Abstract
Cancer reprogramming is an important facilitator of cancer development and survival, with tumor cells exhibiting a preference for aerobic glycolysis beyond oxidative phosphorylation, even under sufficient oxygen supply condition. This metabolic alteration, known as the Warburg effect, serves as a significant indicator of malignant tumor transformation. The Warburg effect primarily impacts cancer occurrence by influencing the aerobic glycolysis pathway in cancer cells. Key enzymes involved in this process include glucose transporters (GLUTs), HKs, PFKs, LDHs, and PKM2. Moreover, the expression of transcriptional regulatory factors and proteins, such as FOXM1, p53, NF-κB, HIF1α, and c-Myc, can also influence cancer progression. Furthermore, lncRNAs, miRNAs, and circular RNAs play a vital role in directly regulating the Warburg effect. Additionally, gene mutations, tumor microenvironment remodeling, and immune system interactions are closely associated with the Warburg effect. Notably, the development of drugs targeting the Warburg effect has exhibited promising potential in tumor treatment. This comprehensive review presents novel directions and approaches for the early diagnosis and treatment of cancer patients by conducting in-depth research and summarizing the bright prospects of targeting the Warburg effect in cancer.
Key words: Warburg effect, Cancer, Regulatory factor, LncRNAs, miRNAs, Traditional Chinese medicine, Small-molecule drug, Drug resistance
Graphical abstract
This review summarizes the molecular mechanism of Warburg effect in cancer, and discusses the strategy of regulating Warburg effect to treat cancer.
1. Introduction
Cancer development is driven by abnormal signaling and metabolic reprogramming. Oncogenes closely relate to metabolic changes in cancer cells. Metabolic reprogramming is often mediated by various carcinogenic signals and controls cancer cell metabolism by altering the expression level and activity of key metabolic enzymes1. Therefore, a complete understanding of the relationship between oncoprotein and specific metabolism of cancer cells is conducive to developing of novel therapeutic strategies. In this manuscript, we begin with the current status of tumor glucose metabolism regulation therapy, address issues in current metabolic regulation research, and propose strategies for complex tumor glucose metabolism.
Glucose is the main energy source of life activity and tumor cell activity. In malignant tumor tissues, the glucose intake is much higher than that in normal tissues2. The metabolic imbalance resulting from restricting glucose absorption, specifically abnormal glycolytic gene expression, closely relates to tumor cell growth and developmen3. Under physiological oxygen concentration and functional mitochondria, tumor cells shift from oxidative phosphorylation to aerobic glycolysis, characterized by high glucose uptake and lactic acid production, known as the Warburg effect4,5. This metabolic pathway is crucial to cancer. It is also the reason that the growth rate of cancer cells is much faster than that of normal cells.
Warburg effect provides many benefits for cancer cells to compete and share energy. Even with sufficient oxygen supply, cancer cells increase energy supply through aerobic glycolysis due to improper proliferation and oncogenes’ influence6. ATP is the main energy source of cell activity. Because of the hypoxia microenvironment and Warburg effect in tumors, tumor cells mainly undergo aerobic glycolysis, and 95% of ATP production is obtained through the glycoside-dependent mechanism7. Although compared with aerobic respiration, the amount of ATP generated during aerobic glycolysis under anaerobic conditions is very small, the expression of GLUT in tumor cells is often significantly increased. Its large amount of expression can meet the energy needs of tumor cells, accelerate the uptake of glucose by tumor cells, provide sufficient raw materials for aerobic glycolysis, and promote the production of ATP8, which meets the needs of rapid tumor growth. Simultaneously, due to the substantial production of lactic acid, the cells reshape the tumor microenvironment, influencing the infiltration of various immune cells and closely correlating with tumor initiation, progression, invasion and metastasis.
Tumor cells at different stages of differentiation mainly rely on different modes of glucose metabolism. Cancer cells exhibit metabolic plasticity at different stages of metastasis. Changes in tumor metabolism can promote the occurrence of drug resistance. The Warburg effect also plays an important role in immune metabolism. LncRNAs, miRNAs, and Circular RNAs play a crucial role in regulating the glucose metabolism process of cancer. They mainly target and calibrate genes that regulate metabolism, causing changes in the metabolism of cancer cells, thereby affecting their growth, diffusion, and metastasis. This manuscript summarizes and analyzes the abnormal effects of glycolytic pathways in cancer, the mechanisms of key glycolytic enzymes and related signaling pathways, and the research progress of targeted drugs. It also explores the potential of combination therapy targeting complementary metabolic pathways to enhance or synergistically inhibit the survival of tumor cells. Metabolomics, an emerging technology, aims to study the abnormal metabolic patterns of tumors and holds promise for diagnosing and providing personalized treatment for tumor metabolic typing. It can also enable researchers to study the reprogramming of drug-resistant tumor cell metabolic networks more efficiently and systematically. At present, the mechanism of cancer occurrence and development is still unclear, especially since the research on its relationship with the Warburg effect is still in its infancy. Abnormal activity of aerobic glycolysis is a crucial feature of tumor cells, revealing its essential role in tumor occurrence and development, providing new ideas for early diagnosis and treatment of cancer, and developing new effective therapeutic targets is a challenging task.
2. Warburg effect: A glorious research history spanning a century
Glucose is an essential nutrient crucial for sustaining the life activities of animals. The energy of cells mainly comes from carbohydrate metabolism, which has two pathways, oxidative phosphorylation and aerobic glycolysis. According to the “Pasteur effect”, low oxygen concentrations favor fermentation, while high oxygen levels inhibit it, promoting aerobic respiration and reducing aerobic glycolysis rates. This supports efficient energy utilization and carbon use for synthesis reactions. Since NADH can enter mitochondria and undergo oxidation, it inhibits lactic acid production. During hypoxia, NADH cannot be oxidized through the respiratory chain, and pyruvate acts as a hydrogen acceptor to reduce to lactate9. Therefore, normal cells are inhibited from aerobic glycolysis under aerobic conditions, mainly utilizing aerobic respiration. In 1923, German biochemist Otto Heinrich Warburg discovered significant differences in the carbohydrate metabolism of tumor cells compared to normal cells5. During the process of malignant proliferation, the energy metabolism pathway was reshaped, and the energy metabolism pathway of cancerous cells changed from the tricarboxylic acid cycle to aerobic glycolysis. Otto Warburg noted that even with sufficient oxygen, tumor cells favor aerobic glycolysis over oxidative phosphorylation for energy production due to its efficiency5. This phenomenon was subsequently coined the term “Warburg effect” by E. Racker in 197610. Warburg effect is considered to be the metabolic reconstruction of tumor cells in order to adapt to the energy needs of proliferation and differentiation. This effect provides a number of ATP for the growth and proliferation of tumor cells, creates a microenvironment fitable for tumor cells to survive, and enhances invasion and metastasis ability, and also helps tumor cells escape from the mechanism of body immunity and cell apoptosis11. Otto Warburg was honored with the Nobel Prize in Physiology or Medicine in 1931 for discovering cellular respiratory oxidotransferase. In 1956, Otto Heineieh Warburg further demonstrated that the Warburg effect originated from “irreversible respiratory system damage” caused by mitochondrial dysfunction. However, during the 20 years from 1952 to 1976, a large number of researchers successively demonstrated that most tumors were not accompanied by mitochondrial dysfunction while aerobic glycolysis was upregulated, opposing Warburg's initial “respiratory injury theory”. However, there are still misunderstandings about the reasons for the Warburg effect and its relationship with mitochondrial oxidative metabolism. From 1964 to 1965, researchers separated type I, II, III and IV hexokinase (HK) isoenzymes by ion exchange chromatography and electrophoresis12,13. In 1965, it was discovered that the absence of phosphofructose can lead to glycogen storage disease type VII14. In 1969, Louis Sokoloff employed the glucose analogue 2-deoxy-d-glucose (2-DG), which can readily penetrate the blood–brain barrier (BBB), in positron emission tomography (PET) imaging to assess glucose utilization in the brain15. Moreover, Louis Sokoloff synthesized the most widely used radioactive tracer 2-[18F] fluoro-2-deoxy-d-glucose (18F-DG or FDG) in 199616. Glucose transporter type 1 (GLUT1), a protein that transports glucose, was first isolated from red blood cells in 197717. The GLUT1 gene sequence was identified in 198518. In the same year, researchers pointed out the importance of HK2's synthesis and metabolism through the pentose phosphate pathway19. In 1994, it was discovered that HK2 was highly expressed in tumors. However, due to the already discovered abnormal glucose metabolism in tumors, researchers were still unable to confirm whether the function of HK2 in tumors was similar to that in normal tissues20. Subsequently, researchers developed an interest in studying HK2 as a potential target in tumors. As one of the metabolic characteristics in tumor cells, Warburg effect has been found for a hundred years. It has been widely studied and deeply discussed in recent 20 years. At present, there are many explanations for the universal Warburg effect in tumor cells. Since 1998, researchers have successively confirmed that cancer cells heavily depend on mitochondrial respiration, and mitochondrial dysfunction can contribute to the development of the Warburg effect21, 22, 23, 24. In 2000, researchers began using 2-DG to evaluate in vitro and in vivo glucose uptake activity without the use of radioactive substances25. Subsequently, it was proven that 2-DG as a glucose metabolism inhibitor mainly inhibits aerobic glycolysis by acting on HK226. In 2004, researchers confirmed that the Warburg effect is not a simple adaptation to hypoxia27. In 2009, Lewis C. Cantley and Craig B. Thompson reviewed the “Warburg effect” in the journal Science, revealing that some signal transduction pathways associated with cell proliferation can also control metabolic pathways responsible for synthesizing nutrients into biomass28. Some cancer-related mutations enable tumor cells to obtain and metabolize nutrients in a way conducive to proliferation rather than efficiently producing ATP. It is suggested that a further understanding of the relationship between cell metabolism and growth control may ultimately be beneficial to the treatment of cancer. In 2011, the Warburg effect is considered a core feature of the malignant phenotype of cancer and is referred to as a “marker of cancer”29. In the same year, PKM2 inhibitor Shikonin was discovered to hold potential for treating tumors by inhibiting the Warburg effect through impeding the final rate-limiting step of aerobic glycolysis30. In 2012, researcher Yi et al.31 revealed in the journal Science that O-linked β-N-acetylglucosamine (O-GlcNAc) induces glycosylation at Ser529 of phosphofructokinase-1 (PFK1) in response to the hypoxic state of cancer cells, inhibits PFK1 activity, and then redirects glucose flux during aerobic glycolysis to the PPP, providing the energy required for cancer cell growth through compensation. Inhibiting glycosylation at PFK1 Ser529 can inhibit cancer cell proliferation and tumor tissue growth. The regulatory mechanism of PFK1 in aerobic glycolysis in cancer has been revealed, suggesting that inhibiting PFK1 glycosylation may provide a new therapeutic approach for combating cancer. In 2014, Professor Yan Ning's research group in Tsinghua University School of Medicine analyzed the crystal structure of human glucose transporter GLUT1 for the first time in the journal Nature, revealing the process of the transport of essential substances in the human body to the cell membrane, which brings therapeutic prospects for manual intervention as a potential target of disease diagnosis or drug development32. In 2015, Bradley A. Webb's team first analyzed the biological crystal structure of mammalian PFK1 tetramers in the journal Nature, which helps people understand the functional changes caused by PFK1 mutations in diseases, and based on this, developed anti-tumor drugs targeting PFK1 to affect the warburg effect33. In 2020, Professor Olivier Pourqui é’s team from Harvard Medical School published an interesting research online in the journal Nature34. The research found the close similarity between the developing tail bud cells and the cancer cells (high pHi and low pHe) showing high Warburg metabolism, suggesting that some tumor cells will reactivate the developmental metabolic program. In 2021, Ming's team of Memorial Sloan Kettering Cancer Center in the United States published a study in the journal Science that offered a fresh perspective on the “Warburg effect”35. This study shows that there is an undiscovered link between Warburg metabolism and the activity of phosphatidylinositol three kinase (PI3K), which is realized by latate dehydrogenase A (LDHA). Tumors cells can use Warburg metabolism to maintain the biological activity of PI3K signaling pathway, thus ensuring the continuous proliferation of cancer cells. This discovery challenges the view in textbooks for a long time. More importantly, this study proposes a promising cancer treatment method. By inhibiting the Warburg “switch” of LDHA, it can inhibit the growth of cancer cells and then be used to treat cancer (Fig. 1A).
Figure 1.
The timeline and classic process of the Warburg effect. (A) The timeline of the Warburg effect from target discovery to drug development. (B) The classic process of the Warburg effect from glucose to lactate. The pyruvate produced by aerobic glycolysis also enters the tricarboxylic acid cycle inhibiting the Warburg effect.
A prominent metabolic alteration in tumor cells is the persistent abnormal activation of aerobic glycolysis, even under aerobic conditions, known as the Warburg effect, also names aerobic glycolysis. The classic Warburg effect in tumor cells consists of several precise steps in succession: firstly, glucose is transported to tumor cells at high speed via GLUTs and sodium–glucose junction transporter 1, and then phosphorylated into glucose-6 phosphate by HK, which is a speed-limiting step that provides direct feedback inhibition to preserve energy. HK combines with mitochondrial membrane and has high affinity for glucose, which is helpful to start aerobic glycolysis at low glucose level. Next, glucose-6-phosphate isomerase (GPI) converts glucose-6-phosphate into fructose-6-phosphate. Fructose-6-phosphate is then phosphorylated to produce fructose-1,6-diphosphate and fructose 2,6-diphosphate, facilitated by PFK1 and PFK2, with the consumption of an ATP molecule. Secondly, aldolase converts fructose-1,6-diphosphate into glyceraldehyde 3-phosphate and dihydroxyacetone phosphate. Glyceraldehyde-3-phosphate is converted from glyceraldehyde-3-phosphate dehydrogenase to glycerate 1,3-diphosphate, and then converted from phosphoglycerate kinase to glycerate 3-phosphate to generate two ATP molecules. Phosphoglycerate mutase isomerizes 3-phosphoglycerate to 2-phosphoglycerate, and then forms phosphoenolpyruvate. Finally, PEP is converted into pyruvate to generate an ATP molecule through the rate limiting step of aerobic glycolysis catalyzed by pyruvate kinase (Fig. 1B).
3. Warburg effect in cancer
Aerobic glycolysis is an adaptive metabolic change in tumor tissue, leading to cancer being considered a metabolic disease. Besides facilitating the rapid production of ATP in tumor tissues, aerobic glycolysis provides tumor cells with essential metabolic intermediates and precursors for phospholipid and nucleic acid synthesis. It also shields cancer cells from chemotherapy drug damage36. The Warburg effect, characterized by low efficiency but rapid metabolism, consumes substantial glucose in the tumor microenvironment. It synthesizes necessary precursor substances for rapid tumor proliferation and secretes lactic acid, creating an immunosuppressive microenvironment conducive to tumor growth and metastasis37. Warburg effect can promote the upregulation of the expression of HK, GLUT, LDH, PK and 3-phosphoinositide-dependent protein kinase 1 (PDK1), improve the tolerance of tumor cells to hypoxia conditions, and then promote the invasion and metastasis of malignant tumors. Therefore, the metabolic mode targeting tumor cells is theoretically a strategy to fundamentally solve the malignant transformation of tumors. Now, after many years of clinical research on anti-tumor strategies for oncogenes fell into difficulties, the metabolic mode of tumor cells has attracted people's attention again. Anti-cancer drugs based on Warburg effect have gradually become a hot spot in tumor research.
Different initial lesions of tumors depend on different metabolic pathways. Ovarian cancer stem cells overexpress genes concerned with glucose uptake, oxidative phosphorylation (OXPHOS) and fat acid oxidation (FAO), and also overexpress the key enzyme for pyruvate to enter TCA cycle, preferring pyruvate to enter TCA, maintaining OXPHOS metabolic characteristics38. In high-grade glioma tissue, glioma stem cells mainly gather in two specific regions, around tumor blood vessels and necrotic areas with poor blood supply. The characteristics of the two microenvironments are completely different, with the former having relatively abundant oxygen content and sugars, while the latter has the opposite. However, as the tumor progresses, the proportion of the latter will gradually increase. On the other hand, based on differences in genetic phenotypes, glioma stem cells can be divided into at least two subtypes, namely the pre neuronal type and the mesenchymal type39. The biological characteristics of glioma stem cells are intricately linked to their microenvironment. Research has found that glioma stem cells clustered in hypoxic necrotic areas are mainly mesenchymal cells with relatively high malignancy, and the hypoxic microenvironment is conducive to maintaining stem cell characteristics such as self-replication and low differentiation. Conversely, glioma stem cells in the surrounding area of tumor blood vessels are mainly neuronal type with lower malignancy. It is worth noting that with the dynamic process of tumor volume growth and angiogenesis, changes in the microenvironment will promote the epithelial mesenchymal transition of glioma stem cells by affecting the glucose metabolism mode of stem cells and the expression of other regulatory factors40. Similar to other tumor cells, glioma stem cells highly depend on aerobic glycolysis to maintain their own energy supply even under relatively normal oxygen levels. Research results from the Anderson Cancer Center in the United States further validate this theory41. The aerobic glycolysis level of brain glioma stem cells, especially mesenchymal glioma stem cells, significantly increased, becoming the primary mode of energy metabolism41. However, compared with ordinary high-grade glioma cells, the glucose uptake level of glioma stem cells is relatively low, and the ATP production is higher than that of high-grade glioma cells, indicating that OXPHOS is still an important energy source of glioma stem cells42. This may be attributed to the heterogeneity of glucose metabolism in glioma stem cells from varying tissue sources or genotypes. OXPHOS and aerobic glycolysis mutually complement each other in glioma stem cells, and the glucose metabolism mode of these cells exhibits heterogeneity across different growth and development stages. Tumor cells at different differentiation stages mainly depend on different glucose metabolism modes43, the level of aerobic glycolysis is also regulated differently.
3.1. Cancer initiation and progression
Over-proliferation is one of the most important malignant behaviors of tumor cells. Most of the energy that tumor cells need to absorb in the process of proliferation mainly comes from the aerobic glycolysis of Warburg effect. Warburg effect plays an essential role in tumorigenesis and development44. Cells have different metabolic characteristics at different stages. Based on the different metabolic characteristics at various stages of the disease, the researchers put forward the hypothesis that the continuous activation of genes promotes the metabolic changes in the process of tumor development45. The first wave of oncogene activation transformed tumor cell metabolism into partial glycolytic phenotype. At this time, the imbalance of tumor cell proliferation and angiogenesis leads to hypoxia of tumor cells, triggering the activation of the second wave of genes, accelerating aerobic glycolysis, leading to the classic “Warburg” phenotype and almost completely inhibited OXPHOS. The imbalance between high energy demand and nutrient shortage of tumor cells induced the third wave of gene expression, and tumor cells supported their survival through glutamine. α-Ketoglutaric acid from glutamine enters TCA, re-establishes OXPHOS, and provides ATP and NADPH. Retrograde signals from these reactivated mitochondria trigger the fourth wave of gene recoding to regulate mitochondrial biogenesis and degradation45. The activation of different oncogenes under different metabolic states further induces the plasticity of tumor metabolism so that tumor cells can make the best use of available metabolic substrates and adapt to the rapidly changing microenvironment. As ovarian cancer progresses, glucose and fatty acid oxidation decrease significantly, leading to increased lactic acid production and a shift towards glycolytic metabolism46. In addition, compared with benign or advanced ovarian cancer cells, the initial cells of ovarian cancer have reduced oxidation of glucose and fatty acids, increased aerobic glycolysis, and tended to aerobic glycolysis phenotype47. Unlike most cells, normal prostate epithelial cells are highly dependent on glycolytic metabolic capacity. In the early stage of prostate cancer, cells gradually changed to OXPHOS-dependent metabolic mode, and in the tumor progression stage, cells preferentially carried out aerobic glycolysis metabolism. Therefore, prostate cancer cells use aerobic glycolysis and OXPHOS to supply energy at different stages of disease progression48. It is worth noting that there are also differences in metabolic characteristics between breast cancer and normal breast tissue, which reflects the heterogeneity and complexity of breast cancer49. Apart from glucose, amino acids like glutamine and serine serve as crucial substrates for cell growth and proliferation. When tumor cells take up glutamine, it undergoes conversion to glutamate through the action of glutaminase (GLS). Differential expression of GLS has been observed in breast cancer tissue, depending on the breast cancer subtype50.
Reprogramming of energy metabolism is considered a hallmark that promotes tumor onset and progression51. Oncogenes drive the imbalance of metabolic pathways, which provides selective advantages for tumor cells, enabling them to proliferate highly in harsh microenvironments and improving their survival rate52. In addition, metabolites produced by energy metabolism Reprogramming interact with signal transduction pathways to promote the occurrence of tumors. Changes in the metabolic program of tumor cells also further affect other cells in the tumor microenvironment, helping to regulate processes closely related to tumor development53. The metabolic effects of tumors reveal key factors in the occurrence and development of tumors, opening up new diagnostic and therapeutic perspectives (Fig. 2A).
Figure 2.
The role of the Warburg effect in cancer. (A) Up-regulation of cancer-promoting factors and down-regulation of tumor suppressor factors enhance the Warburg effect and promote tumor initiation and progression. (B) Up-regulation of cancer-promoting factors and down-regulation of tumor suppressor factors enhanced the Warburg· effect and promoted the metastasis of primary tumors. (C) LPS promotes the polarization of M2 macrophages to produce immune factors and regulate tumor progression. (D). The Warburg effect promotes the transformation of drug-sensitive tumor cells into drug-resistant tumor cells.
3.2. Cancer metastasis
The primary cause of tumor resistance, metastasis and recurrence is the presence of a small number of tumor cells with strong stem cell characteristics, including self-renewal ability and multiple differentiation capability, known as “cancer stem cells (CSCs)”54. Due to the presence of CSCs, tumor cell lines can exhibit certain stemness. The strength of tumor cell lines can reflect their malignancy and potential for malignant events such as metastasis and drug resistance55. The metabolic transformation from oxidative phosphorylation to aerobic glycolysis can affect tumor cell stemness and participate in regulating tumor invasion and distant metastasis. Pancreatic cancer cells showed extensive enhancement of aerobic glycolysis, including overexpression of glycolytic enzymes and increased production of lactic acid, which regulated tumor cell metastasis by promoting epithelial–mesenchymal transformation, tumor angiogenesis and metastasis and colonization of distant organs56. The tumor cells in the process of metastasis isolated from the blood, compared with the tumor cells in the primary tumor or lung metastasis, have the priority of OXPHOS57. Compared with solid tumor cells, tumor cells isolated from ascites of ovarian cancer patients tend to express glycolytic phenotype, suggesting that the glycolytic rate decreases after tumor attachment and growth47. It further supports the concept of metabolic plasticity of cancer cells at different stages of metastasis. In addition, the researchers found that once the tumor cells leave the primary focus, they will have metabolic changes different from the primary focus to facilitate survival after metastasis. The metastatic tumor cells exhibit distinct metabolic advantages depending on the metastatic organ. Breast cancer cells isolated from bone or lung metastases prioritize OXPHOS, whereas breast cancer cells from liver metastases primarily exhibit aerobic glycolysis.
Although aerobic glycolysis is a known feature of tumor cells, the metabolic transformation from oxidative phosphorylation to aerobic glycolysis may regulate tumor invasion and distal metastasis by promoting epithelial–mesenchymal transition, tumor angiogenesis and epigenetic modification, but targeting this metabolic pathway has not yet been successfully translated into clinical practice. The significant metabolic heterogeneity and cellular plasticity observed in solid tumors make it unlikely that metabolic inhibitors will be effective as a single therapy for cancer, but combination therapy with two or more drugs that simultaneously inhibit different metabolic pathways can be considered to reduce drug toxicity and prevent recurrence. In addition, aerobic glycolysis-related inhibitors may contribute to the sensitization of tumor cells and improve the treatment outcomes of drug-resistant relapse, providing new treatment or prevention strategies for combating cancer (Fig. 2B).
3.3. Cancer immune microenvironment
The tumor microenvironment consists of tumor-infiltrating monocytes/macrophages, immunosuppressive cells, fibroblasts, vascular system and inflammatory factors secreted by them. Tumor metabolism and tumor immune microenvironment interact with each other. The metabolism of tumor is mainly aerobic glycolysis. The accumulation of lactic acid produced by aerobic glycolysis will cause the tumor immune microenvironment to be in an acidic state. At the same time, the growth of tumor cells requires a lot of energy, which can also cause the tumor immune microenvironment to be in a low oxygen and low energy state. This hypoxic, low-energy, low-pH microenvironment has a significant impact on the human immune system, which can affect the function of T cells, promote the immune escape of tumor cells, and accelerate the occurrence, development and metastasis of tumor cells58. At the same time, the tumor metabolism process will produce immunosuppressive factors, change the function of immune cells, make the tumor microenvironment in an immunosuppressive state, and further accelerate the occurrence of tumors59.
In the natural immune system, macrophages, as highly plastic cells, play different roles in promoting inflammation and anti-inflammation in the body60, the transition of macrophages from a pro-inflammatory state to a repair state is crucial for promoting inflammation and restoring homeostasis. It was found that after LPS activated M1 macrophages, their glycolytic metabolism increased and the concentration of lactic acid increased, and the accumulated lactic acid could be used as a precursor to lead to the lactic acid modification of histone lysine. The results showed that histone lactate played an important role in the homeostasis regulation of M1 macrophages infected by bacteria61, 62, 63. Lactic acid can participate in gene expression regulation through the apparent modification of histone lactate. The high lactic acid produced by aerobic glycolysis is the essential feature of Warburg effect, which suggests that Warburg effect may participate in the occurrence of disease through epigenetic. Therefore, the discovery of histone lactate provides a new idea for Warburg effect to participate in the molecular mechanism of disease occurrence. Studies have observed that lactic acid produced by malignant tumor cells can regulate the activity of dendritic cells, indicating that tumor-derived lactic acid also affects the immune response of anti-tumor T cells64. These studies show that lactic acid molecules play an important role in the process of inflammatory regulation and immune homeostasis through the apparent modification of histone lactate. In addition, the accumulation of extracellular lactate has a significant impact on the metabolism of tumor cells and the transformation of non-tumor cells into tumor cells. The promoting effect of lactic acid on tumors includes blocking the differentiation and activation of monocytes and T cells65. Research has shown that lactic acid can promote the secretion of vascular endothelial growth factor and the polarization of M2 type macrophages through a hypoxia-inducible factor-1α (HIF-1α) dependent mechanism. Additionally, arginase 1 secreted by polarized M2 type macrophages can transmit signals to tumor cells, promoting the growth of tumor cells66,67.
Recently, the success of immunotherapy has attracted researchers' attention to the tumor immune microenvironment68. There is a close relationship between the key glycolytic enzyme PKM2 and inflammation69, 70, 71. PKM2 can play an important role in immune metabolism by enhancing oxidative phosphorylation and Warburg effect, and building a bridge between energy metabolism and tumor immunity. Therefore, a comprehensive understanding of the interaction between key glycolytic enzymes and immune cells will help to find potential targets for tumor therapy and provide a new scheme for immunotherapy combined with small-molecule targeted metabolic therapy. Relevant studies have shown that in the treatment of prostate cancer, increasing the infiltration of M2 macrophages in the tumor immune microenvironment, inhibiting the expression of dendritic cells, and increasing the level of immunosuppressive cytokines can improve the efficacy of immunotherapy72. In addition, in prostate cancer, as the exposure of tumor cells increases during drug treatment, the tumor microenvironment also changes, making the tumor microenvironment further “evolve” to the immunosuppressive environment, which to some extent explains the reason for the low objective response rate of immune treatment of prostate cancer in the published research73. By altering the tumor immune microenvironment to regulate the function of immune cells, we can achieve the goal of anti-tumor treatment, which provides a novel approach to anti-tumor therapy. Therefore, a comprehensive understanding of the interaction between key glycolytic enzymes and immune cells can help identify potential targets for tumor therapy and provide new solutions for immunotherapy combined with small molecule targeted metabolic therapy (Fig. 2C).
3.4. Cancer drug resistance
Tumor resistance seriously affects clinical efficacy and patient prognosis, and is one of the main reasons for the failure of tumor-targeted therapy. The characteristic aerobic glycolysis of tumor cells is one of the important factors affecting targeted drug resistance74. Glucose metabolism and glutamine metabolism are significantly up-regulated in tumor cells, providing cells with material synthesis precursors, energy and redox power, which is consistent with the characteristics of rapid tumor proliferation. Many studies have shown that metabolism is related to overcoming drug resistance of tumor. The change of tumor metabolism can enhance the occurrence of drug resistance. With the problems of adverse reactions and drug resistance of traditional anti-tumor drugs becoming increasingly prominent, tumor metabolism, as a long-standing proposition, has become the focus of researchers again. Standardized chemotherapy, endocrine therapy, targeted therapy and other comprehensive treatment can make most cancer patients get different degrees of remission, but the following drug resistance is an important reason for the failure of comprehensive treatment. Elevated aerobic glycolysis levels can provide ample ATP and NADPH for tumor cells to repair DNA damage induced by anti-tumor drugs, facilitate intracellular drug efflux, and repair oxidative damage. This collectively promotes drug resistance in tumor cells under conditions of high aerobic glycolysis27. Therefore, it is important to study the characteristics of aerobic glycolysis in order to reverse the drug resistance of cancer.
Currently, researchers have confirmed that increasing aerobic glycolysis levels can enhance drug resistance in various tumor cells56,75. Taxus and anthracycline are the most commonly used chemotherapeutic drugs for breast cancer. Simulating the acidic microenvironment resulting from continuous aerobic glycolysis revealed that breast cancer cell MCF-7 exhibited greater resistance to paclitaxel and doxorubicin under acidic conditions (pH 6.6) than under normal culture conditions (pH 7.4)76. The production of ATP and lactic acid of MCF-7 decreased after treatment with aerobic glycolysis inhibitor, and the sensitivity to paclitaxel and doxorubicin increased significantly76. The weak base property of doxorubicin makes it prone to protonation in the extracellular acidic microenvironment formed by aerobic glycolysis, thus hindering the uptake of drugs by tumor cells. The paclitaxel-resistant breast cancer cell line MDA-MB-435 has higher LDH-A expression level and activity than its parent cells. Suppressing the expression of LDH-A can enhance the drug sensitivity in paclitaxel-resistant cells. Furthermore, the combination of paclitaxel and an LDH-A inhibitor demonstrates a more potent apoptotic effect in paclitaxel-resistant cell lines and other breast cancer cell lines77. Endocrine therapy has significantly improved the prognosis of hormone receptor-positive breast cancer patients, but more than 25% of early patients will relapse within 10 years78. The isolated cultured breast cancer cell MCF-7 is sensitive to endocrine drugs, while the co-culture of MCF-7 and fibroblasts will show resistance to tamoxifen and fluvastatin. Further research found that the energy coupling relationship between tumor cells and CAFs, namely the reverse Warburg effect, was the main cause of endocrine drug resistance. Utilizing mitochondrial toxic drugs (such as arsenic trioxide and metformin) or tyrosine kinase inhibitors (such as dasatinib) can disrupt this energy coupling, shifting the energy metabolism of MCF-7 towards the Warburg effect and restoring sensitivity to endocrine therapy79. This also highlights the important role of tumor microenvironment in the process of drug resistance. Trastuzumab has changed the natural course of HER2-positive breast cancer patients, and is one of the most widely used drugs for targeted treatment of breast cancer, but most patients will have acquired trastuzumab resistance after the initial benefits. Trastuzumab can inhibit the growth of tumor cells by inhibiting the HER2-heat shock factor-1–LDH-A pathway and down-regulating the aerobic glycolysis level of tumor cells. The activation of this pathway leads to the activation of aerobic glycolysis, which is one of the mechanisms contributing to the resistance of HER2-positive breast cancer cells to trastuzumab. The combination of trastuzumab and aerobic glycolysis inhibitor can synergistically inhibit tumor growth in both trastuzumab-resistant and sensitive strains, which may be attributed to stronger aerobic glycolysis inhibition80.
Numerous studies have shown that aerobic glycolysis is closely related to the development of tumor-targeted drug resistance. The increased activity of tumor aerobic glycolysis can cause an increase in lactate dehydrogenase levels, leading to a weakened effect of anti-tumor immunosuppressive agents. The transport proteins, key rate-limiting enzymes, and metabolites in the aerobic glycolysis process can affect tumor progression and drug resistance through different mechanisms. The mechanism of tumor cell resistance to targeted drugs is very complex and is the result of the interaction of various factors, but the specific mechanism has not been fully elucidated. More and more evidence suggests that the aerobic glycolysis process of tumor cells is closely related to targeted drug resistance. The abnormal aerobic glycolysis process of tumor cells helps them to rapidly proliferate to adapt to nutritional constraints and form drug resistance phenotypes. In future anti-tumor treatment research, targeted glycolytic drugs can be combined with existing anti-tumor targeted drugs, which may help overcome tumor resistance and have significant implications for improving treatment effectiveness (Fig. 2D).
4. Regulators of Warburg effect
4.1. Direct regulatory enzymes
Different subtypes of certain enzymes can promote Warburg effect in different types of cancer, thus inducing cancer development, including GLUT 1/2/3/4, HK 1/2/3, GPI, phosphofructokinase-fructose-bisphosphatase (PFKFB) 3/4, aldolase A (ALDOA), triosephosphate isomerase (TPI), phosphoglycerate kinase 1 (PGK1), phosphoglyceric acid mutase-1 (PGAM1), pyruvate kinase M2 (PKM2), Lactate dehydrogenase (LDH) A/B, monocarboxylate transporter 1 (MCT-1), hexosephosphate isomerase (HPI), PFK-L, aldolase (ALD)-A, glyceraldehyde-3-phosphate dehydrogenase (GAPDH), enolase (ENO)-α81,82. Interestingly, a small number of enzymes play an anti-cancer role in inhibiting Warburg effect in cancer, including pyruvate dehydrogenase (PDH), succinate dehydrogenase (SDH), 3-hydroxy-3-methylglutaryl CoA synthase-2 (HMGCS2) (Table 1) (Fig. 3).
Table 1.
Direct regulatory enzymes.
| Name | Regulation on Warburg effect | Regulatory mechanism | Cancer type | Ref. |
|---|---|---|---|---|
| Glucose transporter type 1 (GLUT1) | + | Gluts are glucose transporters on the cell membrane which assistant with glucose entering | Renal cell carcinomas | 82 |
| Glucose Transporter 2 (GLUT2) | Pancreatic ductal adenocarcinomas (PDAC) | 89 | ||
| Glucose Transporter 3 (GLUT3) | Hepatocellular carcinoma | 93 | ||
| Glucose Transporter 4 (GLUT4) | Gastric cancer | 98 | ||
| Hexokinase 1 | + | Hexokinases change glucose from a stable state to an active state, producing “glucose-6-phosphate” | Pancreatic cancer | 104 |
| Hexokinase 2 | Glioblastoma multiforme | 112 | ||
| Hexokinase 3 | Colorectal cancer | 120 | ||
| Glucose-6-phosphate isomerase | + | Rearrangement of “glucose-6-phosphate” to “fructose-6-phosphate” | Neuroblastoma | 127 |
| Phosphofructokinase-1(PFK1) | + | Fructose-6-phosphate is catalyzed to produce fructose-1,6-diphosphate | Osteosarcoma; colon cancer; lung cancer | 130 |
| Phosphofructokinase-fructose-bisphosphatase-3 (PFKFB3) | + | PFKFB3 activates PFK1 to regulate aerobic glycolysis | Pancreatic ductal adenocarcinoma | 131 |
| 6-Phosphofructo-2-kinase/fructose-2,6-bisphosphatase 4 (PFKFB4) | + | PFKFB4 catalyzes the kinase reaction that synthesizes F2,6-BP from fructose-6-phosphate (F6P) and ATP, and it can also hydrolyze F2,6-BP into F6P and inorganic phosphate (Pi) through its phosphatase activity | Breast cancer | 132 |
| Aldolase A (ALDOA) | + | Catalyzes fructose-1,6-diphosphate to “glyceraldehyde 3-phosphate” and “dihydroxyacetone phosphate” | Osteosarcoma | 135 |
| Triosephosphate isomerase (TPI) | + | Catalysis of “dihydroxyacetone phosphate” to “glyceraldehyde 3-phosphate” | Colorectal cancer | 137 |
| Glyceraldehyde-3-phosphate dehydrogenase (GAPDH) | + | Two molecules of “glyceraldehyde 3-phosphate” are oxidized by NAD+ and GAPDH to generate “1,3-diphosphoglyceric acid”. | Breast cancer | 138 |
| Phosphoglycerate kinase 1 | + | Phosphoglycerate kinase catalyzes the conversion of 1,3-diphosphoglycerate to 3-phosphoglycerate | Liver cancer | 139 |
| Phosphoglyceric acid mutase-1 (PGAM1) | + | Phosphoglycerate mutase promotes 3-phosphoglycerate to 2-phosphoglycerate | Non-small cell lung cancer (NSCLC) tissues | 143 |
| Pyruvate kinase M2 | + | Pyruvate kinase catalyzes phosphoenolpyruvate to produce a molecule of ATP and pyruvate | Cervical cancer; brain tumour | 147,148 |
| Lactate dehydrogenase A (LDH A) | + | LDH catalyzes the interconversion between pyruvate and lactate. | Breast cancer | 77 |
| Lactate dehydrogenase B (LDH B) | Colorectal adenocarcinoma | 155 | ||
| Pyruvate dehydrogenase (PDH) | – | PDH catalyzes the oxidative decarboxylation of pyruvate to acetyl-CoA | Renal cell carcinoma | 157 |
| Succinate dehydrogenase (SDH) | – | SDH oxidizes succinic acid to fumaric acid | Renal cell carcinoma | 158 |
| 3-Hydroxy-3-methylglutaryl CoA synthase-2 (HMGCS2) | – | HMGCS2 controls the synthesis of ketone body β-hydroxybutyric acid | Colon cancer | 162 |
| CD147 | + | CD147 inhibits the p53-dependent signaling pathway | Hepatic carcinoma | 166 |
Figure 3.
Regulators of the Warburg effect involve lncRNAs, miRNAs, circRNAs, proteins and drugs.
4.1.1. Glucose transporter (GLUT)
The GLUT family, composed of 14 transmembrane proteins encoded by SLC2A gene, is a carrier protein embedded in the cell membrane. Under the condition of not consuming energy, it regulates the extracellular glucose to enter the cell by promoting diffusion along the concentration gradient83. The transformation of energy metabolism mode of tumor cells requires tumor cells to take more glucose to meet the needs of biosynthesis. As a carrier for glucose to enter cells through cell membranes, glucose transporters play an important role in the process of enhanced aerobic glycolysis of malignant tumor cells84. GLUT1, also known as SLC2A1, exists in almost every cell of the human body. It is the most important glucose transporter in brain, nervous system, muscle and other tissues and organs, and is extremely important for maintaining human normal physiological functions. If the function of GLUT1 is completely lost, it will lead to death. Partial loss of function will lead to brain atrophy, mental retardation, growth retardation, epilepsy and other symptoms. On the other hand, GLUT1 also plays an important role in the metabolism of cancer cells. Cancer cells need to consume excessive glucose to maintain their growth and expansion. A significant excess of GLUT1 in cells often means that there is carcinogenesis85.
Whether in a hypoxia or not, the aerobic glycolysis process in tumors may be driven by HIF-1α, which induces the activity of various enzymes responsible for metabolic switching81. HIF-1 increases the expression of a large number of glycolytic enzymes, which are subtypes of glycolytic enzymes found in non-malignant cells. Renal cell carcinoma (RCC) is the most common type of RCC. Because abnormal hypoxia-inducible factor (HIF) stability (HIF-1 or HIF-2) leads to decreased mitochondrial activity, and the inactivation of von Hippel Lindau tumor suppressor gene in RCC, these cells highly rely on glucose uptake mediated by high affinity GLUT1 for aerobic glycolysis and ATP generation, which ultimately drives tumor development82. A study found that compound STF-31 selectively kills RCC that depends on GLUT1 by directly binding to GLUT1 and hindering glucose uptake in the body, ultimately inhibiting the growth of RCC without causing toxicity to normal tissues82. Glucose transporter isoform GLUT2, also known as Slc2a2, is a glucose transporter that is expressed in liver cells, intestine, kidney and central nervous system. Due to its low affinity and high capacity, it has the ability to promote metabolism and provide metabolites that stimulate the transcription of glucose sensitive genes86,87. Some studies have shown that abnormal expression of GLUT2 is closely related to tumorigenesis, and it may eventually cause high expression of GLUT2 in MCF7 and MDA-MB-231 cells and tumor models through functional involvement in fructose metabolism88. The high expression of GLUT2 in pancreatic ductal adenocarcinoma (PDAC) also indicates a poor clinical prognosis. GLUT2 depends on the production of p38γ induced aerobic glycolysis to ultimately promote the growth of PDAC89. GLUT3, which is highly specifically expressed by neurons, has unique characteristics suitable for cell specific expression and function90. Its regulation is an adaptive response, which can prevent cell damage when the metabolic energy of the brain is reduced. As a carrier of glucose entering cells, GLUT3 is one of the important participants in cell glucose metabolism and tumorigenesis, and its high expression is concerned with the low survival rate of tumor patients91,92. GLUT3 can be up-regulated by CUEDC2 (CUE domain containing protein 2) to enhance aerobic glycolysis and ultimately drive cancer progression. It is worth noting that glucocorticoid receptor regulates its transcription by binding to GLUT3 proximal promoter region and plays a key role in CUEDC2 mediated GLUT3 transcription in PLC cells93.
Glucose transporter type 4 (GLUT4) is expressed in fat and muscle, playing a crucial role in systemic glucose homeostasis94. GLUT4 is effectively sequestered in cells during non stimulation. GLUT4 is interfered in insulin resistance and type 2 diabetes, and glucose uptake in muscle and adipose tissue is reduced95. The upregulation of GLUT4 to induce aerobic glycolysis would lead to drug resistance of targeted therapy in cancer. A study found that the N6 methyl adenosine (m6A) demethylase ALKBH5 promotes the m6A demethylation of GLUT4 mRNA, and increases the stability of GLUT4 mRNA in a YTHDF2-dependent manner, thereby enhancing aerobic glycolysis in drug-resistant breast cancer cells. At the same time, the expression of ALKBH5 is significantly up-regulated in HER2 HER2-targeted treatment of drug-resistant breast cancer cells, suggesting that targeting the ALKBH5/GLUT4 axis has therapeutic potential for HER2 targeted treatment of refractory breast cancer patients96. Krüppel-like transcription factor 8 (KLF8), as a GT box (CACCC) binding double transcription factor, is abnormally expressed in several types of human tumors. Its high expression is significantly related to carcinogenic transformation and tumor progression97. KLF8 activates GLUT4 promoter in a dose-dependent manner, and silencing KLF8 expression significantly reduces glucose uptake, lactate secretion and ATP production. GLUT4 may be a potential transcription target of KLF8. KLF8 regulates aerobic glycolysis by targeting GLUT4 and can be used as a new biomarker for tumor survival and potential therapeutic target98.
