Next Article in Journal
Claudin-4: A New Molecular Target for Epithelial Cancer Therapy
Next Article in Special Issue
S100A8 and S100A9 in Hematologic Malignancies: From Development to Therapy
Previous Article in Journal
Deregulation of Astroglial TASK-1 K+ Channel Decreases the Responsiveness to Perampanel-Induced AMPA Receptor Inhibition in Chronic Epilepsy Rats
Previous Article in Special Issue
IKAROS in Acute Leukemia: A Positive Influencer or a Mean Hater?
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Metabolic Reprogramming and Potential Therapeutic Targets in Lymphoma

1
Division of Hematopathology, Department of Pathology, Duke University School of Medicine, Durham, NC 27710, USA
2
Department of Hematology, Ninth People’s Hospital, Shanghai Jiao-Tong University School of Medicine, Shanghai 200025, China
3
Duke Cancer Institute, Durham, NC 27710, USA
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2023, 24(6), 5493; https://doi.org/10.3390/ijms24065493
Submission received: 11 February 2023 / Revised: 9 March 2023 / Accepted: 10 March 2023 / Published: 13 March 2023
(This article belongs to the Special Issue Molecular Mechanism of Leukemia)

Abstract

:
Lymphoma is a heterogeneous group of diseases that often require their metabolism program to fulfill the demand of cell proliferation. Features of metabolism in lymphoma cells include high glucose uptake, deregulated expression of enzymes related to glycolysis, dual capacity for glycolytic and oxidative metabolism, elevated glutamine metabolism, and fatty acid synthesis. These aberrant metabolic changes lead to tumorigenesis, disease progression, and resistance to lymphoma chemotherapy. This metabolic reprogramming, including glucose, nucleic acid, fatty acid, and amino acid metabolism, is a dynamic process caused not only by genetic and epigenetic changes, but also by changes in the microenvironment affected by viral infections. Notably, some critical metabolic enzymes and metabolites may play vital roles in lymphomagenesis and progression. Recent studies have uncovered that metabolic pathways might have clinical impacts on the diagnosis, characterization, and treatment of lymphoma subtypes. However, determining the clinical relevance of biomarkers and therapeutic targets related to lymphoma metabolism is still challenging. In this review, we systematically summarize current studies on metabolism reprogramming in lymphoma, and we mainly focus on disorders of glucose, amino acids, and lipid metabolisms, as well as dysregulation of molecules in metabolic pathways, oncometabolites, and potential metabolic biomarkers. We then discuss strategies directly or indirectly for those potential therapeutic targets. Finally, we prospect the future directions of lymphoma treatment on metabolic reprogramming.

1. Introduction

The reprogramming of metabolic pathways is shaped by both oncogenes and tumor-suppressor genes to ensure a stabilized supply of intermediate metabolites for the biosynthesis of biomass and generation of ATP in tumor cells [1,2,3,4,5,6]. Deregulated uptake of glucose and amino acids (AA), using nutrients and intermediates of the tricarboxylic acid cycle (TCA cycle) occasionally, NADPH production, increased requirements for nitrogen, alterations in metabolite-driven gene regulation, and interactions between the metabolism and microenvironment are regarded as crucial hallmarks of tumor metabolism [7,8].
Lymphoma is a group of heterogeneous diseases, and 40–50% of patients achieve a complete response after inductive therapy [9,10,11]. High glucose uptake, deregulated expression of glycolytic enzymes, the dual ability for both glycolytic and oxidative metabolism, active glycolysis prevailing over aerobic glucose metabolism without defects in mitochondrial function, and increased glutamine metabolism and fatty acid biosynthesis are notable metabolic effects of lymphoma [12]. This metabolic reprogramming is a dynamic process involving dysregulation of transcription factors and signaling pathways, such as phosphatidylinositol 3-kinase (PI3K)/mammalian target of rapamycin (mTOR), hypoxia-inducible factors (HIF), MYC, p53, and AMP-activated protein kinase (AMPK). Additionally, some viruses, such as Epstein–Barr virus (EBV), also contribute to the metabolic abnormalities associated with lymphoma. Changes in the tumor microenvironment—hypoxia, for example—can also induce metabolic alterations in lymphoma. Under hypoxic conditions, several pathways of cell survival and proliferation, such as glycolysis, glutamine oxidation, bioenergetics, and redox homeostasis, are activated. These metabolic alterations are associated with both tumorigenesis and disease progression [13]. Some dominant mutations in genes encoding metabolic enzymes, which play significant roles in tumorigenesis and tumor progression, cannot be ignored. Thus, therapeutic strategies targeting critical metabolic enzymes, transcription factors, signaling pathways, and the microenvironment are being investigated both in preclinical and clinical research.
In this review, we focus on the metabolic characteristics of lymphoma. We aim to highlight the dysregulation and functional defects in metabolic pathways in lymphoid malignancies and discuss metabolic biomarkers and targeted therapies.

2. Metabolic Alterations in Lymphoma

2.1. Altered Glucose Metabolism

2.1.1. Up-Regulated Glucose Uptake

High glucose uptake is one of the hallmarks of aggressive lymphomas and is associated with poor outcomes in diffuse large B-cell lymphoma (DLBCL) [14,15]. Increased glucose uptake is a feature of malignant cells and can be detected via positron emission tomography/computed tomography (FDG-PET), a non-invasive molecular imaging tool, by analyzing glucose metabolism [16]. FDG-PET has become a standard staging and efficacy evaluation method, especially for DLBCL [7]. Hexokinase 2 (HK2) and, glucose transporter (GLUT) 3 contribute to FDG uptake and higher HK2 expression in DLBCL may contribute to its higher FDG uptake than non-DLBCL [17] (Table 1). GLUT3 was also identified as a glucose transporter of FDG accumulation in primary central nervous system lymphoma [18] (Table 1). In patients with advanced-stage classic Hodgkin lymphoma (cHL), however, higher GLUT1 expression is associated with programmed cell death ligand 1 (PD-L1) and PD-L2 expression and prolonged survival [19] (Table 1, Figure 1).

2.1.2. Aerobic Glycolysis

The increased metabolism of glucose to lactate under aerobic conditions (the Warburg effect) is another feature of cancer cells [6]. Compared to oxidative phosphorylation (OxPhos), which produces 36 ATP molecules per glucose molecule, glycolysis is inefficient and generates only two ATP molecules per glucose molecule. However, aerobic glycolytic metabolism represents ~60% of the energy produced in malignant cells due to their uptake of larger amounts of glucose and higher rates of glycolysis [20]. Some crucial glycolytic enzymes are increased and play essential roles in energy metabolism, which are recognized as potential biomarkers and relevant therapeutic targets in cancers, including aggressive B-cell lymphomas [21,22,23]. HK2 is highly expressed in DLBCL and results in higher FDG uptake [17]. HK2 is also required to promote the growth of DLBCL cells under hypoxia [24] (Table 1). Increased levels of serum lactate dehydrogenase (LDH) in primary cutaneous lymphoma (PCL) have been associated with independent risk factors for progression-free survival (PFS) rate and overall survival (OS) rate [25] (Table 1). The LDH-5 is upregulated in non-Hodgkin lymphoma (NHL), which is directly related to clinical stage, extra-nodal site involvement, and performance status of patients with NHL [26] (Table 1). The excreted lactate is eliminated from the cells via monocarboxylate transporters (MCTs), which have been found to be highly expressed in stromal cells of DLBCL samples [27]. Accordingly, therapeutic strategies that disrupt lactate transport might be promising approaches for treating lymphoma [28,29,30,31] (Figure 1).
Table 1. Enzymes and metabolite alterations in lymphoma.
Table 1. Enzymes and metabolite alterations in lymphoma.
Enzymes /MetabolitesSubtype of LymphomaAlterationMechanismRef
GLUT3DLBCLupregulatedFDG uptake[17]
GLUT3PCNSLupregulatedglucose transporter of FDG accumulation[18]
GLUT1cHLupregulatedbetter clinical outcomes[19]
HK2DLBCLupregulatedFDG uptake, promote cell growth[17,24]
LDHPCLupregulatedaggressive clinical behavior, resistance to chemotherapy, and low survival rates[25]
LDH-5NHLhighly upregulatedcorrelated with HIF-1α cytoplasmic accumulation in NHL cells [26]
MCT1DLBCLhighly expressedan OXPHOS phenotype[27]
SHMTDLBCL make glycine for purine synthesis[32]
IDO 1DLBCL converts tryptophan into kynurenine-pathway metabolites, inhibits T-cell activity, and induces immune tolerance, predict a favorable prognosis[33]
IDO 1HL converts tryptophan into kynurenine-pathway metabolites, inhibits T-cell activity, and induces immune tolerance, predict dismal clinical outcomes[34]
FASNDLBCLoverexpressedaggressive clinical course and therapeutic resistance, shorter OS and PFS[35]
PI3K/AKT/mTORBCL and TCLactivatedrelated to p53, HIF-1α, and MYC[36,37,38,39]
PKM2ALCLinactivatedinduces oncogenesis by phosphorylating nuclear STAT3 and regulating the transcription of genes involved in cell survival and proliferation[40]
POX/PRODHMYC-driven lymphomaexpressedmitochondrial tumor suppressor[41]
GLSBLupregulatedenhances glutamine uptake and glutamine catabolism[41]
ChokαTCLoverexpressedregulation of choline metabolism[42]
SDHBLreduced expressionincreased the intracellular drug concentration[43]
IDH2AITLmutatedincreasing DNA hypermethylation of gene promoters[44,45]
D2HGDHDLBCLmutatedaffects histone and DNA methylation and HIF-1α hydroxylation[46]
Abbreviations: GLUT1, glucose transporter 1; GLUT3, glucose transporter 3; HK2, hexokinase 2; LDH, lactate dehydrogenase; MCT, monocarboxylate transporter; SHMT, serine hydroxymethyltransferase; IDO, indoleamine 2,3-dioxygenase; FASN, fatty acid synthase; PI3K/AKT/mTOR, phosphatidylinositol 3-kinase/Akt serine/threonine kinase 1/mammalian target of rapamycin, PKM2, pyruvate kinase M2; POX/PRODH, proline dehydrogenase; GLS, glutaminase; Chokα, choline kinase-α; SDH, dehydrogenase; IDH2, isocitrate dehydrogenase 2; D2HGDH, D2-hydroxyglutarate dehydrogenase; DLBCL, diffuse large B-cell lymphoma; PCNSL, primary central nervous system lymphoma; cHL, classic Hodgkin lymphoma; PCL, primary cutaneous lymphoma; NHL, non-Hodgkin lymphoma; HL, Hodgkin lymphoma; BCL, B-cell lymphoma; TCL, T-cell lymphoma; ALCL, anaplastic large-cell lymphoma; BL, Burkitt lymphoma; AITL, angioimmunoblastic T-cell lymphoma; FDG, fluorodeoxyglucose; OS, overall survival; PFS, progression free survival; OXPHOS, oxidative phosphorylation; HIF-1α, hypoxia-inducible factors-1 alpha; STAT3, signal transducer and activator of transcription 3.

2.1.3. Mitochondrial Metabolism

As the main metabolic reaction, mitochondrial metabolism reprograms to reserves functions through several mechanisms during aerobic glycolysis. Because it supports bioenergetics and homeostasis, mitochondrial-dependent metabolic reprogramming is crucial for the survival and proliferation of tumor cells during Warburg metabolism [47,48,49] (Figure 1). Although impaired mitochondrial respiration is often construed as the cause of aerobic glycolysis, increasing evidence suggests that both glycolytic and oxidative metabolism exist in tumor cells [50,51], varying among tumor types [20]. Moreover, the fact that glycolysis overwhelms, rather than reduces, mitochondrial function depends on the glycolytic phenotype of the tumor cells [52].
Mitochondria generate most of ATP in lymphoma cells [14]. Glycolytic ATP production differs in DLBCL subtypes with distinct gene expression profiles [15]. Using genome-wide arrays, DLBCL can be subdivided into subsets, mainly including the B cell receptor/proliferation cluster (BCR-DLBCL) and the oxidative phosphorylation cluster (OXPHOS-DLBCL) [53,54,55]. Approximately 30% ATP is produced from glycolysis in cultured OXPHOS-DLBCL cell lines and 45% in BCR-DLBCL cell lines [15]. BCR-DLBCLs have exhibited higher glycolytic flux due to the increased expression of BCR signaling components triggered in a PI3K-dependent manner, making them sensitive to BCR signaling inhibitors.
In contrast, nutrient and energy metabolism in OXPHOS-DLBCL show a significant mitochondrial characteristic with increased activity of components of the mitochondrial electron transport chain and elevated expression of mitochondrial translation factors. Accordingly, OXPHOS is elevated, and there is an increased contribution of the mitochondria to the total cellular energy budget; several glucose-derived carbons and fatty acids are incorporated into the TCA cycle [56,57]. Thus, increased lipogenesis from these carbon substrates is a specific metabolic characteristic of OXPHOS-DLBCL. Although the utilization of fatty-acid-dependent carbons in OXPHOS-DLBCL occurs in parallel to BCR signaling, the former does not display the complete phosphorylation pattern and is resistant to inhibitors of BCR signaling [53,57,58,59]. In contrast, pharmacological inhibition of the mitochondrial translation signal leads to the inhibition of cell proliferation in OXPHOS-DLBCL [56].

2.2. Altered Amino Acid Metabolism

Enhancement of glucose metabolism has been assumed to promote the bioenergetic requirements of tumor cells. However, hyperproliferative tumor cells need additional biosynthetic processes beyond glucose metabolism. Therefore, anaplerosis via pyruvate carboxylation, glutamine metabolism, and other sources, such as fatty acid metabolism, is necessary [60,61,62,63,64]. Under aerobic conditions, glutaminolysis, which is regulated by the activity of the MYC oncogene, may play a considerable part in the biosynthetic pathways [63,64].

2.2.1. Glutamine Metabolism

Glutamine is a non-essential and abundant AA in the human body. It is a major nitrogen donor in the biosynthesis of AA and nucleotides [65] and is essential for the survival and proliferation of malignant cells [2]. Glutamine catabolism (called glutamine decomposition) is upgraded in several glutamine-dependent tumors and is often driven by glutaminase (GLS), which transforms glutamine into glutamate and ammonia [4]. In glutamine-dependent tumors, glutamine carbon is present in aspartate, glutamate, and TCA cycle metabolites through glutaminolysis [60,66,67]. DLBCL consumes mitochondrial pyruvate via glutamate-pyruvate transaminase 2 to enable α-ketoglutarate (α-KG) production as part of glutaminolysis [68]. Thus, glutamine exceeds pyruvate as a carbon source for the TCA cycle in DLBCL. Under states of glutamine starvation, glutamine promotes tumor cell proliferation by providing α-KG (a process termed anaplerosis) to supplement depleted TCA cycle intermediates [67,69,70]. In ibrutinib-resistant MCL cells, GLS is overexpressed, and it correlates well with elevated glutamine dependency and glutaminolysis. In addition, GLS expression is correlated with MYC expression and depletion of GLS, or else glutamine significantly reduced cell growth [71]. The nuclear factor erythroid 2-related factor 2 (NFE2L2, also known as NRF2) is a transcription factor that stimulates the expression of genes, and it is a cellular protector that applies to both normal cells and malignant cells [72]. In lung adenocarcinoma, mutations in Kelch-like ECH-associated protein 1 (KEAP1), which constitutively activates NRF2, are required for glutathione synthesis [3]. Keap1 and Nrf2 expression was associated with high international prognostic index (IPI) scores and advanced clinical stage in DLBCL [73]. It was reported that NRF2 plays a significant role in chemotherapy resistance in hematologic malignancies. NRF2 activation metabolically rewires and elevates pathways involved in glutamine metabolism, making it vulnerable to glutaminase inhibition [3,74,75]. In a MYC-induced human Burkitt lymphoma (BL) cell line, glutamine was imported and metabolized through the TCA cycle under hypoxia. Inhibiting glutamine entry into the TCA cycle can inhibit tumor progression [47,76]. Although glutamine supports tumorigenesis directly, glutamine uptake is often severely restricted in developing tumors [77] (Figure 1).

2.2.2. Proline Metabolism

Proline is the only proteinogenic secondary AA that participates in redox regulation and metabolic signaling. Glutamate generated by proline can be further converted to α-KG or can generate glutamine via glutamine synthetase (GS). Proline dehydrogenase (POX/PRODH) participates in the first step of proline catabolism and is considered a mitochondrial tumor suppressor in MYC-driven lymphoma [41] (Table 1). Proline catabolism induced by POX/PRODH generates electrons to produce reactive oxygen species (ROS) and triggers a broad spectrum of downstream reactions, including cell cycle arrest and initiation of apoptosis [41]. In a human BL cell line, MYC suppressed POX/PRODH expression primarily through up-regulating miR-23b*. Furthermore, MYC markedly increased the enzymes of proline biosynthesis from glutamine [41]. Understanding the relationship between glutamine and proline metabolism might enable us to develop novel therapeutic strategies.

2.2.3. Serine, Glycine, and Serine Hydroxymethyltransferase

Serine and glycine are non-essential AAs involved in various anabolic processes to support tumor cell proliferation [78]. Restricting of dietary serine and glycine can decrease tumor growth in MYC-driven lymphoma animal models [79]. Serine hydroxymethyltransferase (SHMT) transforms serine into glycine and transfers a one-carbon unit to s tetrahydrofolate. DLBCL cells are defective in glycine uptake, therefore depending on SHMT to make glycine for purine synthesis, which renders them particularly sensitive to SHMT inhibitors [32] (Table 1, Figure 1).

2.2.4. Indoleamine 2,3-Dioxygenase 1

Indoleamine 2,3-dioxygenase 1 (IDO 1) is an intracellular enzyme that converts tryptophan into kynurenine-pathway metabolites, inhibits T-cell activity, and induces immune tolerance in the tumor microenvironment. Increased IDO-positive cells can predict a favorable prognosis in patients with DLBCL treated with R-CHOP [33] (Table 1). However, IDO expression is associated with dismal clinical outcomes in HLs [34] (Table 1). In adult T-cell lymphoma (TCL) and cutaneous T-cell lymphoma (CTCL), IDO may also cause immunosuppression and predict prognosis, which indicates the potential of developing novel cancer immunotherapies targeting IDO [80,81]. Epacadostat, an IDO1 inhibitor, can significantly decrease the concentration of kynurenine in cell cultures [80]. Because of the inhibition of IDO on chimeric antigen receptor T (CD19-CART) cells, fludarabine and cyclophosphamide are used frequently before CART cell administration of CART cells, which can downregulate IDO expression and improve the antitumor activity of CD19-CART cells in vivo [82].

2.3. Altered Lipid Metabolism

Elevated fatty acid biosynthesis is one of the hallmarks and an important metabolic inducer in multiple malignancies.

2.3.1. Fatty Acid Synthase

Fatty acid synthase (FASN) plays a critical role in the pathogenesis of lymphoma. Increased expression of FASN is seen in patients with DLBCL and is associated with shorter OS and PFS. Moreover, FASN overexpression in DLBCL is recognized as an independent prognostic factor that predicts an aggressive clinical course and therapeutic resistance [83,84] (Table 1). FASN can induce S6Kinase to facilitate the formation of the USP11-eukaryotic initiation factor 4B (eIF4B) complex for sustained tumorigenic translation in DLBCL. In contrast, eIF4B expression can be reduced at transcription and translation levels when FASN expression is suppressed [83]. FASN is up-regulated in a PI3K-dependent manner in primary effusion lymphoma (PEL) and other types of non-viral B-NHLs [85] (Table 1). FASN is highly and consistently expressed in mantle cell lymphomas (MCLs), as well [86]. Thus, FASN might be a potential therapeutic target for lymphoma therapy. Although inhibiting FASN alone or in combination with PI3K demonstrated a robust decrease in tumor growth, current FASN inhibitors have limited clinical applications because of certain pharmacological limitations [35,87,88] (Table 1, Figure 1).

2.3.2. PPARδ

Peroxisome proliferator-activated receptor delta (PPARδ) is a member of the lipid-activated nuclear receptor family, which regulates the expression of genes involved in glucose, lipid, and cholesterol metabolism [89,90,91,92]. PPARδ can improve metabolic efficiency and tolerance to oxidative stress to make malignant B cells survive under cytotoxic drugs or low glucose and oxygen [91]. PPARδ also regulates the signal transduction process by altering cholesterol metabolism and cytokine signaling. Thus, PPARδ antagonists may also play a therapeutic role in lymphoma [93].

2.3.3. Cholesterol Synthesis

Cancer cells have an increased demand for cholesterol. By targeting and blocking cholesterol uptake, cellular cholesterol can be reduced through the high-affinity, high-density lipoprotein (HDL) receptor scavenger receptor type B-1 (SCARB1). In activated B cell (ABC) DLBCL cell lines, enhanced BCR signaling and the resultant de novo cholesterol synthesis can significantly reduce the ability of HDL-like nanoparticles (HDL NPs) to reduce cellular cholesterol levels, and induce cell death. Thus, ABC DLBCLs are thought to be less sensitive to HDL NPs than GCB DLBCLs. Therefore, by targeting both cellular cholesterol uptake and BCR-associated cholesterol synthesis, cellular cholesterol reduction and apoptosis can be achieved in resistant ABC DLBCL cell lines [94].

