooxygenase-2 inhibition in head and neck cancer cells. Cell Oncol 2012; 35(1): 1–8.
186. Jiang WG, Douglas-Jones AG, Mansel RE. Aberrant expression of 5-lipoxygenase-activating protein (5-LOXAP) has prognostic and survival significance in patients with breast cancer. Prostag Leukotr Ess 2006; 74(2): 125–134.
187. Anderson K, Seed T, Ondrey F, et al. The selective 5-lipoxygenase inhibitor A63162 reduces PC3 proliferation and initiates morphologic changes consistent with secretion. Anticancer Res 1994; 14(5A): 1951–1960.
188. Hostanska K, Daum G, Saller R. Cytostatic and apoptosis-inducing activity of boswellic acids toward malignant cell lines in vitro. Anticancer Res 2002; 22(5): 2853–2862.
189. Chen X, Wang S, Wu N, et al. Overexpression of 5-lipoxygenase in rat and human esophageal adenocarcinoma and inhibitory effects of zileuton and celecoxib on carcinogenesis. Clin Cancer Res 2004; 10: 6703–6709.
190. Carter GW, Young PR, Albert DH, et al. 5-lipoxygenase inhibitory activity of zileuton. J Pharmacol Exp Ther 1991; 256: 929–937.
191. Hoque A, Lippman SM, Wu T-T, et al. Increased 5-lipoxygenase expression and induction of apoptosis by its inhibitors in esophageal cancer: a potential target for prevention. Carcinogenesis 2005; 26(4): 785–791.
192. Goto H, Nishizawa Y, Katayama H, et al. Induction of apoptosis in an estrogen-responsive mouse Leydig tumor cell by leukotriene. Oncol Rep 2007; 17(1): 225–232.
193. Hayashi T, Nishiyama K, Shirahama T. Inhibition of 5-lipoxygenase pathway suppresses the growth of bladder cancer cells. Int J Urol 2006; 13(8): 1086–1091.
194. Wang C, Pei A, Chen J, et al. A natural coumarin derivative esculetin offers neuroprotection on cerebral ischemia/reperfusion injury in mice. J Neurochem 2012; 121(6): 1007–1013.
195. Tomohiro N, Yasuko K, Sei-Itsu M. Inhibitory effect of esculetin on 5-lipoxygenase and leukotriene biosynthesis. Biochim Biophys Acta 1983; 753: 130–132.
196. Park C, Jin C-Y, Kwon HJ, et al. Induction of apoptosis by esculetin in human leukemia U937 cells: roles of Bcl-2 and extracellular-regulated kinase signaling. Toxicol in Vitro 2010; 24(2): 486–494.
197. Shi F, Li T, Liu Z, et al. FOXO1: another avenue for treating digestive malignancy. Semin Cancer Biol 2018; 50: 124–131.
198. Zhang Y, Gan B, Liu D, et al. FOXO family members in cancer. Cancer Biol Ther 2011; 12: 253–259.
199. Fu Z, Tindall D. FOXOs, cancer and regulation of apoptosis. Oncogene 2008; 27: 2312–2319.
200. Brownawell AM, Kops GJ, Macara IG, et al. Inhibition of nuclear import by protein kinase B (Akt) regulates the subcellular distribution and activity of the forkhead transcription factor AFX. Mol Cell Biol 2001; 21(10): 3534–3546.
201. Singh A, Plati J, Khosravi-Far R. Harnessing the tumor suppressor function of FOXO as an alternative therapeutic approach in cancer. Curr Drug Targets 2011; 12(9): 1311–1321.
202. Yadav RK, Chauhan AS, Zhuang L, et al. FOXO transcription factors in cancer metabolism. Semin Cancer Biol 2018; 50: 65–76.
203. Hermeking H, Benzinger A. 14-3-3 proteins in cell cycle regulation. Semin Cancer Biol 2006; 16(3): 183–192.
204. Abeshouse A, Ahn J, Akbani R, et al. The molecular taxonomy of primary prostate cancer. Cell 2015; 163: 1011–1025.
205. Coomans de Brachène A, Demoulin JB. FOXO transcription factors in cancer development and therapy. Cell Mol Life Sci 2016; 73(6): 1159–1172.
206. Eijkelenboom A, Burgering BM. FOXOs: signalling integrators for homeostasis maintenance. Nat Rev Mol Cell Biol 2013; 14(2): 83–97.
207. Li C-F, Zhang W-G, Liu M, et al. Aquaporin 9 inhibits hepatocellular carcinoma through up-regulating FOXO1 expression. Oncotarget 2016; 7: 44161.
208. Kim SY, Yoon J, San Ko Y, et al. Constitutive phosphorylation of the FOXO1 transcription factor in gastric cancer cells correlates with microvessel area and the expressions of angiogenesis-related molecules. BMC Cancer 2011; 11: 264.
209. Potente M, Urbich C, Sasaki Hofmann WK, et al. Involvement of FOXO transcription factors in angiogenesis and postnatal neovascularization. J Clin Invest 2005; 115(9): 2382–2392.
210. Jiang S, Han J, Li T, et al. Curcumin as a potential protective compound against cardiac diseases. Pharmacol Res 2017; 119: 373–383.
211. Zhao Z, Li C, Xi H, et al. Curcumin induces apoptosis in pancreatic cancer cells through the induction of forkhead box O1 and inhibition of the PI3K/Akt pathway. Mol Med Rep 2015; 12(4): 5415–5422.
212. Myatt SS, Brosens JJ, Lam EW-F. Sense and sensitivity: FOXO and ROS in cancer development and treatment. Antioxid Redox Sign 2011; 14: 675–687.
213. Stine ZE, Walton ZE, Altman BJ, et al. MYC, metabolism, and cancer. Cancer Discov 2015; 5: 1024–1039.
214. Wilhelm K, Happel K, Eelen G, et al. FOXO1 couples metabolic activity and growth state in the vascular endothelium. Nature 2016; 529: 216–220.
215. Masui K, Tanaka K, Akhavan D, et al. MTOR complex 2 controls glycolytic metabolism in glioblastoma through FOXO acetylation and upregulation of c-Myc. Cell Metab 2013; 18: 726–739.
216. Zhang X, Rielland M, Yalcin S, et al. Regulation and function of FOXO transcription factors in normal and cancer stem cells: what have we learned. Curr Drug Targets 2011; 12(9): 1267–1283.
217. Farhan M, Wang H, Gaur U, et al. FOXO signaling pathways as therapeutic targets in cancer. Int J Biol Sci 2017; 13(7): 815–827.
218. Fendt S-M. Is there a therapeutic window for metabolism-based cancer therapies. Front Endocrinol 2017; 8: 150.