4.1.2. Hexokinase (HK)
Aerobic glycolysis is a strictly regulated process. Hexokinase is a key enzyme in the glucose metabolism process. It catalyzes the phosphorylation of glucose to 6-phosphate glucose in the first step of aerobic glycolysis, in which HK plays a crucial role. There are five hexokinase isoenzymes found in mammals, HK1, HK2, HK3, HK4 and hexokinase domain containing 199,100. HK1 was mainly distributed in the brain; HK2 was mainly distributed in myocardium, fat and bone; HK3 was mainly distributed in bone marrow, lung and spleen; HK4, also known as glucokinase, regulates insulin secretion in the pancreas and regulates glucose uptake and glycogen synthesis and decomposition in the liver99. The five isoenzymes of hexokinase have specific expression levels in different energy metabolism pathways of different tumor cells. These characteristic metabolic pathways provide possibilities for tumor intervention and treatment101. The most common mutated oncogene in cancer is KRAS. Carcinogenic KRAS changes tumor metabolism by increasing glucose intake and promoting aerobic glycolysis102. KRAS produces two gene products, KRAS4A and KRAS4B, by using the substitute fourth exon103. HK1, as the effector of KRAS4A, can change the activity of HK1 by direct GTP-dependent interaction with KRAS4A104. Since HK1 is a glycolytic enzyme controlling the pathway flow, KRAS4A will have a significant impact on the direct regulation of HK1104. Relevant studies have designed HK1 inhibitors to target KRAS-mediated metabolic reprogramming105. At present, it is generally believed that HK2 has a dual role in tumor cells: one is to induce aerobic glycolysis, and the level of cell aerobic glycolysis is positively correlated with the expression and activity of HK2; The other is to inhibit apoptosis by binding with the volt-dependent anion channel (VDAC) on the outer membrane of mitochondria106. Both HK1 and HK2 can directly bind to VDAC. As a mediator of the release of apoptosis promoting proteins, VDAC1 oligomerization reconstructs the outer membrane of mitochondria to change its permeability so that the channel remains closed, and inhibits the release of cytochrome c to inhibit mitochondrial-mediated apoptosis107,108. HK2 is hardly expressed in normal tissues, but highly expressed in many tumor tissues109, 110, 111. The expression level of HK2 is often related to the rapid proliferation of tumor cells and the prognosis of tumor patients. Interestingly, glioblastoma (GBM) showed an increase in HK2 expression, therapeutic drug resistance and intracranial proliferation compared with normal brain and low-grade glioma, which mainly expressed HK1, and was associated with poor overall survival of GBM patients. Intracranial xenografts of HK2 knockout GBM cells showed decreased proliferation and angiogenesis. It is suggested that targeting HK2 may interfere with the growth and therapeutic sensitivity of GBM112. Studies have shown that when HK2 from tumor cells, which is bound to mitochondria, is introduced into normal liver cells, it significantly increases the aerobic glycolysis rate, the binding ratio of HK2 to mitochondria, and the affinity of mitochondrial-bound HK2 for the substrate. By capturing ATP released by mitochondria, it becomes one of the mechanisms of tumor high efficiency aerobic glycolysis113. In most tumor cells, HK2 protein and its activity significantly increased, ensuring the rapid aerobic glycolysis of tumor cells. The high expression of HK2 protein may be related to gene amplification, expression and regulation. It has been found that activation of HK2 promoter, increase of gene copy number and hypomethylation are important mechanisms of overexpression of HK2 in tumor cells114. Hypoxic conditions, cyclicadenosine monophosphate (cAMP) and p53 can enhance the transcription of HK2 gene115. Demethylation of HK-related regulatory genes with deoxycytidine and DNA demethylase can increase the expression of HK2 mRNA and protein in rat hepatocytes, indicating that the stability of HK-related regulatory genes can affect the expression of HK2116. Insulin can induce HK2 gene transcription of adipocytes and muscle cells, increase mRNA and promote protein synthesis117. The methylation degree of HK2 gene promoter was significantly different between high glycolytic tumor cells and normal cells. There were 18 CpG sites methylation of HK2 gene in normal liver cells, but no methylation of these sites was found in tumor cells, that is, low methylation was closely related to high expression of HK2 in tumor cells116. HK1 and HK2 both have an N-terminal hydrophobic 15 amino acid sequence, which is compatible with amphoteric α-helix and can bind to the outer membrane of mitochondria. HK3 and HK4 lack this sequence and cannot bind to the outer membrane of mitochondria118. Compared with other HK isoenzymes, HK3 has lower protein expression and higher glucose affinity. Compared with normal tissues, the expression of HK3 in colorectal cancer tissues was up-regulated and positively correlated with some metastasis-related genes119. LPS, as a classical inflammatory body activator, can activate inflammatory bodies in cancer cells to enhance metastasis. Its enhanced cancer cell movement depends on the increase of glucose uptake and aerobic glycolysis120. In the presence of LPS, NF-κB activated inflammatory bodies upregulate the expression of nuclear Snail, forming a protein complex with Snail, which binds to the promoter region of HK3 to enhance aerobic glycolysis120. HK3 was significantly up-regulated under LPS treatment, which could protect cells from oxidant-induced death and increase ATP production under hypoxia121. In addition, another key enzyme of aerobic glycolysis, GPI, also performs important functions in cancer. Its inhibition will hinder aerobic glycolysis and activate oxidative phosphorylation. GPI is also called glucose phosphate isomerase or hexose phosphate isomerase, and can also catalyze the exchange of furan isomers of α and β of glucose 6-phosphate. GPI also has the activity of cell division and growth factors in vitro, which has the same effect as autocrine motor factor (AMF)122,123. Overexpression of GPI/AMF is related to the aggressive phenotype, increased mortality and poor prognosis of many cancer types124, 125, 126. Valproic acid, as an inhibitor of histone deacetylase (HDAC), can inhibit the progress of neuroblastoma (NB) by inhibiting E2F transcription factor 1 (E2F1)/GPI signal pathway and ultimately blocking Warburg effect127.
4.1.3. Phosphofructokinase (PFK)
The second step of the aerobic glycolysis pathway is that fructose 6-phosphate and ATP are converted into fructose 1,6-diphosphate and ADP under the action of 6-PFK-1, which is also the second irreversible reaction of the aerobic glycolysis pathway. PFK includes two subtypes: PFK-1 and PFK-2. PFK-1 is the most important rate limiting enzyme in the aerobic glycolysis pathway. It is a tetramer, which is regulated by the allosteric regulation of fructose 1,6-diphosphate, ADP, AMP, fructose 2,6-diphosphate, ATP and citric acid. PFK-2 is a bifunctional enzyme with two independent catalytic centers in the enzyme protein, which can act as 6-phosphate fructose kinase 2 to catalyze the phosphorylation of 6-phosphate fructose C2 to form fructose 2,6-diphosphate, and also act as fructose diphosphatase-2 to hydrolyze fructose 2,6-diphosphate C2 to 6-phosphate fructose, thus completing the mutual conversion of fructose 6-phosphate and fructose 2,6-diphosphate. And then realize the regulation of PFK1128. Selective inhibition of PFK-1 activity can prevent the proliferation of gastric cancer (GC) cells and inhibit the growth of carcinoma in situ129. As a structural homolog of p53, the up-regulation of TAp73 is related to the higher expression of PFKL in tumor cells. By activating the expression of PFK-1/PFK-1, it can enhance glucose consumption and lactate excretion, increase ATP production and enhance antioxidant defense, and promote Warburg effect, which has established that TAp73 may be the key regulator of aerobic glycolysis130. Phosphate fructose kinase-fructose diphosphatase-3 (PFKFB3), as a glycolytic driver, can also activate the key rate-limiting enzyme PFK-1. It has the highest kinase activity in the PFKFB family and is overexpressed in PDAC. It may provide local ATP supply in the plasma membrane of PDAC cells to maintain the plasma membrane calcium ATPase function. PFKFB3 inhibitor PFK15 causes cytotoxic calcium overload by inhibiting the function of PMCA, and ultimately leads to cell death. It is suggested that targeted aerobic glycolysis may be a feasible treatment for PDAC131. Another regulatory enzyme, fructose-6-phosphate 2-kinase/fructose-2,6-diphosphatase 4 (PFKFB4), which synthesizes glycolytic allosteric stimulators, up-regulates the expression of transketolase in S857 phosphorylated steroid receptor coactivator-3 (SRC-3) to drive glucose flux to purine synthesis. Silencing SRC-3 or PFKFB4 can inhibit the growth and metastasis of breast tumors in vivo, suggesting that targeting the PFKFB4–SRC-3 axis may have therapeutic value in breast tumors that are significantly dependent on glucose metabolism132.
4.1.4. Others
Fructose-diphosphate aldolase A (ALDOA) is highly expressed in various types of cancer as a glycolytic enzyme133,134, especially osteosarcoma (OS)135. ALDOA competes with miR-34c-5p and reduces the inhibitory effect of miR-34c-5p on ALDOA, resulting in increased ALDOA expression and aerobic glycolysis. The miR-34c-5p/ALDOA axis may provide a new therapeutic target for OS treatment135. TPI can catalyze the mutual conversion of dihydroxyacetone phosphate (DHAP) and glyceraldehyde-3-phosphate and is highly expressed in tissues with increased aerobic glycolysis ability136,137. The regulatory characteristics of GAPDH are the most different in the Warburg effect process. After metabolic control analysis based on the aerobic glycolysis mathematical model, GAPDH showed high computational flux control coefficient and low reaction free energy, suggesting that GAPDH may be involved in controlling the rate of aerobic glycolysis. The natural product koningic acid, as a specific GAPDH inhibitor, showed its ability to dynamically affect aerobic glycolysis in BT-474 tumor bearing mice, The upstream glycolytic metabolites F 1,6-BP and DHAP accumulated to their peak within 500–1000 min, while the downstream glycolytic metabolites 3 PG and PEP were depleted within the first 10 min and subsequently recovered. At the same time, the tumor size of BT-474 tumor bearing mice was significantly inhibited after KA treatment, suggesting that KA can induce the acute dynamic changes of aerobic glycolysis network in tumors, which is the main way to restrain the proliferation of tumors of breast cancer by influencing aerobic glycolysis mechanism138. PGK1 is the first enzyme to produce ATP in aerobic glycolysis, which is mediated by different post-translational modifications. O-GlcNAc is a universal post-translational modification of protein serine and threonine residues. Compared with adjacent matched tissues, O-GlcNAcyclization of PGK1 is significantly higher in human colon cancer tissues. Blocking the acylation of T255 O-GlcNA on PGK1 can inhibit the Warburg effect, reduce the proliferation of colon cancer cells, and inhibit the growth of tumor in nude mice139 Phosphate glycerate mutase catalyzes the conversion of 3-phosphate glyceride (3-pG) to 2-phosphate glyceride (2-PG) in the late stage of aerobic glycolysis, and its expression can be observed to increase in different cancers140, 141, 142. In non-small cell lung cancer (NSCLC) tissues, mTOR, as a positive regulator of Warburg effect, stimulates the expression of downstream effector PGAM1 through transcriptional activation mediated by HIF-1α, and blocks PGAM1's inhibition of mTOR-dependent aerobic glycolysis, which can hinder the occurrence and proliferation of NSCLC143. There are four pyruvate kinase subtypes that regulate the final step limit of aerobic glycolysis in mammals136,144, while tumor tissues only express the embryonic M2 subtype of pyruvate kinase (PKM2)145,146. Knocking down PKM2 expression will reduce the production of lactic acid and increase the consumption of oxygen, and ultimately lead to the reversal of Warburg effect147. The phosphorylation level of PKM2 S37 is related to the activity of EGFR and ERK1/2 in human GBM samples. Replacing the wild-type PKM2 with a nuclear translocation defect mutant (S37A) can block the Warburg effect promoted by EGFR and the development of brain tumors148. It is suggested that PKM2 expression is necessary for aerobic glycolysis. In the cytoplasm of tumor cells, lactate dehydrogenase (LDH) catalyzes the conversion between pyruvate and lactic acid. The tetramer LDH consists of two subunits encoded by independent genes: LDHA and LDHB149,150. LDHA controls the conversion of pyruvate to lactic acid during cell aerobic glycolysis in muscle or liver151. The promoted expression and activity of LDH-A in paclitaxel-resistant cells are directly related to their sensitivity to oxalate, an aerobic glycolysis inhibitor. The inhibition of LDH-A makes paclitaxel-resistant cells re-sensitive to paclitaxel77, indicating that LDH-A may be a promising therapeutic target to overcome paclitaxel resistance. Lactate dehydrogenase B controls the conversion of lactic acid to pyruvate in the process of cell aerobic glycolysis in the heart and brain150,152. The phosphorylation of LDHB serine 162 mediated by the conservative serine/threonine kinase Aurora-A, which is responsible for the centrosome maturation in G2 and the formation of bipolar spindle in mitosis153,154, significantly increases its activity of reducing pyruvate to lactic acid, thus promoting NAD regeneration, aerobic glycolysis flux and biosynthesis of glycolytic metabolites, so as to promote tumor progression155. Blocking S162 phosphorylation by LDHB-S162A mutant can inhibit aerobic glycolysis and tumor growth in cancer cells and xenotransplantation models155. It reveals the interesting mechanism of LDHB in aerobic glycolysis regulation and tumor progression. Metabolic changes are particularly prominent in RCC, especially in clear cell renal cell carcinoma (ccRCC)156. Compared to its matched adjacent kidney tissue, ccRCC tumor exhibited enhanced aerobic glycolysis, inhibited PDH activity, and displayed invasive tumor characteristics such as minimal glucose oxidation and turnover of TCA cycle in vivo, which provided strong evidence for the Warburg effect in human tumors for the first time157. SDH, as the first TCA cycle enzyme with tumor inhibition characteristics, is responsible for the oxidation of succinate to fumarate in the TCA cycle, and sending electrons into the mitochondrial respiratory chain to produce ATP (complex II). Silencing SDH can maintain the maximum glycolytic flux by consuming extracellular pyruvate, thereby preserving Warburg-like bioenergy characteristics, indicating the metabolic vulnerability of SDH-related malignant tumors in the future158. HMGCS2 is a rate-limiting enzyme in the pathway of ketone body production, which can lead to the production of ketone bodies including β-hydroxybutyrate and participate in the metabolic reprogramming of tumor cells159,160. When the expression of HMGCS2 is reduced, the ketone body production is reduced, which greatly improves the utilization rate of tumor cells for energy in vivo, and promotes the metastasis and progression of tumor149,161,162. CD147 is a transmembrane protein that is over-expressed on the surface of various malignant cells163. Its silencing significantly reduces the aerobic glycolysis rate and lactic acid efflux in cancer cell lines, indicating that CD147 is involved in tumor aerobic glycolysis164,165. CD147 activates the PI3K/Akt signal pathway in Hepatocellular carcinoma (HCC) cells through lactic acid output mediated by monocarboxylate transporter 1 (MCT1), up-regulates GLUT1 and down-regulates TIGAR, PGC1a/TFAM and p53R2, and inhibits mitochondrial biogenesis and oxidative phosphorylation in a p53 dependent manner166. After blocking CD147 and/or down-regulating MCT1, glucose metabolism decreased and the growth of HCC cells was inhibited166.
The process of cancer are affected by various factors. GLUTs, key aerobic glycolysis enzymes (such as HKs, PFKs, PKs) and lactic acid production enzymes (such as LDHs) regulate the glucose metabolism in the body and supply energy for human cells. Abnormal enhancement of aerobic glycolysis is one of the important features in tumor growth, during which GLUTs and key glycolytic enzymes participate in energy metabolism of tumor cells, regulating their growth, infiltration, and invasion. Studying the possible effects of changes in key glycolytic enzymes in tumor cells on cell metabolism, revealing their roles and molecular mechanisms in tumor occurrence and development, not only enriches research ideas on tumor related diseases, but also provides new ideas and treatment targets for early diagnosis and treatment of cancer. At the same time, it can also provide new ideas for discovering new tumor treatment drugs.
4.2. Transcription factors and transcriptional co-regulators
More than ten genes encoding glycolytic enzymes directly lead to Warburg effect, and transcription factors play a direct role in regulating Warburg effect150,167,168 (Table 2). Transcription factor sine oculis homeobox 1 (SIX1) is responsible for regulating the occurrence of body organs169,170. Its overexpression occurs in various types of cancer and is positively correlated with poor prognosis170, 171, 172. The up-regulation of SIX1 is usually accompanied by the down-regulation of microRNA-548a-3p. A study found that SIX1 promotes aerobic glycolysis through HBO1 and AIB1 histone acetyltransferase, and is directly inhibited by microRNA-548a-3p173. It indicates that targeting microRNA-548a-3p/SIX1 axis may be a new strategy for cancer treatment. Transcription factor Yin Yang 1 (YY1) is a multifunctional transcription factor protein that can activate or inhibit genes174, 175, 176. It is overexpressed in prostate cancer. By directly binding to and activating the gene PFKP encoding glycolytic rate-limiting enzyme, YY1 ultimately enhances the Warburg effect and promotes the malignant growth of prostate cancer177. The expression of aerobic glycolysis related-genes SLC2A1 and HK1 is positively correlated with the expression of RUNX1 mRNA. RUNX1 and HK1 serve as adverse prognostic factors for PDAC patients. Overexpression of RUNX1 can increase lactate secretion, glucose uptake, intracellular ATP levels, induce the Warburg effect, and promote tumor growth and metastasis178. Fork-head box protein M1 (FOXM1), as a carcinogenic transcription factor of the forkhead transcription factor superfamily, also participates in the regulation of metabolism in cancer179,180. It can increase the activity of lactate dehydrogenase, lactate production and glucose utilization by up-regulating the expression of LDHA, and ultimately promote the occurrence and metastasis of pancreatic adenocarcinoma181. As a tumor suppressor, p53 participates in metabolism regulation as an important part of p53 response, which not only helps to maintain the dynamic balance of normal cell metabolism, but also helps to control the development of cancer182. Transient metabolic stress, such as the fluctuation of available oxygen and nutrients, will also trigger more adaptive responses involving p53. In this process, p53 induces metabolic remodeling and promotes catabolism, while coordinating to reduce cell proliferation and growth183,184. P53 can inhibit Warburg effect by reducing aerobic glycolysis and promoting oxidative phosphorylation through multiple mechanisms185. The fructose phosphate enzyme, which promotes the third step of aerobic glycolysis pathway, is regulated by various metabolites in the glucose metabolism mechanism186. Its metabolites include ATP, citrate and lactic acid. The fructose phosphate enzyme can directly inhibit PFK1, and PFK1 can be activated by AMP and fructose-2,6-diphosphate. P53 plays a key regulatory role in this signal pathway, that is, p53 reduces the rate of aerobic glycolysis by up-regulating the expression of TIGAR, promotes aerobic glycolysis and transfers its intermediate product to the predetermined pentose phosphate pathway. In addition, p53 has an inhibitory effect on glycolytic enzymes in this pathway187,188. In fibroblasts, p53 not only down-regulates the phosphoglyceride mutase that promotes the conversion of glycerol 3-phosphate to glycerol 2-phosphate189. In addition, it also has a negative regulatory effect on the expression of pyruvate dehydrogenase 2, which reduces the activity of pyruvate dehydrogenase and further prevents the conversion of pyruvate to acetyl coenzyme A190. The co-loss of p53 and p10 will promote tumor development by increasing the selective up-regulation of HK2-mediated aerobic glycolysis by hexokinase HK2191. NF-κB is a multi-member nuclear transcription factor that can participate in a variety of biological processes, including cell proliferation and differentiation, immune and inflammatory reactions. The NF-κB protein family consists of five members: ρ50 (NF-κB1), p52 (NF-κB2), p65NF-κB (RelA), c-Rel (Rel) and RelB, among which p65NF-κB (RelA) is the most widely studied member of the NF-κB family. NF-κB is also an important cancer-promoting factor. NF-κB activation can be observed in various types of tumors192, 193, 194, 195, 196. A variety of carcinogenic factors such as tobacco, alcohol and radiation can cause NF-κB activation, further up-regulate a series of inflammatory factors, and promote cell proliferation and malignant transformation197. As an important transcription factor in the process of epithelial-mesenchymal transformation, Snail can up-regulate cellular aerobic glycolysis, thus promoting the occurrence and development of tumors119,198,199, and it is positively correlated with the expression of the key enzyme HK3 in the first step of aerobic glycolysis and clinical adverse prognosis199. Transcription factor NF-κB can directly bind to snail promoter, start snail transcription and up-regulate its protein expression200. Some studies have shown that Metadherin (MTDH), as an oncogene, can accelerate the progression of colorectal cancer (CRC) by up-regulating the expression of NF-κB p65 and snail, inducing aerobic glycolysis. Knocking down MTDH gene expression may inhibit the occurrence and development of CRC201. ATF4 is a gene expression regulator that can respond to various forms of stress202, and its existence plays a role in the resistance targeted by LDHA. After being knocked down, ATF4 makes melanoma cells sensitive to LDHA inhibitors and inhibits aerobic glycolysis, and finally inhibits tumor cell proliferation203. Peroxisome proliferator-activated receptor γ coactivator-1α (PPARGC1A) is a central regulator of mitochondrial metabolism204. It can up-regulate the expression of pyruvate dehydrogenase E1α 1 subunit and mitochondrial pyruvate vector 1 to induce the Warburg effect, promote the metastatic transmission of cholangiocarcinoma (CCA) cells, suggesting that blocking the PPARGC1A signal axis may inhibit the metastasis of CCA205. A new tumor suppressor, CAB39L, can also inhibit tumor occurrence by promoting LKB1–AMPK–PPARGC1A axis, thus preventing the metabolic transformation driving carcinogenesis206.
Table 2.
Transcription factors and transcriptional co-regulators.
| Name | Classification | Regulation on Warburg effect | Regulatory mechanism | Cancer type | Ref. |
|---|---|---|---|---|---|
| Six1 | Tumor promoter | + | Six1 promotes HBO1 and AIB1 histone acetyltransferases activation thus to increases the expression of many glycolytic genes | Breast cancer | 173 |
| Yin yang 1 (yy1) | Tumor promoter | + | Yy1 directly binds and activates PFKP to stimulate aerobic glycolysis | Prostate cancer | 177 |
| Runx1 | Tumor promoter | + | Runx1 binds to the promoters of SLC2A1 and HK1 to up-regulate their expression to enhance aerobic glycolysis | Pancreatic ductal adenocarcinoma | 178 |
| Forkhead box protein M1 (FOXM1) | Tumor promoter | + | FOXM1enhances aerobic glycolysis via transcriptional regulation of LDHA expression | Pancreatic cancer | 181 |
| P53 | Tumor suppressor | – | 1. p53 inhibits the expression of glycolytic genes, such as HK2, PGM and GLUT; 2. p53 induces expression of TIGAR (TP53-induced aerobic glycolysis and apoptosis regulator) to decrease fructose-2,6-bisphosphate levels to inhibit aerobic glycolysis |
185 | |
| NF-κB | Tumor promoter | + | MTDH up-regulates the expression of NF-κB p65 and snail, induces the Warburg effect | Colorectal cancer | 201 |
| Atf4 | Tumor promoter | + | ATF4 regulates SLC1A5 to increase uptake of glutamine and essential amino acid, thus to activate mTORC1-regulated aerobic glycolysis | Malignant melanoma | 203 |
| Steroid receptor coactivator-3 (SRC-3) | Oncogenic transcriptional coregulator | + | The phosphorylation of SRC-3 by PFKFB4 enhances the expression of transketolase, adenosine monophosphate deaminase 1 (AMPD1), and xanthine dehydrogenase to contribute to the Warburg effect | Breast cancer | 132 |
| Krüppel-like transcription factor 8 (KLF8) | Tumor promoter | + | KLF8 activated the GLUT4 promoter activity to enhance GLUT4 expression | Gastric cancer | 98 |
| Peroxisome proliferator-activated receptor gamma coactivator 1α (PPARGC1A) | Tumor promoter | – | PPARGC1A enhances pyruvate oxidation metabolism through pyruvate dehydrogenase E1 alpha 1 subunit and mitochondrial pyruvate carrier up-regulation to reverse the Warburg effect | Cholangiocarcinoma (CCA). | 205 |
| Tumour suppressive transcriptional co-activator | – | Gastric cancer | 206 |
In recent years, it has been found that the metabolic reprogramming of tumor cells is regulated by many different factors, and the regulation of transcription factors on tumor cell metabolic genes is one of the main mechanisms of tumor cell metabolic reprogramming, among which c-Myc, HIF-1, p53, Fox are the most studied in tumor metabolic Reprogramming. This manuscript reviews the regulatory mechanisms of c-Myc, HIF-1, p53, Fox on aerobic glycolysis with a comprehensive view to providing ideas for understanding tumor metabolic reprogramming and its related molecular mechanisms.
4.3. LncRNAs, miRNAs and circular RNAs
The molecular mechanism of cancer pathogenesis is gradually clarified, and its molecular changes mainly include: abnormal expression of lncRNAs, miRNAs and circular RNAs, gene mutation, gene amplification, gene translocation, and epigenetic changes. Among them, lncRNAs, miRNAs and circular RNAs play an important role in the occurrence and development of cancer and are of great value in the diagnosis, assessment of disease progression and prognosis judgment of cancer207. In this review, we summarized some LncRNAs, miRNAs and circular RNAs that affect the research program of Warburg Effect in the treatment of cancer (Table 3) (Fig. 3).
Table 3.
LncRNAs, miRNAs and Circular RNAs.
| Name | Classification | Regulation on Warburg effect | Regulatory mechanism | Cancer type | Ref. |
|---|---|---|---|---|---|
| KCNQ1OT1 | Oncogenic lncRNA | + | KCNQ1OT1 sponges miR-34c-5p to increase ALDOA expression to contributed to the Warburg effect | Osteosarcoma | 135 |
| miR-34c-5p | Tumor-suppressive miRNA | – | miR-34c-5p inhibits ALDOA expression by directly targeting its 3′UTR. | Osteosarcoma | 135 |
| miR-30d | Tumor-suppressive miRNA | – | miR-30d suppresses aerobic glycolysis by inhibits RUNX1/SLC2A1/HK1 to suppress aerobic glycolysis | Pancreatic ductal adenocarcinoma | 178 |
| LINC00926 | Tumor-suppressive lncRNA | – | LINC00926 negatively regulates PGK1 expression to inhibit aerobic glycolysis | Breast cancer | 211 |
| MIR210HG | Oncogenic lncRNA | + | MIR210HG potentiates the translation of HIF-1α, thus to increase HIF-1α protein level to up-regulate the expression of glycolytic enzymes. | Triple-negative breast cancer (TNBC). | 212 |
| ABHD11-AS1 | Oncogenic lncRNA | + | ABHD11-AS1 increases the occupation of EZH2 and H3K27me3 on KLF4 promoter region to enhance KLF4 expression to promoted the Warburg effect | Non-small-cell lung cancer | 213 |
| HULC | Oncogenic lncRNA | + | HULC directly binds to two glycolytic enzymes, LDHA and pyruvate kinase M2 to promote aerobic glycolysis | Liver cancer | 215 |
| KCNQ1OT1 | Oncogenic lncRNA | + | KCNQ1OT1 directly binds and stabilizing HK2 to promote aerobic glycolysis | Colorectal cancer | 218 |
| SNHG14 | Oncogenic lncRNA | + | SNHG14 increases IRF6 mRNA degradation and inhibits the expression of IRF6 to encourage aerobic glycolysis | Glioma | 222 |
| SOX2OT | Oncogenic lncRNA | + | SOX2OT promotes the expression of PKM2 to promote aerobic glycolysis | Hepatocellular carcinoma (HCC) | 229 |
| NF-kappa B interacting long noncoding RNA (NKILA) | Oncogenic lncRNA | + | NKILA increases the expression of HIF-1α to enhance the Warburg effect | Glioma | 225 |
| IDH1-AS1 | Oncogenic lncRNA | + | IDH1-AS1 promotes c-Myc collaborate with HIF1α to activate the Warburg effect | Cervical cancer | 226 |
| lincRNA-p21 | Oncogenic lncRNA | + | LincRNA-p21 promotes HIF-1α accumulation to enhance aerobic glycolysis | Cervical cancer | 227 |
| EPB41L4A-AS1 | Tumor-suppressive lncRNA | – | EPB41L4A-AS1 interacts and co-localizes with HDAC2 in nucleolus and inhibits the released HDAC2 from nucleolus to hinder aerobic glycolysis | Cervical cancer | 228 |
| MALAT1 | Oncogenic lncRNA | + | MALAT1 activates the mTORC1–4EBP1 axis to reprogram the tumor glucose metabolism | Hepatocellular carcinoma | 230 |
| miR-124 | Tumor-suppressive miRNA | – | Targets polypyrimidine tract-binding protein 1 to block the PKM1/PKM2 regulated aerobic glycolysis | Colorectal cancer | 233 |
| hsa-miR-144-5p | Tumor-suppressive miRNA | – | hsa-miR-144-5p inhibits GLUT-1 expression | Renal cell carcinoma | 235 |
| hsa-miR-186-3p | Oncogenic miRNA | + | hsa-miR-186-3p increases GLUT-1 expression | Renal cell carcinoma | 235 |
| miR-16-1-3p | Tumor-suppressive miRNA | – | miR-16-1-3p suppresses aerobic glycolysis via inhibition of PGK1 | Breast cancer | 239 |
| let-7a-5p | Tumor-suppressive miRNA | – | let-7a-5p inhibits GLUT12 expression to suppress aerobic glycolysis | TNBC | 244 |
| miR-378∗ | Oncogenic miRNA | + | miR-378∗ inhibits the expression of ERRγ and GABPA to increase aerobic glycolysis | Breast cancer | 245 |
| miR-30a-5p | Tumor-suppressive miRNA | – | miR-30a-5p inhibits LDHA expression to disturb aerobic glycolysis | Breast cancer | 246 |
| miR-34a | Tumor-suppressive miRNA | – | miR-34a inhibits LDHA expression to disturb aerobic glycolysis | Cervical cancer | 247 |
| miRNA-27b | Oncogenic miRNA | + | miR-27b inhibits pyruvate dehydrogenase protein X to contribute to the Warburg effect | Breast cancer | 252 |
| miR-155 | Oncogenic miRNA | + | miR-155 upregulates HK2 to contribute to Warburg effect | Breast cancer | 253 |
| miR-210-3p | Oncogenic miRNA | + | miR-210-3p contributes to maintain HIF-1α stabilization and suppressed p53 activity to strengthen the Warburg effect | TNBC | 255 |
| miR-214 | Oncogenic miRNA | + | miR-214 targets the adenosine A2A receptor (A2AR) and PR/SET domain 16 to enhance the Warburg effect | Gastric cancer | 261 |
| miR-3662 | Tumor-suppressive miRNA | – | miR-3662 dampened aerobic glycolysis by targeting HIF-1α | Hepatocellular carcinoma | 262 |
| miR-885-5p | Tumor-suppressive miRNA | – | miR-885-5p directly targets the 3′ UTR of HK2 to inhibit its expression | Hepatocellular carcinoma | 263 |
| miR-532-3p | Tumor-suppressive miRNA | – | miR-532-3p inhibits aerobic glycolysis by targeting HK2 | Ovarian cancer | 264 |
| miR-199a | Tumor-suppressive miRNA | – | miR-199a limits aerobic glycolysis by targeting HK2 and PKM2 | Hepatocellular carcinoma | 265, 266 |
| miR-23a | Oncogenic miRNA | – | miR-23a directly targets PGC-1α and G6PC to decrease the glucose production | Hepatocellular carcinoma | 267 |
| miR-338-3p | Tumor-suppressive miRNA | – | miR-338-3p suppresses the Warburg effect via PKLR inhibition | Hepatocellular carcinoma | 268 |
| miR-23a/b | Oncogenic miRNA | + | miR-23a/b controls the mitochondrial glutaminase expression to regulate the Warburg effect | Prostate cancer | 272 |
| miR-145 | Tumor-suppressive miRNA | – | miR-145 enhance the c-myc/DNMT3A/miR-133b/PKM2 pathway to inhibit the Warburg effect | Ovarian cancer | 273 |
| miR-133b | Tumor-suppressive miRNA | – | miR-133b inhibits its target gene PKM2 expression to impede the Warburg effect | Ovarian cancer | 273 |
| miR-644a | Tumor-suppressive miRNA | – | miR-644a directly targets c-Myc, Akt, IGF1R, and GAPDH to inhibit their expression thus to suppress the Warburg effect | Prostate cancer | 274 |
| miR-34c | Tumor-suppressive miRNA | – | miR-34c suppresses c-Myc expression to inhibit the Warburg effect | Glioblastoma | 275 |
| let-7 | Tumor-suppressive miRNA | – | let-7 inhibits PDK1 protein expression via post-transcriptional regulation | Hepatoma | 282 |
| circRNF20 | Oncogenic Circular RNA | + | circRNF20 acts as the sponge of miR-487a to increase HIF-1α expression | Breast cancer | 286 |
| miR-487a | Tumor-suppressive miRNA | – | miR-487a inhibits the Warburg effect through HIF-1α-HK2 inhibition. | Breast cancer | 286 |
| circRNA_100290 | Oncogenic Circular RNA | + | circRNA_100290 sets GLUT1 free from miR-378a-mediated GLUT1 inhibition | Oral squamous cell carcinoma | 288 |
| miR-378a | Tumor-suppressive miRNA | – | miR-378a inhibits GLUT1 expression to obstruct aerobic glycolysis | Oral squamous cell carcinoma | 288 |
| circNRIP1 | Oncogenic Circular RNA | + | circNRIP1 acts as a microRNA-149-5p sponge to activate AKT1/mTOR pathway | Gastric cancer | 293 |
| miR-149-5p | Tumor-suppressive miRNA | – | miR-149-5p inhibits the AKT1/mTOR pathway to suppress aerobic glycolysis | Gastric cancer | 293 |
| circVPS33B | Oncogenic Circular RNA | + | circVPS33B accelerates the Warburg effect via miR-873-5p/HNRNPK axis | Gastric cancer | 298 |
| miR-873-5p | Tumor-suppressive miRNA | – | miR-873-5p inhibits HNRNPK expression to suppress aerobic glycolysis | Gastric cancer | 298 |
| circATP2B1 | Oncogenic Circular RNA | + | circATP2B1 accelerates aerobic glycolysis via miR-326 thus to activate pyruvate kinase M2 | Gastric cancer | 301 |
| miR-326 | Tumor-suppressive miRNA | – | miR-326 decreases the expression of PKM2 to bother aerobic glycolysis | Gastric cancer | 301 |
| circCUL3 | Oncogenic Circular RNA | + | circCUL3 sponges miR-515-5p to activate STAT3–HK2 pathway | Gastric cancer | 302 |
| miR-515-5p | Tumor-suppressive miRNA | – | miR-515-5p decreases STAT3 expression thus to inhibit HK2 expression | Gastric cancer | 302 |
| circFOXP1 | Oncogenic Circular RNA | + | circFOXP1 increases PKLR mRNA level to enhance the Warburg effect | Gallbladder cancer | 305 |
| miR-370 | Tumor-suppressive miRNA | – | miR-370 decreases PKLR mRNA level | Gallbladder cancer | 305 |
| circECE1 | Oncogenic Circular RNA | + | CircECE1 interacts with c-Myc to prevent c-Myc ubiquitination and degradation | Osteosarcoma | 308 |
| miR-324-5p | Tumor-suppressive miRNA | – | miR-324-5p decreases PKM2 expression to enhance the Warburg effect | Ovarian cancer | 317 |
| lncRNA-NEAT1 | Oncogenic lncRNA | + | lncRNA-NEAT1 inhibits miR-34a to promote LDHA-aerobic glycolysis axis | Cervical cancer | 324 |
| miR-34a | Tumor-suppressive miRNA | – | miR-34a inhibits LDHA activity to impede aerobic glycolysis | Cervical cancer | 324 |
| UCA1 | Oncogenic lncRNA | + | UCA1 elevates the activation of HK2 oncogenes via inhibition of miR-203 activity | Esophagal cancer | 328 |
| miR-203 | Tumor-suppressive miRNA | – | miR-203 suppress HK2 expression to inhibit glucose uptake | Esophagal cancer | 328 |
4.3.1. LncRNAs
LncRNAs refer to non-coding RNAs with a length greater than 200 nucleotides. According to the relationship with protein-coding genes, lncRNAs can be divided into five categories: sense lncRNAs, antisense lncRNAs, bidirectional lncRNAs, intragenic lncRNAs and intergenic lncRNAs208. LncRNAs are involved in almost all aspects of gene regulation, such as gene imprinting, epigenetic regulation, transport between nucleus and cytoplasm, splicing and translation of mRNA208. So far, more than 3000 lncRNAs have been confirmed to participate in a variety of biological processes, such as chromatin imprinting, cell differentiation and tumor aerobic glycolysis209, among which the role in tumor genesis and development includes the regulation of tumor cell proliferation, invasion and metastasis210
Homo sapiens long intergenic non-protein coding RNA 00926 (LINC00926), an lncRNA producing 2.53 kb transcript, negatively regulates the expression of PGK1 by enhancing the ubiquitination of PGK1 mediated by E3 ligase STUB1, inhibits aerobic glycolysis, tumor growth and lung metastasis of breast cancer in vitro and in vivo, and predicts good clinical results of breast cancer211. It suggests that LINC00926/PGK1 may represent a potential signaling pathway that inhibits the growth and progression of breast cancer. In TNBC, a hypoxia-induced lncRNA MIR210HG as a glycolytic regulator can regulate the expression of glycolytic genes and drive the Warburg effect by increasing the translation of HIF-1α mRNA, suggesting that MIR210HG target may be a potential strategy for the treatment of TNBC patients212. LncRNA ABHD11-AS1 is overexpressed in NSCLC and closely associated with a poor prognosis. M6A methyltransferase-like 3 (METTL3) enhances the stability and expression of ABHD11-AS1 transcript through M6A modification, and promotes NSCLC proliferation by inducing Warburg effect213. The expression of lncRNA HULC is up-regulated in hepatocellular carcinoma214, and as a bridging molecule, it can directly bind with LDHA and PKM2 and increase their phosphorylation to promote aerobic glycolysis and enhance the proliferation of hepatocellular carcinoma cells215. LncRNA KCNQ1OT1 is up-regulated in a variety of tumors216,217. It enhances aerobic glycolysis and promotes cancer progression by stabilizing HK2 in CRC218. Interestingly, another kind of lncRNA SNHG14 is overexpressed not only in many kinds of cancers219,220, but also in ischemic brain tissue221. In glioma, Lin28A can stabilize SNHG14 and its simultaneous overexpression222. The up-regulation of SNHG14 will promote the degradation of IRF6 mRNA, further promote the transcription of PKM2 and GLUT1, and finally induce aerobic glycolysis and glioma cell proliferation. It shows the potential strategy of targeting Lin28A/SNHG14/IRF6 axis to regulate Warburg effect in the treatment of glioma222. The cancer genome map database also found that another kind of lncRNA NKILA was overexpressed in glioma patients, and its expression level was negatively correlated with the survival time of patients223, 224, 225. NF-κB interacting lncRNA NKILA enhanced the Warburg effect and angiogenesis of glioma by up-regulating HIF-1α expression in vivo and in vitro, and stimulated the growth of glioma. It suggests that NKILA may be a potential target for glioma. Also acting as tumor promoters are lncRNA IDH1-AS1 and lincRNA-P21, which also promote the progression of cancer by promoting aerobic glycolysis226,227. Interestingly, lncRNA EPB41L4A-AS1 as a tumor suppressor can induce p53 and PGC-1α, its down-regulation and deletion are related to the poor prognosis of cancer patients. Knocking down EPB41L4A-AS1 can promote aerobic glycolysis and glutamine metabolism228. Apart from inducing the Warburg effect to promote tumor cell proliferation, lncRNAs can also facilitate tumor metastasis. In HCC cells, lncRNA SOX2OT promotes tumor metastasis through PKM2-mediated aerobic glycolysis229. LncRNA metastasis-associated lung adenocarcinoma transcript 1 (MALAT1) enhances mTOR-mediated TCF7L2 translation to regulate cancer glucose metabolism, inhibit gluconeogenesis and promote aerobic glycolysis, and ultimately promote HCC progression230.
4.3.2. miRNAs
As endogenous non-coding RNAs that regulate gene expression after mRNA transcription, miRNAs participate in numerous physiological and pathological processes, including cell proliferation, apoptosis and differentiation231. miRNAs play a crucial role in regulating the glucose metabolism of cancer cells. They mainly target genes that regulate metabolism, change the metabolism of cancer cells, and promote the growth, proliferation and metastasis of cancer cells232.
Disorder of miRNA expression can cause various types of cancer. As a tumor suppressor, miR-124 is down-regulated in most patients with CRA or CRC and plays a role in the development of adenoma233. Overexpression of miR-124 can induce autophagy and apoptosis through PTB1/PKM1/PKM2 feedback cascade to participate in the proliferation and development of CRC233. In addition, miR-122, miR-133a and miR-133b can also participate in the apoptosis of CRC cells by regulating the reprogramming of M2 pyruvate kinase isoenzyme and metabolism234. The OS of RCC patients is positively correlated with the level of has-miR-144-5p and negatively correlated with the level of has-miR-186–3p. They may balance the glycolytic state of RCC by regulating GLUT1235. PGK1 is over-expressed in many types of cancer, including breast cancer, and is related to poor prognosis236, 237, 238. MicroRNA-16-1-3p inhibits the growth and metastasis of breast tumors by inhibiting the Warburg effect mediated by PGK1239. GLUT12 is also up-regulated in breast cancer and promotes glucose uptake240. Let-7a-5p, as a tumor suppressor241, 242, 243, can inhibit the Warburg effect by inhibiting the expression of GLUT12, and ultimately inhibit the growth and metastasis of breast cancer244. The expression of miR-378 (∗) as a molecular switch also affects the occurrence and development of breast cancer. MiR-378 targets the mRNA of ERRγ and GABPA at the same time, affects its ability to encode energy metabolism, accelerates the process of tricarboxylic acid circulation, reduces cell oxygen consumption, and ultimately enhances the proliferation of breast cancer cells245. MiR-30a-5p and miR-34a interfere with aerobic glycolysis to treat breast cancer and cervical cancer by inhibiting the expression of LDHA246,247. MiRNA-27b also plays an anti-cancer role in breast cancer by targeting multiple tumor suppressor genes as a carcinogenic miRNA248, 249, 250. Glucose metabolism disorder is often related to the abnormal function of PDH complex components, and its expression level is positively related to the survival rate of breast cancer patients251,252. MicroRNA-27b reduces mitochondrial oxidation and promotes extracellular acidification by inhibiting PDHX, resulting in cell metabolic disorder and promoting the proliferation of breast cancer252. MiR-155 promotes Warburg effect in the development of breast cancer by up-regulating HK2253. TNBC is the subtype with the worst prognosis in breast cancer and shows high metabolic remodeling254. The cancer genome map data showed that the key regulator miR-210-3p in TNBC related to aerobic glycolysis promoted aerobic glycolysis, induced Warburg effect and reduced TNBC cell apoptosis by regulating the downstream aerobic glycolysis genes of HIF-1α and p53255. As a carcinogenic miRNA, miR-214 is up-regulated in different cancers and is induced by hypoxia, its expression level is related to distant metastasis256, 257, 258, 259, 260. MiR-214 promotes aerobic glycolysis by targeting adenosine A2A receptor (A2AR) and PR/SET domain 16, and ultimately improves the migration and proliferation of cancer cells261. In addition to inducing the Warburg effect, which plays a carcinogenic role in the occurrence and development of cancer, miRNAs can also serve as tumor suppressors for cancer treatment. Interestingly, miR-3662 inhibits Warburg effect by targeting HIF-1α262. MiR-885-5p and miR-532-3p inhibit the progression of liver cancer and ovarian cancer by silencing HK2 negatively regulating Warburg effect, respectively263,264. The liver is the primary organ responsible for regulating the metabolism and energy homeostasis of the whole body, and its abnormal energy metabolism may be closely related to the progress of liver cancer. MiR-199a targets both HK2 and PKM2 to limit the aerobic glycolysis process for the treatment of hepatocellular carcinoma265,266. The expression of miR-23a in HCC mouse model and primary human HCC is up-regulated, and its activation can inhibit PGC-1α and G6PC in a targeted manner, resulting in the accumulation of G6P to produce ribose-5-phosphate for nucleotide synthesis through hexose monophosphate shunt pathway, which can meet the nutrition required for the division and growth of liver cancer cells under hypoxia conditions, and promote the development of liver cancer267. MiR-338-3p inhibits the expression of PKLR to alleviate the Warburg effect inhibiting the progression of liver cancer268. C-MYC oncogene can produce c-Myc protein, which regulates miRNA and glucose metabolism enzymes and ultimately promotes cell proliferation269,270. The expression of miR-23a/b in human prostate cancer is significantly down-regulated271. c-MYC regulates glutaminase through transcriptional inhibition of miR-23a/b, up-regulates glutamine catabolism, and promotes the occurrence of cancer272. MiR-145 inhibits miR-133b/PKM2 pathway by targeting c-myc/DNMT3A pathway in ovarian cancer cells, and jointly inhibits aerobic glycolysis273. Also, miR-644a and miR-34c are tumor suppressors that inhibit Warburg effect by targeting c-myc274,275. As one of the cancers with high aerobic glycolysis rate, PDAC is highly sensitive to miRNAs that affect Warburg effect276,277. YT521-B homology domain containing protein YTHDC1 can regulate mRNA splicing to promote the decay of pri-miR-30d and induce the activation of miR-30d178. The interaction between miR-30d and transcription factor RUNX1 subsequently down-regulated the expression of SLC2A1 and HK1 to inhibit aerobic glycolysis, suggesting that miR-30d, as a tumor suppressor, is related to the prognosis of PDAC, and can be used as a potential target and effective prognostic marker for research178. The RNA-binding protein Lin28A and its homologue Lin28B target PDK1 in an anoxic or HIF-1-independent manner by regulating the PI3K-mTOR pathway mediated by let-7278, one of the first miRNAs found in Caenorhabditis elegans279, 280, 281, and finally enhance the aerobic glycolysis of cancer cells in vivo and promote the proliferation of tumor cells282. It provides a potential strategy for the treatment of cancer with abnormal expression of Lin28 and let-7.
4.3.3. Circular RNAs
CircRNA is a kind of closed circular RNA formed by reverse splicing of exon or intron, which widely exists in human cells and is highly conservative in different species, but has its specificity in different tissues, different developmental stages and different diseases. Due to its unique ring structure, lack of 5′ terminal cap and 3′ terminal poly A tail, and it is not easy to be degraded by exonuclease RNaseR, making circRNA has unique advantages in the development and application of new clinical markers. In addition, more and more evidence show that circRNA, like mRNA and lncRNA, is rich in a large number of microRNA (miRNA) binding sites, which can be used as a miRNA sponge to participate in the competitive endogenous RNA (ceRNA) regulatory network, regulate the expression of downstream target genes, and then participate in the regulation of the occurrence and development of various human diseases including cancer283, 284, 285. A large number of studies have shown that circRNA plays an important role in the occurrence and development of many clinical common tumors by posttranscriptional regulation of miRNA and/or lncRNA.
