Potential Therapeutic Targets in Lymphoma: Comparison
Please note this is a comparison between Version 2 by Jessie Wu and Version 1 by Zijun Y. Xu-Monette.

Lymphoma is a heterogeneous group of diseases that often require their metabolism programs 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. 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. 

  • lymphoma
  • metabolism
  • therapy

1. 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 1).
Table 1.
Primary approved drugs and ongoing clinical trials targeting lymphoma metabolism.
Agent Target Mechanism Type of Diseases
methotrexate DHFR folate to THF conversion prophylaxis and treatment of CNS lymphoma
IM156 mitochondrial complex I inhibitor mitochondrial oxidative phosphorylation and NADH oxidation lymphomas (NCT03272256)
IACS-010759 oxidative phosphorylation inhibitor
203][13][14] (Table 2). Accompanied by modulation of glycolysis and expression of HIF1-α, DCA inhibited cell survival in a DL mouse model [204][15]. 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 2). These findings indicate that the mitochondrial translation pathway is a potential therapeutic target for these tumors as well [56][16].
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][17], hypomethylating agents [190][18], depletion of the canonical coenzyme NAD+, and chemoradiotherapy [206][19]. Pharmacological molecules targeting mutant IDH1 and IDH2 enzymes are being explored and assessed for antitumor efficacy.
Table 2.
Targeting lymphoma metabolism via metabolic enzymes, metabolite depletion, and/or signaling pathways.
Pathway Compound Application Development Stage Ref
HK2 2-DG B-NHL cell lines [192][2]
3-BrPA BL cell lines, mice [195,196][6][7]
Pyruvatemitochondrial oxidative phosphorylation and NADH oxidation R/R AML (NCT02882321)
3-BP T-NHL mice (DL) [197][8] AZD-3965 MCT1 mediate the bidirectional transport of lactatein and out of cells DLBCL, BL (NCT01791595)
LDHA FX11 BL cell line [198,[9199]][10] CPI-613 mitochondria oxidative metabolism T-cell NHL (NCT04217317), R/R BL or HGBCL (NCT03793140)
α-Tocopherol T-NHL mice (DL) [ IM156 mitochondria complex I advanced lymphoma (NCT03272256)
199][10]
PDK DCA DL mice [202,203][13][14] IACS-010759 mitochondria complex I R/R AML (NCT02882321)
Mitochondrial protein translation Tigecycline OxPhos-DLBCL cell lines FDA-approved [56][16] ibrutinib BTK inhibitor downstream pro-proliferative kinase of BCR signal
GlutaminaseMCL, CLL/SLL, WM, MZL, cGVHD (NCT02169180, NCT04771507, NCT02604511, NCT04212013, NCT05348096)
BPTES BL mice [47][20 acalabrutinib MCL, CLL/SLL, MZL (NCT02213926, NCT04505254, NCT04646395)
rapalogs mTORC1 inhibitor rapamycin analogues MCL
temsirolimus mTORC1 inhibitor cell cycle arrest in the G1 phase, inhibits tumor angiogenesis by reducing synthesis of VEGF MCL (NCT01078142, NCT01180049), HL (NCT01902160), R/R NHL (NCT01281917)
idelalisib PI3Kδ inhibitor against PI3Kδ isoforms iNHL (NCT01282424), HL (NCT01393106), FL (NCT03568929), CLL (NCT03582098)
copanlisib PI3Kδ and PI3Kα inhibitor against PI3K-α and PI3K-δ isoforms iNHL, DLBCL, MCL, PTCL (NCT05217914, NCT04433182, NCT04939272, NCT03877055, NCT03052933)
duvelisib PI3Kδ and PI3Kγinhibitor against PI3Kδ and PI3Kγ isoforms CLL/SLL, DLBCL, PTCL (NCT02004522, NCT04890236, NCT04803201)
umbralisib PI3Kδ and casein kinase-1 epsilon inhibitor against PI3Kδisoforms and casein kinase-1 epsilon MZL, FL, CLL, WM (NCT03919175, NCT03364231, NCT02535286)
L-asparaginase asparagine inhibit protein biosynthesis in lymphoblasts ALL (NCT01518517, NCT00506597), NK/T cell lymphoma (NCT00854425)
Telaglenastat (CB-839) glutaminase inhibitor glutamine to glutamate conversion NHL (NCT02071888), including MCL, WM, TCL
AG-270 MAT2A production of S-adenosylmethionine advanced lymphoma (NCT03361358, NCT03435250)
Devimistat (CPI-613) lipoate analog mitochondrial oxidative metabolism R/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.

