Next Article in Journal
Organismal Roles of Hsp90
Previous Article in Journal
Identification of Inhibitors of Tubulin Polymerization Using a CRISPR-Edited Cell Line with Endogenous Fluorescent Tagging of β-Tubulin and Histone H1
Previous Article in Special Issue
Metformin in Differentiated Thyroid Cancer: Molecular Pathways and Its Clinical Implications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Action Mechanism of Metformin and Its Application in Hematological Malignancy Treatments: A Review

Department of Hematology and Oncology, Zhongda Hospital, School of Medicine, Southeast University, Nanjing 210009, China
*
Author to whom correspondence should be addressed.
Biomolecules 2023, 13(2), 250; https://doi.org/10.3390/biom13020250
Submission received: 3 November 2022 / Revised: 18 January 2023 / Accepted: 20 January 2023 / Published: 29 January 2023
(This article belongs to the Special Issue Metformin and Cancer)

Abstract

:
Hematologic malignancies (HMs) mainly include acute and chronic leukemia, lymphoma, myeloma and other heterogeneous tumors that seriously threaten human life and health. The common effective treatments are radiotherapy, chemotherapy and hematopoietic stem cell transplantation (HSCT), which have limited options and are prone to tumor recurrence and (or) drug resistance. Metformin is the first-line drug for the treatment of type 2 diabetes (T2DM). Recently, studies identified the potential anti-cancer ability of metformin in both T2DM patients and patients that are non-diabetic. The latest epidemiological and preclinical studies suggested a potential benefit of metformin in the prevention and treatment of patients with HM. The mechanism may involve the activation of the adenosine monophosphate-activated protein kinase (AMPK) signaling pathway by metformin as well as other AMPK-independent pathways to exert anti-cancer properties. In addition, combining current conventional anti-cancer drugs with metformin may improve the efficacy and reduce adverse drug reactions. Therefore, metformin can also be used as an adjuvant therapeutic agent for HM. This paper highlights the anti-hyperglycemic effects and potential anti-cancer effects of metformin, and also compiles the in vitro and clinical trials of metformin as an anti-cancer and chemosensitizing agent for the treatment of HM. The need for future research on the use of metformin in the treatment of HM is indicated.

1. Introduction

Hematological malignancies (HMs) mainly include heterogeneous tumors, such as leukemia, lymphoma and myeloma [1,2], and the mechanisms are abnormal hematological cells or immune cells with blocked differentiation and indefinite proliferation followed by biological organismal dysfunction [3]. HMs are different from solid tumors, and due to their specificity, conventional treatment options mainly include chemotherapy, radiotherapy and hematopoietic stem cell transplantation (HSCT) [4]. Although the current conventional first-line treatment has a certain efficacy, primary or secondary drug resistance still often occurs, resulting in a poor overall efficacy with respect to the tumor [5]. Therefore, the search for new therapeutic targets and the development of new treatments remains an important operation with scientific significance in the research of HMs [6].
Metformin is recognized by the World Health Organization (WHO) as an essential drug for the treatment of patients with diabetes and can be widely used in the treatment of type 2 diabetes (T2DM) alone or in combination with other drugs [7]. The most classic action of metformin is to lower hyperglycemia and reduce its clinical symptoms [8]. Due to its low cost and high clinical value in terms of glycemic control and safety, it is now one of the most prescribed drugs in the market [9].
In 1995, Evans et al. first discovered that metformin not only controlled blood glucose in patients with diabetes, but also reduced the risk of malignancies [10]. Since then, various studies in patients with T2DM have shown that metformin use reduces the risk of many cancers [11,12,13,14,15,16], including prostate cancer [17,18], lung cancer [19,20,21], head and neck cancer [22,23], breast cancer [24,25], pancreatic cancer [26], colorectal cancer [26,27,28], endometrial cancer [29], ovarian cancer [30,31,32] and hepatocellular cancer [33,34]. Therefore, this review focuses on the mechanism of action of metformin in cancer prevention and the potential role of this drug in the treatment of patients with HM.
In addition, several studies have supported the therapeutic benefit of metformin in patients with HM. Preclinical data, clinical trial reports, reviews and meta-analyses have reported the effect of metformin as an anti-cancer agent in HM from various perspectives, but these articles reached different conclusions. In this paper, we further focus on the specific mechanisms involved in metformin in cancer treatments, starting with the anti-hyperglycemic effect action mechanism of metformin. Focusing on a collection of published studies on the anti-cancer mechanism of the action of metformin in HM and clinical studies, we affirm the therapeutic potential of metformin for future use in patients with HM.

2. Metformin Mechanism of Action

Metformin, a natural derivative of guanidine present in French cloves [35], was first reported to be synthesized by Werner and Bell in 1922 [36,37]. Metformin was first introduced as an anti-hypoglycemic effect agent in the treatment of diabetes by Jean Sterne in 1957 [35], and the following decades saw intense medical interest in the effects of metformin. In 1994, the U.S. Food and Drug Administration (FDA) approved the use of metformin for the treatment of T2DM [38]. In 1998, the UK Prospective Diabetes Study (UKPDS) confirmed the safety and efficacy of metformin as an initial treatment for hyperglycemia in patients with T2DM [39]. This opened an era of using metformin as the first-line treatment for T2DM.

2.1. Metformin Anti-Hyperglycemic Effects Mechanism

2.1.1. Systemic Actions of Metformin

Typically, metformin lowers the blood glucose level by inhibiting hepatic glucose production and glucagon signaling, improving insulin sensitivity and increasing peripheral glucose uptake in skeletal muscle, while not affecting glycogenolysis [40,41,42,43,44,45,46,47,48]. These effects normalize chronic hyperglycemia and hyperlipidemia and reduce the damaging effects of metabolic defects on several organs of the body [49,50,51,52].
Recent studies have shown that the gastrointestinal tract and its intestinal microorganisms are also involved in the anti-hyperglycemic effects of metformin [53,54]. Metformin increases the intestinal absorption and utilization of glucose in the intestinal cells, which in turn limits the entry of glucose into the bloodstream [55,56]. Altering the composition of the intestinal flora and the secretion of certain key molecules, such as glucagon-like peptide-1 (GLP1) and growth differentiation factor 15 (GDF15), during treatments can affect the gastrointestinal microenvironment and produce beneficial metabolic effects [57,58]. Figure 1 summarizes the mechanisms described above.

2.1.2. Cellular Actions of Metformin

AMPK-Dependent Mechanisms

Zhou et al. demonstrated that metformin can activate adenosine monophosphate-activated protein kinase (AMPK) in isolated rat hepatocytes, providing a cellular mechanism for the anti-hyperglycemic effect of metformin [59]. AMPK is an energy sensor that acts in numerous metabolic pathways involved in intracellular energy homeostasis [60].
The activation of AMPK involves many different pathways. In 2000, two independent groups reported that the inhibition of electron transport chain I (ETC I) by metformin decreased nicotinamide adenine dinucleotide (NADH) oxidation and oxygen consumption rates, leading to a decrease in adenosine triphosphate (ATP), an increase in the adenosine monophosphate (AMP)/ATP ratio, and, ultimately, the activation of AMPK [61]. In addition, the inhibition of ETC complex 1 can further inhibit oxidative phosphorylation (OXPHOS), which also causes the activation of AMPK [62]. Shaw et al. demonstrated that metformin can promote AMPK activation by mediating liver kinase B1 (LKB1), the upstream kinase for AMPKα phosphorylation at the amino acid threonine (Thr172), by binding AMPK [63]. In lysosomes, metformin binds to presenilin enhancer-2 (PEN2) and activates AMPK via the axis inhibitor (AXIN)/LKB1–v–ATPase–Ragulator pathway [64,65]. Finally, high intracellular calcium concentrations can promote the activation of calcium/calmodulin-dependent protein kinase II (CaMKII), which leads to AMPK phosphorylation [66,67].
AMPK activation can lead to the inhibition of the mammalian target of rapamycin (mTOR) signaling proteins [68,69]. mTOR is considered to be an important regulatory molecule in controlling the growth and proliferation of diabetes and cancer cells [69,70]. Activated AMPK shuts down ATP depletion, restores cellular homeostasis and converts the cell from an anabolic to catabolic state. At the same time, the metabolic activation of insulin receptor (IR) and insulin receptor substrate 1 (IRS1) increases cellular sensitivity relative to insulin [71]. AMPK phosphorylates acetyl coenzyme A carboxylase (ACC), inhibiting the conversion of acetyl coenzyme A(acetyl-CoA) to malonyl-CoA, leading to reduced hepatic lipogenesis and hepatic steatosis, and improving insulin resistance and hyperglycemia [72,73,74].

AMPK-Independent Mechanisms

The same inhibition of gluconeogenesis by metformin has been observed experimentally in mouse hepatocytes lacking the catalytic subunit of AMPK, verifying that metformin can inhibit hepatic gluconeogenesis in an AMPK-independent manner [67].
Some targets can drive metformin effects independently of AMPK, such as ETC I, II and IV; LKB1/AMPK; adenylate cyclase (AC); AMP deaminase; (reduced) nicotinamide adenine dinucleotide phosphate (NADPH) oxidase and mitochondrial glycerophosphate dehydrogenase (mGPD); hexokinase (HK); and voltage-dependent anion channel 1(VDAC1) [75,76,77,78,79,80,81,82,83]. Metformin is able to restrict cytoplasmic access to mitochondrial ATP from these channels, which in turn restricts the cell’s energy supply [84]. For example, metformin can inhibit redox-dependent substrate hepatic gluconeogenesis by blocking ETC IV [85]. Metformin enhances glucose transport to hepatocytes, which is mediated by glucose transporter protein 1 (GLUT1) via the activation of IRS 2, and reduces plasma glucose levels [85]. Metformin similarly promotes Cl efflux from hepatocytes, depolarizes cell membranes and reduces the uptake of gluconeogenic substrates such as alanine and lactate [75]. Figure 2 summarizes the mechanisms described above.