CircRNA microarray sequencing in breast cancer samples showed that circRNF20 transcript was up-regulated and associated with poor prognosis286. CircRNF20, as a miRNA sponge, carries miR-487a to target and promote the promoter binding of HIF-1α and HK2, promote the transcription and expression of HK2, and promote the proliferation of breast cancer cells by inducing Warburg effect286. In oral squamous cell carcinoma, the expression of GLUT1 and hsa_circRNA_100290 (circ_SLC30A7) was significantly up-regulated, and the expression of miR-378a was significantly down-regulated287,288. MiR-378a can directly target and bind circRNA100290 and GLUT1. As ceRNA, circRNA_100290 can counteract the inhibition of GLUT1 mediated by miR-378a, thus inducing Warburg effect and promoting the proliferation of cancer cells288. It is possible that the targeted regulation of circRNA_100290/miR-378a/GLUT1 axis is a potential strategy for oral squamous cell carcinoma. AKT/mTOR pathway, as a classical signal pathway that mediates tumor metabolic homeostasis, promotes tumor growth and metastasis by inducing Warburg effect to maintain energy homeostasis289, 290, 291, 292. CircNRIP1 as microRNA-149-5p sponge can promote the proliferation, migration and invasion of gastric cancer through AKT1/mTOR pathway293. As a DNA-RNA binding protein, the overexpression of heterogeneous nuclear ribonucleoprotein K (HNRNPK) is related to the poor prognosis of various types of cancer294, 295, 296, 297. CircVPS33B, which is up-regulated in invasive GC, as a sponge of miR-873-5p, can increase the expression of HNRNPK, promote glucose uptake and lactate production, induce Warburg effect, and promote the growth of invasive GC298. PKM2, as the rate-limiting enzyme of aerobic glycolysis, can promote the glucose uptake of tumor cells299,300. The up-regulated circATP2B1 expressed in GC targets miR-326-3p/miR-330-5p in the RNA-induced silencing complex dependent manner, reducing their inhibition of PKM2, promoting the aerobic glycolysis process, and enhancing the growth and proliferation of GC301. In other studies, the overexpression of a novel type of circCUL3 was identified in GC through circRNA microarray analysis and qRT-PCR experiment302. CircCUL3 acts as a sponge for miR-3-515p, leading to the activation of signal transducer and activator of transcription 3 (STAT3), which then accelerates the transcription process of the key rate-limiting enzyme HK2 in aerobic glycolysis, ultimately triggering the Warburg effect and promoting the development of GC302. Because of the heterogeneity and lack of effective markers, most patients with gallbladder cancer (GBC) are in advanced stage when diagnosed303,304. Some researchers found that circFOXP1 overexpressed in GBC as a key regulator of Warburg effect is also a prognostic biomarker of GBC progression305. CircFOXP1 can stabilize PKLR mRNA by promoting the nuclear cytoplasmic shuttle of PTBP1 or acting as a sponge of miR-370, and up-regulate its expression to regulate Warburg effect and promote GBC progress305. Interestingly, circRNA can still act as a reused signal pathway regulator in gene expression without relying on acting as a miRNA sponge306,307. The up-regulated expression of CircECE1 in OS inhibits the transcription of thioredoxin binding protein, a negative regulator of aerobic glycolysis, by activating the proto-oncogene c-Myc, and finally regulates the Warburg effect to promote the occurrence and development of OS308.
4.3.4. Crosstalk among the lncRNAs, miRNAs and circular RNAs
The miR-210 derived from pancreatic cancer stem cells has been identified as a carcinogenic miRNA that can enhance the resistance of cancer cells to GEM309. DLEU2L, as the ceRNA of miR-210-3p, is expressed at a low level in pancreatic cancer tissue. It can down regulate its expression by combining with miR-210-3p, block AKT/mTOR signal transduction and inhibit Warburg effect, and finally inhibit the proliferation, migration and invasion of pancreatic cancer cells310. H19, one of the most extensively characterized lncRNAs related to tumor progression, is overexpressed in various tumors and functions as an oncogene311, 312, 313, 314, 315. H19 can promote glucose consumption and lactic acid production by inhibiting miR-519d-3p/LDHA axis in GC, induce aerobic glycolysis, accelerate proliferation and immune escape of GC cells316. It suggests that targeting the H19/miR-519d-3p/LDHA axis may be an effective strategy for the treatment of GC. H19 can also directly bind with miR-324-5p and inhibit it. An active saponin monomer, ginsenoside 20(S)-Rg3, extracted from red ginseng, can block H19's competitive inhibition of miR-324-5p, further enhance the inhibition of PKM2, and finally inhibit the Warburg effect and hinder the occurrence of ovarian cancer cells317. LncRNA nuclear-rich transcripts 1 (NEAT1) plays carcinogenic roles in different types of tumors as tumor promoters318, 319, 320, 321, 322, 323. Interestingly, NEAT1 is positively correlated with 5-Fu resistance in cervical cancer. The up-regulated expression of miR-34a can knock down NEAT1, bind with LDHA and inhibit the rate of cell aerobic glycolysis, and finally sensitize 5-Fu resistant cervical cancer cells324. Esophageal cancer (EC), one of the five deadliest cancers325, is often characterized by early diagnosis, late invasion and dysphagia326,327. The expression of lncRNA urothelial carcinoma associated 1 (UCA1) was up-regulated in EC, positively correlated with HK2, and negatively correlated with miR-203 expression328. The overexpression of UCA1 can promote the activation of HK2 oncogene by inhibiting the activity of miR-203, improve the glucose uptake rate and produce lactic acid, induce Warburg effect and promote the development of EC328.
4.4. Oncogenic protein/tumor suppressor
Mitochondria, as the central platform of cell metabolism, can provide the ATP needed by cancer cells to meet capacity requirements329,330. Warburg effect is often accompanied by mitochondrial imbalance or dysfunction331,332. Compared with normal tissues, hypoxia activated E3 ligase SIAH2 in breast cancer tissues degrades NRF1, downregulates the expression of nuclear-encoded mitochondrial gene, which is related to poor clinical prognosis, thus enhancing aerobic glycolysis333. HIFs, a key factor in the adaptive response of cells to hypoxia, can up-regulate the genes encoding metabolic pathway enzymes334, 335, 336, and the overexpression of HIFs is often associated with high aerobic glycolysis and poor clinical prognosis337, 338, 339, 340, 341. It can increase the production of anaerobic ATP and inhibit tumor cell death through metabolic reprogramming342, 343, 344. Rho-related BTB domain-containing protein 3 (RHOBTB3), as a novel scaffold protein, is an important ATPase345, which can promote the degradation of HIFα under the condition of hypoxia or not, thus inhibiting the aerobic glycolysis process and inhibiting the development of tumor346. In addition, SIRT3, the main mitochondrial deacetylase in the Sirtuins family347, inhibits aerobic glycolysis by inhibiting HIF1α-mediated metabolic reprogramming and inhibits the growth and proliferation of tumor348. Tumor suppressor Ras association domain family 1A (RASSF1A) can also promote its stability and enhance the activation of Warburg effect by binding with HIF-1α349. Interestingly, HIF-1α can also activate its transcription through RASSF1A binding. RASSF1A–HIF-1α forms a feedforward loop to drive aerobic glycolysis349.
Glucose metabolism can eliminate free radicals in mitochondria through pyruvate, which has a certain resistance to necrosis and apoptosis of cancer cells in hypoxic environment350. Necrosis is regulated by the formation and phosphorylation of receptor interacting protein kinase, receptor interacting protein 1–receptor interacting protein three complex. It can enhance the resistance of cancer cells to receptor interacting protein dependent necrosis by participating in the Warburg effect and regulating the aerobic glycolysis pathway351. High dimodulin expression, as a key factor in the formation of the three-dimensional structure of tumor cells, not only promotes increased glucose utilization by cancer cells but also accelerates the aerobic glycolysis process. It promotes the occurrence and development of cancer through the activation of the Warburg effect352. In some specific types of cancer, such as colorectal cancer, cancer cells do not decompose glucose into pyruvate and enter the tricarboxylic acid cycle to supply energy as normal cells do, but take glucose for metabolic synthesis through a higher rate of aerobic glycolysis. It is considered to provide tumor cells with an evolutionary advantage and provide more biosynthetic substances for the growth and reproduction of tumor cells353. Colon cancer-1 (MACC1) was first discovered in colon cancer. Its upstream and downstream constitute a delicate regulatory environment supporting its carcinogenic effect354. MACC1 promotes Warburg effect by regulating PI3K/protein kinase B (Akt) signal pathway, thus accelerating the occurrence and development of colorectal cancer355. PTEN, as a tumor suppressor, is often mutated or deleted in a variety of cancer types356, 357, 358, 359. Its overexpression also depends on PI3K to block metabolic transformation and inhibit the occurrence of aerobic glycolysis360. The effector mTOR downstream of AKT can also provide ATP and other nutritional requirements for cancer cells by promoting central carbon metabolism361. The overexpression of glycolytic enzyme GPI in many human cancers is associated with poor prognosis362. The metabolism of cancer cells is characterized by extensive aerobic glycolysis dependence, but the single drug use of aerobic glycolysis inhibitors may promote the metabolic reprogramming of cancer cells mediated by mTORC1/S6K1 to avoid fragile aerobic glycolysis dependence. Combining the inhibition of mTORC1 on the basis of inhibition of aerobic glycolysis may be a potential strategy to deal with the phenomenon of aerobic glycolysis resistance. A study found that the silencing of GPI and the combination of mTORC1 inhibitor Everoliumus can reduce the growth of xenograft tumors by blocking metabolic reprogramming361. It is noteworthy that the high basic level of phosphorylated p70 S6 kinase and ribosomal protein S6 (P-S6) is negatively correlated with the sensitivity of cancer cells to aerobic glycolysis inhibitor 2-deoxyglucose (2DG), which can be used as a predictor of the response of cancer cells to aerobic glycolysis inhibition361. As a metabolic sensor, AMPK is usually involved in the negative regulation of mTOR to maintain cell energy homeostasis363,364. Its deletion accelerated the occurrence of HIF-1-induced aerobic glycolysis under normoxic stable conditions, and promoted the progress of Myc-mediated lymphoma365.
Interestingly, there is no enhancement of Warburg effect in circulating chronic lymphoblastic leukemia (CLL) cells366,367, but after contact with the matrix microenvironment, it promotes the metabolism of CLL into aerobic glycolysis through Notch–c-Myc signal transduction368, which contributes to the drug resistance of cancer cells369. Tumor suppressor follicular protein (FLCN) can also interact with AMPK and inhibit its activation of HIF370,371, which hinders AMPK-mediated Warburg effect and ultimately inhibits the occurrence and development of cancer372. PKM2 expressed only in normal embryonic stage will be re-expressed in most cancer cells146. PKM2 expression is necessary for Warburg effect. PKM2 K305 acetylation will reduce its enzyme activity and promote its degradation through HSC70 chaperone-mediated autophagy (CMA), thus reversing the aerobic glycolysis process373. The double specific phosphatase Cdc25A, which is up-regulated in various types of cancer374, can mediate the dephosphorylation of PKM25 to promote aerobic glycolysis and tumorigenesis375. Protein arginine N-methyltransferase 6 (PRMT6) promotes PKM2 nuclear relocation and inhibits ERK/PKM2 axis-driven aerobic glycolysis by methylation of CRAF at arginine 100 and interference of RAS/RAF signal transduction to ERK376. The deletion of PKM2-mediated tumorigenicity and sorafenib resistance in HCC mouse models and clinical samples, which was reversed by the aerobic glycolysis inhibitor 2DG376, Suggesting that targeting the PRMT6–ERK–PKM2 regulatory axis could be a potential strategy for treating sorafenib resistance events in HCC. Under the conditions of up-regulation of K-Ras G12V and B-Raf V600E expression and activation of EGFR, ERK can also drive phosphorylation of PGK1 S203 and lead to mitochondrial translocation, and induce PGK1 to act as a protein kinase in TCA cycle to promote Warburg effect and induce cancer377. Another study found that when the downstream signal molecule of Kras/ERK pathway, guanosine triphosphatase ADP ribosylation factor 6 (ARF6) was knocked down, Warburg effect could be inhibited and pancreatic cancer cell proliferation could be reduced378, which could be used as a new prognostic marker for pancreatic cancer. A study has found that HCC is often accompanied by epigenetic changes, in which the expression of lymphatic specific helicase (HELLS) is significantly up-regulated in HCC and induces aerobic glycolysis379. The depletion of HELLS will reverse the Warburg effect, ultimately inhibiting the growth and metastasis of HCC in vivo and in vitro379. In addition, a new epigenetic regulatory modification method, N-methyladenosine (mA) modification of RNA, has also been found to have anticancer efficacy in CRC patients by affecting aerobic glycolysis. METTL3, as a clinical oncogene of CRC, interacts with HK2 and GLUT1 through the mA-IGF2BP2/3 dependent mechanism and stabilizes their expression, ultimately activating the glycolytic pathway. Targeting METTL3 and its mA modification may be a potential strategy for treating CRC380. Interestingly, the expression level of lncRNA LINRIS is negatively correlated with OS in CRC patients. LINRIS blocks the K139 ubiquitination of the mA “reader” IGF2BP2, preventing its degradation. Knocking down LINRIS can correspondingly hinder the downstream effect of IGF2BP2, cause metabolic changes in CRC cells, and inhibit MYC-mediated aerobic glycolysis process381.
As a key transcription factor of aerobic glycolysis, the proto-oncogene c-Myc is directly involved in regulating the expression of genes related to glucose metabolism382,383. Soft tissue sarcoma is divided into more than 70 subtypes due to its strong heterogeneity384,385, the targeted treatment strategy is not optimistic, and the response rate to cytotoxic chemotherapy is also low386. Some researchers found that fructose-1,6-diphosphatase 2 (FBP2) is missing in most STS subtypes. Forcing FBP2 expression can simultaneously inhibit aerobic glycolysis and c-Myc-mediated TFAM expression, hinder mitochondrial biogenesis and respiration, thus inhibiting the proliferation of STS387. In addition, the recovery of FBP1 expression can also treat breast cancer and ccRCC by inhibiting Warburg effect388,389. The E3 ubiquitin ligase RING finger protein 6 (RNF6) overexpressed in pancreatic cancer (PC) can promote the degradation of MAD1 through the ubiquitin–proteasome pathway, upregulate the expression of c-Myc, and promote the Warburg effect in PC390. It is suggested that RNF6 may be a new biomarker in PC. SIRT6, a member of NAD-dependent protein deacetylase family, can also regulate ribosomal metabolism by inhibiting MYC transcription activity391, 392, 393, and its expression level is negatively correlated with poor prognosis393. It is suggested that SIRT6 may be a molecular switch regulating Warburg effect. In addition, p53, the most common mutant gene in cancer, can also regulate the balance between respiratory and glycolytic pathways, while wild-type p53 in cancer often destroys the synthesis of cytochrome c oxidase 2 (SCO2) gene downstream of Warburg effect394. The deletion of p53 increased the expression of SCO2 and promoted the metabolic transformation to Warburg effect394. It provides a research direction for revealing the molecular mechanism of p53 regulation of the regulating Warburg effect in cancer. CDK8 is an oncogene that promotes the proliferation of cancer cells in various types of cancer395, 396, 397, 398, 399. Its activity is necessary for the aerobic glycolysis cascade reaction. The deletion of CDK8 can make colorectal cancer cells sensitive to aerobic glycolysis inhibitors400. TNBC is the most heterogeneous breast cancer, often accompanied by poor clinical prognosis401. Isocitrate dehydrogenase 2 (IDH2) is overexpressed in TNBC and induces aerobic glycolysis through serine biosynthesis pathway in the presence of phosphoglycerate dehydrogenase and phosphate aminotransferase402. This study found that phosphoglycerate dehydrogenase and phosphate aminotransferase as IDH2 synthetic dosage lethal partners may be potential targets for the treatment of TNBC in the future (Table 4) (Fig. 4).
Table 4.
Oncogenic protein/Tumor suppressor.
| Name | Classification | Effect on Warburg effect | Regulatory mechanism | Cancer type | Ref. |
|---|---|---|---|---|---|
| p38γ MAPK | Oncogenic protein | + | Up-regulation of p38γ increases the expression of PFKFB3 and GLUT2 to enhance aerobic glycolysis | Pancreatic ductal adenocarcinomas (PDAC) | 89 |
| CUEDC2 (CUE domain-containing protein 2) | Oncogenic protein | + | CUEDC2 promotes the expression of GLUT3 and LDHA | Hepatocellular carcinoma | 93 |
| TAp73 | Oncogenic protein | + | Ap73 increases the expression of phosphofructokinase-1, liver type (PFKL) to promote aerobic glycolysis | Osteosarcoma; colon cancer; lung cancer | 130 |
| ERK2 | Oncogenic protein | + | ERK2 binds and phosphorylates PKM2 at Ser 37 to promote its nuclear translocation, therefore up-regulate c-Myc expression | Glioblastoma | 148 |
| AKT | Oncogenic protein | + | The PI3K/AKT pathway activates the transport of glucose, decreases glycogen synthesis via regulation of HK and PFK. | Hepatocellular carcinoma | |
| TGF-β1 | Oncogenic protein | + | TGF-β1 activates Smad, p38 MAPK, PI3K/AKT pathways as well as HIF-1α up-regulation. | ||
| Aurora-A | Oncogenic protein | + | Aurora-A phosphorylates LDHB at serine 162 and increases LDHB activity to contribute to the Warburg effect | Colorectal adenocarcinoma | 155 |
| CD147 | Oncogenic protein | + | CD147 promotes the expression of MCT1 and increases p53 degradation to facilitate the Warburg effect. | Hepatic carcinoma | 166 |
| Calcium binding protein 39-like (CAB39L) | Tumor-suppressor | – | CAB39L interacts with LKB1–STRAD complex and then activation of the LKB1-AMPKα/β-PGC1α pathway to enhance the mitochondrial OXPHOS | Gastric cancer | 206 |
| LKB-1 | Tumor-suppressor | – | LKB1 activates AMPK-PGC1α pathway to inhibit aerobic glycolysis | Gastric cancer | 206 |
| METTL3 | Oncogenic protein | + | METTL3 installs the m6A modification and enhances ABHD11-AS1 transcript stability to increase its expression thus to enhance the Warburg effect | Non-small-cell lung cancer | 213 |
| IRF6 | Tumor-suppressor | – | IRF6 inhibited the transcription of PKM2 and GLUT1, thereby impairing aerobic glycolysis | Glioma | 222 |
| Stat3 | Tumor-suppressor | – | Stat3 activates miR-23a to regulate PGC-1α and G6PC thus to decrease the glucose production | Hepatocellular carcinoma | 267 |
| mTOR complex 2 | Oncogenic protein | + | mTORC2 up-regulates the cellular level c-Myc to control the aerobic glycolysis through mtorc2-HDAC–FOXO pathway | Glioblastoma | 275 |
| Lin28 A/B | Oncogenic protein | + | Lin28 A/B increases PDK1 protein expression and PI3K–AKT activation to promote aerobic glycolysis | Hepatoma | 282 |
| SIAH2 | Oncogenic protein | + | SIAH2 degrades NRF1 (nuclear respiratory factor 1) via ubiquitination it on lysine 230 thus to inhibit the expression of pyruvate dehydrogenase beta to enhance the Warburg effect | Breast cancer | 333 |
| Rho-related BTB domain-containing protein 3 (RHOBTB3) | Tumor-suppressor | – | RHOBTB3 interacts with the hydroxylase PHD2 to promote HIFα hydroxylation, ubiquitination and degradation to repress the Warburg effect | Renal carcinoma | 346 |
| Metastasis associated with the colon cancer 1 (MACC1) | Oncogenic protein | + | MACC1 promotes the Warburg effect via PI3K/AKT signaling pathway | Gastric cancer | 355 |
| PTEN | Tumor-suppressor | – | PTEN inhibits aerobic glycolysis via PFKFB3 (6-phosphofructo-2-kinase/fructose-2,6-biphosphatase isoform 3) inhibition and PI3K/mtor/PKM2 suppression | 360 | |
| mTOR complex 1 | Oncogenic protein | + | mTORC1 directs the increased glucose flux back to aerobic glycolysis through the pentose phosphate pathway | Ovarian cancer | 361 |
| AMPK | Tumor-suppressor | – | AMPKα suppresses aerobic glycolysis through HIF-1α inhibition | Lymphomas | 365 |
| Notch | Oncogenic protein | + | Notch promotes the Warburg effect via c-Myc activation | Chronic lymphocytic leukemia (CLL) | 369 |
| Folliculin (FLCN) | Tumor-suppressor | – | FLCN is an AMPK binding partner and inhibits HIF-dependent aerobic glycolysis | Endometrial adenocarcinoma | 372 |
| HSC70 | Tumor-suppressor | – | PKM2 interacts with HSC70 and promotes its lysosomal-dependent degradation. | Cervical cancer | 373 |
| CDC25A | Oncogenic protein | + | CDC25A-mediated PKM2 dephosphorylation could up-regulation of the expression of the glycolytic genes | Glioblastoma | 375 |
| N-Methyltransferase 6 (PRMT6) | Tumor-suppressor | + | PRMT6 methylates CRAF to inhibit ERK-mediated PKM2 activation | Hepatocellular carcinoma | 376 |
| EGFR | Oncogenic protein | + | The activation of EGFR induces ERK-dependent phosphoglycerate kinase 1 S203 phosphorylation to promote its mitochondrial translocation, Mitochondrial PGK1 phosphorylate pyruvate dehydrogenase kinase 1 at T338 to inhibit the pyruvate dehydrogenase (PDH) complex to enhance the Warburg effect. | Glioma | 377 |
| ADP ribosylation factor 6 (ARF6) | Oncogenic protein | + | ARF6 maintains the activation of ERK/c-Myc axis to enhance the Warburg effect | Pancreatic cancer | 378 |
| Helicase, lymphoid-specific (HELLS) | Oncogenic protein | + | HELLS inhibits multiple tumor suppressor genes expression to facilitate the Warburg effect | Hepatocellular carcinoma | 379 |
| Gluconeogenic isozyme fructose-1,6-bisphosphatase 2 (FBP2) | Tumor-suppressor | – | Nuclear FBP2 represses c-Myc-dependent TFAM expression to block the Warburg effect | Sarcoma | 387 |
| Fructose-1,6-bisphosphatase 1 (FBP1) | Tumor-suppressor | – | FBP1 interacts with the HIF inhibitory domain to inhibit nuclear HIF function | Renal carcinoma | 389 |
| E3 ubiquitin ligase RING-finger protein 6 (RNF6) | Oncogenic protein | + | RNF6 enhances c-Myc-mediated aerobic glycolysis. | Pancreatic cancer | 390 |
| SIRT6 | Tumor-suppressor | – | SIRT6 suppresses aerobic glycolysis through MYC inhibition | Pancreatic and colorectal cancers | 393 |
| SIRT3 | Tumor-suppressor | – | SIRT3 inhibits the Warburg effect through PHD regulated HIF-1α destabilization | Breast cancers | 348 |
| Ras association domain family 1A (RASSF1A) | Oncogenic protein | + | RASSF1A regulates HIF-1α stability to increase HIF-1α mediated the transcription of genes associated with aerobic glycolysis | Cervical cancer | 349 |
| Synthesis of cytochrome c oxidase 2 (SCO2) | Tumor-suppressor | – | SCO2 regulates the cytochrome c oxidase (COX) complex to inhibit the Warburg effect | Human colon cancer | 394 |
| CDK8 | Oncogenic protein | + | CDK8 promotes the expression of multiple glycolytic genes | Colorectal carcinoma | 400 |
| Isocitrate dehydrogenase 2 (IDH2) | Oncogenic protein | + | IDH2 regulates the serine biosynthesis pathway to affect the TCA cycle and aerobic glycolysis | Triple-negative breast cancer (TNBC) | 402 |
Figure 4.
Protein–protein interactions regulating the Warburg effect. Mechanisms of protein–protein interactions include phosphorylation, methylation and ubiquitination. One protein can also promote the transcription or translation of another protein to increase the expression of the interacting partner.
The maintenance of cellular metabolic homeostasis relies on a multi-level and multi-dimensional three-dimensional regulatory network between the body and cells, as well as within cells, known as the cellular metabolic signaling network. This network is very complex, including various growth factor signaling pathways (such as EGFR), tumor suppressors (such as p53), oncogenic factors (such as mTOR), HIF-1α, c-MYC, and AMPK. The metabolic pathways within cells are interdependent and mutually constrained, maintaining the steady state of cellular metabolism.
4.5. Mutations, tumor microenvironment and others
Mutations are significant triggers for tumor occurrence and development, as well as a key factors contributing to the disordered the Warburg effect. Wild-type IDH1/2 can promote aerobic glycolysis to enhance the survival and proliferation of cancer cells, while in IDH mutant (IDH (mt)) tumor cells, the expression of certain essential glycolytic genes is down-regulated403, and the key glycolytic gene LDHA showed low expression and hypermethylation in IDH (mt) GBM404,405. The tumor suppressor gene wild-type p53 suppresses the Warburg effect in tumors by regulating the energy metabolism related genes SCO2, TIGAR, GLS2 and Parkin406, 407, 408, 409. As the most common mutated gene, wild-type p53 became mutp53 after mutation410. The latter induced Warburg effect by promoting GLUT1 translocation to the plasma membrane after the activation of RhoA/ROCK pathway411. The inhibition of RhoA/ROCK/GLUT1 signal pathway can reverse the Warburg effect and inhibit the occurrence of cancer411. In addition, splicing factor 3b subunit 1 (SF3B1) plays an important role in the correct splicing of mRNA412, 413, 414, 415. However, hot spot gene mutations of SF3B1 often occur in different types of human cancer416, 417, 418, 419, 420. SF3B1 K700E mutation causes abnormal splicing of PPP2R5A in PDAC, and promotes the proliferation of PDAC cells through the activation of aerobic glycolysis regulator c-Myc through post-translation regulation421. SF3B1 knockout or PP2A activator FTY-720 can significantly inhibit the SF3B1 K700E mutant PDAC model in vitro and in vivo421.
Importantly, many factors in the microenvironment affect the metabolism of tumor cells. Hypoxia is the main factor that affects the metabolic state of tumor cells. Hypoxia can induce the stabilization of HIFs, improve glucose uptake by up-regulating GLUT1 and GLUT3, up-regulate the key enzymes of aerobic glycolysis, accelerate the aerobic glycolysis process, up-regulate pyruvate dehydrogenase kinase, inhibit pyruvate dehydrogenase activity and inhibit OXPHOS, make cells more dependent on aerobic glycolysis pathway to produce ATP, and finally complete the transformation from OXPHOS to aerobic hydrolysis422. Tumor cells that are over-dependent on aerobic glycolysis metabolism pathway, when exposed to areas with perfect blood supply and sufficient oxygen, the energy metabolism mode changes to OXPHOS, and the ability of metastasis and invasion is enhanced. The change of nutritional components in tumor microenvironment will also affect the metabolic mode of tumor cells. HeLa cells are mainly dependent on aerobic glycolysis pathway in the presence of glucose. However, when the culture environment contains only galactose, HeLa cells decompose galactose through OXPHOS to maintain their own proliferation423. The rapid proliferation of tumor cells needs to consume a lot of glucose and oxygen, resulting in a relative lack of nutrition and the formation of local hypoxia microenvironment. Hypoxia, low pH value, nutrient deficiency and other physical and chemical conditions are conducive to the proliferation and metastasis of tumor cells424. These characteristic physical and chemical conditions are involved in regulating tumor energy metabolism and reprogramming, and are also indispensable for tumor cell survival425. The rate of cell growth and catabolism is regulated by various levels of cell regulatory systems, including gene expression, post-transcriptional regulation, translation, and post-translational modification, which ultimately affect metabolic flux426. Metabolic transformation is usually associated with oxygen restriction, genetic disturbance and overflow metabolism427. For example, fructose 1,6-bisphosphate (FBP), a glycolytic metabolite, interferes with respiratory activity and regulates metabolic flux through transcription factors426. The inhibition of Warburg effect of cancer cells will lead to the crisis of their cell bioenergy and further induce the death of cancer cells428.
More and more evidence shows that tumor cells can regulate the expression level of substrate proteins through ubiquitination modification, thereby affecting the activation or inhibition of proteins on Warburg effect. Ubiquitination is a process of protein post-translational modification, characterized by the covalent linkage between ubiquitin molecules and protein substrates. Ubiquitination modification process is reversible and dynamic. It mediates the occurrence and development of tumors by participating in regulating gene transcription, inflammatory response, DNA damage repair and other biological processes429. The regulatory mechanism of ubiquitination/de ubiquitination on Warburg effect in tumor cells is complex, and its mechanism is an important regulatory point for normal cells and tumor cells430. In recent years, with the deepening of research on autophagy, the close relationship between autophagy and “Warburg effect” has garnered significant interest from researchers431. Autophagy and tumor energy metabolism are considered to be highly related to clinical practice and become the focus of cancer transformation research432. In tumor microenvironment, tumor cells and interstitial cells play an important role in tumor genesis and metastasis through the interaction of growth factors and various cytokines. Tumor cells “hijack” peripheral interstitial cells or cancer-associated fibroblast (CAF), and use the high-energy metabolites produced by CAF to maintain the metabolic needs of tumor cells. The activation of oncogene and the large amount of reactive oxygen species secreted by tumor cells can induce CAF to produce oxidative stress, and then induce autophagy of CAF433. ROS can also destroy mitochondrial metabolic enzymes, thus inhibiting TCA cycle, causing damage to mitochondrial function, and increasing CAF aerobic glycolysis434. The high-energy metabolites such as lactic acid and ketone produced by CAF autophagy and the “Warburg effect”, in turn, provide fuel and necessary “chemical building materials” for the mitochondrial biosynthesis and oxidative phosphorylation of tumor cells, resulting in an immortal vicious cycle of tumor cells. This process is called “epithelial mesenchymal metabolic coupling”, also known as “reverse Warburg effect”435, 436, 437. In addition, autophagy can activate the aerobic glycolysis pathway to promote cell glucose metabolism and ultimately promote its proliferation and transformation to tumor during tumor development437. In addition, proinflammatory cytokines, tumor necrosis factor-α and IL-17 will also stimulate and strengthen the aerobic glycolysis of cancer cells and inhibit the oxidative phosphorylation of mitochondria, providing more powerful energy support for the development and metastasis of cancer438 (Table 5).
Table 5.
Mutations, tumor microenvironment and others.
| Name | Classification | Regulation on Warburg effect | Regulatory mechanism | Cancer type | Ref. |
|---|---|---|---|---|---|
| LPS | LPS activates TLR4 to enhance aerobic glycolysis via uPFK2 activation | ||||
| KRAS4A | Oncogenic protein (mutation) | + | KRAS4A targets HKI to initiate the Warburg effect | Pancreatic cancer | 104 |
| Glucose | – | High glucose promotes the acetylation of PKM2 at K305 via the acetyltransferase PCAF and decreases PKM2 enzyme activity and promotes its lysosomal-dependent degradation. | Cervical cancer | 373 | |
| K-Ras G12V | Oncogenic protein (mutation) | + | The activation of K-Ras G12V and B-Raf V600E induce ERK-dependent activation of the warburg effect. | Glioma | 377 |
| B-Raf V600E | Oncogenic protein (mutation) | + | |||
| Mutant isocitrate dehydrogenase 1 (IDH1) | Tumor suppressor (mutation) | – | IDH(mt) in gliomas downregulated the expresson of LDHA and many HIF1α-responsive genes to limit the aerobic glycolysis. | Gliomas | 403 |
| Mutant isocitrate dehydrogenase 2 (IDH2) | |||||
| Mutant p53 | Oncogenic protein (mutation) | + | Mutp53 promotes the Warburg effect through RhoA/ROCK/GLUT1 signalling | Breast cancer | 411 |
| SF3B1 K700E | Oncogenic protein (mutation) | + | MutSF3B1 stimulates the aerobic glycolysis through PP2A-c-Myc activation | Pancreatic ductal adenocarcinoma | 421 |
| Fructose 1,6-bisphosphate (FBP) | Intracellular metabolite | FBP senses the flux of the upper aerobic glycolysis, and regulates the lower glycolytic flux by the allosteric activation of pyruvate kinase | 426 |
5. Targeting Warburg effect as cancer strategies
5.1. Classic strategies
5.1.1. Synthetic small molecule drugs
5.1.1.1. GLUT inhibitors
The proliferation of tumor cells may depend on specific GLUTs or particular subtypes of key glycolytic enzymes. It may provide new clues for the treatment of cancer by taking this as the target of action, using specific drug targeted therapy, regulating the glucose metabolism of tumor cells, and inhibiting the growth of cancer cells (Table 6) (Fig. 5A). It is noteworthy that high GLUT-1 expression is associated with high invasion and tumor incidence of various types of cancer439, 440, 441. A study reported the pharmacological identification results of a series of compounds, in which STF-31 inhibits glucose uptake mediated by GLUT1, selectively inhibits the proliferation and growth of RCCs, and has no adverse reactions82. In addition, researchers have designed and synthesized a more selective GLUT1 inhibitor BAY-876. In vitro PK data shows good metabolic stability in liver cells, while in vivo PK data shows high oral bioavailability and a longer end half-life, suggesting that BAY-876 may be a highly potential selective GLUT1 inhibitor in tumor treatment442. As an effective glucose uptake inhibitor, fasentin can selectively target GLUT-1/GLUT-4 transporter and has anti-angiogenic activity443,444. Upregulation of GLUT-4 expression in low-glucose microenvironment can significantly promote the invasion and metastasis of tumor cells445, 446, 447, 448, 449. The combination of GLUT4 inhibitor Ritonavir and HNF4A agonist Benfluorex can inhibit glucose uptake and proliferation of HCC cells by mediating AMPK2/HNF4A/BORIS/GLUT4 signal pathway450. In addition, high GLUT-1 expression is associated with chemoresistance of various types of cancer451,452. After being treated with GLUT-1 inhibitor WZB117, imatinib-resistant cells exert synergistic growth inhibition by reducing AKT phosphorylation and Bcl-2 expression, overcoming the imatinib-resistance of gastrointestinal stromal tumors453. Cytochalasin B, as a pan inhibitor of GLUT, can hinder the transport of glucose and glucosamine in liver cancer cells by inhibiting GLUT-1–4 targets454, 455, 456. The structure of single transporter human glucose transporter 1 (hGLUT1) has been elucidated, and the Phe amide derived inhibitors GLUT-i1 and GLUT-i2 bind to the specific glucose substrate binding site of hGLUT1 and inhibit its glucose transport function457. Their co-complex structure provides important structural insights for designing more selective hGLUT1 inhibitors. A series of 1-pyrazolo [3,4-d] pyrimidine derivatives also showed good binding to GLUTs, with compound 3 showing significant inhibitory effects on GLUT1 and good selectivity for GLUT2, while exhibiting good metabolic stability and pharmacokinetic characteristics458. Some new glucose uptake inhibitors chromopynone-1 and glupin can selectively target GLUT1/3 to regulate cancer metabolism and ultimately inhibit the growth of various types of tumor cells459,460. As a glucose uptake inhibitor, glutor can also inhibit aerobic glycolysis flux by targeting GLUT1/2/3, and ultimately play a role in inhibiting growth in many cancer cell lines461. Interestingly, NV-5440 can regulate cell metabolism by simultaneously targeting mTORC1 and GLUT1, effectively inhibit glucose uptake, and inhibit the proliferation of breast cancer cells462. Its pharmacokinetics in vivo shows that NV-5440 has good stability in vivo462. This new dual target inhibitor provides a promising research direction for future Warburg effect inhibitor development strategies.
Table 6.
Synthetic small molecule drugs.
| Name in the literature | Chemical structure | Mechanism | Cancer cell line (activity) | Ref. |
|---|---|---|---|---|
| STF-31 | ![]() |
Inhibit glucose uptake mediated by GLUT1 | GLUT1: IC50 = 1 μmol/L; RCC4 |
82 |
| BAY-876 | ![]() |
Inhibit GLUT1 | GLUT1: IC50 = 2 nmol/L; SKOV-3; OVCAR-3 |
442 |
| Fasentin | ![]() |
Inhibit GLUT1/4 | GLUT4: IC50 = 68 μmol/L; PPC-1; DU145; U937 |
443 |
| Ritonavir | ![]() |
Mediate AMPK2/HNF4A/BORIS/GLUT4 signal pathway | Huh-7; HCC-LM3; Hepg2; SMMC-7721; PLC/PRF/5 |
450 |
| WZB117 | ![]() |
Reduce AKT phosphorylation and Bcl-2 expression by inhibiting GLUT1 | GLUT1: IC50 = 10 μmol/L A549; MCF7; GIST-T1; GIST-T1/IM-R |
453 |
| Cytochalasin B | ![]() |
Hinder the transport of glucose and glucosamine by inhibiting GLUT 1–4 targets | B16F10: 0.4 μmol/L | 454, 455 |
| GLUT-i1 | ![]() |
Inhibit hGLUT1 | Glut1: IC50 = 0.267 ± 0.133 μmol/L; Glut2: IC50 = 56 ± 13.6 μmol/L; Glut3: IC50 = 5.2 ± 1.1 μmol/L; Glut4: IC50 = 0.195 ± 0.066 μmol/L |
457 |
| GLUT-i2 | ![]() |
Inhibit hGLUT1 | Glut1: IC50 = 0.140 ± 0.072 μmol/L; Glut2: IC50 > 30 μmol/L; Glut3: IC50 = 7.7 ± 1.35 μmol/L; Glut4: IC50 = 0.090 ± 0.08 μmol/L |
457 |
| Compound 3 | ![]() |
Inhibit GLUT1/2 | Glut1: IC50 = 7 nmol/L; Glut2: IC50 = 1.1 μmol/L; Glut3: IC50 = 40 nmol/L |
458 |
| Chromopynone-1 | ![]() |
Inhibit GLUT1/3 | IC50 = 412 ± 120 nmol/L; UM-UC-3: GI50 = 4 nmol/L; MIA PaCa-2: GI50 = 4 nmol/L |
459 |
| Glupin | ![]() |
Inhibit GLUT1/3 | DLD-1: IC50 = 59.6 ± 8.4 nmol/L; DLD-1 GLUT1−/−: IC50 = 11.4 ± 1.6 nmol/L; UM-UC-3: IC50 = 32 nmol/L; MIA PaCa-2: IC50 = 61 nmol/L; WSU-NHL: IC50 = 62 nmol/L; SU-DHL-6: IC50 = 92 nmol/L; SK-N-SH: IC50 = 84 nmol/L |
460 |
| Glutor | ![]() |
Inhibit GLUT1/2/3 | IC50 = 19 ± 2 nmol/L; MIA PaCa2: GI50 = 0.6 μmol/L; HCT116: GI50 = 3.8 μmol/L |
461 |
| NV-5440 | ![]() |
Inhibit GLUT1 and mTORC1 | pS6K1T389: IC50 = 70 nmol/L; MCF7: IC50 = 36 nmol/L |
462 |
| C3k | ![]() |
Inhibit PKM2 | HCT116: IC50 = 0.18 μmol/L; Hela: IC50 = 0.29 μmol/L |
466 |
| Benserazide | ![]() |
Inhibit PKM2 | 10 μmol/L; SK-MEL-5; SK-MEL-28 |
469 |
| Docetaxel | ![]() |
Inhibit PKM2 | PC/DX25: IC50 = 316.4 nmol/L | 470 |
| Caffeic acid phenethyl ester | ![]() |
Inhibit PKM2 | PC/DX25: IC50 = 15.7 μmol/L | 470 |
| ML-265 | ![]() |
Activate PKM2 | H1299: IC50 = 92 nmol/L | 472 |
| Compound 5 | ![]() |
Activate PKM2 | AC50 = 15 nmol/L; U87: IC50 = 3.03 ± 0.54 μmol/L; U118: IC50 = 1.80 ± 0.28 μmol/L; SF126: IC50 = 7.49 ± 0.45 μmol/L; SHG44: IC50 = 5.24 ± 2.46 μmol/L; U251: IC50 = 7.14 ± 0.06 μmol/L; C6: IC50 = 1.66 ± 0.01 μmol/L; 3T3: IC50 = 7.93 ± 3.21 μmol/L |
473 |
| Compound 16 | ![]() |
Activate PKM2 | U118 tumor xenograft (50 mg/kg) | 473 |
| 2-DG | ![]() |
Inhibit HK2 | MCF-7 | 474, 476 |
| 3-Bromopyruvic acid | ![]() |
Inhibit HK2 | MCF-7: 80 μmol/L; MDA-MB-231: 160 μmol/L |
475, 476, 477, 479 |
| Ketoconazole | ![]() |
Inhibit HK2 | U87: EC50 = 9.3 μmol/L; T89G EC50 = 9.5 μmol/L; U251: EC50 = 10.1 μmol/L; NHA: EC50 = 73 μmol/L |
480 |
| Posaconazole | ![]() |
Inhibit HK2 | U87: EC50 = 9.7 μmol/L; T89G EC50 = 10.2 μmol/L; U251: EC50 = 11.6 μmol/L; NHA: EC50 = 80.3 μmol/L |
480 |
| N-Bromoacetylethanolamine phosphate | ![]() |
Inhibit PFK2 |
Ki = 0.24 mmol/L; IC50 = 2.2 mmol/L |
483 |
| 3PO | ![]() |
Inhibit PFKFB3 | PFKFB3: IC50 = 22.9 μmol/L | 484 |
| PFK15 | ![]() |
Inhibit PFKFB3 | PFKFB3: IC50 = 207 nmol/L | 490 |
| PQP | ![]() |
Inhibit PFKFB3 | T24; UM-UC-3 |
491 |
| N4A | ![]() |
Inhibit PFKFB3 | PFKFB3: Ki = 1.29 ± 0.26 μmol/L; IC50 = 2.97 ± 0.16 μmol/L; HeLa: GI50 = 14.2 ± 1.5 μmol/L |
492 |
| YN1 | ![]() |
Inhibit PFKFB3 | PFKFB3: Ki = 0.24 ± 0.03 μmol/L; IC50 = 0.67 ± 0.08 μmol/L; HeLa: GI50 = 8.2 ± 0.8 μmol/L |
492 |
| PFK158 | ![]() |
Inhibit PFKFB3 | PFKFB3: IC50 = 137 nmol/L; (NCT02044861) | 493 |
| compound 26 | ![]() |
Inhibit PFKFB3 | PFKFB3: IC50 = 11 nmol/L; A549: GI50 = 281 nmol/L |
495 |
| KAN0438757 | ![]() |
Inhibit PFKFB3 | U2OS: 10 μmol/L | 496 |
| 5MPN | ![]() |
Inhibit PFKFB4 | H460 | 499 |
| Metformin | ![]() |
Activate AMPK pathway | 5 mmol/L; MDA-231; MCF-7; U2OS |
500 |
| 2,3-Dihydro-2-(naphthalen-1-yl) quinazoline-4 (1H)-one (DHNQ) | ![]() |
Inhibit PI3K | Colo-205: IC50 = 0.25 ± 0.05 μmol/L; HCT-116: IC50 = 0.3 ± 0.1 μmol/L; Fr-2: IC50 = 4.2 ± 1.3 μmol/L; |
502 |
| Canagliflozin | ![]() |
Inhibit sodium-glucose cotransporter 2 | 10 μmol/L; Huh7; Hepg2 |
503 |
| Romidepsin | ![]() |
Inhibit HDAC and reduce ATP production | U87: IC50 = 18.4 μmol/L | 504 |
| Panobinostat | ![]() |
Inhibit FAO and hinder the occurrence of OXPHOS | U87: IC50 = 0.78 μmol/L | 504 |
| Vorinostat | ![]() |
Inhibit FAO and hinder the occurrence of OXPHOS | U87: IC50 = 2.6 μmol/L | 504 |
| CX-4945 | ![]() |
Downregulation of TAP73 | AGS-1: 10 μmol/L | 506 |
| 1,3-Benzodioxane derivative compound 10 | ![]() |
Inhibit LDHA | LDHA: IC50 = 47.20 μmol/L PANC-1: GI50 = 12.19 μmol/L |
507 |
Figure 5.
Multiple drugs targeting the Warburg effect for treating cancer. (A) Synthetic small molecule drug regulates the Warburg effect for inhibiting cancer progression. (B) Traditional Chinese Medicine targets the regulator of the Warburg effect for cancer treatment. (C) The combination strategy regulates the Warburg effect for interfering with cancer progression.
