Next Article in Journal
Systematic Identification of MACC1-Driven Metabolic Networks in Colorectal Cancer
Next Article in Special Issue
Regulation of Long Non-Coding RNAs by Plant Secondary Metabolites: A Novel Anticancer Therapeutic Approach
Previous Article in Journal
Outcome of Stage IV Completely Necrotic Wilms Tumour and Local Stage III Treated According to the SIOP 2001 Protocol
Previous Article in Special Issue
Argonaute Proteins Take Center Stage in Cancers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Involvement of Long Non-Coding RNAs in Glucose Metabolism in Cancer

1
Research Unit of Non-Coding RNAs and Genome Editing, Division of Oncology, Department of Internal Medicine, Comprehensive Cancer Center Graz, Medical University of Graz, 8036 Graz, Austria
2
BioTechMed-Graz, 8010 Graz, Austria
3
Department of Experimental Therapeutics, The University of Texas MD Anderson Cancer Center, Houston, TX 77030, USA
*
Author to whom correspondence should be addressed.
Cancers 2021, 13(5), 977; https://doi.org/10.3390/cancers13050977
Submission received: 27 January 2021 / Revised: 21 February 2021 / Accepted: 23 February 2021 / Published: 26 February 2021
(This article belongs to the Collection The Role of Non-coding RNA in Cancer)

Abstract

:

Simple Summary

Long non-coding RNAs (lncRNAs) are a heterogenous group of transcripts that regulate various cellular functions. They are implicated in all hallmarks of cancer, including metabolic alterations. Through the modulation of expression of oncogenic or tumor-suppressive genes, alteration of various signaling pathways, protein stability, and upregulation of metabolic enzymes, lncRNAs enhance glucose uptake in cancer and, thus, favor cancer progression. These transcripts represent crucial regulators of cancer glucose metabolism and, as such, they are potential clinical biomarkers and therapeutic targets. This review aims to provide an overview of the lncRNAs involved in cancer glucose metabolism and summarizes their underlying molecular mechanisms.

Abstract

The rapid and uncontrolled proliferation of cancer cells is supported by metabolic reprogramming. Altered glucose metabolism supports cancer growth and progression. Compared with normal cells, cancer cells show increased glucose uptake, aerobic glycolysis and lactate production. Byproducts of adjusted glucose metabolism provide additional benefits supporting hallmark capabilities of cancer cells. Long non-coding RNAs (lncRNAs) are a heterogeneous group of transcripts of more than 200 nucleotides in length. They regulate numerous cellular processes, primarily through physical interaction with other molecules. Dysregulated lncRNAs are involved in all hallmarks of cancer including metabolic alterations. They may upregulate metabolic enzymes, modulate the expression of oncogenic or tumor-suppressive genes and disturb metabolic signaling pathways favoring cancer progression. Thus, lncRNAs are not only potential clinical biomarkers for cancer diagnostics and prediction but also possible therapeutic targets. This review summarizes the lncRNAs involved in cancer glucose metabolism and highlights their underlying molecular mechanisms.

1. Introduction

Cancers are characterized by uncontrolled and rapid cellular proliferation. To satisfy the accompanying increased demand for energy, cancer cells have to reprogram their metabolism, especially glucose metabolism [1,2]. Normal cells commonly acquire energy via glycolysis in the cytosol and oxidative phosphorylation in the mitochondria, which produces more ATP per molecule of glucose. When oxygen is deficient, cells primarily rely on glycolysis rather than on oxygen-consuming mitochondrial oxidative phosphorylation [3]. However, cancer cells favor glycolysis even in the presence of oxygen. This phenomenon is known as aerobic glycolysis or the Warburg effect [4]. Despite being known for decades, it was only recently acknowledged as a hallmark of cancer [5]. Cancer cells compensate for the net energy loss of glycolysis by an increased glucose uptake and eventual lactate production. Lactate lowers the surrounding pH, which additionally aids cancer invasion and metastasis, while glycolysis intermediates act as substrates that are used in the synthesis of anabolic components including nucleic acids, fatty acids, membrane phospholipids and proteins [3,6,7]. Compared with oxidative phosphorylation, glycolysis produces smaller amounts of reactive oxygen species, which can induce senescence and apoptosis of cancer cells [8]. Therefore, the Warburg effect provides great benefits to cancer thereby supporting its malignant features.
Long non-coding RNAs (lncRNAs) represent a heterogeneous group of RNA transcripts that are more than 200 nucleotides in length [9,10]. They comprise a large portion of the transcriptome but usually do not translate into proteins. However, lncRNAs share features with messenger RNA (mRNA) such as RNA polymerase-dependent transcription and post-transcriptional RNA processing including capping, splicing and polyadenylation [11,12]. LncRNAs exert important cellular functions through different molecular mechanisms, above all by interacting with various molecules [13]. By means of binding to DNA, RNA and protein, lncRNAs regulate gene expression via controlling the chromatin structure, methylation status, sequestration of miRNA, assembly or disruption of protein complexes including transcription factors and post-translational modifications such as phosphorylation and ubiquitination (Figure 1) [14,15,16]. Based on their genomic localization and connection to protein coding genes, lncRNAs can be classified as intergenic, intronic, enhancer, sense, antisense and bidirectional. Besides their genomic context, lncRNAs can also be categorized based on their functionality into signaling, guide, scaffold or decoy molecules [10,17]. Indeed, their functions are largely dependent on their subcellular localization [18].
Multiple studies have documented an aberrant lncRNA expression in various cancers where they act as oncogenes or tumor suppressors [19,20,21]. A few of these contributions to cancers are induced by lncRNAs via their ability to regulate glucose metabolism. They upregulate metabolic enzymes, disturb metabolic signaling pathways and modulate the expression of oncogenic or tumor-suppressive genes (Figure 2). In addition to potentially being useful as clinical biomarkers for cancer diagnostics and prediction, lncRNAs are also increasingly being considered as potential therapeutic targets [22].
The aim of this review is to provide an overview of the lncRNAs involved in cancer glucose metabolism. Additionally, we summarize the molecular mechanisms by which they exert these effects.

2. Methods

This review article is based on a literature review using the PubMed database. The following terms were used to search for the articles: ‘glucose metabolism’, ‘long-non-coding RNA’, ‘lncRNA’, ‘cancer’. Cancer studies in which the involvement of lncRNA in glucose metabolism was confirmed were used for further evaluation and were incorporated into this review.

3. Results

A total of 63 articles meeting the entry criteria is included in this review article. The main findings are interpreted through two main sections. The first section summarizes the pathophysiological processes of cancer regulated by lncRNAs and their influence on glucose metabolism. The second section summarizes the general findings according to the types of cancers.

3.1. LncRNAs Regulates Pathophysiological Processes in Cancer Via the Influence of Glucose Metabolism

LncRNAs are dysregulated in various pathophysiological processes. A few of these processes include interactions with the tumor microenvironment and drug resistance.
Numerous studies have documented altered expression of various lncRNAs under hypoxic conditions. Hypoxic stress, in turn, increases the rates of glucose consumption and lactate production [23,24]. For example, lncRNA AC020978 is upregulated in non-small cell lung cancer (NSCLC) under glucose deprivation and hypoxia [25]. Similarly, the lncRNA HOXA transcript at the distal tip (HOTTIP) promotes glycolysis under hypoxic conditions via the miR-615-3p/HMGB3 axis in NSCLC cells [26]. In ovarian cancer, gastric carcinoma high expressed transcript 1 (GHET1) interacts with the E3 ubiquitin ligase von Hippel-Lindau (VHL). This consequently prevents the VHL-mediated degradation of HIF1α. Increased levels of HIF1α promote the glucose uptake and lactate production in ovarian cancer cells [27]. In addition, the HIF1α/lncRNA Retinoic Acid Early Transcript 1K (RAET1K)/miR-100-5p axis modulates hypoxia-induced glycolysis in hepatocellular carcinoma (HCC) cells and affects HCC progression [28].
Moreover, extracellular cytokines may increase the expression of lncRNAs. Urothelial carcinoma-associated 1 (UCA1) is upregulated by TGF-β1 in HCC. UCA1 then upregulates metabolic enzymes and increases glucose uptake and lactate production [29]. Interestingly, even microorganisms may modify glucose metabolism in cancer via lncRNAs. Fusobacterium nucleatum, a member of the normal flora in the oral cavity, activates lncRNA ENO1-IT1 transcription that eventually induces glucose metabolism in colorectal cancer (CRC) [30].
Furthermore, drug resistance is one of the biggest challenges in cancer treatment. Understanding the contributing factors is of the utmost importance in achieving better clinical outcomes [31]. Emerging studies correlate many lncRNAs with drug resistance across different types of cancers. UCA1 is upregulated in acute myeloid leukemia (AML) following adriamycin (ADR)-based chemotherapy. It contributes to HIF1α-dependent glycolysis in AML [32]. This lncRNA is also upregulated in radioresistant cervical cancer cell lines and promotes glycolysis. In cytarabine-resistant AML cell lines, HOX antisense intergenic RNA myeloid 1 (HOTAIRM1) increases glucose consumption, lactate production and cellular proliferation [33]. In addition, lncRNA H19 overexpression in ginsenoside 20(S)-Rg3-treated ovarian cancer cells increases glucose consumption, lactate production and pyruvate kinase isozyme M2 (PKM2) expression [34]. Another study proved increased glucose consumption and lactate secretion in cetuximab-resistant CRC cells, suggesting that glucose metabolism might be involved in cetuximab resistance. In the resistant cell lines, LINC00973 was upregulated [35].

3.2. LncRNAs and Glucose Metabolism in Different Cancer Entities

3.2.1. Gastrointestinal Cancers

Gastrointestinal cancers rank among the most commonly diagnosed cancers worldwide. In 2018, almost six million new cases were reported and around four million people died [36]. Moreover, these cancers are among the most commonly studied for the roles of lncRNAs [37,38,39]. The gastrointestinal tract comprises the organs that mediate digestion, suggesting that altered glucose metabolism may be an important factor in the development and progression of cancers of this system [40]. See Table 1 for a summary of lncRNA involvement in glucose metabolism in gastrointestinal cancers.