Cite article

Cite article

Cite article

OR

Download to reference manager

If you have citation software installed, you can download article citation data to the citation manager of your choice

Share options

Share

Share this article

Share with email
EMAIL ARTICLE LINK
Share on social media

Share access to this article

Sharing links are not relevant where the article is open access and not available if you do not have a subscription.

For more information view the Sage Journals article sharing page.

Information, rights and permissions

Information

Published In

Article first published online: October 7, 2020
Issue published: October 2020

Keywords

  1. Cancer metabolism
  2. glycolysis
  3. oxidative phosphorylation
  4. pentose phosphate pathway
  5. cancer therapy

Rights and permissions

© The Author(s) 2020.
Creative Commons License (CC BY-NC 4.0)
This article is distributed under the terms of the Creative Commons Attribution-NonCommercial 4.0 License (https://creativecommons.org/licenses/by-nc/4.0/) which permits non-commercial use, reproduction and distribution of the work without further permission provided the original work is attributed as specified on the SAGE and Open Access pages (https://us.sagepub.com/en-us/nam/open-access-at-sage).
Request permissions for this article.
PubMed: 33028168

Authors

Affiliations

Pegah Farhadi
Medical Biology Research Center, School of Medicine, Kermanshah University of Medical Sciences, Kermanshah, Iran
Reza Yarani
Translational Type 1 Diabetes Research, Department of Clinical Research, Steno Diabetes Center Copenhagen, Gentofte, Denmark
Sadat Dokaneheifard
Sylvester Comprehensive Cancer Center, Department of Human Genetics, University of Miami Miller School of Medicine, Miami, FL, USA
Kamran Mansouri
Medical Biology Research Center, School of Medicine, Kermanshah University of Medical Sciences, Kermanshah, Iran
Department of Molecular Medicine, School of Medicine, Kermanshah University of Medical Sciences, Kermanshah, Iran

Notes

Kamran Mansouri, Medical Biology Research Center, School of Medicine, Kermanshah University of Medical Sciences, Sorkheh Ligeh Blvd, P.O. Box 1568, Kermanshah 6715847141, Iran. Emails: [email protected]; [email protected]

Author contributions

All authors contributed equally to this work.

Metrics and citations

Metrics

Journals metrics

This article was published in Tumor Biology.

VIEW ALL JOURNAL METRICS

Article usage*

Total views and downloads: 15103

*Article usage tracking started in December 2016


Articles citing this one

Receive email alerts when this article is cited

Web of Science: 0

Crossref: 0

There are no citing articles to show.

Figures and tables

Figures & Media

Tables

View Options

View options

PDF/ePub

View PDF/ePub

Get access

Access options

If you have access to journal content via a personal subscription, university, library, employer or society, select from the options below:


Alternatively, view purchase options below:

Purchase 24 hour online access to view and download content.

Access journal content via a DeepDyve subscription or find out more about this option.