5.1.1.2. PKM2 inhibitors and agonists
PKM2 is one of the key enzymes of aerobic glycolysis. In recent years, many studies have found that its expression is enhanced in tumor tissue. In lung adenocarcinoma cells, PKM2 knockout significantly inhibited cell proliferation, glucose absorption, ATP and fatty acid synthesis, and enhanced mitochondrial respiratory capacity. More importantly, PKM2 knockout inhibits matrix metalloprotein 2 and VEGF, indicating that reducing PKM2 expression can inhibit tumor growth, metastasis and invasion431. It is suggested that PKM2 inhibitors may exert anticancer effects by inhibiting aerobic glycolysis, and the research and development of PKM2 inhibitors are promising anticancer strategies in the future. In various types of cancer, the use of chemotherapy drug cisplatin often leads to resistance, and PKM2 inhibitor Shikonin can reverse cisplatin resistance by inducing necrotic apoptosis, thereby improving the efficacy of cisplatin in treating cancer463, 464, 465. A novel naphthoquinone derivative C3k exhibits better inhibitory activity against PKM2 targets than Shikonin, and C3k exhibits significant anti proliferative activity in multiple types of tumor cell lines with higher PKM2 levels466, 467, 468. Interestingly, benserazide, a dopamine decarboxylase inhibitor for Parkinson's disease, can also specifically bind and block PKM2 enzyme activity, inhibit aerobic glycolysis, and thus inhibit the growth of melanoma469. The combination therapy of docetaxel and caffeic acid phenethyl ester can also inhibit the expression of PKM2, thereby impeding the metabolic process of tumor cells, inducing cell apoptosis, and ultimately inhibiting the proliferation of prostate cancer cells470. These findings provide more ideas for the combined use of drugs to treat tumors. It is worth noting that PKM2 expression appears to be crucial for tumor formation471, but studies have also shown that using PKM2 expression can replace PKM1's shRNA, thereby reducing lactate production and increasing oxygen consumption, reversing the Warburg effect147. There is a hypothesis that increasing PKM2 activity to PKM1 level may also have anti-proliferative effects. ML-265, as a PKM2 activator, reduces PGAM1 phosphorylation by activating PKM2 in vitro and in vivo models of lung cancer, ultimately inhibiting tumor growth472. A series of parthenolide dimers have also been found to have PKM2 activation activity. Among them, compounds 5 and 16 inhibit STAT3 signaling pathway, ultimately inhibit GBM cell proliferation and metastasis, and inhibit tumor growth while promoting the formation of PKM2 tetramer without affecting the expression of total PKM2473.
5.1.1.3. HK2 inhibitors
2-DG and 3-bromopyruvate are classic inhibitors targeting tumor glycolytic enzyme HK2474,475. After using fasentin, 2-DG and 3-bromopyruvate in cancer cells treated with mitochondrial respiratory regulators, glucose uptake and aerobic glycolysis were inhibited, but the survival rate of irradiated cells was not improved476. It is suggested that enhanced aerobic glycolysis is the reason for inhibiting mitochondrial respiratory related radioresistance. Interestingly, 3-bromopyruvate, the second-generation aerobic glycolysis inhibitor, has a stronger anticancer effect. Some studies have found that 3-bromopyruvate can eliminate tumor tissue in all rat models of advanced liver cancer without recurrence and adverse reactions477. Cisplatin is often resistant to chemotherapy drugs because of ATP-dependent multidrug resistance phenotyp. The addition of 3-bromopyruvate, a multidrug resistance reversal regulator478, reverses the drug resistance of cisplatin by causing ATP depletion, which has stronger synergistic anti-cancer effect than the use of cisplatin alone479. Drug repositioning research not only greatly reduces the development cycle of new drugs, but also reduces economic costs, and is increasingly receiving attention from the pharmaceutical industry480. Azole Antifungal, ketoconazole and posaconazole, were found to be able to target HK2 as tumor metabolism inhibitors to inhibit tumor metabolism, induce apoptosis, and finally play an effective anti-cancer role in GBM in vivo and in vitro models481. Unfortunately, most HK2 inhibitors are currently ineffective and have adverse reactions during clinical trials, making them unable to enter clinical practice482.
5.1.1.4. PFK inhibitors
The first PFK2 inhibitor discovered is N-bromoacetylethanolamine phosphate, a non-competitive inhibition relative to ATP. Unfortunately, this reagent has no inhibitory activity on the phosphorylation of fructose 6-phosphate and is nonspecific for PFK2483. The researchers designed and synthesized the PFKFB3 inhibitor 3PO based on the homologous model of PFKFB3 isozyme484, which has good anti-cancer activity in various types of tumor cells485, 486, 487. Due to the poor pharmacokinetic characteristics and low efficacy of 3PO488, subsequent researchers have improved its efficacy by loading nanocarriers into multiple types of cells489. A new PFKFB3 inhibitor PFK15 with higher selectivity and stronger inhibitory activity was obtained by modifying the 3PO structure, which has better metabolic stability490. Another PFKFB3 inhibitor PQP exhibits an additive growth inhibitory effect when combined with LDHA inhibitors, although its inhibitory activity on cancer cell growth is weaker than PFK15491. In 2011, the crystal structure of PFKFB3 was confirmed, and two compounds N4A and YN1, which are more selective to PFKFB3, were virtual screened and identified to inhibit aerobic glycolysis in human cervical cancer and human breast cancer cells through targeted inhibition of PFKFB3, leading to cell death492. It is worth noting that a new PFKFB3 inhibitor PFK158 is currently undergoing a phase I clinical trial targeting patients with advanced solid malignant tumors, and a phase II clinical trial targeting leukemia patients is also being prepared490. PFK158 has also been found to play an anticancer role in gynecological cancer and mesophylioma493,494. Further weak screening for PFKFB3 kinase optimization revealed a more selective PFKFB3 inhibitor compound 26495. PFKFB3 has been proven to be involved in homologous recombination repair of DNA double-strand breaks496. The new PFKFB3 inhibitor KAN0438757 inhibits the incorporation of deoxynucleotides during DNA repair by targeting PFKFB3, inhibits HR protein recruitment and reduces HR activity, and ultimately inhibits U2OS cell proliferation and growth496. A study proved that the level of PFKFB4 was negatively correlated with disease-free survival (DFS) and OS in tumor samples of breast cancer patients, which was associated with poor prognosis497. Knocking down PFKFB4 can inhibit tumor cell proliferation and metastasis, and overcome resistance to chemotherapy drugs498. Although research progress on PFKFB4 inhibitors has been slow, some studies have found that a compound 5MPN exerts anti-proliferative effects by inhibiting the PFKFB4 target in H460 adenocarcinoma cells499. The addition of 5MPN can also improve the sensitivity of carfilzomib to tumors499.
5.1.1.5. The others
Metformin, as a hypoglycemic drug, has become an effective treatment for cancer. It can limit the function of mitochondria, and participate in activating AMPK pathway, reducing the use of glucose, thus inhibiting the Warburg effect500,501. Warburg effect has played a significant role in promoting the growth, proliferation, diffusion and metastasis of colorectal cancer cells, creating suitable environmental conditions for the proliferation and metastasis of cancer. 2,3-Dihydro-2-(naphthalen-1-yl) quinazoline-4 (1H)-one (DHNQ), as a new PI3K pathway inhibitor, down-regulates the metabolic flux of colorectal cancer cells at the gene level, thus inhibiting Warburg effect to limit the proliferation, metastasis and apoptosis of colorectal cancer cells502. Interestingly, the new anti-diabetes drug sodium-glucose cotransporter 2 inhibitor canagliflozin has been repurposed for the treatment of liver cancer, which mainly inhibits the uptake of glucose by liver cancer cells in a glucose-dependent manner to inhibit angiogenesis in the tumor, impede the aerobic glycolysis process of liver cancer, and ultimately suppress the growth of cancer cells503. The HDAC inhibitor romidepsin can reduce ATP production in a c-Myc dependent manner, while the FAO inhibitors panobinostat or vorinostat can also hinder the occurrence of OXPHOS. The combination of HDAC inhibitors and FAO inhibitors can exert a stronger synergistic effect in GBM compared to single therapy504. The structural homolog of tumor suppressor factor p53, TAP73, is overexpressed in human gastric adenocarcinoma cell line AGS-1130. Casein kinase 2 (CK2) inhibitor CX-4945 inhibits AGS-1 cell uptake of glucose and release of lactate by inhibiting TAP73 function, inhibits aerobic glycolysis, and ultimately inhibits gastric cancer growth505,506. Recently, researchers have discovered a novel potential LDHA inhibitor, 1,3-benzodioxane derivative, which has good anti-proliferative activity against various types of tumor cell lines in vitro and may become a potential candidate drug for LDHA inhibitors507.
5.1.2. Traditional Chinese medicine (TCM)
Traditional Chinese medicine plays an irreplaceable role in the comprehensive treatment of tumors, but its efficacy mechanism still needs to be further clarified. The regulation of traditional Chinese medicine on tumor glucose metabolism is a feasible new direction. The diversified components of traditional Chinese medicine make it have the characteristics of multiple pathways and multiple effects. It has unique advantages in the treatment of complex and multifactorial tumor syndrome.
5.1.2.1. Compound recipe and simple recipe
Compound recipe is a classic way of clinical application of traditional Chinese medicine. A few studies have reported the effect of compound recipe on tumor glucose metabolism, which deserves more active and in-depth research. Shenmai liquid was used to treat S180 tumor-bearing mice with spleen deficiency. Shenmai liquid could significantly inhibit tumor growth and reduce the total LDH activity of tumor tissue; Shenmai liquid did not affect the activity of LDH1, but decreased the activity of LDH5, suggesting that the reduction of pyruvate to lactic acid was inhibited508. The serum containing Xiaochaihu decoction can promote the differentiation of human hepatoma cell line SMMC-7721 and block the cell in G0/G1 phase. The mechanism is related to the reduction of LDH activity in cells and the up-regulation of SDH activity509. Pharmacological studies of single drugs are simpler compared to compound drugs and can consider the characteristics of multiple components and multiple effects of traditional Chinese medicine more comprehensively than single drugs. Therefore, the study of single drug in tumor metabolism is slightly more than that of complex drug. Ginseng oil can reduce the SDH content of SGC-823 gastric cancer cells, suggesting that the tricarboxylic acid cycle may be damaged510. Danshen combined with 5-fluorouracil can reduce the microvessel density of Lewis lung cancer tissue in mice. Its mechanism is related to the reduction of HIF-1α mRNA expression in tumor tissue and the down-regulation of VEGF content in serum511 (Fig. 5B).
5.1.2.2. Monomer of traditional Chinese medicine
Monomer is the most basic component of a single drug. The study of effective monomers of traditional Chinese medicine helps to understand the material basis and mechanism of action of traditional Chinese medicine. Because of its clear structure and ease to accurately quantify, the research on the role and mechanism of monomer in tumor metabolism is more in-depth than that of compound and single drugs (Table 7) (Fig. 5B). The water extract BZL101 of Scutellaria barbata D. Don has selective cytotoxicity to breast cancer cells, which is mainly due to inducing breast cancer cells to produce a large number of ROS, which then leads to DNA damage. The excessive activation of poly (ADP ribose) polymerase (PARP) caused by DNA damage further inhibits cell aerobic glycolysis, which shows that LDH activity and lactic acid production are reduced, which ultimately leads to ATP reduction and necrosis512, 513, 514. Arctigenin selectively induced the necrosis of human non-small cell lung cancer A549 cells which were deprived of glucose. The mechanism was that the inhibition of mitochondrial respiration reduced the cellular ATP content, resulting in the increase of ROS. Glucose deprivation was simulated with aerobic glycolysis inhibitor 2-DG, which had synergistic effect with arctigenin, and had selectivity to tumor cells, but low toxicity to normal cells515.
Table 7.
Traditional Chinese medicine.
| Name in the literature | Natural source | Chemical structure | Cancer cell line (activity) | Mechanism | Ref. |
|---|---|---|---|---|---|
| BZL101 | Scutellaria barbata D. Don | Skbr3 (0.5 mg/mL); BT474 (0.5 mg/mL); U937 (IC50 ˜ 10 μg/mL) |
Produce ROS, the excessive activation of PARP caused by DNA damage further inhibits cell aerobic glycolysis | 512, 561 | |
| Arctigenin | Arctium lappa L. | ![]() |
A549 (IC50 ˜ 10 μmol/L, glucose-deprived, 8 h) | Inhibition of mitochondrial respiration reduced the cellular ATP content, resulting in the increase of ROS | 515 |
| Epoxycaryophyllene | Annona squamosa L. Bark. | ![]() |
Transform PKM2 into PKM1 by inhibiting aerobic glycolysis | 517 | |
| Oleanolic acid | Ligustrumlucidum; Araliachinensis | ![]() |
PC-3; MCF-7 |
Regulate c-Myc-mediated PKM selective splicing by inhibiting the mtor pathway, resulting in a decrease in PKM2 and an increase in PKM1 | 518 |
| Shikonin | Lithospermum erythrorhizon Sieb. Et Zucc. | ![]() |
MCF-7 (PKM2 IC50 = 0.8 μmol/L) | Selectively inhibit the activity of PKM2 | 30 |
| Alkannin | Lithospermum erythrorhizon Sieb. Et Zucc. | ![]() |
MCF-7 (PKM2 IC50 = 0.9 μmol/L) | Selectively inhibit the activity of PKM2 | 30 |
| Cannabinoids | Cannabis sativa L. | ![]() |
Panc1 | Inhibit the Akt/c-myc pathway to reduce the expression and activity of PKM2 | 519 |
| Arsenic trioxide | Arsenolite/Arsenopyrite; Realgar; Orpiment | As2O3 | HEK293T; SGC7901; HK2 Kd = 12.3 nmol/L |
Inhibit HK2 | 520 |
| Resveratrol | Reynoutria japonica Houtt.; Senna tora (L.) Roxb.; Morus alba L. | ![]() |
HCC-LM3 (IC50 = 59.46 μmol/L); Bel-7402 (IC50 = 81.57 μmol/L); Huh-7 (IC50 = 111.37 μmol/L) |
Inhibit aerobic glycolysis, ultimately promotes mitochondrial apoptosis, inhibits cell proliferation | 521 |
| Oroxylin A | Scutellaria baicalensis Georgi | ![]() |
MDA-MB-231; MCF-7 |
Induce dissociation of HK2 from the mitochondria and inhibits aerobic glycolysis by SIRT3-mediated deacetylation of cyclophilin D | 524 |
| Methyl-jasmonate | Styrax grandiflorus Griff.; Tunisian rosemary oil | ![]() |
CT-26 | Promotion of HK2 dissociation from mitochondrial VDAC and the reduction of intracellular ATP level | 525 |
| Epigallocatechin gallate | Green tea | ![]() |
Tca8113 (40 μmol/L); Tscca (40 μmol/L) |
Reduce the expression of HK2 protein by inhibiting AKT pathway | 527 |
| Curcumin | Curcuma aromatica Salisb.; C. longa L.; C. zedoaria (Berg.) Rosc.; Acorus calamus L. | ![]() |
HCT116 (IC50≈40 μmol/L); HT29 (IC50≈50 μmol/L) |
Down-regulating the expression and activity of HK2 | 530 |
| Quercetin | Styphnolobium japonicum (L.) Schott | ![]() |
SMMC-7721 (50 μmol/L); Bel-7402 (50 μmol/L) |
Reduce HK2 and Akt-mTOR Pathway | 531 |
| Ginsenoside Rh4 | Ginseng | ![]() |
KYSE 150 (IC50≈60 μmol/L, 24 h) | Suppresses the expression of PD-L1 via targeting AKT | 532 |
| Cardamonin | Alpinia katsumadai Hayata | ![]() |
SKOV3 (20 μmol/L) | Inhibition of aerobic glycolysis by mTOR; Down-regulate HIF-1α-mediated cell metabolism |
533, 539 |
| Halofuginone | Dichroa febrifuga Lour. | ![]() |
SW480 (IC50 = 24.83 nmol/L); HCT116 (IC50 = 5.82 nmol/L); SW620 (IC50 = 40.76 nmol/L); HT29 (IC50 = 40.76 nmol/L); DLD-1 (IC50 = 60.89 nmol/L) |
Suppression of Akt/mTORC1 signaling and glucose metabolism | 533 |
| Salvianolic acid B | Salvia miltiorrhiza Bunge | ![]() |
Cal27 (IC50 = 297 μmol/L); HN4 (IC50 = 201 μmol/L); Leuk1 (IC50 = 294 μmol/L) |
PI3K/AKT/HIF-1α signal pathway | 535 |
| Dihydroartemisinin | Artemisia annua L. | ![]() |
Lncap (IC50≈30 μmol/L) | PI3K/AKT/HIF-1α signal pathway | 536 |
| Tanshinone ⅡA | Salvia miltiorrhiza Bunge | ![]() |
Siha (IC50≈0.75 mg/L, 48 h); Hela (IC50≈2 mg/L, 48 h); C33a (IC50≈4 mg/L, 48 h); CAL27 (IC50≈2 μmol/L); SCC15 (IC50≈5 μmol/L) |
PI3K/AKT/HIF-1α signal pathway; Reducing AKT/c-Myc signal |
537, 555 |
| Ginsenoside Rg3 | Ginseng | ![]() |
Atp4a−/− mice | PI3K/AKT/HIF-1α signal pathway | 538 |
| Β-Asarone | Rhizoma Acori Tatarinowii | ![]() |
MGC803 (IC50 = 39.92 μg/mL); SGC7901 (IC50 = 84.6 μg/mL); MKN74 (IC50 = 96.22 μg/mL) |
Reduces the expression of PDK1, PDK4, HIF1α and c-Myc | 540 |
| Eleutheroside E | Acanthopanax senticosus | ![]() |
Inhibit the Ras-related protein RAP-1A and aerobic glycolysis | 543, 544 | |
| Chrysin | Oroxylum indicum (L.) Vent. | ![]() |
DU145 | Promote the degradation of HIF-1α, inhibit the expression of VEGF | 545 |
| Wogonin | Scutellaria baicalensis Georgi | ![]() |
HCT116 (RF = 2.05) | Inhibit the activation of PI3K/AKT and HIF-1α signal pathway mediated by hypoxia | 546 |
| Apigenin | Apium graveolens L. Var. Dulce DC. | ![]() |
CD18 (50 μmol/L) S2-013 (50 μmol/L) |
Down-regulation of GLUT1 expression depended on the inhibition of HIF-1α | 549 |
| Silibinin | Silybum marianum | ![]() |
Hepg2 (IC50 = 68 μmol/L) | Target GLUT-4 to inhibit glucose uptake | 550, 551, 552 |
| Xanthohumol | Humulus lupulus L. | ![]() |
U87 (IC50 ˜ 5 μmol/L, 72 h); T98G (IC50 ˜ 5 μmol/L, 72 h); LN229 (IC50 ˜ 5 μmol/L, 72 h); |
Down-regulating c-Myc and HK2 | 554 |
| Andrographolide | Andrographis paniculata (Burm. F.) Nees | ![]() |
C2C12 | Inhibition of NF-κB pathway and glycolytic enzyme HK2 | 556 |
| Polydatin | Polygonum cuspidatum Sieb. Et Zucc. | ![]() |
4T1 (IC50 = 66.56 μmol/L); MCF-7 (IC50 = 103.1 μmol/L) |
Inhibit ROS/PI3K/AKT/HIF-1α/HK2 signal axis | 557 |
| Betulinic Acid | Betula platyphylla Suk. | ![]() |
MCF-7 (IC50 = 19.06 μmol/L); MDA-MB-231 (IC50 = 48.55 μmol/L) |
Suppresses aerobic glycolysis via caveolin-1/NF-kappab/c-Myc pathway | 558 |
| Astragaloside IV | Astragalus membranaceus (Fisch) Bge. Var. Mongholicus (Bge.) Hsiao; Astragalus membranaceus (Fisch) Bge. | ![]() |
MDA-MB-231/ADR (40 μg/mL); BT-549/ADR (40 μg/mL); MDA-MB-549/ADR (40 μg/mL) |
Inhibit hsa_circ_0001982-mir-206/mir-613 axis | 560 |
Previous studies have shown that PKM2 can play a role as a transcriptional coactivator. Once PKM2 enters the nucleus, it can promote the transcription of target genes, thus promoting the growth of cancer cells and positive feedback regulated aerobic glycolysis516. Epoxycaryophyllene is a natural compound widely distributed in plants. It can transform PKM2 into PKM1 by inhibiting aerobic glycolysis, thus exerting its unique anti-cancer effect517. At present, clinical research on the treatment of colorectal cancer with epoxycaryophyllene is being carried out. With the in-depth study of the mechanism of Warburg effect of this drug, it may become a new choice for the treatment of colorectal cancer. Inhibition of aerobic glycolysis is one of the anti-tumor mechanisms of oleanolic acid. Oleanolic acid regulates c-Myc-mediated PKM selective splicing by inhibiting the mTOR pathway, resulting in a decrease in PKM2 and an increase in PKM1, and promoting metabolism from aerobic glycolysis to aerobic respiration518. Shikonin and its enantiomer alkannin can selectively inhibit the activity of PKM2, and their inhibition of aerobic glycolysis is not affected by the drug sensitivity of tumor, suggesting that they have certain application potential in the treatment of drug-resistant tumors30. Cannabinoids play an anti-tumor role by inhibiting the Akt/c-myc pathway to reduce the expression and activity of PKM2, thereby inhibiting aerobic glycolysis and glutamine absorption519.
HK2 is highly expressed in most cancer tissues, and studying its mechanism in the glycolytic pathway will be a focus of tumor treatment. Arsenic trioxide (ATO) has anti-tumor effect, and inhibiting its activity by binding with HK2 is the mechanism of its inducing apoptosis. Overexpression of HK2 can resist the apoptosis of human gastric cancer SGC7901 cells induced by arsenic. The metabonomic analysis showed that ATO could play a role in promoting apoptosis by inhibiting the activity of HK2 and inhibiting the aerobic glycolysis of tumor520. Resveratrol inhibits aerobic glycolysis, ultimately promotes mitochondrial apoptosis, inhibits cell proliferation, and reduces the resistance of hepatocellular carcinoma to sorafenib, which has a certain synergistic effect521,522. It is worth noting that the mechanism of oroxylin A inhibiting tumor aerobic glycolysis is related to cell type and oxygen environment523,524. Similarly, the anti-tumor activity of methyl jasmonate is also related to the promotion of HK2 dissociation from mitochondrial VDAC and the reduction of intracellular ATP level525,526. Epigallocatechin gallate can inhibit the anchored independent growth of human tongue squamous cell carcinoma in a dose-dependent manner, reduce the expression of HK2 protein by inhibiting AKT pathway to weaken aerobic glycolysis, and inhibit the binding of HK2 with mitochondria to promote apoptosis527, 528, 529. Curcumin has the effects of growth inhibition and apoptosis induction on human colorectal cancer cells, and HK2 is its important target: by down-regulating the expression and activity of HK2 protein, curcumin down-regulates the aerobic glycolysis of tumor cells, thus inhibiting the production of ATP530. It also provides a new perspective for the application of anti-inflammatory Chinese medicine in tumor prevention and treatment.
At present, many traditional Chinese medicine compounds have been found to have regulatory effects on mTOR. Quercetin and ginsenoside Rh4 reduce the expression of aerobic glycolysis-related proteins and inhibit aerobic glycolysis through AKT–mTOR pathway531,532. Inhibition of aerobic glycolysis by mTOR is a key step in cardamonin -induced autophagy533. Halofuginone is a derivative of halofuginone alkaloid extracted from the Chinese herbal medicine Dichroa febrifuga Lour., it exerts its potential anticancer effect by inhibiting Akt/mTORC1 signal transduction and inhibiting aerobic glycolysis pathway. In the tumor cells of colorectal cancer patients treated with halofuginone, mTORC1 and phosphorylated Akt are significantly inhibited, further leading to rapid reduction of hexokinase 2 and glucose transporter 1, thus achieving the effect of inhibiting tumor cell growth534. In general, halofuginone inhibits glucose uptake and aerobic glycolysis in colorectal cancer cells by regulating Akt/mTORC1 signal pathway, thus inhibiting the growth of cancer cells in vitro and in vivo534.
HIF is an important node of cancer metabolism, and also a research hotspot of cancer metabolism targeted by traditional Chinese medicine. The regulation of Warburg effect by HIF is closely related to PI3K/AKT pathway. Studies have shown that salvianolic acid B, dihydroartemisinin, tanshinone IIA and ginsenoside Rg3 exert antitumor activity by regulating abnormal glucose metabolism through PI3K/AKT/HIF-1α signal pathway535, 536, 537, 538. Some traditional Chinese medicine compounds have been proved to have multiple targets and pathways in regulating cancer metabolism. For example, cardamonin can not only down-regulate HIF-1α-mediated cell metabolism, but also have a significant impact on glucose uptake and lactic acid production and efflux539. β-Asarone reduces the expression of several key genes (PDK1, PDK4, HIF1α, c-Myc, etc.) in aerobic glycolysis, achieving chemosensitivity and inhibition of tumor aerobic glycolysis540. It shows that the metabolism of cachexia cannot be separated from the original energy metabolism pathway of cells, and the advantages of multi-molecular target of traditional Chinese medicine may play an advantage in cancer metabolism. ROS is not only a potential mutagen of cancer, but also a factor that stimulates aerobic glycolysis541. Carcinogenic K-Ras inhibits mitochondrial respiration by down-regulating mitochondrial respiratory chain complex I, and up-regulates NOX1 production to promote ROS production and HIF accumulation542. The effective components eleutheroside E of Acanthopanax senticosus can inhibit the Ras-related protein RAP-1A, aerobic glycolysis, and maintain the normal nerve activity of mice with ischemia and hypoxia, but whether it can play a role in tumor cells remains to be further studied543,544. Chrysin can promote the degradation of HIF-1α, inhibit the expression of VEGF, and inhibit the angiogenesis of human prostate cancer DU145 transplanted tumor545. In vitro and in vivo experiments showed that wogonin could inhibit the activation of PI3K/AKT and HIF-1α signal pathway mediated by hypoxia, reduce cell glucose uptake, inhibit the expression of glycolytic enzymes HK2, PDK1 and LDHA, and inhibit tumor aerobic glycolysis, thus reversing the hypoxia-induced cisplatin resistance of human colon cancer cells546. The inhibition of apigenin on the proliferation and apoptosis of tumor cells was accompanied by the down-regulation of GLUT1 expression. Further research found that the inhibition of celery on GLUT1 expression could reduce the glucose uptake of tumor cells, which depended on the inhibition of HIF-1α to a certain extent547, 548, 549. Silibinin can also selectively target GLUT-4 to inhibit glucose uptake and exert anti-angiogenic activity and inhibit the growth of cancer cells in a variety of cancers550, 551, 552. In tumor cells overexpressed with Myc, there will be simultaneous transcription of Myc and HIFα sensitive genes, which will eventually stimulate the conversion of hypoxic aerobic glycolysis553. Xanthohumol leads to aerobic glycolysis inhibition by down-regulating c-Myc and subsequent HK2 inhibition554. Tanshinone ⅡA can also inhibit oral squamous cell carcinoma by reducing AKT/c-Myc signal mediated aerobic glycolysis555.
NF-κB is not only a cancer-promoting factor, but also plays an important role in inflammation and cancer metabolism. It has always been an important target of anticancer drug research. For further exploration of commonly used anti-inflammatory drugs in clinical practice, it may be possible to find effective compounds targeting NF-κB or other inflammatory targets. Inflammation can cause the activation of IKK kinase family, release the transcription factor NF-κB, and ultimately induce the expression of various target genes including HIF1α556. Andrographis paniculata and Reynoutria japonica houtt are commonly used in clinical anti-inflammatory treatment. Andrographolide can significantly reduce the levels of IL-1β and IL-6, and down-regulate the expression of iNOS and COX-2. Its anti-inflammatory effect is related to the inhibition of NF-κB pathway and glycolytic enzyme HK2, but whether it can affect tumor cells requires further investigation556. Polydatin inhibits glycolytic phenotype by inhibiting ROS/PI3K/AKT/HIF-1α/HK2 signal axis and enhances its anti-cancer effect557. Betulinic acid inhibits aerobic glycolysis of breast cancer cells by regulating Cav-1/NF-κB/c-Myc pathway558. In addition, in the study of the reversal mechanism of astragaloside IV on precancerous lesions of gastric cancer, it was found that astragaloside IV can reduce abnormal aerobic glycolysis by regulating the expression of p53, TIGAR and other related glycolytic proteins559. It could also reverse the drug resistance of TNBC by blocking the hsa_circ_0001982-miR-206/miR-613 axis and inhibiting aerobic glycolysis560.
5.2. Combination strategies
Tumor metabolism involves multiple genes and pathways with complex regulation. Inhibition of a single signal pathway or a single target may not be sufficient to treat tumor, and may even lead to drug resistance. Activation of bypass pathway is one of the important reasons. To overcome the drug resistance caused by the activation of the bypass, at present, the way of combining multiple signal pathway inhibitors horizontally is mostly adopted. In addition, because each signal pathway will ultimately affect the cell behavior by regulating the downstream metabolism, the longitudinal combination of signal pathway and metabolic pathway inhibitors may also produce enhanced anti-tumor effect. Therefore, exploring the complementary metabolic pathway targeted by combination therapy may enhance or synergistically inhibit the survival of tumor cells. With the introduction of the hybrid OXPHOS/glycolytic metabolism model, a new idea for the comprehensive treatment of tumor metabolism has been proposed. In vitro experiments confirmed that metformin combined with aerobic glycolysis inhibitor 2-DG can activate AMPK and reduce the phosphorylation of mTORC1-regulated protein A, and induce cancer cell death at both cellular and animal levels, providing a preliminary basis for comprehensive targeted inhibition500. It is suggested that combined inhibition of aerobic glycolysis and OXPHOS activity in tumor cells, along with precise blocking of their metabolic pathways, may eliminate the metabolic plasticity of tumor cells and enhance the efficacy of tumor metabolic regulation therapy. Metformin and fasting therapy can cooperate to inhibit tumor growth through PP2A–GSK3β–MCL-1 pathway, but their safety requires further experimental confirmation562. In addition, the combination of chemotherapy drugs and metabolic-targeted therapeutic drugs can reduce the dose of chemotherapy drugs and the adverse reactions of chemotherapy. On the other hand, it can regulate the metabolic mode of tumor cells, overcome the chemotherapy resistance of tumor cells, and improve the treatment efficacy of tumor patients. This will be a promising treatment strategy563. For example, shikone, a small molecule active substance, can prevent tissue plasminogen activator (TPA) mediated skin cells from transforming into cancer at an early stage by inhibiting PKM2. It can repair the damage of mitochondrial function caused by TPA, reduce the production of lactic acid, an aerobic glycolysis marker, and also overcome the resistance of bladder cancer to cisplatin through mediating necrosis463,564. The AMPK inducer AICAR and methotrexate (MTX) often produce drug resistance when used alone in breast cancer. One study found that the combination of the two can block the aerobic glycolysis process by promoting mitochondrial oxidation, inhibit the occurrence of warburg effect, and finally reverse the phenomenon of slow proliferation of breast cancer cells565 (Fig. 5C).
5.3. New strategies
With the development of modern molecular biology technology and the development of targeted gene research, the importance of aerobic glycolysis pathway in tumor cell growth, invasion and other processes has been ignored for a long time. In recent years, fluorodeoxyglucose positron emission tomography (FDG-PET) technology based on the “Warburg effect” has made the research on tumor energy metabolism rise again. Using this PET imaging technology to detect patients, we can judge the location of the primary tumor and metastatic tumor in the body by where the glucose uptake in the body is significantly enhanced, providing a means for efficient and specific diagnosis and efficacy monitoring of tumors566. As a highly metastatic and fatal subtype of breast cancer, triple-negative breast cancer (TNBC) remains an unresolved challenge in clinical identification and diagnostic imaging401,567,568. GLUT1 inhibitor probe provides a potential use for aerobic glycolysis diagnostic imaging of TNBC569. However, due to the low contrast of imaging due to impaired physiological aerobic glycolysis, FDG/PET diagnostic results are not ideal, and more sensitive and accurate identification is still required570. GLUT1 and GLUT3 transport glucose analogues FDG into tumor cells571, while the decreased sensitivity of FDG-PET detection in the diagnosis of liver cancer leads to abnormal expression of GLUT2 in liver cancer tissue572. Unfortunately, the role and mechanism of GLUT2 in the development of liver cancer are still unclear. PET/CT integrates FDG-PET images with CT images to make tumor localization more accurate. Due to differences in GLUT-1 expression on various types of cancer cell membranes, its applicability is limited. Mucous adenocarcinoma, signet ring cell carcinoma and poorly differentiated cancer tissue have low expression of GLUT-1, and the uptake of glucose is less than that of intestinal cancer559. In addition, the tumor size, stage and the degree of gastric expansion will affect the FDG uptake. Compared with gastroscopy and enhanced CT, its sensitivity and specificity are poor. Although PET/CT has poor sensitivity in detecting primary lesions, it is of great significance in evaluating microsatellite instability of cancer, asymptomatic progressive cancer and neoadjuvant treatment of cancer, especially neoadjuvant treatment573. In order to improve the sensitivity and specificity of PET/CT, there are currently new tracers such as 18F-FLT targeting proliferation, 18F-FMISO targeting tumor hypoxia, and radiolabeled choline derivatives targeting phosphate metabolism574. magnetic resonance imaging (MRI) with new sequence has outstanding performance in primary tumor and lymph node staging, and PET/MRI has more advantages in detecting tumor, lymph node and distant metastasis575.
Warburg effect has been confirmed in different types of tumors, and FDG-PET has been used to detect cancer cells and applied to clinical practice. However, there are potential limitations when FDG-PET is used to detect certain types of cancer, such as colorectal cancer, that is, there are some false positives, which may lead to some colorectal cancer patients being wrongly judged by FDG-PET when detecting disease recurrence. Nevertheless, with the advent of “one-stop shop” imaging and contrast enhanced FDGPET, the integration of new PET trackers and PET magnetic resonance imaging (MRI), it may play an increasing role in the evaluation and management of cancer patients. The metabolic reprogramming property of tumor cells can enhance the aerobic glycolysis process, which is mainly manifested in the tumor microenvironment and is closely related to the down-regulation of mitochondrial H+-ATPase catalytic subunit. It can be used as a pre-indicator of patients by detecting the glucose metabolism in tumor cells. The further study on the relationship between FDG-PET, PET-MRI and cancer cell metabolic reprogramming characteristics and Warburg effect will provide a new direction and idea for clinical detection of cancer (Fig. 6A).
Figure 6.
New and potential strategies for cancer treatment targeting the Warburg effect. (A) New strategies include fluorodeoxyglucose positron emission tomography (FDG-PET) technology, FDG-PET images with CT images (PET/CT) technology, and positron emission tomography magnetic resonance imaging (PET-MRI) technology. (B) Potential strategies involve synthetic lethal combination therapy and metabolomics which is used for tumor typing.
5.4. Potential strategies
Nowadays, cancer targeted therapy has become an inevitable trend of precision medicine. However, the target of targeted therapy mainly focuses on cancer mutation genes and pathways, which severely limits the types of drug targets. Mutations that lead to the development of cancer will also bring about weaknesses that can be used for treatment. Drug treatment of cancer also depends on such a concept. Large-scale cancer genome sequencing has classified mutations of various cancer types, which can be used to explore the vulnerability of cancer. These gene changes include functional acquisition mutations (gene amplification, translocation or mutation) and functional loss mutations (gene function is impaired due to missense mutations or deletions). The concept of synthetic lethality, that is, the presence of cancer gene mutations is usually associated with a new vulnerability that can be targeted for treatment, can help us greatly expand the arsenal of potential cancer drug targets. Cancer cells usually maintain metabolic activity and resist anti-glycolytic therapy through metabolic reprogramming. Therefore, synthetic lethal combination therapy is crucial to reduce the therapeutic effect of Warburg metabolism. A study found that chemical synthetic lethal agents for the loss of VHL in renal cancer are usually lethal to cancer cells with high GLUT1 levels and need aerobic glycolysis82, suggesting that the method of screening compounds with lethal synthesis may also be applicable to other tumor types with loss of tumor suppressor gene function or functional gain of oncogene.
Additionally, individualized precise treatment can be tailored based on the different molecular typing and metabolic characteristics of tumor cells. As a new technology, metabonomics aims to study the abnormal metabolic patterns of tumors, and has shown great potential in the diagnosis of metabolic typing and individualized treatment of tumors. According to the metabolic differences between breast cancer tumor cells and stromal cells, researchers divided them into four metabolic phenotypes: Warburg type (tumor cells show aerobic glycolysis characteristics, while stromal cells do not); reverse Warburg type (stromal cells have glycolytic characteristics, but tumor cells do not); mixed metabolic type (both tumor cells and stromal cells have glycolytic characteristics); no marker type (tumor cells and stromal cells have no aerobic glycolysis characteristics500. According to the different metabolic phenotypes of tumor cells, pancreatic cancer can be divided into Warburg type, reverse Warburg type, lipid-dependent type and glutamine metabolic type576. The research and application of metabolic typing provide a metabolic basis for more effective tumor metabolic intervention. Currently, research on metabolism primarily focuses on in vitro cultured cells and tissues. The culture environment of in vitro cultured cells cannot accurately reflect the true situation of the in vivo environment. During the processing of in vitro tissues, rapid changes such as ischemia and hypoxia can lead to drastic changes in cell metabolism, and it is difficult to reflect the in situ metabolic changes of living tissue cells in animals. Therefore, it is urgent to develop new metabolite labeling techniques, in situ non-destructive testing of proteins and metabolites, and micro detection techniques. Glucose metabolism markers can also be applied to targeted tumor therapy by modifying tumor cells and immune cells577, 578, 579, 580. Among them, based on the self-modification of tumor cells, specific antigens can be artificially added to the surface of tumor cells to increase their targeting. For example, known haptens can be added to the surface of tumor cells, which can activate anti-tumor immunity through hapten carriers577. Additionally, it can link antigen target molecules of marketed tumor-specific antibody drugs, potentially addressing the dilemma of TNBC and other tumor types without drugs available579; Alternatively, by modifying specific antigens and using their paired antibodies to develop CAR-T cells, personalized tumor immunotherapy can be achieved580. As T cell and NK cell immunotherapy advances, modifications based on glucose metabolism markers can enhance target selectivity for CAR-T, TCR-T, and CAR NK cell therapies. However, due to the complexity of metabolic markers and the lack of cross-validation, a long road remains toward achieving fully standardized treatment.
Cancer cells meet their energy and biosynthesis needs by altering their own metabolic characteristics and enhancing aerobic glycolysis, which can promote cancer cell proliferation, reduce drug-induced cell apoptosis, and thus develop cancer resistance. A large amount of evidence suggests that metabolic disorders in cancer cells are closely related to cancer resistance during cancer treatment. For example, LDHA is associated with paclitaxel/trastuzumab resistance in the treatment of breast cancer581. PDK3 is associated with hypoxia-induced drug resistance in the treatment of cervical and colon cancer582. Fatty acid synthase (FAS) is associated with docetaxel/trastuzumab/doxorubicin resistance in the treatment of breast cancer, and gemcitabine and radiotherapy resistance in the treatment of pancreatic cancer80. GLS is associated with cisplatin resistance in the treatment of gastric cancer77. The molecular mechanisms underlying cancer drug resistance caused by metabolic disorders in cancer cells are extremely complex. The combination of GLUT1 inhibitor WZB117 and cisplatin or paclitaxel can also demonstrate synergistic anticancer effects583. Ritonavir can also inhibit the proliferation of primary myeloma cells and increase the sensitivity of cancer cells to doxorubicin584. The combination of HK2 inhibitor 2-DG and ABT-737 can improve ABT-737 resistance585. In clinical trials, the combination of PDK inhibitor dichloroacetate (DCA) and omeprazole has shown synergistic anticancer efficacy586. GLS1 inhibitor dimethyl-2[-5-phenylacetyl-1,2,4-thiadiazo-2-yl] ethyl sulfide (BPTES) can improve the therapeutic efficacy of mutant IDH1 patients587. The combination of FAS inhibitor G28UCMM, trastuzumab and lapatinib showed a good synergistic effect588. The aerobic glycolysis of malignant tumor cells is also closely related to tumor radiation resistance589,590. A study has confirmed that in hypoxic malignant tumors, reducing lactate content and increasing oxygen consumption based on the Warburg effect can increase their sensitivity to fractionated radiotherapy and improve patient prognosis591. These research results indicate a close relationship between metabolic disorders in cancer cells and resistance to cancer. Targeting key enzymes in the metabolic process of cancer cells can be used to improve cancer resistance and enhance the efficacy of chemotherapy drugs in cancer patients. However, the molecular mechanism underlying cancer resistance caused by metabolic disorders is not fully understood and further research is needed. The combination of cancer chemotherapy and anti-metabolic therapy, combined with individualized treatment concepts, will further improve the efficacy of cancer patients (Fig. 6B).
6. Discussions
Activation of oncogenes and the deletion of tumor suppressors promote metabolic reprogramming in cancer, leading to increased nutrient absorption for energy and biosynthetic pathways. Thus, metabolic reprogramming is considered a characteristic of cancer29. Cancer cells utilize a higher rate of aerobic glycolysis to ingest glucose for metabolic synthesis, providing an evolutionary advantage and more biosynthetic substances for tumor cell growth and reproduction353. This energy metabolism characteristic, known as the Warburg effect, plays a crucial role in maintaining high proliferation rates and increasing anti-apoptotic characteristics in cancer cells28.
During cancer occurrence and development, the Warburg effect influences the evolution of cancer by affecting the energy metabolism of cancer cells, particularly the aerobic glycolysis pathway, and by participating in the expression of common transcription regulatory factors and proteins such as FOXM1, p53, NF-κB, HIF1α, and c-Myc. Key enzymes involved in the Warburg effect include GLUTs, HKs, PFKs, LDHs, and PKM2, which significantly contribute to cancer occurrence and development. High expression of these key glycolytic enzymes in patients is associated with distant metastasis, deeper tumor invasion, and later clinical stages.
LncRNAs, miRNAs, and circular RNAs play crucial roles in regulating the glucose metabolism of cancer cells by targeting genes involved in metabolism, thereby promoting cancer cell growth, proliferation and metastasis232. Limiting the energy intake of tumor cells as a strategy for cancer treatment and prevention has garnered increasing interest among scientists. This concept has shown remarkable results in animal and human experiments592,593. The principle of taking energy metabolism as an anti-tumor strategy mainly depends on the difference of energy production mechanism between normal cells and transformed cells594. However, the differences in energy production mechanisms between normal cells and cancer cells must be considered when using energy metabolism as an anti-tumor strategy. Tumor cells are highly vulnerable to limitations in energy supply, making them an attractive target for further research on factors or substances that can influence the Warburg effect and affect the growth, reproduction, and metastasis of cancer cells.
Traditional Chinese medicine offers multiple channels and effects in the comprehensive treatment of tumors due to the cooperation, synergism, and network interactions between its diversified components. However, it also poses challenges in terms of efficacy, pharmacological mechanism, and drug interactions. Tumor cells exhibit metabolic heterogeneity due to differences in genotypes, types, degrees of differentiation, stages, and microenvironments. The metabolic coupling between tumor cells and the tumor microenvironment can induce changes in tumor cell metabolism, enabling effective adaptation to external stress and facilitating malignant progression. Therefore, targeted tumor metabolic therapy requires accurate diagnosis and individualized typing detection. Treatment should be selected based on the metabolic dependence and specific metabolic phenotypes of tumor cells, leading to different intervention measures and the transformation of metabolic phenotypes to achieve the best therapeutic effect.
Metabolic reprogramming is also an important feature of tumor drug resistance. Understanding the metabolic process underlying drug resistance in tumor cells can shed light on the molecular mechanisms involved. Uncontrolled metabolic processes have been found to overcome drug resistance in tumor cells, and inhibitors targeting metabolic enzymes have shown significant effects in reversing tumor drug resistance, with some entering phase I clinical trials. Discovering molecular markers related to drug resistance can accurately predict drug resistance, metastasis, and prognosis in tumors. Metabonomics has enabled more efficient and systematic studies on the reprogramming of metabolic networks in drug-resistant tumor cells.
The Warburg effect, originally used to describe increased lactic acid production in cancer, has enduring research significance and has been found to play important roles in regulating various life activities, including immunity, angiogenesis, pluripotency, reproduction, pathogen infection, macrophage polarization, and T cell activation595, 596, 597, 598. Exploring the multi-target comprehensive treatment potential of the Warburg effect is worth further investigation599, 600. By reducing the activity of enzymes related to aerobic glycolysis or controlling energy supply through dietary means, tumor cells can be exposed to a prolonged energy-deficient microenvironment. Understanding aerobic glycolysis has confirmed the potential for targeted therapy of metabolic changes in tumor cells, and recent studies have identified numerous potential pathways and targets for cancer treatment. Each tumor type has its own unique metabolic pathway, with most tumors exhibiting characteristics of the Warburg effect. While research on targeted therapy has focused mainly on key enzymes and rate-limiting enzymes in aerobic glycolysis and glutamine metabolism, limited research has been conducted on key enzymes in the TCA cycle. It is believed that more potential therapeutic targets will be discovered in the future. Challenges in targeted tumor metabolism include the non-specific tropism of metabolic inhibitors and their potential inhibition of immune cells, making combination therapy with metabolic inhibitors more effective in clinical practice.