1.1. Targeting Metabolic Pathways

1.1.1. 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][1]. In NHL cell lines, 2-DG can inhibit cell proliferation both under hypoxia and normoxia (Table 2). When combined with methylprednisolone, 2-DG synergistically inhibits cell proliferation by downregulating HIF-1α and c-Myc [192][2]. 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][3]. 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][4]. 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][5]. 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][6][7] (Table 2). 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][8] (Table 2). 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][9][10] (Table 2). 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][10] (Table 2). 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][11][12]. 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,
]
Glutamine uptake
L-asparaginase
NHL
cell lines
[207][21]
SHMT1/2 SHIN1 DLBCL cell lines [33][22]
FASN orlistat MCL cell lines [86,208][23][24]
T-NHL cell lines, mice [209][25]
C75 DLBCL, PEL, and B-NHL cell lines [84,[2685]][27]
NA BaP BL patients [210][28]
PPARα Fenofibrate B-NHL mice [211][29]
Choline kinase CK37 T-NHL mice [42][30]
HIF-1α PX-478 PEL cell lines [115][31]
PCI-24781(HDACi) DLBCL Phase 1/2 [212][32]
SAHA(HDACi) B-NHL cell lines, mice [213][33]
MYC 10058-F4 DLBCL cell lines [214,215,216][34][35][36]
PI3K LY294002 B-NHL cell lines [217][37]
  AZD8835 ABC-DLBCL cell lines [218][38]
AKT Akti1/2 PEL cell lines [219][39]
AZD-5363 PTEN-deficient DLBCL cell lines, mice [218][38]
MK-2206 ABC-DLBCL mice [220][40]
NaB (HDACi) BL cell lines [221][41]
mTOR Rapamycin ALCL, NHL cell lines [36,131][42][43]
mTOR C1/2 AZD-2014 MCL cell lines [222][44]
Dual inhibitor of PI3K and mTOR NVP-BEZ235 PEL mice [223][45]
PF-04091502 PEL cell lines [219][39]
Bimiralisib (PQR309) DLBCL, MCL, SMZL, CLL, HL, and ALCL cell lines, mice [224][46]
AMPK AICAR MCL, SMZL, FL, and CLL cell lines, mice and patients [159,225,[226,48][227,49][228,229][47]50][51][52]
Metformin B and T-NHL cell lines, mice [157,230][53][54]
Phenformin PTEN-deficient T-cell lymphomas cell lines [230][54]
Abbreviations: 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.

1.1.2. 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][20] (Table 2). 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][55]. In NHL cell lines, L-ASNase exerts cytotoxicity by depriving the cells of glutamine, resulting in the suppression of cell growth and survival [207][21] (Table 1 and Table 2).

1.1.3. Targeting Lipid Metabolism

Pharmacological inhibition using the FASN-specific inhibitor C75 triggered caspase-dependent apoptosis in DLBCL cell lines [84][26] (Table 2). 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][27] (Table 2). 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][23] (Table 2). Orlistat can also interfere with the ubiquitination of NOXA protein and induce apoptosis, thereby offering new strategies to kill bortezomib-resistant MCL cells [208][24]. TCLs, when treated with orlistat in vitro, manifest tumor-specific inhibition of cell survival and induction of apoptosis [209][25]. 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][25] (Table 2). 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][28] (Table 2). 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][29] (Table 2).

1.2. Targeting Oncogenic Regulators

1.2.1. 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][31] (Table 2). 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][32] (Table 2). 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][33]. 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][56] (Figure 1).
Figure 1. 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.

1.2.2. 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][37]. However, 10058-F4 is ineffective in vivo due to its poor bioavailability and rapid metabolism and clearance in other cancers [214,215,216][34][35][36] (Table 2). 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][57]. 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][58][59][60]. 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][61].
The oncogenic form of HSP90 optimizes several MYC metabolic pathways, including the production of nucleotides [121][62]. 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.