Colorectal Cancer

A common lncRNA-induced disarragement of glucose metabolism in colorectal cancer (CRC) is mediated via the c-MYC oncogene. Many genes that are important for cell growth and cancer progression are targeted by c-MYC including those that regulate metabolism [41]. Glycolysis-associated lncRNA of colorectal cancer (GLCC1), also referred to as AF339830, is an oncogenic lncRNA in CRC. It is associated with poor overall survival (OS) of CRC patients (p = 0.0023). A higher GLCC1 expression leads to increased celllular proliferation and an upregulated glycolysis pathway while experimental downregulation results in a significant reduction of cellular proliferation and tumor growth in mouse models. In addition, GLCC1 stabilizes c-MYC and prevents its degradation by combining it with a heat shock protein 90 (HSP90) chaperon. This further increases the transcriptional level of lactate dehydrogenase A (LDHA), which then activates glycolytic metabolism. It has also been found that glucose starvation significantly increases GLCC1 expression, suggesting that GLCC1 may be a glucose starvation-induced oncogenic lncRNA in CRC [42]. The same physicial interaction with c-MYC and resulting effects have been observed for LINC00504 [43]. Similarly, oncogenic long intergenic non-coding RNA for IGF2BP2 Stability (LINRIS) is important for glycolysis maintainance in CRC. It interacts with insulin-like growth factor 2 mRNA-binding protein 2 (IGF2BP2) and blocks its degradation through the ubiquitination-autophagic pathway. In turn, c-MYC, a downstream target of IGF2BP2, is stabilized. Consequently, the expression of MYC-related metabolic enzymes glucose trasporter 1 (GLUT-1), pyruvate kinase isozyme M2 (PKM2) and LDHA is increased [44].
In contrast, tumor-suppressive lncRNA RAD51 Antisense RNA1 (RAD51-AS1) regulates CRC progression by acting as a competing endogenous RNA (ceRNA) for miR-29b-3p and miR-29c-3p. As a consequence of their upregulation, their common target N-myc downstream-regulated gene 2 (NDRG2) is upregulated. Eventually, this leads to increased levels of hexokinase 2 (HK2) and GLUT-1 as well as elevated glucose consumption and lactate production. Interestingly, RAD51-AS1 expression is not significantly associated with a worse OS of patients (logrank p = 0.35). However, decreased NDRG2 expression is associated with a poor OS (logrank p = 0.043) [45]. A newly identified oncogenic lncRNA in renal cancer [46,47], lncRNA Activated in RCC with Sunitinib Resistance (lncARSR), is also highly expressed in CRC tissues and predicts poor disease-free survival (DFS) (hazard ratio (HR), 2.484; 95% CI, 1.132–6.232; p = 0.028) and OS (HR, 3.250; 95% CI, 1.596–6.716; p = 0.001). In the molecular context, lncARSR sponges miR-34a-5p and, consequently, enhances hexokinase 1 (HK1)-mediated aerobic glycolysis in vitro and in vivo [48].
Wu and colleagues [35] investigated the underlying differences between cetuximab-resistant and cetuximab-sensitive CRC cells. They proved an increased glucose consumption and lactate secretion in cetuximab-resistant cells, suggesting that glucose metabolism might be involved in cetuximab resistance. In addition, lncRNA LINC00973 was upregulated in the resistant cell lines. Upon siRNA-mediated LINC00973 downregulation, the resistant cells exhibited increased apoptosis, reduced cell viability and decreased glucose consumption and lactate production [35].
Huang et al. [49] confirmed the upregulation of lncRNA FEZF1 antisense RNA 1 (FEZF1-AS1) in two expanded CRC cohorts. They revealed that FEZF1-AS1 expression is associated with CRC cell proliferation, metastasis and a poor OS (logrank p = 0.004) and DFS (logrank p = 0.019). Moreover, they found that FEZF1-AS1 binds to cytoplasmic and nuclear PKM2 protein thereby increasing its stability and promoting its activity. Cytoplasmic PKM2 promotes pyruvate kinase activity and lactate production (aerobic glycolysis), whereas nuclear PKM2 acts as a protein kinase and activates STAT3 signaling. A multivariate analysis showed that FEZF1-AS1 expression is an independent risk factor for the survival of patients with CRC (HR, 2.240; 95% CI, 1.028–4.878; p = 0.042) [49].
A recent study found that microbiota might interact with lncRNAs to modulate glucose metabolism in CRC. Specifically, Fusobacterium nucleatum (F. nucleatum) induces glucose metabolism and colorectal carcinogenesis via lncRNA enolase1-intronic transcript 1 (ENO1-IT1). F. nucleatum activates lncRNA ENO1-IT1 transcription by increasing the binding efficiency of the transcription factor specificity protein 1 (SP1) to the promoter region of ENO1-IT1. Upregulated ENO1-IT1 mediates the histone modification of several genes including the alpha-enolase glycolytic enzyme ENO1 via KAT7 histone acetyltransferase [30].
LncRNA colorectal neoplasia differentially expressed (CRNDE) is upregulated in CRC. Treatment with insulin and insulin-like growth factors (IGF) represses CRNDE nuclear transcripts. However, inhibitors against the PI3K/Akt/mTOR or Raf/MAPK pathway have opposite effects, suggesting that CRNDE is a downstream target of these signaling pathways. Additionally, CRNDE promotes metabolic changes in CRC mainly by upregulating the genes responsible for glucose-6-phosphate anabolic reactions (GYS1, G6PC3, ISYNA1 and G6PD). CRNDE also positively regulates GLUT-4 and carbohydrate-responsive element-binding protein (MLXIPL) levels, leading to an increased glucose uptake and lactate production [50].

Hepatocellular Carcinoma

Liver-specific cytoplasmic long intergenic non-coding RNA LINC01554 is downregulated in hepatocellular carcinoma (HCC). The downregulation of this tumor-suppressive lncRNA is associated with adjacent organ invasion, tumor size, advanced tumor stage and an overall poor outcome in patients with HCC [65]. LINC01554 promotes proteasomal degradation of PKM2, an important enzyme at the late stage of glycolysis, and inhibits the Akt/mTOR signaling pathway, one of the key signalling pathways that mediates cellular biosynthesis and aerobic glycolysis in cancer cells. Mechanistically, LINC01554 expression is inhibited by miR-365a-3p at the transcriptional level [51]. The effect of lncRNA metastasis-associated lung adenocarcinoma transcript 1 (MALAT1) on glucose metabolism is essential for its oncogenic activity in HCC. It upregulates the metabolic transcription factor TCF7L2 via the mTORC1 pathway, which then enhances the expression of glycolytic genes and downregulates enzymes involved in gluconeogenesis. In HCC, gluconeogenesis inhibits aerobic glycolysis and has a tumor-suppressive role. MALAT1 overexpression results in a higher glucose uptake and increased lactate production [52]. HOX transcript antisense RNA (HOTAIR) is another upregulated lncRNA in HCC that activates the mTOR signaling pathway. It promotes glycolysis and lactate production by upregulating and directly binding to GLUT-1 [55]. Additionally, HOTAIR might act as a decoy of miR-130a-3p, which normally targets HIF1α [56].
Wang and colleagues [28] demonstrated that the HIF1α/lncRNA Retinoic Acid Early Transcript 1K (RAET1K)/miR-100-5p axis modulates hypoxia-induced glycolysis in HCC cells and affects HCC progression. The expression of lncRNA RAET1K, HIF1α and LDHA is upregulated in HCC tissue specimens while the miR-100-5p expression level is decreased. Mechanistically, RAET1K downregulates the expression of miR-100-5p by acting as a sponge while HIF1α binds the promoter region of the RAET1K and activate its transcription. RAET1K silencing significantly suppressed HCC cell proliferation and invasion. It also suppressed the hypoxia-induced increase in lactate concentration and glucose uptake while miR-100-5p inhibition reversed the effects of RAET1K silencing on hypoxia-induced glycolysis in HCC cells [28]. Upregulated lncRNA five prime to Xist (Ftx) is associated with aggressive clinicopathological features of HCC. In addition, Ftx enhances the activity and expression of key enzymes in glucose metabolism GLUT-1 and GLUT-4. Mechanistically, Ftx directly regulates the transcriptional and post-transcriptional expression of PPARγ. The PPARγ pathway is responsible for the Ftx-induced aerobic glycolysis in HCC [53].
It has been shown that TGF-β1 upregulates the expression of lncRNA urothelial carcinoma-associated 1 (UCA1) in HCC, which in turn significantly increases the expression level of HK2. Consequently, lactate production, glucose uptake and ATP production are increased while the OS is decreased (p = 0.0152) [29]. LncRNA SOX2OT promotes glucose metabolism by PKM2, HK2 and LDHA activation and increases the metastatic potential of HCC. SOX2OT physically binds to miR-122-5p, which also has a PKM2 targeting site and contributes to the PKM2 induction [58]. LncRNA DEAD/H box protein 11 antisense RNA 1 (DDX11-AS1) directly binds to tumor-suppressive miR-195-5p, thus allowing MACC1, a key regulator of the hepatocyte growth factor-HGF receptor (HGFR) pathway, to be overexpressed in HCC. As a result of DDX11-AS1 overexpression, glucose consumption, lactate production, cellular proliferation, invasion and migration are increased [57].
Yu and colleagues [54] investigated the correlation between HCC and blood glucose dysregulation. HCC cells treated with a high glucose concentration were shown to have an increased expression of miR-483-3p and a decreased expression of lncRNA maternally expressed gene 3 (MEG3). Both of them led to suppressed endoplasmic reticulum protein 29 (ERp29) expression, which resulted in an increased HCC cell proliferation and migration. They found that MEG3 mediates protective effects via binding to miR-483-3p. When MEG3 is downregulated, overexpressed miR-483-3p suppresses ERp29 expression. Normally, ERp29 decreases cell proliferation and prevents epithelial-mesenchymal transition [54].

Gastric Cancer

Xia et al. [59] investigated the role of lncRNA DLX6 antisense RNA 1 (DLX6-AS1) in gastric cancer. DLX6-AS1 exhibits oncogenic effects on multiple cancers including gastric cancer [66,67,68,69,70,71,72,73]. A higher DLX6-AS1 expression is positively correlated with tumor size, lymph node involvement and tumor-node-metastasis (TNM) staging. In the molecular context, this lncRNA downregulates miR-4290, which targets 3-phosphoinositide-dependent protein kinase 1 (PDK1). PDK1 upregulation significantly increases intracellular levels of glucose uptake, ATP and lactate production. Aerobic glycolysis was efficiently elevated by PDK1 overexpression while the mitochondrial respiration was inhibited [59]. In addition, lncRNA DiGeorge syndrome critical region gene 9 (DGCR9) has an increased expression in gastric cancer. DGCR9 increases cellular proliferation, migration and glucose uptake. DGCR9 is positively associated with lymph node invasion and the TNM stage in gastric cancer patients [60].

Pancreatic Cancer

LncRNA HOTAIR is overexpressed in patients with pancreatic adenocarcinoma. HOTAIR alters cancer cell energy metabolism in pancreatic adenocarcinoma by the upregulation of HK2. This leads to the enhancement of tumor cell proliferation, glucose uptake, increased ATP and lactate production [61].

Esophageal Cancer

Zhao et al. [62] investigated the function of long intergenic non-protein coding RNA 184 (LINC00184) on glycolysis and mitochondrial oxidative phosphorylation of esophageal cancer cells. They found that highly expressed LINC00184 increases esophageal cancer cell proliferation, migration, invasion and colony formation. LINC00184 recruits DNA methyltransferase 1 (DNMT1) to tumor-suppressive phosphatase and the tensin homolog (PTEN) promotor region. Silenced PTEN, via Akt phosphorylation, leads to increased glycolysis and decreased mitochondrial oxidative phosphorylation [62]. Another lncRNA that was found to play a role in esophageal cancer is the lncRNA UCA1. UCA1 mediates oncogenic effects by regulating glucose metabolism in esophageal cancer. It significantly suppresses the degradation of HK2 by sponging miR-203. Upregulated UCA1 increases the glucose uptake, lactate output and extracellular acidification rate (ECAR) value [63].

Gallbladder Cancer

Sun and colleagues [64] reported that lncRNA plasmacytoma variant translocation 1 (PVT1) acts as an oncogenic lncRNA in gallbladder cancer (GBC) by sponging and repressing miR-143, allowing HK2 to be overexpressed. The upregulation of HK2 is responsible for the higher cellular glucose consumption and lactate production, which gives GBC cells an edge for proliferation, migration and invasion [64]. In addition, PVT1 has been found to be overexpressed in multiple other cancers including CRC, non-small cell lung cancer (NSCLC) and pancreatic cancer [74,75,76]. However, its implications in glucose metabolism disruption in these cancer entities has yet to be investigated.

3.2.2. Genitourinary Cancers

Genitourinary (GU) cancers comprise of a group of cancers in urinary and male reproductive systems [77]. They affect a large number of patients each year with prostate cancer being the most commonly diagnosed cancer and the second leading cause of death in males [78]. Table 2 provides and overview of lncRNA involvement in glucose metabolism in genitourinary cancers.

Kidney Cancer

Kidney-specific lncRNA FoxO-induced long non-coding RNA 1 (FILNC1) is downregulated in renal cell carcinoma correlating with poor clinical outcomes. Upon glucose starvation, the transcription factor FoxO upregulates FILNC1 expression. In turn, FILNC1 interacts with c-MYC mRNA-binding protein (AUF1), which prevents the interaction of AUF1 with c-MYC mRNA. This leads to the downregulation of c-MYC protein levels. However, in renal cancer, this pathway is altered. FILNC1 is downregulated leading to increased c-MYC protein levels thereby increasing the glucose uptake, lactate production and tumor development [79].

Bladder Cancer

Duan et al. [80] observed that the lncRNA cancer susceptibility candidate 8 (CASC8) is significantly downregulated in bladder cancer and established its association with an advanced stage of bladder cancer. Normally, CASC8 binds to the fibroblast growth factor receptor 1 (FGFR1), which in turn prevents FGFR1-mediated LDHA phosphorylation. Consequently, the conversion of pyruvate into lactate is attenuated, glycolysis is reduced and bladder cancer cell proliferation is slowed down [80]. As already observed in HCC and esophageal cancers, lncRNA UCA1 also promotes glycolysis in bladder cancer cells. It regulates glycolysis by upregulating HK2 expression through mTOR and its downstream effector STAT3, which is a direct transcriptional activator for HK2 [81]. An additional mechanism of positive HK2 regulation is mediated through miR-143 downregulation by the active mTOR [82].