2.3.4. Choline Metabolism

Choline kinase-α (Chokα), an enzyme involved in the regulation of choline metabolism, leads to the phosphorylation of choline to phosphocholine [95]. The overexpression of Chokα is frequently observed in cancers and is associated with insufficient histological differentiation and poor prognosis [96]. Dysregulation of choline metabolism in TCLs is related to the overexpression of Chokα in tumor cells [42]. Chokα, as a regulatory gene, possesses carcinogenic activity and could be a potential therapeutic target for hematological malignancies [42] (Table 1 and Table 2).

3. Regulation of Altered Metabolism in Lymphoma

Metabolic reprogramming in lymphoma, including glucose metabolism, fatty acid synthesis, and glutaminolysis, is a dynamic process driven by the dysregulation of HIF, MYC, p53, PI3K/mTOR, and AMPK signaling pathways, as well as EBV.

3.1. HIF

HIF-1α is a subunit of a heterodimeric transcription factor HIF-1that controls the expression of more than 200 genes affecting large-scale cellular processes, including glucose transport, cellular metabolism, angiogenesis, cell proliferation, apoptosis, and erythropoiesis, in response to cellular hypoxia [97,98,99,100]. In addition, PI3K/AKT/mTOR, RAS-mitogen-activated protein kinase, extracellular signal-regulated kinase, erbB2 families, MYC, and ROS can also independently increase the expression of HIF-1α expression [101,102,103,104,105]. In lymph node biopsies of DLBCL and FL patients, stabilization and upregulation of HIF-1α were observed [106]. Furthermore, in patients with DLBCL treated with R-CHOP-like regimens, the expression of HIF-1α protein is a significant independent prognostic factor for superior survival [107,108].
HIF-1α can also regulate the expression of genes encoding essential enzymes of aerobic glycolysis in BLs [65,109,110,111]. In a MYC-driven BL cell line, HIF-1α, together with MYC, regulates the HK2 and PDK1 expression [112]. Lymphoblastoid cells also showed a Warburg effect regulated by HIF-1α [109,112]. In anaplastic lymphoma kinase (ALK)-rearranged anaplastic large-cell lymphoma (ALCL), hypoxia pathways, regulated by ALK via STAT3- and C/EBPβ-dependent transcription, are significantly enriched [113]. HIF-1α is dispensable for cell growth in ALCL, whereas HIF-2α is required for both cell growth and internal environment maintenance in ALCL [114]. Moreover, HIF-1α plays a crucial role in the pathogenesis of PEL, an aggressive B-cell lymphoma with a poor prognosis caused by KSHV, which promotes the activity of HIF-1α and, in turn, activates the KSHV gene. HIF-1α can also induce aerobic and anaerobic glycolysis and lipid biosynthesis in PEL cell lines [115].
HIF-1α activation resulted in global translation inhibition coupled with deregulation of mitochondrial function during hypoxic stress in DLBCLs. However, the translational inhibition by HIF-1α is not complete, and the expression levels of hypoxic targets, such as GLUT1 and HK2, are refractory to translational inhibition in DLBCL cells [24] (Figure 2).

3.2. MYC

The MYC oncogene, which plays a crucial role in several human cancers, including B- and T-cell malignancies, is regarded as a master regulator of cell metabolism and proliferation [116]. As a transcription factor encoded by MYC, c-Myc (MYC) can associate altered cellular metabolism with tumorigenesis. Most metabolic genes are housekeeping genes, which have an open chromatin state, and their transcription extension is also enhanced by c-Myc [117].
In lymphoma, MYC activation occurs through various molecular processes, including mutations, amplification, translocations, altered intracellular localization of the MYC protein, and miRNA-dependent mechanisms [118,119,120]. The oncogenic HSP90, which optimizes several MYC metabolic programs, is indispensable to fulfill biomass, energetic, and secretory demands of B-cell lymphomas and, specifically, to support the metabolic process driven by MYC [121]. MYC overexpression, caused by gene rearrangements and mutations, has been observed in various hematological malignancies, especially in invasive B-cell lymphomas, such as DLBCLs and BLs [122,123,124]. MYC protein, the primary regulator of the Warburg effect, regulates HK2 and PDK1 genes cooperatively with HIF-1α. In a BL cell model, the expression level of genes involved in glucose metabolism was decreased after the loss of MYC function [110,125]. In BLs, the MYC protein not only promotes glucose uptake that induces aerobic glycolysis, but it also enhances glutamine uptake and glutamine catabolism by up-regulating mitochondrial GLS [41] (Table 1). In addition, MYC promotes cell proliferation in malignancies through various biosynthetic pathways, particularly ornithine and polyamine biosynthesis pathways [63,64,126]. Through the downregulation of GLS via the upregulation of miR-23b* and an increase in the expression of the enzymes of proline biosynthesis from glutamine, MYC suppresses the expression of the mitochondrial tumor suppressor POX/PRODH [41]. In addition, by upregulating the key enzyme phosphate cytidylyltransferase 1 in DLBCLs, MYC promotes choline metabolism and inhibits mitophagy-dependent necroptosis [127] (Figure 2).

3.3. The PI3K/mTOR Pathway

Through activating mutations, gene amplifications, or rearrangements in PIK3CA and receptor tyrosine kinases or inactivating mutations in tumor suppressor genes, the PI3K/AKT/mTOR pathway, which regulates cell metabolism, proliferation, apoptosis, and autophagy, is altered in several malignancies [128,129,130,131].
PI3K is a lipid kinase that participates in intracellular signal transduction. Multiple pathways mediated by PI3K-δ and/or PI3K-γ induce cell survival, proliferation, and differentiation in hematologic malignancies [7,132,133]. AKT, a downstream effector of PI3K, is an essential inducer of the glycolytic phenotype, which maintains the survival of cancer cells that depend on glycolysis [134,135]. By up-regulating GLUT1 expression, activating glycolytic enzymes, modulating HK2 expression, and activating and interacting with mitochondria, AKT stimulates glucose uptake and glycolysis [136]. mTOR is a serine/threonine kinase downstream of PI3K-AKT that activates the transcription of multiple genes involved in cancer cell metabolism through mTOR complexes 1 and 2 (mTORC1 and 2) [137]. mTORC1, which is sensitive to rapamycin and environmental alterations [138], is a vital regulator controlling a range of metabolic processes, including glycolysis and mitochondrial metabolism [139,140]. NHLs, characterized by enhanced metabolic activities, are highly associated with aberrant activation of the PI3K/AKT/mTOR signaling pathway [141]. Activation of the PI3K/AKT/mTOR pathway is recurrent in both B-cell lymphomas and TCLs and appears to be related to p53, HIF-1α, and MYC [36,37,38,39,128,142]. Thus, pharmacologic inhibition of the PI3K/AKT/mTOR pathway can be employed for lymphoma therapy (Table 1, Figure 2).

3.4. p53

Tumor suppressor p53 regulates numerous signaling pathways in response to diverse cellular stresses, including DNA damage, loss of normal cell contact, abnormal oncogenic events, and hypoxia in some normal cellular processes [143,144].
In NHLs, p53 inhibits glucose uptake by inhibiting the NF-κB pathway and regulating the flux of glucose in glycolysis. Loss of function of p53 gives rise to the activation of GLUT-3 transcription through the NF-κB pathway [145]. In addition, p53 can activate the synthesis of cytochrome c oxidase-2, resulting in the accumulation of ATP and providing negative feedback to the glycolysis pathway. In addition, the p53-inducible enzyme TIGAR can also give negative feedback to glycolysis activity [146].
TP53 is one of the most important factors correlated with dismal survival and mediates cell resistance to chemotherapy in patients with DLBCL. More than 20% of patients with DLBCL have TP53 mutations, most of which can disrupt protein function and lead to disease progression [147,148]. The adverse outcomes of patients with DLBCLs who exhibit MYC rearrangement, MYC expression alone, or concurrent BCL2 expression can be augmented by TP53 mutations [149] (Figure 2).

3.5. AMPK

AMPK, a highly conserved Ser/Thr protein kinase complex composed of catalytic α, regulatory β, and γ subunits, plays a crucial role as a master regulator of cellular energy homeostasis [150]. AMPK deregulates aerobic glycolysis and tumor cell growth in vivo. Blocking AMPK signaling actives the metabolic reprogramming of cancer cells and induces normoxic HIF-1α stabilization, resulting in the Warburg effect and affecting tumor progression in vivo [151]. Defective AMPKα signaling leads to the rewiring of metabolic pathways to favor cell growth pathways, resulting in the functioning of a tumor-suppressor pathway [152,153]. High expression levels of two AMPKβ subunits are significantly associated with prolonged five-year OS in patients with NHLs [154]. In c-Myc transgenic mice, deletion of the AMPKα1 isoform promoted B-cell lymphoma [151]. In addition, the reduction of the AMPKβ1 subunit promotes the development of p53-mutant TCLs [155] (Figure 2). In contrast, AMPK activators (metformin and 5-aminoimidazole-1-β-4-carboxamide-1-β-D-ribofuranoside [AICAR]) were required for capsaicin-induced mTOR complex 1 (mTORC1) inhibition, BCL2 downregulation, and Bax upregulation in osteosarcoma cells [156]. AMPK activation could block lymphoma cell growth via inhibition of the mTOR pathway and the induction of autophagy [157]. For instance, high-dose metformin, which plays an important role in B-cell development, could influence the AMPK signaling and reduce ATP production and inhibit cell proliferation in DLBCL [158]. AICAR had a selective antitumor activity in MCL cell lines. Moreover, AICAR was highly synergistic with rituximab both in vitro and in vivo [159]. Aside from activating AMPK, AICAR can also inhibit mTORC1. When combined with a mTORC1 inhibitor, AICAR enhances the efficacy of rapamycin, which indicated a synergistic effect [160]. In solid tumors, combined treatment of rapamycin and AICAR showed an additive effect on tumor size reduction compared with use of each drug alone [161], which might be theoretically possible to imitate in lymphoma therapy. Therefore, AMPK signaling, regardless alone or combined, might be a potential target in the treatment of DLBCL.

3.6. SIRT4

SIRT4, a sirtuin localized in the mitochondria, is a tumor suppressor present in various tumor models. The positive expression of SIRT4 in B-cell lymphoma downregulates glutamine uptake and inhibits cell growth, whereas deletion of SIRT4 in the Eµ-myc models up-regulates glutamine consumption and accelerates tumorigenesis [162].

3.7. EBV

EBV infects almost every adult with limited symptoms. However, in the state of immune suppression, EBV can cause lymphomas [163]. EBV infection is supposed to participate in the development of a variety of malignant lymphomas and lymphoproliferative disorders (LPDs), such as HL, BL, some types of DLBCL, primary effusion lymphoma, EBV-positive T-cell LPDs, and so on [164]. It has been estimated that nearly half of all HLs contain EBV DNA. Latent membrane protein (LMP) 2A, one of expressed proteins, plays an important role in enabling EBV and establishes a lifelong latent infection and the development of EBV-associated diseases [165]. Epstein–Barr nuclear antigen 2 (EBNA2), together with its key host target MYC, induce metabolism pathways needed for B cell remodeling and proliferation [166]. EBV infection can induces glucose uptake, OXPHOS, purine and pyrimidine, amino acid, and fatty acid biosynthesis pathways in lymphoma development [167]. EBV replication induces B cell expansion in germinal center, actives cytidine deaminase, and MYC translocation, which is essential to the development and progression of BL [168]. Lactate export mediated by MCTs, one part of the Warburg effect, is crucial for continued cell proliferation. EBV infection of B lymphocytes could directly promote the induction of MCT1 and MCT4 through the viral proteins EBNA2 and LMP1 [163]. Dual MCT1/4 inhibition induced lymphoblastoid cell death through the electron transport chain complex I inhibitors [163]. Pharmacological inhibition of lactate export program could render lymphoma cells hypersensitive to metformin and phenformin (electron transport chain complex I inhibitors), which might be a therapeutic approach for targeting EBV related lymphomas [163]. Ectopic expression of LMP1 in EBV-mediated B-cell growth transformation induced FASN [169]. Targeting lipogenesis might be an attractive strategy in treating LMP1-positive EBV-associated lymphomas. NF-κB activation is typical of transformed B lymphocytes, including herpes virus transformed lymphoblasts. Inhibition of the NF-κB pathway in the EBV transformed B-cells lowers glucose uptake and triggers its autophagy-induced death [170]. Besides, EBV infected B cells show upregulation of HIF-1α, MYC, p53, PI3K/mTOR, and AMPK signaling pathways [164], which are all dysregulated in lymphoma metabolic reprogramming.

4. Oncometabolites in Lymphoma

Oncometabolites, found in serum, plasma, urine, saliva, and tumor tissue samples, can be evaluated as cancer biomarkers. They refer to metabolites in tumor cells that are significantly upregulated compared with those in normal cells [171]. Oncometabolites are identified based on oncogenic driver mutations of genes encoding some metabolic enzymes [172]. A major mode of action of these oncometabolites is to act as competitive inhibitors of metabolite-dependent enzymes due to their structural similarity to intermediate metabolites [171]. Therefore, mitochondria have a strong effect on controlling the spatial structure of chromatin and give rise to the occurrence of tumors by producing oncometabolites.

4.1. Glycolytic Enzymes

Recent studies have shown that increased GAPDH expression can induce NF-κB-dependent HIF-1α activation and can play a crucial role in the angiogenesis and invasiveness of NHLs [22]. In ALCL, the enzymatically inactive isoform of PKM2, regulated by ALK, can induce oncogenesis by phosphorylating nuclear STAT3 and regulating the transcription of genes involved in cell survival and proliferation [40] (Table 1).

4.2. SDH

Some TCA enzymes, such as succinate dehydrogenase (SDH), can act as classical tumor suppressors. Dysfunction of SDH can result in the accumulation of metabolic intermediate and exhaustion of aspartate. Aspartate generation depends on the activity of pyruvate carboxylase (PC), which catalyzes the carboxylation of pyruvate to oxaloacetate. SDH-defective cells increase PC expression to synthesize aspartate from glucose to manage TCA cycle truncation and produce aspartate [173,174]. Reduced SDH expression was observed in both doxorubicin- and vincristine-exposed Ramos cells (BL cell line), with a prolonged duration in culture, and this increased the intracellular drug concentration [43] (Table 1).

4.3. IDH2, D2HG, and D2HGDH

α-KG, an intermediate metabolite of the TCA cycle, coordinates epigenetic plasticity and leads to malignant behavior and tumor cell survival [175,176]. Under mitochondrial dysfunction caused by the TCA cycle or mutations in the genes encoding electron transport chain enzymes, some glutamine-derived α-KGs undergo reductive carboxylation to support redox homeostasis and biosynthesis [177]. Isocitrate dehydrogenase 1 or 2 (IDH1 or IDH2) frequently undergo gene mutations in AML [178,179,180,181,182]. However, in angioimmunoblastic TCLs, IDH2 mutations play a crucial role in the pathogenesis by increasing DNA hypermethylation of gene promoters [44,45] (Table 1). Recurrent somatic point mutations in the coding regions of IDH1 and IDH2 generate 2-hydroxyglutarate (2HG) through the downregulation of α-KG expression. D-2HG is the exclusive 2HG stereoisomer produced by IDH1 and IDH2 mutants, which can accumulate in high concentrations and inhibit α-KG-dependent dioxygenases, including enzymes involved in DNA and histone demethylation, leading to hypermethylation of CpG islands and histones in DNA [183,184,185,186,187,188,189,190]. D2-HG can be transformed into α-KG by D2-hydroxyglutarate dehydrogenase (D2HGDH). By elevating α-KG levels, wild-type D2HGDH affects histone and DNA methylation and HIF-1α hydroxylation, whereas mutant-type D2HGDH has the opposite effect in DLBCLs [46] (Table 1).

5. Therapeutic Strategies Targeting Metabolism

In the past decade, progress in targeting lymphoma metabolism therapeutically has been limited. Only a few metabolism-based drugs for lymphoma have been successfully developed, including those which already have been approved for clinical application and ongoing clinical trials (Table 2).

5.1. Targeting Metabolic Pathways

5.1.1. Targeting Glucose Metabolism

Pharmaceutical inhibition of glycolytic enzymes may provide a novel therapeutic strategy for NHLs. HK2 induces the conversion of glucose to glucose-6-phosphate, which is the first step in glycolysis. Two-deoxyglucose (2-DG), an analog of glucose, once bound and phosphorylated by HK2, is not further metabolized, thereby suppressing HK2 [191]. In NHL cell lines, 2-DG can inhibit cell proliferation both under hypoxia and normoxia (Table 3). When combined with methylprednisolone, 2-DG synergistically inhibits cell proliferation by downregulating HIF-1α and c-Myc [192]. Because 2-DG-induced toxicity is regulated by members of the BCL2 family, it can be enhanced by antagonizing BCL2-positive lymphoma cell lines. 2-DG can also induce GADD153/CHOP expression, which is a marker of endoplasmic reticulum stress and an activator of Bim [193]. A glycolysis inhibitor that targets Mcl-1 can restore the sensitivity of lymphoma cells (primary Ramos cells or Eµ-myc) to ABT-737-induced apoptosis [52]. In MCL cells, 2-DG and glucose restriction via anti-glycolytic drugs inhibit TRAIL-induced cell death, indicating that mitochondrial metabolism directly regulates the sensitivity of tumor cells to apoptosis [194]. 3-bromopyruvate (3-BrPA) cannot only restrain tumor glycolysis acting through the hexokinase step, but also hampers mitochondrial ATP production. Both in vitro and in vivo, 3-BrPA demonstrated a significant positive tumor response in Raji-lymphoma-bearing mice [195,196] (Table 3). With augmented induction of apoptosis, 3-bromopyruvate (3-BP), a pyruvate inhibitor and brominated derivative of pyruvate, can inhibit metabolism and survival in Dalton’s lymphoma (DL) cells [197] (Table 3). FX11, a competitive small-molecule inhibitor of LDHA, inhibits cell proliferation and induces death in P493 BL cells by reducing ATP levels and inducing significant oxidative stress [198,199] (Table 3). Interestingly, in lymphoma-bearing mice, α-tocopherol, the most active component of vitamin E, also contributes to keeping cell proliferation in check by downregulating LDHA, PKC-α, and c-Myc expression [199] (Table 3). Inhibition of MCT1 using AZD3965 that blocked lactate efflux led to the accumulation of glycolytic intermediates in vitro and significant downregulation of tumor proliferation in vivo in the Raji BL model. Moreover, when combined with doxorubicin or rituximab, enhanced cell proliferation inhibition and cell death were observed in vitro and in vivo [200,201]. Dichochloroacetate (DCA), a by-product of drinking water disinfection, blocks phosphorylation of PDK at the mitochondrial membrane level, and thus, glycolysis is downregulated due to the activation of the PDH [202,203] (Table 3). Accompanied by modulation of glycolysis and expression of HIF1-α, DCA inhibited cell survival in a DL mouse model [204]. Taken together, these data suggest that glycolytic enzymes with deregulated expression levels are potential biomarkers and have become potential therapeutic targets in NHLs. Tigecycline, approved by FDA, is selectively toxic to OxPhos-DLBCL cell lines and primary tumors and can pharmacologically disturb the mitochondrial translation pathway (Table 3). These findings indicate that the mitochondrial translation pathway is a potential therapeutic target for these tumors as well [56].
The presence of IDH1 or IDH2 mutations and accumulation of 2-HG may result in dependence on specific pathways and the introduction of therapeutic vulnerability. For instance, tumors that harbor IDH mutations are more sensitive to electron-transport-chain inhibitors [205], hypomethylating agents [190], depletion of the canonical coenzyme NAD+, and chemoradiotherapy [206]. Pharmacological molecules targeting mutant IDH1 and IDH2 enzymes are being explored and assessed for antitumor efficacy.
Table 3. Targeting lymphoma metabolism via metabolic enzymes, metabolite depletion, and/or signaling pathways.
Table 3. Targeting lymphoma metabolism via metabolic enzymes, metabolite depletion, and/or signaling pathways.
PathwayCompoundApplicationDevelopment StageRef
HK22-DGB-NHLcell lines[192]
3-BrPABLcell lines, mice[195,196]
Pyruvate3-BPT-NHLmice (DL)[197]
LDHAFX11BLcell line[198,199]
α-TocopherolT-NHLmice (DL)[199]
PDKDCADLmice[202,203]
Mitochondrial protein translation TigecyclineOxPhos-DLBCL cell linesFDA-approved[56]
GlutaminaseBPTESBLmice [47]
Glutamine uptakeL-asparaginaseNHLcell lines[207]
SHMT1/2SHIN1DLBCLcell lines[33]
FASNorlistatMCLcell lines[86,208]
T-NHLcell lines, mice[209]
C75DLBCL, PEL, and B-NHLcell lines[84,85]
NABaPBLpatients[210]
PPARαFenofibrateB-NHLmice[211]
Choline kinaseCK37T-NHLmice[42]
HIF-1αPX-478PELcell lines[115]
PCI-24781(HDACi)DLBCLPhase 1/2[212]
SAHA(HDACi)B-NHLcell lines, mice[213]
MYC10058-F4DLBCLcell lines[214,215,216]
PI3KLY294002B-NHLcell lines[217]
AZD8835ABC-DLBCLcell lines[218]
AKTAkti1/2PELcell lines[219]
AZD-5363PTEN-deficient DLBCLcell lines, mice[218]
MK-2206ABC-DLBCLmice[220]
NaB (HDACi)BLcell lines[221]
mTORRapamycinALCL, NHLcell lines[36,131]
mTOR C1/2AZD-2014MCLcell lines[222]
Dual inhibitor of PI3K and mTORNVP-BEZ235PELmice[223]
PF-04091502PELcell lines[219]
Bimiralisib (PQR309)DLBCL, MCL, SMZL, CLL, HL, and ALCLcell lines, mice[224]
AMPKAICAR MCL, SMZL, FL, and CLLcell lines, mice and patients[159,225,226,227,228,229]
MetforminB and T-NHLcell lines, mice[157,230]
PhenforminPTEN-deficient T-cell lymphomascell lines[230]
Abbreviation: HK2, hexokinase 2; LDHA, lactate dehydrogenase A; PDK, pyruvate dehydrogenase kinase; SHMT1/2, serine hydroxymethyltransferase 1/2; FASN, fatty acid synthase; PPARα, peroxisome proliferator-activated receptor alpha; HIF-1α, hypoxia-inducible factor 1 alpha; PI3K, phosphatidylinositol 3-kinase; mTOR, mammalian target of rapamycin; AMPK; DG, deoxyglucose; BrPA, bromopyruvate; BP, bromopyruvate; DCA, dichloroacetate; BPTES, bis-2-(5-phenylacetamido-1,2,4-diathiazol-2-yl) ethyl sulfide; SHIN1, serine hydroxymethyltransferase 1 inhibitor; HDACi, histone deacetylase inhibitor; SAHA, suberoylanilide hydroxamic acid; AICAR, 5-aminoimidazole-4-carboxamide-1-β-D-ribofuranoside; B-NHL, B-cell non-Hodgkin lymphoma; BL, Burkitt lymphoma; T-NHL, T-cell non-Hodgkin lymphoma; DL, Dalton’s lymphoma; DLBCL, diffuse large B-cell lymphoma; MCL, mantle cell lymphomas; PEL, primary effusion lymphoma; ABC, activated B cell; SMZL, splenic marginal zone lymphoma; FL, follicular lymphoma; CLL, chronic lymphocytic leukemia; ALCL, anaplastic large-cell lymphoma.