2.2. Metformin Anti-Cancer Mechanism

2.2.1. Activation of LKB1-AMPK-mTOR Signaling Pathway

It is commonly believed that metformin exerts its anti-cancer activity via the LKB1–AMPK–mTOR pathway, and the overexpression of mTOR in this pathway is often associated with the development of numerous diseases such as tumors, among others [86,87,88,89,90]. Metformin activates AMPK after the inhibition of ETC I, inhibits the expression of mTOR and further inhibits the expression of more important factors mediating downstream pro-carcinogenic pathways and the process of tumor development, such as nuclear factor kappa B (NF-κB), interleukin-6 (IL6), mitogen-activated protein kinase (MAPK), Ras and c-MYC [91,92,93]. In animal or in vitro model experiments, high doses (~mM) of metformin inhibited mitochondrial respiratory chain complex I, leading to an increase in the AMP/ATP ratio and prompting the phosphorylation and activation of the upstream kinase LKB1, while low doses of metformin (~μM) activated AMPK via the lysosomal pathway [65]. The mammalian target of rapamycin complex 1 (mTORC1) activation by metformin in the lysosomal pathway via the Ras-associated guanosine triphosphatase (GTPase) completes the regulation of AMPK, which is independent of the AMP/ATP ratio [65,94,95,96]. The activation of the LKB1–AMPK–mTOR signaling pathway can usually mediate the following cellular responses.
  • Regulation of epigenetic modifying enzyme activity [97]: AMPK directly and indirectly regulates histone acetylation and alters gene expression patterns via the epigenetic regulation of various chromatin functions [98,99]. Activated AMPK phosphorylates a variety of substrates including histone acetyltransferases (HATs), histone deacetylases (HDACs) and deoxyribonucleic acid (DNA) methyltransferases (DNMTs), often leading to their inhibition [98,99,100];
  • Activation of p53: Activated P53 promotes apoptosis, cycle arrest, autophagy and the inhibition of the phosphoinositide 3-kinase (PI3K)–protein kinase B (AKT) signaling pathway [101,102]. The effect of P53 in tumor apoptosis was also associated with the reduced expression of transcription factor specificity protein 1 (Sp1), Sp3, Sp4 and oncogenes (B-cell lymphoma-2 (BCL-2), mTOR, vascular endothelial growth factor (VEGF) and MYC, etc.) [103,104];
  • Activation or inhibition of the expression of other cancer-related signaling pathways:
    3.1
    Activation of the AMPK-forkhead box O1/3 (FOXO1/3) signaling pathway [105,106,107]: FOXO is a proliferation-associated transcription factor, and the activation of FOXO3a has been shown to be required for the pro-apoptotic and chemotherapeutic effects of metformin in a variety of tumor models [108,109];
    3.2
    Suppression of Hippo pathway expression [110,111]: In some (cancer) models, metformin has been shown to repress the activity or expression of yes-associated protein (YAP)/tea domain transcription factor (TAZ), which are two central effectors of the Hippo pathway that are mediated by AMPK [112,113,114,115,116,117];
    3.3
    Inhibition of the receptor tyrosine kinase (TK) pathway: this includes epidermal growth factor receptor (EGFR) signaling, further targeting downstream effectors AKT, mTOR, extracellular regulated protein kinases (ERK), etc. [118];
    3.4
    Inhibition of the PI3K/AKT pathway: in vitro and in vivo experiments showed that the PI3K/AKT signaling pathway inhibited by metformin stimulated the expression of phosphatase and tensin homolog (PTEN) and inositol triphosphate 3 (IP3) [119];
    3.5
    Affecting natural killer (NK) cell signaling pathways: the anti-cancer properties of specific immune regulation, the inhibition of cell proliferation and the induction of cell cycle arrest are exerted via the inhibition of tumor cell metastasis, endothelial cell proliferation and the alteration of NK cells’ ligand expression on the surface of tumor cells [120,121,122,123];
    3.6
    Stimulation of the Sirtuin1 (SIRT1) signaling pathway: the upregulation of SIRT1 by AMPK can lead to the amelioration of oxidative stress and a reduction in DNA damage [124,125];
  • Inhibition of cancer stem cells (CSCs): CSCs are cancer cells with unlimited renewal capacity. A number of studies have shown that metformin inhibits the biological activity of CSCs in a variety of tumors, including gastric, endometrial and ovarian cancers [126]. Subsequently, metformin activates additional signaling pathways for targeting CSCs via the activation of AMPK, which include PI3K–AKT–mTOR, insulin-insulin growth factor1 (IGF1), MAPK, sonic hedgehog (Shh), Wnt, TGFB, Notch, Hippo and NFkB pathways [126];
  • Reducing IRS1 phosphorylation [127]: activated AMPK inhibits hyperinsulinemia-associated tumor activity by reducing the circulating insulin levels and targeting the insulin–IGF1–PI3K signaling axis [127,128];
  • Mitigation of hypoxia and other tumor responses caused by hypoxia: Metformin leads to slow tumor growth by decreasing the expression of hypoxia-inducible factor-1α (HIF-1α), which in turn decreases the expression of HIF1-targeted genes [129,130,131,132]. The inhibition of HIF-1α simultaneously suppresses the immunosuppressive activity of myeloid-derived suppressor cells (MDSCs) and improves T-cell immunity in the tumor microenvironment (TME) [133,134,135]. These immune responses are ultimately also involved in the inhibition of angiogenesis [136];
  • Inhibition of adipogenesis: The cancer cells themselves require more nutrients and energy; thus, the rate of adipogenesis is higher [137]. AMPK inhibits lipogenesis by targeting the activity or expression of many lipogenic enzymes. On the one hand, AMPK phosphorylates and inhibits 3-hydroxy-3-methylglutaryl CoA reductase (HMGCR), which catalyzes the rate-limiting step in cholesterol synthesis. On the other hand, AMPK phosphorylates and inactivates acetyl CoA carboxylase (ACC), the main enzyme involved in the biosynthesis of fatty acid and HMGCR, resulting in the inhibition of cholesterol biosynthesis [138]. In addition, AMPK also phosphorylates sterol regulatory element binding protein-1c (SREBP-1c) at Ser-372, which restricts its cleavage and nuclear translocation. This process results in the downregulation of SREBP-1c target genes, including those encoding ACC1 and fatty acid synthase (FASN), and leads to reduced lipogenesis [139];
  • Activation of ataxia telangiectasia-mutated gene (ATM), which leads to the activation of DNA damage repair pathways and the inhibition of tumor growth [140].

2.2.2. Activation of AMPK-Independent Signaling Pathways

Some of the pathways controlled by metformin can also be modified in a manner that is not dependent on AMPK activation. Metformin can promote crosstalk between apoptosis and autophagy in multiple ways to accomplish anti-tumor effects [141]. These include downregulating the signal transducer and the activator of transcription 3 (STAT3)/Bcl2/Beclin1 signaling pathway activity in cancer cells, disrupting unfolded protein response (UPR) transcription, blocking the oncogenic Wnt signaling pathway, regulating caspase enzyme activity, increasing reactive oxygen species (ROS) production by cancer cells, inhibiting the activation of mTORC1 by Rag GTPases and inhibiting glutaminase activity [142,143,144,145,146,147,148,149]. Shen et al. found the preferential inhibition of cell-cycle-related proteins among the broad inhibitory effects of metformin on protein synthesis [150]. Specifically, the metformin activation of the p53-regulated in the development DNA damage response 1 (REDD1) axis caused an independent inhibition of AMPK by mTOR, leading to a decrease in cell-cycle protein D1 (cyclin D1) expression, which in turn inhibited the G0/G1 cell cycle [66,151]. In addition, metformin induces the expression of dicer, an enzyme that processes microRNAs (miRNAs) precursors into mature miRNAs, as a means of regulating miRNAs to exert anti-cancer effects [152]. Numerous experiments have shown that metformin inhibits tumorigenesis and tumor growth by altering the expression of specific miRNAs in human breast, pancreatic, prostate, renal cell, lung, hepatocellular, bile duct, small intestine and human oral cancers [153,154,155,156,157,158,159,160,161,162]. The repression of different miRNAs can directly or indirectly induce apoptosis, cycle arrest and other phenomena. Metformin reduces stressor-induced ROS production, activates endogenous repair systems, prevents ROS toxicity, protects cells and ameliorates genomic instability and possible cancer risk [163]. Metformin reduces the expression of the P-glycoprotein (P-gp) encoded by multidrug resistance1 (MDR1), and, via this action, it blocks drug efflux from cancer cells, allowing metformin to be used as an adjuvant relative to classical chemotherapy [164].

2.2.3. Altered Energy Metabolism of Tumors

Cancer cells have a different pathway for energy production. Even under aerobic conditions, the tricarboxylic acid cycle (TCA cycle) and OXPHOS are restricted, and they favor lactic acid fermentation, called the “Warburg effect” [165,166]. Many studies observed that mitochondrial OXPHOS is the main source of ATP in most cancer tissues [167]. Metformin can reduce the availability of glucose to cancer cells by causing a metabolic shift from OXPHOS to glycolysis via the inhibition of mitochondrial function [61]. Metformin targets mitochondrial glycerol 3-phosphate dehydrogenase (mGPDH), causing the inhibition of mitochondrial function while impairing the production of dihydroxyacetone phosphate (DHAP), which can reduce the process of glucose production from glycerol [80]. In addition, metformin targets mitochondrial integrity by modulating calcium flux in cancer cells [168]. Inducing endoplasmic reticulum stress and releasing calcium into the cytoplasm of cells induces higher calcium uptake by mitochondria, ultimately leading to mitochondrial swelling.