Prostate Cancer

Prostate cancer gene expression marker 1 (PCGEM1), a lncRNA highly associated with prostate tumors, significantly induces genes involved in multiple metabolic pathways such as aerobic glycolysis, pentose phosphate shunt and NADPH generation used for redox homeostasis. Mechanistically, PCGEM1 binds directly to the target promoter of c-MYC where it functions as a coactivator and mediates enhanced c-MYC transcription [83].

3.2.3. Female Reproductive System Cancers

Ovarian Cancer

Sood et al. [84] investigated the role of lncRNA ceruloplasmin (NRCP) in ovarian cancer metabolism. They found that NRCP is highly upregulated in ovarian tumors while its experimental knockdown leads to a decreased glycolysis and an increased mitochondrial respiration in cancer cells (Table 3). This, in turn, results in significantly increased cancer cell apoptosis and decreased cell proliferation. Mechanistically, NRCP acts as an intermediate binder between STAT1 and RNA polymerase II thereby increasing their activities. This leads to an elevated expression of the downstream target genes glucose-6-phosphate isomerase (GPI), aldolase A (ALDOA) and aldolase C (ALDOC) [84]. Furthermore, lncRNA gastric carcinoma high expressed transcript 1 (GHET1) is overexpressed in ovarian cancers and it is positively correlated with the tumor size and metastasis in patients. The overexpression of GEHT1 promotes proliferation and colony formation of ovarian cancer cells. GHET1 interacts with the E3 ubiquitin ligase von Hippel-Lindau (VHL), which consequently prevents the VHL-mediated degradation of HIF1α. Increased levels of HIF1α promote the glucose uptake and lactate production of ovarian cancer cells [27]. The important roles of lncRNAs in the regulation of HIF1α in cancer have been demonstrated previously [23]. It was observed that lncRNA H19 overexpression in ginsenoside 20(S)-Rg3-treated ovarian cancer cells enhances glucose consumption, lactate production and PKM2 expression. H19 directly sponges miR-324-5p to enhance the Warburg effect. Unaffected miR-324-5p inhibits PKM2 and negatively regulates glucose uptake and lactate production in ovarian cancer cells and inhibits cell proliferation [34].

Cervical Cancer

Yao and colleagues [85] showed that lncRNA UCA1 has increased expression levels in radioresistant cervical cancer cell lines and promotes glycolysis. In addition, they found that this mechanism is regulated by the HK2 enzyme. These findings suggest that UCA1 plays an important role in the regulation of radioresistance in cervical cancer via the HK2/glycolytic pathway [85].

3.2.4. Breast Cancer

Lin and colleagues [86] demonstrated that lncRNA LINC00538 (YIYA) promotes cancer glycolysis, cell proliferation and tumor growth in breast cancer (Table 4).
Mechanistically, YIYA associates with the cytosolic cyclin-dependent kinase CDK6. It regulates the CDK6-dependent phosphorylation of the fructose bisphosphatase PFK2 (PFKFB3) in a cell cycle-independent manner and the catalysis of glucose 6-phosphate to fructose-2,6-bisphosphate/fructose-1,6-bisphosphate. Both lncRNA YIYA and CDK6 are required to maintain enhanced glycolysis in breast cancer [86]. Another study proved that upregulated lncRNA breast cancer anti-estrogen resistance 4 (BCAR4) in triple negative breast cancer is transcriptionally targeted by Yes-associated protein (YAP) and it is required for YAP-promoted glycolysis through GLI2-dependent Hedgehog signaling. YAP is normally inhibited by the tumor-suppressive Hippo pathway, which is commonly disregulated in many cancers [87]. BCAR4 associates with the Hedgehog signalling effector GLI2 and promotes the transcription of glycolytic activators HK2 and PFKFB3. The individual or simulanteous upregulation of BCAR4 and GLI2 results in increased glucose uptake and lactate production. The overexpression of YAP (p = 0.0166) and BCAR4 (p = 0.0173) correlates with a poor recurrence-free survival (RFS) for breast cancer patients [88]. LINC00346 is also overexpressed in breast cancer cells. Its experimental knockdown results in reduced cell proliferation and glycolysis as well as induced apoptosis through the upregulation of miR-148a/b, which is a direct target of LINC00346. The lower glycolysis rate is due to a repressed GLUT-1 expression [89].

3.2.5. Hematologic Cancer

Wang et al. [90] found that lncRNA antisense non-coding RNA at the INK4 locus (ANRIL) is overexpressed in acute myeloid leukemia (AML) patients and promotes disease development and progression via the modulation of glucose metabolism. This lncRNA is commonly involved in multiple solid tumors [91] (Table 5). The authors found that ANRIL increases AML cell survival via the AdipoR1 pathway, which is an important regulator in cell senescence and metabolism including the Warburg effect. ANRIL and AdipoR1 upregulation results in the increased expression of GLUT-1, which facilitates the transportation of glucose across the plasma membrane, and LDHA, the enzyme that converts pyruvate to lactate. In addition, AdipoR1 upregulates AMPK and SIRT1, the main regulators of cell senescence and cell metabolism. Thus, they are important targets in AML treatment [90]. Moreover, lncRNA in non-homologous end joining (NHEJ) pathway 1 (LINP1) is significantly overexpressed in AML patients. It promotes the malignant phenotype of AML cells and stimulates glucose metabolism [92]. Previous studies have confirmed that LINP1 functions as a scaffold for interaction with several important regulators in NHEJ-mediated DNA repair [93,94]. LINP1 upregulates the expression levels of GLUT-1 and LDHA by upregulating HNF4α, which in turn activates the AMPK/WNT5A signaling pathway. The knockdown of LINP1 expression remarkably suppresses the glucose uptake and AML cell maintenance [92]. Oncogenic lncRNA UCA1 is also upregulated in AML following adriamycin (ADR)-based chemotherapy. Mechanistically, UCA1 acts as a ceRNA of miR-125a and directly binds to it. Consequently, HK2, a target of miR-125a, is upregulated while the miR-125-mediated inhibition of a HIF1α-dependent glycolysis is blocked [32]. HOX antisense intergenic RNA myeloid 1 (HOTAIRM1) is the lncRNA that is highly specific for maturing myeloid cells and is well-known for its implication in the development of AML [95,96]. In cytarabine-resistant AML cell lines, HOTAIRM1 enhances cell viability, reduces apoptosis and enhances glucose consumption and lactate production. These effects are mediated via the activated Wnt/β-catenin pathway [33].
Wang et al. [97] found that lncRNA protein disulfide isomerase family A member 3 pseudogene 1 (PDIA3P) is highly expressed in multiple myeloma and correlates with the OS of patients. They found that PDIA3P regulates multiple myeloma growth and drug resistance through glucose 6-phosphate dehydrogenase (G6PD) and the pentose phosphate pathway (PPP). Specifically, PDIA3P interacts with c-MYC thereby enhancing its transactivation and binding to the G6PD promoter, thus stimulating G6PD expression and PPP flux [97]. Additionally, lncRNA MALAT1 is upregulated in multiple myeloma serums and cells. It is responsible for increased cell viability, invasion and glycolysis. MALAT1 sponges miR-1271-5p, inhibiting its functions. In turn, SRY-Box 13 (SOX13), a miR-1271-5p target, is upregulated [98]. SOX13 is a member of the SRY-related high-mobility group box (Sox) transcription factor family and is known to be involved in the progression of several tumors including glioma [99].

3.2.6. Brain Cancers

Guo and colleagues [100] observed that the lncRNA X-inactive specific transcript (XIST) is upregulated in glioblastoma cell lines and tissue samples (Table 6). Knockdown experiments highlighted the importance of this lncRNA in glioblastoma cell viability, invasion, resistance to apoptosis and enhanced glucose metabolism. XIST acts as a ceRNA for miR-126. By sponging miR-126, insulin receptor substrate 1 (IRS1) is upregulated along with the PI3K/Akt pathway. In addition, protein levels of GLUT-1 and GLUT-3 receptors as well as cellular glucose uptake and lactate production are elevated [100]. This lncRNA is also involved in several other cancer entities including bladder cancer, cervical cancer and pancreatic cancer, always exerting its functions via sponging various miRNAs and thereby promoting cancer progression [101,102,103]. However, a potential mechanism of XIST in glucose metabolism disruption in these cancers is yet to be investigated.
Furthermore, overexpressed LINC00174 leads to an increased cell proliferation, tumor growth, migration, invasion and glycolysis of glioma cells. This lncRNA interacts with the miR-152-3p/SLC2A1 axis. LINC00174 accelerates glioma carcinogenesis via sponging miR-152-3p. Consequently, SLC2A1 (GLUT-1), a downstream target of miR-152-3p, is upregulated. SLC2A1 is a crucial transporter for glucose uptake and plays an important role in glycolysis [104]. LncRNA small nucleolar RNA host gene 9 (SNHG9) is also elevated in glioblastoma tissues and it is associated with a lower OS (p = 0.0106) and PFS (p = 0.0014) in patients with glioma. The overexpression of SNHG9 increases the cell proliferation and aerobic glycolysis. Mechanistically, SNHG9 downregulates miR-199a-5p, leading to the upregulation of the Wnt2 axis in glioblastoma cells [105]. The upregulation of lncRNA SNHG5 in glioma positively correlates with glucose uptake, migration and invasion of glioma cells. By acting as a sponge for miR-205, which targets E2F transcription factor 3 (E2F3), SNHG5 enhances glioma growth [106]. Similarly, overexpressed lncRNA SNHG1 promotes glioma progression in vitro and in vivo. SNHG1 regulates pleckstrin homology like domain family A member 1 (PHLDA1) expression by sponging miR-194, leading to an increased glioma cell glucose uptake, proliferation, migration, invasion, angiogenesis and in vivo tumor growth [107].
The expression level of lncRNA TP53 target 1 (TP53TG1) is significantly higher in human glioma tissues or cell lines compared with normal brain tissue. TP53TG1 is shown to be important in cellular proliferation and migration. It also promotes the upregulation of several glucose metabolism related genes such as glucose-regulated protein 78 (GRP78), isocitrate dehydrogenase 1 (IDH1) and LDHA. Interestingly, these effects are more striking in situations of glucose deprivation than in high glucose conditions [108]. Similarly, increased levels of Metastasis Associated in Colon Cancer 1 Antisense RNA 1 (MACC1-AS1), particularly under glucose deprivation, result in increased cell viability and inhibited apoptosis. High levels of MACC1-AS1 in glioma tissues and cells are associated with a poor prognosis of glioma patients (p = 0.0106). MACC1-AS1 overexpression significantly increases the levels of GLUT-1, HK2, G6PD, MCT1, ATP, lactate and NAPDH. These effects of MACC1-AS1 overexpression were found to be mediated via the AMPK pathway [109].

3.2.7. Osteosarcoma

Yang and colleagues [110] investigated the oncogenic mechanisms of lncRNA taurine-upregulated gene 1 (TUG1) in osteosarcoma (Table 7). They reported inhibited glucose consumption, lactate production and cell viability of osteosarcoma cells upon TUG1 knockdown. However, an increased expression of TUG1 markedly increased the expression of HK2 [110]. Similarly, upregulated lncRNA Potassium Voltage-Gated Channel Subfamily Q Member 1 Overlapping Transcript 1 (KCNQ1OT1) facilitates proliferation and decreases the apoptosis of osteosarcoma cells. KCNQ1OT1 contributes to the Warburg effect by stimulating ALDOA expression. KCNQ1OT1 sponges miR-34c-5p, which directly targets and inhibits ALDOA expression [111]. Similar to its effect in GBC, PVT1 upregulation has been observed in osteosarcoma where it is associated with a poor survival of patients (p = 0.011). Glucose uptake, lactate production and the expression of HK2 are all increased in osteosarcoma cells with an overexpression of PVT1 and decreased by PVT1 knockdown [112].
In contrast, lncRNA heart and neural crest derivatives expressed 2-antisense RNA 1 (HAND2-AS1) has a decreased expression in osteosarcoma compared with normal tissue and functions as a tumor suppressor. The disruption of lncRNA HAND2-AS1 results in osteosarcoma cell proliferation through the alteration of glucose metabolism. It leads to an increased expression of GLUT-1, consequently enhancing glucose uptake and the proliferation of osteosarcoma cell lines. Interestingly, HAND2-AS1 is not associated with metastasis formation [113]. Another study has shown that HAND2-AS1 regulates osteosarcoma metabolism via sequestering fructose-1,6-bisphosphatase 1 (FBP1) from binding to HIF1α thereby increasing the HIF1α expression [114]. Additionally, a tumor-suppressive role of this lncRNA has recently been identified in endometrioid carcinoma that acts via different mechanisms [115]. The potential involvement in glucose metabolism alteration should be investigated.