5.1.2. Targeting Amino Acid Metabolism

Glutamine metabolism plays a crucial role in cell survival and proliferation under glucose deficiency and hypoxia, making it sensitive to the glutaminase inhibitor Bis-2-[5-(phenylacetamido)-1,3,4-thiadiazol-2-yl]ethyl sulfide (BPTES) [47] (Table 3). L-Asparaginase (L-ASNase), which exhibits some glutaminase activity, can hydrolyze extracellular glutamine to glutamate and ammonia and prevent glutamine from entering the cell [231]. In NHL cell lines, L-ASNase exerts cytotoxicity by depriving the cells of glutamine, resulting in the suppression of cell growth and survival [207] (Table 2 and Table 3).

5.1.3. Targeting Lipid Metabolism

Pharmacological inhibition using the FASN-specific inhibitor C75 triggered caspase-dependent apoptosis in DLBCL cell lines [84] (Table 3). FASN, up-regulated in PEL and other types of non-viral B-NHLs in a PI3K dependent manner, is sensitive to C75 as well [85] (Table 3). FASN is also highly and consistently expressed in MCLs. MCL cell lines, in which FASN is highly and consistently expressed, exhibited significant apoptosis when treated with orlistat, an anti-obesity drug, which is also an inhibitor of FASN and is approved by the FDA [86] (Table 3). Orlistat can also interfere with the ubiquitination of NOXA protein and induce apoptosis, thereby offering new strategies to kill bortezomib-resistant MCL cells [208]. TCLs, when treated with orlistat in vitro, manifest tumor-specific inhibition of cell survival and induction of apoptosis [209]. Orlistat-induced tumor growth retardation decreased tumor cell survival and chemosensitization to cisplatin, and a prolonged life span was observed in tumor-bearing mice [209] (Table 3). BaP, a redeployed drug that combines bezafibrate and medroxyprogesterone acetate, whose complete mechanism was not fully understood, was extended to the clinic and demonstrated efficacy with low toxicity in clinical trials involving patients with BL (ISRCTN34303497) [210] (Table 3). Interestingly, recent findings indicate that obesity is associated with an increased risk of developing malignant lymphomas. In wild-type B-cell lymphoma mice, tumor size was significantly associated with the depletion of white adipose tissues (WAT) and elevated levels of lipid metabolites. Thus, tumor growth can be significantly suppressed with PPARα agonists and the lipid-lowering drug fenofibrate [211] (Table 3).

5.2. Targeting Oncogenic Regulators

5.2.1. HIF-1α Inhibitors

HIF-1α plays an important role in the pathogenesis of lymphoma. Inhibition of HIF-1α may have a therapeutic effect on lymphoma. PX-478 is a small-molecule inhibitor of HIF-1α and inhibits lymphoma cell growth even under normoxia [115] (Table 3). PCI-24781 is a broad-spectrum histone deacetylase inhibitor (HDACi) that can promote the accumulation of HIF-1α and induce initial autophagy of lymphoma cells. Long-term incubation with HDACi can inhibit the expression of HIF-1α [212] (Table 3). Suberoylanilide hydroxamic acid (SAHA, vorinostat), a second-generation HDACi, can inhibit cell proliferation and induce apoptosis by inhibiting the expression of HIF-1α in B-cell lymphoma in vitro and in vivo [213]. Bortezomib is the first proteasome inhibitor used for treating MCLs by suppressing the transcription and expression of HIF-1-specific target genes. Although HIF-1α is not destroyed by bortezomib, heterodimeric HIF-1 cannot transactivate the target genes [232] (Figure 3).

5.2.2. MYC Inhibitors

Most MYC inhibitors are intended to interfere with the MYC–MAX interaction. Because MYC lacks hydrophobic pockets, its structure makes it difficult to target the MYC protein directly. Therefore, inhibition of c-Myc is an attractive strategy to inhibit glucose metabolism and the proliferation of lymphoma cells. In DLBCL cell lines, including GCB and ABC, cell proliferation was inhibited by the c-Myc inhibitor 10058-F4, as a single agent and in combination with 2-DG [217]. However, 10058-F4 is ineffective in vivo due to its poor bioavailability and rapid metabolism and clearance in other cancers [214,215,216] (Table 3). Thus, no small-molecule inhibitor targeting the MYC gene is currently tested in clinical trials. Recently, Bhatt et al. cultured DLBCL cells in three-dimensional matrices and discovered a reduced proliferative activity and altered metabolism that is consistent with in vivo tumor cell function [233]. Development of three-dimensional conditions that stabilize tumor cell function towards in vivo phenotype might be helpful for the understanding of lymphoma metabolic process and make progress in relevant research.
A therapeutic approach targeting MYC phosphorylation and degradation is a promising way of treating cancers addicted to high MYC protein levels [234,235,236]. The enzyme 1α/X-box binding protein 1/stearoyl-CoA-desaturase 1 (IRE1α/XBP1/SCD1) axis plays a protective role in counterbalancing anabolism mediated by MYC overexpression. Genetic and pharmacological inhibition of XBP1 leads to MYC-dependent apoptosis, which can be alleviated by exogenous unsaturated fatty acids. In addition, IRE1α inhibition can enhance the cytotoxic effects of standard chemotherapy against BL with MYC overexpression [237].
The oncogenic form of HSP90 optimizes several MYC metabolic pathways, including the production of nucleotides [121]. Given the importance of MYC in driving metabolic reprogramming in lymphoma progress, oncogenic HSP90 inhibitors could reverse the immunosuppressive effects on the lymphoma microenvironment and thus potentially improve efficacy of lymphoma immunotherapy.

5.2.3. The PI3K/mTOR Pathway

PI3K Inhibitors

Treatment with the PI3K inhibitor LY294002, either as a single agent or in combination with 2-DG, reduces cell survival, FDG uptake, and cell growth [217] (Table 3). Inhibition of PI3K-α and -δ by copanlisib demonstrated manageable safety and significant efficacy in relapsed or refractory (R/R) indolent lymphoma patients who were heavily pretreated. High response rates were associated with increased expression of PI3K/BCR signaling pathway genes [238] (Table 2). Another PI3K-α/δ inhibitor, AZD8835, demonstrated remarkable potency in ABC-like DLBCL models by inhibiting NF-κB signaling and had a synergistic effect with the Bruton’s tyrosine kinase (BTK) inhibitor ibrutinib both in vitro and in vivo [218] (Table 3). A similar anti-tumor effect was observed in the ABC DLBCL mouse xenograft model treated with a combination of the PI3Kδ inhibitor idelalisib and BTK inhibitor ONO/GS-4059 [220]. Duvelisib, an oral inhibitor of PI3K-δ/γ isoforms, demonstrated acceptable safety and promising clinical activity in patients with R/R TCLs [239]. In vitro, phospho-AKT (pAKT) showed a synergistic effect with duvelisib in TCL cell lines [239] (Table 2 and Table 3, Figure 3).

Akt Inhibitors

Akti 1/2, an AKT inhibitor that reduces the rate of lactate production in hypoxic conditions, displays improved cytotoxicity to PEL cells [219]. In combination with the 2-DG, Akti 1/2 demonstrated strong synergistic cytotoxicity toward PEL cells and shifted metabolism from aerobic glycolysis toward oxidative respiration [219] (Table 3). Through the downregulation of MYC, the AKT inhibitor AZD5363 induced apoptosis in PTEN-deficient DLBCLs [218] (Table 3). Combined with the PI3Kδ inhibitor idelalisib, the AKT inhibitor MK-2206 could increase the sensitivity of tumor cells to idelalisib in an ABC DLBCL mouse xenograft [220] (Table 3). In BL cell lines, by regulating AKT phosphorylation and MYC protein expression, cell proliferation was inhibited by HDACi (sodium butyrate, NaB) combined with VP-16, which indicates that the PI3K/AKT pathway is a target of HDACi as well [221] (Table 3). Hypoglycemic agents, such as metformin, phenformin, and the AMPK activator acadesine, have strong antitumor effects on T-cell specific PTEN-deficient (tPTEN−/−) lymphoma cells [230] (Figure 3).

mTOR Inhibitors

Rapamycin inhibits the growth of TCLs by decreasing the glycolytic rate and glucose utilization both in vitro and in vivo. Rapamycin-treated cells displayed reduced sensitivity to low-glucose conditions but continued to rely on OXPHOS, which can be reversed by an OXPHOS inhibitor [131] (Table 3). Besides, accumulating evidence indicates that rapamycin may serve as a potential glucocorticoid (GC) sensitizer in lymphomas through genetic prevention of 4E-BP1 phosphorylation [240,241]. Rapamycin also led to a reduction in GLUT1 mRNA expression, FDG uptake, and cell survival of NHL cell lines [36] (Table 3). A dual mTORC1/2 kinase inhibitor AZD-2014 can inhibit cellular energy metabolism via the TCA cycle and further inhibit glycolysis in MCL cells [222] (Table 3, Figure 3).

Dual Inhibitors of PI3K and mTOR

Dual targeting of PI3K and mTOR by the inhibitor NVP-BEZ235 led to improved results compared to the application of rapamycin alone in PEL [223] (Table 3). PF-04691502, another dual inhibitor of PI3K and mTOR, which has a synergistic effect with 2-DG, demonstrated increased cytotoxicity in PEL cells under hypoxic conditions or in a glycolytic phenotype [219] (Table 2).
Bimiralisib (PQR309), an oral, novel, selective dual PI3K/mTOR inhibitor, had anti-lymphoma activity as a single agent or in combination with ibrutinib, lenalidomide, and rituximab in vitro. Increased transcripts coding for the BCR pathway can improve the activity of bimiralisib. Bimiralisib showed activity in lymphoma cells with primary or secondary resistance to idelalisib and appeared to be a novel and promising therapeutic compound [224] (Table 3).

5.2.4. AMPK Activators

Activation of AMPK may represent an important therapeutic strategy for lymphoma treatment due to its essential role in its pathway [242,243,244]. Metformin, belonging to the biguanide class of oral hypoglycemic agents, could potentially inhibit cell growth in lymphomas as an AMPK activator both in vitro and in vivo (Table 3). The activation of AMPK promoted by metformin was associated with inhibition of the mTOR pathway without involving AKT [157]. Acadesine (AICAR) is a cell-permeable nucleoside analog that can be metabolically transformed to AICA ribotide (ZMP) by the cells. AICAR can mimic a low-energy state and can regulate the anticancer effects via different mechanisms [242,245,246,247,248]. AICAR has been shown to induce selective cell apoptosis, cell proliferation inhibition, and cell cycle arrest in several hematological malignancies [159,226,227,228,229] (Table 3). AICAR showed an acceptable safety profile in a phase I/II clinical trial involving patients with R/R chronic lymphocytic leukemia [249]. In MCL patients, AICAR showed a cytotoxic effect when combined with rituximab [159]. AICAR triggers the activation of the AMPK pathway, inhibits the downstream mTOR cascade, and forces MCL cells to enter caspase-dependent apoptosis. The selective BH3-mimetic agent ABT-199 targeting BCL2 increases the sensitivity to AICAR in BCL2-overexpressing MCL cells [159,225] (Table 3, Figure 3).

6. Potential Metabolic Biomarkers of Lymphoma

Metabolomics has already been used to investigate the pathogenesis of diseases and explore new biomarkers for disease diagnosis, treatments, and prognosis. The PI3K/AKT/mTOR signaling pathway, for instance, has been shown to participate in the cell pro-survival and metabolic reprogramming involving fatty acid metabolism, glycolysis, and tricarboxylic acid cycle in B-cell lymphoma [250]. Wang et al. generated a metabolic gene panel, using 13 metabolism-associated genes (MAGs) in DLBCL-related metabolic pathways, which divided patients with DLBCL into one of two risk groups. The combination of metabolic gene signatures and other prognostic factors showed a superior prognostic power than IPI and other standard prognostic clinical variables [250]. Based on the expression patterns of 92 prognosis associated MAGs mining from public database, He et al. divided DLBCL patients into two metabolic clusters with significantly different prognoses [251]. A prognostic risk model was constructed based on 14 genes selected from 92 MAGs, and then the DLBCL patients were classified into two risk groups, which was superior to the IPI [251]. These nomograms would be useful in clinical practice and future clinical trials. Serum samples from 100 DLBCL patients and 100 matched healthy controls were analyzed by an untargeted mass-spectrometry-based metabolomics platform. The results of this combined study indicate that 2-AG might play a potential role in the pathogenesis and progression in patients with DLBCL [252]. The differences in serum metabolites between BLs and normal mice were analyzed by nuclear magnetic resonance-based metabolomics. Glutamate, glycerol, and choline had a high accuracy of diagnosis, which might provide non-invasive approaches for the diagnosis and prognosis of patients with BL [253]. Serum metabolomics analysis can identify high-risk DLBCL patients with failure of immunochemotherapy, which might provide another novel, non-invasive way for the diagnosis and prognosis of lymphoma [254]. A study on the skin and plasma of CTCL mice indicated that aberrant metabolites and metabolic pathways were essential metabolic features of CTCLs, whereas accumulative cytidine-5′-triphosphate in adjacent non-involved skin tissues led to CTCL further development [255]. Changes in lipid profiles provide new biological insights into how MYC regulates cellular metabolism in MYC-induced lymphoma [256] (Table 4).

7. Conclusions and Perspectives

Lymphoma is a hematological malignancy with complex terminology and variable clinical outcomes and continues to be an area of unmet need. Metabolic reprogramming in lymphoma, involving glycolysis, lipid metabolism, and glutaminolysis, is a complicated process regulated by numerous tumor oncogenes, suppressors, and pathways. These regulators mainly target the most crucial enzymes of metabolic pathways, leading to a shift in metabolic phenotypes and accumulation of aberrant metabolites, some of which are related to the biological behavior of lymphoma cells and may serve as potential molecular biomarkers.
However, the molecular heterogeneity of DLBCL brings great challenges to precision therapy. Over the last two decades, molecular and genetic therapeutics have been used in patients, including those with R/R lymphomas. Studies have indicated that immune checkpoints also play crucial roles in metabolic process. Checkpoint signals, on the one hand, can regulate metabolic process [257]. For instance, PD-L1 can activate the PI3K/AKT/mTOR pathway, enhance glucose uptake, and stimulate glycolysis by the tumor cells [258]. On the other hand, metabolism modulates the tumor response to immune checkpoint inhibitor (ICI). For example, obesity patients are recognized to have enhanced PD-1 expression, which is associated with better efficiency to ICIs in solid tumors [259]. Hence, metabolism itself could be a checkpoint that limits immune-mediated tumor destruction [257]. Glutamine metabolism inhibitors have not only improve the immunosuppressive microenvironment, but also effectively reversed treatment resistance to ICI when combined with an ICI [260]. Therefore, combining metabolic inhibitors with checkpoint inhibitors is expected to improve the efficacy of future treatment landscape of lymphoma. Moreover, metabolic reprogramming is an important characteristic of immune cell activation. Accumulation of lactic acid from tumor glycolysis drives macrophages toward the M2 phenotype, which promotes tumor progression [251]. Signal regulatory protein alpha (SIRPα), expressed on macrophages, triggers a “don’t eat me” signal by interacting with the cluster of differentiation 47 (CD47) and further enables immune evasion for lymphoma cells [261]. Antibodies, fusion proteins, small molecule inhibitors, and peptides targeting the CD47-SIRPα axis, together with PD-1(IBI322, HX009), might be promising immunotherapeutic strategies in future lymphoma therapy [262]. A better comprehension of the clinical relevance of metabolic process and microenvironment would help us overcome the barriers to lymphoma treatment and achieve better efficacy and prognosis.

Author Contributions

Y.P., T.L., Z.Y.X.-M. and K.H.Y. designed, wrote, and edited the article; all authors contributed vital strategies, participated in discussions, provided scientific input, and proved the manuscript; K.H.Y. is the corresponding author of the manuscript; Y.P. and T.L. contributed equally to this manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the National Cancer Institute/National Institutes of Health (R01CA233490 and R01CA187415 to KHY). K.H.Y. is supported by Institutional Research and Development Fund and a Gundersen Lutheran Medical Foundation Award.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