2.2.4. Improved Immune and Inflammatory Response

Inflammatory cells are essential players in the tumor process, promoting the proliferation, survival and migration of tumor cells [169]. In addition to targeting cancer cells, metformin has been shown to target immune cells in the tumor microenvironment, such as CD8+ T cell, regulatory cells (Tregs), MDSCs and tumor-associated macrophages (TAMs), which may contribute to the expression of anti-tumor activities [170].
The NF-κB signaling pathway is a key regulator of immunity and plays an important role in tumor progression. Metformin inhibits the NF-κB signaling pathway and reduces the levels of inflammatory cytokines including tumor necrosis factor-α (TNF-α) and IL-6 [171]. In addition, metformin inhibits ETC I, further inhibiting ROS and interleukin-1β (IL-1β) production via lipopolysaccharide (LPS)-activated macrophages [172]. Metformin can directly and indirectly alter the function of CD8+ T cells, blocking T cell depletion by increasing the number of CD8+ T lymphocytes (CD8+ TILs) and protecting them from apoptosis and exhaustion [173,174]. In addition, metformin can reduce programmed cell death protein 1 ligand 1 (PD-L1) in tumor cells via multiple pathways, leading to enhanced cytotoxic T lymphocyte (CTL) activity [175]. In addition, metformin helps the body produce sustained anti-tumor immunity by acting on multiple immune pathways.

2.2.5. Reduced Tumor Vascular Metastasis and Invasion

Several studies investigated the mechanism by which metformin inhibits cancer invasion and metastasis by inhibiting the EMT signaling pathway [176]. Metformin inhibits the epithelial to mesenchymal transition (EMT) signaling pathway via multiple pathways, including blocking cyclooxygenase-2 (COX-2)-mediated prostaglandin E2 (PGE2) production, reducing Snail protein expression, inhibiting B-cell-specific Moloney murine leukemia virus integration region 1 (Bmi-1, also known as RNF51 or PCGF4) and reducing IL-6 [177,178,179,180,181,182]. In addition, metformin inhibits angiogenesis and attenuates tumor migration by attenuating angiogenic stimulation, downregulating VEGF expression, inhibiting HIF-1α-induced angiogenesis-related factor expression and reducing the signaling of platelet-derived growth factor B (PDGF-B) and PDGF-receptorβ (PDGF-Rβ) [183,184,185,186]. Figure 3 summarizes the mechanisms described above.

3. Therapeutic Application of Metformin in Hematological Malignancies

A large number of different types of studies have shown that metformin is a potential therapeutic target for patients with leukemia, lymphoma and multiple myeloma, and it is expected that the pleiotropic effects of this drug could act on multiple targets [53,187,188,189,190,191,192,193,194]. Herein, we discuss experimental ex vivo studies and clinical observations of metformin in different hematologic malignancies. We compiled tables of preclinical applications of metformin in malignant hematological during the last decade, divided into in vivo animal experiments (Table 1), in vitro cellular experiments in leukemia (Table 2), lymphoma (Table 3) and myeloma (Table 4).

3.1. Leukemia

3.1.1. Acute Myeloid Leukemia (AML)

AML is the most common acute leukemia in adults and is characterized by the uncontrolled proliferation of leukemic cells that inhibit normal bone marrow hematopoiesis following varying degrees of differentiation blockage with respect to bone marrow progenitor cells [239].
Since metformin was shown to interfere with the proliferation and clonal activity of AML cells without affecting normal CD34+ HSCs, numerous experiments have begun to further explore the anti-AML effects of metformin [240]. Metformin accelerates AML cell apoptosis by inhibiting OXPHOS, decreasing FOXM1, activating the LKB1–AMPK signaling pathway and dephosphorylating eukaryotic initiation factor 4E binding protein 1 (4E-BP1) [213,224,241,242]. In acute progranulocytic leukemia (APL) cell lines, metformin induced c-Myc degradation and apoptosis via the MEK–ERK signaling pathway [243]. In addition, metformin combined with sorafenib, cytarabine and venetoclax showed synergistic enhancement effects in AML treatment, and these effects were demonstrated by mTOR inhibition and the regulation of apoptosis-related proteins [196,199,221]. Clinically, a retrospective hospital cohort study found that although metformin users did not perform better than non-users in overall survival (OS) and disease-free status, they performed better than insulin and oral diabetes medication-treated AML patients [244].

3.1.2. Chronic Myeloid Leukemia (CML)

CML is characterized by the presence of the Philadelphia (Ph) chromosome with a reciprocal translocation between chromosomes 9 and 22, and this translocation generates the BCR/ABL fusion gene [245]. A tyrosine kinase is produced by the BCR/ABL fusion gene, which drives tumorigenesis and prevents apoptosis by activating downstream signals [246].
After AMPK activation, metformin inhibited the proliferation and clonal activity of different CML cell lines, including the T315I BCR-ABL mutant CML line expressing imatinib resistance [247]. With regard to drug combinations, experiments have shown that the combination of nilotinib and metformin is more effective than the combination of nilotinib and c-Jun N-terminal kinase (JNK) inhibitors [208].
An observational study of patients with newly diagnosed chronic-phase CML showed that the proportion of patients reaching complete cytogenetic response (CCyR) was 100% and 73.6% in the metformin and non-metformin groups, respectively, with similar numbers reaching major molecular response (MMR) and complete molecular response (CMR) between the two groups. The median time for achieving MMR and CMR was shorter in the metformin group compared to the non-metformin group (11.1 months vs. 19.5 months; 37.4 months vs. NR), indicating that metformin used together with TKI therapy may increase the proportion of CML patients achieving CCyR and shorten the time to MMR and CMR [248].

3.1.3. Acute Lymphocytic Leukemia (ALL)

ALL is aggressive, with the clonal proliferation of immature lymphocytes at different stages of differentiation that, if not controlled, can cause a destructive infiltration of the bone marrow [249]. In adults, 75% of cases develop from the precursors of the B-cell lineage and the rest consist of malignant T-cell precursors [250].
Metformin increases the chemosensitivity of ALL by inhibiting the AKT–mTOR pathway via the activation of AMPK [251,252]. In T-ALL cells, metformin-stimulated AMPK inhibited mTOR to trigger cellular autophagic responses by a mechanism involving miRNA-19 overexpression [225,253]. Metformin-triggered apoptotic effects also involve the AMPK-dependent activation of the endoplasmic reticulum (ER) stress/UPR cell death pathway and the apoptotic mediators including inositol-requiring enzyme1 (IRE1) and C/EBP-homologous protein (CHOP) [225]. In addition to AMPK, the expression of other important metabolic kinases is also regulated by metformin in ALL cells. Metformin regulates the expression of protein kinase C-ε(PKCε) and δ in responding cells, leading to the downregulation of PKCδ and control of PKCε [254]. The altered balance between PKCε and PKCδ has important implications for tumor cell survival, as both kinases control the nuclear factor erythroid 2-related factor (Nrf2) transcription factor [255]. In adult T-cell leukemia (ATL), metformin activates LKB1–AMPK and inhibits leukemic proliferation by reducing Tax expression [256]. Leukemic stem cells (LSCs) are rare cells with leukemic origins that have an intrinsic resistance mechanism relative to chemotherapy [257]. In T-ALL cells, metformin can target the Hoescht 33342low side population and the CD34+CD7CD4 subpopulation, which are rich in LSCs [251].
In ALL, metformin can be used in combination with anthracyclines (doxorubicin and adriamycin), dexamethasone, imatinib, proviral integration sites for Moloney murine leukemia virus (PIM)-1/2 kinase inhibitors, thiazolidinediones and other drugs to exert synergistic effects, reducing tumor cell growth and survival by stimulating various signaling pathways and showing a cumulative effect [193,203,222,225,251,252]. The use of metformin in ALL treatment can help reduce the dose of adriamycin that is necessary to prolong remission and reduce the cardiotoxicity of anthracyclines [258]. There are, of course, different phenomena. Studies have shown that the combination of metformin and cisplatin exerts an antagonistic effect in ALL, suggesting that metformin and oncology drugs are not synergistic in one way or the other [206].
In a prospective study of 102 patients with new-onset Ph B-cell leukemia (NCT03118128), metformin combined with chemotherapy was shown to be effective in ALL patients with elevated ABCB1 gene expression levels (p = 0.025), and in the metformin user group the drug was protective against treatment failure (odds ratio (OR) 0.07, 95% confidence interval (CI) 0.0037–1.53) and early relapse (OR 0.05, 95% CI 0.0028–1.153) in the metformin user group [259]. A phase 1 study (2018) exploring metformin in combination with induction chemotherapy for relapsed/refractory (R/R) ALL showed that the combination of metformin and vincristine, dexamethasone, PEG-asparaginase and doxorubicin (VXLD) was tolerable in the 14 included patients; the overall treatment induced ER stress; and toxic reactions occurred at all dose levels of metformin [260]. In the randomized controlled trial of an intensive insulin regimen in patients with hyperglycemic ALL (NCT00500240), the trial was terminated early, although this ALL clinical outcome was invalid. However, a secondary analysis showed that the use of metformin and/or thiazolidinedione was associated with increased progression-free survival (PFS) compared to other drug combinations and may improve the patient prognosis [261]. In a small phase I clinical trial in patients with ALL, the addition of metformin to standard chemotherapy was well tolerated. The unselected use of metformin during this treatment was the only variable considered to have a significant adverse prognosis (p = 0.55). The Cox regression showed that the addition of metformin reduced the risk of recurrence by 56% [262]. A retrospective analysis reported on the safety of the metformin treatment for hyperglycemia control in 17 patients with ALL during the induction period. The data showed no significant toxicity of metformin in all patients, with only one patient failing to control glycemia, suggesting that metformin can be used in clinical ALL patients for glycemic control [263].