Overall, cancer cells exhibit accelerated growth, proliferation, and metastasis pathways through the involvement of oncogenes, key enzymes, metabolic pathways, and other metabolic mechanisms. The mechanism of cancer occurrence and development, particularly its relationship with the Warburg effect, remains unclear and requires further exploration. In-depth studies on the relationship between the Warburg effect and cancer can provide new directions and approaches for the early diagnosis and treatment of cancer patients.
Acknowledgments
This work was supported in part by National Natural Science Foundation of China (Grant No. 82172649, 82003580, and 82173666), and Shenzhen science and technology research and development funds (Grant No. JCYJ20210324094612035, China).
Footnotes
Peer review under the responsibility of Chinese Pharmaceutical Association and Institute of Materia Medica, Chinese Academy of Medical Sciences.
Author contributions
Bo Liu, Dahong Yao and Jin Zhang conceived the project and supervised the project. Jin Zhang provided the new idea. Minru Liao drafted the tables and manuscript. Lifeng Wu draw the figures. Jin Zhang and Zhiwen Wang proofread the structure and figures. Bo Liu and Chaodan Luo polished the manuscript. All the authors read and approved the final manuscript.
Conflicts of interest
The authors have no conflicts of interest to declare.
References
- 1.Chen F., Chen J., Yang L., Liu J., Zhang X., Zhang Y., et al. Extracellular vesicle-packaged HIF-1α-stabilizing lncRNA from tumour-a ssociated macrophages regulates aerobic glycolysis of breast cancer cells. Nat Cell Biol. 2019;21:498–510. doi: 10.1038/s41556-019-0299-0. [DOI] [PubMed] [Google Scholar]
- 2.Ampferl R., Rodemann H.P., Mayer C., Hofling T.T.A., Dittmann K. Glucose starvation impairs DNA repair in tumour cells selectively by blocking histone acetylation. Radiother Oncol. 2018;126:465–470. doi: 10.1016/j.radonc.2017.10.020. [DOI] [PubMed] [Google Scholar]
- 3.Supabphol S., Seubwai W., Wongkham S., Saengboonmee C. High glucose: an emerging association between diabetes mellitus and cancer progression. J Mol Med (Berl) 2021;99:1175–1193. doi: 10.1007/s00109-021-02096-w. [DOI] [PubMed] [Google Scholar]
- 4.Feng H., Wang X., Chen J., Cui J., Gao T., Gao Y., et al. Nuclear imaging of glucose metabolism: beyond 18F-FDG. Contrast Media Mol Imaging. 2019;2019 doi: 10.1155/2019/7954854. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Warburg O. On the origin of cancer cells. Science. 1956;123:309–314. doi: 10.1126/science.123.3191.309. [DOI] [PubMed] [Google Scholar]
- 6.Liberti M.V., Locasale J.W. The Warburg effect: how does it benefit cancer cells?. Trends Biochem Sci. 2016;41:211–218. doi: 10.1016/j.tibs.2015.12.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Hu C., Pang B., Lin G., Zhen Y., Yi H. Energy metabolism manipulates the fate and function of tumour myeloid-derived suppressor cells. Br J Cancer. 2020;122:23–29. doi: 10.1038/s41416-019-0644-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Ancey P.-B., Contat C., Meylan E. Glucose transporters in cancer—from tumor cells to the tumor microen vironment. FEBS J. 2018;285:2926–2943. doi: 10.1111/febs.14577. [DOI] [PubMed] [Google Scholar]
- 9.Lopez-Lazaro M. The warburg effect: why and how do cancer cells activate glycolysis in the presence of oxygen?. Anti Cancer Agents Med Chem. 2008;8:305–312. doi: 10.2174/187152008783961932. [DOI] [PubMed] [Google Scholar]
- 10.Racker E. Why do tumor cells have a high aerobic glycolysis?. J Cell Physiol. 1976;89:697–700. doi: 10.1002/jcp.1040890429. [DOI] [PubMed] [Google Scholar]
- 11.Warburg O. Über den Stoffwechsel der Carcinomzelle. Naturwissenschaften. 1924;12:1131–1137. [Google Scholar]
- 12.Gonzalez C., Ureta T., Sanchez R., Niemeyer H. Multiple molecular forms of ATP:hexose 6-phosphotransferase from rat liver. Biochem Biophys Res Commun. 1964;16:347–352. doi: 10.1016/0006-291x(64)90038-5. [DOI] [PubMed] [Google Scholar]
- 13.Katzen H.M., Schimke R.T. Multiple forms of hexokinase in the rat: tissue distribution, age dependency, and properties. Proc Natl Acad Sci U S A. 1965;54:1218–1225. doi: 10.1073/pnas.54.4.1218. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Tarui S., Okuno G., Ikura Y., Tanaka T., Suda M., Nishikawa M. Phosphofructokinase deficiency in skeletal muscle. a new type of glycogenosis. Biochem Biophys Res Commun. 1965;19:517–523. doi: 10.1016/0006-291x(65)90156-7. [DOI] [PubMed] [Google Scholar]
- 15.Sokoloff L. Relation between physiological function and energy metabolism in the central nervous system. J Neurochem. 1977;29:13–26. doi: 10.1111/j.1471-4159.1977.tb03919.x. [DOI] [PubMed] [Google Scholar]
- 16.Schmidt K.C., Lucignani G., Sokoloff L. Fluorine-18-fluorodeoxyglucose PET to determine regional cerebral glucose utilization: a re-examination. J Nucl Med. 1996;37:394–399. [PubMed] [Google Scholar]
- 17.Kasahara M., Hinkle P.C. Reconstitution and purification of the d-glucose transporter from huma n erythrocytes. J Biol Chem. 1982;252:7384–7390. [PubMed] [Google Scholar]
- 18.Mueckler M., Caruso C., Baldwin S.A., Panico M., Blench I., Morris H.R., et al. Sequence and structure of a human glucose transporter. Science. 1985;229:941–945. doi: 10.1126/science.3839598. [DOI] [PubMed] [Google Scholar]
- 19.Wilson J.E. In: Regulation of carbohydrate metabolism. Beitner R., editor. CRC Press; Boca Raton: 1985. Regulation of mammalian hexokinase activity; pp. 45–85. [Google Scholar]
- 20.Shinohara Y., Yamamoto K., Kogure K., Ichihara J., Terada H. Steady state transcript levels of the type II hexokinase and type 1 glucose transporter in human tumor cell lines. Cancer Lett. 1994;82:27–32. doi: 10.1016/0304-3835(94)90142-2. [DOI] [PubMed] [Google Scholar]
- 21.Chandel N.S., Maltepe E., Goldwasser E., Mathieu C.E., Simon M.C., Schumacker P.T. Mitochondrial reactive oxygen species trigger hypoxia-induced transcription. Proc Natl Acad Sci U S A. 1998;95:11715–11720. doi: 10.1073/pnas.95.20.11715. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Agani F.H., Pichiule P., Chavez J.C., LaManna J.C. The role of mitochondria in the regulation of hypoxia-inducible factor 1 expression during hypoxia. J Biol Chem. 2000;275:35863–35867. doi: 10.1074/jbc.M005643200. [DOI] [PubMed] [Google Scholar]
- 23.King A., Selak M.A., Gottlieb E. Succinate dehydrogenase and fumarate hydratase: linking mitochondrial dysfunction and cancer. Oncogene. 2006;25:4675–4682. doi: 10.1038/sj.onc.1209594. [DOI] [PubMed] [Google Scholar]
- 24.Ratcliffe P.J. Fumarate hydratase deficiency and cancer: activation of hypoxia signaling?. Cancer Cell. 2007;11:303–305. doi: 10.1016/j.ccr.2007.03.015. [DOI] [PubMed] [Google Scholar]
- 25.Ueyama A., Sato T., Yoshida H., Magata K., Koga N. Nonradioisotope assay of glucose uptake activity in rat skeletal muscle using enzymatic measurement of 2-deoxyglucose 6-phosphate in vitro and in vivo. Biol Signals Recept. 2000;9:267–274. doi: 10.1159/000014649. [DOI] [PubMed] [Google Scholar]
- 26.Aft R.L., Zhang F.W., Gius D. Evaluation of 2-deoxy-d-glucose as a chemotherapeutic agent: mechanism of cell death. Br J Cancer. 2002;87:805–812. doi: 10.1038/sj.bjc.6600547. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Gatenby R.A., Gillies R.J. Why do cancers have high aerobic glycolysis?. Nat Rev Cancer. 2004;4:891–899. doi: 10.1038/nrc1478. [DOI] [PubMed] [Google Scholar]
- 28.Vander Heiden M.G., Cantley L.C., Thompson C.B. Understanding the Warburg effect: the metabolic requirements of cell proliferation. Science. 2009;324:1029–1033. doi: 10.1126/science.1160809. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Hanahan D., Weinberg R.A. Hallmarks of cancer: the next generation. Cell. 2011;144:646–674. doi: 10.1016/j.cell.2011.02.013. [DOI] [PubMed] [Google Scholar]
- 30.Chen J., Xie J., Jiang Z., Wang B., Wang Y., Hu X. Shikonin and its analogs inhibit cancer cell glycolysis by targeting tumor pyruvate kinase-M2. Oncogene. 2011;30:4297–4306. doi: 10.1038/onc.2011.137. [DOI] [PubMed] [Google Scholar]
- 31.Yi W., Clark P.M., Mason D.E., Keenan M.C., Hill C., Goddard W.A., 3rd, et al. Phosphofructokinase 1 glycosylation regulates cell growth and metabolism. Science. 2012;337:975–980. doi: 10.1126/science.1222278. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Deng D., Xu C., Sun P., Wu J., Yan C., Hu M., et al. Crystal structure of the human glucose transporter GLUT1. Nature. 2014;510:121–125. doi: 10.1038/nature13306. [DOI] [PubMed] [Google Scholar]
- 33.Webb B.A., Forouhar F., Szu F.E., Seetharaman J., Tong L., Barber D.L. Structures of human phosphofructokinase-1 and atomic basis of cancer-associated mutations. Nature. 2015;523:111–114. doi: 10.1038/nature14405. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Oginuma M., Harima Y., Tarazona O.A., Diaz-Cuadros M., Michaut A., Ishitani T., et al. Intracellular pH controls WNT downstream of glycolysis in amniote embryos. Nature. 2020;584:98–101. doi: 10.1038/s41586-020-2428-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Xu K., Yin N., Peng M., Stamatiades E.G., Shyu A., Li P., et al. Glycolysis fuels phosphoinositide 3-kinase signaling to bolster T cell immunity. Science. 2021;371:405–410. doi: 10.1126/science.abb2683. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Epstein T., Gatenby R.A., Brown J.S. The Warburg effect as an adaptation of cancer cells to rapid fluctuations in energy demand. PLoS One. 2017;12 doi: 10.1371/journal.pone.0185085. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Pavlova N.N., Zhu J., Thompson C.B. The hallmarks of cancer metabolism: still emerging. Cell Metabol. 2022;34:355–377. doi: 10.1016/j.cmet.2022.01.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Pasto A., Bellio C., Pilotto G., Ciminale V., Silic-Benussi M., Guzzo G., et al. Cancer stem cells from epithelial ovarian cancer patients privilege oxidative phosphorylation, and resist glucose deprivation. Oncotarget. 2014;5:4305–4319. doi: 10.18632/oncotarget.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Nakano I. Stem cell signature in glioblastoma: therapeutic development for a moving target. J Neurosurg. 2015;122:324–330. doi: 10.3171/2014.9.JNS132253. [DOI] [PubMed] [Google Scholar]
- 40.Codrici E., Enciu A.M., Popescu I.D., Mihai S., Tanase C. Glioma stem cells and their microenvironments: providers of challenging therapeutic targets. Stem Cell Int. 2016;2016 doi: 10.1155/2016/5728438. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Zhou Y., Zhou Y., Shingu T., Feng L., Chen Z., Ogasawara M., et al. Metabolic alterations in highly tumorigenic glioblastoma cells: preference for hypoxia and high dependency on glycolysis. J Biol Chem. 2011;286:32843–32853. doi: 10.1074/jbc.M111.260935. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Vlashi E., Lagadec C., Vergnes L., Matsutani T., Masui K., Poulou M., et al. Metabolic state of glioma stem cells and nontumorigenic cells. Proc Natl Acad Sci U S A. 2011;108:16062–16067. doi: 10.1073/pnas.1106704108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Nakano I. Therapeutic potential of targeting glucose metabolism in glioma stem cells. Expert Opin Ther Targets. 2014;18:1233–1236. doi: 10.1517/14728222.2014.944899. [DOI] [PubMed] [Google Scholar]
- 44.Christofk H.R., Vander Heiden M.G., Harris M.H., Ramanathan A., Gerszten R.E., Wei R., et al. The M2 splice isoform of pyruvate kinase is important for cancer metab olism and tumour growth. Nature. 2008;452:230–233. doi: 10.1038/nature06734. [DOI] [PubMed] [Google Scholar]
- 45.Smolkova K., Plecita-Hlavata L., Bellance N., Benard G., Rossignol R., Jezek P. Waves of gene regulation suppress and then restore oxidative phosphorylation in cancer cells. Int J Biochem Cell Biol. 2011;43:950–968. doi: 10.1016/j.biocel.2010.05.003. [DOI] [PubMed] [Google Scholar]
- 46.Anderson A.S., Roberts P.C., Frisard M.I., McMillan R.P., Brown T.J., Lawless M.H., et al. Metabolic changes during ovarian cancer progression as targets for sphingosine treatment. Exp Cell Res. 2013;319:1431–1442. doi: 10.1016/j.yexcr.2013.02.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Anderson A.S., Roberts P.C., Frisard M.I., Hulver M.W., Schmelz E.M. Ovarian tumor-initiating cells display a flexible metabolism. Exp Cell Res. 2014;328:44–57. doi: 10.1016/j.yexcr.2014.08.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Elia I., Schmieder R., Christen S., Fendt S.M. Organ-specific cancer metabolism and its potential for therapy. Handb Exp Pharmacol. 2016;233:321–353. doi: 10.1007/164_2015_10. [DOI] [PubMed] [Google Scholar]
- 49.Gong Y., Ji P., Yang Y.S., Xie S., Yu T.J., Xiao Y., et al. Metabolic-pathway-based subtyping of triple-negative breast cancer reveals potential therapeutic targets. Cell Metabol. 2021;33:51–64.e9. doi: 10.1016/j.cmet.2020.10.012. [DOI] [PubMed] [Google Scholar]
- 50.Edwards D.N., Ngwa V.M., Raybuck A.L., Wang S., Hwang Y., Kim L.C., et al. Selective glutamine metabolism inhibition in tumor cells improves antitumor T lymphocyte activity in triple-negative breast cancer. J Clin Invest. 2021;131 doi: 10.1172/JCI140100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Han L., Dong L., Leung K., Zhao Z., Li Y., Gao L., et al. METTL16 drives leukemogenesis and leukemia stem cell self-renewal by reprogramming BCAA metabolism. Cell Stem Cell. 2023;30:52–68 e13. doi: 10.1016/j.stem.2022.12.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Tao L., Moreno-Smith M., Ibarra-Garcia-Padilla R., Milazzo G., Drolet N.A., Hernandez B.E., et al. CHAF1A blocks neuronal differentiation and promotes neuroblastoma oncogenesis via metabolic reprogramming. Adv Sci. 2021;8 doi: 10.1002/advs.202005047. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Sun L., Suo C., Li S.T., Zhang H., Gao P. Metabolic reprogramming for cancer cells and their microenvironment: beyond the Warburg effect. Biochim Biophys Acta Rev Cancer. 2018;1870:51–66. doi: 10.1016/j.bbcan.2018.06.005. [DOI] [PubMed] [Google Scholar]
- 54.de Sousa e Melo F., Kurtova A.V., Harnoss J.M., Kljavin N., Hoeck J.D., Hung J., et al. A distinct role for Lgr5+ stem cells in primary and metastatic colon cancer. Nature. 2017;543:676–680. doi: 10.1038/nature21713. [DOI] [PubMed] [Google Scholar]
- 55.Maccalli C., Rasul K.I., Elawad M., Ferrone S. The role of cancer stem cells in the modulation of anti-tumor immune responses. Semin Cancer Biol. 2018;53:189–200. doi: 10.1016/j.semcancer.2018.09.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Yang J., Ren B., Yang G., Wang H., Chen G., You L., et al. The enhancement of glycolysis regulates pancreatic cancer metastasis. Cell Mol Life Sci. 2020;77:305–321. doi: 10.1007/s00018-019-03278-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.LeBleu V.S., O'Connell J.T., Gonzalez Herrera K.N., Wikman H., Pantel K., Haigis M.C., et al. PGC-1alpha mediates mitochondrial biogenesis and oxidative phosphorylation in cancer cells to promote metastasis. Nat Cell Biol. 2014;16:992–1003. doi: 10.1038/ncb3039. 1-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Xia L., Oyang L., Lin J., Tan S., Han Y., Wu N., et al. The cancer metabolic reprogramming and immune response. Mol Cancer. 2021;20:28. doi: 10.1186/s12943-021-01316-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Chen D., Zhang X., Li Z., Zhu B. Metabolic regulatory crosstalk between tumor microenvironment and tumor-associated macrophages. Theranostics. 2021;11:1016–1030. doi: 10.7150/thno.51777. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Wynn T.A., Chawla A., Pollard J.W. Macrophage biology in development, homeostasis and disease. Nature. 2013;496:445–455. doi: 10.1038/nature12034. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Lu Z., Li X., Yang P., Mu G., He L., Song C., et al. Heparin-binding protein enhances NF-kappaB pathway-mediated inflammatory gene transcription in M1 macrophages via lactate. Inflammation. 2021;44:48–56. doi: 10.1007/s10753-020-01263-4. [DOI] [PubMed] [Google Scholar]
- 62.Li J., Li M., Ye K., Jiang Q., Wang M., Wen X., et al. Chemical profile of Xian-He-Cao-Chang-Yan formula and its effects on ulcerative colitis. J Ethnopharmacol. 2021;267 doi: 10.1016/j.jep.2020.113517. [DOI] [PubMed] [Google Scholar]
- 63.He Q., Yin J., Zou B., Guo H. WIN55212-2 alleviates acute lung injury by inhibiting macrophage glycolysis through the miR-29b-3p/FOXO3/PFKFB3 axis. Mol Immunol. 2022;149:119–128. doi: 10.1016/j.molimm.2022.06.005. [DOI] [PubMed] [Google Scholar]
- 64.Brown M., Roulson J.A., Hart C.A., Tawadros T., Clarke N.W. Arachidonic acid induction of Rho-mediated transendothelial migration in prostate cancer. Br J Cancer. 2014;110:2099–2108. doi: 10.1038/bjc.2014.99. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Apostolova P., Pearce E.L. Lactic acid and lactate: revisiting the physiological roles in the tumor microenvironment. Trends Immunol. 2022;43:969–977. doi: 10.1016/j.it.2022.10.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Zhou H.C., Xin-Yan Y., Yu W.W., Liang X.Q., Du X.Y., Liu Z.C., et al. Lactic acid in macrophage polarization: the significant role in inflammation and cancer. Int Rev Immunol. 2022;41:4–18. doi: 10.1080/08830185.2021.1955876. [DOI] [PubMed] [Google Scholar]
- 67.Fang X., Zhao P., Gao S., Liu D., Zhang S., Shan M., et al. Lactate induces tumor-associated macrophage polarization independent of mitochondrial pyruvate carrier-mediated metabolism. Int J Biol Macromol. 2023;237 doi: 10.1016/j.ijbiomac.2023.123810. [DOI] [PubMed] [Google Scholar]
- 68.Chen M., Liu H., Li Z., Ming A.L., Chen H. Mechanism of PKM2 affecting cancer immunity and metabolism in tumor microenvironment. J Cancer. 2021;12:3566–3574. doi: 10.7150/jca.54430. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Sun X., Yao L., Liang H., Wang D., He Y., Wei Y., et al. Intestinal epithelial PKM2 serves as a safeguard against experimental colitis via activating beta-catenin signaling. Mucosal Immunol. 2019;12:1280–1290. doi: 10.1038/s41385-019-0197-6. [DOI] [PubMed] [Google Scholar]
- 70.Ouyang X., Han Y., Qu G., Li M., Wu N., Liu H., et al. Metabolic regulation of T cell development by Sin1-mTORC2 is mediated by pyruvate kinase M2. J Mol Cell Biol. 2019;11:93–106. doi: 10.1093/jmcb/mjy065. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Angiari S., Runtsch M.C., Sutton C.E., Palsson-McDermott E.M., Kelly B., Rana N., et al. Pharmacological activation of pyruvate kinase M2 inhibits CD4+ T cell pathogenicity and suppresses autoimmunity. Cell Metabol. 2020;31:391–405 e8. doi: 10.1016/j.cmet.2019.10.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Dallos M.C., Drake C.G. Blocking PD-1/PD-L1 in genitourinary malignancies: to immunity and beyond. Cancer J. 2018;24:20–30. doi: 10.1097/PPO.0000000000000302. [DOI] [PubMed] [Google Scholar]
- 73.Turajlic S., Litchfield K., Xu H., Rosenthal R., McGranahan N., Reading J.L., et al. Insertion-and-deletion-derived tumour-specific neoantigens and the immunogenic phenotype: a pan-cancer analysis. Lancet Oncol. 2017;18:1009–1021. doi: 10.1016/S1470-2045(17)30516-8. [DOI] [PubMed] [Google Scholar]
- 74.Cao Y. Adipocyte and lipid metabolism in cancer drug resistance. J Clin Invest. 2019;129:3006–3017. doi: 10.1172/JCI127201. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Sun L., Lu T., Tian K., Zhou D., Yuan J., Wang X., et al. Alpha-Enolase promotes gastric cancer cell proliferation and metastasis via regulating AKT signaling pathway. Eur J Pharmacol. 2019;845:8–15. doi: 10.1016/j.ejphar.2018.12.035. [DOI] [PubMed] [Google Scholar]
- 76.Tavares-Valente D., Baltazar F., Moreira R., Queiros O. Cancer cell bioenergetics and pH regulation influence breast cancer cell resistance to paclitaxel and doxorubicin. J Bioenerg Biomembr. 2013;45:467–475. doi: 10.1007/s10863-013-9519-7. [DOI] [PubMed] [Google Scholar]
- 77.Zhou M., Zhao Y., Ding Y., Liu H., Liu Z., Fodstad O., et al. Warburg effect in chemosensitivity: targeting lactate dehydrogenase-A re-sensitizes taxol-resistant cancer cells to taxol. Mol Cancer. 2010;9:33. doi: 10.1186/1476-4598-9-33. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Cuzick J., Sestak I., Baum M., Buzdar A., Howell A., Dowsett M., et al. Effect of anastrozole and tamoxifen as adjuvant treatment for early-stage breast cancer: 10-year analysis of the ATAC trial. Lancet Oncol. 2010;11:1135–1141. doi: 10.1016/S1470-2045(10)70257-6. [DOI] [PubMed] [Google Scholar]
- 79.Martinez-Outschoorn U.E., Goldberg A., Lin Z., Ko Y.H., Flomenberg N., Wang C., et al. Anti-estrogen resistance in breast cancer is induced by the tumor microenvironment and can be overcome by inhibiting mitochondrial function in epithelial cancer cells. Cancer Biol Ther. 2011;12:924–938. doi: 10.4161/cbt.12.10.17780. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Zhao Y., Liu H., Liu Z., Ding Y., Ledoux S.P., Wilson G.L., et al. Overcoming trastuzumab resistance in breast cancer by targeting dysregulated glucose metabolism. Cancer Res. 2011;71:4585–4597. doi: 10.1158/0008-5472.CAN-11-0127. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Al Tameemi W., Dale T.P., Al-Jumaily R.M.K., Forsyth N.R. Hypoxia-modified cancer cell metabolism. Front Cell Dev Biol. 2019;7:4. doi: 10.3389/fcell.2019.00004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Chan D.A., Sutphin P.D., Nguyen P., Turcotte S., Lai E.W., Banh A., et al. Targeting GLUT1 and the Warburg effect in renal cell carcinoma by chemical synthetic lethality. Sci Transl Med. 2011;3 doi: 10.1126/scitranslmed.3002394. 94ra70. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Pao S.S., Paulsen I.T., Saier M.H., Jr. Major facilitator superfamily. Microbiol Mol Biol Rev. 1998;62:1–34. doi: 10.1128/mmbr.62.1.1-34.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Tilekar K., Upadhyay N., Iancu C.V., Pokrovsky V., Choe J.Y., Ramaa C.S. Power of two: combination of therapeutic approaches involving glucose transporter (GLUT) inhibitors to combat cancer. Biochim Biophys Acta Rev Cancer. 2020;1874 doi: 10.1016/j.bbcan.2020.188457. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Thorens B., Mueckler M. Glucose transporters in the 21st century. Am J Physiol Endocrinol Metab. 2010;298:E141–E145. doi: 10.1152/ajpendo.00712.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Thorens B. GLUT2, glucose sensing and glucose homeostasis. Diabetologia. 2015;58:221–232. doi: 10.1007/s00125-014-3451-1. [DOI] [PubMed] [Google Scholar]
- 87.Leturque A., Brot-Laroche E., Le Gall M., Stolarczyk E., Tobin V. The role of GLUT2 in dietary sugar handling. J Physiol Biochem. 2005;61:529–537. doi: 10.1007/BF03168378. [DOI] [PubMed] [Google Scholar]
- 88.Hamann I., Krys D., Glubrecht D., Bouvet V., Marshall A., Vos L., et al. Expression and function of hexose transporters GLUT1, GLUT2, and GLUT5 in breast cancer-effects of hypoxia. FASEB J. 2018;32:5104–5118. doi: 10.1096/fj.201800360R. [DOI] [PubMed] [Google Scholar]
- 89.Wang F., Qi X.M., Wertz R., Mortensen M., Hagen C., Evans J., et al. p38gamma MAPK is essential for aerobic glycolysis and pancreatic tumorigenesis. Cancer Res. 2020;80:3251–3264. doi: 10.1158/0008-5472.CAN-19-3281. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Simpson I.A., Dwyer D., Malide D., Moley K.H., Travis A., Vannucci S.J. The facilitative glucose transporter GLUT3: 20 years of distinction. Am J Physiol Endocrinol Metab. 2008;295:E242–E253. doi: 10.1152/ajpendo.90388.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Libby C.J., Gc S., Benavides G.A., Fisher J.L., Williford S.E., Zhang S., et al. A role for GLUT3 in glioblastoma cell invasion that is not recapitulated by GLUT1. Cell Adhes Migrat. 2021;15:101–115. doi: 10.1080/19336918.2021.1903684. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.Tsai T.H., Yang C.C., Kou T.C., Yang C.E., Dai J.Z., Chen C.L., et al. Overexpression of GLUT3 promotes metastasis of triple-negative breast cancer by modulating the inflammatory tumor microenvironment. J Cell Physiol. 2021;236:4669–4680. doi: 10.1002/jcp.30189. [DOI] [PubMed] [Google Scholar]
- 93.Zhong X., Tian S., Zhang X., Diao X., Dong F., Yang J., et al. CUE domain-containing protein 2 promotes the Warburg effect and tumorigenesis. EMBO Rep. 2017;18:809–825. doi: 10.15252/embr.201643617. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.Govers R. Molecular mechanisms of GLUT4 regulation in adipocytes. Diabetes Metab. 2014;40:400–410. doi: 10.1016/j.diabet.2014.01.005. [DOI] [PubMed] [Google Scholar]
- 95.Gould G., Brodsky F., Bryant N. Building GLUT4 vesicles: CHC22 Clathrin's human touch. Trends Cell Biol. 2020;30:705–719. doi: 10.1016/j.tcb.2020.05.007. [DOI] [PubMed] [Google Scholar]
- 96.Liu H., Lyu H., Jiang G., Chen D., Ruan S., Liu S., et al. ALKBH5-mediated m6A demethylation of GLUT4 mRNA promotes glycolysis and resistance to HER2-targeted therapy in breast cancer. Cancer Res. 2022;82:3974–3986. doi: 10.1158/0008-5472.CAN-22-0800. [DOI] [PubMed] [Google Scholar]
- 97.Wang J., Hu J.L., Cao R.B., Ding Q., Peng G., Fei S.J., et al. Small hairpin RNA-mediated Krüppel-like factor 8 gene knockdown inhibits invasion of nasopharyngeal carcinoma. Oncol Lett. 2015;9:2515–2519. doi: 10.3892/ol.2015.3099. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.Mao A., Zhou X., Liu Y., Ding J., Miao A., Pan G. KLF8 is associated with poor prognosis and regulates glycolysis by targeting GLUT4 in gastric cancer. J Cell Mol Med. 2019;23:5087–5097. doi: 10.1111/jcmm.14378. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Middleton R.J. Hexokinases and glucokinases. Biochem Soc Trans. 1990;18:180–183. doi: 10.1042/bst0180180. [DOI] [PubMed] [Google Scholar]
- 100.Ureta T. The comparative isozymology of vertebrate hexokinases. Comp Biochem Physiol B. 1982;71:549–555. doi: 10.1016/0305-0491(82)90461-8. [DOI] [PubMed] [Google Scholar]
- 101.Wilson J.E. Isozymes of mammalian hexokinase: structure, subcellular localization and metabolic function. J Exp Biol. 2003;206:2049–2057. doi: 10.1242/jeb.00241. [DOI] [PubMed] [Google Scholar]
- 102.Kimmelman A. Metabolic dependencies in RAS-driven cancers. Clin Cancer Res. 2015;21:1828–1834. doi: 10.1158/1078-0432.CCR-14-2425. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.Tsai F.D., Lopes M.S., Zhou M., Court H., Ponce O., Fiordalisi J.J., et al. K-Ras4A splice variant is widely expressed in cancer and uses a hybrid membrane-targeting motif. Proc Natl Acad Sci U S A. 2015;112:779–784. doi: 10.1073/pnas.1412811112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104.Amendola C.R., Mahaffey J.P., Parker S.J., Ahearn I.M., Chen W.C., Zhou M., et al. KRAS4A directly regulates hexokinase 1. Nature. 2019;576:482–486. doi: 10.1038/s41586-019-1832-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105.Hay N. Reprogramming glucose metabolism in cancer: can it be exploited for cancer therapy?. Nat Rev Cancer. 2016;16:635–649. doi: 10.1038/nrc.2016.77. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106.Shoshan-Barmatz V., Mizrachi D. VDAC1: from structure to cancer therapy. Front Oncol. 2012;2:164. doi: 10.3389/fonc.2012.00164. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107.Azoulay-Zohar H., Israelson A., Abu-Hamad S., Shoshan-Barmatz V. In self-defence: hexokinase promotes voltage-dependent anion channel closure and prevents mitochondria-mediated apoptotic cell death. Biochem J. 2004;377:347–355. doi: 10.1042/BJ20031465. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.Shimizu S., Konishi A., Kodama T., Tsujimoto Y. BH4 domain of antiapoptotic Bcl-2 family members closes voltage-dependent anion channel and inhibits apoptotic mitochondrial changes and cell death. Proc Natl Acad Sci U S A. 2000;97:3100–3105. doi: 10.1073/pnas.97.7.3100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109.Shangguan X., He J., Ma Z., Zhang W., Ji Y., Shen K., et al. SUMOylation controls the binding of hexokinase 2 to mitochondria and protects against prostate cancer tumorigenesis. Nat Commun. 2021;12:1812. doi: 10.1038/s41467-021-22163-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110.Li H., Song J., He Y., Liu Y., Liu Z., Sun W., et al. CRISPR/Cas9 screens reveal that hexokinase 2 enhances cancer stemness and tumorigenicity by activating the ACSL4-fatty acid β-oxidation pathway. Adv Sci. 2022;9 doi: 10.1002/advs.202105126. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Wang Q., Guo X., Li L., Gao Z., Su X., Ji M., et al. N-Methyladenosine METTL3 promotes cervical cancer tumorigenesis and Warburg effect through YTHDF1/HK2 modification. Cell Death Dis. 2020;11:911. doi: 10.1038/s41419-020-03071-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112.Wolf A., Agnihotri S., Micallef J., Mukherjee J., Sabha N., Cairns R., 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. doi: 10.1084/jem.20101470. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113.Rempel A., Mathupala S.P., Griffin C.A., Hawkins A.L., Pedersen P.L. Glucose catabolism in cancer cells: amplification of the gene encoding type II hexokinase. Cancer Res. 1996;56:2468–2471. [PubMed] [Google Scholar]
- 114.Pedersen P.L., Mathupala S., Rempel A., Geschwind J.F., Ko Y.H. Mitochondrial bound type II hexokinase: a key player in the growth and survival of many cancers and an ideal prospect for therapeutic intervention. Biochim Biophys Acta. 2002;1555:14–20. doi: 10.1016/s0005-2728(02)00248-7. [DOI] [PubMed] [Google Scholar]
- 115.Suh D.H., Kim M.A., Kim H., Kim M.K., Kim H.S., Chung H.H., et al. Association of overexpression of hexokinase II with chemoresistance in epithelial ovarian cancer. Clin Exp Med. 2014;14:345–353. doi: 10.1007/s10238-013-0250-9. [DOI] [PubMed] [Google Scholar]
- 116.Goel A., Mathupala S., Pedersen P. Glucose metabolism in cancer. Evidence that demethylation events play a role in activating type II hexokinase gene expression. J Biol Chem. 2003;278:15333–15340. doi: 10.1074/jbc.M300608200. [DOI] [PubMed] [Google Scholar]
- 117.Printz R.L., Koch S., Potter L.R., O'Doherty R.M., Tiesinga J.J., Moritz S., et al. Hexokinase II mRNA and gene structure, regulation by insulin, and evolution. J Biol Chem. 1993;268:5209–5219. [PubMed] [Google Scholar]
- 118.Mathupala S., Ko Y., Pedersen P. Hexokinase II: cancer's double-edged sword acting as both facilitator and gatekeeper of malignancy when bound to mitochondria. Oncogene. 2006;25:4777–4786. doi: 10.1038/sj.onc.1209603. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119.Pudova E.A., Kudryavtseva A.V., Fedorova M.S., Zaretsky A.R., Shcherbo D.S., Lukyanova E.N., et al. HK3 overexpression associated with epithelial–mesenchymal transition in colorectal cancer. BMC Genom. 2018;19:113. doi: 10.1186/s12864-018-4477-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 120.Wu X., Qian S., Zhang J., Feng J., Luo K., Sun L., et al. Lipopolysaccharide promotes metastasis via acceleration of glycolysis by the nuclear factor-κB/snail/hexokinase3 signaling axis in colorectal cancer. Cancer Metabol. 2021;9:23. doi: 10.1186/s40170-021-00260-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121.Wyatt E., Wu R., Rabeh W., Park H.W., Ghanefar M., Ardehali H. Regulation and cytoprotective role of hexokinase III. PLoS One. 2010;5 doi: 10.1371/journal.pone.0013823. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 122.Watanabe H., Takehana K., Date M., Shinozaki T., Raz A. Tumor cell autocrine motility factor is the neuroleukin/phosphohexose isomerase polypeptide. Cancer Res. 1996;56:2960–2963. [PubMed] [Google Scholar]
- 123.Liotta L., Mandler R., Murano G., Katz D., Gordon R., Chiang P., et al. Tumor cell autocrine motility factor. Proc Natl Acad Sci U S A. 1986;83:3302–3306. doi: 10.1073/pnas.83.10.3302. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124.Jiang W.G., Raz A., Douglas-Jones A., Mansel R.E. Expression of autocrine motility factor (AMF) and its receptor, AMFR, in human breast cancer. J Histochem Cytochem. 2006;54:231–241. doi: 10.1369/jhc.5A6785.2005. [DOI] [PubMed] [Google Scholar]
- 125.Gong W., Jiang Y., Wang L., Wei D., Yao J., Huang S., et al. Expression of autocrine motility factor correlates with the angiogenic phenotype of and poor prognosis for human gastric cancer. Clin Cancer Res. 2005;11:5778–5783. doi: 10.1158/1078-0432.CCR-05-0214. [DOI] [PubMed] [Google Scholar]
- 126.Lucarelli G., Rutigliano M., Sanguedolce F., Galleggiante V., Giglio A., Cagiano S., et al. Increased expression of the autocrine motility factor is associated with poor prognosis in patients with clear cell-renal cell carcinoma. Medicine. 2015;94:e2117. doi: 10.1097/MD.0000000000002117. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 127.Fang E., Wang J., Hong M., Zheng L., Tong Q. Valproic acid suppresses Warburg effect and tumor progression in neuroblastoma. Biochem Biophys Res Commun. 2019;508:9–16. doi: 10.1016/j.bbrc.2018.11.103. [DOI] [PubMed] [Google Scholar]
- 128.Al Hasawi N., Alkandari M.F., Luqmani Y.A. Phosphofructokinase: a mediator of glycolytic flux in cancer progression. Crit Rev Oncol Hematol. 2014;92:312–321. doi: 10.1016/j.critrevonc.2014.05.007. [DOI] [PubMed] [Google Scholar]
- 129.Wang T.A., Zhang X.D., Guo X.Y., Xian S.L., Lu Y.F. 3-Bromopyruvate and sodium citrate target glycolysis, suppress survivin, and induce mitochondrial-mediated apoptosis in gastric cancer cells and inhibit gastric orthotopic transplantation tumor growth. Oncol Rep. 2016;35:1287–1296. doi: 10.3892/or.2015.4511. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 130.Li L., Li L., Li W., Chen T., Bin Z., Zhao L., et al. TAp73-induced phosphofructokinase-1 transcription promotes the Warburg effect and enhances cell proliferation. Nat Commun. 2018;9:4683. doi: 10.1038/s41467-018-07127-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131.Richardson D.A., Sritangos P., James A.D., Sultan A., Bruce J.I.E. Metabolic regulation of calcium pumps in pancreatic cancer: role of phosphofructokinase-fructose-bisphosphatase-3 (PFKFB3) Cancer Metabol. 2020;8:2. doi: 10.1186/s40170-020-0210-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 132.Dasgupta S., Rajapakshe K., Zhu B., Nikolai B.C., Yi P., Putluri N., et al. Metabolic enzyme PFKFB4 activates transcriptional coactivator SRC-3 to drive breast cancer. Nature. 2018;556:249–254. doi: 10.1038/s41586-018-0018-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 133.Lessa R.C., Campos A.H., Freitas C.E., Silva F.R., Kowalski L.P., Carvalho A.L., et al. Identification of upregulated genes in oral squamous cell carcinomas. Head Neck. 2013;35:1475–1481. doi: 10.1002/hed.23169. [DOI] [PubMed] [Google Scholar]
- 134.Peng Y., Li X., Wu M., Yang J., Liu M., Zhang W., et al. New prognosis biomarkers identified by dynamic proteomic analysis of colorectal cancer. Mol Biosyst. 2012;8:3077–3088. doi: 10.1039/c2mb25286d. [DOI] [PubMed] [Google Scholar]
- 135.Shen Y., Xu J., Pan X., Zhang Y., Weng Y., Zhou D., et al. LncRNA KCNQ1OT1 sponges miR-34c-5p to promote osteosarcoma growth via ALDOA enhanced aerobic glycolysis. Cell Death Dis. 2020;11:278. doi: 10.1038/s41419-020-2485-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 136.Altenberg B., Greulich K.O. Genes of glycolysis are ubiquitously overexpressed in 24 cancer classes. Genomics. 2004;84:1014–1020. doi: 10.1016/j.ygeno.2004.08.010. [DOI] [PubMed] [Google Scholar]
- 137.Pekel G., Ari F. Therapeutic targeting of cancer metabolism with triosephosphate isomerase. Chem Biodivers. 2020;17 doi: 10.1002/cbdv.202000012. [DOI] [PubMed] [Google Scholar]
- 138.Liberti M.V., Dai Z., Wardell S.E., Baccile J.A., Liu X., Gao X., et al. A predictive model for selective targeting of the Warburg effect through GAPDH inhibition with a natural product. Cell Metabol. 2017;26 doi: 10.1016/j.cmet.2017.08.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 139.Nie H., Ju H., Fan J., Shi X., Cheng Y., Cang X., et al. O-GlcNAcylation of PGK1 coordinates glycolysis and TCA cycle to promote tumor growth. Nat Commun. 2020;11:36. doi: 10.1038/s41467-019-13601-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 140.Ren F., Wu H., Lei Y., Zhang H., Liu R., Zhao Y., et al. Quantitative proteomics identification of phosphoglycerate mutase 1 as a novel therapeutic target in hepatocellular carcinoma. Mol Cancer. 2010;9:81. doi: 10.1186/1476-4598-9-81. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 141.Li C., Shu F., Lei B., Lv D., Zhang S., Mao X. Expression of PGAM1 in renal clear cell carcinoma and its clinical significance. Int J Clin Exp Pathol. 2015;8:9410–9415. [PMC free article] [PubMed] [Google Scholar]
- 142.Peng X., Gong F., Chen Y., Qiu M., Cheng K., Tang J., et al. Proteomics identification of PGAM1 as a potential therapeutic target for urothelial bladder cancer. J Proteonomics. 2016;132:85–92. doi: 10.1016/j.jprot.2015.11.027. [DOI] [PubMed] [Google Scholar]
- 143.Sun Q., Li S., Wang Y., Peng H., Zhang X., Zheng Y., et al. Phosphoglyceric acid mutase-1 contributes to oncogenic mTOR-mediated tumor growth and confers non-small cell lung cancer patients with poor prognosis. Cell Death Differ. 2018;25:1160–1173. doi: 10.1038/s41418-017-0034-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 144.Majumder P., Febbo P., Bikoff R., Berger R., Xue Q., McMahon L., et al. mTOR inhibition reverses Akt-dependent prostate intraepithelial neoplasia through regulation of apoptotic and HIF-1-dependent pathways. Nat Med. 2004;10:594–601. doi: 10.1038/nm1052. [DOI] [PubMed] [Google Scholar]
- 145.Mazurek S., Boschek C.B., Hugo F., Eigenbrodt E. Pyruvate kinase type M2 and its role in tumor growth and spreading. Semin Cancer Biol. 2005;15:300–308. doi: 10.1016/j.semcancer.2005.04.009. [DOI] [PubMed] [Google Scholar]
- 146.Dombrauckas J.D., Santarsiero B.D., Mesecar A.D. Structural basis for tumor pyruvate kinase M2 allosteric regulation and catalysis. Biochemistry. 2005;44:9417–9429. doi: 10.1021/bi0474923. [DOI] [PubMed] [Google Scholar]
- 147.Christofk H.R., Vander Heiden M.G., Harris M.H., Ramanathan A., Gerszten R.E., Wei R., et al. The M2 splice isoform of pyruvate kinase is important for cancer metabolism and tumour growth. Nature. 2008;452:230–233. doi: 10.1038/nature06734. [DOI] [PubMed] [Google Scholar]
- 148.Yang W., Zheng Y., Xia Y., Ji H., Chen X., Guo F., et al. ERK1/2-dependent phosphorylation and nuclear translocation of PKM2 promotes the Warburg effect. Nat Cell Biol. 2012;14:1295–1304. doi: 10.1038/ncb2629. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 149.Pavlova N.N., Thompson C.B. The emerging hallmarks of cancer metabolism. Cell Metabol. 2016;23:27–47. doi: 10.1016/j.cmet.2015.12.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 150.Doherty J.R., Cleveland J.L. Targeting lactate metabolism for cancer therapeutics. J Clin Invest. 2013;123:3685–3692. doi: 10.1172/JCI69741. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 151.Dang C.V., Semenza G.L. Oncogenic alterations of metabolism. Trends Biochem Sci. 1999;24:68–72. doi: 10.1016/s0968-0004(98)01344-9. [DOI] [PubMed] [Google Scholar]
- 152.Read J.A., Winter V.J., Eszes C.M., Sessions R.B., Brady R.L. Structural basis for altered activity of M- and H-isozyme forms of human lactate dehydrogenase. Proteins. 2001;43:175–185. doi: 10.1002/1097-0134(20010501)43:2<175::aid-prot1029>3.0.co;2-#. [DOI] [PubMed] [Google Scholar]
- 153.Yan M., Wang C., He B., Yang M., Tong M., Long Z., et al. Aurora-A kinase: a potent oncogene and target for cancer therapy. Med Res Rev. 2016;36:1036–1079. doi: 10.1002/med.21399. [DOI] [PubMed] [Google Scholar]