K.H.Y. received research support from Roche Molecular System, Gilead Sciences Pharmaceutical, Seattle Genetics, Dai Sanyo Pharmaceutical, Adaptive Biotechnology, Incyte Pharmaceutical, and HTG Molecular Diagnostics.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Nabi, K.; Le, A. The Intratumoral Heterogeneity of Cancer Metabolism. Adv. Exp. Med. Biol. 2021, 1311, 149–160. [Google Scholar] [PubMed]
  2. DeBerardinis, R.J.; Lum, J.J.; Hatzivassiliou, G.; Thompson, C.B. The biology of cancer: Metabolic reprogramming fuels cell growth and proliferation. Cell Metab. 2008, 7, 11–20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Stine, Z.E.; Schug, Z.T.; Salvino, J.M.; Dang, C.V. Targeting cancer metabolism in the era of precision oncology. Nat. Rev. Drug Discov. 2022, 21, 141–162. [Google Scholar] [CrossRef]
  4. Stine, Z.E.; Walton, Z.E.; Altman, B.J.; Hsieh, A.L.; Dang, C.V. MYC, Metabolism and Cancer. Cancer Discov. 2015, 5, 1024–1039. [Google Scholar] [CrossRef] [Green Version]
  5. Brunner, J.S.; Finley, L.W.S. Metabolic determinants of tumour initiation. Nat. Rev. Endocrinol. 2022, 19, 134–150. [Google Scholar] [CrossRef]
  6. 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. [Google Scholar] [CrossRef]
  7. Hanahan, D.; Weinberg, R.A. Hallmarks of cancer: The next generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef] [Green Version]
  8. Pavlova, N.N.; Zhu, J.; Thompson, C.B. The hallmarks of cancer metabolism: Still emerging. Cell Metab. 2022, 34, 355–377. [Google Scholar] [CrossRef]
  9. Cheson, B.D.; Fisher, R.I.; Barrington, S.F.; Cavalli, F.; Schwartz, L.H.; Zucca, E.; Lister, T.A. Recommendations for initial evaluation, staging, and response assessment of Hodgkin and non-Hodgkin lymphoma: The Lugano classification. J. Clin. Oncol. 2014, 32, 3059–3068. [Google Scholar] [CrossRef]
  10. Ricard, F.; Barrington, S.; Korn, R.; Brueggenwerth, G.; Trotman, J.; Cheson, B.; Salles, G.; Schwartz, L.; Goldmacher, G.; Jarecha, R.; et al. Application of the lugano classification for initial evaluation, staging, and response assessment of hodgkin and non-hodgkin lymphoma: The prolog consensus initiative (part 2-technical). J. Nucl. Med. 2022, 64, 239–243. [Google Scholar] [CrossRef] [PubMed]
  11. Ricard, F.; Cheson, B.; Barrington, S.; Trotman, J.; Schmid, A.; Brueggenwerth, G.; Salles, G.; Schwartz, L.; Goldmacher, G.; Jarecha, R.; et al. Application of the lugano classification for initial evaluation, staging, and response assessment of hodgkin and non-hodgkin lymphoma: The prolog consensus initiative (part 1-clinical). J. Nucl. Med. 2022, 64, 102–108. [Google Scholar] [CrossRef]
  12. Calvo-Vidal, M.N.; Cerchietti, L. The metabolism of lymphomas. Curr. Opin. Hematol. 2013, 20, 345–354. [Google Scholar] [CrossRef]
  13. Maxwell, S.A.; Mousavi-Fard, S. Non-Hodgkin’s B-cell lymphoma: Advances in molecular strategies targeting drug resistance. Exp. Biol. Med. 2013, 238, 971–990. [Google Scholar] [CrossRef] [PubMed]
  14. Cho, S.F.; Chang, C.C.; Liu, Y.C.; Chang, C.S.; Hsiao, H.H.; Liu, T.C.; Huang, C.T.; Lin, S.F. Utilization of 18F-FDG PET/CT as a staging tool in patients with newly diagnosed lymphoma. Kaohsiung J. Med. Sci. 2015, 31, 130–137. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Chihara, D.; Oki, Y.; Onoda, H.; Taji, H.; Yamamoto, K.; Tamaki, T.; Morishima, Y. High maximum standard uptake value (SUVmax) on PET scan is associated with shorter survival in patients with diffuse large B cell lymphoma. Int. J. Hematol. 2011, 93, 502–508. [Google Scholar] [CrossRef]
  16. Zaucha, J.M.; Chauvie, S.; Zaucha, R.; Biggii, A.; Gallamini, A. The role of PET/CT in the modern treatment of Hodgkin lymphoma. Cancer Treat. Rev. 2019, 77, 44–56. [Google Scholar] [CrossRef] [PubMed]
  17. Shim, H.K.; Lee, W.W.; Park, S.Y.; Kim, H.; So, Y.; Kim, S.E. Expressions of glucose transporter Types 1 and 3 and hexokinase-II in diffuse large B-cell lymphoma and other B-cell non-Hodgkin’s lymphomas. Nucl. Med. Biol. 2009, 36, 191–197. [Google Scholar] [CrossRef] [PubMed]
  18. Takahashi, Y.; Akahane, T.; Yamamoto, D.; Nakamura, H.; Sawa, H.; Nitta, K.; Ide, W.; Hashimoto, I.; Kamada, H. Correlation between positron emission tomography findings and glucose transporter 1, 3 and L-type amino acid transporter 1 mRNA expression in primary central nervous system lymphomas. Mol. Clin. Oncol. 2014, 2, 525–529. [Google Scholar] [CrossRef] [Green Version]
  19. Koh, Y.W.; Han, J.H.; Park, S.Y.; Yoon, D.H.; Suh, C.; Huh, J. GLUT1 as a Prognostic Factor for Classical Hodgkin’s Lymphoma: Correlation with PD-L1 and PD-L2 Expression. J. Pathol. Transl. Med. 2017, 51, 152–158. [Google Scholar] [CrossRef] [Green Version]
  20. 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. [Google Scholar] [CrossRef] [Green Version]
  21. Lee, S.C.; Marzec, M.; Liu, X.; Wehrli, S.; Kantekure, K.; Ragunath, P.N.; Nelson, D.S.; Delikatny, E.J.; Glickson, J.D.; Wasik, M.A. Decreased lactate concentration and glycolytic enzyme expression reflect inhibition of mTOR signal transduction pathway in B-cell lymphoma. NMR Biomed. 2013, 26, 106–114. [Google Scholar] [CrossRef] [PubMed]
  22. Chiche, J.; Pommier, S.; Beneteau, M.; Mondragón, L.; Meynet, O.; Zunino, B.; Mouchotte, A.; Verhoeyen, E.; Guyot, M.; Pagès, G.; et al. GAPDH enhances the aggressiveness and the vascularization of non-Hodgkin’s B lymphomas via NF-κB-dependent induction of HIF-1α. Leukemia 2015, 29, 1163–1176. [Google Scholar] [CrossRef] [PubMed]
  23. Nakajima, K.; Kawashima, I.; Koshiisi, M.; Kumagai, T.; Suzuki, M.; Suzuki, J.; Mitsumori, T.; Kirito, K. Glycolytic enzyme hexokinase II is a putative therapeutic target in B-cell malignant lymphoma. Exp. Hematol. 2019, 78, 46–55.e43. [Google Scholar] [CrossRef]
  24. Bhalla, K.; Jaber, S.; Nahid, M.N.; Underwood, K.; Beheshti, A.; Landon, A.; Bhandary, B.; Bastian, P.; Evens, A.M.; Haley, J.; et al. Role of hypoxia in Diffuse Large B-cell Lymphoma: Metabolic repression and selective translation of HK2 facilitates development of DLBCL. Sci. Rep. 2018, 8, 744. [Google Scholar] [CrossRef] [Green Version]
  25. Liu, W.P.; Song, Y.Q.; Zheng, W.; Wang, X.P.; Ding, N.; Zhu, J. Aggressive behavior and elevated lactate dehydrogenase at baseline confer inferior prognosis in patients with primary cutaneous lymphoma. Clin. Lymphoma Myeloma Leuk. 2013, 13, 534–540. [Google Scholar] [CrossRef]
  26. Lu, R.; Jiang, M.; Chen, Z.; Xu, X.; Hu, H.; Zhao, X.; Gao, X.; Guo, L. Lactate dehydrogenase 5 expression in Non-Hodgkin lymphoma is associated with the induced hypoxia regulated protein and poor prognosis. PLoS ONE 2013, 8, e74853. [Google Scholar] [CrossRef]
  27. Gooptu, M.; Whitaker-Menezes, D.; Sprandio, J.; Domingo-Vidal, M.; Lin, Z.; Uppal, G.; Gong, J.; Fratamico, R.; Leiby, B.; Dulau-Florea, A.; et al. Mitochondrial and glycolytic metabolic compartmentalization in diffuse large B-cell lymphoma. Semin. Oncol. 2017, 44, 204–217. [Google Scholar] [CrossRef]
  28. Doherty, J.R.; Cleveland, J.L. Targeting lactate metabolism for cancer therapeutics. J. Clin. Investig. 2013, 123, 3685–3692. [Google Scholar] [CrossRef]
  29. Parks, S.K.; Chiche, J.; Pouysségur, J. Disrupting proton dynamics and energy metabolism for cancer therapy. Nat. Rev. Cancer 2013, 13, 611–623. [Google Scholar] [CrossRef] [PubMed]
  30. Le Floch, R.; Chiche, J.; Marchiq, I.; Naiken, T.; Ilc, K.; Murray, C.M.; Critchlow, S.E.; Roux, D.; Simon, M.P.; Pouysségur, J. CD147 subunit of lactate/H+ symporters MCT1 and hypoxia-inducible MCT4 is critical for energetics and growth of glycolytic tumors. Proc. Natl. Acad. Sci. USA 2011, 108, 16663–16668. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Baenke, F.; Dubuis, S.; Brault, C.; Weigelt, B.; Dankworth, B.; Griffiths, B.; Jiang, M.; Mackay, A.; Saunders, B.; Spencer-Dene, B.; et al. Functional screening identifies MCT4 as a key regulator of breast cancer cell metabolism and survival. J. Pathol. 2015, 237, 152–165. [Google Scholar] [CrossRef] [PubMed]
  32. Ducker, G.S.; Ghergurovich, J.M.; Mainolfi, N.; Suri, V.; Jeong, S.K.; Hsin-Jung Li, S.; Friedman, A.; Manfredi, M.G.; Gitai, Z.; Kim, H.; et al. Human SHMT inhibitors reveal defective glycine import as a targetable metabolic vulnerability of diffuse large B-cell lymphoma. Proc. Natl. Acad. Sci. USA 2017, 114, 11404–11409. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Nam, S.J.; Kim, S.; Paik, J.H.; Kim, T.M.; Heo, D.S.; Kim, C.W.; Jeon, Y.K. An increase in indoleamine 2,3-dioxygenase-positive cells in the tumor microenvironment predicts favorable prognosis in patients with diffuse large B-cell lymphoma treated with rituximab, cyclophosphamide, doxorubicin, vincristine, and prednisolone. Leuk. Lymphoma 2016, 57, 1956–1960. [Google Scholar] [CrossRef]
  34. Choe, J.Y.; Yun, J.Y.; Jeon, Y.K.; Kim, S.H.; Park, G.; Huh, J.R.; Oh, S.; Kim, J.E. Indoleamine 2,3-dioxygenase (IDO) is frequently expressed in stromal cells of Hodgkin lymphoma and is associated with adverse clinical features: A retrospective cohort study. BMC Cancer 2014, 14, 335. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Dashnamoorthy, R.; Abermil, N.; Behesti, A.; Kozlowski, P.; Lansigan, F.; Kinlaw, W.B.; Gartenhaus, R.; Jones, G.; Hlatky, L.; Evens, A.M. The Lipid Addiction of Diffuse Large B-Cell Lymphoma (DLBCL) and Potential Treatment Strategies with Novel Fatty Acid Synthase (FASN) Small Molecule Inhibitors. Blood 2014, 124, 4490. [Google Scholar] [CrossRef]
  36. Argyriou, P.; Papageorgiou, S.G.; Panteleon, V.; Psyrri, A.; Bakou, V.; Pappa, V.; Spathis, A.; Economopoulou, P.; Papageorgiou, E.; Economopoulos, T.; et al. Hypoxia-inducible factors in mantle cell lymphoma: Implication for an activated mTORC1→HIF-1α pathway. Ann. Hematol. 2011, 90, 315–322. [Google Scholar] [CrossRef]
  37. Sander, S.; Calado, D.P.; Srinivasan, L.; Köchert, K.; Zhang, B.; Rosolowski, M.; Rodig, S.J.; Holzmann, K.; Stilgenbauer, S.; Siebert, R.; et al. Synergy between PI3K signaling and MYC in Burkitt lymphomagenesis. Cancer Cell 2012, 22, 167–179. [Google Scholar] [CrossRef] [Green Version]
  38. Schmitz, R.; Young, R.M.; Ceribelli, M.; Jhavar, S.; Xiao, W.; Zhang, M.; Wright, G.; Shaffer, A.L.; Hodson, D.J.; Buras, E.; et al. Burkitt lymphoma pathogenesis and therapeutic targets from structural and functional genomics. Nature 2012, 490, 116–120. [Google Scholar] [CrossRef] [Green Version]
  39. Yamamoto-Sugitani, M.; Kuroda, J.; Shimura, Y.; Nagoshi, H.; Chinen, Y.; Ohshiro, M.; Mizutani, S.; Kiyota, M.; Nakayama, R.; Kobayashi, T.; et al. Comprehensive cytogenetic study of primary cutaneous gamma-delta T-cell lymphoma by means of spectral karyotyping and genome-wide single nucleotide polymorphism array. Cancer Genet. 2012, 205, 459–464. [Google Scholar] [CrossRef]
  40. Hwang, S.R.; Murga-Zamalloa, C.; Brown, N.; Basappa, J.; McDonnell, S.R.; Mendoza-Reinoso, V.; Basrur, V.; Wilcox, R.; Elenitoba-Johnson, K.; Lim, M.S. Pyrimidine tract-binding protein 1 mediates pyruvate kinase M2-dependent phosphorylation of signal transducer and activator of transcription 3 and oncogenesis in anaplastic large cell lymphoma. Lab. Investig. 2017, 97, 962–970. [Google Scholar] [CrossRef] [Green Version]
  41. Liu, W.; Le, A.; Hancock, C.; Lane, A.N.; Dang, C.V.; Fan, T.W.; Phang, J.M. Reprogramming of proline and glutamine metabolism contributes to the proliferative and metabolic responses regulated by oncogenic transcription factor c-MYC. Proc. Natl. Acad. Sci. USA 2012, 109, 8983–8988. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Xiong, J.; Bian, J.; Wang, L.; Zhou, J.Y.; Wang, Y.; Zhao, Y.; Wu, L.L.; Hu, J.J.; Li, B.; Chen, S.J.; et al. Dysregulated choline metabolism in T-cell lymphoma: Role of choline kinase-α and therapeutic targeting. Blood Cancer J. 2015, 5, 287. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Kusao, I.; Troelstrup, D.; Shiramizu, B. Possible Mitochondria-Associated Enzymatic Role in Non-Hodgkin Lymphoma Residual Disease. Cancer Growth Metastasis 2008, 1, 3–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Lemonnier, F.; Cairns, R.A.; Inoue, S.; Li, W.Y.; Dupuy, A.; Broutin, S.; Martin, N.; Fataccioli, V.; Pelletier, R.; Wakeham, A.; et al. The IDH2 R172K mutation associated with angioimmunoblastic T-cell lymphoma produces 2HG in T cells and impacts lymphoid development. Proc. Natl. Acad. Sci. USA 2016, 113, 15084–15089. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Wang, C.; McKeithan, T.W.; Gong, Q.; Zhang, W.; Bouska, A.; Rosenwald, A.; Gascoyne, R.D.; Wu, X.; Wang, J.; Muhammad, Z.; et al. IDH2R172 mutations define a unique subgroup of patients with angioimmunoblastic T-cell lymphoma. Blood 2015, 126, 1741–1752. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Lin, A.P.; Abbas, S.; Kim, S.W.; Ortega, M.; Bouamar, H.; Escobedo, Y.; Varadarajan, P.; Qin, Y.; Sudderth, J.; Schulz, E.; et al. D2HGDH regulates alpha-ketoglutarate levels and dioxygenase function by modulating IDH2. Nat. Commun. 2015, 6, 7768. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Le, A.; Lane, A.N.; Hamaker, M.; Bose, S.; Gouw, A.; Barbi, J.; Tsukamoto, T.; Rojas, C.J.; Slusher, B.S.; Zhang, H.; et al. Glucose-independent glutamine metabolism via TCA cycling for proliferation and survival in B cells. Cell Metab. 2012, 15, 110–121. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. 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. [Google Scholar] [CrossRef] [Green Version]
  49. Zaugg, K.; Yao, Y.; Reilly, P.T.; Kannan, K.; Kiarash, R.; Mason, J.; Huang, P.; Sawyer, S.K.; Fuerth, B.; Faubert, B.; et al. Carnitine palmitoyltransferase 1C promotes cell survival and tumor growth under conditions of metabolic stress. Genes Dev. 2011, 25, 1041–1051. [Google Scholar] [CrossRef] [Green Version]
  50. Marin-Valencia, I.; Yang, C.; Mashimo, T.; Cho, S.; Baek, H.; Yang, X.L.; Rajagopalan, K.N.; Maddie, M.; Vemireddy, V.; Zhao, Z.; et al. Analysis of tumor metabolism reveals mitochondrial glucose oxidation in genetically diverse human glioblastomas in the mouse brain in vivo. Cell Metab. 2012, 15, 827–837. [Google Scholar] [CrossRef] [Green Version]
  51. Moreno-Sánchez, R.; Rodríguez-Enríquez, S.; Saavedra, E.; Marín-Hernández, A.; Gallardo-Pérez, J.C. The bioenergetics of cancer: Is glycolysis the main ATP supplier in all tumor cells? Biofactors 2009, 35, 209–225. [Google Scholar] [CrossRef] [PubMed]
  52. Meynet, O.; Bénéteau, M.; Jacquin, M.A.; Pradelli, L.A.; Cornille, A.; Carles, M.; Ricci, J.E. Glycolysis inhibition targets Mcl-1 to restore sensitivity of lymphoma cells to ABT-737-induced apoptosis. Leukemia 2012, 26, 1145–1147. [Google Scholar] [CrossRef] [Green Version]
  53. Monti, S.; Savage, K.J.; Kutok, J.L.; Feuerhake, F.; Kurtin, P.; Mihm, M.; Wu, B.; Pasqualucci, L.; Neuberg, D.; Aguiar, R.C.; et al. Molecular profiling of diffuse large B-cell lymphoma identifies robust subtypes including one characterized by host inflammatory response. Blood 2005, 105, 1851–1861. [Google Scholar] [CrossRef] [Green Version]
  54. Alizadeh, A.A.; Eisen, M.B.; Davis, R.E.; Ma, C.; Lossos, I.S.; Rosenwald, A.; Boldrick, J.C.; Sabet, H.; Tran, T.; Yu, X.; et al. Distinct types of diffuse large B-cell lymphoma identified by gene expression profiling. Nature 2000, 403, 503–511. [Google Scholar] [CrossRef] [PubMed]
  55. Sha, C.; Barrans, S.; Cucco, F.; Bentley, M.A.; Care, M.A.; Cummin, T.; Kennedy, H.; Thompson, J.S.; Uddin, R.; Worrillow, L.; et al. Molecular High-Grade B-Cell Lymphoma: Defining a Poor-Risk Group That Requires Different Approaches to Therapy. J. Clin. Oncol. 2019, 37, 202–212. [Google Scholar] [CrossRef]
  56. Norberg, E.; Lako, A.; Chen, P.H.; Stanley, I.A.; Zhou, F.; Ficarro, S.B.; Chapuy, B.; Chen, L.; Rodig, S.; Shin, D.; et al. Differential contribution of the mitochondrial translation pathway to the survival of diffuse large B-cell lymphoma subsets. Cell Death Differ. 2017, 24, 251–262. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Caro, P.; Kishan, A.U.; Norberg, E.; Stanley, I.A.; Chapuy, B.; Ficarro, S.B.; Polak, K.; Tondera, D.; Gounarides, J.; Yin, H.; et al. Metabolic signatures uncover distinct targets in molecular subsets of diffuse large B cell lymphoma. Cancer Cell 2012, 22, 547–560. [Google Scholar] [CrossRef] [Green Version]
  58. Monti, S.; Chapuy, B.; Takeyama, K.; Rodig, S.J.; Hao, Y.; Yeda, K.T.; Inguilizian, H.; Mermel, C.; Currie, T.; Dogan, A.; et al. Integrative analysis reveals an outcome-associated and targetable pattern of p53 and cell cycle deregulation in diffuse large B cell lymphoma. Cancer Cell 2012, 22, 359–372. [Google Scholar] [CrossRef] [Green Version]
  59. Chen, L.; Monti, S.; Juszczynski, P.; Daley, J.; Chen, W.; Witzig, T.E.; Habermann, T.M.; Kutok, J.L.; Shipp, M.A. SYK-dependent tonic B-cell receptor signaling is a rational treatment target in diffuse large B-cell lymphoma. Blood 2008, 111, 2230–2237. [Google Scholar] [CrossRef] [Green Version]
  60. DeBerardinis, R.J.; Mancuso, A.; Daikhin, E.; Nissim, I.; Yudkoff, M.; Wehrli, S.; Thompson, C.B. Beyond aerobic glycolysis: Transformed cells can engage in glutamine metabolism that exceeds the requirement for protein and nucleotide synthesis. Proc. Natl. Acad. Sci. USA 2007, 104, 19345–19350. [Google Scholar] [CrossRef] [Green Version]
  61. Wishart, D.S. Metabolomics for Investigating Physiological and Pathophysiological Processes. Physiol. Rev. 2019, 99, 1819–1875. [Google Scholar] [CrossRef] [PubMed]