3.1.4. Chronic Lymphocytic Leukemia (CLL)

CLL is a common leukemia in western countries, and metformin has been found to be cytotoxic to both resting and activated CLL cells [264]. The cytostatic effect of metformin on activated CLL cells was manifested by the reduced expression of proliferation and proliferation-related molecules. In these cells, metformin inhibited the activation-induced upregulation of chemokine receptors and adhesion molecules that can synergistically stimulate mitogenic activity and cell homing in lymphoid tissues [265]. Metformin inhibits the activation of transcription factors in the CLL pro-survival and pro-activation pathways by activating AMPK and reducing glucose metabolism, further impeding CLL cell survival, activation and clonal expansion [265,266,267,268,269,270]. CLL cells are more dependent on the protective activity of BCL-2 anti-apoptotic protein than normal cells. This also explains the higher sensitivity of CLL cells to metformin than normal lymphocytes [271]. Metformin modulates the glycolytic capacity of CLL cells and limits the availability of extracellular glucose, thereby increasing the sensitivity of CLL cells relative to the cytostatic effects of metformin [264]. The combination of classical anti-CLL drugs fludarabine, ritonavir and ABT-737 with metformin may enhance its cytotoxicity by altering mitochondrial activities and other mechanisms of action, and can exert synergistic effects [53,272].

3.2. Multiple Myeloma (MM)

MM is a malignant plasma cell proliferative disease with major symptoms including elevated blood calcium levels, renal insufficiency, anemia and bone damage (CRAB) [273]. In vitro and in vivo experiments demonstrated a direct anti-tumor effect of metformin on myeloma [92,191,194,201,202,235,237,274].
Metformin inhibits the proliferation of myeloma cells by stimulating apoptosis and cell-cycle arrest by a mechanism involving the dual inhibition of the AMPK activation-mediated mTORC1 and mTORC2 pathways [201]. In terms of cell proliferation, metformin inhibits the expression of a range of signaling pathways such as IGF1, PI3K, AKT and downstream mammalian rapamycin [190]. Metformin reduces insulin and IGF1 production, increases insulin sensitivity and is thought to reduce the progression of monoclonal gammopathy of undetermined significance (MGUS) relative to multiple myeloma [275,276]. Metformin specifically reduced IL-6R expression via AMPK, is mediated by mTOR and showed synergistic effects with common anti-myeloma drugs [92]. Hypoxia stimulates the migration and homing of malignant plasma cells relative to new bone marrow niches, promoting osteolytic bone destruction and MM-induced angiogenesis [277,278,279,280]. Metformin prevents HIF-1α accumulation in hypoxia and slows disease progression by inhibiting the activity of ETC I [131,281,282]. In addition, cellular experiments demonstrated that metformin could exert anti-MM cellular effects by downregulating the STAT3 signaling pathway; activating the ERK1/2, MAPK and NF-κB pathways; blocking glucose-regulated protein 78 (GRP78) promoter reporter activation; inhibiting UPR activation; and so on [194,238,283].
The combination of metformin and dexamethasone synergistically eliminates MM cells by reducing AKT–mTOR signaling and inhibits cell proliferation [191]. In animal models, metformin significantly synergized with ritonavir to reduce the tumor burden of multiple myeloma [53]. Concomitant treatments with metformin and bortezomib inhibit the effect of UPR on GRP78, impair autophagosome formation and increase apoptosis [194].
Several studies have shown that being overweight, having high insulin levels and a history of diabetes are poor prognostic indicators for patients with multiple myeloma [284]. A retrospective study of 1240 patients with MM evaluated the impact of diabetes mellitus (DM) and antidiabetic drugs on clinical outcomes in MM, and the data suggested that metformin is associated with reduced death in progressive MM and that it improves MM-related prognosis [192]. In a retrospective study of 3287 U.S. veterans with diabetes diagnosed with MGUS, metformin users demonstrated the potential for a lower risk of progression relative to multiple myeloma than non-users (hazard ratio (HR) 0.47, 95%CI 0.25–0.87) [189]. These findings were extended by a case control study by Boursi et al. Using information from a U.K. database, Boursi et al. found that antidiabetic drugs may be associated with a reduced risk of active MM in patients with MGUS diabetes and with a reduced but not significant risk of MM among patients with diabetes exposed to metformin (OR 0.39, 95%CI 0.14–1.13), with the effect being observed mainly in patients who had been on treatment for at least 2 years, rather than those who had been on treatment for a shorter period of time [190]. Of course, there are studies that have come to different conclusions. One study of 2363 MGUS patients and 9193 controls showed no significant reductions in the risk of MGUS in patients with diabetes treated with metformin alone compared to those without diabetes (OR 0.77, 95%CI 0.56–1.05) [285].

3.3. Lymphoma

The incidence of lymphoma is currently increasing year by year [286]. In patients who have relapsed or are resistant to conventional chemotherapy, especially those with an aggressive subtype, the prognosis remains suboptimal. Shi et al. provided evidence of the ex vivo activity of metformin acting on human lymphoma cells, demonstrating that the metformin activation of AMPK inhibits the growth of B- and T-cell lymphomas via the inhibition of mTOR signaling, suggesting that metformin can be used in the treatment of lymphoma [193]. At the same time, metformin can enhance the tumor suppressive effect of common drugs for treatments such as asdriamycin, dacarbazine, venetoclax and rituximab [200,230,287].
Diffuse large B-cell lymphoma (DLBCL) is the most common histologic subtype of non-Hodgkin’s lymphoma (NHL) [288]. The mechanisms by which metformin affects DLBCL cell proliferation are multifactorial. A portion of DLBCL survival and progression is highly dependent on B-cell receptor (BCR) signaling [289,290,291]. Metformin blocks cholesterol synthesis by inhibiting the key cholesterol synthesis-related enzyme hydroxy-3-methylglutaryl coenzyme A synthetase 1 (HMGCS1), which ultimately leads to the intracellular inhibition of cholesterol-dependent BCR and its downstream signaling, resulting in a blockage of BCR signaling [292]. Metformin acts on lymphoma cells to produce an increase in reactive oxygen species, and it promotes the loss of mitochondrial membrane permeability [200]. Metformin induced a G1 cell-cycle block and showed a decrease in cellular proteins C-MYC, cyclin dependent kinase 2 (CDK2), E2F and proliferating cell nuclear antigen (PCNA) [200]. Burkitt’s lymphoma (BL) is a rare and aggressive subtype of NHL [293]. Metformin was found to strongly affect the viability of DAUDI cells (a human BL cell line), and peripheral blood lymphocytes (PBL) were unaffected by treatments, suggesting that metformin has the ability to selectively target cancer cells [228]. BL exhibits a specific metabolic shift towards aerobic glycolysis, leading to a strong change in metformin that can kill BL cells by determining glucose use [228].
The results of clinical studies on the role of metformin in lymphoma are mixed. A study by Wang et al. found no evidence of significant improvements in event-free (EFS), lymphoma-specific (LSS) and OS-associated metformin use in terms of prognostic improvement in newly diagnosed DLBCL and follicular lymphoma (FL) [294]. In addition, two large retrospective studies and one case–control study involving the action of metformin in patients with lymphoma showed similar results in that metformin exposure as a pre-chemotherapy agent had no significant effects on lymphoma survival [188,295,296].
There are other studies that disagreed with the above observations. A retrospective study on metformin action in patients with DLBCL in T2DM showed that the use of metformin did help significantly improve CR rates and ORR, in addition to PFS. In addition, the metformin group showed a trend toward improved OS, although the difference did not reach a statistically significant level [292]. A retrospective analysis of DLBCL patients receiving first-line chemotherapy showed that the use of metformin improved the PFS from 60 to 90 months and OS from 71 to 100 months in patients with diabetes compared with those not using metformin [200]. In a 10-year study in Taiwan on whether the use of metformin affects the risk of NHL, the risk of NHL was significantly lower in patients with T2DM treated with metformin-based antidiabetic drugs compared with non-metformin antidiabetic drugs [297]. One team also found significantly higher overall long-term survival for lymphoma in the metformin group (5.89 vs. 1.29 years, p < 0.001) than in the non-metformin group. Metformin was associated with a significant positive impact in improving overall survival in patients with type 2 diabetes and lymphoma [298]. In patients with DLBCL with DM, metformin improved the PFS (94 vs. 55.4 mon, p = 0.007) and OS (100 vs. 70.5 mon, p = 0.031) compared with other anti-diabetic agents [299]. A prospective study by Dr. Zhao’s group also found that the use of metformin prolonged survival in DLBCL [300].

3.4. Ongoing Studies

According to ClinicalTrials.gov, there are currently 17 trials of metformin and HM. Metformin is being investigated in HM mainly as an adjuvant therapy. Five of these seventeen studies have been completed. These studies are shown specifically in Table 5.