- 154.Hannak E., Kirkham M., Hyman A., Oegema K. Aurora-A kinase is required for centrosome maturation in Caenorhabditis elegans. J Cell Biol. 2001;155:1109–1116. doi: 10.1083/jcb.200108051. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 155.Cheng A., Zhang P., Wang B., Yang D., Duan X., Jiang Y., et al. Aurora-A mediated phosphorylation of LDHB promotes glycolysis and tumor progression by relieving the substrate-inhibition effect. Nat Commun. 2019;10:5566. doi: 10.1038/s41467-019-13485-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 156.Hakimi A.A., Reznik E., Lee C.H., Creighton C.J., Brannon A.R., Luna A., et al. An integrated metabolic atlas of clear cell renal cell carcinoma. Cancer Cell. 2016;29:104–116. doi: 10.1016/j.ccell.2015.12.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 157.Courtney K.D., Bezwada D., Mashimo T., Pichumani K., Vemireddy V., Funk A.M., et al. Isotope tracing of human clear cell renal cell carcinomas demonstrates suppressed glucose oxidation in vivo. Cell Metabol. 2018;28:793–800 e2. doi: 10.1016/j.cmet.2018.07.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 158.Cardaci S., Zheng L., MacKay G., van den Broek N.J., MacKenzie E.D., Nixon C., et al. Pyruvate carboxylation enables growth of SDH-deficient cells by supporting aspartate biosynthesis. Nat Cell Biol. 2015;17:1317–1326. doi: 10.1038/ncb3233. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 159.Kim J., Li C., Weiss H., Zhou Y., Liu C., Wang Q., et al. Regulation of ketogenic enzyme HMGCS2 by Wnt/β-catenin/PPARγ pathway in intestinal cells. Cells. 2019;8:1106. doi: 10.3390/cells8091106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 160.Grabacka M., Pierzchalska M., Dean M., Reiss K. Regulation of ketone body metabolism and the role of PPARα. Int J Mol Sci. 2016;17:2093. doi: 10.3390/ijms17122093. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 161.Grabacka M.M., Wilk A., Antonczyk A., Banks P., Walczyk-Tytko E., Dean M., et al. Fenofibrate induces ketone body production in melanoma and glioblastoma cells. Front Endocrinol. 2016;7:5. doi: 10.3389/fendo.2016.00005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 162.Puchalska P., Crawford P.A. Multi-dimensional roles of ketone bodies in fuel metabolism, signaling, and therapeutics. Cell Metabol. 2017;25:262–284. doi: 10.1016/j.cmet.2016.12.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 163.Li Y., Xu J., Chen L., Zhong W.D., Zhang Z., Mi L., et al. HAb18G (CD147), a cancer-associated biomarker and its role in cancer detection. Histopathology. 2009;54:677–687. doi: 10.1111/j.1365-2559.2009.03280.x. [DOI] [PubMed] [Google Scholar]
- 164.Ke X., Fei F., Chen Y., Xu L., Zhang Z., Huang Q., et al. Hypoxia upregulates CD147 through a combined effect of HIF-1α and Sp1 to promote glycolysis and tumor progression in epithelial solid tumors. Carcinogenesis. 2012;33:1598–1607. doi: 10.1093/carcin/bgs196. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 165.Le Floch R., Chiche J., Marchiq I., Naiken T., Ilc K., Murray C., et al. CD147 subunit of lactate/H+ symporters MCT1 and hypoxia-inducible MCT4 is critical for energetics and growth of glycolytic tumors. Proc Natl Acad Sci U S A. 2011;108:16663–16668. doi: 10.1073/pnas.1106123108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 166.Huang Q., Li J., Xing J., Li W., Li H., Ke X., et al. CD147 promotes reprogramming of glucose metabolism and cell proliferation in HCC cells by inhibiting the p53-dependent signaling pathway. J Hepatol. 2014;61:859–866. doi: 10.1016/j.jhep.2014.04.035. [DOI] [PubMed] [Google Scholar]
- 167.Ngo H., Tortorella S.M., Ververis K., Karagiannis T.C. The Warburg effect: molecular aspects and therapeutic possibilities. Mol Biol Rep. 2015;42:825–834. doi: 10.1007/s11033-014-3764-7. [DOI] [PubMed] [Google Scholar]
- 168.Yeung S., Pan J., Lee M. Roles of p53, MYC and HIF-1 in regulating glycolysis—the seventh hallmark of cancer. Cell Mol Life Sci. 2008;65:3981–3999. doi: 10.1007/s00018-008-8224-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 169.Kumar J. The sine oculis homeobox (SIX) family of transcription factors as regulators of development and disease. Cell Mol. 2009;66:565–583. doi: 10.1007/s00018-008-8335-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 170.Wu W., Ren Z., Li P., Yu D., Chen J., Huang R., et al. Six1: a critical transcription factor in tumorigenesis. Int J Cancer. 2015;136:1245–1253. doi: 10.1002/ijc.28755. [DOI] [PubMed] [Google Scholar]
- 171.Coletta R., Christensen K., Micalizzi D., Jedlicka P., Varella-Garcia M., Ford H. Six1 overexpression in mammary cells induces genomic instability and is sufficient for malignant transformation. Cancer Res. 2008;68:2204–2213. doi: 10.1158/0008-5472.CAN-07-3141. [DOI] [PubMed] [Google Scholar]
- 172.Ng K.T., Man K., Sun C.K., Lee T.K., Poon R.T., Lo C.M., et al. Clinicopathological significance of homeoprotein Six1 in hepatocellular carcinoma. Br J Cancer. 2006;95:1050–1055. doi: 10.1038/sj.bjc.6603399. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 173.Li L., Liang Y., Kang L., Liu Y., Gao S., Chen S., et al. Transcriptional regulation of the Warburg effect in cancer by SIX1. Cancer Cell. 2018;33 doi: 10.1016/j.ccell.2018.01.010. [DOI] [PubMed] [Google Scholar]
- 174.Shi Y., Seto E., Chang L.S., Shenk T. Transcriptional repression by YY1, a human GLI-Kruppel-related protein, and relief of repression by adenovirus E1A protein. Cell. 1991;67:377–388. doi: 10.1016/0092-8674(91)90189-6. [DOI] [PubMed] [Google Scholar]
- 175.Seto E., Shi Y., Shenk T. YY1 is an initiator sequence-binding protein that directs and activates transcription in vitro. Nature. 1991;354:241–245. doi: 10.1038/354241a0. [DOI] [PubMed] [Google Scholar]
- 176.Sarvagalla S., Kolapalli S.P., Vallabhapurapu S. The two sides of YY1 in cancer: a friend and a foe. Front Oncol. 2019;9:1230. doi: 10.3389/fonc.2019.01230. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 177.Xu C., Tsai Y.H., Galbo P.M., Gong W., Storey A.J., Xu Y., et al. Cistrome analysis of YY1 uncovers a regulatory axis of YY1:BRD2/4-PFKP during tumorigenesis of advanced prostate cancer. Nucleic Acids Res. 2021;49:4971–4988. doi: 10.1093/nar/gkab252. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 178.Hou Y., Zhang Q., Pang W., Hou L., Liang Y., Han X., et al. YTHDC1-mediated augmentation of miR-30d in repressing pancreatic tumorigenesis via attenuation of RUNX1-induced transcriptional activation of Warburg effect. Cell Death Differ. 2021;28:3105–3124. doi: 10.1038/s41418-021-00804-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 179.Ye H., Kelly T.F., Samadani U., Lim L., Rubio S., Overdier D.G., et al. Hepatocyte nuclear factor 3/fork head homolog 11 is expressed in proliferating epithelial and mesenchymal cells of embryonic and adult tissues. Mol Cell Biol. 1997;17:1626–1641. doi: 10.1128/mcb.17.3.1626. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 180.Koo C.Y., Muir K.W., Lam E.W. FOXM1: from cancer initiation to progression and treatment. Biochim Biophys Acta. 2012;1819:28–37. doi: 10.1016/j.bbagrm.2011.09.004. [DOI] [PubMed] [Google Scholar]
- 181.Cui J., Shi M., Xie D., Wei D., Jia Z., Zheng S., et al. FOXM1 promotes the warburg effect and pancreatic cancer progression via transactivation of LDHA expression. Clin Cancer Res. 2014;20:2595–2606. doi: 10.1158/1078-0432.CCR-13-2407. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 182.Donehower L.A., Harvey M., Slagle B.L., McArthur M.J., Montgomery C.A., Jr., Butel J.S., et al. Mice deficient for p53 are developmentally normal but susceptible to spontaneous tumours. Nature. 1992;356:215–221. doi: 10.1038/356215a0. [DOI] [PubMed] [Google Scholar]
- 183.Jones R.G., Plas D.R., Kubek S., Buzzai M., Mu J., Xu Y., et al. AMP-activated protein kinase induces a p53-dependent metabolic checkpoint. Mol Cell. 2005;18:283–293. doi: 10.1016/j.molcel.2005.03.027. [DOI] [PubMed] [Google Scholar]
- 184.Maddocks O.D., Berkers C.R., Mason S.M., Zheng L., Blyth K., Gottlieb E., et al. Serine starvation induces stress and p53-dependent metabolic remodelling in cancer cells. Nature. 2013;493:542–546. doi: 10.1038/nature11743. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 185.Kruiswijk F., Labuschagne C.F., Vousden K.H. p53 in survival, death and metabolic health: a lifeguard with a licence to kill. Nat Rev Mol Cell Biol. 2015;16:393–405. doi: 10.1038/nrm4007. [DOI] [PubMed] [Google Scholar]
- 186.Liu X., Jiang J., Jin X., Liu Y., Xu C., Zhang J., et al. Simultaneous determination of YZG-331 and its metabolites in monkey blood by liquid chromatography–tandem mass spectrometry. J Pharm Biomed Anal. 2021;193 doi: 10.1016/j.jpba.2020.113720. [DOI] [PubMed] [Google Scholar]
- 187.Li H., Jogl G. Structural and biochemical studies of TIGAR (TP53-induced glycolysis and apoptosis regulator) J Biol Chem. 2009;284:1748–1754. doi: 10.1074/jbc.M807821200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 188.Bensaad K., Tsuruta A., Selak M.A., Vidal M.N., Nakano K., Bartrons R., et al. TIGAR, a p53-inducible regulator of glycolysis and apoptosis. Cell. 2006;126:107–120. doi: 10.1016/j.cell.2006.05.036. [DOI] [PubMed] [Google Scholar]
- 189.Kondoh H., Lleonart M., Gil J., Wang J., Degan P., Peters G., et al. Glycolytic enzymes can modulate cellular life span. Cancer Res. 2005;65:177–185. [PubMed] [Google Scholar]
- 190.Contractor T., Harris C.R. p53 negatively regulates transcription of the pyruvate dehydrogenase kinase Pdk2. Cancer Res. 2012;72:560–567. doi: 10.1158/0008-5472.CAN-11-1215. [DOI] [PubMed] [Google Scholar]
- 191.Wang L., Xiong H., Wu F., Zhang Y., Wang J., Zhao L., et al. Hexokinase 2-mediated Warburg effect is required for PTEN- and p53-deficiency-driven prostate cancer growth. Cell Rep. 2014;8:1461–1474. doi: 10.1016/j.celrep.2014.07.053. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 192.Plewka D., Plewka A., Miskiewicz A., Morek M., Bogunia E. Nuclear factor-kappa B as potential therapeutic target in human colon cancer. J Cancer Res Therapeut. 2018;14:516–520. doi: 10.4103/0973-1482.180607. [DOI] [PubMed] [Google Scholar]
- 193.Song R., Song H., Liang Y., Yin D., Zhang H., Zheng T., et al. Reciprocal activation between ATPase inhibitory factor 1 and NF-kappaB drives hepatocellular carcinoma angiogenesis and metastasis. Hepatology. 2014;60:1659–1673. doi: 10.1002/hep.27312. [DOI] [PubMed] [Google Scholar]
- 194.Wu J.Y., Lin S.S., Hsu F.T., Chung J.G. Fluoxetine inhibits DNA repair and NF-κB-modulated metastatic potential in non-small cell lung cancer. Anticancer Res. 2018;38:5201–5210. doi: 10.21873/anticanres.12843. [DOI] [PubMed] [Google Scholar]
- 195.Neil J.R., Schiemann W.P. Altered TAB1:I kappaB kinase interaction promotes transforming growth factor beta-mediated nuclear factor-kappaB activation during breast cancer progression. Cancer Res. 2008;68:1462–1470. doi: 10.1158/0008-5472.CAN-07-3094. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 196.Pantuck A., An J., Liu H., Rettig M. NF-kappaB-dependent plasticity of the epithelial to mesenchymal transition induced by Von Hippel-Lindau inactivation in renal cell carcinomas. Cancer Res. 2010;70:752–761. doi: 10.1158/0008-5472.CAN-09-2211. [DOI] [PubMed] [Google Scholar]
- 197.Fan Y., Mao R., Yang J. NF-κB and STAT3 signaling pathways collaboratively link inflammation to cancer. Protein Cell. 2013;4:176–185. doi: 10.1007/s13238-013-2084-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 198.Lee S.Y., Jeon H.M., Ju M.K., Kim C.H., Yoon G., Han S.I., et al. Wnt/Snail signaling regulates cytochrome c oxidase and glucose metabolism. Cancer Res. 2012;72:3607–3617. doi: 10.1158/0008-5472.CAN-12-0006. [DOI] [PubMed] [Google Scholar]
- 199.Dong C., Yuan T., Wu Y., Wang Y., Fan T., Miriyala S., et al. Loss of FBP1 by Snail-mediated repression provides metabolic advantages in basal-like breast cancer. Cancer Cell. 2013;23:316–331. doi: 10.1016/j.ccr.2013.01.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 200.Wu Y., Deng J., Rychahou P.G., Qiu S., Evers B.M., Zhou B.P. Stabilization of snail by NF-kappaB is required for inflammation-induced cell migration and invasion. Cancer Cell. 2009;15:416–428. doi: 10.1016/j.ccr.2009.03.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 201.El-Ashmawy N.E., El-Zamarany E.A., Khedr E.G., Abo-Saif M.A. Effect of modification of MTDH gene expression on colorectal cancer aggressiveness. Gene. 2019;698:92–99. doi: 10.1016/j.gene.2019.02.069. [DOI] [PubMed] [Google Scholar]
- 202.Harding H.P., Zhang Y., Zeng H., Novoa I., Lu P.D., Calfon M., et al. An integrated stress response regulates amino acid metabolism and resistance to oxidative stress. Mol Cell. 2003;11:619–633. doi: 10.1016/s1097-2765(03)00105-9. [DOI] [PubMed] [Google Scholar]
- 203.Pathria G., Scott D.A., Feng Y., Sang Lee J., Fujita Y., Zhang G., et al. Targeting the Warburg effect via LDHA inhibition engages ATF4 signaling for cancer cell survival. EMBO J. 2018;37 doi: 10.15252/embj.201899735. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 204.Lin J., Handschin C., Spiegelman B.M. Metabolic control through the PGC-1 family of transcription coactivators. Cell Metabol. 2005;1:361–370. doi: 10.1016/j.cmet.2005.05.004. [DOI] [PubMed] [Google Scholar]
- 205.Dan L., Wang C., Ma P., Yu Q., Gu M., Dong L., et al. PGC1alpha promotes cholangiocarcinoma metastasis by upregulating PDHA1 and MPC1 expression to reverse the Warburg effect. Cell Death Dis. 2018;9:466. doi: 10.1038/s41419-018-0494-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 206.Li W., Wong C.C., Zhang X., Kang W., Nakatsu G., Zhao Q., et al. CAB39L elicited an anti-Warburg effect via a LKB1–AMPK–PGC1alpha axis to inhibit gastric tumorigenesis. Oncogene. 2018;37:6383–6398. doi: 10.1038/s41388-018-0402-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 207.Chen Y., Gao D.Y., Huang L. In vivo delivery of miRNAs for cancer therapy: challenges and strategies. Adv Drug Deliv Rev. 2015;81:128–141. doi: 10.1016/j.addr.2014.05.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 208.Mattick J.S., Makunin I.V. Non-coding RNA. Hum Mol Genet. 2006;15(Spec No 1):R17–R29. doi: 10.1093/hmg/ddl046. [DOI] [PubMed] [Google Scholar]
- 209.Sun W., Li A.Q., Zhou P., Jiang Y.Z., Jin X., Liu Y.R., et al. DSCAM-AS1 regulates the G1/S cell cycle transition and is an independent prognostic factor of poor survival in luminal breast cancer patients treated with endocrine therapy. Cancer Med. 2018;7:6137–6146. doi: 10.1002/cam4.1603. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 210.Wilusz J.E., Sunwoo H., Spector D.L. Long noncoding RNAs: functional surprises from the RNA world. Genes Dev. 2009;23:1494–1504. doi: 10.1101/gad.1800909. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 211.Chu Z., Huo N., Zhu X., Liu H., Cong R., Ma L., et al. FOXO3A-induced LINC00926 suppresses breast tumor growth and metastasis through inhibition of PGK1-mediated Warburg effect. Mol Ther. 2021;29:2737–2753. doi: 10.1016/j.ymthe.2021.04.036. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 212.Du Y., Wei N., Ma R., Jiang S.H., Song D. Long noncoding RNA MIR210HG promotes the warburg effect and tumor growth by enhancing HIF-1alpha translation in triple-negative breast cancer. Front Oncol. 2020;10 doi: 10.3389/fonc.2020.580176. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 213.Xue L., Li J., Lin Y., Liu D., Yang Q., Jian J., et al. m6 A transferase METTL3-induced lncRNA ABHD11-AS1 promotes the Warburg effect of non-small-cell lung cancer. J Cell Physiol. 2021;236:2649–2658. doi: 10.1002/jcp.30023. [DOI] [PubMed] [Google Scholar]
- 214.Panzitt K., Tschernatsch M.M., Guelly C., Moustafa T., Stradner M., Strohmaier H.M., et al. Characterization of HULC, a novel gene with striking up-regulation in hepatocellular carcinoma, as noncoding RNA. Gastroenterology. 2007;132:330–342. doi: 10.1053/j.gastro.2006.08.026. [DOI] [PubMed] [Google Scholar]
- 215.Wang C., Li Y., Yan S., Wang H., Shao X., Xiao M., et al. Interactome analysis reveals that lncRNA HULC promotes aerobic glycolysis through LDHA and PKM2. Nat Commun. 2020;11:3162. doi: 10.1038/s41467-020-16966-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 216.Liu H., Chen R., Kang F., Lai H., Wang Y. KCNQ1OT1 promotes ovarian cancer progression via modulating MIR-142-5p/CAPN10 axis. Mol Genet Genomic Med. 2020;8 doi: 10.1002/mgg3.1077. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 217.Chen L., Xiong Y., Yan C., Zhou W., Endo Y., Xue H., et al. LncRNA KCNQ1OT1 accelerates fracture healing via modulating miR-701-3p/FGFR3 axis. FASEB J. 2020;34:5208–5222. doi: 10.1096/fj.201901864RR. [DOI] [PubMed] [Google Scholar]
- 218.Chen C., Wei M., Wang C., Sun D., Liu P., Zhong X., et al. Long noncoding RNA KCNQ1OT1 promotes colorectal carcinogenesis by enhancing aerobic glycolysis via hexokinase-2. Aging (Albany NY) 2020;12:11685–11697. doi: 10.18632/aging.103334. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 219.Liu G., Ye Z., Zhao X., Ji Z. SP1-induced up-regulation of lncRNA SNHG14 as a ceRNA promotes migration and invasion of clear cell renal cell carcinoma by regulating N-WASP. Am J Cancer Res. 2017;7:2515–2525. [PMC free article] [PubMed] [Google Scholar]
- 220.Dong H., Wang W., Mo S., Liu Q., Chen X., Chen R., et al. Long non-coding RNA SNHG14 induces trastuzumab resistance of breast cancer via regulating PABPC1 expression through H3K27 acetylation. J Cell Mol Med. 2018;22:4935–4947. doi: 10.1111/jcmm.13758. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 221.Qi X., Shao M., Sun H., Shen Y., Meng D., Huo W. Long non-coding RNA SNHG14 promotes microglia activation by regulating miR-145-5p/PLA2G4A in cerebral infarction. Neuroscience. 2017;348:98–106. doi: 10.1016/j.neuroscience.2017.02.002. [DOI] [PubMed] [Google Scholar]
- 222.Lu J., Liu X., Zheng J., Song J., Liu Y., Ruan X., et al. Lin28A promotes IRF6-regulated aerobic glycolysis in glioma cells by stabilizing SNHG14. Cell Death Dis. 2020;11:447. doi: 10.1038/s41419-020-2650-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 223.Lu Z., Li Y., Wang J., Che Y., Sun S., Huang J., et al. Long non-coding RNA NKILA inhibits migration and invasion of non-small cell lung cancer via NF-kappaB/Snail pathway. J Exp Clin Cancer Res. 2017;36:54. doi: 10.1186/s13046-017-0518-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 224.Bian D., Gao C., Bao K., Song G. The long non-coding RNA NKILA inhibits the invasion-metastasis cascade of malignant melanoma via the regulation of NF-κB. Am J Cancer Res. 2017;7:28–40. [PMC free article] [PubMed] [Google Scholar]
- 225.Zhang W., Guo Q., Liu G., Zheng F., Chen J., Huang D., et al. NKILA represses nasopharyngeal carcinoma carcinogenesis and metastasis by NF-kappaB pathway inhibition. PLoS Genet. 2019;15 doi: 10.1371/journal.pgen.1008325. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 226.Xiang S., Gu H., Jin L., Thorne R.F., Zhang X.D., Wu M. LncRNA IDH1-AS1 links the functions of c-Myc and HIF1alpha via IDH1 to regulate the Warburg effect. Proc Natl Acad Sci U S A. 2018;115:E1465–E1474. doi: 10.1073/pnas.1711257115. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 227.Yang F., Zhang H., Mei Y., Wu M. Reciprocal regulation of HIF-1alpha and lincRNA-p21 modulates the Warburg effect. Mol Cell. 2014;53:88–100. doi: 10.1016/j.molcel.2013.11.004. [DOI] [PubMed] [Google Scholar]
- 228.Liao M., Liao W., Xu N., Li B., Liu F., Zhang S., et al. LncRNA EPB41L4A-AS1 regulates glycolysis and glutaminolysis by mediating nucleolar translocation of HDAC2. EBioMedicine. 2019;41:200–213. doi: 10.1016/j.ebiom.2019.01.035. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 229.Liang Y., Zhang D., Zheng T., Yang G., Wang J., Meng F., et al. lncRNA–SOX2OT promotes hepatocellular carcinoma invasion and metastasis through miR-122-5p-mediated activation of PKM2. Oncogenesis. 2020;9:54. doi: 10.1038/s41389-020-0242-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 230.Malakar P., Stein I., Saragovi A., Winkler R., Stern-Ginossar N., Berger M., et al. Long noncoding RNA MALAT1 regulates cancer glucose metabolism by enhancing mTOR-mediated translation of TCF7L2. Cancer Res. 2019;79:2480–2493. doi: 10.1158/0008-5472.CAN-18-1432. [DOI] [PubMed] [Google Scholar]
- 231.Vriens M.R., Weng J., Suh I., Huynh N., Guerrero M.A., Shen W.T., et al. MicroRNA expression profiling is a potential diagnostic tool for thyroid cancer. Cancer. 2012;118:3426–3432. doi: 10.1002/cncr.26587. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 232.Zhang L.F., Jiang S., Liu M.F. MicroRNA regulation and analytical methods in cancer cell metabolism. Cell Mol Life Sci. 2017;74:2929–2941. doi: 10.1007/s00018-017-2508-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 233.Taniguchi K., Sugito N., Kumazaki M., Shinohara H., Yamada N., Nakagawa Y., et al. MicroRNA-124 inhibits cancer cell growth through PTB1/PKM1/PKM2 feedback cascade in colorectal cancer. Cancer Lett. 2015;363:17–27. doi: 10.1016/j.canlet.2015.03.026. [DOI] [PubMed] [Google Scholar]
- 234.Liu G., Li Y.I., Gao X. Overexpression of microRNA-133b sensitizes non-small cell lung cancer cells to irradiation through the inhibition of glycolysis. Oncol Lett. 2016;11:2903–2908. doi: 10.3892/ol.2016.4316. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 235.Morais M., Dias F., Nogueira I., Leao A., Goncalves N., Araujo L., et al. Cancer cells' metabolism dynamics in renal cell carcinoma patients' outcome: influence of GLUT-1-related hsa-miR-144 and hsa-miR-186. Cancers. 2021;13:1733. doi: 10.3390/cancers13071733. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 236.Ahmad S.S., Glatzle J., Bajaeifer K., Buhler S., Lehmann T., Konigsrainer I., et al. Phosphoglycerate kinase 1 as a promoter of metastasis in colon cancer. Int J Oncol. 2013;43:586–590. doi: 10.3892/ijo.2013.1971. [DOI] [PubMed] [Google Scholar]
- 237.Hu H., Zhu W., Qin J., Chen M., Gong L., Li L., et al. Acetylation of PGK1 promotes liver cancer cell proliferation and tumorigenesis. Hepatology. 2017;65:515–528. doi: 10.1002/hep.28887. [DOI] [PubMed] [Google Scholar]
- 238.Fu D., He C., Wei J., Zhang Z., Luo Y., Tan H., et al. PGK1 is a potential survival biomarker and invasion promoter by regulating the HIF-1alpha-mediated epithelial–mesenchymal transition process in breast cancer. Cell Physiol Biochem. 2018;51:2434–2444. doi: 10.1159/000495900. [DOI] [PubMed] [Google Scholar]
- 239.Ye T., Liang Y., Zhang D., Zhang X. MicroRNA-16-1-3p represses breast tumor growth and metastasis by inhibiting PGK1-mediated Warburg effect. Front Cell Dev Biol. 2020;8 doi: 10.3389/fcell.2020.615154. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 240.Pujol-Gimenez J., de Heredia F.P., Idoate M.A., Airley R., Lostao M.P., Evans A.R. Could GLUT12 be a potential therapeutic target in cancer treatment? A preliminary report. J Cancer. 2015;6:139–143. doi: 10.7150/jca.10429. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 241.Liu H., Jiang B. Let-7a-5p represses proliferation, migration, invasion and epithelial–mesenchymal transition by targeting Smad2 in TGF-b2-induced human lens epithelial cells. J Bio Sci. 2020;45:59. [PubMed] [Google Scholar]
- 242.Liu T.P., Huang C.C., Yeh K.T., Ke T.W., Wei P.L., Yang J.R., et al. Down-regulation of let-7a-5p predicts lymph node metastasis and prognosis in colorectal cancer: implications for chemotherapy. Surg Oncol. 2016;25:429–434. doi: 10.1016/j.suronc.2016.05.016. [DOI] [PubMed] [Google Scholar]
- 243.Chen C., Liu X., Chen C., Chen Q., Dong Y., Hou B. Clinical significance of let-7a-5p and miR-21-5p in patients with breast cancer. Ann Clin Lab Sci. 2019;49:302–308. [PubMed] [Google Scholar]
- 244.Shi Y., Zhang Y., Ran F., Liu J., Lin J., Hao X., et al. Let-7a-5p inhibits triple-negative breast tumor growth and metastasis through GLUT12-mediated warburg effect. Cancer Lett. 2020;495:53–65. doi: 10.1016/j.canlet.2020.09.012. [DOI] [PubMed] [Google Scholar]
- 245.Eichner L.J., Perry M.C., Dufour C.R., Bertos N., Park M., St-Pierre J., et al. miR-378(∗) mediates metabolic shift in breast cancer cells via the PGC-1beta/ERRgamma transcriptional pathway. Cell Metabol. 2010;12:352–361. doi: 10.1016/j.cmet.2010.09.002. [DOI] [PubMed] [Google Scholar]
- 246.Li L., Kang L., Zhao W., Feng Y., Liu W., Wang T., et al. miR-30a-5p suppresses breast tumor growth and metastasis through inhibition of LDHA-mediated Warburg effect. Cancer Lett. 2017;400:89–98. doi: 10.1016/j.canlet.2017.04.034. [DOI] [PubMed] [Google Scholar]
- 247.Zhang R., Su J., Xue S.L., Yang H., Ju L.L., Ji Y., et al. HPV E6/p53 mediated down-regulation of miR-34a inhibits Warburg effect through targeting LDHA in cervical cancer. Am J Cancer Res. 2016;6:312–320. [PMC free article] [PubMed] [Google Scholar]
- 248.Jin L., Wessely O., Marcusson E.G., Ivan C., Calin G.A., Alahari S.K. Prooncogenic factors miR-23b and miR-27b are regulated by Her2/Neu, EGF, and TNF-alpha in breast cancer. Cancer Res. 2013;73:2884–2896. doi: 10.1158/0008-5472.CAN-12-2162. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 249.Wang Y., Rathinam R., Walch A., Alahari S.K. ST14 (suppression of tumorigenicity 14) gene is a target for miR-27b, and the inhibitory effect of ST14 on cell growth is independent of miR-27b regulation. J Biol Chem. 2009;284:23094–23106. doi: 10.1074/jbc.M109.012617. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 250.Wang Y., Liang H., Zhou G., Hu X., Liu Z., Jin F., et al. HIC1 and miR-23∼27∼24 clusters form a double-negative feedback loop in breast cancer. Cell Death Differ. 2017;24:421–432. doi: 10.1038/cdd.2016.136. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 251.Chen B., Liu Y., Jin X., Lu W., Liu J., Xia Z., et al. MicroRNA-26a regulates glucose metabolism by direct targeting PDHX in colorectal cancer cells. BMC Cancer. 2014;14:443. doi: 10.1186/1471-2407-14-443. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 252.Eastlack S.C., Dong S., Ivan C., Alahari S.K. Suppression of PDHX by microRNA-27b deregulates cell metabolism and promotes growth in breast cancer. Mol Cancer. 2018;17:100. doi: 10.1186/s12943-018-0851-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 253.Jiang S., Zhang L.F., Zhang H.W., Hu S., Lu M.H., Liang S., et al. A novel miR-155/miR-143 cascade controls glycolysis by regulating hexokinase 2 in breast cancer cells. EMBO J. 2012;31:1985–1998. doi: 10.1038/emboj.2012.45. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 254.Emens L.A. Breast cancer immunotherapy: facts and hopes. Clin Cancer Res. 2018;24:511–520. doi: 10.1158/1078-0432.CCR-16-3001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 255.Du Y., Wei N., Ma R., Jiang S., Song D. A miR-210-3p regulon that controls the Warburg effect by modulating HIF-1alpha and p53 activity in triple-negative breast cancer. Cell Death Dis. 2020;11:731. doi: 10.1038/s41419-020-02952-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 256.Wang Z., Cai H., Lin L., Tang M., Cai H. Upregulated expression of microRNA-214 is linked to tumor progression and adverse prognosis in pediatric osteosarcoma. Pediatr Blood Cancer. 2014;61:206–210. doi: 10.1002/pbc.24763. [DOI] [PubMed] [Google Scholar]
- 257.Narducci M.G., Arcelli D., Picchio M.C., Lazzeri C., Pagani E., Sampogna F., et al. MicroRNA profiling reveals that miR-21, miR486 and miR-214 are upregulated and involved in cell survival in Sezary syndrome. Cell Death Dis. 2011;2:e151. doi: 10.1038/cddis.2011.32. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 258.Xin R., Bai F., Feng Y., Jiu M., Liu X., Bai F., et al. MicroRNA-214 promotes peritoneal metastasis through regulating PTEN negatively in gastric cancer. Clin Res Hepatol Gastroenterol. 2016;40:748–754. doi: 10.1016/j.clinre.2016.05.006. [DOI] [PubMed] [Google Scholar]
- 259.Zhang K.C., Xi H.Q., Cui J.X., Shen W.S., Li J.Y., Wei B., et al. Hemolysis-free plasma miR-214 as novel biomarker of gastric cancer and is correlated with distant metastasis. Am J Cancer Res. 2015;5:821–829. [PMC free article] [PubMed] [Google Scholar]
- 260.Yang T.S., Yang X.H., Wang X.D., Wang Y.L., Zhou B., Song Z.S. MiR-214 regulate gastric cancer cell proliferation, migration and invasion by targeting PTEN. Cancer Cell Int. 2013;13:68. doi: 10.1186/1475-2867-13-68. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 261.Yang L., Zhang W., Wang Y., Zou T., Zhang B., Xu Y., et al. Hypoxia-induced miR-214 expression promotes tumour cell proliferation and migration by enhancing the Warburg effect in gastric carcinoma cells. Cancer Lett. 2018;414:44–56. doi: 10.1016/j.canlet.2017.11.007. [DOI] [PubMed] [Google Scholar]
- 262.Chen Z., Zuo X., Zhang Y., Han G., Zhang L., Wu J., et al. MiR-3662 suppresses hepatocellular carcinoma growth through inhibition of HIF-1alpha-mediated Warburg effect. Cell Death Dis. 2018;9:549. doi: 10.1038/s41419-018-0616-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 263.Xu F., Yan J.J., Gan Y., Chang Y., Wang H.L., He X.X., et al. miR-885-5p negatively regulates warburg effect by silencing hexokinase 2 in liver cancer. Mol Ther Nucleic Acids. 2019;18:308–319. doi: 10.1016/j.omtn.2019.09.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 264.Zhou Y., Zheng X., Lu J., Chen W., Li X., Zhao L. Ginsenoside 20(S)-Rg3 inhibits the Warburg effect via modulating DNMT3A/miR-532-3p/HK2 pathway in ovarian cancer cells. Cell Physiol Biochem. 2018;45:2548–2559. doi: 10.1159/000488273. [DOI] [PubMed] [Google Scholar]
- 265.Zhang L.F., Lou J.T., Lu M.H., Gao C., Zhao S., Li B., et al. Suppression of miR-199a maturation by HuR is crucial for hypoxia-induced glycolytic switch in hepatocellular carcinoma. EMBO J. 2015;34:2671–2685. doi: 10.15252/embj.201591803. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 266.Guo W., Qiu Z., Wang Z., Wang Q., Tan N., Chen T., et al. MiR-199a-5p is negatively associated with malignancies and regulates glycolysis and lactate production by targeting hexokinase 2 in liver cancer. Hepatology. 2015;62:1132–1144. doi: 10.1002/hep.27929. [DOI] [PubMed] [Google Scholar]
- 267.Wang B., Hsu S.H., Frankel W., Ghoshal K., Jacob S.T. Stat3-mediated activation of microRNA-23a suppresses gluconeogenesis in hepatocellular carcinoma by down-regulating glucose-6-phosphatase and peroxisome proliferator-activated receptor gamma, coactivator 1 alpha. Hepatology. 2012;56:186–197. doi: 10.1002/hep.25632. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 268.Nie H., Li J., Yang X.M., Cao Q.Z., Feng M.X., Xue F., et al. Mineralocorticoid receptor suppresses cancer progression and the Warburg effect by modulating the miR-338-3p-PKLR axis in hepatocellular carcinoma. Hepatology. 2015;62:1145–1159. doi: 10.1002/hep.27940. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 269.Eilers M., Eisenman R.N. Myc's broad reach. Genes Dev. 2008;22:2755–2766. doi: 10.1101/gad.1712408. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 270.Li F., Wang Y., Zeller K.I., Potter J.J., Wonsey D.R., O'Donnell K.A., et al. Myc stimulates nuclearly encoded mitochondrial genes and mitochondrial biogenesis. Mol Cell Biol. 2005;25:6225–6234. doi: 10.1128/MCB.25.14.6225-6234.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 271.Porkka K.P., Pfeiffer M.J., Waltering K.K., Vessella R.L., Tammela T.L., Visakorpi T. MicroRNA expression profiling in prostate cancer. Cancer Res. 2007;67:6130–6135. doi: 10.1158/0008-5472.CAN-07-0533. [DOI] [PubMed] [Google Scholar]
- 272.Gao P., Tchernyshyov I., Chang T.C., Lee Y.S., Kita K., Ochi T., et al. c-Myc suppression of miR-23a/b enhances mitochondrial glutaminase expression and glutamine metabolism. Nature. 2009;458:762–765. doi: 10.1038/nature07823. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 273.Li J., Zhang S., Zou Y., Wu L., Pei M., Jiang Y. miR-145 promotes miR-133b expression through c-myc and DNMT3A-mediated methylation in ovarian cancer cells. J Cell Physiol. 2020;235:4291–4301. doi: 10.1002/jcp.29306. [DOI] [PubMed] [Google Scholar]
- 274.Ebron J.S., Shankar E., Singh J., Sikand K., Weyman C.M., Gupta S., et al. MiR-644a disrupts oncogenic transformation and Warburg effect by direct modulation of multiple genes of tumor-promoting pathways. Cancer Res. 2019;79:1844–1856. doi: 10.1158/0008-5472.CAN-18-2993. [DOI] [PubMed] [Google Scholar]
- 275.Masui K., Tanaka K., Akhavan D., Babic I., Gini B., Matsutani T., et al. mTOR complex 2 controls glycolytic metabolism in glioblastoma through FoxO acetylation and upregulation of c-Myc. Cell Metabol. 2013;18:726–739. doi: 10.1016/j.cmet.2013.09.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 276.Ruscetti M., Morris JPt, Mezzadra R., Russell J., Leibold J., Romesser P.B., et al. Senescence-induced vascular remodeling creates therapeutic vulnerabilities in pancreas cancer. Cell. 2021;184:4838–4839. doi: 10.1016/j.cell.2021.07.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 277.Ying H., Kimmelman A.C., Lyssiotis C.A., Hua S., Chu G.C., Fletcher-Sananikone E., et al. Oncogenic Kras maintains pancreatic tumors through regulation of anabolic glucose metabolism. Cell. 2012;149:656–670. doi: 10.1016/j.cell.2012.01.058. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 278.Zhu H., Shyh-Chang N., Segre A.V., Shinoda G., Shah S.P., Einhorn W.S., et al. The Lin28/let-7 axis regulates glucose metabolism. Cell. 2011;147:81–94. doi: 10.1016/j.cell.2011.08.033. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 279.Abbott A.L., Alvarez-Saavedra E., Miska E.A., Lau N.C., Bartel D.P., Horvitz H.R., et al. The let-7 microRNA family members mir-48, mir-84, and mir-241 function together to regulate developmental timing in Caenorhabditis elegans. Dev Cell. 2005;9:403–414. doi: 10.1016/j.devcel.2005.07.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 280.Ambros V., Horvitz H.R. Heterochronic mutants of the nematode Caenorhabditis elegans. Science. 1984;226:409–416. doi: 10.1126/science.6494891. [DOI] [PubMed] [Google Scholar]
- 281.Nimmo R.A., Slack F.J. An elegant miRror: microRNAs in stem cells, developmental timing and cancer. Chromosoma. 2009;118:405–418. doi: 10.1007/s00412-009-0210-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 282.Ma X., Li C., Sun L., Huang D., Li T., He X., et al. Lin28/let-7 axis regulates aerobic glycolysis and cancer progression via PDK1. Nat Commun. 2014;5:5212. doi: 10.1038/ncomms6212. [DOI] [PubMed] [Google Scholar]
- 283.Xiang Z., Xu C., Wu G., Liu B., Wu D. CircRNA–UCK2 increased TET1 inhibits proliferation and invasion of prostate cancer cells via sponge miRNA-767-5p. Open Med. 2019;14:833–842. doi: 10.1515/med-2019-0097. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 284.Feng Y., Yang Y., Zhao X., Fan Y., Zhou L., Rong J., et al. Circular RNA circ0005276 promotes the proliferation and migration of prostate cancer cells by interacting with FUS to transcriptionally activate XIAP. Cell Death Dis. 2019;10:792. doi: 10.1038/s41419-019-2028-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 285.Shan G., Shao B., Liu Q., Zeng Y., Fu C., Chen A., et al. circFMN2 sponges miR-1238 to promote the expression of LIM-homeobox gene 2 in prostate cancer cells. Mol Ther Nucleic Acids. 2020;21:133–146. doi: 10.1016/j.omtn.2020.05.008. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
- 286.Cao L., Wang M., Dong Y., Xu B., Chen J., Ding Y., et al. Circular RNA circRNF20 promotes breast cancer tumorigenesis and Warburg effect through miR-487a/HIF-1alpha/HK2. Cell Death Dis. 2020;11:145. doi: 10.1038/s41419-020-2336-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 287.Chen L., Zhang S., Wu J., Cui J., Zhong L., Zeng L., et al. circRNA_100290 plays a role in oral cancer by functioning as a sponge of the miR-29 family. Oncogene. 2017;36:4551–4561. doi: 10.1038/onc.2017.89. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
- 288.Chen X., Yu J., Tian H., Shan Z., Liu W., Pan Z., et al. Circle RNA hsa_circRNA_100290 serves as a ceRNA for miR-378a to regulate oral squamous cell carcinoma cells growth via glucose transporter-1 (GLUT1) and glycolysis. J Cell Physiol. 2019;234:19130–19140. doi: 10.1002/jcp.28692. [DOI] [PubMed] [Google Scholar]
- 289.Xia P., Xu X.Y. PI3K/Akt/mTOR signaling pathway in cancer stem cells: from basic research to clinical application. Am J Cancer Res. 2015;5:1602–1609. [PMC free article] [PubMed] [Google Scholar]
- 290.Su M.A., Huang Y.T., Chen I.T., Lee D.Y., Hsieh Y.C., Li C.Y., et al. An invertebrate Warburg effect: a shrimp virus achieves successful replication by altering the host metabolome via the PI3K–Akt–mTOR pathway. PLoS Pathog. 2014;10 doi: 10.1371/journal.ppat.1004196. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 291.Lu C.L., Qin L., Liu H.C., Candas D., Fan M., Li J.J. Tumor cells switch to mitochondrial oxidative phosphorylation under radiation via mTOR-mediated hexokinase II inhibition—a Warburg-reversing effect. PLoS One. 2015;10 doi: 10.1371/journal.pone.0121046. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 292.Polivka J., Jr., Janku F. Molecular targets for cancer therapy in the PI3K/AKT/mTOR pathway. Pharmacol Ther. 2014;142:164–175. doi: 10.1016/j.pharmthera.2013.12.004. [DOI] [PubMed] [Google Scholar]
- 293.Zhang X., Wang S., Wang H., Cao J., Huang X., Chen Z., et al. Circular RNA circNRIP1 acts as a microRNA-149-5p sponge to promote gastric cancer progression via the AKT1/mTOR pathway. Mol Cancer. 2019;18:20. doi: 10.1186/s12943-018-0935-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 294.Wang Z., Qiu H., He J., Liu L., Xue W., Fox A., et al. The emerging roles of hnRNPK. J Cell Physiol. 2020;235:1995–2008. doi: 10.1002/jcp.29186. [DOI] [PubMed] [Google Scholar]
- 295.Wiesmann N., Strozynski J., Beck C., Zimmermann N., Mendler S., Gieringer R., et al. Knockdown of hnRNPK leads to increased DNA damage after irradiation and reduces survival of tumor cells. Carcinogenesis. 2017;38:321–328. doi: 10.1093/carcin/bgx006. [DOI] [PubMed] [Google Scholar]
- 296.Chen X., Gu P., Xie R., Han J., Liu H., Wang B., et al. Heterogeneous nuclear ribonucleoprotein K is associated with poor prognosis and regulates proliferation and apoptosis in bladder cancer. J Cell Mol Med. 2017;21:1266–1279. doi: 10.1111/jcmm.12999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 297.Gao T., Liu X., He B., Nie Z., Zhu C., Zhang P., et al. Exosomal lncRNA 91H is associated with poor development in colorectal cancer by modifying HNRNPK expression. Cancer Cell Int. 2018;18:11. doi: 10.1186/s12935-018-0506-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 298.Lu Y., Cheng J., Cai W., Zhuo H., Wu G., Cai J. Inhibition of circRNA circVPS33B reduces Warburg effect and tumor growth through regulating the miR-873-5p/HNRNPK axis in infiltrative gastric cancer. OncoTargets Ther. 2021;14:3095–3108. doi: 10.2147/OTT.S292575. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 299.Liu F., Ma F., Wang Y., Hao L., Zeng H., Jia C., et al. PKM2 methylation by CARM1 activates aerobic glycolysis to promote tumorigenesis. Nat Cell Biol. 2017;19:1358–1370. doi: 10.1038/ncb3630. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 300.Gordin D., Shah H., Shinjo T., St-Louis R., Qi W., Park K., et al. Characterization of glycolytic enzymes and pyruvate kinase M2 in type 1 and 2 diabetic nephropathy. Diabetes Care. 2019;42:1263–1273. doi: 10.2337/dc18-2585. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 301.Zhao X., Tian Z., Liu L. circATP2B1 promotes aerobic glycolysis in gastric cancer cells through regulation of the miR-326 gene cluster. Front Oncol. 2021;11 doi: 10.3389/fonc.2021.628624. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 302.Pu Z., Xu M., Yuan X., Xie H., Zhao J. Circular RNA circCUL3 accelerates the Warburg effect progression of gastric cancer through regulating the STAT3/HK2 axis. Mol Ther Nucleic Acids. 2020;22:310–318. doi: 10.1016/j.omtn.2020.08.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 303.Kakaei F., Beheshtirouy S., Nejatollahi S.M., Zarrintan S., Mafi M.R. Surgical treatment of gallbladder carcinoma: a critical review. Updates Surg. 2015;67:339–351. doi: 10.1007/s13304-015-0328-x. [DOI] [PubMed] [Google Scholar]
- 304.Lazcano-Ponce E.C., Miquel J.F., Munoz N., Herrero R., Ferrecio C., Wistuba, et al. Epidemiology and molecular pathology of gallbladder cancer. CA Cancer J Clin. 2001;51:349–364. doi: 10.3322/canjclin.51.6.349. [DOI] [PubMed] [Google Scholar]