3.2.8. Melanoma

H19 acts as an oncogenic lncRNA in many human cancers [116,117]. Xu et al. [118] identified its molecular mechanisms in melanoma. They found that H19 acts as a sponge for miR-106a-5p, leading to the upregulation of E2F3 transcription factor expression in melanoma cells. E2F3 promotes melanoma cell glucose metabolism and growth [118].

3.2.9. Lung Cancer

Huang and colleagues [25] proved that lncRNA AC020978, which is upregulated in non-small cell lung cancers (NSCLC) (Table 8), has a significant correlation with advanced stage and poor clinical outcomes. AC020978 promotes cell growth and induces metabolic reprogramming. It is upregulated under glucose deprivation and hypoxia. Mechanistically, AC020978 directly interacts with PKM2 and enhances the stability of this protein, which is crucial for the Warburg effect. In addition, AC020978 promotes the nuclear translocation of PKM2. It thereby acts as a protein kinase and interacts with many tumor-associated genes to accelerate carcinogenesis including HIF1α, Oct-4, STAT3 and CTNNB1 that encode β-catenin [25]. Likewise, the lncRNA HOXA transcript at the distal tip (HOTTIP) promotes glycolysis under hypoxic conditions via the miR-615-3p/HMGB3 axis in NSCLC cells. Hypoxic stress increases the rates of glucose consumption and lactate production. Glycolysis rate is increased by the substantial increase of HK2 enzyme expression compared with control cells grown in a normoxic environment. In addition, hypoxia increases HOTTIP and high mobility group box 3 (HMGB3) expression while it suppresses miR-615-3p expression. Mechanistically, HOTTIP acts as a molecular sponge for miR-615-3p leading to an increased expression of its target HMGB3 [26]. HMGB3 is a well-known oncogenic member of chromatin-binding proteins that can regulate gene expression in various cancers via transcription factor binding modulation [119,120,121]. Moreover, lncRNA MALAT1 is upregulated in NSCLC tissues and cells. It is responsible for enhanced cell proliferation, colony formation, reduced cell apoptosis in vitro and increased tumor growth in vivo. MALAT1 knockdown decreases glucose consumption and lactate production. It mediates these effects by binding to miR-613 thereby inhibiting its interaction with copper metabolism MURR1 domain-containing 8 (COMMD8) [122]. Research on COMMD8 in cancers is scarce. However, in addition to this research, one more study has shown that the overexpression of COMMD8 mitigates malignant NSCLC features [123]. Another oncogenic lncRNA in NSCLC is brain cytoplasmic RNA 1 (BCYRN1). BCYRN1 induces glycolysis and upregulates the expression levels of PKM2 in NSCLC cells. These effects are mediated by reducing the miR‑149 expression levels [124].
LINC00551 has a low expression in lung adenocarcinoma and correlates with a worse OS of lung cancer patients (HR = 0.67, logrank p = 0.0014). It regulates the glycolytic pathway by suppressing glucose uptake and lactate production. Mechanistically, LINC00551 inhibits glycolysis by impairing the c-MYC-mediated transcription of glycolytic enzymes particularly PKM2 [125].

3.2.10. Head and Neck Cancer

LncRNA ANRIL is also highly expressed in patients with nasopharyngeal carcinoma (Table 9). Overexpression is associated with advanced stage cancer, a shorter OS (p = 0.027), DFS (p = 0.033), frequent recurrences, increased cell proliferation and transformation via a stem-like cancer cell induction. ANRIL reprograms glucose metabolism to meet the cancer’s increased needs for energy by increasing the glucose uptake. It does that by affecting the mTOR signal pathway, which in turn increases the expression of GLUT-1 and LDHA, essential genes in glucose metabolism [126].
Cheng and colleagues [127] identified a novel lncRNA lnc-p23154 associated with oral squamous cell carcinoma (OSCC). The upregulation of lnc-p23154 correlates with OSCC cell migration and invasion in vitro and in vivo. Additionally, they found that lnc-p23154 is involved in OSCC glycolysis where it upregulates GLUT-1 expression. Lnc-p23154 is primarily located in the nucleus and binds to the promoter region of miR-378a-3p, which represses GLUT-1 expression. Therefore, lnc-p23154 plays an important role in GLUT-1-mediated glycolysis by inhibiting miR-378a-3p transcription and accelerating OSCC metastasis [127]. Moreover, lncRNA UCA1 is upregulated in tongue squamous cell carcinoma (TSCC) and promotes proliferation, migration, invasion and glycolysis metabolism of TSCC cells. UCA1 was suggested to sponge miR-138-5p whose downstream target is CCR7 [128].

4. Conclusions

To maintain the characteristics of uncontrolled and accelerated proliferation, cancer cells reprogram their metabolism. This provides cancer with numerous benefits enabling its growth and progression. As glucose is the main source of energy, it is excessively metabolized by cancer cells leading to increased lactate production. Byproducts of altered glucose metabolism are substrates essential for the building of new cellular components and they lead to a lower surrounding pH thereby aiding metastasis formation. Metabolic disarrangement is associated with poor outcomes in certain cancers. A few of the key players of glucose metabolism such as LDAH and PKM2 could be used in patient stratifications and the prediction of the disease outcomes.
LncRNAs are a heterogenous group of transcripts capable of mediating various cellular functions including the regulation of glucose metabolism. Different molecular mechanisms of lncRNAs have been described. Among the most common is the regulation of the expression of metabolic enzymes such as GLUTs, LDAH, PKM2 and HK2. Namely, ANRIL, UCA1, HOTAIR and H19 are a few of the most familiar lncRNAs that target metabolic enzymes. In addition, lncRNAs may sponge to different miRNA and mRNA thus modulating the expression of numerous oncogenic or tumor-suppressive genes. Proteins such as c-MYC and HIF1α are frequently stabilized and/or upregulated whereas, in the same way, lncRNAs alter signaling pathways favoring cancer progression. For instance, mTOR/STAT3 is commonly dysregulated in various cancers.
In conclusion, this review highlights lncRNAs as crucial regulators of cancer glucose metabolism. A blockade of these mechanisms may alleviate cancer progression, therefore rendering lncRNAs a promising therapeutic target.