4. Adverse Effects of Metformin

Metformin can cause various adverse reactions, which cannot be ignored in the treatment process. Digestive disturbances (nausea, vomiting, diarrhea, stomach upset, etc.) are the most common side effects of metformin and, while severe symptoms may lead to discontinuation in 5–10% of metformin users, these side effects are usually transient and mild [301,302,303]. This can be easily resolved by gradually titrating the dose, taking the drug with food or switching to an extended-release formulation [304]. The effects of metformin on gastrointestinal function may lead to vitamin B12 deficiency [305]. A cross-sectional study found a significant correlation between metformin ≥ 1500 mg/d and vitamin B12 deficiency [306]. This effect was more common in patients who had been on high doses for three years [307]. It can be reversed by discontinuation or calcium supplementation [308].
The most serious adverse effect is lactic acidosis, but this complication is very rare, with an incidence of 1/30,000, and it is mainly seen in patients with diabetes with renal and hepatic abnormalities [309]. Metformin increases plasma lactic acid levels by inhibiting mitochondrial respiration in the liver and inhibiting the conversion of lactic acid into glucose in the liver [310,311]. Therefore, excessive metformin can lead to a severe hypoglycemia. The true role of metformin in the development of metformin-associated lactic acidosis (MALA) in the clinical setting is unclear. Some trials have shown that the incidence of lactic acidosis in metformin users is not significantly different from that in patients treated with other oral anti-hyperglycemic effects agents [312]. The risk factors for MALA include impaired renal function resulting in poor metformin clearance, impaired liver function resulting in poor lactate clearance and increased lactate production, such as sepsis [313]. Overall, mortality in patients taking metformin or even developing lactic acidosis appears to be related to the underlying disease rather than metformin being a toxic drug [314]. As metformin is excreted by the kidneys, chronic kidney disease is considered a major risk factor for lactic acidosis [75,315]. There is insufficient evidence to support the safe use of metformin in people with eGFR values below 30 mL/min/1.73 m2. Therefore, it is generally recommended to discontinue metformin if renal function is below this level [316].
In-depth studies have identified several unexpected risks associated with metformin, such as birth defects in offspring, cognitive deficits in rodents, acute hepatitis, cholestasis and hemolytic anemia [317,318,319]. These require further confirmation of adverse effect results and revision of clinical guidelines for precisely defined indications and doses of metformin. A fact sheet issued to the public by the UK Teratology Information Service states that there are no known risks associated with the use of metformin during pregnancy [320]. However, there is limited evidence that metformin can affect anthropometric parameters in offspring with unknown long-term consequences. Therefore, the risks and benefits of continuing metformin during pregnancy should be carefully evaluated and individualized [321]. Metformin therapy appears to be safe for most patients in the early stages of COVID-19 infection. However, immediate discontinuation of metformin as soon as COVID-19 infection becomes severe is a reasonable and practical preventive measure for metformin. This approach may reduce the likelihood of life-threatening complications from COVID-19 or lactic acidosis events associated with metformin [322].

5. Limitations of Metformin in Cancer Treatment

Although some studies support the effectiveness and feasibility of metformin for the treatment of cancer, metformin is primarily used in patients with diabetes. In clinical use, the feasibility of metformin alone or in combination with other cancer treatment regimens may be limited to patients with non-diabetes because it may result in adverse effects that are intolerable to some patients [323]. For example, patients receiving both metformin and radiation therapy experienced increased local toxicity, increased frequency of treatment interruptions and desquamation or dermatitis [324,325,326]. The report also identified time-related biases in observational studies that tended to exaggerate metformin’s anti-tumor effects [327,328]. The anti-tumor effects of metformin increased with increasing doses, as compiled from experimental data. Many experiments have used metformin at concentrations much higher than the conventional dose for diabetes treatment [329]. These high doses are not suitable for actual clinical use due to potential drug toxicity. At the same time, the efficacy of metformin as an anti-cancer drug is highly dependent on the glucose concentration in the tumor microenvironment [330,331].
In addition, many factors influence the effectiveness and responsiveness of metformin in cancer therapy. For example, few studies on metformin transporters, OCT, PMAT and MATE have provided insight into the accumulation of metformin within cancer cells to determine the specificity for cancer cells and whether normal non-cancerous cells are immune to these effects of metformin. The tissue expression of transporters mediating metformin uptake varies between normal and tumor cells and may be affected by various drugs. Poor uptake of metformin in target cells may limit its potential to treat cancer [90]. The heterogeneity of transporter expression in different cancers and how to obtain safe and effective doses of metformin in the treatment of cancer are areas of interest [332]. As with other chemotherapeutic treatment modalities, the long-term administration of metformin may be challenged by treatment-related drug resistance. Reports show that MCF7 breast cancer cells become cross-resistant to metformin and tamoxifen after long-term metformin therapy and are dependent on constitutive activation of Akt-Snail1-E-cadherin signaling [333]. Existing studies have shown inconsistent marker expression and survival outcomes for metformin as an anti-cancer agent in different settings. Variations in study design and potential biases, such as time affected by prior treatment and other relevant confounding factors, contributed to the variability of results [334].
Multiple factors influence the sensitivity of cancer cells to metformin, such as genetics, microenvironment, metabolic environment, biodistribution and tissue specificity [170]. It was demonstrated that the culture medium altered the sensitivity of cancer cells to metformin, as cells cultured in DMEM required up to 10 mM metformin to inhibit proliferation, whereas tumor cells cultured in an RPMI medium required a lower dose of metformin. At low glucose concentrations in cells lacking AMPK/LKB1, metformin is more sensitive to growth inhibition. Dietary restriction with intermittent fasting has been shown to enhance the response to metformin and metformin appears to impair tumor growth only when administered during fasting-induced hypoglycemia [335]. These are all issues that should be considered when treating cancer with metformin.

6. Conclusions

Metformin appears to be a new hope for HM treatment, and metformin acts as a pleiotropic compound that works together to prevent tumorigenesis via various pathways. The anti-cancer mechanism introduces the modulatory effects of metformin on multiple signaling pathways, including the activation of AMPK-dependent and AMPK-independent pathways, the improvement of cancer cell microenvironment, the amelioration of inflammatory responses and a reduction in tumor vascular metastasis. In addition, combining current anti-cancer drugs with metformin may improve their efficacy and reduce adverse drug reactions. Therefore, metformin can also be further investigated as an adjuvant therapeutic agent for HM. The field of oncology is among the most rigorous fields that use the gold standard of evidence in randomized controlled trials. Therefore, large, randomized, double-blind and placebo-controlled studies are necessary in order to find the best combination strategy by identifying the optimal dose of metformin, combination of drugs to avoid adverse events and potential biomarkers to implement this strategy in a precise oncological context. Ultimately, the efficacy of metformin in HM will be tested and the utility of the drug will be clarified.