  62. Yang, Y.; Fan, T.W.; Lane, A.N.; Higashi, R.M. Chloroformate derivatization for tracing the fate of Amino acids in cells and tissues by multiple stable isotope resolved metabolomics (mSIRM). Anal. Chim. Acta 2017, 976, 63–73. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Gao, P.; Tchernyshyov, I.; Chang, T.C.; Lee, Y.S.; Kita, K.; Ochi, T.; Zeller, K.I.; De Marzo, A.M.; Van Eyk, J.E.; Mendell, J.T.; et al. C-Myc suppression of miR-23a/b enhances mitochondrial glutaminase expression and glutamine metabolism. Nature 2009, 458, 762–765. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Wise, D.R.; DeBerardinis, R.J.; Mancuso, A.; Sayed, N.; Zhang, X.Y.; Pfeiffer, H.K.; Nissim, I.; Daikhin, E.; Yudkoff, M.; McMahon, S.B.; et al. Myc regulates a transcriptional program that stimulates mitochondrial glutaminolysis and leads to glutamine addiction. Proc. Natl. Acad. Sci. USA 2008, 105, 18782–18787. [Google Scholar] [CrossRef] [Green Version]
  65. DeBerardinis, R.J.; Cheng, T. Q’s next: The diverse functions of glutamine in metabolism, cell biology and cancer. Oncogene 2010, 29, 313–324. [Google Scholar] [CrossRef] [Green Version]
  66. Wise, D.R.; Thompson, C.B. Glutamine addiction: A new therapeutic target in cancer. Trends Biochem. Sci. 2010, 35, 427–433. [Google Scholar] [CrossRef] [Green Version]
  67. Altman, B.J.; Stine, Z.E.; Dang, C.V. From Krebs to clinic: Glutamine metabolism to cancer therapy. Nat. Rev. Cancer 2016, 16, 619–634. [Google Scholar] [CrossRef] [Green Version]
  68. Wei, P.; Bott, A.J.; Cluntun, A.A.; Morgan, J.T.; Cunningham, C.N.; Schell, J.C.; Ouyang, Y.; Ficarro, S.B.; Marto, J.A.; Danial, N.N.; et al. Mitochondrial pyruvate supports lymphoma proliferation by fueling a glutamate pyruvate transaminase 2-dependent glutaminolysis pathway. Sci. Adv. 2022, 8, eabq0117. [Google Scholar] [CrossRef]
  69. Weinberg, F.; Hamanaka, R.; Wheaton, W.W.; Weinberg, S.; Joseph, J.; Lopez, M.; Kalyanaraman, B.; Mutlu, G.M.; Budinger, G.R.; Chandel, N.S. Mitochondrial metabolism and ROS generation are essential for Kras-mediated tumorigenicity. Proc. Natl. Acad. Sci. USA 2010, 107, 8788–8793. [Google Scholar] [CrossRef] [Green Version]
  70. Cruzat, V.; Macedo Rogero, M.; Noel Keane, K.; Curi, R.; Newsholme, P. Glutamine: Metabolism and Immune Function, Supplementation and Clinical Translation. Nutrients 2018, 10, 1564. [Google Scholar] [CrossRef] [Green Version]
  71. Li, L.; Nie, L.; Jordan, A.; Cai, Q.; Liu, Y.; Li, Y.; Che, Y.; Vargas, J.; Chen, Z.; Leeming, A.; et al. Targeting glutaminase is therapeutically effective in ibrutinib-resistant mantle cell lymphoma. Haematologica 2022. [Google Scholar] [CrossRef]
  72. Zimta, A.A.; Cenariu, D.; Irimie, A.; Magdo, L.; Nabavi, S.M.; Atanasov, A.G.; Berindan-Neagoe, I. The Role of Nrf2 Activity in Cancer Development and Progression. Cancers 2019, 11, 1755. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Yi, X.; Zhao, Y.; Xue, L.; Zhang, J.; Qiao, Y.; Jin, Q.; Li, H. Expression of Keap1 and Nrf2 in diffuse large B-cell lymphoma and its clinical significance. Exp. Ther. Med. 2018, 16, 573–578. [Google Scholar] [CrossRef] [Green Version]
  74. Mukhopadhyay, S.; Goswami, D.; Adiseshaiah, P.P.; Burgan, W.; Yi, M.; Guerin, T.M.; Kozlov, S.V.; Nissley, D.V.; McCormick, F. Undermining Glutaminolysis Bolsters Chemotherapy While NRF2 Promotes Chemoresistance in KRAS-Driven Pancreatic Cancers. Cancer Res. 2020, 80, 1630–1643. [Google Scholar] [CrossRef] [Green Version]
  75. Rushworth, S.A.; Macewan, D.J. The role of nrf2 and cytoprotection in regulating chemotherapy resistance of human leukemia cells. Cancers 2011, 3, 1605–1621. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Xiang, Y.; Stine, Z.E.; Xia, J.; Lu, Y.; O’Connor, R.S.; Altman, B.J.; Hsieh, A.L.; Gouw, A.M.; Thomas, A.G.; Gao, P.; et al. Targeted inhibition of tumor-specific glutaminase diminishes cell-autonomous tumorigenesis. J. Clin. Investig. 2015, 125, 2293–2306. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Ganguly, K.; Bhatia, R.; Rauth, S.; Kisling, A.; Atri, P.; Thompson, C.; Vengoji, R.; Ram Krishn, S.; Shinde, D.; Thomas, V.; et al. Mucin 5AC Serves as the Nexus for β-Catenin/c-Myc Interplay to Promote Glutamine Dependency During Pancreatic Cancer Chemoresistance. Gastroenterology 2022, 162, 253–268.e213. [Google Scholar] [CrossRef] [PubMed]
  78. Locasale, J.W. Serine, glycine and one-carbon units: Cancer metabolism in full circle. Nat. Rev. Cancer 2013, 13, 572–583. [Google Scholar] [CrossRef] [Green Version]
  79. Maddocks, O.D.K.; Athineos, D.; Cheung, E.C.; Lee, P.; Zhang, T.; van den Broek, N.J.F.; Mackay, G.M.; Labuschagne, C.F.; Gay, D.; Kruiswijk, F.; et al. Modulating the therapeutic response of tumours to dietary serine and glycine starvation. Nature 2017, 544, 372–376. [Google Scholar] [CrossRef] [Green Version]
  80. Maliniemi, P.; Laukkanen, K.; Väkevä, L.; Dettmer, K.; Lipsanen, T.; Jeskanen, L.; Bessede, A.; Oefner, P.J.; Kadin, M.E.; Ranki, A. Biological and clinical significance of tryptophan-catabolizing enzymes in cutaneous T-cell lymphomas. Oncoimmunology 2017, 6, e1273310. [Google Scholar] [CrossRef] [Green Version]
  81. Masaki, A.; Ishida, T.; Maeda, Y.; Suzuki, S.; Ito, A.; Takino, H.; Ogura, H.; Totani, H.; Yoshida, T.; Kinoshita, S.; et al. Prognostic Significance of Tryptophan Catabolism in Adult T-cell Leukemia/Lymphoma. Clin. Cancer Res. 2015, 21, 2830–2839. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Ninomiya, S.; Narala, N.; Huye, L.; Yagyu, S.; Savoldo, B.; Dotti, G.; Heslop, H.E.; Brenner, M.K.; Rooney, C.M.; Ramos, C.A. Tumor indoleamine 2,3-dioxygenase (IDO) inhibits CD19-CAR T cells and is downregulated by lymphodepleting drugs. Blood 2015, 125, 3905–3916. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Kapadia, B.; Nanaji, N.M.; Bhalla, K.; Bhandary, B.; Lapidus, R.; Beheshti, A.; Evens, A.M.; Gartenhaus, R.B. Fatty Acid Synthase induced S6Kinase facilitates USP11-eIF4B complex formation for sustained oncogenic translation in DLBCL. Nat. Commun. 2018, 9, 829. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Uddin, S.; Hussain, A.R.; Ahmed, M.; Bu, R.; Ahmed, S.O.; Ajarim, D.; Al-Dayel, F.; Bavi, P.; Al-Kuraya, K.S. Inhibition of fatty acid synthase suppresses c-Met receptor kinase and induces apoptosis in diffuse large B-cell lymphoma. Mol. Cancer Ther. 2010, 9, 1244–1255. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Bhatt, A.P.; Jacobs, S.R.; Freemerman, A.J.; Makowski, L.; Rathmell, J.C.; Dittmer, D.P.; Damania, B. Dysregulation of fatty acid synthesis and glycolysis in non-Hodgkin lymphoma. Proc. Natl. Acad. Sci. USA 2012, 109, 11818–11823. [Google Scholar] [CrossRef] [Green Version]
  86. Gelebart, P.; Zak, Z.; Anand, M.; Belch, A.; Lai, R. Blockade of fatty acid synthase triggers significant apoptosis in mantle cell lymphoma. PLoS ONE 2012, 7, e33738. [Google Scholar] [CrossRef] [Green Version]
  87. Dashnamoorthy, R.; Lansigan, F.; Davis, W.L.; Kinlaw, W.B., III; Gartenhaus, R., III; Evens, A.M. Targeting The Interactions Of Fatty Acid Metabolism With PI3K/mTOR and MAPK As a Novel Therapeutic Strategy In Diffuse Large B-Cell Lymphoma (DLBCL). Blood 2013, 122, 5133. [Google Scholar] [CrossRef]
  88. Röhrig, F.; Schulze, A. The multifaceted roles of fatty acid synthesis in cancer. Nat. Rev. Cancer 2016, 16, 732–749. [Google Scholar] [CrossRef]
  89. Spaner, D.E.; Lee, E.; Shi, Y.; Wen, F.; Li, Y.; Tung, S.; McCaw, L.; Wong, K.; Gary-Gouy, H.; Dalloul, A.; et al. PPAR-alpha is a therapeutic target for chronic lymphocytic leukemia. Leukemia 2013, 27, 1090–1099. [Google Scholar] [CrossRef] [Green Version]
  90. Harmon, G.S.; Lam, M.T.; Glass, C.K. PPARs and lipid ligands in inflammation and metabolism. Chem. Rev. 2011, 111, 6321–6340. [Google Scholar] [CrossRef] [Green Version]
  91. Wang, X.; Wang, G.; Shi, Y.; Sun, L.; Gorczynski, R.; Li, Y.J.; Xu, Z.; Spaner, D.E. PPAR-delta promotes survival of breast cancer cells in harsh metabolic conditions. Oncogenesis 2016, 5, e232. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Khozoie, C.; Borland, M.G.; Zhu, B.; Baek, S.; John, S.; Hager, G.L.; Shah, Y.M.; Gonzalez, F.J.; Peters, J.M. Analysis of the peroxisome proliferator-activated receptor-β/δ (PPARβ/δ) cistrome reveals novel co-regulatory role of ATF4. BMC Genom. 2012, 13, 665. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Sun, L.; Shi, Y.; Wang, G.; Wang, X.; Zeng, S.; Dunn, S.E.; Fairn, G.D.; Li, Y.J.; Spaner, D.E. PPAR-delta modulates membrane cholesterol and cytokine signaling in malignant B cells. Leukemia 2018, 32, 184–193. [Google Scholar] [CrossRef]
  94. Rink, J.S.; Yang, S.; Cen, O.; Taxter, T.; McMahon, K.M.; Misener, S.; Behdad, A.; Longnecker, R.; Gordon, L.I.; Thaxton, C.S. Rational Targeting of Cellular Cholesterol in Diffuse Large B-Cell Lymphoma (DLBCL) Enabled by Functional Lipoprotein Nanoparticles: A Therapeutic Strategy Dependent on Cell of Origin. Mol. Pharm. 2017, 14, 4042–4051. [Google Scholar] [CrossRef] [PubMed]
  95. Glunde, K.; Bhujwalla, Z.M.; Ronen, S.M. Choline metabolism in malignant transformation. Nat. Rev. Cancer 2011, 11, 835–848. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Arlauckas, S.P.; Popov, A.V.; Delikatny, E.J. Choline kinase alpha-Putting the ChoK-hold on tumor metabolism. Prog. Lipid Res. 2016, 63, 28–40. [Google Scholar] [CrossRef] [Green Version]
  97. Kierans, S.J.; Taylor, C.T. Regulation of glycolysis by the hypoxia-inducible factor (HIF): Implications for cellular physiology. J. Physiol. 2021, 599, 23–37. [Google Scholar] [CrossRef]
  98. Viola, A.; Munari, F.; Sánchez-Rodríguez, R.; Scolaro, T.; Castegna, A. The Metabolic Signature of Macrophage Responses. Front. Immunol. 2019, 10, 1462. [Google Scholar] [CrossRef] [Green Version]
  99. Gonzalez-King, H.; García, N.A.; Ontoria-Oviedo, I.; Ciria, M.; Montero, J.A.; Sepúlveda, P. Hypoxia Inducible Factor-1α Potentiates Jagged 1-Mediated Angiogenesis by Mesenchymal Stem Cell-Derived Exosomes. Stem Cells 2017, 35, 1747–1759. [Google Scholar] [CrossRef] [Green Version]
  100. Eyrich, N.W.; Potts, C.R.; Robinson, M.H.; Maximov, V.; Kenney, A.M. Reactive Oxygen Species Signaling Promotes Hypoxia-Inducible Factor 1α Stabilization in Sonic Hedgehog-Driven Cerebellar Progenitor Cell Proliferation. Mol. Cell Biol. 2019, 39, e00268-18. [Google Scholar] [CrossRef] [Green Version]
  101. Majumder, P.K.; Febbo, P.G.; Bikoff, R.; Berger, R.; Xue, Q.; McMahon, L.M.; Manola, J.; Brugarolas, J.; McDonnell, T.J.; Golub, T.R.; 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. [Google Scholar] [CrossRef] [PubMed]
  102. Laughner, E.; Taghavi, P.; Chiles, K.; Mahon, P.C.; Semenza, G.L. HER2 (neu) signaling increases the rate of hypoxia-inducible factor 1alpha (HIF-1alpha) synthesis: Novel mechanism for HIF-1-mediated vascular endothelial growth factor expression. Mol. Cell Biol. 2001, 21, 3995–4004. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Fukuda, R.; Hirota, K.; Fan, F.; Jung, Y.D.; Ellis, L.M.; Semenza, G.L. Insulin-like growth factor 1 induces hypoxia-inducible factor 1-mediated vascular endothelial growth factor expression, which is dependent on MAP kinase and phosphatidylinositol 3-kinase signaling in colon cancer cells. J. Biol. Chem. 2002, 277, 38205–38211. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Gao, P.; Zhang, H.; Dinavahi, R.; Li, F.; Xiang, Y.; Raman, V.; Bhujwalla, Z.M.; Felsher, D.W.; Cheng, L.; Pevsner, J.; et al. HIF-dependent antitumorigenic effect of antioxidants in vivo. Cancer Cell 2007, 12, 230–238. [Google Scholar] [CrossRef] [Green Version]
  105. Gupta, D.; Crosby, M.E.; Almasan, A.; Macklis, R.M. Regulation of CD20 expression by radiation-induced changes in intracellular redox status. Free Radic Biol. Med. 2008, 44, 614–623. [Google Scholar] [CrossRef] [Green Version]
  106. Evens, A.M.; Schumacker, P.T.; Helenowski, I.B.; Singh, A.T.; Dokic, D.; Keswani, A.; Kordeluk, E.; Raji, A.; Winter, J.N.; Jovanovic, B.D.; et al. Hypoxia inducible factor-alpha activation in lymphoma and relationship to the thioredoxin family. Br. J. Haematol. 2008, 141, 676–680. [Google Scholar] [CrossRef] [Green Version]
  107. Evens, A.M.; Sehn, L.H.; Farinha, P.; Nelson, B.P.; Raji, A.; Lu, Y.; Brakman, A.; Parimi, V.; Winter, J.N.; Schumacker, P.T.; et al. Hypoxia-inducible factor-1 {alpha} expression predicts superior survival in patients with diffuse large B-cell lymphoma treated with R-CHOP. J. Clin. Oncol. 2010, 28, 1017–1024. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Powell, J.R.; Dojcinov, S.; King, L.; Wosniak, S.; Gerry, S.; Casbard, A.; Bailey, H.; Gallop-Evans, E.; Maughan, T. Prognostic significance of hypoxia inducible factor-1α and vascular endothelial growth factor expression in patients with diffuse large B-cell lymphoma treated with rituximab. Leuk. Lymphoma 2013, 54, 959–966. [Google Scholar] [CrossRef]
  109. Darekar, S.; Georgiou, K.; Yurchenko, M.; Yenamandra, S.P.; Chachami, G.; Simos, G.; Klein, G.; Kashuba, E. Epstein-Barr virus immortalization of human B-cells leads to stabilization of hypoxia-induced factor 1 alpha, congruent with the Warburg effect. PLoS ONE 2012, 7, e42072. [Google Scholar] [CrossRef]
  110. Dang, C.V. MYC, metabolism, cell growth, and tumorigenesis. Cold Spring Harb Perspect. Med. 2013, 3, a014217. [Google Scholar] [CrossRef] [Green Version]
  111. Lewis, B.C.; Prescott, J.E.; Campbell, S.E.; Shim, H.; Orlowski, R.Z.; Dang, C.V. Tumor induction by the c-Myc target genes rcl and lactate dehydrogenase A. Cancer Res. 2000, 60, 6178–6183. [Google Scholar]
  112. Mushtaq, M.; Darekar, S.; Klein, G.; Kashuba, E. Different Mechanisms of Regulation of the Warburg Effect in Lymphoblastoid and Burkitt Lymphoma Cells. PLoS ONE 2015, 10, e0136142. [Google Scholar] [CrossRef] [PubMed]
  113. Martinengo, C.; Poggio, T.; Menotti, M.; Scalzo, M.S.; Mastini, C.; Ambrogio, C.; Pellegrino, E.; Riera, L.; Piva, R.; Ribatti, D.; et al. ALK-dependent control of hypoxia-inducible factors mediates tumor growth and metastasis. Cancer Res. 2014, 74, 6094–6106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Marzec, M.; Liu, X.; Wong, W.; Yang, Y.; Pasha, T.; Kantekure, K.; Zhang, P.; Woetmann, A.; Cheng, M.; Odum, N.; et al. Oncogenic kinase NPM/ALK induces expression of HIF1α mRNA. Oncogene 2011, 30, 1372–1378. [Google Scholar] [CrossRef] [PubMed]
  115. Shrestha, P.; Davis, D.A.; Veeranna, R.P.; Carey, R.F.; Viollet, C.; Yarchoan, R. Hypoxia-inducible factor-1 alpha as a therapeutic target for primary effusion lymphoma. PLoS Pathog. 2017, 13, e1006628. [Google Scholar] [CrossRef] [Green Version]
  116. Dang, C.V. MYC on the path to cancer. Cell 2012, 149, 22–35. [Google Scholar] [CrossRef] [Green Version]
  117. Sabò, A.; Kress, T.R.; Pelizzola, M.; de Pretis, S.; Gorski, M.M.; Tesi, A.; Morelli, M.J.; Bora, P.; Doni, M.; Verrecchia, A.; et al. Selective transcriptional regulation by Myc in cellular growth control and lymphomagenesis. Nature 2014, 511, 488–492. [Google Scholar] [CrossRef] [Green Version]
  118. Ott, G.; Rosenwald, A.; Campo, E. Understanding MYC-driven aggressive B-cell lymphomas: Pathogenesis and classification. Blood 2013, 122, 3884–3891. [Google Scholar] [CrossRef] [Green Version]
  119. Leucci, E.; Cocco, M.; Onnis, A.; De Falco, G.; van Cleef, P.; Bellan, C.; van Rijk, A.; Nyagol, J.; Byakika, B.; Lazzi, S.; et al. MYC translocation-negative classical Burkitt lymphoma cases: An alternative pathogenetic mechanism involving miRNA deregulation. J. Pathol. 2008, 216, 440–450. [Google Scholar] [CrossRef]
  120. Ruzinova, M.B.; Caron, T.; Rodig, S.J. Altered subcellular localization of c-Myc protein identifies aggressive B-cell lymphomas harboring a c-MYC translocation. Am. J. Surg. Pathol. 2010, 34, 882–891. [Google Scholar] [CrossRef]
  121. Calvo-Vidal, M.N.; Zamponi, N.; Krumsiek, J.; Stockslager, M.A.; Revuelta, M.V.; Phillip, J.M.; Marullo, R.; Tikhonova, E.; Kotlov, N.; Patel, J.; et al. Oncogenic HSP90 Facilitates Metabolic Alterations in Aggressive B-cell Lymphomas. Cancer Res. 2021, 81, 5202–5216. [Google Scholar] [CrossRef] [PubMed]
  122. Wang, C.; Li, D.; Zhang, L.; Jiang, S.; Liang, J.; Narita, Y.; Hou, I.; Zhong, Q.; Zheng, Z.; Xiao, H.; et al. RNA Sequencing Analyses of Gene Expression during Epstein-Barr Virus Infection of Primary B Lymphocytes. J. Virol. 2019, 93, e00226-19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Klein, G. Burkitt lymphoma—A stalking horse for cancer research? Semin. Cancer Biol. 2009, 19, 347–350. [Google Scholar] [CrossRef] [PubMed]
  124. Elenitoba-Johnson, K.S.; Jenson, S.D.; Abbott, R.T.; Palais, R.A.; Bohling, S.D.; Lin, Z.; Tripp, S.; Shami, P.J.; Wang, L.Y.; Coupland, R.W.; et al. Involvement of multiple signaling pathways in follicular lymphoma transformation: p38-mitogen-activated protein kinase as a target for therapy. Proc. Natl. Acad. Sci. USA 2003, 100, 7259–7264. [Google Scholar] [CrossRef] [Green Version]
  125. Denko, N.C. Hypoxia, HIF1 and glucose metabolism in the solid tumour. Nat. Rev. Cancer 2008, 8, 705–713. [Google Scholar] [CrossRef]
  126. Wang, R.; Dillon, C.P.; Shi, L.Z.; Milasta, S.; Carter, R.; Finkelstein, D.; McCormick, L.L.; Fitzgerald, P.; Chi, H.; Munger, J.; et al. The transcription factor Myc controls metabolic reprogramming upon T lymphocyte activation. Immunity 2011, 35, 871–882. [Google Scholar] [CrossRef] [Green Version]