- 305.Wang S., Zhang Y., Cai Q., Ma M., Jin L.Y., Weng M., et al. Circular RNA FOXP1 promotes tumor progression and Warburg effect in gallbladder cancer by regulating PKLR expression. Mol Cancer. 2019;18:145. doi: 10.1186/s12943-019-1078-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 306.Hu W., Bi Z.Y., Chen Z.L., Liu C., Li L.L., Zhang F., et al. Emerging landscape of circular RNAs in lung cancer. Cancer Lett. 2018;427:18–27. doi: 10.1016/j.canlet.2018.04.006. [DOI] [PubMed] [Google Scholar]
- 307.Kristensen L.S., Hansen T.B., Veno M.T., Kjems J. Circular RNAs in cancer: opportunities and challenges in the field. Oncogene. 2018;37:555–565. doi: 10.1038/onc.2017.361. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 308.Shen S., Yao T., Xu Y., Zhang D., Fan S., Ma J. CircECE1 activates energy metabolism in osteosarcoma by stabilizing c-Myc. Mol Cancer. 2020;19:151. doi: 10.1186/s12943-020-01269-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 309.Yang Z., Zhao N., Cui J., Wu H., Xiong J., Peng T. Exosomes derived from cancer stem cells of gemcitabine-resistant pancreatic cancer cells enhance drug resistance by delivering miR-210. Cell Oncol. 2020;43:123–136. doi: 10.1007/s13402-019-00476-6. [DOI] [PubMed] [Google Scholar]
- 310.Xu F., Wu H., Xiong J., Peng T. Long non-coding RNA DLEU2L targets miR-210-3p to suppress gemcitabine resistance in pancreatic cancer cells via BRCA2 regulation. Front Mol Biosci. 2021;8 doi: 10.3389/fmolb.2021.645365. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 311.Wang J., Xie S., Yang J., Xiong H., Jia Y., Zhou Y., et al. The long noncoding RNA H19 promotes tamoxifen resistance in breast cancer via autophagy. J Hematol Oncol. 2019;12:81. doi: 10.1186/s13045-019-0747-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 312.Zhang Y., Huang W., Yuan Y., Li J., Wu J., Yu J., et al. Long non-coding RNA H19 promotes colorectal cancer metastasis via binding to hnRNPA2B1. J Exp Clin Cancer Res. 2020;39:141. doi: 10.1186/s13046-020-01619-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 313.Yoshimura H., Matsuda Y., Yamamoto M., Michishita M., Takahashi K., Sasaki N., et al. Reduced expression of the H19 long non-coding RNA inhibits pancreatic cancer metastasis. Lab Invest. 2018;98:814–824. doi: 10.1038/s41374-018-0048-1. [DOI] [PubMed] [Google Scholar]
- 314.Yan J., Zhang Y., She Q., Li X., Peng L., Wang X., et al. Long noncoding RNA H19/miR-675 axis promotes gastric cancer via FADD/Caspase 8/Caspase 3 signaling pathway. Cell Physiol Biochem. 2017;42:2364–2376. doi: 10.1159/000480028. [DOI] [PubMed] [Google Scholar]
- 315.Lecerf C., Le Bourhis X., Adriaenssens E. The long non-coding RNA H19: an active player with multiple facets to sustain the hallmarks of cancer. Cell Mol Life Sci. 2019;76:4673–4687. doi: 10.1007/s00018-019-03240-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 316.Sun L., Li J., Yan W., Yao Z., Wang R., Zhou X., et al. H19 promotes aerobic glycolysis, proliferation, and immune escape of gastric cancer cells through the microRNA-519d-3p/lactate dehydrogenase A axis. Cancer Sci. 2021;112:2245–2259. doi: 10.1111/cas.14896. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 317.Zheng X., Zhou Y., Chen W., Chen L., Lu J., He F., et al. Ginsenoside 20(S)-Rg3 prevents PKM2-targeting miR-324-5p from H19 sponging to antagonize the Warburg effect in ovarian cancer cells. Cell Physiol Biochem. 2018;51:1340–1353. doi: 10.1159/000495552. [DOI] [PubMed] [Google Scholar]
- 318.Huang X., Zhong R., He X., Deng Q., Peng X., Li J., et al. Investigations on the mechanism of progesterone in inhibiting endometrial cancer cell cycle and viability via regulation of long noncoding RNA NEAT1/microRNA-146b-5p mediated Wnt/beta-catenin signaling. IUBMB Life. 2019;71:223–234. doi: 10.1002/iub.1959. [DOI] [PubMed] [Google Scholar]
- 319.Li W., Zhang Z., Liu X., Cheng X., Zhang Y., Han X., et al. The FOXN3–NEAT1–SIN3A repressor complex promotes progression of hormonally responsive breast cancer. J Clin Invest. 2017;127:3421–3440. doi: 10.1172/JCI94233. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 320.An J., Lv W., Zhang Y. LncRNA NEAT1 contributes to paclitaxel resistance of ovarian cancer cells by regulating ZEB1 expression via miR-194. OncoTargets Ther. 2017;10:5377–5390. doi: 10.2147/OTT.S147586. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 321.He Z., Dang J., Song A., Cui X., Ma Z., Zhang Z. NEAT1 promotes colon cancer progression through sponging miR-495-3p and activating CDK6 in vitro and in vivo. J Cell Physiol. 2019;234:19582–19591. doi: 10.1002/jcp.28557. [DOI] [PubMed] [Google Scholar]
- 322.Wang C.L., Wang D., Yan B.Z., Fu J.W., Qin L. Long non-coding RNA NEAT1 promotes viability and migration of gastric cancer cell lines through up-regulation of microRNA-17. Eur Rev Med Pharmacol Sci. 2018;22:4128–4137. doi: 10.26355/eurrev_201807_15405. [DOI] [PubMed] [Google Scholar]
- 323.Cora D., Re A., Caselle M., Bussolino F. MicroRNA-mediated regulatory circuits: outlook and perspectives. Phys Biol. 2017;14 doi: 10.1088/1478-3975/aa6f21. [DOI] [PubMed] [Google Scholar]
- 324.Shao X., Zheng X., Ma D., Liu Y., Liu G. Inhibition of lncRNA-NEAT1 sensitizes 5-Fu resistant cervical cancer cells through de-repressing the microRNA-34a/LDHA axis. Biosci Rep. 2021;41 doi: 10.1042/BSR20200533. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 325.Chen W., Zheng R., Baade P.D., Zhang S., Zeng H., Bray F., et al. Cancer statistics in China, 2015. CA Cancer J Clin. 2016;66:115–132. doi: 10.3322/caac.21338. [DOI] [PubMed] [Google Scholar]
- 326.Jemal A., Center M.M., DeSantis C., Ward E.M. Global patterns of cancer incidence and mortality rates and trends. Cancer Epidemiol Biomarkers Prev. 2010;19:1893–1907. doi: 10.1158/1055-9965.EPI-10-0437. [DOI] [PubMed] [Google Scholar]
- 327.Tang W.R., Fang J.Y., Wu K.S., Shi X.J., Luo J.Y., Lin K. Epidemiological characteristics and prediction of esophageal cancer mortality in China from 1991 to 2012. Asian Pac J Cancer Prev APJCP. 2014;15:6929–6934. doi: 10.7314/apjcp.2014.15.16.6929. [DOI] [PubMed] [Google Scholar]
- 328.Liu H.E., Shi H.H., Luo X.J. Upregulated long noncoding RNA UCA1 enhances Warburg effect via miR-203/HK2 axis in esophagal cancer. JAMA Oncol. 2020;2020 doi: 10.1155/2020/8847687. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 329.Vyas S., Zaganjor E., Haigis M.C. Mitochondria and cancer. Cell. 2016;166:555–566. doi: 10.1016/j.cell.2016.07.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 330.Wallace D.C. Mitochondria and cancer. Nat Rev Cancer. 2012;12:685–698. doi: 10.1038/nrc3365. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 331.Lu J., Tan M., Cai Q. The Warburg effect in tumor progression: mitochondrial oxidative metabolism as an anti-metastasis mechanism. Cancer Lett. 2015;356:156–164. doi: 10.1016/j.canlet.2014.04.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 332.Fresquet V., Garcia-Barchino M.J., Larrayoz M., Celay J., Vicente C., Fernandez-Galilea M., et al. Endogenous retroelement activation by epigenetic therapy reverses the Warburg effect and elicits mitochondrial-mediated cancer cell death. Cancer Discov. 2021;11:1268–1285. doi: 10.1158/2159-8290.CD-20-1065. [DOI] [PubMed] [Google Scholar]
- 333.Ma B., Cheng H., Mu C., Geng G., Zhao T., Luo Q., et al. The SIAH2–NRF1 axis spatially regulates tumor microenvironment remodeling for tumor progression. Nat Commun. 2019;10:1034. doi: 10.1038/s41467-019-08618-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 334.Gordan J.D., Thompson C.B., Simon M.C. HIF and c-Myc: sibling rivals for control of cancer cell metabolism and proliferation. Cancer Cell. 2007;12:108–113. doi: 10.1016/j.ccr.2007.07.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 335.Boutin A.T., Weidemann A., Fu Z., Mesropian L., Gradin K., Jamora C., et al. Epidermal sensing of oxygen is essential for systemic hypoxic response. Cell. 2008;133:223–234. doi: 10.1016/j.cell.2008.02.038. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 336.Wang G.L., Semenza G.L. General involvement of hypoxia-inducible factor 1 in transcriptional response to hypoxia. Proc Natl Acad Sci U S A. 1993;90:4304–4308. doi: 10.1073/pnas.90.9.4304. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 337.Seagroves T.N., Ryan H.E., Lu H., Wouters B.G., Knapp M., Thibault P., et al. Transcription factor HIF-1 is a necessary mediator of the pasteur effect in mammalian cells. Mol Cell Biol. 2001;21:3436–3444. doi: 10.1128/MCB.21.10.3436-3444.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 338.Semenza G.L. HIF-1 mediates metabolic responses to intratumoral hypoxia and oncogenic mutations. J Clin Invest. 2013;123:3664–3671. doi: 10.1172/JCI67230. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 339.Semenza G.L. Oxygen sensing, homeostasis, and disease. N Engl J Med. 2011;365:537–547. doi: 10.1056/NEJMra1011165. [DOI] [PubMed] [Google Scholar]
- 340.Brugarolas J., Kaelin W.G., Jr. Dysregulation of HIF and VEGF is a unifying feature of the familial hamartoma syndromes. Cancer Cell. 2004;6:7–10. doi: 10.1016/j.ccr.2004.06.020. [DOI] [PubMed] [Google Scholar]
- 341.Rankin E.B., Giaccia A.J. The role of hypoxia-inducible factors in tumorigenesis. Cell Death Differ. 2008;15:678–685. doi: 10.1038/cdd.2008.21. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 342.Klimova T., Chandel N.S. Mitochondrial complex III regulates hypoxic activation of HIF. Cell Death Differ. 2008;15:660–666. doi: 10.1038/sj.cdd.4402307. [DOI] [PubMed] [Google Scholar]
- 343.An W.G., Kanekal M., Simon M.C., Maltepe E., Blagosklonny M.V., Neckers L.M. Stabilization of wild-type p53 by hypoxia-inducible factor 1alpha. Nature. 1998;392:405–408. doi: 10.1038/32925. [DOI] [PubMed] [Google Scholar]
- 344.Graeber T.G., Osmanian C., Jacks T., Housman D.E., Koch C.J., Lowe S.W., et al. Hypoxia-mediated selection of cells with diminished apoptotic potential in solid tumours. Nature. 1996;379:88–91. doi: 10.1038/379088a0. [DOI] [PubMed] [Google Scholar]
- 345.Espinosa E.J., Calero M., Sridevi K., Pfeffer S.R. RhoBTB3: a Rho GTPase-family ATPase required for endosome to Golgi transport. Cell. 2009;137:938–948. doi: 10.1016/j.cell.2009.03.043. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 346.Zhang C.S., Liu Q., Li M., Lin S.Y., Peng Y., Wu D., et al. RHOBTB3 promotes proteasomal degradation of HIFalpha through facilitating hydroxylation and suppresses the Warburg effect. Cell Res. 2015;25:1025–1042. doi: 10.1038/cr.2015.90. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 347.Verdin E., Hirschey M.D., Finley L.W., Haigis M.C. Sirtuin regulation of mitochondria: energy production, apoptosis, and signaling. Trends Biochem Sci. 2010;35:669–675. doi: 10.1016/j.tibs.2010.07.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 348.Finley L.W., Carracedo A., Lee J., Souza A., Egia A., Zhang J., et al. SIRT3 opposes reprogramming of cancer cell metabolism through HIF1alpha destabilization. Cancer Cell. 2011;19:416–428. doi: 10.1016/j.ccr.2011.02.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 349.Dabral S., Muecke C., Valasarajan C., Schmoranzer M., Wietelmann A., Semenza G.L., et al. A RASSF1A–HIF1alpha loop drives Warburg effect in cancer and pulmonary hypertension. Nat Commun. 2019;10:2130. doi: 10.1038/s41467-019-10044-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 350.Huang C.Y., Kuo W.T., Huang Y.C., Lee T.C., Yu L.C.H. Resistance to hypoxia-induced necroptosis is conferred by glycolytic P yruvate scavenging of mitochondrial superoxide in colorectal cancer cells. Cell Death Dis. 2013;4:e622. doi: 10.1038/cddis.2013.149. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 351.Cho Y.S., Challa S., Moquin D., Genga R., Ray T.D., Guildford M., et al. Phosphorylation-driven assembly of the RIP1–RIP3 complex regulates pro grammed necrosis and virus-induced inflammation. Cell. 2009;137:1112–1123. doi: 10.1016/j.cell.2009.05.037. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 352.Nam S.O., Yotsumoto F., Miyata K., Fukagawa S., Yamada H., Kuroki M., et al. Warburg effect regulated by amphiregulin in the development of colorec tal cancer. Cancer Med. 2015;4:575–587. doi: 10.1002/cam4.416. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 353.Fang S., Fang X. Advances in glucose metabolism research in colorectal cancer. Biomed Rep. 2016;5:289–295. doi: 10.3892/br.2016.719. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 354.Wu Z.Z., Chen L.S., Zhou R., Bin J.P., Liao Y.L., Liao W.J. Metastasis-associated in colon cancer-1 in gastric cancer: beyond metastasis. World J Gastroenterol. 2016;22:6629–6637. doi: 10.3748/wjg.v22.i29.6629. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 355.Liu J., Pan C., Guo L., Wu M., Guo J., Peng S., et al. A new mechanism of trastuzumab resistance in gastric cancer: MACC1 promotes the Warburg effect via activation of the PI3K/AKT signaling pathway. J Hematol Oncol. 2016;9:76. doi: 10.1186/s13045-016-0302-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 356.Maehama T., Dixon J.E. The tumor suppressor, PTEN/MMAC1, dephosphorylates the lipid second messenger, phosphatidylinositol 3,4,5-trisphosphate. J Biol Chem. 1998;273:13375–13378. doi: 10.1074/jbc.273.22.13375. [DOI] [PubMed] [Google Scholar]
- 357.Bonneau D., Longy M. Mutations of the human PTEN gene. Hum Mutat. 2000;16:109–122. doi: 10.1002/1098-1004(200008)16:2<109::AID-HUMU3>3.0.CO;2-0. [DOI] [PubMed] [Google Scholar]
- 358.Cantley L.C., Neel B.G. New insights into tumor suppression: PTEN suppresses tumor formation by restraining the phosphoinositide 3-kinase/AKT pathway. Proc Natl Acad Sci U S A. 1999;96:4240–4245. doi: 10.1073/pnas.96.8.4240. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 359.Simpson L., Parsons R. PTEN: life as a tumor suppressor. Exp Cell Res. 2001;264:29–41. doi: 10.1006/excr.2000.5130. [DOI] [PubMed] [Google Scholar]
- 360.Garcia-Cao I., Song M.S., Hobbs R.M., Laurent G., Giorgi C., de Boer V.C., et al. Systemic elevation of PTEN induces a tumor-suppressive metabolic state. Cell. 2012;149:49–62. doi: 10.1016/j.cell.2012.02.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 361.Pusapati R.V., Daemen A., Wilson C., Sandoval W., Gao M., Haley B., et al. mTORC1-dependent metabolic reprogramming underlies escape from glycolysis addiction in cancer cells. Cancer Cell. 2016;29:548–562. doi: 10.1016/j.ccell.2016.02.018. [DOI] [PubMed] [Google Scholar]
- 362.Hirono Y., Fushida S., Yonemura Y., Yamamoto H., Watanabe H., Raz A. Expression of autocrine motility factor receptor correlates with disease progression in human gastric cancer. Br J Cancer. 1996;74:2003–2007. doi: 10.1038/bjc.1996.667. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 363.Hardie D.G. AMP-activated protein kinase: an energy sensor that regulates all aspects of cell function. Genes Dev. 2011;25:1895–1908. doi: 10.1101/gad.17420111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 364.Gwinn D.M., Shackelford D.B., Egan D.F., Mihaylova M.M., Mery A., Vasquez D.S., et al. AMPK phosphorylation of raptor mediates a metabolic checkpoint. Mol Cell. 2008;30:214–226. doi: 10.1016/j.molcel.2008.03.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 365.Faubert B., Boily G., Izreig S., Griss T., Samborska B., Dong Z., et al. AMPK is a negative regulator of the Warburg effect and suppresses tumor growth in vivo. Cell Metabol. 2013;17:113–124. doi: 10.1016/j.cmet.2012.12.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 366.Jitschin R., Hofmann A.D., Bruns H., Giessl A., Bricks J., Berger J., et al. Mitochondrial metabolism contributes to oxidative stress and reveals therapeutic targets in chronic lymphocytic leukemia. Blood. 2014;123:2663–2672. doi: 10.1182/blood-2013-10-532200. [DOI] [PubMed] [Google Scholar]
- 367.Hosnijeh F.S., Lan Q., Rothman N., San Liu C., Cheng W.L., Nieters A., et al. Mitochondrial DNA copy number and future risk of B-cell lymphoma in a nested case-control study in the prospective EPIC cohort. Blood. 2014;124:530–535. doi: 10.1182/blood-2013-10-532085. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 368.Samudio I., Fiegl M., Andreeff M. Mitochondrial uncoupling and the Warburg effect: molecular basis for the reprogramming of cancer cell metabolism. Cancer Res. 2009;69:2163–2166. doi: 10.1158/0008-5472.CAN-08-3722. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 369.Jitschin R., Braun M., Qorraj M., Saul D., Le Blanc K., Zenz T., et al. Stromal cell-mediated glycolytic switch in CLL cells involves Notch–c-Myc signaling. Blood. 2015;125:3432–3436. doi: 10.1182/blood-2014-10-607036. [DOI] [PubMed] [Google Scholar]
- 370.Baba M., Hong S.B., Sharma N., Warren M.B., Nickerson M.L., Iwamatsu A., et al. Folliculin encoded by the BHD gene interacts with a binding protein, FNIP1, and AMPK, and is involved in AMPK and mTOR signaling. Proc Natl Acad Sci U S A. 2006;103:15552–15557. doi: 10.1073/pnas.0603781103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 371.Behrends C., Sowa M.E., Gygi S.P., Harper J.W. Network organization of the human autophagy system. Nature. 2010;466:68–76. doi: 10.1038/nature09204. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 372.Yan M., Gingras M.C., Dunlop E.A., Nouet Y., Dupuy F., Jalali Z., et al. The tumor suppressor folliculin regulates AMPK-dependent metabolic transformation. J Clin Invest. 2014;124:2640–2650. doi: 10.1172/JCI71749. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 373.Lv L., Li D., Zhao D., Lin R., Chu Y., Zhang H., et al. Acetylation targets the M2 isoform of pyruvate kinase for degradation through chaperone-mediated autophagy and promotes tumor growth. Mol Cell. 2011;42:719–730. doi: 10.1016/j.molcel.2011.04.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 374.Boutros R., Lobjois V., Ducommun B. CDC25 phosphatases in cancer cells: key players? Good targets? Nat Rev Cancer. 2007;7:495–507. doi: 10.1038/nrc2169. [DOI] [PubMed] [Google Scholar]
- 375.Liang J., Cao R., Zhang Y., Xia Y., Zheng Y., Li X., et al. PKM2 dephosphorylation by Cdc25A promotes the Warburg effect and tumorigenesis. Nat Commun. 2016;7 doi: 10.1038/ncomms12431. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 376.Wong T.L., Ng K.Y., Tan K.V., Chan L.H., Zhou L., Che N., et al. CRAF methylation by PRMT6 regulates aerobic glycolysis-driven hepatocarcinogenesis via ERK-dependent PKM2 nuclear relocalization and activation. Hepatology. 2020;71:1279–1296. doi: 10.1002/hep.30923. [DOI] [PubMed] [Google Scholar]
- 377.Li X., Jiang Y., Meisenhelder J., Yang W., Hawke D.H., Zheng Y., et al. Mitochondria-translocated PGK1 functions as a protein kinase to coordinate glycolysis and the TCA cycle in tumorigenesis. Mol Cell. 2016;61:705–719. doi: 10.1016/j.molcel.2016.02.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 378.Liang C., Qin Y., Zhang B., Ji S., Shi S., Xu W., et al. ARF6, induced by mutant Kras, promotes proliferation and Warburg effect in pancreatic cancer. Cancer Lett. 2017;388:303–311. doi: 10.1016/j.canlet.2016.12.014. [DOI] [PubMed] [Google Scholar]
- 379.Law C.T., Wei L., Tsang F.H., Chan C.Y., Xu I.M., Lai R.K., et al. HELLS regulates chromatin remodeling and epigenetic silencing of multiple tumor suppressor genes in human hepatocellular carcinoma. Hepatology. 2019;69:2013–2030. doi: 10.1002/hep.30414. [DOI] [PubMed] [Google Scholar]
- 380.Shen C., Xuan B., Yan T., Ma Y., Xu P., Tian X., et al. m6A-dependent glycolysis enhances colorectal cancer progression. Mol Cancer. 2020;19:72. doi: 10.1186/s12943-020-01190-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 381.Wang Y., Lu J.H., Wu Q.N., Jin Y., Wang D.S., Chen Y.X., et al. LncRNA LINRIS stabilizes IGF2BP2 and promotes the aerobic glycolysis in colorectal cancer. Mol Cancer. 2019;18:174. doi: 10.1186/s12943-019-1105-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 382.Dang C.V. MYC on the path to cancer. Cell. 2012;149:22–35. doi: 10.1016/j.cell.2012.03.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 383.Dang C.V., Le A., Gao P. MYC-induced cancer cell energy metabolism and therapeutic opportunities. Clin Cancer Res. 2009;15:6479–6483. doi: 10.1158/1078-0432.CCR-09-0889. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 384.Cancer Genome Atlas Research Network Electronic address edsc, Cancer Genome Atlas Research N. Comprehensive and integrated genomic characterization of adult soft tissue sarcomas. Cell. 2017;171:950–965 e28. doi: 10.1016/j.cell.2017.10.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 385.Fletcher C.D. The evolving classification of soft tissue tumours—an update based on the new 2013 WHO classification. Histopathology. 2014;64:2–11. doi: 10.1111/his.12267. [DOI] [PubMed] [Google Scholar]
- 386.Linch M., Miah A.B., Thway K., Judson I.R., Benson C. Systemic treatment of soft-tissue sarcoma-gold standard and novel therapies. Nat Rev Clin Oncol. 2014;11:187–202. doi: 10.1038/nrclinonc.2014.26. [DOI] [PubMed] [Google Scholar]
- 387.Huangyang P., Li F., Lee P., Nissim I., Weljie A.M., Mancuso A., et al. Fructose-1,6-bisphosphatase 2 inhibits sarcoma progression by restraining mitochondrial biogenesis. Cell Metabol. 2020;31:174–188 e7. doi: 10.1016/j.cmet.2019.10.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 388.Dong C., Yuan T., Wu Y., Wang Y., Fan T.W., Miriyala S., et al. Loss of FBP1 by Snail-mediated repression provides metabolic advantages in basal-like breast cancer. Cancer Cell. 2013;23:316–331. doi: 10.1016/j.ccr.2013.01.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 389.Li B., Qiu B., Lee D.S., Walton Z.E., Ochocki J.D., Mathew L.K., et al. Fructose-1,6-bisphosphatase opposes renal carcinoma progression. Nature. 2014;513:251–255. doi: 10.1038/nature13557. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 390.Qiu Y., Zhu H., Xu D., Feng Q., Wen C., Du Y., et al. RING-finger protein 6 enhances c-Myc-mediated Warburg effect by promoting MAD1 degradation to facilitate pancreatic cancer metastasis. Am J Cancer Res. 2021;11:2025–2043. [PMC free article] [PubMed] [Google Scholar]
- 391.Finkel T., Deng C.X., Mostoslavsky R. Recent progress in the biology and physiology of sirtuins. Nature. 2009;460:587–591. doi: 10.1038/nature08197. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 392.Zhong L., D'Urso A., Toiber D., Sebastian C., Henry R.E., Vadysirisack D.D., et al. The histone deacetylase Sirt6 regulates glucose homeostasis via Hif1alpha. Cell. 2010;140:280–293. doi: 10.1016/j.cell.2009.12.041. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 393.Sebastian C., Zwaans B.M., Silberman D.M., Gymrek M., Goren A., Zhong L., et al. The histone deacetylase SIRT6 is a tumor suppressor that controls cancer metabolism. Cell. 2012;151:1185–1199. doi: 10.1016/j.cell.2012.10.047. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 394.Matoba S., Kang J.G., Patino W.D., Wragg A., Boehm M., Gavrilova O., et al. p53 regulates mitochondrial respiration. Science. 2006;312:1650–1653. doi: 10.1126/science.1126863. [DOI] [PubMed] [Google Scholar]
- 395.Clark A.D., Oldenbroek M., Boyer T.G. Mediator kinase module and human tumorigenesis. Crit Rev Biochem Mol Biol. 2015;50:393–426. doi: 10.3109/10409238.2015.1064854. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 396.Firestein R., Bass A.J., Kim S.Y., Dunn I.F., Silver S.J., Guney I., et al. CDK8 is a colorectal cancer oncogene that regulates beta-catenin activity. Nature. 2008;455:547–551. doi: 10.1038/nature07179. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 397.Firestein R., Shima K., Nosho K., Irahara N., Baba Y., Bojarski E., et al. CDK8 expression in 470 colorectal cancers in relation to beta-catenin activation, other molecular alterations and patient survival. Int J Cancer. 2010;126:2863–2873. doi: 10.1002/ijc.24908. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 398.Galbraith M.D., Donner A.J., Espinosa J.M. CDK8: a positive regulator of transcription. Transcription. 2010;1:4–12. doi: 10.4161/trns.1.1.12373. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 399.Kapoor A., Goldberg M.S., Cumberland L.K., Ratnakumar K., Segura M.F., Emanuel P.O., et al. The histone variant macroH2A suppresses melanoma progression through regulation of CDK8. Nature. 2010;468:1105–1109. doi: 10.1038/nature09590. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 400.Galbraith M.D., Andrysik Z., Pandey A., Hoh M., Bonner E.A., Hill A.A., et al. CDK8 kinase activity promotes glycolysis. Cell Rep. 2017;21:1495–1506. doi: 10.1016/j.celrep.2017.10.058. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 401.Liao M., Zhang J., Wang G., Wang L., Liu J., Ouyang L., et al. Small-molecule drug discovery in triple negative breast cancer: curren t situation and future directions. J Med Chem. 2021;64:2382–2418. doi: 10.1021/acs.jmedchem.0c01180. [DOI] [PubMed] [Google Scholar]
- 402.Barnabas G.D., Lee J.S., Shami T., Harel M., Beck L., Selitrennik M., et al. Serine biosynthesis is a metabolic vulnerability in IDH2-driven breast cancer progression. Cancer Res. 2021;81:1443–1456. doi: 10.1158/0008-5472.CAN-19-3020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 403.Chesnelong C., Chaumeil M.M., Blough M.D., Al-Najjar M., Stechishin O.D., Chan J.A., et al. Lactate dehydrogenase A silencing in IDH mutant gliomas. Neuro Oncol. 2014;16:686–695. doi: 10.1093/neuonc/not243. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 404.Baumgart E., Fahimi H.D., Stich A., Volkl A. L-Lactate dehydrogenase A4- and A3B isoforms are bona fide peroxisomal enzymes in rat liver. Evidence for involvement in intraperoxisomal NADH reoxidation. J Biol Chem. 1996;271:3846–3855. doi: 10.1074/jbc.271.7.3846. [DOI] [PubMed] [Google Scholar]
- 405.Walenta S., Mueller-Klieser W.F. Lactate: mirror and motor of tumor malignancy. Semin Radiat Oncol. 2004;14:267–274. doi: 10.1016/j.semradonc.2004.04.004. [DOI] [PubMed] [Google Scholar]
- 406.Vousden K.H., Ryan K.M. p53 and metabolism. Nat Rev Cancer. 2009;9:691–700. doi: 10.1038/nrc2715. [DOI] [PubMed] [Google Scholar]
- 407.Levine A.J., Puzio-Kuter A.M. The control of the metabolic switch in cancers by oncogenes and tumor suppressor genes. Science. 2010;330:1340–1344. doi: 10.1126/science.1193494. [DOI] [PubMed] [Google Scholar]
- 408.Aylon Y., Oren M. New plays in the p53 theater. Curr Opin Genet Dev. 2011;21:86–92. doi: 10.1016/j.gde.2010.10.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 409.Li T., Kon N., Jiang L., Tan M., Ludwig T., Zhao Y., et al. Tumor suppression in the absence of p53-mediated cell-cycle arrest, apoptosis, and senescence. Cell. 2012;149:1269–1283. doi: 10.1016/j.cell.2012.04.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 410.Stambolsky P., Tabach Y., Fontemaggi G., Weisz L., Maor-Aloni R., Siegfried Z., et al. Modulation of the vitamin D3 response by cancer-associated mutant p53. Cancer Cell. 2010;17:273–285. doi: 10.1016/j.ccr.2009.11.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 411.Zhang C., Liu J., Liang Y., Wu R., Zhao Y., Hong X., et al. Tumour-associated mutant p53 drives the Warburg effect. Nat Commun. 2013;4:2935. doi: 10.1038/ncomms3935. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 412.Alsafadi S., Houy A., Battistella A., Popova T., Wassef M., Henry E., et al. Cancer-associated SF3B1 mutations affect alternative splicing by promoting alternative branchpoint usage. Nat Commun. 2016;7 doi: 10.1038/ncomms10615. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 413.Darman R.B., Seiler M., Agrawal A.A., Lim K.H., Peng S., Aird D., et al. Cancer-associated SF3B1 hotspot mutations induce cryptic 3′ splice site selection through use of a different branch point. Cell Rep. 2015;13:1033–1045. doi: 10.1016/j.celrep.2015.09.053. [DOI] [PubMed] [Google Scholar]
- 414.Kesarwani A.K., Ramirez O., Gupta A.K., Yang X., Murthy T., Minella A.C., et al. Cancer-associated SF3B1 mutants recognize otherwise inaccessible cryptic 3′ splice sites within RNA secondary structures. Oncogene. 2017;36:1123–1133. doi: 10.1038/onc.2016.279. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 415.Liu Z., Yoshimi A., Wang J., Cho H., Chun-Wei Lee S., Ki M., et al. Mutations in the RNA splicing factor SF3B1 promote tumorigenesis through MYC stabilization. Cancer Discov. 2020;10:806–821. doi: 10.1158/2159-8290.CD-19-1330. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 416.Biankin A.V., Waddell N., Kassahn K.S., Gingras M.C., Muthuswamy L.B., Johns A.L., et al. Pancreatic cancer genomes reveal aberrations in axon guidance pathway genes. Nature. 2012;491:399–405. doi: 10.1038/nature11547. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 417.Furney S.J., Pedersen M., Gentien D., Dumont A.G., Rapinat A., Desjardins L., et al. SF3B1 mutations are associated with alternative splicing in uveal melanoma. Cancer Discov. 2013;3:1122–1129. doi: 10.1158/2159-8290.CD-13-0330. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 418.Maguire S.L., Leonidou A., Wai P., Marchio C., Ng C.K., Sapino A., et al. SF3B1 mutations constitute a novel therapeutic target in breast cancer. J Pathol. 2015;235:571–580. doi: 10.1002/path.4483. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 419.Malcovati L., Karimi M., Papaemmanuil E., Ambaglio I., Jadersten M., Jansson M., et al. SF3B1 mutation identifies a distinct subset of myelodysplastic syndrome with ring sideroblasts. Blood. 2015;126:233–241. doi: 10.1182/blood-2015-03-633537. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 420.Wang L., Brooks A.N., Fan J., Wan Y., Gambe R., Li S., et al. Transcriptomic characterization of SF3B1 mutation reveals its pleiotropic effects in chronic lymphocytic leukemia. Cancer Cell. 2016;30:750–763. doi: 10.1016/j.ccell.2016.10.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 421.Yang J.Y., Huo Y.M., Yang M.W., Shen Y., Liu D.J., Fu X.L., et al. SF3B1 mutation in pancreatic cancer contributes to aerobic glycolysis and tumor growth through a PP2A–c-Myc axis. Mol Oncol. 2021;15:3076–3090. doi: 10.1002/1878-0261.12970. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 422.Justus C.R., Sanderlin E.J., Yang L.V. Molecular connections between cancer cell metabolism and the tumor microenvironment. Int J Mol Sci. 2015;16:11055–11086. doi: 10.3390/ijms160511055. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 423.Rossignol R., Gilkerson R., Aggeler R., Yamagata K., Remington S.J., Capaldi R.A. Energy substrate modulates mitochondrial structure and oxidative capacity in cancer cells. Cancer Res. 2004;64:985–993. doi: 10.1158/0008-5472.can-03-1101. [DOI] [PubMed] [Google Scholar]
- 424.Vaupel P., Schmidberger H., Mayer A. The Warburg effect: essential part of metabolic reprogramming and central contributor to cancer progression. Int J Radiat Biol. 2019;95:912–919. doi: 10.1080/09553002.2019.1589653. [DOI] [PubMed] [Google Scholar]
- 425.Logozzi M., Mizzoni D., Angelini D.F., Di Raimo R., Falchi M., Battistini L., et al. Microenvironmental pH and exosome levels interplay in human cancer cell lines of different histotypes. Cancers. 2018;10:370. doi: 10.3390/cancers10100370. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 426.Shimizu K., Matsuoka Y. Regulation of glycolytic flux and overflow metabolism depending on the source of energy generation for energy demand. Biotechnol Adv. 2019;37:284–305. doi: 10.1016/j.biotechadv.2018.12.007. [DOI] [PubMed] [Google Scholar]
- 427.Fong S.S., Nanchen A., Palsson B.O., Sauer U. Latent pathway activation and increased pathway capacity enable Escherichia coli adaptation to loss of key metabolic enzymes. J Biol Chem. 2006;281:8024–8033. doi: 10.1074/jbc.M510016200. [DOI] [PubMed] [Google Scholar]
- 428.Kee H.J., Cheong J.H. Tumor bioenergetics: an emerging avenue for cancer metabolism targeted therapy. BMB Rep. 2014;47:158–166. doi: 10.5483/BMBRep.2014.47.3.273. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 429.Huang Q., Zhang X. Emerging roles and research tools of atypical ubiquitination. Proteomics. 2020;20 doi: 10.1002/pmic.201900100. [DOI] [PubMed] [Google Scholar]
- 430.Deng L., Meng T., Chen L., Wei W., Wang P. The role of ubiquitination in tumorigenesis and targeted drug discovery. Signal Transduct Targeted Ther. 2020;5:11. doi: 10.1038/s41392-020-0107-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 431.Liu Y., Liu Q., Chen D., Matsuura A., Xiang L., Qi J. Inokosterone from Gentiana rigescens Franch extends the longevity of yeast and mammalian cells via antioxidative stress and mitophagy induction. Antioxidants. 2022;11:214. doi: 10.3390/antiox11020214. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 432.Sikder S., Mondal A., Das C., Kundu T.K. Autophagy in cancer: a metabolic perspective. Subcell Biochem. 2022;100:143–172. doi: 10.1007/978-3-031-07634-3_5. [DOI] [PubMed] [Google Scholar]
- 433.Yuan M., Tu B., Li H., Pang H., Zhang N., Fan M., et al. Cancer-associated fibroblasts employ NUFIP1-dependent autophagy to sec rete nucleosides and support pancreatic tumor growth. Nat Can (Ott) 2022;3:945–960. doi: 10.1038/s43018-022-00426-6. [DOI] [PubMed] [Google Scholar]
- 434.Dong D., Yao Y., Song J., Sun L., Zhang G. Cancer-associated fibroblasts regulate bladder cancer invasion and metabolic phenotypes through autophagy. Dis Markers. 2021;2021 doi: 10.1155/2021/6645220. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 435.Wilde L., Roche M., Domingo-Vidal M., Tanson K., Philp N., Curry J., et al. Metabolic coupling and the reverse Warburg effect in cancer: implications for novel biomarker and anticancer agent development. Semin Oncol. 2017;44:198–203. doi: 10.1053/j.seminoncol.2017.10.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 436.Pavlides S., Whitaker-Menezes D., Castello-Cros R., Flomenberg N., Witkiewicz A.K., Frank P.G., et al. The reverse Warburg effect: aerobic glycolysis in cancer associated fibroblasts and the tumor stroma. Cell Cycle. 2009;8:3984–4001. doi: 10.4161/cc.8.23.10238. [DOI] [PubMed] [Google Scholar]
- 437.Jena B.C., Das C.K., Banerjee I., Bharadwaj D., Majumder R., Das S., et al. TGF-β1 induced autophagy in cancer associated fibroblasts during hypoxia contributes EMT and glycolysis via MCT4 upregulation. Exp Cell Res. 2022;417 doi: 10.1016/j.yexcr.2022.113195. [DOI] [PubMed] [Google Scholar]
- 438.Straus D.S. TNFalpha and IL-17 cooperatively stimulate glucose metabolism and growth factor production in human colorectal cancer cells. Mol Cancer. 2013;12:78. doi: 10.1186/1476-4598-12-78. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 439.Liu Y., Cao Y., Zhang W., Bergmeier S., Qian Y., Akbar H., et al. A small-molecule inhibitor of glucose transporter 1 downregulates glycolysis, induces cell-cycle arrest, and inhibits cancer cell growth in vitro and in vivo. Mol Cancer Therapeut. 2012;11:1672–1682. doi: 10.1158/1535-7163.MCT-12-0131. [DOI] [PubMed] [Google Scholar]
- 440.Cho M.H., Park C.K., Park M., Kim W.K., Cho A., Kim H. Clinicopathologic features and molecular characteristics of glucose metabolism contributing to 18F-fluorodeoxyglucose uptake in gastrointestinal stromal tumors. PLoS One. 2015;10 doi: 10.1371/journal.pone.0141413. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 441.Berlth F., Monig S., Pinther B., Grimminger P., Maus M., Schlosser H., et al. Both GLUT-1 and GLUT-14 are independent prognostic factors in gastric adenocarcinoma. Ann Surg Oncol. 2015;22(Suppl 3):S822–S831. doi: 10.1245/s10434-015-4730-x. [DOI] [PubMed] [Google Scholar]
- 442.Siebeneicher H., Cleve A., Rehwinkel H., Neuhaus R., Heisler I., Muller T., et al. Identification and optimization of the first highly selective GLUT1 inhibitor BAY-876. ChemMedChem. 2016;11:2261–2271. doi: 10.1002/cmdc.201600276. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 443.Guda M.R., Labak C.M., Omar S.I., Asuthkar S., Airala S., Tuszynski J., et al. GLUT1 and TUBB4 in glioblastoma could be efficacious targets. Cancers. 2019;11:1308. doi: 10.3390/cancers11091308. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 444.Wood T.E., Dalili S., Simpson C.D., Hurren R., Mao X., Saiz F.S., et al. A novel inhibitor of glucose uptake sensitizes cells to FAS-induced cell death. Mol Cancer Therapeut. 2008;7:3546–3555. doi: 10.1158/1535-7163.MCT-08-0569. [DOI] [PubMed] [Google Scholar]
- 445.Watson R.T., Pessin J.E. Intracellular organization of insulin signaling and GLUT4 translocation. Recent Prog Horm Res. 2001;56:175–193. doi: 10.1210/rp.56.1.175. [DOI] [PubMed] [Google Scholar]
- 446.Bryant N.J., Govers R., James D.E. Regulated transport of the glucose transporter GLUT4. Nat Rev Mol Cell Biol. 2002;3:267–277. doi: 10.1038/nrm782. [DOI] [PubMed] [Google Scholar]
- 447.Govers R. Molecular mechanisms of GLUT4 regulation in adipocytes. Diabetes Metab. 2014;40:400–410. doi: 10.1016/j.diabet.2014.01.005. [DOI] [PubMed] [Google Scholar]
- 448.Mishra R.K., Wei C., Hresko R.C., Bajpai R., Heitmeier M., Matulis S.M., et al. In silico modeling-based identification of glucose transporter 4 (GLUT4)-selective inhibitors for cancer therapy. J Biol Chem. 2015;290:14441–14453. doi: 10.1074/jbc.M114.628826. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 449.Abel E.D., Peroni O., Kim J.K., Kim Y.B., Boss O., Hadro E., et al. Adipose-selective targeting of the GLUT4 gene impairs insulin action in muscle and liver. Nature. 2001;409:729–733. doi: 10.1038/35055575. [DOI] [PubMed] [Google Scholar]
- 450.Huang Y., Xian L., Liu Z., Wei L., Qin L., Xiong Y., et al. AMPKalpha2/HNF4A/BORIS/GLUT4 pathway promotes hepatocellular carcinoma cell invasion and metastasis in low glucose microenviroment. Biochem Pharmacol. 2022;203 doi: 10.1016/j.bcp.2022.115198. [DOI] [PubMed] [Google Scholar]
- 451.Wang Y.D., Li S.J., Liao J.X. Inhibition of glucose transporter 1 (GLUT1) chemosensitized head and neck cancer cells to cisplatin. Technol Cancer Res Treat. 2013;12:525–535. doi: 10.7785/tcrt.2012.500343. [DOI] [PubMed] [Google Scholar]
- 452.Abouzeid A.H., Patel N.R., Rachman I.M., Senn S., Torchilin V.P. Anti-cancer activity of anti-GLUT1 antibody-targeted polymeric micelles co-loaded with curcumin and doxorubicin. J Drug Target. 2013;21:994–1000. doi: 10.3109/1061186X.2013.840639. [DOI] [PubMed] [Google Scholar]
- 453.Shima T., Taniguchi K., Tokumaru Y., Inomata Y., Arima J., Lee S.W., et al. Glucose transporter-1 inhibition overcomes imatinib resistance in gastrointestinal stromal tumor cells. Oncol Rep. 2022;47:9345. doi: 10.3892/or.2021.8218. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 454.Ebstensen R., Plagemann P. Cytochalasin B: inhibition of glucose and glucosamine transport. Proc Natl Acad Sci USA. 1972;69:1430–1434. doi: 10.1073/pnas.69.6.1430. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 455.Burant C., Bell G. Mammalian facilitative glucose transporters: evidence for similar substrate recognition sites in functionally monomeric proteins. Biochemistry. 1992;31:10414–10420. doi: 10.1021/bi00157a032. [DOI] [PubMed] [Google Scholar]
- 456.Li Q., Manolescu A., Ritzel M., Yao S., Slugoski M., Young J., et al. Cloning and functional characterization of the human GLUT7 isoform SLC2A7 from the small intestine. Am J Physiol Gastrointest Liver Physiol. 2004;287:G236–G242. doi: 10.1152/ajpgi.00396.2003. [DOI] [PubMed] [Google Scholar]
- 457.Kapoor K., Finer-Moore J., Pedersen B., Caboni L., Waight A., Hillig R., et al. Mechanism of inhibition of human glucose transporter GLUT1 is conserved between cytochalasin B and phenylalanine amides. Proc Natl Acad Sci U S A. 2016;113:4711–4716. doi: 10.1073/pnas.1603735113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 458.Siebeneicher H., Bauser M., Buchmann B., Heisler I., Müller T., Neuhaus R., et al. Identification of novel GLUT inhibitors. Bioorg Med Chem Lett. 2016;26:1732–1737. doi: 10.1016/j.bmcl.2016.02.050. [DOI] [PubMed] [Google Scholar]