Author Contributions

Conceptualization, writing of original draft: Y.Z.; review and editing of the manuscript: Y.Z., F.Z., J.G., L.Z. and B.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by a Key Medical of Jiangsu Province grant funded by the Jiangsu Provincial Health and Family Planning Commission, grant number ZDXKB2016020.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liu, J.; Bell, C.; Campbell, V.; DeBacker, J.; Tamberg, E.; Lee, C.; Mehta, S. Noninvasive Ventilation in Patients With Hematologic Malignancy. J. Intensive Care Med. 2017, 885066617690725. [Google Scholar]
  2. Al-Dorzi, H.M.; Al Orainni, H.; Al Eid, F.; Tlayjeh, H.; Itani, A.; Al Hejazi, A.; Arabi, Y.M. Characteristics and predictors of mortality of patients with hematologic malignancies requiring invasive mechanical ventilation. Ann. Thorac. Med. 2017, 12, 259–265. [Google Scholar] [CrossRef]
  3. Zhang, J.; Liu, Y.; Li, Q.; Xu, A.; Hu, Y.; Sun, C. Ferroptosis in hematological malignancies and its potential network with abnormal tumor metabolism. Biomed. Pharmacother. 2022, 148, 112747. [Google Scholar] [CrossRef] [PubMed]
  4. Eppert, K.; Takenaka, K.; Lechman, E.R.; Waldron, L.; Nilsson, B.; van Galen, P.; Metzeler, K.H.; Poeppl, A.; Ling, V.; Beyene, J.; et al. Stem cell gene expression programs influence clinical outcome in human leukemia. Nat. Med. 2011, 17, 1086–1093. [Google Scholar] [CrossRef]
  5. Westin, J.; Sehn, L.H. CAR T cells as a second-line therapy for large B-cell lymphoma: A paradigm shift? Blood 2022, 139, 2737–2746. [Google Scholar] [CrossRef] [PubMed]
  6. Lv, H.; Zhen, C.; Liu, J.; Shang, P. β-Phenethyl Isothiocyanate Induces Cell Death in Human Osteosarcoma through Altering Iron Metabolism, Disturbing the Redox Balance, and Activating the MAPK Signaling Pathway. Oxid. Med. Cell. Longev. 2020, 2020, 5021983. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. World Health Organization. Model List of Essential Medicines. 22nd List. 2021. Available online: https://www.who.int/publications/i/item/WHO-MHP-HPS-EML-2021.02 (accessed on 2 November 2022).
  8. DeFronzo, R.A.; Goodman, A.M.; Group, M.M.S. Efficacy of metformin in patients with non-insulin-dependent diabetes mellitus. N. Engl. J. Med. 1995, 333, 541–549. [Google Scholar] [CrossRef] [Green Version]
  9. Maruthur, N.M.; Tseng, E.; Hutfless, S.; Wilson, L.M.; Suarez-Cuervo, C.; Berger, Z.; Chu, Y.; Iyoha, E.; Segal, J.B.; Bolen, S. Diabetes Medications as Monotherapy or Metformin-Based Combination Therapy for Type 2 Diabetes: A Systematic Review and Meta-analysis. Ann. Intern. Med. 2016, 164, 740–751. [Google Scholar] [CrossRef]
  10. Evans, J.M.; Donnelly, L.A.; Emslie-Smith, A.M.; Alessi, D.R.; Morris, A.D. Metformin and reduced risk of cancer in diabetic patients. BMJ 2005, 330, 1304–1305. [Google Scholar] [CrossRef] [Green Version]
  11. Zhang, Z.J.; Li, S. The prognostic value of metformin for cancer patients with concurrent diabetes: A systematic review and meta-analysis. Diabetes Obes. Metab. 2014, 16, 707–710. [Google Scholar] [CrossRef]
  12. Zhang, P.; Li, H.; Tan, X.; Chen, L.; Wang, S. Association of metformin use with cancer incidence and mortality: A meta-analysis. Cancer Epidemiol. 2013, 37, 207–218. [Google Scholar] [CrossRef] [PubMed]
  13. Noto, H.; Goto, A.; Tsujimoto, T.; Noda, M. Cancer risk in diabetic patients treated with metformin: A systematic review and meta-analysis. PLoS ONE 2012, 7, e33411. [Google Scholar] [CrossRef]
  14. Franciosi, M.; Lucisano, G.; Lapice, E.; Strippoli, G.F.; Pellegrini, F.; Nicolucci, A. Metformin therapy and risk of cancer in patients with type 2 diabetes: Systematic review. PLoS ONE 2013, 8, e71583. [Google Scholar] [CrossRef] [PubMed]
  15. Gandini, S.; Puntoni, M.; Heckman-Stoddard, B.M.; Dunn, B.K.; Ford, L.; DeCensi, A.; Szabo, E. Metformin and cancer risk and mortality: A systematic review and meta-analysis taking into account biases and confounders. Cancer Prev. Res. 2014, 7, 867–885. [Google Scholar] [CrossRef] [Green Version]
  16. Yin, M.; Zhou, J.; Gorak, E.J.; Quddus, F. Metformin is associated with survival benefit in cancer patients with concurrent type 2 diabetes: A systematic review and meta-analysis. Oncologist 2013, 18, 1248–1255. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Deng, D.; Yang, Y.; Tang, X.; Skrip, L.; Qiu, J.; Wang, Y.; Zhang, F. Association between metformin therapy and incidence, recurrence and mortality of prostate cancer: Evidence from a meta-analysis. Diabetes Metab. Res. Rev. 2015, 31, 595–602. [Google Scholar] [CrossRef] [PubMed]
  18. Raval, A.D.; Thakker, D.; Vyas, A.; Salkini, M.; Madhavan, S.; Sambamoorthi, U. Impact of metformin on clinical outcomes among men with prostate cancer: A systematic review and meta-analysis. Prostate Cancer Prostatic Dis. 2015, 18, 110–121. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Zhu, N.; Zhang, Y.; Gong, Y.I.; He, J.; Chen, X. Metformin and lung cancer risk of patients with type 2 diabetes mellitus: A meta-analysis. Biomed. Rep. 2015, 3, 235–241. [Google Scholar] [CrossRef] [Green Version]
  20. Sakoda, L.C.; Ferrara, A.; Achacoso, N.S.; Peng, T.; Ehrlich, S.F.; Quesenberry, C.P., Jr.; Habel, L.A. Metformin use and lung cancer risk in patients with diabetes. Cancer Prev. Res. 2015, 8, 174–179. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Lin, J.J.; Gallagher, E.J.; Sigel, K.; Mhango, G.; Galsky, M.D.; Smith, C.B.; LeRoith, D.; Wisnivesky, J.P. Survival of patients with stage IV lung cancer with diabetes treated with metformin. Am. J. Respir. Crit. Care Med. 2015, 191, 448–454. [Google Scholar] [CrossRef] [Green Version]
  22. Sandulache, V.C.; Hamblin, J.S.; Skinner, H.D.; Kubik, M.W.; Myers, J.N.; Zevallos, J.P. Association between metformin use and improved survival in patients with laryngeal squamous cell carcinoma. Head Neck 2014, 36, 1039–1043. [Google Scholar] [CrossRef] [PubMed]
  23. Yen, Y.C.; Lin, C.; Lin, S.W.; Lin, Y.S.; Weng, S.F. Effect of metformin on the incidence of head and neck cancer in diabetics. Head Neck 2015, 37, 1268–1273. [Google Scholar] [CrossRef] [PubMed]
  24. Col, N.F.; Ochs, L.; Springmann, V.; Aragaki, A.K.; Chlebowski, R.T. Metformin and breast cancer risk: A meta-analysis and critical literature review. Breast Cancer Res. Treat. 2012, 135, 639–646. [Google Scholar] [CrossRef] [PubMed]
  25. Xu, H.; Chen, K.; Jia, X.; Tian, Y.; Dai, Y.; Li, D.; Xie, J.; Tao, M.; Mao, Y. Metformin Use Is Associated With Better Survival of Breast Cancer Patients With Diabetes: A Meta-Analysis. Oncologist 2015, 20, 1236–1244. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Wang, Z.; Lai, S.T.; Xie, L.; Zhao, J.D.; Ma, N.Y.; Zhu, J.; Ren, Z.G.; Jiang, G.L. Metformin is associated with reduced risk of pancreatic cancer in patients with type 2 diabetes mellitus: A systematic review and meta-analysis. Diabetes Res. Clin. Pract. 2014, 106, 19–26. [Google Scholar] [CrossRef]
  27. Fransgaard, T.; Thygesen, L.C.; Gögenur, I. Metformin Increases Overall Survival in Patients with Diabetes Undergoing Surgery for Colorectal Cancer. Ann. Surg. Oncol. 2016, 23, 1569–1575. [Google Scholar] [CrossRef]
  28. Mei, Z.B.; Zhang, Z.J.; Liu, C.Y.; Liu, Y.; Cui, A.; Liang, Z.L.; Wang, G.H.; Cui, L. Survival benefits of metformin for colorectal cancer patients with diabetes: A systematic review and meta-analysis. PLoS ONE 2014, 9, e91818. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Nevadunsky, N.S.; Van Arsdale, A.; Strickler, H.D.; Moadel, A.; Kaur, G.; Frimer, M.; Conroy, E.; Goldberg, G.L.; Einstein, M.H. Metformin use and endometrial cancer survival. Gynecol. Oncol. 2014, 132, 236–240. [Google Scholar] [CrossRef]
  30. Dilokthornsakul, P.; Chaiyakunapruk, N.; Termrungruanglert, W.; Pratoomsoot, C.; Saokaew, S.; Sruamsiri, R. The effects of metformin on ovarian cancer: A systematic review. Int. J. Gynecol. Cancer 2013, 23, 1544–1551. [Google Scholar] [CrossRef]
  31. Romero, I.L.; McCormick, A.; McEwen, K.A.; Park, S.; Karrison, T.; Yamada, S.D.; Pannain, S.; Lengyel, E. Relationship of type II diabetes and metformin use to ovarian cancer progression, survival, and chemosensitivity. Obstet. Gynecol. 2012, 119, 61–67. [Google Scholar] [CrossRef] [Green Version]
  32. Kumar, S.; Meuter, A.; Thapa, P.; Langstraat, C.; Giri, S.; Chien, J.; Rattan, R.; Cliby, W.; Shridhar, V. Metformin intake is associated with better survival in ovarian cancer: A case-control study. Cancer 2013, 119, 555–562. [Google Scholar] [CrossRef]
  33. Zhang, Z.J.; Zheng, Z.J.; Shi, R.; Su, Q.; Jiang, Q.; Kip, K.E. Metformin for liver cancer prevention in patients with type 2 diabetes: A systematic review and meta-analysis. J. Clin. Endocrinol. Metab. 2012, 97, 2347–2353. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Ma, S.J.; Zheng, Y.X.; Zhou, P.C.; Xiao, Y.N.; Tan, H.Z. Metformin use improves survival of diabetic liver cancer patients: Systematic review and meta-analysis. Oncotarget 2016, 7, 66202–66211. [Google Scholar] [CrossRef] [Green Version]