Author Contributions

Conceptualization, A.B. and M.P.; writing—original draft preparation, A.B.; writing—review and editing, A.B., D.A.B., F.P. and M.P.; visualization, A.B.; supervision, M.P; funding acquisition, A.B. and M.P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by BioTechMed-Graz (Lab Rotation Program 2020, to A.B.). M.P research was supported by the Austrian Science Fund FWF (DK-MCD, W 1226). Open Access Funding by the Austrian Science Fund (FWF).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Martinez-Outschoorn, U.E.; Peiris-Pagés, M.; Pestell, R.G.; Sotgia, F.; Lisanti, M.P. Cancer metabolism: A therapeutic perspective. Nat. Rev. Clin. Oncol. 2017, 14, 11–31. [Google Scholar] [CrossRef]
  2. DeBerardinis, R.J.; Chandel, N.S. Fundamentals of cancer metabolism. Sci. Adv. 2016, 2, e1600200. [Google Scholar] [CrossRef] [Green Version]
  3. Fadaka, A.; Ajiboye, B.; Ojo, O.; Adewale, O.; Olayide, I.; Emuowhochere, R. Biology of glucose metabolization in cancer cells. J. Oncol. Sci. 2017, 3, 45–51. [Google Scholar] [CrossRef]
  4. Warburg, O. On the Origin of Cancer Cells. Science 1956, 123, 309–314. [Google Scholar] [CrossRef]
  5. Hanahan, D.; Weinberg, R.A. Hallmarks of Cancer: The Next Generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef] [Green Version]
  6. Jiang, B. Aerobic glycolysis and high level of lactate in cancer metabolism and microenvironment. Genes Dis. 2017, 4, 25–27. [Google Scholar] [CrossRef]
  7. Peppicelli, S.; Bianchini, F.; Calorini, L. Extracellular acidity, a “reappreciated” trait of tumor environment driving malignancy: Perspectives in diagnosis and therapy. Cancer Metastasis Rev. 2014, 33, 823–832. [Google Scholar] [CrossRef]
  8. Heiden, M.G.V.; Cantley, L.C.; Thompson, C.B. Understanding the Warburg Effect: The Metabolic Requirements of Cell Proliferation. Science 2009, 324, 1029–1033. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Ling, H.; Vincent, K.; Pichler, M.; Fodde, R.; Berindan-Neagoe, I.; Slack, F.; Calin, G. Junk DNA and the long non-coding RNA twist in cancer genetics. Oncogene 2015, 34, 5003–5011. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Hombach, S.; Kretz, M. Non-coding RNAs: Classification, Biology and Functioning. Adv. Exp. Med. Biol. 2016, 937, 3–17. [Google Scholar] [CrossRef] [PubMed]
  11. Melé, M.; Mattioli, K.; Mallard, W.; Shechner, D.M.; Gerhardinger, C.; Rinn, J.L. Chromatin environment, transcriptional regulation, and splicing distinguish lincRNAs and mRNAs. Genome Res. 2017, 27, 27–37. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Deveson, I.W.; Hardwick, S.A.; Mercer, T.R.; Mattick, J.S. The Dimensions, Dynamics, and Relevance of the Mammalian Noncoding Transcriptome. Trends Genet. 2017, 33, 464–478. [Google Scholar] [CrossRef]
  13. Pichler, M.; Rodriguez-Aguayo, C.; Nam, S.Y.; Dragomir, M.P.; Bayraktar, R.; Anfossi, S.; Knutsen, E.; Ivan, C.; Fuentes-Mattei, E.; Kil Lee, S.; et al. Therapeutic potential of FLANC, a novel primate-specific long non-coding RNA in colorectal cancer. Gut 2020, 69, 1818–1831. [Google Scholar] [CrossRef] [PubMed]
  14. Mercer, T.R.; Dinger, M.E.; Mattick, J.S. Long non-coding RNAs: Insights into functions. Nat. Rev. Genet. 2009, 10, 155–159. [Google Scholar] [CrossRef] [PubMed]
  15. Schmitz, S.U.; Grote, P.; Herrmann, B.G. Mechanisms of long noncoding RNA function in development and disease. Cell. Mol. Life Sci. 2016, 73, 2491–2509. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Kazimierczyk, M.; Kasprowicz, M.K.; Kasprzyk, M.E.; Wrzesinski, J. Human Long Noncoding RNA Interactome: Detection, Characterization and Function. Int. J. Mol. Sci. 2020, 21, 1027. [Google Scholar] [CrossRef] [Green Version]
  17. Laurent, G.S.; Wahlestedt, C.; Kapranov, P. The Landscape of long noncoding RNA classification. Trends Genet. 2015, 31, 239–251. [Google Scholar] [CrossRef] [Green Version]
  18. Carlevaro-Fita, J.; Johnson, R. Global Positioning System: Understanding Long Noncoding RNAs through Subcellular Localization. Mol. Cell 2019, 73, 869–883. [Google Scholar] [CrossRef] [Green Version]
  19. Fu, P.-F.; Zheng, X.; Fan, X.; Lin, A.-F. Role of cytoplasmic lncRNAs in regulating cancer signaling pathways. J. Zhejiang Univ. Sci. B 2019, 20, 1–8. [Google Scholar] [CrossRef]
  20. Guzel, E.; Okyay, T.M.; Yalcinkaya, B.; Karacaoglu, S.; Gocmen, M.; Akcakuyu, M.H. Tumor suppressor and oncogenic role of long non-coding RNAs in cancer. North. Clin. Istanb. 2020, 7, 81. [Google Scholar] [CrossRef]
  21. Barth, D.A.; Juracek, J.; Slaby, O.; Pichler, M.; Calin, G.A. lncRNA and Mechanisms of Drug Resistance in Cancers of the Genitourinary System. Cancers 2020, 12, 2148. [Google Scholar] [CrossRef]
  22. Gupta, S.C.; Tripathi, Y.N. Potential of long non-coding RNAs in cancer patients: From biomarkers to therapeutic targets. Int. J. Cancer 2017, 140, 1955–1967. [Google Scholar] [CrossRef] [PubMed]
  23. Barth, D.A.; Prinz, F.; Teppan, J.; Jonas, K.; Klec, C.; Pichler, M. Long-noncoding RNA (IncRNA) in the regulation of hypoxia-inducible factor (HIF) in cancer. Noncoding RNA 2020, 6, 27. [Google Scholar] [CrossRef]
  24. Eales, K.L.; Hollinshead, K.E.R.; Tennant, D.A. Hypoxia and metabolic adaptation of cancer cells. Oncogenesis 2016, 5, e190. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Hua, Q.; Mi, B.; Xu, F.; Wen, J.; Zhao, L.; Liu, J.; Huang, G. Hypoxia-induced lncRNA-AC020978 promotes proliferation and glycolytic metabolism of non-small cell lung cancer by regulating PKM2/HIF-1α axis. Theranostics 2020, 10, 4762–4778. [Google Scholar] [CrossRef]
  26. Shi, J.; Wang, H.; Feng, W.; Huang, S.; An, J.; Qiu, Y.; Wu, K. Long non-coding RNA HOTTIP promotes hypoxia-induced glycolysis through targeting miR-615-3p/HMGB3 axis in non-small cell lung cancer cells. Eur. J. Pharmacol. 2019, 862, 172615. [Google Scholar] [CrossRef]
  27. Liu, D.; Li, H. Long non-coding RNA GEHT1 promoted the proliferation of ovarian cancer cells via modulating the protein stability of HIF1α. Biosci. Rep. 2019, 39, 20181650. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Zhou, Y.; Huang, Y.; Hu, K.; Zhang, Z.; Yang, J.; Wang, Z. HIF1A activates the transcription of lncRNA RAET1K to modulate hypoxia-induced glycolysis in hepatocellular carcinoma cells via miR-100-5p. Cell Death Dis. 2020, 11, 1–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Hu, M.-L.; Wang, X.-Y.; Chen, W.-M. TGF-β1 upregulates the expression of lncRNA UCA1 and its downstream HXK2 to promote the growth of hepatocellular carcinoma. Eur. Rev. Med. Pharmacol. Sci. 2018, 22, 4846–4854. [Google Scholar]
  30. Hong, J.; Guo, F.; Lu, S.-Y.; Shen, C.; Ma, D.; Zhang, X.; Xie, Y.; Yan, T.; Yu, T.; Sun, T.; et al. F. nucleatum targets lncRNA ENO1-IT1 to promote glycolysis and oncogenesis in colorectal cancer. Gut 2020. [Google Scholar] [CrossRef]
  31. Wang, X.; Zhang, H.; Chen, X. Drug resistance and combating drug resistance in cancer. Cancer Drug Resist. 2019, 2, 141–160. [Google Scholar] [CrossRef] [Green Version]
  32. Zhang, Y.; Liu, Y.; Xu, X. Knockdown of LncRNA-UCA1 suppresses chemoresistance of pediatric AML by inhibiting glycolysis through the microRNA-125a/hexokinase 2 pathway. J. Cell. Biochem. 2018, 119, 6296–6308. [Google Scholar] [CrossRef] [PubMed]
  33. Chen, L.; Hu, N.; Wang, C.; Zhao, H. HOTAIRM1 knockdown enhances cytarabine-induced cytotoxicity by suppression of glycolysis through the Wnt/β-catenin/PFKP pathway in acute myeloid leukemia cells. Arch. Biochem. Biophys. 2020, 680, 108244. [Google Scholar] [CrossRef] [PubMed]
  34. Zheng, X.; Zhou, Y.; Chen, W.; Chen, L.; Lu, J.; He, F.; Li, X.; Zhao, L. 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. [Google Scholar] [CrossRef] [PubMed]
  35. Jing, C.; Ma, R.; Cao, H.; Wang, Z.; Liu, S.; Chen, D.; Wu, Y.; Zhang, J.; Wu, J. Long noncoding RNA and mRNA profiling in cetuximab-resistant colorectal cancer cells by RNA sequencing analysis. Cancer Med. 2019, 8, 1641–1651. [Google Scholar] [CrossRef] [PubMed]
  36. Bray, F.; Ferlay, J.; Soerjomataram, I.; Siegel, R.L.; Torre, L.A.; Jemal, A. Global cancer statistics 2018: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J. Clin. 2018, 68, 394–424. [Google Scholar] [CrossRef] [Green Version]
  37. Smolle, M.; Uranitsch, S.; Gerger, A.; Pichler, M.; Haybaeck, J. Current Status of Long Non-Coding RNAs in Human Cancer with Specific Focus on Colorectal Cancer. Int. J. Mol. Sci. 2014, 15, 13993–14013. [Google Scholar] [CrossRef] [Green Version]
  38. Deng, H.; Wang, J.M.; Li, M.; Tang, R.; Tang, K.; Su, Y.; Hou, Y.; Zhang, J. Long non-coding RNAs: New biomarkers for prognosis and diagnosis of colon cancer. Tumor Biol. 2017, 39, 101042831770633. [Google Scholar] [CrossRef] [Green Version]
  39. Rigoutsos, I.; Kil Lee, S.; Nam, S.Y.; Anfossi, S.; Pasculli, B.; Pichler, M.; Jing, Y.; Rodriguez-Aguayo, C.; Telonis, A.G.; Rossi, S.; et al. N-BLR, a primate-specific non-coding transcript leads to colorectal cancer invasion and migration. Genome Biol. 2017, 18, 1–21. [Google Scholar] [CrossRef] [Green Version]
  40. Sawayama, H.; Ishimoto, T.; Sugihara, H.; Miyanari, N.; Miyamoto, Y.; Baba, Y.; Yoshida, N.; Baba, H. Clinical impact of the Warburg effect in gastrointestinal cancer (Review). Int. J. Oncol. 2014, 45, 1345–1354. [Google Scholar] [CrossRef]
  41. Dang, C.V.; Le, A.; Gao, P. MYC-Induced Cancer Cell Energy Metabolism and Therapeutic Opportunities. Clin. Cancer Res. 2009, 15, 6479–6483. [Google Scholar] [CrossRef] [Green Version]
  42. Tang, J.; Yan, T.; Bao, Y.; Shen, C.; Yu, C.; Zhu, X.; Tian, X.; Guo, F.; Liang, Q.; Liu, Q.; et al. LncRNA GLCC1 promotes colorectal carcinogenesis and glucose metabolism by stabilizing c-Myc. Nat. Commun. 2019, 10, 1–15. [Google Scholar] [CrossRef]
  43. Feng, J.; Ma, J.; Liu, S.; Wang, J.; Chen, Y. A noncoding RNA LINC00504 interacts with c-Myc to regulate tumor metabolism in colon cancer. J. Cell. Biochem. 2019, 120, 14725–14734. [Google Scholar] [CrossRef] [PubMed]
  44. Wang, Y.; Lu, J.-H.; Wu, Q.-N.; Jin, Y.; Wang, D.-S.; Chen, Y.-X.; Liu, J.; Luo, X.-J.; Meng, Q.; Pu, H.-Y.; et al. LncRNA LINRIS stabilizes IGF2BP2 and promotes the aerobic glycolysis in colorectal cancer. Mol. Cancer 2019, 18, 1–18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Li, C.; Wang, P.; Du, J.; Chen, J.; Liu, W.; Ye, K. LncRNA RAD51-AS1 / miR -29b/c-3p/ NDRG2 crosstalk repressed proliferation, invasion and glycolysis of colorectal cancer. IUBMB Life 2021, 73, 286–298. [Google Scholar] [CrossRef] [PubMed]
  46. Qu, L.; Wu, Z.; Li, Y.; Xu, Z.; Liu, B.; Liu, F.; Bao, Y.; Wu, D.; Liu, J.; Wang, A.; et al. A feed-forward loop between lncARSR and YAP activity promotes expansion of renal tumour-initiating cells. Nat. Commun. 2016, 7, 12692. [Google Scholar] [CrossRef]