1.2.3. Targeting 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][37] (Table 2). 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][63] (Table 1). 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][38] (Table 2). 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][40]. Duvelisib, an oral inhibitor of PI3K-δ/γ isoforms, demonstrated acceptable safety and promising clinical activity in patients with R/R TCLs [239][64]. In vitro, phospho-AKT (pAKT) showed a synergistic effect with duvelisib in TCL cell lines [239][64] (Table 1 and Table 2, Figure 1).

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][39]. 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][39] (Table 2). Through the downregulation of MYC, the AKT inhibitor AZD5363 induced apoptosis in PTEN-deficient DLBCLs [218][38] (Table 2). 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][40] (Table 2). 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][41] (Table 2). 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][54] (Figure 1).

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][43] (Table 2). 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][65][66]. Rapamycin also led to a reduction in GLUT1 mRNA expression, FDG uptake, and cell survival of NHL cell lines [36][42] (Table 2). 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][44] (Table 2, Figure 1).

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][45] (Table 2). 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][39] (Table 1).
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][46] (Table 2).

1.2.4. 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][67][68][69]. 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 2). The activation of AMPK promoted by metformin was associated with inhibition of the mTOR pathway without involving AKT [157][53]. 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][67][70][71][72][73]. 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][47][49][50][51][52] (Table 2). AICAR showed an acceptable safety profile in a phase I/II clinical trial involving patients with R/R chronic lymphocytic leukemia [249][74]. In MCL patients, AICAR showed a cytotoxic effect when combined with rituximab [159][47]. 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][47][48] (Table 2, Figure 1).

2. 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][75]. 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][75]. 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][76]. 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][76]. 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][77]. 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][78]. 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][79]. 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][80]. Changes in lipid profiles provide new biological insights into how MYC regulates cellular metabolism in MYC-induced lymphoma [256][81] (Table 3).

References

  1. 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.
  2. 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.
  3. 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.
  4. 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.
  5. 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.
  6. 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.
  7. 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.
  8. 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.
  9. 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.
  10. 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.
  11. 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.
  12. 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.
  13. 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.
  14. 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.
  15. 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.
  16. 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.
  17. 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.
  18. 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.
  19. 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.
  20. 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.
  21. 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.
  22. 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.
  23. 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.
  24. 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.
  25. 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.
  26. 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.
  27. 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.
  28. 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.
  29. 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.
  30. 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.
  31. 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.
  32. 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.
  33. 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.
  34. 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.
  35. 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.
  36. Whitfield, J.R.; Beaulieu, M.E.; Soucek, L. Strategies to Inhibit Myc and Their Clinical Applicability. Front. Cell Dev. Biol. 2017, 5, 10.
  37. 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.
  38. 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.
  39. 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.
  40. 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.
  41. 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.
  42. 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.
  43. 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.
  44. 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.
  45. 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.
  46. 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.
  47. 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.
  48. 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.
  49. 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.
  50. 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.
  51. 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.
  52. 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.
  53. 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.
  54. 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.
  55. 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.
  56. 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.
  57. Bhatt, R.; Ravi, D.; Evens, A.M.; Parekkadan, B. Scaffold-mediated switching of lymphoma metabolism in culture. Cancer Metab. 2022, 10, 15.
  58. 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.
  59. 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.
  60. 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.
  61. 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.
  62. 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.
  63. 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.
  64. 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.
  65. 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.
  66. 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.
  67. Drew, B.G.; Kingwell, B.A. Acadesine, an adenosine-regulating agent with the potential for widespread indications. Expert Opin. Pharmacother. 2008, 9, 2137–2144.
  68. 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.
  69. Vakana, E.; Platanias, L.C. AMPK in BCR-ABL expressing leukemias. Regulatory effects and therapeutic implications. Oncotarget 2011, 2, 1322–1328.
  70. 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.
  71. 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.
  72. 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.
  73. 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.
  74. 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.
  75. 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.
  76. 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.
  77. 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.
  78. 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.
  79. 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.
  80. 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.
  81. 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.
More