  127. Xiong, J.; Wang, L.; Fei, X.C.; Jiang, X.F.; Zheng, Z.; Zhao, Y.; Wang, C.F.; Li, B.; Chen, S.J.; Janin, A.; et al. MYC is a positive regulator of choline metabolism and impedes mitophagy-dependent necroptosis in diffuse large B-cell lymphoma. Blood Cancer J. 2017, 7, e582. [Google Scholar] [CrossRef] [Green Version]
  128. Wong, K.K.; Engelman, J.A.; Cantley, L.C. Targeting the PI3K signaling pathway in cancer. Curr. Opin. Genet. Dev. 2010, 20, 87–90. [Google Scholar] [CrossRef] [Green Version]
  129. Westin, J.R. Status of PI3K/Akt/mTOR pathway inhibitors in lymphoma. Clin. Lymphoma Myeloma Leuk. 2014, 14, 335–342. [Google Scholar] [CrossRef] [Green Version]
  130. Carnero, A. The PKB/AKT pathway in cancer. Curr. Pharm. Des 2010, 16, 34–44. [Google Scholar] [CrossRef] [Green Version]
  131. Kittipongdaja, W.; Wu, X.; Garner, J.; Liu, X.; Komas, S.M.; Hwang, S.T.; Schieke, S.M. Rapamycin Suppresses Tumor Growth and Alters the Metabolic Phenotype in T-Cell Lymphoma. J. Investig. Dermatol 2015, 135, 2301–2308. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Shapouri-Moghaddam, A.; Mohammadian, S.; Vazini, H.; Taghadosi, M.; Esmaeili, S.A.; Mardani, F.; Seifi, B.; Mohammadi, A.; Afshari, J.T.; Sahebkar, A. Macrophage plasticity, polarization, and function in health and disease. J. Cell Physiol. 2018, 233, 6425–6440. [Google Scholar] [CrossRef] [PubMed]
  133. Schmid, M.C.; Avraamides, C.J.; Dippold, H.C.; Franco, I.; Foubert, P.; Ellies, L.G.; Acevedo, L.M.; Manglicmot, J.R.; Song, X.; Wrasidlo, W.; et al. Receptor tyrosine kinases and TLR/IL1Rs unexpectedly activate myeloid cell PI3kγ, a single convergent point promoting tumor inflammation and progression. Cancer Cell 2011, 19, 715–727. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Cairns, R.A.; Harris, I.S.; Mak, T.W. Regulation of cancer cell metabolism. Nat. Rev. Cancer 2011, 11, 85–95. [Google Scholar] [CrossRef] [Green Version]
  135. Fan, Y.; Dickman, K.G.; Zong, W.X. Akt and c-Myc differentially activate cellular metabolic programs and prime cells to bioenergetic inhibition. J. Biol. Chem. 2010, 285, 7324–7333. [Google Scholar] [CrossRef] [Green Version]
  136. Robey, R.B.; Hay, N. Is Akt the “Warburg kinase”?-Akt-energy metabolism interactions and oncogenesis. Semin. Cancer Biol. 2009, 19, 25–31. [Google Scholar] [CrossRef] [Green Version]
  137. Inoki, K.; Corradetti, M.N.; Guan, K.L. Dysregulation of the TSC-mTOR pathway in human disease. Nat. Genet. 2005, 37, 19–24. [Google Scholar] [CrossRef]
  138. Duvel, K.; Yecies, J.L.; Menon, S.; Raman, P.; Lipovsky, A.I.; Souza, A.L.; Triantafellow, E.; Ma, Q.; Gorski, R.; Cleaver, S.; et al. Activation of a metabolic gene regulatory network downstream of mTOR complex 1. Mol. Cell 2010, 39, 171–183. [Google Scholar] [CrossRef] [Green Version]
  139. Laplante, M.; Sabatini, D.M. Regulation of mTORC1 and its impact on gene expression at a glance. J. Cell Sci. 2013, 126 Pt 8, 1713–1719. [Google Scholar] [CrossRef] [Green Version]
  140. Guertin, D.A.; Sabatini, D.M. Defining the role of mTOR in cancer. Cancer Cell 2007, 12, 9–22. [Google Scholar] [CrossRef] [Green Version]
  141. Blachly, J.S.; Baiocchi, R.A. Targeting PI3-kinase (PI3K), AKT and mTOR axis in lymphoma. Br. J. Haematol. 2014, 167, 19–32. [Google Scholar] [CrossRef]
  142. Rasul, A.; Ding, C.; Li, X.; Khan, M.; Yi, F.; Ali, M.; Ma, T. Dracorhodin perchlorate inhibits PI3K/Akt and NF-κB activation, up-regulates the expression of p53, and enhances apoptosis. Apoptosis 2012, 17, 1104–1119. [Google Scholar] [CrossRef] [PubMed]
  143. Vousden, K.H.; Prives, C. Blinded by the Light: The Growing Complexity of p53. Cell 2009, 137, 413–431. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Feng, Z.; Levine, A.J. The regulation of energy metabolism and the IGF-1/mTOR pathways by the p53 protein. Trends Cell Biol. 2010, 20, 427–434. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Xi, Y.; Zhang, Y.; Pan, J.; Chen, S.; Lu, S.; Shen, F.; Huang, Z. Triptolide dysregulates glucose uptake via inhibition of IKKβ-NF-κB pathway by p53 activation in cardiomyocytes. Toxicol. Lett. 2020, 318, 1–11. [Google Scholar] [CrossRef]
  146. Yeung, S.J.; Pan, J.; Lee, M.H. Roles of p53, MYC and HIF-1 in regulating glycolysis-the seventh hallmark of cancer. Cell Mol. Life Sci. 2008, 65, 3981–3999. [Google Scholar] [CrossRef]
  147. Xu-Monette, Z.Y.; Medeiros, L.J.; Li, Y.; Orlowski, R.Z.; Andreeff, M.; Bueso-Ramos, C.E.; Greiner, T.C.; McDonnell, T.J.; Young, K.H. Dysfunction of the TP53 tumor suppressor gene in lymphoid malignancies. Blood 2012, 119, 3668–3683. [Google Scholar] [CrossRef] [Green Version]
  148. Xu-Monette, Z.Y.; Wu, L.; Visco, C.; Tai, Y.C.; Tzankov, A.; Liu, W.M.; Montes-Moreno, S.; Dybkaer, K.; Chiu, A.; Orazi, A.; et al. Mutational profile and prognostic significance of TP53 in diffuse large B-cell lymphoma patients treated with R-CHOP: Report from an International DLBCL Rituximab-CHOP Consortium Program Study. Blood 2012, 120, 3986–3996. [Google Scholar] [CrossRef]
  149. Wang, X.J.; Medeiros, L.J.; Bueso-Ramos, C.E.; Tang, G.; Wang, S.; Oki, Y.; Desai, P.; Khoury, J.D.; Miranda, R.N.; Tang, Z.; et al. P53 expression correlates with poorer survival and augments the negative prognostic effect of MYC rearrangement, expression or concurrent MYC/BCL2 expression in diffuse large B-cell lymphoma. Mod. Pathol. 2017, 30, 194–203. [Google Scholar] [CrossRef] [Green Version]
  150. Garcia, D.; Shaw, R.J. AMPK: Mechanisms of Cellular Energy Sensing and Restoration of Metabolic Balance. Mol. Cell 2017, 66, 789–800. [Google Scholar] [CrossRef] [Green Version]
  151. Faubert, B.; Boily, G.; Izreig, S.; Griss, T.; Samborska, B.; Dong, Z.; Dupuy, F.; Chambers, C.; Fuerth, B.J.; Viollet, B.; et al. AMPK is a negative regulator of the Warburg effect and suppresses tumor growth in vivo. Cell Metab. 2013, 17, 113–124. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Hardie, D.G. AMP-activated protein kinase: An energy sensor that regulates all aspects of cell function. Genes Dev. 2011, 25, 1895–1908. [Google Scholar] [CrossRef] [Green Version]
  153. Shackelford, D.B.; Shaw, R.J. The LKB1-AMPK pathway: Metabolism and growth control in tumour suppression. Nat. Rev. Cancer 2009, 9, 563–575. [Google Scholar] [CrossRef] [PubMed]
  154. Hoffman, A.E.; Demanelis, K.; Fu, A.; Zheng, T.; Zhu, Y. Association of AMP-activated protein kinase with risk and progression of non-Hodgkin lymphoma. Cancer Epidemiol. Biomark. Prev. 2013, 22, 736–744. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Houde, V.P.; Donzelli, S.; Sacconi, A.; Galic, S.; Hammill, J.A.; Bramson, J.L.; Foster, R.A.; Tsakiridis, T.; Kemp, B.E.; Grasso, G.; et al. AMPK β1 reduces tumor progression and improves survival in p53 null mice. Mol. Oncol. 2017, 11, 1143–1155. [Google Scholar] [CrossRef] [Green Version]
  156. Ying, H.; Wang, Z.; Zhang, Y.; Yang, T.Y.; Ding, Z.H.; Liu, S.Y.; Shao, J.; Liu, Y.; Fan, X.B. Capsaicin induces apoptosis in human osteosarcoma cells through AMPK-dependent and AMPK-independent signaling pathways. Mol. Cell Biochem. 2013, 384, 229–237. [Google Scholar] [CrossRef]
  157. Shi, W.Y.; Xiao, D.; Wang, L.; Dong, L.H.; Yan, Z.X.; Shen, Z.X.; Chen, S.J.; Chen, Y.; Zhao, W.L. Therapeutic metformin/AMPK activation blocked lymphoma cell growth via inhibition of mTOR pathway and induction of autophagy. Cell Death Dis. 2012, 3, e275. [Google Scholar] [CrossRef] [Green Version]
  158. Singh, A.R.; Gu, J.J.; Zhang, Q.; Torka, P.; Sundaram, S.; Mavis, C.; Hernandez-Ilizaliturri, F.J. Metformin sensitizes therapeutic agents and improves outcome in pre-clinical and clinical diffuse large B-cell lymphoma. Cancer Metab. 2020, 8, 10. [Google Scholar] [CrossRef]
  159. Montraveta, A.; Xargay-Torrent, S.; López-Guerra, M.; Rosich, L.; Pérez-Galán, P.; Salaverria, I.; Beà, S.; Kalko, S.G.; de Frias, M.; Campàs, C.; et al. Synergistic anti-tumor activity of acadesine (AICAR) in combination with the anti-CD20 monoclonal antibody rituximab in in vivo and in vitro models of mantle cell lymphoma. Oncotarget 2014, 5, 726–739. [Google Scholar] [CrossRef] [Green Version]
  160. Mukhopadhyay, S.; Chatterjee, A.; Kogan, D.; Patel, D.; Foster, D.A. 5-Aminoimidazole-4-carboxamide-1-β-4-ribofuranoside (AICAR) enhances the efficacy of rapamycin in human cancer cells. Cell Cycle 2015, 14, 3331–3339. [Google Scholar] [CrossRef] [Green Version]
  161. Liang, S.; Medina, E.A.; Li, B.; Habib, S.L. Preclinical evidence of the enhanced effectiveness of combined rapamycin and AICAR in reducing kidney cancer. Mol. Oncol. 2018, 12, 1917–1934. [Google Scholar] [CrossRef] [Green Version]
  162. Jeong, S.M.; Lee, A.; Lee, J.; Haigis, M.C. SIRT4 protein suppresses tumor formation in genetic models of Myc-induced B cell lymphoma. J. Biol. Chem. 2014, 289, 4135–4144. [Google Scholar] [CrossRef] [Green Version]
  163. Bonglack, E.N.; Messinger, J.E.; Cable, J.M.; Ch’ng, J.; Parnell, K.M.; Reinoso-Vizcaíno, N.M.; Barry, A.P.; Russell, V.S.; Dave, S.S.; Christofk, H.R.; et al. Monocarboxylate transporter antagonism reveals metabolic vulnerabilities of viral-driven lymphomas. Proc. Natl. Acad. Sci. USA 2021, 118, e2022495118. [Google Scholar] [CrossRef] [PubMed]
  164. Piccaluga, P.P.; Weber, A.; Ambrosio, M.R.; Ahmed, Y.; Leoncini, L. Epstein-Barr Virus-Induced Metabolic Rearrangements in Human B-Cell Lymphomas. Front. MicroBiol. 2018, 9, 1233. [Google Scholar] [CrossRef]
  165. Portis, T.; Dyck, P.; Longnecker, R. Epstein-Barr Virus (EBV) LMP2A induces alterations in gene transcription similar to those observed in Reed-Sternberg cells of Hodgkin lymphoma. Blood 2003, 102, 4166–4178. [Google Scholar] [CrossRef] [PubMed]
  166. Burton, E.M.; Gewurz, B.E. Epstein-Barr virus oncoprotein-driven B cell metabolism remodeling. PLoS Pathog. 2022, 18, e1010254. [Google Scholar]
  167. Mrozek-Gorska, P.; Buschle, A.; Pich, D.; Schwarzmayr, T.; Fechtner, R.; Scialdone, A.; Hammerschmidt, W. Epstein-Barr virus reprograms human B lymphocytes immediately in the prelatent phase of infection. Proc. Natl. Acad. Sci. USA 2019, 116, 16046–16055. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Mawson, A.R.; Majumdar, S. Malaria, Epstein-Barr virus infection and the pathogenesis of Burkitt’s lymphoma. Int. J. Cancer 2017, 141, 1849–1855. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Hulse, M.; Johnson, S.M.; Boyle, S.; Caruso, L.B.; Tempera, I. Epstein-Barr Virus-Encoded Latent Membrane Protein 1 and B-Cell Growth Transformation Induce Lipogenesis through Fatty Acid Synthase. J. Virol. 2021, 9, e01857-20. [Google Scholar] [CrossRef]
  170. Pujals, A.; Favre, L.; Pioche-Durieu, C.; Robert, A.; Meurice, G.; Le Gentil, M.; Chelouah, S.; Martin-Garcia, N.; Le Cam, E.; Guettier, C.; et al. Constitutive autophagy contributes to resistance to TP53-mediated apoptosis in Epstein-Barr virus-positive latency III B-cell lymphoproliferations. Autophagy 2015, 11, 2275–2287. [Google Scholar] [CrossRef] [Green Version]
  171. Nowicki, S.; Gottlieb, E. Oncometabolites: Tailoring our genes. FEBS J. 2015, 282, 2796–2805. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Zong, W.X.; Rabinowitz, J.D.; White, E. Mitochondria and Cancer. Mol. Cell 2016, 61, 667–676. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Cardaci, S.; Zheng, L.; MacKay, G.; van den Broek, N.J.; MacKenzie, E.D.; Nixon, C.; Stevenson, D.; Tumanov, S.; Bulusu, V.; Kamphorst, J.J.; et al. Pyruvate carboxylation enables growth of SDH-deficient cells by supporting aspartate biosynthesis. Nat. Cell Biol. 2015, 17, 1317–1326. [Google Scholar] [CrossRef] [PubMed]
  174. Lussey-Lepoutre, C.; Hollinshead, K.E.; Ludwig, C.; Menara, M.; Morin, A.; Castro-Vega, L.J.; Parker, S.J.; Janin, M.; Martinelli, C.; Ottolenghi, C.; et al. Loss of succinate dehydrogenase activity results in dependency on pyruvate carboxylation for cellular anabolism. Nat. Commun. 2015, 6, 8784. [Google Scholar] [CrossRef] [Green Version]
  175. Carey, B.W.; Finley, L.W.; Cross, J.R.; Allis, C.D.; Thompson, C.B. Intracellular α-ketoglutarate maintains the pluripotency of embryonic stem cells. Nature 2015, 518, 413–416. [Google Scholar] [CrossRef] [Green Version]
  176. Chin, R.M.; Fu, X.; Pai, M.Y.; Vergnes, L.; Hwang, H.; Deng, G.; Diep, S.; Lomenick, B.; Meli, V.S.; Monsalve, G.C.; et al. The metabolite α-ketoglutarate extends lifespan by inhibiting ATP synthase and TOR. Nature 2014, 510, 397–401. [Google Scholar] [CrossRef] [Green Version]
  177. Mullen, A.R.; Wheaton, W.W.; Jin, E.S.; Chen, P.H.; Sullivan, L.B.; Cheng, T.; Yang, Y.; Linehan, W.M.; Chandel, N.S.; DeBerardinis, R.J. Reductive carboxylation supports growth in tumour cells with defective mitochondria. Nature 2011, 481, 385–388. [Google Scholar] [CrossRef] [Green Version]
  178. Wang, F.; Travins, J.; DeLaBarre, B.; Penard-Lacronique, V.; Schalm, S.; Hansen, E.; Straley, K.; Kernytsky, A.; Liu, W.; Gliser, C.; et al. Targeted inhibition of mutant IDH2 in leukemia cells induces cellular differentiation. Science 2013, 340, 622–626. [Google Scholar] [CrossRef]
  179. Losman, J.A.; Looper, R.E.; Koivunen, P.; Lee, S.; Schneider, R.K.; McMahon, C.; Cowley, G.S.; Root, D.E.; Ebert, B.L.; Kaelin, W.G., Jr. (R)-2-hydroxyglutarate is sufficient to promote leukemogenesis and its effects are reversible. Science 2013, 339, 1621–1625. [Google Scholar] [CrossRef]
  180. Dang, L.; Yen, K.; Attar, E.C. IDH mutations in cancer and progress toward development of targeted therapeutics. Ann. Oncol. 2016, 27, 599–608. [Google Scholar] [CrossRef] [Green Version]
  181. Mardis, E.R.; Ding, L.; Dooling, D.J.; Larson, D.E.; McLellan, M.D.; Chen, K.; Koboldt, D.C.; Fulton, R.S.; Delehaunty, K.D.; McGrath, S.D.; et al. Recurring mutations found by sequencing an acute myeloid leukemia genome. N. Engl. J. Med. 2009, 361, 1058–1066. [Google Scholar] [CrossRef] [Green Version]
  182. Ward, P.S.; Patel, J.; Wise, D.R.; Abdel-Wahab, O.; Bennett, B.D.; Coller, H.A.; Cross, J.R.; Fantin, V.R.; Hedvat, C.V.; Perl, A.E.; et al. The common feature of leukemia-associated IDH1 and IDH2 mutations is a neomorphic enzyme activity converting alpha-ketoglutarate to 2-hydroxyglutarate. Cancer Cell 2010, 17, 225–234. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Dang, L.; White, D.W.; Gross, S.; Bennett, B.D.; Bittinger, M.A.; Driggers, E.M.; Fantin, V.R.; Jang, H.G.; Jin, S.; Keenan, M.C.; et al. Cancer-associated IDH1 mutations produce 2-hydroxyglutarate. Nature 2009, 462, 739–744. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Chowdhury, R.; Yeoh, K.K.; Tian, Y.M.; Hillringhaus, L.; Bagg, E.A.; Rose, N.R.; Leung, I.K.; Li, X.S.; Woon, E.C.; Yang, M.; et al. The oncometabolite 2-hydroxyglutarate inhibits histone lysine demethylases. EMBO Rep. 2011, 12, 463–469. [Google Scholar] [CrossRef] [Green Version]
  185. Janke, R.; Iavarone, A.T.; Rine, J. Oncometabolite D-2-Hydroxyglutarate enhances gene silencing through inhibition of specific H3K36 histone demethylases. Elife 2017, 6, e22451. [Google Scholar] [CrossRef] [PubMed]
  186. Koivunen, P.; Lee, S.; Duncan, C.G.; Lopez, G.; Lu, G.; Ramkissoon, S.; Losman, J.A.; Joensuu, P.; Bergmann, U.; Gross, S.; et al. Transformation by the (R)-enantiomer of 2-hydroxyglutarate linked to EGLN activation. Nature 2012, 483, 484–488. [Google Scholar] [CrossRef] [Green Version]
  187. Xu, W.; Yang, H.; Liu, Y.; Yang, Y.; Wang, P.; Kim, S.H.; Ito, S.; Yang, C.; Wang, P.; Xiao, M.T.; et al. Oncometabolite 2-hydroxyglutarate is a competitive inhibitor of α-ketoglutarate-dependent dioxygenases. Cancer Cell 2011, 19, 17–30. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Figueroa, M.E.; Abdel-Wahab, O.; Lu, C.; Ward, P.S.; Patel, J.; Shih, A.; Li, Y.; Bhagwat, N.; Vasanthakumar, A.; Fernandez, H.F.; et al. Leukemic IDH1 and IDH2 mutations result in a hypermethylation phenotype, disrupt TET2 function, and impair hematopoietic differentiation. Cancer Cell 2010, 18, 553–567. [Google Scholar] [CrossRef] [Green Version]
  189. Lu, C.; Ward, P.S.; Kapoor, G.S.; Rohle, D.; Turcan, S.; Abdel-Wahab, O.; Edwards, C.R.; Khanin, R.; Figueroa, M.E.; Melnick, A.; et al. IDH mutation impairs histone demethylation and results in a block to cell differentiation. Nature 2012, 483, 474–478. [Google Scholar] [CrossRef] [Green Version]
  190. Turcan, S.; Fabius, A.W.; Borodovsky, A.; Pedraza, A.; Brennan, C.; Huse, J.; Viale, A.; Riggins, G.J.; Chan, T.A. Efficient induction of differentiation and growth inhibition in IDH1 mutant glioma cells by the DNMT Inhibitor Decitabine. Oncotarget 2013, 4, 1729–1736. [Google Scholar] [CrossRef] [Green Version]
  191. Geschwind, J.F.; Georgiades, C.S.; Ko, Y.H.; Pedersen, P.L. Recently elucidated energy catabolism pathways provide opportunities for novel treatments in hepatocellular carcinoma. Expert Rev. Anticancer Ther. 2004, 4, 449–457. [Google Scholar] [CrossRef]
  192. Pang, Y.Y.; Wang, T.; Chen, F.Y.; Wu, Y.L.; Shao, X.; Xiao, F.; Huang, H.H.; Zhong, H.; Zhong, J.H. Glycolytic inhibitor 2-deoxy-d-glucose suppresses cell proliferation and enhances methylprednisolone sensitivity in non-Hodgkin lymphoma cells through down-regulation of HIF-1α and c-MYC. Leuk. Lymphoma 2015, 56, 1821–1830. [Google Scholar] [CrossRef] [PubMed]
  193. Zagorodna, O.; Martin, S.M.; Rutkowski, D.T.; Kuwana, T.; Spitz, D.R.; Knudson, C.M. 2-deoxyglucose-induced toxicity is regulated by Bcl-2 family members and is enhanced by antagonizing Bcl-2 in lymphoma cell lines. Oncogene 2012, 31, 2738–2749. [Google Scholar] [CrossRef] [Green Version]