  35. Bailey, C.J. Metformin: Historical overview. Diabetologia 2017, 60, 1566–1576. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Watanabe, C. Studies in the metabolism changes induced by administration of guanidine bases: I. Influence of injected guanidine hydrochloride upon blood sugar content. J. Biol. Chem. 1918, 33, 253–265. [Google Scholar] [CrossRef]
  37. Werner, E.A.; Bell, J. CCXIV.—The preparation of methylguanidine, and of ββ-dimethylguanidine by the interaction of dicyanodiamide, and methylammonium and dimethylammonium chlorides respectively. J. Chem. Soc. Trans. 1922, 121, 1790–1794. [Google Scholar] [CrossRef]
  38. Alušík, Š.; Paluch, Z. Metformin: The past, presence, and future. Minerva Med. 2015, 106, 233–238. [Google Scholar] [PubMed]
  39. King, P.; Peacock, I.; Donnelly, R. The UK prospective diabetes study (UKPDS): Clinical and therapeutic implications for type 2 diabetes. Br. J. Clin. Pharmacol. 1999, 48, 643–648. [Google Scholar] [CrossRef] [PubMed]
  40. Rojas, L.B.; Gomes, M.B. Metformin: An old but still the best treatment for type 2 diabetes. Diabetol. Metab. Syndr. 2013, 5, 6. [Google Scholar] [CrossRef] [Green Version]
  41. Agius, L.; Ford, B.E.; Chachra, S.S. The Metformin Mechanism on Gluconeogenesis and AMPK Activation: The Metabolite Perspective. Int. J. Mol. Sci. 2020, 21, 3240. [Google Scholar] [CrossRef]
  42. Jones, V.C.; Dietze, E.C.; Jovanovic-Talisman, T.; McCune, J.S.; Seewaldt, V.L. Metformin and Chemoprevention: Potential for Heart-Healthy Targeting of Biologically Aggressive Breast Cancer. Front. Public Health 2020, 8, 509714. [Google Scholar] [CrossRef] [PubMed]
  43. Nasri, H.; Rafieian-Kopaei, M. Metformin: Current knowledge. J. Res. Med. Sci. Off. J. Isfahan Univ. Med. Sci. 2014, 19, 658. [Google Scholar]
  44. Sliwinska, A.; Drzewoski, J. Molecular action of metformin in hepatocytes: An updated insight. Curr. Diabetes Rev. 2015, 11, 175–181. [Google Scholar] [CrossRef]
  45. Ruderman, N.B.; Carling, D.; Prentki, M.; Cacicedo, J.M. AMPK, insulin resistance, and the metabolic syndrome. J. Clin. Investig. 2013, 123, 2764–2772. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Natali, A.; Ferrannini, E. Effects of metformin and thiazolidinediones on suppression of hepatic glucose production and stimulation of glucose uptake in type 2 diabetes: A systematic review. Diabetologia 2006, 49, 434–441. [Google Scholar] [CrossRef] [Green Version]
  47. Petersen, M.C.; Vatner, D.F.; Shulman, G.I. Regulation of hepatic glucose metabolism in health and disease. Nat. Rev. Endocrinol. 2017, 13, 572–587. [Google Scholar] [CrossRef] [Green Version]
  48. Stumvoll, M.; Nurjhan, N.; Perriello, G.; Dailey, G.; Gerich, J.E. Metabolic effects of metformin in non-insulin-dependent diabetes mellitus. N. Engl. J. Med. 1995, 333, 550–554. [Google Scholar] [CrossRef]
  49. Holman, R.R.; Paul, S.K.; Bethel, M.A.; Matthews, D.R.; Neil, H.A.W. 10-year follow-up of intensive glucose control in type 2 diabetes. N. Engl. J. Med. 2008, 359, 1577–1589. [Google Scholar] [CrossRef] [Green Version]
  50. Hong, J.; Zhang, Y.; Lai, S.; Lv, A.; Su, Q.; Dong, Y.; Zhou, Z.; Tang, W.; Zhao, J.; Cui, L.; et al. Effects of metformin versus glipizide on cardiovascular outcomes in patients with type 2 diabetes and coronary artery disease. Diabetes Care 2013, 36, 1304–1311. [Google Scholar] [CrossRef] [Green Version]
  51. Zhao, M.; Li, X.W.; Chen, Z.; Hao, F.; Tao, S.X.; Yu, H.Y.; Cheng, R.; Liu, H. Neuro-Protective Role of Metformin in Patients with Acute Stroke and Type 2 Diabetes Mellitus via AMPK/Mammalian Target of Rapamycin (mTOR) Signaling Pathway and Oxidative Stress. Med. Sci. Monit. 2019, 25, 2186–2194. [Google Scholar] [CrossRef]
  52. Zhou, J.; Massey, S.; Story, D.; Li, L. Metformin: An Old Drug with New Applications. Int. J. Mol. Sci. 2018, 19, 2863. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Adekola, K.U.; Dalva Aydemir, S.; Ma, S.; Zhou, Z.; Rosen, S.T.; Shanmugam, M. Investigating and targeting chronic lymphocytic leukemia metabolism with the human immunodeficiency virus protease inhibitor ritonavir and metformin. Leuk. Lymphoma 2015, 56, 450–459. [Google Scholar] [CrossRef] [PubMed]
  54. Shin, N.R.; Lee, J.C.; Lee, H.Y.; Kim, M.S.; Whon, T.W.; Lee, M.S.; Bae, J.W. An increase in the Akkermansia spp. population induced by metformin treatment improves glucose homeostasis in diet-induced obese mice. Gut 2014, 63, 727–735. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. McCreight, L.J.; Bailey, C.J.; Pearson, E.R. Metformin and the gastrointestinal tract. Diabetologia 2016, 59, 426–435. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Minamii, T.; Nogami, M.; Ogawa, W. Mechanisms of metformin action: In and out of the gut. J. Diabetes Investig. 2018, 9, 701–703. [Google Scholar] [CrossRef] [PubMed]
  57. Maida, A.; Lamont, B.; Cao, X.; Drucker, D. Metformin regulates the incretin receptor axis via a pathway dependent on peroxisome proliferator-activated receptor-α in mice. Diabetologia 2011, 54, 339–349. [Google Scholar] [CrossRef] [PubMed]
  58. Coll, A.P.; Chen, M.; Taskar, P.; Rimmington, D.; Patel, S.; Tadross, J.A.; Cimino, I.; Yang, M.; Welsh, P.; Virtue, S. GDF15 mediates the effects of metformin on body weight and energy balance. Nature 2020, 578, 444–448. [Google Scholar] [CrossRef]
  59. Zhou, G.; Myers, R.; Li, Y.; Chen, Y.; Shen, X.; Fenyk-Melody, J.; Wu, M.; Ventre, J.; Doebber, T.; Fujii, N. Role of AMP-activated protein kinase in mechanism of metformin action. J. Clin. Investig. 2001, 108, 1167–1174. [Google Scholar] [CrossRef]
  60. Andrzejewski, S.; Siegel, P.M.; St-Pierre, J. Metabolic Profiles Associated With Metformin Efficacy in Cancer. Front. Endocrinol. 2018, 9, 372. [Google Scholar] [CrossRef] [Green Version]
  61. El-Mir, M.-Y.; Nogueira, V.; Fontaine, E.; Avéret, N.; Rigoulet, M.; Leverve, X. Dimethylbiguanide inhibits cell respiration via an indirect effect targeted on the respiratory chain complex I. J. Biol. Chem. 2000, 275, 223–228. [Google Scholar] [CrossRef] [Green Version]
  62. Vander Heiden, M.G. Targeting cancer metabolism: A therapeutic window opens. Nat. Rev. Drug Discov. 2011, 10, 671–684. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Shaw, R.J.; Lamia, K.A.; Vasquez, D.; Koo, S.-H.; Bardeesy, N.; DePinho, R.A.; Montminy, M.; Cantley, L.C. The kinase LKB1 mediates glucose homeostasis in liver and therapeutic effects of metformin. Science 2005, 310, 1642–1646. [Google Scholar] [CrossRef] [PubMed]
  64. Ma, T.; Tian, X.; Zhang, B.; Li, M.; Wang, Y.; Yang, C.; Wu, J.; Wei, X.; Qu, Q.; Yu, Y.; et al. Low-dose metformin targets the lysosomal AMPK pathway through PEN2. Nature 2022, 603, 159–165. [Google Scholar] [CrossRef] [PubMed]
  65. Zhang, C.S.; Li, M.; Ma, T.; Zong, Y.; Cui, J.; Feng, J.W.; Wu, Y.Q.; Lin, S.Y.; Lin, S.C. Metformin Activates AMPK through the Lysosomal Pathway. Cell Metab. 2016, 24, 521–522. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Ben Sahra, I.; Regazzetti, C.; Robert, G.; Laurent, K.; Le Marchand-Brustel, Y.; Auberger, P.; Tanti, J.F.; Giorgetti-Peraldi, S.; Bost, F. Metformin, independent of AMPK, induces mTOR inhibition and cell-cycle arrest through REDD1. Cancer Res. 2011, 71, 4366–4372. [Google Scholar] [CrossRef] [Green Version]
  67. Foretz, M.; Hébrard, S.; Leclerc, J.; Zarrinpashneh, E.; Soty, M.; Mithieux, G.; Sakamoto, K.; Andreelli, F.; Viollet, B. Metformin inhibits hepatic gluconeogenesis in mice independently of the LKB1/AMPK pathway via a decrease in hepatic energy state. J. Clin. Investig. 2010, 120, 2355–2369. [Google Scholar] [CrossRef] [Green Version]
  68. Inoki, K.; Kim, J.; Guan, K.-L. AMPK and mTOR in cellular energy homeostasis and drug targets. Annu. Rev. Pharmacol. Toxicol. 2012, 52, 381–400. [Google Scholar] [CrossRef]
  69. Xu, J.; Ji, J.; Yan, X.-H. Cross-talk between AMPK and mTOR in regulating energy balance. Crit. Rev. Food Sci. Nutr. 2012, 52, 373–381. [Google Scholar] [CrossRef]
  70. Zachariah Tom, R.; Garcia-Roves, P.M.; Sjögren, R.J.; Jiang, L.Q.; Holmström, M.H.; Deshmukh, A.S.; Vieira, E.; Chibalin, A.V.; Björnholm, M.; Zierath, J.R. Effects of AMPK activation on insulin sensitivity and metabolism in leptin-deficient ob/ob mice. Diabetes 2014, 63, 1560–1571. [Google Scholar] [CrossRef] [Green Version]
  71. O’neill, H.M. AMPK and exercise: Glucose uptake and insulin sensitivity. Diabetes Metab. J. 2013, 37, 1–21. [Google Scholar] [CrossRef] [Green Version]