  47. Qu, L.; Ding, J.; Chen, C.; Wu, Z.; Liu, B.; Gao, Y.; Chen, W.; Liu, F.; Sun, W.; Li, X.-F.; et al. Exosome-transmitted lncARSR promotes sunitinib resistance in renal cancer by acting as a competing endogenous RNA. Cancer Cell 2016, 29, 653–668. [Google Scholar] [CrossRef]
  48. Li, S.; Zhu, K.; Liu, L.; Gu, J.; Niu, H.; Guo, J. lncARSR sponges miR-34a-5p to promote colorectal cancer invasion and metastasis via hexokinase-1-mediated glycolysis. Cancer Sci. 2020, 111, 3938–3952. [Google Scholar] [CrossRef]
  49. Bian, Z.; Zhang, J.; Li, M.; Feng, Y.; Wang, X.; Zhang, J.; Yao, S.; Jin, G.; Du, J.; Han, W.; et al. LncRNA–FEZF1-AS1 Promotes Tumor Proliferation and Metastasis in Colorectal Cancer by Regulating PKM2 Signaling. Clin. Cancer Res. 2018, 24, 4808–4819. [Google Scholar] [CrossRef] [Green Version]
  50. Ellis, B.C.; Graham, L.D.; Molloy, P.L. CRNDE, a long non-coding RNA responsive to insulin/IGF signaling, regulates genes involved in central metabolism. Biochim. Biophys. Acta (BBA) Bioenerg. 2014, 1843, 372–386. [Google Scholar] [CrossRef] [Green Version]
  51. Zheng, Y.-L.; Li, L.; Jia, Y.-X.; Zhang, B.-Z.; Li, J.-C.; Zhu, Y.-H.; Li, M.-Q.; He, J.-Z.; Zeng, T.-T.; Ban, X.-J.; et al. LINC01554-Mediated Glucose Metabolism Reprogramming Suppresses Tumorigenicity in Hepatocellular Carcinoma via Downregulating PKM2 Expression and Inhibiting Akt/mTOR Signaling Pathway. Theranostics 2019, 9, 796–810. [Google Scholar] [CrossRef]
  52. Malakar, P.; Stein, I.; Saragovi, A.; Winkler, R.; Stern-Ginossar, N.; Berger, M.; Pikarsky, E.; Karni, R. Long Noncoding RNA MALAT1 Regulates Cancer Glucose Metabolism by Enhancing mTOR-Mediated Translation of TCF7L2. Cancer Res. 2019, 79, 2480–2493. [Google Scholar] [CrossRef] [Green Version]
  53. Li, X.; Zhao, Q.; Qi, J.; Wang, W.; Zhang, D.; Li, Z.; Qin, C. lncRNA Ftx promotes aerobic glycolysis and tumor progression through the PPARγ pathway in hepatocellular carcinoma. Int. J. Oncol. 2018, 53, 551–566. [Google Scholar] [CrossRef] [PubMed]
  54. Li, X.; Cheng, T.; He, Y.; Zhou, S.; Wang, Y.; Zhang, K.; Yu, P. High glucose regulates ERp29 in hepatocellular carcinoma by LncRNA MEG3-miRNA 483-3p pathway. Life Sci. 2019, 232, 116602. [Google Scholar] [CrossRef]
  55. Wei, S.; Fan, Q.; Yang, L.; Zhang, X.; Ma, Y.; Zong, Z.; Hua, X.; Su, N.; Sun, H.; Li, H.; et al. Promotion of glycolysis by HOTAIR through GLUT1 upregulation via mTOR signaling. Oncol. Rep. 2017, 38, 1902–1908. [Google Scholar] [CrossRef] [Green Version]
  56. Hu, M.; Fu, Q.; Jing, C.; Zhang, X.; Qin, T.; Pan, Y. LncRNA HOTAIR knockdown inhibits glycolysis by regulating miR-130a-3p/HIF1A in hepatocellular carcinoma under hypoxia. Biomed. Pharmacother. 2020, 125, 109703. [Google Scholar] [CrossRef] [PubMed]
  57. Wan, T.; Zheng, J.; Yao, R.; Yang, S.; Zheng, W.; Zhou, P. LncRNA DDX11-AS1 accelerates hepatocellular carcinoma progression via the miR-195-5p/MACC1 pathway. Ann. Hepatol. 2021, 20, 100258. [Google Scholar] [CrossRef] [PubMed]
  58. Liang, Y.; Zhang, D.; Zheng, T.; Yang, G.; Wang, J.; Meng, F.; Liu, Y.; Zhang, G.; Zhang, L.; Han, J.; et al. lncRNA-SOX2OT promotes hepatocellular carcinoma invasion and metastasis through miR-122-5p-mediated activation of PKM2. Oncogenesis 2020, 9, 54. [Google Scholar] [CrossRef]
  59. Qian, Y.; Song, W.; Wu, X.; Hou, G.; Wang, H.; Hang, X.; Xia, T. DLX6 Antisense RNA 1 Modulates Glucose Metabolism and Cell Growth in Gastric Cancer by Targeting microRNA-4290. Dig. Dis. Sci. 2021, 66, 460–473. [Google Scholar] [CrossRef]
  60. Ni, C.; Yang, P.; Guo, J.; Ye, M. Role of DiGeorge syndrome critical region gene 9, a long noncoding RNA, in gastric cancer. OncoTargets Ther. 2018, 11, 2259–2267. [Google Scholar] [CrossRef] [Green Version]
  61. Ma, Y.; Hu, M.; Zhou, L.; Ling, S.; Li, Y.; Kong, B.; Huang, P. Long non‑coding RNA HOTAIR promotes cancer cell energy metabolism in pancreatic adenocarcinoma by upregulating hexokinase‑2. Oncol. Lett. 2019, 18, 2212–2219. [Google Scholar] [CrossRef]
  62. Li, W.; Huang, K.; Wen, F.; Cui, G.; Guo, H.; He, Z.; Zhao, S. LINC00184 silencing inhibits glycolysis and restores mitochondrial oxidative phosphorylation in esophageal cancer through demethylation of PTEN. EBioMedicine 2019, 44, 298–310. [Google Scholar] [CrossRef] [Green Version]
  63. 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. J. Oncol. 2020, 2020, 1–11. [Google Scholar] [CrossRef] [PubMed]
  64. Chen, J.; Yu, Y.; Li, H.; Hu, Q.; Chen, X.; He, Y.; Xue, C.; Ren, F.; Ren, Z.; Li, J.; et al. Long non-coding RNA PVT1 promotes tumor progression by regulating the miR-143/HK2 axis in gallbladder cancer. Mol. Cancer 2019, 18, 1–16. [Google Scholar] [CrossRef] [Green Version]
  65. Ding, Y.; Sun, Z.; Zhang, S.; Chen, Y.; Zhou, B.; Li, G.; Sun, Q.; Zhou, D.; Ge, Y.; Yan, S.; et al. Down-regulation of Long Non-coding RNA LINC01554 in Hepatocellular Cancer and its Clinical Significance. J. Cancer 2020, 11, 3369–3374. [Google Scholar] [CrossRef]
  66. Guo, J.; Chen, Z.; Jiang, H.; Yu, Z.; Peng, J.; Xie, J.; Li, Z.; Wu, W.; Cheng, Z.; Xiao, K. The lncRNA DLX6-AS1 promoted cell proliferation, invasion, migration and epithelial-to-mesenchymal transition in bladder cancer via modulating Wnt/β-catenin signaling pathway. Cancer Cell Int. 2019, 19, 312. [Google Scholar] [CrossRef] [Green Version]
  67. Li, J.; Li, P.; Zhao, W.; Yang, R.; Chen, S.; Bai, Y.; Dun, S.; Chen, X.; Du, Y.; Wang, Y.; et al. Expression of long non-coding RNA DLX6-AS1 in lung adenocarcinoma. Cancer Cell Int. 2015, 15, 1–8. [Google Scholar] [CrossRef] [Green Version]
  68. Li, X.; Zhang, H.; Wu, X. Long noncoding RNA DLX6-AS1 accelerates the glioma carcinogenesis by competing endogenous sponging miR-197-5p to relieve E2F1. Gene 2019, 686, 1–7. [Google Scholar] [CrossRef] [PubMed]
  69. Tian, Y.; Wang, Y.-R.; Jia, S.-H. Knockdown of long noncoding RNA DLX6-AS1 inhibits cell proliferation and invasion of cervical cancer cells by downregulating FUS. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 7307–7313. [Google Scholar]
  70. Yang, J.; Ye, Z.; Mei, D.; Gu, H.; Zhang, J. Long noncoding RNA DLX6-AS1 promotes tumorigenesis by modulating miR-497-5p/FZD4/ FZD6/Wnt/β-catenin pathway in pancreatic cancer. Cancer Manag. Res. 2019, 11, 4209–4221. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Zeng, X.; Hu, Z.; Ke, X.; Tang, H.; Wu, B.; Wei, X.; Liu, Z. Long noncoding RNA DLX6-AS1 promotes renal cell carcinoma progression via miR-26a/PTEN axis. Cell Cycle 2017, 16, 2212–2219. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Zhang, L.; He, X.; Jin, T.; Gang, L.; Jin, Z. Long non-coding RNA DLX6-AS1 aggravates hepatocellular carcinoma carcinogenesis by modulating miR-203a/MMP-2 pathway. Biomed. Pharmacother. 2017, 96, 884–891. [Google Scholar] [CrossRef] [PubMed]
  73. Zhang, J.-J.; Xu, W.-R.; Chen, B.; Wang, Y.-Y.; Yang, N.; Wang, L.-J.; Zhang, Y.-L. The up-regulated lncRNA DLX6-AS1 in colorectal cancer promotes cell proliferation, invasion and migration via modulating PI3K/AKT/mTOR pathway. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 8321–8331. [Google Scholar] [PubMed]
  74. Liu, H.; Ye, D.; Chen, A.; Tan, D.; Zhang, W.; Jiang, W.; Wang, M.; Zhang, X. A pilot study of new promising non-coding RNA diagnostic biomarkers for early-stage colorectal cancers. Clin. Chem. Lab. Med. 2019, 57, 1073–1083. [Google Scholar] [CrossRef]
  75. Qiu, C.; Li, S.; Sun, D.; Yang, S. lncRNA PVT1 accelerates progression of non‑small cell lung cancer via targeting miRNA‑526b/EZH2 regulatory loop. Oncol. Lett. 2020, 19, 1267–1272. [Google Scholar] [CrossRef] [Green Version]
  76. Huang, F.; Chen, W.; Peng, J.; Li, Y.; Zhuang, Y.; Zhu, Z.; Shao, C.; Yang, W.; Yao, H.; Zhang, S. LncRNA PVT1 triggers Cyto-protective autophagy and promotes pancreatic ductal adenocarcinoma development via the miR-20a-5p/ULK1 Axis. Mol. Cancer 2018, 17, 1–16. [Google Scholar] [CrossRef]
  77. Bukowski, R.M.M. Genitourinary Oncology: Current Status and Future Challenges. Front. Oncol. 2011, 1, 32. [Google Scholar] [CrossRef] [Green Version]
  78. Miller, K.D.; Nogueira, L.; Mariotto, A.B.; Rowland, J.H.; Yabroff, K.R.; Alfano, C.M.; Jemal, A.; Kramer, J.L.; Siegel, R.L. Cancer treatment and survivorship statistics, 2019. CA Cancer J. Clin. 2019, 69, 363–385. [Google Scholar] [CrossRef] [Green Version]
  79. Xiao, Z.-D.; Han, L.; Lee, H.; Zhuang, L.; Zhang, Y.; Baddour, J.; Nagrath, D.; Wood, C.G.; Gu, J.; Wu, X.; et al. Energy stress-induced lncRNA FILNC1 represses c-Myc-mediated energy metabolism and inhibits renal tumor development. Nat. Commun. 2017, 8, 1–13. [Google Scholar] [CrossRef]
  80. Hu, R.; Zhong, P.; Xiong, L.; Duan, L. Long Noncoding RNA Cancer Susceptibility Candidate 8 Suppresses the Proliferation of Bladder Cancer Cells via Regulating Glycolysis. DNA Cell Biol. 2017, 36, 767–774. [Google Scholar] [CrossRef]
  81. Jiang, S.; Zhang, L.-F.; Zhang, H.-W.; Hu, S.; Lu, M.-H.; Liang, S.; Li, B.; Li, Y.; Li, D.; Wang, E.-D.; et al. A novel miR-155/miR-143 cascade controls glycolysis by regulatinghexokinase 2in breast cancer cells. EMBO J. 2012, 31, 1985–1998. [Google Scholar] [CrossRef] [Green Version]
  82. Li, Z.; Li, X.; Wu, S.; Xue, M.; Chen, W. Long non-coding RNA UCA1 promotes glycolysis by upregulating hexokinase 2 through the mTOR –STAT3/microRNA143 pathway. Cancer Sci. 2014, 105, 951–955. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Hung, C.-L.; Wang, L.-Y.; Yu, Y.-L.; Chen, H.-W.; Srivastava, S.; Petrovics, G.; Kung, H.-J. A long noncoding RNA connects c-Myc to tumor metabolism. Proc. Natl. Acad. Sci. USA 2014, 111, 18697–18702. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Rupaimoole, R.; Lee, J.; Haemmerle, M.; Ling, H.; Previs, R.A.; Pradeep, S.; Wu, S.Y.; Ivan, C.; Ferracin, M.; Dennison, J.B.; et al. Long Noncoding RNA Ceruloplasmin Promotes Cancer Growth by Altering Glycolysis. Cell Rep. 2015, 13, 2395–2402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Fan, L.; Huang, C.; Li, J.; Gao, T.; Lin, Z.; Yao, T. Long non-coding RNA urothelial cancer associated 1 regulates radioresistance via the hexokinase 2/glycolytic pathway in cervical cancer. Int. J. Mol. Med. 2018, 42, 2247–2259. [Google Scholar] [CrossRef] [Green Version]
  86. Xing, Z.; Zhang, Y.; Liang, K.; Yan, L.; Xiang, Y.; Li, C.; Hu, Q.; Jin, F.; Putluri, V.; Putluri, N.; et al. Expression of Long Noncoding RNA YIYA Promotes Glycolysis in Breast Cancer. Cancer Res. 2018, 78, 4524–4532. [Google Scholar] [CrossRef] [Green Version]
  87. Yu, F.-X.; Zhao, B.; Guan, K.L. Hippo Pathway in Organ Size Control, Tissue Homeostasis, and Cancer. Cell 2015, 163, 811–828. [Google Scholar] [CrossRef] [Green Version]