  194. Robinson, G.L.; Dinsdale, D.; Macfarlane, M.; Cain, K. Switching from aerobic glycolysis to oxidative phosphorylation modulates the sensitivity of mantle cell lymphoma cells to TRAIL. Oncogene 2012, 31, 4996–5006. [Google Scholar] [CrossRef] [PubMed]
  195. Ko, Y.H.; Pedersen, P.L.; Geschwind, J.F. Glucose catabolism in the rabbit VX2 tumor model for liver cancer: Characterization and targeting hexokinase. Cancer Lett. 2001, 173, 83–91. [Google Scholar] [CrossRef]
  196. Schaefer, N.G.; Geschwind, J.F.; Engles, J.; Buchanan, J.W.; Wahl, R.L. Systemic administration of 3-bromopyruvate in treating disseminated aggressive lymphoma. Transl. Res. 2012, 159, 51–57. [Google Scholar] [CrossRef] [PubMed]
  197. Yadav, S.; Pandey, S.K.; Kumar, A.; Kujur, P.K.; Singh, R.P.; Singh, S.M. Antitumor and chemosensitizing action of 3-bromopyruvate: Implication of deregulated metabolism. Chem. Biol. Interact. 2017, 270, 73–89. [Google Scholar] [CrossRef] [PubMed]
  198. Le, A.; Cooper, C.R.; Gouw, A.M.; Dinavahi, R.; Maitra, A.; Deck, L.M.; Royer, R.E.; Vander Jagt, D.L.; Semenza, G.L.; Dang, C.V. Inhibition of lactate dehydrogenase A induces oxidative stress and inhibits tumor progression. Proc. Natl. Acad. Sci. USA 2010, 107, 2037–2042. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Nilsson, L.M.; Forshell, T.Z.; Rimpi, S.; Kreutzer, C.; Pretsch, W.; Bornkamm, G.W.; Nilsson, J.A. Mouse genetics suggests cell-context dependency for Myc-regulated metabolic enzymes during tumorigenesis. PLoS Genet. 2012, 8, e1002573. [Google Scholar] [CrossRef] [Green Version]
  200. Curtis, N.J.; Mooney, L.; Hopcroft, L.; Michopoulos, F.; Whalley, N.; Zhong, H.; Murray, C.; Logie, A.; Revill, M.; Byth, K.F.; et al. Pre-clinical pharmacology of AZD3965, a selective inhibitor of MCT1, DLBCL, NHL and Burkitt’s lymphoma anti-tumor activity. Oncotarget 2017, 8, 69219–69236. [Google Scholar] [CrossRef] [Green Version]
  201. Beloueche-Babari, M.; Wantuch, S.; Casals Galobart, T.; Koniordou, M.; Parkes, H.G.; Arunan, V.; Chung, Y.L.; Eykyn, T.R.; Smith, P.D.; Leach, M.O. MCT1 Inhibitor AZD3965 Increases Mitochondrial Metabolism, Facilitating Combination Therapy and Noninvasive Magnetic Resonance Spectroscopy. Cancer Res. 2017, 77, 5913–5924. [Google Scholar] [CrossRef] [Green Version]
  202. Schultz, I.R.; Shangraw, R.E. Effect of short-term drinking water exposure to dichloroacetate on its pharmacokinetics and oral bioavailability in human volunteers: A stable isotope study. Toxicol. Sci. 2006, 92, 42–50. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Chu, X.; Schwartz, R.; Diamond, M.P.; Raju, R.P. A Combination Treatment Strategy for Hemorrhagic Shock in a Rat Model Modulates Autophagy. Front. Med. 2019, 6, 281. [Google Scholar] [CrossRef]
  204. Kumar, A.; Kant, S.; Singh, S.M. Novel molecular mechanisms of antitumor action of dichloroacetate against T cell lymphoma: Implication of altered glucose metabolism, pH homeostasis and cell survival regulation. Chem. Biol. Interact. 2012, 199, 29–37. [Google Scholar] [CrossRef]
  205. Grassian, A.R.; Parker, S.J.; Davidson, S.M.; Divakaruni, A.S.; Green, C.R.; Zhang, X.; Slocum, K.L.; Pu, M.; Lin, F.; Vickers, C.; et al. IDH1 mutations alter citric acid cycle metabolism and increase dependence on oxidative mitochondrial metabolism. Cancer Res. 2014, 74, 3317–3331. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Tateishi, K.; Wakimoto, H.; Iafrate, A.J.; Tanaka, S.; Loebel, F.; Lelic, N.; Wiederschain, D.; Bedel, O.; Deng, G.; Zhang, B.; et al. Extreme Vulnerability of IDH1 Mutant Cancers to NAD+ Depletion. Cancer Cell 2015, 28, 773–784. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Sugimoto, K.; Suzuki, H.I.; Fujimura, T.; Ono, A.; Kaga, N.; Isobe, Y.; Sasaki, M.; Taka, H.; Miyazono, K.; Komatsu, N. A clinically attainable dose of L-asparaginase targets glutamine addiction in lymphoid cell lines. Cancer Sci. 2015, 106, 1534–1543. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Dengler, M.A.; Weilbacher, A.; Gutekunst, M.; Staiger, A.M.; Vöhringer, M.C.; Horn, H.; Ott, G.; Aulitzky, W.E.; van der Kuip, H. Discrepant NOXA (PMAIP1) transcript and NOXA protein levels: A potential Achilles’ heel in mantle cell lymphoma. Cell Death Dis. 2014, 5, e1013. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  209. Bastos, D.C.; Paupert, J.; Maillard, C.; Seguin, F.; Carvalho, M.A.; Agostini, M.; Coletta, R.D.; Noël, A.; Graner, E. Effects of fatty acid synthase inhibitors on lymphatic vessels: An in vitro and in vivo study in a melanoma model. Lab. Investig. 2017, 97, 194–206. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. Molyneux, E.; Merrick, B.; Khanim, F.L.; Banda, K.; Dunn, J.A.; Iqbal, G.; Bunce, C.M.; Drayson, M.T. Bezafibrate and medroxyprogesterone acetate in resistant and relapsed endemic Burkitt lymphoma in Malawi; an open-label, single-arm, phase 2 study (ISRCTN34303497). Br. J. Haematol. 2014, 164, 888–890. [Google Scholar] [CrossRef]
  211. Huang, J.; Das, S.K.; Jha, P.; Al Zoughbi, W.; Schauer, S.; Claudel, T.; Sexl, V.; Vesely, P.; Birner-Gruenberger, R.; Kratky, D.; et al. The PPARα agonist fenofibrate suppresses B-cell lymphoma in mice by modulating lipid metabolism. BioChim. Biophys. Acta 2013, 1831, 1555–1565. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Bhalla, S.; Evens, A.M.; Prachand, S.; Schumacker, P.T.; Gordon, L.I. Paradoxical regulation of hypoxia inducible factor-1α (HIF-1α) by histone deacetylase inhibitor in diffuse large B-cell lymphoma. PLoS ONE 2013, 8, e81333. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Yang, B.; Yu, D.; Liu, J.; Yang, K.; Wu, G.; Liu, H. Antitumor activity of SAHA, a novel histone deacetylase inhibitor, against murine B cell lymphoma A20 cells in vitro and in vivo. Tumour Biol. 2015, 36, 5051–5061. [Google Scholar] [CrossRef] [PubMed]
  214. Jung, K.Y.; Wang, H.; Teriete, P.; Yap, J.L.; Chen, L.; Lanning, M.E.; Hu, A.; Lambert, L.J.; Holien, T.; Sundan, A.; et al. Perturbation of the c-Myc-Max protein-protein interaction via synthetic α-helix mimetics. J. Med. Chem. 2015, 58, 3002–3024. [Google Scholar] [CrossRef] [Green Version]
  215. Huang, M.J.; Cheng, Y.C.; Liu, C.R.; Lin, S.; Liu, H.E. A small-molecule c-Myc inhibitor, 10058-F4, induces cell-cycle arrest, apoptosis, and myeloid differentiation of human acute myeloid leukemia. Exp. Hematol. 2006, 34, 1480–1489. [Google Scholar] [CrossRef]
  216. Whitfield, J.R.; Beaulieu, M.E.; Soucek, L. Strategies to Inhibit Myc and Their Clinical Applicability. Front. Cell Dev. Biol. 2017, 5, 10. [Google Scholar] [CrossRef] [Green Version]
  217. Broecker-Preuss, M.; Becher-Boveleth, N.; Bockisch, A.; Dührsen, U.; Müller, S. Regulation of glucose uptake in lymphoma cell lines by c-MYC- and PI3K-dependent signaling pathways and impact of glycolytic pathways on cell viability. J. Transl. Med. 2017, 15, 158. [Google Scholar] [CrossRef] [Green Version]
  218. Erdmann, T.; Klener, P.; Lynch, J.T.; Grau, M.; Vočková, P.; Molinsky, J.; Tuskova, D.; Hudson, K.; Polanska, U.M.; Grondine, M.; et al. Sensitivity to PI3K and AKT inhibitors is mediated by divergent molecular mechanisms in subtypes of DLBCL. Blood 2017, 130, 310–322. [Google Scholar] [CrossRef] [Green Version]
  219. Mediani, L.; Gibellini, F.; Bertacchini, J.; Frasson, C.; Bosco, R.; Accordi, B.; Basso, G.; Bonora, M.; Calabrò, M.L.; Mattiolo, A.; et al. Reversal of the glycolytic phenotype of primary effusion lymphoma cells by combined targeting of cellular metabolism and PI3K/Akt/ mTOR signaling. Oncotarget 2016, 7, 5521–5537. [Google Scholar] [CrossRef]
  220. Yahiaoui, A.; Meadows, S.A.; Sorensen, R.A.; Cui, Z.H.; Keegan, K.S.; Brockett, R.; Chen, G.; Quéva, C.; Li, L.; Tannheimer, S.L. PI3Kδ inhibitor idelalisib in combination with BTK inhibitor ONO/GS-4059 in diffuse large B cell lymphoma with acquired resistance to PI3Kδ and BTK inhibitors. PLoS ONE 2017, 12, e0171221. [Google Scholar] [CrossRef] [Green Version]
  221. Ferreira, A.C.; Robaina, M.C.; Rezende, L.M.; Severino, P.; Klumb, C.E. Histone deacetylase inhibitor prevents cell growth in Burkitt’s lymphoma by regulating PI3K/Akt pathways and leads to upregulation of miR-143, miR-145, and miR-101. Ann. Hematol. 2014, 93, 983–993. [Google Scholar] [PubMed]
  222. Sekihara, K.; Saitoh, K.; Han, L.; Ciurea, S.; Yamamoto, S.; Kikkawa, M.; Kazuno, S.; Taka, H.; Kaga, N.; Arai, H.; et al. Targeting mantle cell lymphoma metabolism and survival through simultaneous blockade of mTOR and nuclear transporter exportin-1. Oncotarget 2017, 8, 34552–34564. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Bhatt, A.P.; Bhende, P.M.; Sin, S.H.; Roy, D.; Dittmer, D.P.; Damania, B. Dual inhibition of PI3K and mTOR inhibits autocrine and paracrine proliferative loops in PI3K/Akt/mTOR-addicted lymphomas. Blood 2010, 115, 4455–4463. [Google Scholar] [CrossRef] [PubMed]
  224. Tarantelli, C.; Gaudio, E.; Arribas, A.J.; Kwee, I.; Hillmann, P.; Rinaldi, A.; Cascione, L.; Spriano, F.; Bernasconi, E.; Guidetti, F.; et al. PQR309 Is a Novel Dual PI3K/mTOR Inhibitor with Preclinical Antitumor Activity in Lymphomas as a Single Agent and in Combination Therapy. Clin. Cancer Res. 2018, 24, 120–129. [Google Scholar] [CrossRef] [Green Version]
  225. Choudhary, G.S.; Al-Harbi, S.; Mazumder, S.; Hill, B.T.; Smith, M.R.; Bodo, J.; Hsi, E.D.; Almasan, A. MCL-1 and BCL-xL-dependent resistance to the BCL-2 inhibitor ABT-199 can be overcome by preventing PI3K/AKT/mTOR activation in lymphoid malignancies. Cell Death Dis. 2015, 6, e1593. [Google Scholar] [CrossRef] [Green Version]
  226. Campàs, C.; Santidrián, A.F.; Domingo, A.; Gil, J. Acadesine induces apoptosis in B cells from mantle cell lymphoma and splenic marginal zone lymphoma. Leukemia 2005, 19, 292–294. [Google Scholar] [CrossRef] [Green Version]
  227. Campàs, C.; Lopez, J.M.; Santidrián, A.F.; Barragán, M.; Bellosillo, B.; Colomer, D.; Gil, J. Acadesine activates AMPK and induces apoptosis in B-cell chronic lymphocytic leukemia cells but not in T lymphocytes. Blood 2003, 101, 3674–3680. [Google Scholar] [CrossRef]
  228. Santidrián, A.F.; González-Gironès, D.M.; Iglesias-Serret, D.; Coll-Mulet, L.; Cosialls, A.M.; de Frias, M.; Campàs, C.; González-Barca, E.; Alonso, E.; Labi, V.; et al. AICAR induces apoptosis independently of AMPK and p53 through up-regulation of the BH3-only proteins BIM and NOXA in chronic lymphocytic leukemia cells. Blood 2010, 116, 3023–3032. [Google Scholar] [CrossRef] [Green Version]
  229. Sengupta, T.K.; Leclerc, G.M.; Hsieh-Kinser, T.T.; Leclerc, G.J.; Singh, I.; Barredo, J.C. Cytotoxic effect of 5-aminoimidazole-4-carboxamide-1-beta-4-ribofuranoside (AICAR) on childhood acute lymphoblastic leukemia (ALL) cells: Implication for targeted therapy. Mol. Cancer 2007, 6, 46. [Google Scholar] [CrossRef] [Green Version]
  230. Rosilio, C.; Lounnas, N.; Nebout, M.; Imbert, V.; Hagenbeek, T.; Spits, H.; Asnafi, V.; Pontier-Bres, R.; Reverso, J.; Michiels, J.F.; et al. The metabolic perturbators metformin, phenformin and AICAR interfere with the growth and survival of murine PTEN-deficient T cell lymphomas and human T-ALL/T-LL cancer cells. Cancer Lett. 2013, 336, 114–126. [Google Scholar] [CrossRef]
  231. Offman, M.N.; Krol, M.; Patel, N.; Krishnan, S.; Liu, J.; Saha, V.; Bates, P.A. Rational engineering of L-asparaginase reveals importance of dual activity for cancer cell toxicity. Blood 2011, 117, 1614–1621. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  232. Abd-Aziz, N.; Stanbridge, E.J.; Shafee, N. Bortezomib attenuates HIF-1- but not HIF-2-mediated transcriptional activation. Oncol. Lett. 2015, 10, 2192–2196. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  233. Bhatt, R.; Ravi, D.; Evens, A.M.; Parekkadan, B. Scaffold-mediated switching of lymphoma metabolism in culture. Cancer Metab. 2022, 10, 15. [Google Scholar] [CrossRef]
  234. Welcker, M.; Orian, A.; Jin, J.; Grim, J.E.; Harper, J.W.; Eisenman, R.N.; Clurman, B.E. The Fbw7 tumor suppressor regulates glycogen synthase kinase 3 phosphorylation-dependent c-Myc protein degradation. Proc. Natl. Acad. Sci. USA 2004, 101, 9085–9090. [Google Scholar] [CrossRef] [Green Version]
  235. Bahram, F.; von der Lehr, N.; Cetinkaya, C.; Larsson, L.G. c-Myc hot spot mutations in lymphomas result in inefficient ubiquitination and decreased proteasome-mediated turnover. Blood 2000, 95, 2104–2110. [Google Scholar] [CrossRef]
  236. Popov, N.; Wanzel, M.; Madiredjo, M.; Zhang, D.; Beijersbergen, R.; Bernards, R.; Moll, R.; Elledge, S.J.; Eilers, M. The ubiquitin-specific protease USP28 is required for MYC stability. Nat. Cell Biol. 2007, 9, 765–774. [Google Scholar] [CrossRef]
  237. Xie, H.; Tang, C.H.; Song, J.H.; Mancuso, A.; Del Valle, J.R.; Cao, J.; Xiang, Y.; Dang, C.V.; Lan, R.; Sanchez, D.J.; et al. IRE1α RNase-dependent lipid homeostasis promotes survival in Myc-transformed cancers. J. Clin. Investig. 2018, 128, 1300–1316. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  238. Dreyling, M.; Santoro, A.; Mollica, L.; Leppä, S.; Follows, G.A.; Lenz, G.; Kim, W.S.; Nagler, A.; Panayiotidis, P.; Demeter, J.; et al. Phosphatidylinositol 3-Kinase Inhibition by Copanlisib in Relapsed or Refractory Indolent Lymphoma. J. Clin. Oncol. 2017, 35, 3898–3905. [Google Scholar] [CrossRef]
  239. Horwitz, S.M.; Koch, R.; Porcu, P.; Oki, Y.; Moskowitz, A.; Perez, M.; Myskowski, P.; Officer, A.; Jaffe, J.D.; Morrow, S.N.; et al. Activity of the PI3K-δ,γ inhibitor duvelisib in a phase 1 trial and preclinical models of T-cell lymphoma. Blood 2018, 131, 888–898. [Google Scholar] [CrossRef] [Green Version]
  240. Gu, L.; Gao, J.; Li, Q.; Zhu, Y.P.; Jia, C.S.; Fu, R.Y.; Chen, Y.; Liao, Q.K.; Ma, Z. Rapamycin reverses NPM-ALK-induced glucocorticoid resistance in lymphoid tumor cells by inhibiting mTOR signaling pathway, enhancing G1 cell cycle arrest and apoptosis. Leukemia 2008, 22, 2091–2096. [Google Scholar] [CrossRef] [PubMed]
  241. Gu, L.; Xie, L.; Zuo, C.; Ma, Z.; Zhang, Y.; Zhu, Y.; Gao, J. Targeting mTOR/p70S6K/glycolysis signaling pathway restores glucocorticoid sensitivity to 4E-BP1 null Burkitt Lymphoma. BMC Cancer 2015, 15, 529. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  242. Drew, B.G.; Kingwell, B.A. Acadesine, an adenosine-regulating agent with the potential for widespread indications. Expert Opin. Pharmacother. 2008, 9, 2137–2144. [Google Scholar] [CrossRef]
  243. Van Den Neste, E.; Van den Berghe, G.; Bontemps, F. AICA-riboside (acadesine), an activator of AMP-activated protein kinase with potential for application in hematologic malignancies. Expert Opin. Investig. Drugs 2010, 19, 571–578. [Google Scholar] [CrossRef] [PubMed]
  244. Vakana, E.; Platanias, L.C. AMPK in BCR-ABL expressing leukemias. Regulatory effects and therapeutic implications. Oncotarget 2011, 2, 1322–1328. [Google Scholar] [CrossRef] [Green Version]
  245. Liu, X.; Chhipa, R.R.; Pooya, S.; Wortman, M.; Yachyshin, S.; Chow, L.M.; Kumar, A.; Zhou, X.; Sun, Y.; Quinn, B.; et al. Discrete mechanisms of mTOR and cell cycle regulation by AMPK agonists independent of AMPK. Proc. Natl. Acad. Sci. USA 2014, 111, E435–E444. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  246. Ly, P.; Kim, S.B.; Kaisani, A.A.; Marian, G.; Wright, W.E.; Shay, J.W. Aneuploid human colonic epithelial cells are sensitive to AICAR-induced growth inhibition through EGFR degradation. Oncogene 2013, 32, 3139–3146. [Google Scholar] [CrossRef] [Green Version]
  247. Jose, C.; Bellance, N.; Chatelain, E.H.; Benard, G.; Nouette-Gaulain, K.; Rossignol, R. Antiproliferative activity of levobupivacaine and aminoimidazole carboxamide ribonucleotide on human cancer cells of variable bioenergetic profile. Mitochondrion 2012, 12, 100–109. [Google Scholar] [CrossRef]
  248. Rattan, R.; Giri, S.; Singh, A.K.; Singh, I. 5-Aminoimidazole-4-carboxamide-1-beta-D-ribofuranoside inhibits cancer cell proliferation in vitro and in vivo via AMP-activated protein kinase. J. Biol. Chem. 2005, 280, 39582–39593. [Google Scholar] [CrossRef] [Green Version]
  249. Van Den Neste, E.; Cazin, B.; Janssens, A.; González-Barca, E.; Terol, M.J.; Levy, V.; Pérez de Oteyza, J.; Zachee, P.; Saunders, A.; de Frias, M.; et al. Acadesine for patients with relapsed/refractory chronic lymphocytic leukemia (CLL): A multicenter phase I/II study. Cancer Chemo. Ther. Pharmacol. 2013, 71, 581–591. [Google Scholar] [CrossRef] [Green Version]
  250. Wang, H.; Shao, R.; Liu, W.; Tang, H.; Lu, Y. Identification of a prognostic metabolic gene signature in diffuse large B-cell lymphoma. J. Cell Mol. Med. 2021, 25, 7066–7077. [Google Scholar] [CrossRef]
  251. He, J.; Chen, Z.; Xue, Q.; Sun, P.; Wang, Y.; Zhu, C.; Shi, W. Identification of molecular subtypes and a novel prognostic model of diffuse large B-cell lymphoma based on a metabolism-associated gene signature. J. Transl. Med. 2022, 20, 186. [Google Scholar] [CrossRef]
  252. Zhang, J.; Medina-Cleghorn, D.; Bernal-Mizrachi, L.; Bracci, P.M.; Hubbard, A.; Conde, L.; Riby, J.; Nomura, D.K.; Skibola, C.F. The potential relevance of the endocannabinoid, 2-arachidonoylglycerol, in diffuse large B-cell lymphoma. Oncoscience 2016, 3, 31–41. [Google Scholar] [CrossRef] [Green Version]