- 459.Karageorgis G., Reckzeh E., Ceballos J., Schwalfenberg M., Sievers S., Ostermann C., et al. Chromopynones are pseudo natural product glucose uptake inhibitors targeting glucose transporters GLUT-1 and -3. Nat Chem. 2018;10:1103–1111. doi: 10.1038/s41557-018-0132-6. [DOI] [PubMed] [Google Scholar]
- 460.Ceballos J., Schwalfenberg M., Karageorgis G., Reckzeh E., Sievers S., Ostermann C., et al. Synthesis of indomorphan pseudo-natural product inhibitors of glucose transporters GLUT-1 and -3. Angew Chem Int Ed. 2019;58:17016–17025. doi: 10.1002/anie.201909518. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 461.Reckzeh E., Karageorgis G., Schwalfenberg M., Ceballos J., Nowacki J., Stroet M., et al. Inhibition of glucose transporters and glutaminase synergistically impairs tumor cell growth. Cell Chem Biol. 2019;26 doi: 10.1016/j.chembiol.2019.06.005. [DOI] [PubMed] [Google Scholar]
- 462.Kang S., O'Neill D., Machl A., Lumpkin C., Galda S., Sengupta S., et al. Discovery of small-molecule selective mTORC1 inhibitors via direct inhibition of glucose transporters. Cell Chem Biol. 2019;26 doi: 10.1016/j.chembiol.2019.05.009. [DOI] [PubMed] [Google Scholar]
- 463.Wang Y., Hao F., Nan Y., Qu L., Na W., Jia C., et al. PKM2 inhibitor shikonin overcomes the cisplatin resistance in bladder cancer by inducing necroptosis. Int J Biol Sci. 2018;14:1883–1891. doi: 10.7150/ijbs.27854. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 464.Zhou S., Li D., Xiao D., Wu T., Hu X., Zhang Y., et al. Inhibition of PKM2 enhances sensitivity of olaparib to ovarian cancer cells and induces DNA damage. Int J Biol Sci. 2022;18:1555–1568. doi: 10.7150/ijbs.62947. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 465.Dai Y., Liu Y., Li J., Jin M., Yang H., Huang G. Shikonin inhibited glycolysis and sensitized cisplatin treatment in non-small cell lung cancer cells via the exosomal pyruvate kinase M2 pathway. Bioengineered. 2022;13:13906–13918. doi: 10.1080/21655979.2022.2086378. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 466.Ning X., Qi H., Li R., Li Y., Jin Y., McNutt M., et al. Discovery of novel naphthoquinone derivatives as inhibitors of the tumor cell specific M2 isoform of pyruvate kinase. Eur J Med Chem. 2017;138:343–352. doi: 10.1016/j.ejmech.2017.06.064. [DOI] [PubMed] [Google Scholar]
- 467.Yang H., Zhu R., Zhao X., Liu L., Zhou Z., Zhao L., et al. Sirtuin-mediated deacetylation of hnRNP A1 suppresses glycolysis and growth in hepatocellular carcinoma. Oncogene. 2019;38:4915–4931. doi: 10.1038/s41388-019-0764-z. [DOI] [PubMed] [Google Scholar]
- 468.Ning X., Qi H., Li R., Jin Y., McNutt M., Yin Y. Synthesis and antitumor activity of novel 2,3-didithiocarbamate substituted naphthoquinones as inhibitors of pyruvate kinase M2 isoform. J Enzym Inhib Med Chem. 2018;33:126–129. doi: 10.1080/14756366.2017.1404591. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 469.Zhou Y., Huang Z., Su J., Li J., Zhao S., Wu L., et al. Benserazide is a novel inhibitor targeting PKM2 for melanoma treatment. Int J Cancer. 2020;147:139–151. doi: 10.1002/ijc.32756. [DOI] [PubMed] [Google Scholar]
- 470.Fu Y.K., Wang B.J., Tseng J.C., Huang S.H., Lin C.Y., Kuo Y.Y., et al. Combination treatment of docetaxel with caffeic acid phenethyl ester suppresses the survival and the proliferation of docetaxel-resistant prostate cancer cells via induction of apoptosis and metabolism interference. J Biomed Sci. 2022;29:16. doi: 10.1186/s12929-022-00797-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 471.Mazurek S. Pyruvate kinase type M2: a key regulator of the metabolic budget system in tumor cells. Int J Biochem Cell Biol. 2011;43:969–980. doi: 10.1016/j.biocel.2010.02.005. [DOI] [PubMed] [Google Scholar]
- 472.Jiang J.K., Walsh M.J., Brimacombe K.R., Anastasiou D., Yu Y., Israelsen W.J., et al. Probe reports from the NIH molecular libraries program. National Center for Biotechnology Information (US); Bethesda (MD): 2010. ML265: a potent PKM2 activator induces tetramerization and reduces tumor formation and size in a mouse xenograft model. 2012 Mar 16 [updated 2013 May 8] [Internet] [PubMed] [Google Scholar]
- 473.Ding Y., Xue Q., Liu S., Hu K., Wang D., Wang T., et al. Identification of parthenolide dimers as activators of pyruvate kinase M2 in xenografts of glioblastoma multiforme in vivo. J Med Chem. 2020;63:1597–1611. doi: 10.1021/acs.jmedchem.9b01328. [DOI] [PubMed] [Google Scholar]
- 474.Zhu Z., Jiang W., McGinley J.N., Thompson H.J. 2-Deoxyglucose as an energy restriction mimetic agent: effects on mammary carcinogenesis and on mammary tumor cell growth in vitro. Cancer Res. 2005;65:7023–7030. doi: 10.1158/0008-5472.CAN-05-0453. [DOI] [PubMed] [Google Scholar]
- 475.Chen Y., Wei L., Zhang X., Liu X., Chen Y., Zhang S., et al. 3-Bromopyruvate sensitizes human breast cancer cells to TRAIL-induced apoptosis via the phosphorylated AMPK-mediated upregulation of DR5. Oncol Rep. 2018;40:2435–2444. doi: 10.3892/or.2018.6644. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 476.Bhatt A.N., Chauhan A., Khanna S., Rai Y., Singh S., Soni R., et al. Transient elevation of glycolysis confers radio-resistance by facilitating DNA repair in cells. BMC Cancer. 2015;15:335. doi: 10.1186/s12885-015-1368-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 477.Ko Y., Smith B., Wang Y., Pomper M., Rini D., Torbenson M., et al. Advanced cancers: eradication in all cases using 3-bromopyruvate therapy to deplete ATP. Biochem Biophys Res Commun. 2004;324:269–275. doi: 10.1016/j.bbrc.2004.09.047. [DOI] [PubMed] [Google Scholar]
- 478.Wu L., Xu J., Yuan W., Wu B., Wang H., Liu G., et al. The reversal effects of 3-bromopyruvate on multidrug resistance in vitro and in vivo derived from human breast MCF-7/ADR cells. PLoS One. 2014;9 doi: 10.1371/journal.pone.0112132. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 479.Ihrlund L.S., Hernlund E., Khan O., Shoshan M.C. 3-Bromopyruvate as inhibitor of tumour cell energy metabolism and chemopotentiator of platinum drugs. Mol Oncol. 2008;2:94–101. doi: 10.1016/j.molonc.2008.01.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 480.Jin M.Z., Jin W.L. The updated landscape of tumor microenvironment and drug repurposing. Signal Transduct Targeted Ther. 2020;5:166. doi: 10.1038/s41392-020-00280-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 481.Agnihotri S., Mansouri S., Burrell K., Li M., Mamatjan Y., Liu J., et al. Ketoconazole and posaconazole selectively target HK2-expressing glioblastoma cells. Clin Cancer Res. 2019;25:844–855. doi: 10.1158/1078-0432.CCR-18-1854. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 482.Garcia S.N., Guedes R.C., Marques M.M. Unlocking the potential of HK2 in cancer metabolism and therapeutics. Curr Med Chem. 2019;26:7285–7322. doi: 10.2174/0929867326666181213092652. [DOI] [PubMed] [Google Scholar]
- 483.Sakakibara R., Kitajima S., Hartman F.C., Uyeda K. Hexose phosphate binding sites of fructose-6-phosphate,2-kinase:fructose-2,6-bisphosphatase. Interaction with N-bromoacetylethanolamine phosphate and 3-bromo-1,4-dihydroxy-2-butanone 1,4-bisphosphate. J Biol Chem. 1984;259:14023–14028. [PubMed] [Google Scholar]
- 484.Clem B., Telang S., Clem A., Yalcin A., Meier J., Simmons A., et al. Small-molecule inhibition of 6-phosphofructo-2-kinase activity suppresses glycolytic flux and tumor growth. Mol Cancer Therapeut. 2008;7:110–120. doi: 10.1158/1535-7163.MCT-07-0482. [DOI] [PubMed] [Google Scholar]
- 485.Kotowski K., Supplitt S., Wiczew D., Przystupski D., Bartosik W., Saczko J., et al. 3PO as a selective inhibitor of 6-phosphofructo-2-kinase/fructose-2,6-biphosphatase 3 in A375 human melanoma cells. Anticancer Res. 2020;40:2613–2625. doi: 10.21873/anticanres.14232. [DOI] [PubMed] [Google Scholar]
- 486.Pisarsky L., Bill R., Fagiani E., Dimeloe S., Goosen R.W., Hagmann J., et al. Targeting metabolic symbiosis to overcome resistance to anti-angiogenic therapy. Cell Rep. 2016;15:1161–1174. doi: 10.1016/j.celrep.2016.04.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 487.Lea M.A., Altayyar M., desBordes C. Inhibition of growth of bladder cancer cells by 3-(3-pyridinyl)-1-(4-pyridinyl)-2-propen-1-one in combination with other compounds affecting glucose metabolism. Anticancer Res. 2015;35:5889–5899. [PubMed] [Google Scholar]
- 488.Wang Y., Qu C., Liu T., Wang C. PFKFB3 inhibitors as potential anticancer agents: mechanisms of action, current developments, and structure-activity relationships. Eur J Med Chem. 2020;203 doi: 10.1016/j.ejmech.2020.112612. [DOI] [PubMed] [Google Scholar]
- 489.Akter S., Clem B.F., Lee H.J., Chesney J., Bae Y. Block copolymer micelles for controlled delivery of glycolytic enzyme inhibitors. Pharm Res (N Y) 2012;29:847–855. doi: 10.1007/s11095-011-0613-4. [DOI] [PubMed] [Google Scholar]
- 490.Clem B., O'Neal J., Tapolsky G., Clem A., Imbert-Fernandez Y., Kerr D., et al. Targeting 6-phosphofructo-2-kinase (PFKFB3) as a therapeutic strategy against cancer. Mol Cancer Therapeut. 2013;12:1461–1470. doi: 10.1158/1535-7163.MCT-13-0097. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 491.Lea M., 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;36:1479–1488. [PubMed] [Google Scholar]
- 492.Seo M., Kim J.D., Neau D., Sehgal I., Lee Y.H. Structure-based development of small molecule PFKFB3 inhibitors: a framework for potential cancer therapeutic agents targeting the Warburg effect. PLoS One. 2011;6 doi: 10.1371/journal.pone.0024179. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 493.Mondal S., Roy D., Sarkar Bhattacharya S., Jin L., Jung D., Zhang 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. doi: 10.1002/ijc.31868. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 494.Sarkar Bhattacharya S., Thirusangu P., Jin L., Roy D., Jung D., Xiao Y., et al. PFKFB3 inhibition reprograms malignant pleural mesothelioma to nutrient stress-induced macropinocytosis and ER stress as independent binary adaptive responses. Cell Death Dis. 2019;10:725. doi: 10.1038/s41419-019-1916-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 495.Boyd S., Brookfield J.L., Critchlow S.E., Cumming I.A., Curtis N.J., Debreczeni J., et al. Structure-based design of potent and selective inhibitors of the metabolic kinase PFKFB3. J Med Chem. 2015;58:3611–3625. doi: 10.1021/acs.jmedchem.5b00352. [DOI] [PubMed] [Google Scholar]
- 496.Gustafsson N.M.S., Farnegardh K., Bonagas N., Ninou A.H., Groth P., Wiita E., et al. Targeting PFKFB3 radiosensitizes cancer cells and suppresses homologous recombination. Nat Commun. 2018;9:3872. doi: 10.1038/s41467-018-06287-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 497.Yao L., Wang L., Cao Z., Hu X., Shao Z. High expression of metabolic enzyme PFKFB4 is associated with poor prognosis of operable breast cancer. Cancer Cell Int. 2019;19:165. doi: 10.1186/s12935-019-0882-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 498.Feng C., Li Y., Li K., Lyu Y., Zhu W., Jiang H., et al. PFKFB4 is overexpressed in clear-cell renal cell carcinoma promoting pentose phosphate pathway that mediates sunitinib resistance. J Exp Clin Cancer Res. 2021;40:308. doi: 10.1186/s13046-021-02103-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 499.Oudaert I., Van der Vreken A., Maes A., De Bruyne E., De Veirman K., Vanderkerken K., et al. Metabolic cross-talk within the bone marrow milieu: focus on multiple myeloma. Exp Hematol. 2022;11:49. doi: 10.1186/s40164-022-00303-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 500.Cheong J.H., Park E.S., Liang J., Dennison J.B., Tsavachidou D., Nguyen-Charles C., et al. Dual inhibition of tumor energy pathway by 2-deoxyglucose and metformin is effective against a broad spectrum of preclinical cancer models. Mol Cancer Therapeut. 2011;10:2350–2362. doi: 10.1158/1535-7163.MCT-11-0497. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 501.Zhu J., Zheng Y., Zhang H., Sun H. Targeting cancer cell metabolism: the combination of metformin and 2-deoxyglucose regulates apoptosis in ovarian cancer cells via p38 MAPK/JNK signaling pathway. Am J Transl Res. 2016;8:4812–4821. [PMC free article] [PubMed] [Google Scholar]
- 502.Hussain A., Qazi A.K., Mupparapu N., Guru S.K., Kumar A., Sharma P.R., et al. Modulation of glycolysis and lipogenesis by novel PI3K selective molecule represses tumor angiogenesis and decreases colorectal cancer growth. Cancer Lett. 2016;374:250–260. doi: 10.1016/j.canlet.2016.02.030. [DOI] [PubMed] [Google Scholar]
- 503.Kaji K., Nishimura N., Seki K., Sato S., Saikawa S., Nakanishi K., et al. Sodium glucose cotransporter 2 inhibitor canagliflozin attenuates liver cancer cell growth and angiogenic activity by inhibiting glucose uptake. Int J Cancer. 2018;142:1712–1722. doi: 10.1002/ijc.31193. [DOI] [PubMed] [Google Scholar]
- 504.Nguyen T.T.T., Zhang Y., Shang E., Shu C., Torrini C., Zhao J., et al. HDAC inhibitors elicit metabolic reprogramming by targeting super-enhancers in glioblastoma models. J Clin Invest. 2020;130:3699–3716. doi: 10.1172/JCI129049. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 505.Lu H., Yan C., Quan X.X., Yang X., Zhang J., Bian Y., et al. CK2 phosphorylates and inhibits TAp73 tumor suppressor function to promote expression of cancer stem cell genes and phenotype in head and neck cancer. Neoplasia. 2014;16:789–800. doi: 10.1016/j.neo.2014.08.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 506.Tang S., Yuan Y., Liu Z., He Y., Pan D. Casein kinase 2 inhibitor CX-4945 elicits an anti-Warburg effects through the downregulation of TAp73 and inhibits gastric tumorigenesis. Biochem Biophys Res Commun. 2020;530:686–691. doi: 10.1016/j.bbrc.2020.07.116. [DOI] [PubMed] [Google Scholar]
- 507.Annas D., Cheon S.Y., Yusuf M., Bae S.J., Ha K.T., Park K.H. Synthesis and initial screening of lactate dehydrogenase inhibitor activity of 1,3-benzodioxole derivatives. Sci Rep. 2020;10 doi: 10.1038/s41598-020-77056-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 508.Huang T., Chen Z., Li Y. Experimental study of the anti-tumor activity and effect of Shenmai Liquid on glycometabolism. Chin Tradit Pat Med. 2003;25:390–393. [Google Scholar]
- 509.Lai N., Huang X.S., Chen J. Experimental study on the differentiation effect of SMMC-7721 cells induced by Xiaochaihu Decoction drug serum. J Sichuan Tradit Chin Med. 2013;31:55–57. [Google Scholar]
- 510.Wang M.X., Li F.W., Li X.G. Effects of ginseng oil on the chemical components of SGC-823 gastric cancer cells in vitro. China J Chin Mater Med. 1992;17:110–112. [PubMed] [Google Scholar]
- 511.Feng J., Xie P., Qin Y.H. Effects ofthe expression of HIF-1a,VEGF and MVD of Lewis lung carcinoma in mice of Danshen united 5-Fu. West China J Pharm Sci. 2014;29:42–44. [Google Scholar]
- 512.Fong S., Shoemaker M., Cadaoas J., Lo A., Liao W., Tagliaferri M., et al. Molecular mechanisms underlying selective cytotoxic activity of BZL101, an extract of Scutellaria barbata, towards breast cancer cells. Cancer Biol Ther. 2008;7:577–586. doi: 10.4161/cbt.7.4.5535. [DOI] [PubMed] [Google Scholar]
- 513.Klawitter J., Klawitter J., Gurshtein J., Corby K., Fong S., Tagliaferri M., et al. Bezielle (BZL101)-induced oxidative stress damage followed by redistribution of metabolic fluxes in breast cancer cells: a combined proteomic and metabolomic study. Int J Cancer. 2011;129:2945–2957. doi: 10.1002/ijc.25965. [DOI] [PubMed] [Google Scholar]
- 514.Chen V., Staub R.E., Fong S., Tagliaferri M., Cohen I., Shtivelman E. Bezielle selectively targets mitochondria of cancer cells to inhibit glycolysis and OXPHOS. PLoS One. 2012;7 doi: 10.1371/journal.pone.0030300. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 515.Gu Y., Qi C., Sun X., Ma X., Zhang H., Hu L., et al. Arctigenin preferentially induces tumor cell death under glucose deprivation by inhibiting cellular energy metabolism. Biochem Pharmacol. 2012;84:468–476. doi: 10.1016/j.bcp.2012.06.002. [DOI] [PubMed] [Google Scholar]
- 516.Feng J., Li J., Wu L., Yu Q., Ji J., Wu J., et al. Emerging roles and the regulation of aerobic glycolysis in hepatocellular carcinoma. J Exp Clin Cancer Res. 2020;39:126. doi: 10.1186/s13046-020-01629-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 517.Hsu F.Y., Wang S.K., Duh C.Y. Xeniaphyllane-derived terpenoids from soft coral Sinularia nanolobata. Mar Drugs. 2018;16:40. doi: 10.3390/md16020040. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 518.Liu J., Wu N., Ma L., Liu M., Liu G., Zhang Y., et al. Oleanolic acid suppresses aerobic glycolysis in cancer cells by switching pyruvate kinase type M isoforms. PLoS One. 2014;9 doi: 10.1371/journal.pone.0091606. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 519.Dando I., Donadelli M., Costanzo C., Dalla Pozza E., D'Alessandro A., Zolla L., et al. Cannabinoids inhibit energetic metabolism and induce AMPK-dependent autophagy in pancreatic cancer cells. Cell Death Dis. 2013;4:e664. doi: 10.1038/cddis.2013.151. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 520.Zhang H.N., Yang L., Ling J.Y., Czajkowsky D.M., Wang J.F., Zhang X.W., et al. Systematic identification of arsenic-binding proteins reveals that hexokinase-2 is inhibited by arsenic. Proc Natl Acad Sci U S A. 2015;112:15084–15089. doi: 10.1073/pnas.1521316112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 521.Dai W., Wang F., Lu J., Xia Y., He L., Chen K., et al. By reducing hexokinase 2, resveratrol induces apoptosis in HCC cells addicted to aerobic glycolysis and inhibits tumor growth in mice. Oncotarget. 2015;6:13703–13717. doi: 10.18632/oncotarget.3800. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 522.Jung K.H., Lee J.H., Thien Quach C.H., Paik J.Y., Oh H., Park J.W., et al. Resveratrol suppresses cancer cell glucose uptake by targeting reactive oxygen species-mediated hypoxia-inducible factor-1alpha activation. J Nucl Med. 2013;54:2161–2167. doi: 10.2967/jnumed.112.115436. [DOI] [PubMed] [Google Scholar]
- 523.Dai Q., Yin Q., Wei L., Zhou Y., Qiao C., Guo Y., et al. Oroxylin A regulates glucose metabolism in response to hypoxic stress with the involvement of hypoxia-inducible factor-1 in human hepatoma HepG2 cells. Mol Carcinog. 2016;55:1275–1289. doi: 10.1002/mc.22369. [DOI] [PubMed] [Google Scholar]
- 524.Wei L., Zhou Y., Dai Q., Qiao C., Zhao L., Hui H., et al. Oroxylin A induces dissociation of hexokinase II from the mitochondria and inhibits glycolysis by SIRT3-mediated deacetylation of cyclophilin D in breast carcinoma. Cell Death Dis. 2013;4 doi: 10.1038/cddis.2013.131. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 525.Goldin N., Arzoine L., Heyfets A., Israelson A., Zaslavsky Z., Bravman T., et al. Methyl jasmonate binds to and detaches mitochondria-bound hexokinase. Oncogene. 2008;27:4636–4643. doi: 10.1038/onc.2008.108. [DOI] [PubMed] [Google Scholar]
- 526.Klippel S., Jakubikova J., Delmore J., Ooi M., McMillin D., Kastritis E., et al. Methyljasmonate displays in vitro and in vivo activity against multiple myeloma cells. Br J Haematol. 2012;159:340–351. doi: 10.1111/j.1365-2141.2012.09253.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 527.Gao F., Li M., Liu W.B., Zhou Z.S., Zhang R., Li J.L., et al. Epigallocatechin gallate inhibits human tongue carcinoma cells via HK2-mediated glycolysis. Oncol Rep. 2015;33:1533–1539. doi: 10.3892/or.2015.3727. [DOI] [PubMed] [Google Scholar]
- 528.Lu Q.Y., Zhang L., Yee J.K., Go V.W., Lee W.N. Metabolic consequences of LDHA inhibition by epigallocatechin gallate and oxamate in MIA PaCa-2 pancreatic cancer cells. Metabolomics. 2015;11:71–80. doi: 10.1007/s11306-014-0672-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 529.Gu J.W., Makey K.L., Tucker K.B., Chinchar E., Mao X., Pei I., et al. EGCG, a major green tea catechin suppresses breast tumor angiogenesis and growth via inhibiting the activation of HIF-1alpha and NFkappaB, and VEGF expression. Vasc Cell. 2013;5:9. doi: 10.1186/2045-824X-5-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 530.Wang K., Fan H., Chen Q., Ma G., Zhu M., Zhang X., et al. Curcumin inhibits aerobic glycolysis and induces mitochondrial-mediated apoptosis through hexokinase II in human colorectal cancer cells in vitro. Anti Cancer Drugs. 2015;26:15–24. doi: 10.1097/CAD.0000000000000132. [DOI] [PubMed] [Google Scholar]
- 531.Wu H., Pan L., Gao C., Xu H., Li Y., Zhang L., et al. Quercetin inhibits the proliferation of glycolysis-addicted HCC cells by reducing hexokinase 2 and Akt–mTOR pathway. Molecules. 2019;24:1993. doi: 10.3390/molecules24101993. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 532.Deng X., Zhao J., Qu L., Duan Z., Fu R., Zhu C., et al. Ginsenoside Rh4 suppresses aerobic glycolysis and the expression of PD-L1 via targeting AKT in esophageal cancer. Biochem Pharmacol. 2020;178 doi: 10.1016/j.bcp.2020.114038. [DOI] [PubMed] [Google Scholar]
- 533.Shi D., Zhao D., Niu P., Zhu Y., Zhou J., Chen H. Glycolysis inhibition via mTOR suppression is a key step in cardamonin-induced autophagy in SKOV3 cells. BMC Compl Alternative Med. 2018;18:317. doi: 10.1186/s12906-018-2380-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 534.Chen G.Q., Tang C.F., Shi X.K., Lin C.Y., Fatima S., Pan X.H., et al. Halofuginone inhibits colorectal cancer growth through suppression of Akt/mTORC1 signaling and glucose metabolism. Oncotarget. 2015;6:24148–24162. doi: 10.18632/oncotarget.4376. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 535.Wei J., Wu J., Xu W., Nie H., Zhou R., Wang R., et al. Salvianolic acid B inhibits glycolysis in oral squamous cell carcinoma via targeting PI3K/AKT/HIF-1alpha signaling pathway. Cell Death Dis. 2018;9:599. doi: 10.1038/s41419-018-0623-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 536.Zhu W., Li Y., Zhao D., Li H., Zhang W., Xu J., et al. Dihydroartemisinin suppresses glycolysis of LNCaP cells by inhibiting PI3K/AKT pathway and downregulating HIF-1alpha expression. Life Sci. 2019;233 doi: 10.1016/j.lfs.2019.116730. [DOI] [PubMed] [Google Scholar]
- 537.Liu Z., Zhu W., Kong X., Chen X., Sun X., Zhang W., et al. Tanshinone IIA inhibits glucose metabolism leading to apoptosis in cervical cancer. Oncol Rep. 2019;42:1893–1903. doi: 10.3892/or.2019.7294. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 538.Liu W., Pan H.F., Yang L.J., Zhao Z.M., Yuan D.S., Liu Y.L., et al. Panax ginseng C.A. Meyer (Rg3) ameliorates gastric precancerous lesions in Atp4a–/– mice via inhibition of glycolysis through PI3K/AKT/miRNA-21 pathway. Evid Based Complement Alternat Med. 2020;2020 doi: 10.1155/2020/2672648. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 539.Jin J., Qiu S., Wang P., Liang X., Huang F., Wu H., et al. Cardamonin inhibits breast cancer growth by repressing HIF-1alpha-dependent metabolic reprogramming. J Exp Clin Cancer Res. 2019;38:377. doi: 10.1186/s13046-019-1351-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 540.Tao H., Ding X., Wu J., Liu S., Sun W., Nie M., et al. beta-Asarone increases chemosensitivity by inhibiting tumor glycolysis in gastric cancer. Evid Based Complement Alternat Med. 2020;2020 doi: 10.1155/2020/6981520. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 541.Balamurugan K. HIF-1 at the crossroads of hypoxia, inflammation, and cancer. Int J Cancer. 2016;138:1058–1066. doi: 10.1002/ijc.29519. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 542.Lee G., Won H.S., Lee Y.M., Choi J.W., Oh T.I., Jang J.H., et al. Oxidative dimerization of PHD2 is responsible for its inactivation and contributes to metabolic reprogramming via HIF-1alpha activation. Sci Rep. 2016;6 doi: 10.1038/srep18928. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 543.Zhou Y., Cheng C., Baranenko D., Wang J., Li Y., Lu W. Effects of Acanthopanax senticosus on brain injury induced by simulated spatial radiation in mouse model based on pharmacokinetics and comparative proteomics. Int J Mol Sci. 2018;19:159. doi: 10.3390/ijms19010159. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 544.Song C., Li S., Duan F., Liu M., Shan S., Ju T., et al. The therapeutic effect of Acanthopanax senticosus components on radiation-induced brain injury based on the pharmacokinetics and neurotransmitters. Molecules. 2022;27:1106. doi: 10.3390/molecules27031106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 545.Fu B., Xue J., Li Z., Shi X., Jiang B.H., Fang J. Chrysin inhibits expression of hypoxia-inducible factor-1alpha through reducing hypoxia-inducible factor-1alpha stability and inhibiting its protein synthesis. Mol Cancer Therapeut. 2007;6:220–226. doi: 10.1158/1535-7163.MCT-06-0526. [DOI] [PubMed] [Google Scholar]
- 546.Wang H., Zhao L., Zhu L.T., Wang Y., Pan D., Yao J., et al. Wogonin reverses hypoxia resistance of human colon cancer HCT116 cells via downregulation of HIF-1alpha and glycolysis, by inhibiting PI3K/Akt signaling pathway. Mol Carcinog. 2014;53(Suppl 1):E107–E118. doi: 10.1002/mc.22052. [DOI] [PubMed] [Google Scholar]
- 547.Fang J., Bao Y.Y., Zhou S.H., Fan J. Apigenin inhibits the proliferation of adenoid cystic carcinoma via suppression of glucose transporter-1. Mol Med Rep. 2015;12:6461–6466. doi: 10.3892/mmr.2015.4233. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 548.Melstrom L.G., Salabat M.R., Ding X.Z., Milam B.M., Strouch M., Pelling J.C., et al. Apigenin inhibits the GLUT-1 glucose transporter and the phosphoinositide 3-kinase/Akt pathway in human pancreatic cancer cells. Pancreas. 2008;37:426–431. doi: 10.1097/MPA.0b013e3181735ccb. [DOI] [PubMed] [Google Scholar]
- 549.Melstrom L.G., Salabat M.R., Ding X.Z., Strouch M.J., Grippo P.J., Mirzoeva S., et al. Apigenin down-regulates the hypoxia response genes: HIF-1alpha, GLUT-1, and VEGF in human pancreatic cancer cells. J Surg Res. 2011;167:173–181. doi: 10.1016/j.jss.2010.10.041. [DOI] [PubMed] [Google Scholar]
- 550.Singh R.P., Dhanalakshmi S., Agarwal C., Agarwal R. Silibinin strongly inhibits growth and survival of human endothelial cells via cell cycle arrest and downregulation of survivin, Akt and NF-kappaB: implications for angioprevention and antiangiogenic therapy. Oncogene. 2005;24:1188–1202. doi: 10.1038/sj.onc.1208276. [DOI] [PubMed] [Google Scholar]
- 551.Zhan T., Digel M., Kuch E.M., Stremmel W., Fullekrug J. Silybin and dehydrosilybin decrease glucose uptake by inhibiting GLUT proteins. J Cell Biochem. 2011;112:849–859. doi: 10.1002/jcb.22984. [DOI] [PubMed] [Google Scholar]
- 552.Zappavigna S., Vanacore D., Lama S., Potenza N., Russo A., Ferranti P., et al. Silybin-induced apoptosis occurs in parallel to the increase of ceramides synthesis and mirnas secretion in human hepatocarcinoma cells. Int J Mol Sci. 2019;20:2190. doi: 10.3390/ijms20092190. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 553.Li Y., Sun X.X., Qian D.Z., Dai M.S. Molecular crosstalk between MYC and HIF in cancer. Front Cell Dev Biol. 2020;8 doi: 10.3389/fcell.2020.590576. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 554.Yuan J., Peng G., Xiao G., Yang Z., Huang J., Liu Q., et al. Xanthohumol suppresses glioblastoma via modulation of hexokinase 2-mediated glycolysis. J Cancer. 2020;11:4047–4058. doi: 10.7150/jca.33045. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 555.Li M., Gao F., Zhao Q., Zuo H., Liu W., Li W. Tanshinone IIA inhibits oral squamous cell carcinoma via reducing Akt–c-Myc signaling-mediated aerobic glycolysis. Cell Death Dis. 2020;11:381. doi: 10.1038/s41419-020-2579-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 556.Remels A.H., Gosker H.R., Verhees K.J., Langen R.C., Schols A.M. TNF-alpha-induced NF-kappaB activation stimulates skeletal muscle glycolytic metabolism through activation of HIF-1alpha. Endocrinology. 2015;156:1770–1781. doi: 10.1210/en.2014-1591. [DOI] [PubMed] [Google Scholar]
- 557.Zhang T., Zhu X., Wu H., Jiang K., Zhao G., Shaukat A., et al. Targeting the ROS/PI3K/AKT/HIF-1alpha/HK2 axis of breast cancer cells: combined administration of polydatin and 2-deoxy-d-glucose. J Cell Mol Med. 2019;23:3711–3723. doi: 10.1111/jcmm.14276. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 558.Jiao L., Wang S., Zheng Y., Wang N., Yang B., Wang D., et al. Betulinic acid suppresses breast cancer aerobic glycolysis via caveolin-1/NF-kappaB/c-Myc pathway. Biochem Pharmacol. 2019;161:149–162. doi: 10.1016/j.bcp.2019.01.016. [DOI] [PubMed] [Google Scholar]
- 559.Zhang C., Cai T., Zeng X., Cai D., Chen Y., Huang X., et al. Astragaloside IV reverses MNNG-induced precancerous lesions of gastric carcinoma in rats: regulation on glycolysis through miRNA-34a/LDHA pathway. Phytother Res. 2018;32:1364–1372. doi: 10.1002/ptr.6070. [DOI] [PubMed] [Google Scholar]
- 560.Li H., Xia Z., Liu L., Pan G., Ding J., Liu J., et al. Astragalus IV undermines multi-drug resistance and glycolysis of MDA-MB-231/ADR cell line by depressing hsa_circ_0001982-miR-206/miR-613 axis. Cancer Manag Res. 2021;13:5821–5833. doi: 10.2147/CMAR.S297008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 561.Cha Y.Y., Lee E.O., Lee H.J., Park Y.D., Ko S.G., Kim D.H., et al. Methylene chloride fraction of Scutellaria barbata induces apoptosis in human U937 leukemia cells via the mitochondrial signaling pathway. Clin Chim Acta. 2004;348:41–48. doi: 10.1016/j.cccn.2004.04.013. [DOI] [PubMed] [Google Scholar]
- 562.Elgendy M., Ciro M., Hosseini A., Weiszmann J., Mazzarella L., Ferrari E., et al. Combination of hypoglycemia and metformin impairs tumor metabolic plasticity and growth by modulating the PP2A–GSK3beta–MCL-1 axis. Cancer Cell. 2019;35:798–815 e5. doi: 10.1016/j.ccell.2019.03.007. [DOI] [PubMed] [Google Scholar]
- 563.Zhao Y., Butler E.B., Tan M. Targeting cellular metabolism to improve cancer therapeutics. Cell Death Dis. 2013;4:e532. doi: 10.1038/cddis.2013.60. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 564.Li W., Liu J., Zhao Y. PKM2 inhibitor shikonin suppresses TPA-induced mitochondrial malfunction and proliferation of skin epidermal JB6 cells. Mol Carcinog. 2014;53:403–412. doi: 10.1002/mc.21988. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 565.Fodor T., Szanto M., Abdul-Rahman O., Nagy L., Der A., Kiss B., et al. Combined treatment of MCF-7 cells with AICAR and methotrexate, arrests cell cycle and reverses Warburg metabolism through AMP-activated protein kinase (AMPK) and FOXO1. PLoS One. 2016;11 doi: 10.1371/journal.pone.0150232. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 566.Zhu L., Ploessl K., Zhou R., Mankoff D., Kung H.F. Metabolic imaging of glutamine in cancer. J Nucl Med. 2017;58:533–537. doi: 10.2967/jnumed.116.182345. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 567.Liao M., Qin R., Huang W., Zhu H.-P., Peng F., Han B., et al. Targeting regulated cell death (RCD) with small-molecule compounds in triple-negative breast cancer: a revisited perspective from molecular mechanisms to targeted therapies. J Hematol Oncol. 2022;15:44. doi: 10.1186/s13045-022-01260-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 568.Keren L., Bosse M., Marquez D., Angoshtari R., Jain S., Varma S., et al. A structured tumor-immune microenvironment in triple negative breast cancer revealed by multiplexed ion beam imaging. Cell. 2018;174 doi: 10.1016/j.cell.2018.08.039. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 569.Ulaner G.A., Castillo R., Goldman D.A., Wills J., Riedl C.C., Pinker-Domenig K., et al. 18F-FDG-PET/CT for systemic staging of newly diagnosed triple-negative breast cancer. Eur J Nucl Med Mol Imag. 2016;43:1937–1944. doi: 10.1007/s00259-016-3402-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 570.Tekade R.K., Sun X. The Warburg effect and glucose-derived cancer theranostics. Drug Discov Today. 2017;22:1637–1653. doi: 10.1016/j.drudis.2017.08.003. [DOI] [PubMed] [Google Scholar]
- 571.Maassen T., Vardanyan S., Brosig A., Merz H., Ranjbar M., Kakkassery V., et al. Monosomy-3 alters the expression profile of the glucose transporters GLUT1-3 in uveal melanoma. Int J Mol Sci. 2020;21:9345. doi: 10.3390/ijms21249345. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 572.Lei Y., Hu Q., Gu J. Expressions of carbohydrate response element binding protein and glucose transporters in liver cancer and clinical significance. Pathol Oncol Res. 2020;26:1331–1340. doi: 10.1007/s12253-019-00708-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 573.Coupe N.A., Karikios D., Chong S., Yap J., Ng W., Merrett N., et al. Metabolic information on staging FDG-PET-CT as a prognostic tool in the evaluation of 97 patients with gastric cancer. Ann Nucl Med. 2014;28:128–135. doi: 10.1007/s12149-013-0791-8. [DOI] [PubMed] [Google Scholar]
- 574.Cayvarli H., Bekis R., Akman T., Altun D. The role of 18F-FDG PET/CT in the evaluation of gastric cancer recurrence. Mol Imaging Radionucl Ther. 2014;23:76–83. doi: 10.4274/mirt.83803. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 575.Matthews R., Choi M. Clinical utility of positron emission tomography magnetic resonance imaging (PET-MRI) in gastrointestinal cancers. Diagnostics. 2016;6:35. doi: 10.3390/diagnostics6030035. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 576.Liang C., Qin Y., Zhang B., Ji S., Shi S., Xu W., et al. Energy sources identify metabolic phenotypes in pancreatic cancer. Acta Biochim Biophys Sin. 2016;48:969–979. doi: 10.1093/abbs/gmw097. [DOI] [PubMed] [Google Scholar]
- 577.Zhao Y., Li S., Lv J., Liu Y., Chen Y., Liu Y., et al. Generation of triacyl lipopeptide-modified glycoproteins by metabolic glycoengineering as the neoantigen to boost anti-tumor immune response. Theranostics. 2021;11:7425–7438. doi: 10.7150/thno.60211. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 578.Tomas R.M.F., Gibson M.I. Covalent cell surface recruitment of chemotherapeutic polymers enhances selectivity and activity. Chem Sci. 2021;12:4557–4569. doi: 10.1039/d0sc06580c. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 579.Wang H., Wang R., Cai K., He H., Liu Y., Yen J., et al. Selective in vivo metabolic cell-labeling-mediated cancer targeting. Nat Chem Biol. 2017;13:415–424. doi: 10.1038/nchembio.2297. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 580.Liu C.G., Wang Y., Liu P., Yao Q.L., Zhou Y.Y., Li C.F., et al. Aptamer-T cell targeted therapy for tumor treatment using sugar metabolism and click chemistry. ACS Chem Biol. 2020;15:1554–1565. doi: 10.1021/acschembio.0c00164. [DOI] [PubMed] [Google Scholar]
- 581.Liu H., Liu Y., Zhang J.T. A new mechanism of drug resistance in breast cancer cells: fatty acid synthase overexpression-mediated palmitate overproduction. Mol Cancer Therapeut. 2008;7:263–270. doi: 10.1158/1535-7163.MCT-07-0445. [DOI] [PubMed] [Google Scholar]
- 582.Wang J.B., Erickson J.W., Fuji R., Ramachandran S., Gao P., Dinavahi R., et al. Targeting mitochondrial glutaminase activity inhibits oncogenic transformation. Cancer Cell. 2010;18:207–219. doi: 10.1016/j.ccr.2010.08.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 583.Monti E., Gariboldi M.B. HIF-1 as a target for cancer chemotherapy, chemosensitization and chemoprevention. Curr Mol Pharmacol. 2011;4:62–77. doi: 10.2174/1874467211104010062. [DOI] [PubMed] [Google Scholar]
- 584.McBrayer S.K., Cheng J.C., Singhal S., Krett N.L., Rosen S.T., Shanmugam M. Multiple myeloma exhibits novel dependence on GLUT4, GLUT8, and GLUT11: implications for glucose transporter-directed therapy. Blood. 2012;119:4686–4697. doi: 10.1182/blood-2011-09-377846. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 585.Yamaguchi R., Janssen E., Perkins G., Ellisman M., Kitada S., Reed J.C. Efficient elimination of cancer cells by deoxyglucose-ABT-263/737 combination therapy. PLoS One. 2011;6 doi: 10.1371/journal.pone.0024102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 586.Ishiguro T., Ishiguro R., Ishiguro M., Iwai S. Co-treatment of dichloroacetate, omeprazole and tamoxifen exhibited synergistically antiproliferative effect on malignant tumors: in vivo experiments and a case report. Hepato-Gastroenterology. 2012;59:994–996. doi: 10.5754/hge10507. [DOI] [PubMed] [Google Scholar]
- 587.Seltzer M.J., Bennett B.D., Joshi A.D., Gao P., Thomas A.G., Ferraris D.V., et al. Inhibition of glutaminase preferentially slows growth of glioma cells with mutant IDH1. Cancer Res. 2010;70:8981–8987. doi: 10.1158/0008-5472.CAN-10-1666. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 588.Puig T., Aguilar H., Cufi S., Oliveras G., Turrado C., Ortega-Gutierrez S., et al. A novel inhibitor of fatty acid synthase shows activity against HER2+ breast cancer xenografts and is active in anti-HER2 drug-resistant cell lines. Breast Cancer Res. 2011;13:R131. doi: 10.1186/bcr3077. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 589.Liu G., Li Y.I., Gao X. Overexpression of microRNA-133b sensitizes non-small cell lung cancer cells to irradiation through the inhibition of glycolysis. Oncol Lett. 2016;11:2903–2908. doi: 10.3892/ol.2016.4316. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 590.Shen H., Hau E., Joshi S., Dilda P.J., McDonald K.L. Sensitization of glioblastoma cells to irradiation by modulating the glucose metabolism. Mol Cancer Therapeut. 2015;14:1794–1804. doi: 10.1158/1535-7163.MCT-15-0247. [DOI] [PubMed] [Google Scholar]
- 591.Leung E., Cairns R.A., Chaudary N., Vellanki R.N., Kalliomaki T., Moriyama E.H., et al. Metabolic targeting of HIF-dependent glycolysis reduces lactate, increases oxygen consumption and enhances response to high-dose single-fraction radiotherapy in hypoxic solid tumors. BMC Cancer. 2017;17:418. doi: 10.1186/s12885-017-3402-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 592.Omar H.A., Berman-Booty L., Weng J.R. Energy restriction: stepping stones towards cancer therapy. Future Oncol. 2012;8:1503–1506. doi: 10.2217/fon.12.142. [DOI] [PubMed] [Google Scholar]
- 593.Hursting S.D., Dunlap S.M., Ford N.A., Hursting M.J., Lashinger L.M. Calorie restriction and cancer prevention: a mechanistic perspective. Cancer Metabol. 2013;1:10. doi: 10.1186/2049-3002-1-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 594.Cairns R.A., Harris I.S., Mak T.W. Regulation of cancer cell metabolism. Nat Rev Cancer. 2011;11:85–95. doi: 10.1038/nrc2981. [DOI] [PubMed] [Google Scholar]
- 595.Kornberg M.D. The immunologic Warburg effect: evidence and therapeutic opportunities in autoimmunity. Wiley Interdiscip Rev Syst Biol Med. 2020;12:e1486. doi: 10.1002/wsbm.1486. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 596.Cha Y., Han M.J., Cha H.J., Zoldan J., Burkart A., Jung J.H., et al. Metabolic control of primed human pluripotent stem cell fate and function by the miR-200c–SIRT2 axis. Nat Cell Biol. 2017;19:445–456. doi: 10.1038/ncb3517. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 597.Ferraro E., Germano M., Mollace R., Mollace V., Malara N. HIF-1, the Warburg effect, and macrophage/microglia polarization potential role in COVID-19 pathogenesis. Oxid Med Cell Longev. 2021;2021 doi: 10.1155/2021/8841911. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 598.Kantarci H., Gou Y., Riley B.B. The Warburg effect and lactate signaling augment FGF–MAPK to promote sensory-neural development in the otic vesicle. Elife. 2020;9 doi: 10.7554/eLife.56301. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 599.Chen X.S., Li L.Y., Guan Y.D., Yang J.M., Cheng Y. Anticancer strategies based on the metabolic profile of tumor cells: therapeutic targeting of the Warburg effect. Acta Pharmacol Sin. 2016;37:1013–1019. doi: 10.1038/aps.2016.47. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 600.Yu W., Lin R., He X., Yang X., Zhang H., Hu C., et al. Self-propelled nanomotor reconstructs tumor microenvironment through synergistic hypoxia alleviation and glycolysis inhibition for promoted anti-metastasis. Acta Pharm Sin B. 2021;11:2924–2936. doi: 10.1016/j.apsb.2021.04.006. [DOI] [PMC free article] [PubMed] [Google Scholar]















































