  88. Zheng, X.; Han, H.; Liu, G.; Ma, Y.; Pan, R.; Sang, L.; Li, R.; Yang, L.; Marks, J.R.; Wang, W.; et al. Lnc RNA wires up Hippo and Hedgehog signaling to reprogramme glucose metabolism. EMBO J. 2017, 36, 3325–3335. [Google Scholar] [CrossRef] [PubMed]
  89. Li, Y.; Li, H.; Wang, W.; Yu, X.; Xu, Q. LINC00346 regulates glycolysis by modulation of glucose transporter 1 in breast cancer cells. Mol. Cell. Probes 2020, 54, 101667. [Google Scholar] [CrossRef]
  90. Sun, L.-Y.; Li, X.-J.; Sun, Y.-M.; Huang, W.; Fang, K.; Han, C.; Chen, Z.-H.; Luo, X.-Q.; Chen, Y.-Q.; Wang, W.-T. LncRNA ANRIL regulates AML development through modulating the glucose metabolism pathway of AdipoR1/AMPK/SIRT1. Mol. Cancer 2018, 17, 1–6. [Google Scholar] [CrossRef]
  91. Congrains, A.; Kamide, K.; Ohishi, M.; Rakugi, H. ANRIL: Molecular Mechanisms and Implications in Human Health. Int. J. Mol. Sci. 2013, 14, 1278–1292. [Google Scholar] [CrossRef] [Green Version]
  92. Shi, J.; Dai, R.; Chen, Y.; Guo, H.; Han, Y.; Zhang, Y. LncRNA LINP1 regulates acute myeloid leukemia progression via HNF4α/AMPK/WNT5A signaling pathway. Hematol. Oncol. 2019, 37, 474–482. [Google Scholar] [CrossRef]
  93. Zhang, Y.; He, Q.; Hu, Z.; Feng, Y.; Fan, L.; Tang, Z.; Yuan, J.; Shan, W.; Li, C.; Hu, X.; et al. Long noncoding RNA LINP1 regulates repair of DNA double-strand breaks in triple-negative breast cancer. Nat. Struct. Mol. Biol. 2016, 23, 522–530. [Google Scholar] [CrossRef] [PubMed]
  94. Wang, X.; Liu, H.; Shi, L.; Yu, X.; Gu, Y.; Sun, X. LINP1 facilitates DNA damage repair through non-homologous end joining (NHEJ) pathway and subsequently decreases the sensitivity of cervical cancer cells to ionizing radiation. Cell Cycle 2018, 17, 439–447. [Google Scholar] [CrossRef]
  95. Díaz-Beyá, M.; Brunet, S.; Nomdedéu, J.; Pratcorona, M.; Cordeiro, A.; Gallardo, D.; Escoda, L.; Tormo, M.; Heras, I.; Ribera, J.M.; et al. The lincRNA HOTAIRM1, located in the HOXA genomic region, is expressed in acute myeloid leukemia, impacts prognosis in patients in the intermediate-risk cytogenetic category, and is associated with a distinctive microRNA signature. Oncotarget 2015, 6, 31613–31627. [Google Scholar] [CrossRef] [Green Version]
  96. Hu, N.; Chen, L.; Li, Q.; Zhao, H. LncRNA HOTAIRM1 is involved in the progression of acute myeloid leukemia through targeting miR-148b. RSC Adv. 2019, 9, 10352–10359. [Google Scholar] [CrossRef] [Green Version]
  97. Yang, X.; Ye, H.; He, M.; Zhou, X.; Sun, N.; Guo, W.; Lin, X.; Huang, H.; Lin, Y.; Yao, R.; et al. LncRNA PDIA3P interacts with c-Myc to regulate cell proliferation via induction of pentose phosphate pathway in multiple myeloma. Biochem. Biophys. Res. Commun. 2018, 498, 207–213. [Google Scholar] [CrossRef]
  98. Liu, N.; Feng, S.; Li, H.; Chen, X.; Bai, S.; Liu, Y. Long non-coding RNA MALAT1 facilitates the tumorigenesis, invasion and glycolysis of multiple myeloma via miR-1271-5p/SOX13 axis. J. Cancer Res. Clin. Oncol. 2020, 146, 367–379. [Google Scholar] [CrossRef] [Green Version]
  99. He, Z.; Ruan, X.; Liu, X.; Zheng, J.; Liu, Y.; Liu, L.; Ma, J.; Shao, L.; Wang, D.; Shen, S.; et al. FUS/circ_002136/miR-138-5p/SOX13 feedback loop regulates angiogenesis in Glioma. J. Exp. Clin. Cancer Res. 2019, 38, 1–19. [Google Scholar] [CrossRef] [PubMed]
  100. Cheng, Z.; Luo, C.; Guo, Z. LncRNA-XIST/microRNA-126 sponge mediates cell proliferation and glucose metabolism through the IRS1/PI3K/Akt pathway in glioma. J. Cell. Biochem. 2019, 121, 2170–2183. [Google Scholar] [CrossRef]
  101. Xu, R.; Zhu, X.; Chen, F.; Huang, C.; Ai, K.; Wu, H.; Zhang, L.; Zhao, X. LncRNA XIST/miR-200c regulates the stemness properties and tumourigenicity of human bladder cancer stem cell-like cells. Cancer Cell Int. 2018, 18, 1–10. [Google Scholar] [CrossRef]
  102. Zhu, H.; Zheng, T.; Yu, J.; Zhou, L.; Wang, L. LncRNA XIST accelerates cervical cancer progression via upregulating Fus through competitively binding with miR-200a. Biomed. Pharmacother. 2018, 105, 789–797. [Google Scholar] [CrossRef]
  103. Wei, W.; Liu, Y.; Lu, Y.; Yang, B.; Tang, L. LncRNA XIST Promotes Pancreatic Cancer Proliferation Through miR-133a/EGFR. J. Cell. Biochem. 2017, 118, 3349–3358. [Google Scholar] [CrossRef]
  104. Shi, J.; Zhang, Y.; Qin, B.; Wang, Y.; Zhu, X. Long non-coding RNA LINC00174 promotes glycolysis and tumor progression by regulating miR-152-3p/SLC2A1 axis in glioma. J. Exp. Clin. Cancer Res. 2019, 38, 1–12. [Google Scholar] [CrossRef] [Green Version]
  105. Zhang, H.; Qin, D.; Jiang, Z.; Zhang, J. SNHG9/miR-199a-5p/Wnt2 Axis Regulates Cell Growth and Aerobic Glycolysis in Glioblastoma. J. Neuropathol. Exp. Neurol. 2019, 78, 939–948. [Google Scholar] [CrossRef] [PubMed]
  106. Li, X.; Liu, L.; Luo, Y.; Cui, S.; Chen, W.; Zeng, A.; Shi, Y.; Luo, L. Long non-coding RNA SNHG5 promotes glioma progression via miR-205/E2F3 axis. Biosci. Rep. 2019, 39. [Google Scholar] [CrossRef] [Green Version]
  107. Liu, L.; Shi, Y.; Shi, J.; Wang, H.; Sheng, Y.; Jiang, Q.; Chen, H.; Li, X.; Dong, J. The long non-coding RNA SNHG1 promotes glioma progression by competitively binding to miR-194 to regulate PHLDA1 expression. Cell Death Dis. 2019, 10, 1–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Chen, X.; Gao, Y.; Li, D.; Cao, Y.; Hao, B. LncRNA-TP53TG1 Participated in the Stress Response Under Glucose Deprivation in Glioma. J. Cell. Biochem. 2017, 118, 4897–4904. [Google Scholar] [CrossRef]
  109. Zheng, D.; Che, D.; Lin, F.; Wang, X.; Lu, L.; Chen, J.; Xu, X. LncRNA MACC1-AS1/MACC1 enhances the progression of glioma via regulating metabolic plasticity. Cell Cycle 2020, 19, 2286–2297. [Google Scholar] [CrossRef] [PubMed]
  110. Han, X.; Yang, Y.; Sun, Y.; Qin, L.; Yang, Y. LncRNA TUG1 affects cell viability by regulating glycolysis in osteosarcoma cells. Gene 2018, 674, 87–92. [Google Scholar] [CrossRef]
  111. Shen, Y.; Xu, J.; Pan, X.; Zhang, Y.; Weng, Y.; Zhou, D.; He, S. LncRNA KCNQ1OT1 sponges miR-34c-5p to promote osteosarcoma growth via ALDOA enhanced aerobic glycolysis. Cell Death Dis. 2020, 11, 1–14. [Google Scholar] [CrossRef] [PubMed]
  112. Song, J.; Wu, X.; Liu, F.; Li, M.; Sun, Y.; Wang, Y.; Wang, C.; Zhu, K.; Jia, X.; Wang, B.; et al. Long non-coding RNA PVT1 promotes glycolysis and tumor progression by regulating miR-497/HK2 axis in osteosarcoma. Biochem. Biophys. Res. Commun. 2017, 490, 217–224. [Google Scholar] [CrossRef] [PubMed]
  113. Chen, S.; Xu, X.; Lu, S.; Hu, B. Long non‑coding RNA HAND2‑AS1 targets glucose metabolism and inhibits cancer cell proliferation in osteosarcoma. Oncol. Lett. 2019, 18, 1323–1329. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Kang, Y.; Zhu, X.; Xu, Y.; Tang, Q.; Huang, Z.; Zhao, Z.; Lu, J.; Song, G.; Xu, H.; Deng, C.; et al. Energy stress-induced lncRNA HAND2-AS1 represses HIF1α-mediated energy metabolism and inhibits osteosarcoma progression. Am. J. Cancer Res. 2018, 8, 526–537. [Google Scholar] [PubMed]
  115. Yang, X.; Wang, C.C.; Lee, W.Y.W.; Trovik, J.; Chung, T.K.H.; Kwong, J. Long non-coding RNA HAND2-AS1 inhibits invasion and metastasis in endometrioid endometrial carcinoma through inactivating neuromedin U. Cancer Lett. 2018, 413, 23–34. [Google Scholar] [CrossRef] [PubMed]
  116. Ghafouri-Fard, S.; Esmaeili, M.; Taheri, M. H19 lncRNA: Roles in tumorigenesis. Biomed. Pharmacother. 2020, 123, 109774. [Google Scholar] [CrossRef]
  117. Wang, J.; Zhao, L.; Shang, K.; Liu, F.; Che, J.; Li, H.; Cao, B. Long non-coding RNA H19, a novel therapeutic target for pancreatic cancer. Mol. Med. 2020, 26, 1–7. [Google Scholar] [CrossRef] [Green Version]
  118. Luan, W.; Zhou, Z.; Ni, X.; Xia, Y.; Wang, J.; Yan, Y.; Xu, B. Long non-coding RNA H19 promotes glucose metabolism and cell growth in malignant melanoma via miR-106a-5p/E2F3 axis. J. Cancer Res. Clin. Oncol. 2018, 144, 531–542. [Google Scholar] [CrossRef]
  119. Gong, Y.; Cao, Y.; Song, L.; Zhou, J.; Wang, C.; Wu, B. HMGB3 characterization in gastric cancer. Genet. Mol. Res. 2013, 12, 6032–6039. [Google Scholar] [CrossRef]
  120. Elgamal, O.A.; Park, J.-K.; Gusev, Y.; Azevedo-Pouly, A.C.P.; Jiang, J.; Roopra, A.; Schmittgen, T.D. Tumor Suppressive Function of mir-205 in Breast Cancer Is Linked to HMGB3 Regulation. PLoS ONE 2013, 8, e76402. [Google Scholar] [CrossRef]
  121. Zou, Z.; Gao, J.; Zhang, H.; Lin, Z.; Zhang, Y.; Luo, X.; Liu, C.; Xie, J.; Cai, C. Increased expression of HMGB3: A novel independent prognostic marker of worse outcome in patients with esophageal squamous cell carcinoma. Int. J. Clin. Exp. Pathol. 2015, 8, 345–352. [Google Scholar]
  122. Wang, S.; Wang, T.; Liu, D.; Kong, H. LncRNA MALAT1 Aggravates the Progression of Non-Small Cell Lung Cancer by Stimulating the Expression of COMMD8 via Targeting miR-613. Cancer Manag. Res. 2020, 12, 10735–10747. [Google Scholar] [CrossRef]
  123. Zhang, R.; Niu, Z.; Pei, H.; Peng, Z. Long noncoding RNA LINC00657 induced by SP1 contributes to the non-small cell lung cancer progression through targeting miR-26b-5p/COMMD8 axis. J. Cell. Physiol. 2020, 235, 3340–3349. [Google Scholar] [CrossRef] [PubMed]
  124. Lang, N.; Wang, C.; Zhao, J.; Shi, F.; Wu, T.; Cao, H. Long non‑coding RNA BCYRN1 promotes glycolysis and tumor progression by regulating the miR‑149/PKM2 axis in non‑small‑cell lung cancer. Mol. Med. Rep. 2020, 21, 1509–1516. [Google Scholar] [CrossRef] [Green Version]
  125. Wang, L.; Wang, H.; Wu, B.; Zhang, C.; Yu, H.; Li, X.; Wang, Q.; Shi, X.; Fan, C.; Wang, D.; et al. Long Noncoding RNA LINC00551 Suppresses Glycolysis and Tumor Progression by Regulating c-Myc-Mediated PKM2 Expression in Lung Adenocarcinoma. OncoTargets Ther. 2020, 13, 11459–11470. [Google Scholar] [CrossRef]
  126. Zou, Z.W.; Ma, C.; Medoro, L.; Chen, L.; Wang, B.; Gupta, R.; Liu, T.; Yang, X.Z.; Chen, T.T.; Wang, R.Z.; et al. LncRNA ANRIL is up-regulated in nasopharyngeal carcinoma and promotes the cancer progression via increasing proliferation, reprograming cell glucose metabolism and inducing side-population stem-like cancer cells. Oncotarget 2016, 7, 61741–61754. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Wang, Y.; Zhang, X.; Wang, Z.; Hu, Q.; Wu, J.; Li, Y.; Ren, X.; Wu, T.; Tao, X.; Chen, X.; et al. LncRNA-p23154 promotes the invasion-metastasis potential of oral squamous cell carcinoma by regulating Glut1-mediated glycolysis. Cancer Lett. 2018, 434, 172–183. [Google Scholar] [CrossRef]
  128. Shi, T.T.; Li, R.; Zhao, L. Long noncoding RNA UCA1 regulates CCR7 expression to promote tongue squamous cell carcinoma progression by sponging miR-138-5p. Neoplasma 2021, 67, 1256–1265. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of a long non-coding RNA (lncRNA) interactome. LncRNAs interact with various molecules to perform important cellular functions (created with BioRender).