  253. Yang, F.; Du, J.; Zhang, H.; Ruan, G.; Xiang, J.; Wang, L.; Sun, H.; Guan, A.; Shen, G.; Liu, Y.; et al. Serum Metabolomics of Burkitt Lymphoma Mouse Models. PLoS ONE 2017, 12, e0170896. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  254. Stenson, M.; Pedersen, A.; Hasselblom, S.; Nilsson-Ehle, H.; Karlsson, B.G.; Pinto, R.; Andersson, P.O. Serum nuclear magnetic resonance-based metabolomics and outcome in diffuse large B-cell lymphoma patients-a pilot study. Leuk. Lymphoma 2016, 57, 1814–1822. [Google Scholar] [CrossRef] [PubMed]
  255. Le, Y.; Shen, X.; Kang, H.; Wang, Q.; Li, K.; Zheng, J.; Yu, Y. Accelerated, untargeted metabolomics analysis of cutaneous T-cell lymphoma reveals metabolic shifts in plasma and tumor adjacent skins of xenograft mice. J. Mass Spectrom 2018, 53, 739. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  256. Eberlin, L.S.; Gabay, M.; Fan, A.C.; Gouw, A.M.; Tibshirani, R.J.; Felsher, D.W.; Zare, R.N. Alteration of the lipid profile in lymphomas induced by MYC overexpression. Proc. Natl. Acad. Sci. USA 2014, 111, 10450–10455. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  257. Leone, R.D.; Powell, J.D. Metabolism of immune cells in cancer. Nat. Rev. Cancer 2020, 20, 516–531. [Google Scholar] [CrossRef]
  258. Deshpande, R.P.; Sharma, S.; Watabe, K. The Confounders of Cancer Immunotherapy: Roles of Lifestyle, Metabolic Disorders and Sociological Factors. Cancers 2020, 12, 2983. [Google Scholar] [CrossRef]
  259. Liu, Y.; Xu, R.; Gu, H.; Zhang, E.; Qu, J.; Cao, W.; Huang, X.; Yan, H.; He, J.; Cai, Z. Metabolic reprogramming in macrophage responses. Biomark. Res. 2021, 9, 1. [Google Scholar] [CrossRef]
  260. Dai, X.; Bu, X.; Gao, Y.; Guo, J.; Hu, J.; Jiang, C.; Zhang, Z.; Xu, K.; Duan, J.; He, S.; et al. Energy status dictates PD-L1 protein abundance and anti-tumor immunity to enable checkpoint blockade. Mol. Cell 2021, 81, 2317–2331.e6. [Google Scholar] [CrossRef]
  261. Chao, M.P.; Alizadeh, A.A.; Tang, C.; Myklebust, J.H.; Varghese, B.; Gill, S.; Jan, M.; Cha, A.C.; Chan, C.K.; Tan, B.T.; et al. Anti-CD47 antibody synergizes with rituximab to promote phagocytosis and eradicate non-Hodgkin lymphoma. Cell 2010, 142, 699–713. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  262. Zhao, P.; Xie, L.; Yu, L.; Wang, P. Targeting CD47-SIRPα axis for Hodgkin and non-Hodgkin lymphoma immunotherapy. Genes Dis. 2023, in press. [Google Scholar] [CrossRef]
Figure 1. Metabolic pathways in lymphoma. Lymphoma cells take up glucose and glutamine also to fuel the tricarboxylic acid (TCA) cycle, oxidative phosphorylation (OXPHOS), the pentose phosphate pathway (nucleotide synthesis), and the syntheses of hexosamine (involved in the synthesis of glycosylated molecules), amino acids (proteins), and lipids. Lymphoma cells are also able to take up lactate, free fatty acids, and ketones released predominantly by surrounding catabolic cells, which can be used to replenish TCA-cycle intermediates and fuel OXPHOS. Together, these pathways generate sufficient levels of cellular components to support cell proliferation (necessary nutrient transporters and metabolic enzymes are highlighted in red). Abbreviations: MCT1, monocarboxylate transporter 1; MCT4, monocarboxylate transporter 4; GLUT1/GLUT4, glucose transporter 1/4; HK2, hexokinase; GPI, glucose-6-phosphate isomerase; PFK, phosphofructokinase; ALDA, aldolase A; TPI, triosephosphate isomerase; GAPDH, glyceraldehyde-3-phosphate dehydrogenase; PHGDH, phosphoglycerate dehydrogenase; SHMT1, serine hydroxymethyltransferase 1; PGK1, phosphoglycerate kinase 1; PGM, phosphoglycerate mutase; ENO1, enolase-1; PKM2, pyruvate kinase M2; LDHA, lactate dehydrogenase A; PDH, pyruvate dehydrogenase; PDK1, pyruvate dehydrogenase kinase 1; PC, pyruvate carboxylase; CS, citrate lyase; IDH1/2, isocitrate dehydrogenase1/2; GLS, glutaminase; GS, glutamine synthetase; SDH, succinate dehydrogenase; FH, fumarate hydratase; P5CS, Δ (1)-pyrroline-5-carboxylate (P5C) synthase; P5CDH, P5C dehydrogenase; LAT1, L-type amino-acid transporter 1; CK, choline kinase; ACLY, ATP citrate lyase; ACC, acetyl CoA carboxylase; FASN, fatty acid synthase; PYCR, pyrroline-5-carboxylate reductase; POX/PRODH, proline dehydrogenase; TCA, tricarboxylic acid.
Figure 1. Metabolic pathways in lymphoma. Lymphoma cells take up glucose and glutamine also to fuel the tricarboxylic acid (TCA) cycle, oxidative phosphorylation (OXPHOS), the pentose phosphate pathway (nucleotide synthesis), and the syntheses of hexosamine (involved in the synthesis of glycosylated molecules), amino acids (proteins), and lipids. Lymphoma cells are also able to take up lactate, free fatty acids, and ketones released predominantly by surrounding catabolic cells, which can be used to replenish TCA-cycle intermediates and fuel OXPHOS. Together, these pathways generate sufficient levels of cellular components to support cell proliferation (necessary nutrient transporters and metabolic enzymes are highlighted in red). Abbreviations: MCT1, monocarboxylate transporter 1; MCT4, monocarboxylate transporter 4; GLUT1/GLUT4, glucose transporter 1/4; HK2, hexokinase; GPI, glucose-6-phosphate isomerase; PFK, phosphofructokinase; ALDA, aldolase A; TPI, triosephosphate isomerase; GAPDH, glyceraldehyde-3-phosphate dehydrogenase; PHGDH, phosphoglycerate dehydrogenase; SHMT1, serine hydroxymethyltransferase 1; PGK1, phosphoglycerate kinase 1; PGM, phosphoglycerate mutase; ENO1, enolase-1; PKM2, pyruvate kinase M2; LDHA, lactate dehydrogenase A; PDH, pyruvate dehydrogenase; PDK1, pyruvate dehydrogenase kinase 1; PC, pyruvate carboxylase; CS, citrate lyase; IDH1/2, isocitrate dehydrogenase1/2; GLS, glutaminase; GS, glutamine synthetase; SDH, succinate dehydrogenase; FH, fumarate hydratase; P5CS, Δ (1)-pyrroline-5-carboxylate (P5C) synthase; P5CDH, P5C dehydrogenase; LAT1, L-type amino-acid transporter 1; CK, choline kinase; ACLY, ATP citrate lyase; ACC, acetyl CoA carboxylase; FASN, fatty acid synthase; PYCR, pyrroline-5-carboxylate reductase; POX/PRODH, proline dehydrogenase; TCA, tricarboxylic acid.
Ijms 24 05493 g001
Figure 2. Regulation of lymphoma metabolism by oncogenes and tumor suppressors. Oncogenic signaling (labels in red) regulates the acquisition of abundant nutrients, including glucose and glutamine, and their utilization to support biosynthetic pathways. Intermediates from glycolysis and the TCA cycle supply biosynthetic pathways to produce the macromolecules necessary for cell proliferation. Tumor suppressors (in blue), such as AMPK, p53, and PTEN, act at various nodes to oppose biosynthetic metabolism. Abbreviations: HIF-1α, hypoxia-inducible factor 1 alpha; mTOR, mammalian target of rapamycin; LDHA, lactate dehydrogenase A; GLUT3, glucose transporter 3; HK2, hexokinase; GPI, glucose-6-phosphate isomerase; PFK, phosphofructokinase; GAPDH, glyceraldehyde-3-phosphate dehydrogenase; PKM, pyruvate kinase M; PDK1, pyruvate dehydrogenase kinase 1; PDH, pyruvate dehydrogenase; TCA, tricarboxylic acid; IDH1/2, isocitrate dehydrogenase1/2; POX/PRODH, proline dehydrogenase; PYCR, pyrroline-5-carboxylate reductase; P5C, Δ (1)-pyrroline-5-carboxylate; P5CS, P5C synthase; P5CDH, P5C dehydrogenase; GSA, glutamic semialdehyde; GS, glutamine synthetase; GLS, glutaminase; LAT1, L-type amino-acid transporter 1; FASN, fatty acid synthase; PI3K, phosphatidylinositol 3-kinase; SREBP-1, sterol-regulatory element binding protein-1; AMPK, 5ʹ-AMP-activated protein kinase; NF-κB, nuclear factor kappa B; ROS, reactive oxygen species; 2-HG, 2-hydroxyglutarate; PHDs, prolyl hydroxylases; eLF-4E, eukaryotic initiation factor 4E; 4E-BP1, 4E binding protein 1; BCR, B cell receptor.
Figure 2. Regulation of lymphoma metabolism by oncogenes and tumor suppressors. Oncogenic signaling (labels in red) regulates the acquisition of abundant nutrients, including glucose and glutamine, and their utilization to support biosynthetic pathways. Intermediates from glycolysis and the TCA cycle supply biosynthetic pathways to produce the macromolecules necessary for cell proliferation. Tumor suppressors (in blue), such as AMPK, p53, and PTEN, act at various nodes to oppose biosynthetic metabolism. Abbreviations: HIF-1α, hypoxia-inducible factor 1 alpha; mTOR, mammalian target of rapamycin; LDHA, lactate dehydrogenase A; GLUT3, glucose transporter 3; HK2, hexokinase; GPI, glucose-6-phosphate isomerase; PFK, phosphofructokinase; GAPDH, glyceraldehyde-3-phosphate dehydrogenase; PKM, pyruvate kinase M; PDK1, pyruvate dehydrogenase kinase 1; PDH, pyruvate dehydrogenase; TCA, tricarboxylic acid; IDH1/2, isocitrate dehydrogenase1/2; POX/PRODH, proline dehydrogenase; PYCR, pyrroline-5-carboxylate reductase; P5C, Δ (1)-pyrroline-5-carboxylate; P5CS, P5C synthase; P5CDH, P5C dehydrogenase; GSA, glutamic semialdehyde; GS, glutamine synthetase; GLS, glutaminase; LAT1, L-type amino-acid transporter 1; FASN, fatty acid synthase; PI3K, phosphatidylinositol 3-kinase; SREBP-1, sterol-regulatory element binding protein-1; AMPK, 5ʹ-AMP-activated protein kinase; NF-κB, nuclear factor kappa B; ROS, reactive oxygen species; 2-HG, 2-hydroxyglutarate; PHDs, prolyl hydroxylases; eLF-4E, eukaryotic initiation factor 4E; 4E-BP1, 4E binding protein 1; BCR, B cell receptor.
Ijms 24 05493 g002
Figure 3. Therapeutic strategies targeting metabolism in lymphoma. Several pathways of the bioenergetic and anabolic metabolism of malignant cells harbor targets for the treatment of cancer. In general, agents that disrupt these pathways would be expected to result in deficiencies in energy and materials needed for cell proliferation and survival, forming the basis for their use as anticancer therapies. Abbreviations: MCT1/4, monocarboxylate transporter 1/4; GULT1, glucose transporter 1; HK2, hexokinase 2; LDHA, lactate dehydrogenase A; PDH, pyruvate dehydrogenase; PDK, pyruvate dehydrogenase kinase; DCA, dichochloroacetate; IDH2, isocitrate dehydrogenase; α-KG, α-ketoglutarate; TCA, tricarboxylic acid; GLS, glutaminase; BPTES, bis-2-(5-phenylacetamido-1,2,4-diathiazol-2-yl) ethyl sulfide; LAT1, L-type amino-acid transporter 1; FASN, fatty acid synthase; ACC, acetyl-CoA carboxylase; HDACi, histone deacetylase inhibitor; HIF-1α, hypoxia-inducible factor 1 alpha; mTOR, mammalian target of rapamycin; AMPK, 5ʹ-AMP-activated protein kinase; AICAR, 5-aminoimidazole-4-carboxamide-1-β-D-ribofuranoside; PI3K, phosphatidylinositol 3-kinase; BCR, B cell receptor; CK, choline kinase. In the metabolic pathways, transporters are in green color, key metabolic enzymes are in red color, metabolites are in blue color, and inhibitors are in black color.
Figure 3. Therapeutic strategies targeting metabolism in lymphoma. Several pathways of the bioenergetic and anabolic metabolism of malignant cells harbor targets for the treatment of cancer. In general, agents that disrupt these pathways would be expected to result in deficiencies in energy and materials needed for cell proliferation and survival, forming the basis for their use as anticancer therapies. Abbreviations: MCT1/4, monocarboxylate transporter 1/4; GULT1, glucose transporter 1; HK2, hexokinase 2; LDHA, lactate dehydrogenase A; PDH, pyruvate dehydrogenase; PDK, pyruvate dehydrogenase kinase; DCA, dichochloroacetate; IDH2, isocitrate dehydrogenase; α-KG, α-ketoglutarate; TCA, tricarboxylic acid; GLS, glutaminase; BPTES, bis-2-(5-phenylacetamido-1,2,4-diathiazol-2-yl) ethyl sulfide; LAT1, L-type amino-acid transporter 1; FASN, fatty acid synthase; ACC, acetyl-CoA carboxylase; HDACi, histone deacetylase inhibitor; HIF-1α, hypoxia-inducible factor 1 alpha; mTOR, mammalian target of rapamycin; AMPK, 5ʹ-AMP-activated protein kinase; AICAR, 5-aminoimidazole-4-carboxamide-1-β-D-ribofuranoside; PI3K, phosphatidylinositol 3-kinase; BCR, B cell receptor; CK, choline kinase. In the metabolic pathways, transporters are in green color, key metabolic enzymes are in red color, metabolites are in blue color, and inhibitors are in black color.
Ijms 24 05493 g003
Table 2. Primary approved drugs and ongoing clinical trials targeting lymphoma metabolism.
Table 2. Primary approved drugs and ongoing clinical trials targeting lymphoma metabolism.
AgentTargetMechanismType of Diseases
methotrexateDHFRfolate to THF conversionprophylaxis and treatment of CNS lymphoma
IM156mitochondrial complex I inhibitormitochondrial oxidative phosphorylation and NADH oxidationlymphomas (NCT03272256)
IACS-010759oxidative phosphorylation inhibitormitochondrial oxidative phosphorylation and NADH oxidationR/R AML (NCT02882321)
AZD-3965MCT1mediate the bidirectional transport of lactatein and out of cellsDLBCL, BL (NCT01791595)
CPI-613mitochondriaoxidative metabolismT-cell NHL (NCT04217317), R/R BL or HGBCL (NCT03793140)
IM156mitochondriacomplex Iadvanced lymphoma (NCT03272256)
IACS-010759mitochondriacomplex IR/R AML (NCT02882321)
ibrutinibBTK inhibitordownstream pro-proliferative kinase of BCR signalMCL, CLL/SLL, WM, MZL, cGVHD (NCT02169180, NCT04771507, NCT02604511, NCT04212013, NCT05348096)
acalabrutinibMCL, CLL/SLL, MZL (NCT02213926, NCT04505254, NCT04646395)
rapalogsmTORC1 inhibitorrapamycin analoguesMCL
temsirolimusmTORC1 inhibitorcell cycle arrest in the G1 phase, inhibits tumor angiogenesis by reducing synthesis of VEGFMCL (NCT01078142, NCT01180049), HL (NCT01902160), R/R NHL (NCT01281917)
idelalisibPI3Kδ inhibitoragainst PI3Kδ isoformsiNHL (NCT01282424), HL (NCT01393106), FL (NCT03568929), CLL (NCT03582098)
copanlisibPI3Kδ and PI3Kα inhibitoragainst PI3K-α and PI3K-δ isoformsiNHL, DLBCL, MCL, PTCL (NCT05217914, NCT04433182, NCT04939272, NCT03877055, NCT03052933)
duvelisibPI3Kδ and PI3Kγinhibitoragainst PI3Kδ and PI3Kγ isoformsCLL/SLL, DLBCL, PTCL (NCT02004522, NCT04890236, NCT04803201)
umbralisibPI3Kδ and casein kinase-1 epsilon inhibitoragainst PI3Kδisoforms and casein kinase-1 epsilonMZL, FL, CLL, WM (NCT03919175, NCT03364231, NCT02535286)
L-asparaginaseasparagineinhibit protein biosynthesis in lymphoblastsALL (NCT01518517, NCT00506597), NK/T cell lymphoma (NCT00854425)
Telaglenastat (CB-839)glutaminase inhibitorglutamine to glutamate conversionNHL (NCT02071888), including MCL, WM, TCL
AG-270MAT2Aproduction of S-adenosylmethionineadvanced lymphoma (NCT03361358, NCT03435250)
Devimistat (CPI-613)lipoate analogmitochondrial oxidative metabolismR/R T-cell NHL (NCT04217317)
Abbreviations: DHFR, dihydrofolate reductase; MCT1, monocarboxylate transporter 1; BTK, Bruton’s tyrosine kinase; mTOR, mammalian target of rapamycin; PI3K, phosphoinositide 3-kinases; MAT2A, methionine adenosyltransferase 2A; THF, tetrahydrofolate; NADH, nicotinamide adenine dinucleotide (NAD) + hydrogen (H); BCR, B-cell receptor; VEGF, vascular endothelial growth factor; CNS, central nervous system; R/R, relapsed or refractory; AML, acute myeloid leukemia; DLBCL, diffuse large B-cell lymphoma; BL, Burkitt lymphoma; NHL, non-Hodgkin lymphoma; HGBCL, high grade B-cell lymphoma; MCL, mantle cell lymphoma; CLL/SLL, chronic lymphocytic leukemia/small lymphocytic lymphoma; WM, Waldenstrom macroglobulinemia; MZL, marginal zone lymphoma; cGVHD, chronic graft versus host disease; HL, Hodgkin lymphoma; iNHL, indolent non-Hodgkin lymphoma; FL, follicular lymphoma; PTCL, peripheral T-cell lymphoma; ALL, acute lymphocytic leukemia; TCL, T-cell lymphoma.
Table 4. Potential metabolic biomarkers in NHLs.
Table 4. Potential metabolic biomarkers in NHLs.
SubtypeControlMetabolitesType of Samples OriginMethodClinical RelevanceRef
DLBCLhealthy people2-AGserumpatientsLC-MS/MSpathogenesis or progression[252]
BLNormal miceglutamate, glycerol, and cholineserummiceNMR and MSdiagnosis and prognosis[253]
Refractory or early relapse DLBCL patientscured patients↑lysine, arginine, cadaverine, 2-HBserumpatientsNMRhigh-risk of failing to immunochemotherapy[254]
CTCLcontrol samples↑GLT
↓adenosine monophosphate
↑CTP
↑prostaglandins, pyrimidine, mevalonate pathway
↓tryptophan
↑PRPP
skin and plasmamiceUHPLC-QTOFprogress of carcinogenesis, leads to CTCL further development.[255]
MYC-induced lymphomasnormal tissue↑glycerophosphoglycerols, cardiolipins and monounsaturated fatty acidstissue and cellspatients and cell linesDESI-MSIMYC regulates cellular metabolism in cancer[256]
Abbreviations: DLBCL, diffuse large B-cell lymphoma; BL, Burkitt lymphoma; CTCL, cutaneous T-cell lymphoma; 2-AG, 2-arachidonoylglycerol; 2-HB, 2-hydroxybutyrate; GLT, L-glutamate; CTP, cytidine-5′-triphosphate; PRPP, 5-phospho-α-D-ribose 1-diphosphate; LC-MS/MS, liquid chromatography-tandem mass spectrometry/mass spectrometry; NMR, nuclear magnetic resonance; MS, mass spectrometry; UHPLC-QTOF, ultra-high-performance liquid chromatography–quadrupole time-of-flight; DESI-MSI, desorption electrospray ionization mass spectrometry imaging.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Pang, Y.; Lu, T.; Xu-Monette, Z.Y.; Young, K.H. Metabolic Reprogramming and Potential Therapeutic Targets in Lymphoma. Int. J. Mol. Sci. 2023, 24, 5493. https://doi.org/10.3390/ijms24065493

AMA Style

Pang Y, Lu T, Xu-Monette ZY, Young KH. Metabolic Reprogramming and Potential Therapeutic Targets in Lymphoma. International Journal of Molecular Sciences. 2023; 24(6):5493. https://doi.org/10.3390/ijms24065493

Chicago/Turabian Style

Pang, Yuyang, Tingxun Lu, Zijun Y. Xu-Monette, and Ken H. Young. 2023. "Metabolic Reprogramming and Potential Therapeutic Targets in Lymphoma" International Journal of Molecular Sciences 24, no. 6: 5493. https://doi.org/10.3390/ijms24065493

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