Figure 1. Schematic representation of a long non-coding RNA (lncRNA) interactome. LncRNAs interact with various molecules to perform important cellular functions (created with BioRender).
Cancers 13 00977 g001
Figure 2. Schematic representation of glucose metabolism regulation in cancer by lncRNAs (created with BioRender.com).
Figure 2. Schematic representation of glucose metabolism regulation in cancer by lncRNAs (created with BioRender.com).
Cancers 13 00977 g002
Table 1. Examples of lncRNAs implicated in glucose metabolism alteration in gastrointestinal (GI) cancers.
Table 1. Examples of lncRNAs implicated in glucose metabolism alteration in gastrointestinal (GI) cancers.
lncRNADeregulationFunctions and TargetsReference
AF339830 (GLCC1)Upregulated in CRCStabilizes c-MYC; upregulates LDHA; increases glycolysis[42]
LINRISUpregulated in CRCPrevents IGF2BP2 degradation; stabilizies c-MYC; upregulates GLUT-1, PKM2 and LDHA expession[44]
LINC00973Upregulated in cetuximab-resistant CRC cellsIncreases glucose consumption and lactate production[35]
LINC00504Upregulated in colon cancerIncreases c-MYC transcriptional activity; increases glycolysis, fatty acid synthesis, maintains redox status[43]
FEZF1-AS1Upregulated in CRCStabilizes and promotes the activity of PKM2; upregulates glycolysis and STAT3 signaling[49]
ENO1-IT1Fusobacterium nucleatum-activated lncRNA in CRCUpregulates the alpha-enolase glycolytic enzyme[30]
RAD51-AS1Downregulated in CRCceRNA for miR-29b-3p and miR-29c-3p; upregulates NDRG2; increases the levels of HK2 and GLUT-1, glucose uptake and lactate production[45]
lncARSRUpregulated in CRCSponges miR-34a-5p; upregulates HK1-mediated aerobic glycolysis[48]
CRNDEUpregulated in CRCUpregulates G6PD, GLUT-4 and MLXIPL; increases glucose uptake and lactate production[50]
LINC01554Downregulated in HCCPromotes PKM2 degradation; inhibits Akt/mTOR signaling[51]
MALAT1Upregulated in HCCUpregulates TCF7L2 via the mTORC1 pathway; upregulates the expression of glycolytic enzymes[52]
FtxUpregulated in HCCEnhances the activity and expression of GLUT-1 and GLUT-4 via PPARγ[53]
MEG3Downregulated in high glucose-treated HCCSponges miR-483-3p; suppresses ERp29 expression[54]
UCA1Upregulated in HCCIncreases the expression of HK2; enhances glucose uptake, lactate and ATP production[29]
RAET1KUpregulated in HCCSponges miR-100-5p; upregulates HIF1α; suppresses hypoxia-induced glucose uptake and lactate secretion[28]
HOTAIRUpregulated in HCCUpregulates GLUT-1; activates mTOR signaling pathway; decoy for miR-130a-3p (HIF1α target); promotes glycolysis and lactate production[55,56]
DDX11-AS1Upregulated in HCCSponges miR-195-5p; increases MACC1 expression; increases glucose consumption and lactate production[57]
SOX2OTUpregulated in HCCSponges miR-122-5p; activates PKM2, HK2 and LDHA[58]
DLX6-AS1Upregulated in gastric cancerSponges miR-4290; upregulates PDK1; elevates glycolysis rates; inhibits mitochondrial respiration[59]
DGCR9Upregulated in gastric cancerIncreases glucose uptake[60]
HOTAIRUpregulated in pancreatic cancerUpregulates HK2; increases glucose uptake, ATP and lactate production[61]
LINC00184Upregulated in esophageal cancerSilences PTEN; increases glycolysis and decreases mitochondrial oxidative phosphorylation[62]
UCA1Upregulated in esophageal cancerSponges miR-203; suppresses HK2 degradation; increases glucose uptake and lactate output[63]
PVT1Upregulated in gallbladder cancerSponges and repress miR-143; upregulates HK2; increases glucose uptake and lactate output[64]
Table 2. Examples of lncRNAs implicated in glucose metabolism alteration in GU cancers.
Table 2. Examples of lncRNAs implicated in glucose metabolism alteration in GU cancers.
lncRNADeregulationFunctions and TargetsReference
FILNC1Downregulated in renal cell cancerInteracts with AUF1 (c-MYC mRNA-binding protein); downregulates c-MYC protein levels; decreases the glucose update, lactate production and tumor development[79]
CASC8Downregulated in bladder cancerBinds to the fibroblast growth factor receptor 1 (FGFR1); prevents FGFR1-mediated lactate dehydrogenase A phosphorylation; decreases glycolysis and cell proliferation[80]
UCA1Upregulated in bladder cancerUpregulates HK2 through mTOR and its downstream effector STAT3; upregulates HK2 by active mTOR downregulation of miR143[81,82]
PCGEM1Upregulated in prostate cancerBinds to the c-MYC promotor; upregulates c-MYC; induces aerobic glycolysis, pentose phosphate shunt and NADPH[83]
Table 3. Examples of lncRNAs implicated in glucose metabolism alteration in female reproductive system cancers.
Table 3. Examples of lncRNAs implicated in glucose metabolism alteration in female reproductive system cancers.
lncRNADeregulationFunctions and TargetsReference
NRCPUpregulated in ovarian cancersActs as an intermediate binder between STAT1 and RNA polymerase II; increases the expression of glucose-6-phosphate isomerase (GPI), aldolase A (ALDOA) and aldolase C (ALDOC)[84]
GHET1Upregulated in ovarian cancersInteracts with the E3 ubiquitin ligase von Hippel-Lindau (VHL); prevents VHL-mediated HIF1α-degradation; promotes glucose uptake and lactate production[27]
H19Upregulated in ginsenoside 20(S)-Rg3-treated ovarian cancersSponges miR-324-5p; upregulates PKM2; increases glucose uptake and lactate production[34]
UCA1Upregulated in radioresistant cervical cancerUpregulates HK2[85]
Table 4. Examples of lncRNAs implicated in glucose metabolism alteration in breast cancer.
Table 4. Examples of lncRNAs implicated in glucose metabolism alteration in breast cancer.
lncRNADeregulationFunctions and TargetsReference
BCAR4Upregulated in triple negative breast cancerInteracts with GLI2; upregulates HK2 and PFKFB3; increases glucose uptake and lactate production[88]
YIYAUpregulated in breast cancerInteracts with CDK6; increases phosphorylation PFKFB3; catalysis of glucose 6-phosphate to fructose bisphosphate[86]
LINC00346Upregulated in breast cancer cellsSponges miR-148a/b; enhances GLUT-1 expression[89]
Table 5. Examples of lncRNAs implicated in glucose metabolism alteration in hematologic cancer.
Table 5. Examples of lncRNAs implicated in glucose metabolism alteration in hematologic cancer.
lncRNADeregulationFunctions and TargetsReference
ANRILUpregulated in acute myeloid leukemia (AML)Increases the expression of GLUT-1 and LDAH, upregulates AdipoR1, AMPK and SIRT1[90]
LINP1Upregulated in AMLIncreases the expression of GLUT-1 and LDHA; upregulates HNF4α; activates the AMPK/WNT5A signaling pathway[92]
UCA1Upregulated in AMLceRNA of miR-125a; upregulates HK2; blocks HIF1α-dependent glycolysis[32]
HOTAIRM1Upregulated in AMLActivates the Wnt/β-catenin pathway; enhances glucose uptake and lactate production[33]
PDIA3PUpregulated in multiple myelomaInteracts with c-MYC; upregulates G6PD expression and PPP flux[97]
MALAT1Upregulated in multiple myelomaTargets miR-1271-5p; increases SOX13 expression[98]
Table 6. Examples of lncRNAs implicated in glucose metabolism alteration in brain cancers.
Table 6. Examples of lncRNAs implicated in glucose metabolism alteration in brain cancers.
lncRNADeregulationFunctions and TargetsReference
XISTUpregulated in glioblastomaSponges miRNA126; upregulates the IRS1 and PI3K/Akt pathways; upregulates the GLUT-1 and GLUT-3 receptors; increases cellular glucose uptake and lactate production[100]
LINC00174Upregulated in gliomaSponges miR-1523-3p; upregulates SLC2A1[104]
SNHG9Upregulated in glioblastomaDownregulates miR-199a-5p; upregulates the Wnt2 axis[105]
SNHG1Upregulated in gliomaSponges miR-194; upregulates PHLDA1[107]
SNHG5Upregulated in gliomaSponges miR-205; upregulates E2F transcription factor 3[106]
TP53TG1Upregulated in gliomaUpregulates GRP78, LDHA and IDH1[108]
MACC1-AS1Upregulated in gliomaIncreases the levels of GLUT-1, HK2, G6PD, MCT1, ATP, lactate and NAPDH[109]
Table 7. Examples of lncRNAs implicated in glucose metabolism alteration in osteosarcoma.
Table 7. Examples of lncRNAs implicated in glucose metabolism alteration in osteosarcoma.
lncRNADeregulationFunctions and TargetsReference
TUG1Upregulated in osteosarcomaIncreases HK2 expression, glucose consumption, lactate production[110]
KCNQ1OT1Upregulated in osteosarcomaSponges miR-34c-5p; upregulates aldolase A (ALDOA) expression[111]
PVT1Upregulated in osteosarcomaIncreases HK2 expression, glucose consumption, lactate production[112]
HAND2-AS1Downregulated in osteosarcomaDecreases GLUT-1 and glucose update[113,114]
Table 8. Examples of lncRNAs implicated in glucose metabolism alteration in lung cancers.
Table 8. Examples of lncRNAs implicated in glucose metabolism alteration in lung cancers.
lncRNADeregulationFunctions and TargetsReference
AC020978Upregulated in NSCLCInteracts with and stabilizes PKM2; promotes PKM2 nuclear translocation; upregulates HIF1α, Oct-4, STAT3 and β-catenin[25]
HOTTIPUpregulated in NSCLCSponges miR-615-3p; increases HMGB3 expression; upregulates HK2 expression; increases glucose consumption and lactate production[26]
MALAT1Upregulated in NSCLCSponges miR-613; upregulates COMMD8; increases glucose consumption and lactate production[122]
BCYRN1Upregulated in NSCLCReduces miR‑149 expression; upregulates PKM2 expression; induces glycolysis[124]
LINC00551Downregulated in lung adenocarcinomaDecreases c-MYC expression; decreases PKM2; inhibits glycolysis[125]
Table 9. Examples of lncRNAs implicated in glucose metabolism alteration in head and neck cancer.
Table 9. Examples of lncRNAs implicated in glucose metabolism alteration in head and neck cancer.
lncRNADeregulationFunctions and TargetsReference
ANRILUpregulated in nasopharyngeal carcinomaUpregulates mTOR signal pathway; increases GLUT-1 and LDHA expression[126]
lnc-p23154Upregulated in oral squamous cell carcinomaSponges miR-378a-3p; upregulates GLUT-1 expression[127]
UCA1Upregulated in tongue squamous cell carcinomaSponges miR-138-5p; upregulates CCR7[128]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Balihodzic, A.; Barth, D.A.; Prinz, F.; Pichler, M. Involvement of Long Non-Coding RNAs in Glucose Metabolism in Cancer. Cancers 2021, 13, 977. https://doi.org/10.3390/cancers13050977

AMA Style

Balihodzic A, Barth DA, Prinz F, Pichler M. Involvement of Long Non-Coding RNAs in Glucose Metabolism in Cancer. Cancers. 2021; 13(5):977. https://doi.org/10.3390/cancers13050977

Chicago/Turabian Style

Balihodzic, Amar, Dominik A. Barth, Felix Prinz, and Martin Pichler. 2021. "Involvement of Long Non-Coding RNAs in Glucose Metabolism in Cancer" Cancers 13, no. 5: 977. https://doi.org/10.3390/cancers13050977

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

Article Metrics

Back to TopTop