  72. Fullerton, M.D.; Galic, S.; Marcinko, K.; Sikkema, S.; Pulinilkunnil, T.; Chen, Z.-P.; O’neill, H.M.; Ford, R.J.; Palanivel, R.; O’brien, M. Single phosphorylation sites in Acc1 and Acc2 regulate lipid homeostasis and the insulin-sensitizing effects of metformin. Nat. Med. 2013, 19, 1649–1654. [Google Scholar] [CrossRef] [Green Version]
  73. Ford, R.J.; Fullerton, M.D.; Pinkosky, S.L.; Day, E.A.; Scott, J.W.; Oakhill, J.S.; Bujak, A.L.; Smith, B.K.; Crane, J.D.; Blümer, R.M.; et al. Metformin and salicylate synergistically activate liver AMPK, inhibit lipogenesis and improve insulin sensitivity. Biochem. J. 2015, 468, 125–132. [Google Scholar] [CrossRef] [PubMed]
  74. Boudaba, N.; Marion, A.; Huet, C.; Pierre, R.; Viollet, B.; Foretz, M. AMPK Re-Activation Suppresses Hepatic Steatosis but its Downregulation Does Not Promote Fatty Liver Development. EBioMedicine 2018, 28, 194–209. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Viollet, B.; Guigas, B.; Garcia, N.S.; Leclerc, J.; Foretz, M.; Andreelli, F. Cellular and molecular mechanisms of metformin: An overview. Clin. Sci. 2012, 122, 253–270. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Foretz, M.; Guigas, B.; Bertrand, L.; Pollak, M.; Viollet, B. Metformin: From mechanisms of action to therapies. Cell Metab. 2014, 20, 953–966. [Google Scholar] [CrossRef] [Green Version]
  77. Drahota, Z.; Palenickova, E.; Endlicher, R.; Milerova, M.; Brejchova, J.; Vosahlikova, M.; Svoboda, P.; Kazdova, L.; Kalous, M.; Cervinkova, Z.; et al. Biguanides inhibit complex I, II and IV of rat liver mitochondria and modify their functional properties. Physiol. Res. 2014, 63, 1–11. [Google Scholar] [CrossRef]
  78. Miller, R.A.; Chu, Q.; Xie, J.; Foretz, M.; Viollet, B.; Birnbaum, M.J. Biguanides suppress hepatic glucagon signalling by decreasing production of cyclic AMP. Nature 2013, 494, 256–260. [Google Scholar] [CrossRef] [Green Version]
  79. Ouyang, J.; Parakhia, R.A.; Ochs, R.S. Metformin activates AMP kinase through inhibition of AMP deaminase. J. Biol. Chem. 2011, 286, 1–11. [Google Scholar] [CrossRef] [Green Version]
  80. Madiraju, A.K.; Erion, D.M.; Rahimi, Y.; Zhang, X.M.; Braddock, D.T.; Albright, R.A.; Prigaro, B.J.; Wood, J.L.; Bhanot, S.; MacDonald, M.J.; et al. Metformin suppresses gluconeogenesis by inhibiting mitochondrial glycerophosphate dehydrogenase. Nature 2014, 510, 542–546. [Google Scholar] [CrossRef] [Green Version]
  81. Marini, C.; Salani, B.; Massollo, M.; Amaro, A.; Esposito, A.I.; Orengo, A.M.; Capitanio, S.; Emionite, L.; Riondato, M.; Bottoni, G.; et al. Direct inhibition of hexokinase activity by metformin at least partially impairs glucose metabolism and tumor growth in experimental breast cancer. Cell Cycle 2013, 12, 3490–3499. [Google Scholar] [CrossRef] [Green Version]
  82. Salani, B.; Marini, C.; Rio, A.D.; Ravera, S.; Massollo, M.; Orengo, A.M.; Amaro, A.; Passalacqua, M.; Maffioli, S.; Pfeffer, U. Metformin impairs glucose consumption and survival in Calu-1 cells by direct inhibition of hexokinase-II. Sci. Rep. 2013, 3, 2070. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Picone, P.; Vilasi, S.; Librizzi, F.; Contardi, M.; Nuzzo, D.; Caruana, L.; Baldassano, S.; Amato, A.; Mulè, F.; San Biagio, P.L.; et al. Biological and biophysics aspects of metformin-induced effects: Cortex mitochondrial dysfunction and promotion of toxic amyloid pre-fibrillar aggregates. Aging 2016, 8, 1718–1734. [Google Scholar] [CrossRef]
  84. Shoshan-Barmatz, V.; Anand, U.; Nahon-Crystal, E.; Di Carlo, M.; Shteinfer-Kuzmine, A. Adverse Effects of Metformin From Diabetes to COVID-19, Cancer, Neurodegenerative Diseases, and Aging: Is VDAC1 a Common Target? Front. Physiol. 2021, 12, 730048. [Google Scholar] [CrossRef] [PubMed]
  85. LaMoia, T.E.; Butrico, G.M.; Kalpage, H.A.; Goedeke, L.; Hubbard, B.T.; Vatner, D.F.; Gaspar, R.C.; Zhang, X.-M.; Cline, G.W.; Nakahara, K. Metformin, phenformin, and galegine inhibit complex IV activity and reduce glycerol-derived gluconeogenesis. Proc. Natl. Acad. Sci. USA 2022, 119, e2122287119. [Google Scholar] [CrossRef] [PubMed]
  86. Chen, Y.Q.; Chen, G. Combined therapeutic effect and molecular mechanisms of metformin and cisplatin in human lung cancer xenografts in nude mice. J. Cancer Res. Ther. 2015, 11, 324–330. [Google Scholar] [CrossRef] [PubMed]
  87. Lee, B.B.; Kim, Y.; Kim, D.; Cho, E.Y.; Han, J.; Kim, H.K.; Shim, Y.M.; Kim, D.H. Metformin and tenovin-6 synergistically induces apoptosis through LKB1-independent SIRT1 down-regulation in non-small cell lung cancer cells. J. Cell. Mol. Med. 2019, 23, 2872–2889. [Google Scholar] [CrossRef] [Green Version]
  88. Zhang, Y.; Feng, X.; Li, T.; Yi, E.; Li, Y. Metformin synergistic pemetrexed suppresses non-small-cell lung cancer cell proliferation and invasion in vitro. Cancer Med. 2017, 6, 1965–1975. [Google Scholar] [CrossRef]
  89. Saraei, P.; Asadi, I.; Kakar, M.A.; Moradi-Kor, N. The beneficial effects of metformin on cancer prevention and therapy: A comprehensive review of recent advances. Cancer Manag. Res. 2019, 11, 3295. [Google Scholar] [CrossRef] [Green Version]
  90. Mallik, R.; Chowdhury, T.A. Metformin in cancer. Diabetes Res. Clin. Pr. 2018, 143, 409–419. [Google Scholar] [CrossRef] [PubMed]
  91. Heckman-Stoddard, B.M.; Gandini, S.; Puntoni, M.; Dunn, B.K.; DeCensi, A.; Szabo, E. Repurposing old drugs to chemoprevention: The case of metformin. Semin. Oncol. 2016, 43, 123–133. [Google Scholar] [CrossRef] [Green Version]
  92. Mishra, A.K.; Dingli, D. Metformin inhibits IL-6 signaling by decreasing IL-6R expression on multiple myeloma cells. Leukemia 2019, 33, 2695–2709. [Google Scholar] [CrossRef] [PubMed]
  93. Sekino, N.; Kano, M.; Matsumoto, Y.; Sakata, H.; Akutsu, Y.; Hanari, N.; Murakami, K.; Toyozumi, T.; Takahashi, M.; Otsuka, R.; et al. Antitumor effects of metformin are a result of inhibiting nuclear factor kappa B nuclear translocation in esophageal squamous cell carcinoma. Cancer Sci. 2018, 109, 1066–1074. [Google Scholar] [CrossRef]
  94. Menon, S.; Dibble, C.C.; Talbott, G.; Hoxhaj, G.; Valvezan, A.J.; Takahashi, H.; Cantley, L.C.; Manning, B.D. Spatial control of the TSC complex integrates insulin and nutrient regulation of mTORC1 at the lysosome. Cell 2014, 156, 771–785. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Bar-Peled, L.; Schweitzer, L.D.; Zoncu, R.; Sabatini, D.M. Ragulator is a GEF for the rag GTPases that signal amino acid levels to mTORC1. Cell 2012, 150, 1196–1208. [Google Scholar] [CrossRef] [Green Version]
  96. Bar-Peled, L.; Chantranupong, L.; Cherniack, A.D.; Chen, W.W.; Ottina, K.A.; Grabiner, B.C.; Spear, E.D.; Carter, S.L.; Meyerson, M.; Sabatini, D.M. A Tumor suppressor complex with GAP activity for the Rag GTPases that signal amino acid sufficiency to mTORC1. Science 2013, 340, 1100–1106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Bridgeman, S.C.; Ellison, G.C.; Melton, P.E.; Newsholme, P.; Mamotte, C.D.S. Epigenetic effects of metformin: From molecular mechanisms to clinical implications. Diabetes Obes. Metab. 2018, 20, 1553–1562. [Google Scholar] [CrossRef] [Green Version]
  98. Salminen, A.; Kauppinen, A.; Kaarniranta, K. AMPK/Snf1 signaling regulates histone acetylation: Impact on gene expression and epigenetic functions. Cell. Signal. 2016, 28, 887–895. [Google Scholar] [CrossRef]
  99. Vancura, A.; Vancurova, I. Metformin induces protein acetylation in cancer cells. Oncotarget 2017, 8, 39939–39940. [Google Scholar] [CrossRef]
  100. Hu, T.; Chung, Y.M.; Guan, M.; Ma, M.; Ma, J.; Berek, J.S.; Hu, M.C. Reprogramming ovarian and breast cancer cells into non-cancerous cells by low-dose metformin or SN-38 through FOXO3 activation. Sci. Rep. 2014, 4, 5810. [Google Scholar] [CrossRef] [Green Version]
  101. Pernicova, I.; Korbonits, M. Metformin—Mode of action and clinical implications for diabetes and cancer. Nat. Rev. Endocrinol. 2014, 10, 143–156. [Google Scholar] [CrossRef]
  102. Yu, X.; Mao, W.; Zhai, Y.; Tong, C.; Liu, M.; Ma, L.; Yu, X.; Li, S. Anti-tumor activity of metformin: From metabolic and epigenetic perspectives. Oncotarget 2017, 8, 5619. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Safe, S.; Nair, V.; Karki, K. Metformin-induced anticancer activities: Recent insights. Biol. Chem. 2018, 399, 321–335. [Google Scholar] [CrossRef] [PubMed]
  104. Nair, V.; Sreevalsan, S.; Basha, R.; Abdelrahim, M.; Abudayyeh, A.; Rodrigues Hoffman, A.; Safe, S. Mechanism of metformin-dependent inhibition of mammalian target of rapamycin (mTOR) and Ras activity in pancreatic cancer: Role of specificity protein (Sp) transcription factors. J. Biol. Chem. 2014, 289, 27692–27701. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Zou, J.; Hong, L.; Luo, C.; Li, Z.; Zhu, Y.; Huang, T.; Zhang, Y.; Yuan, H.; Hu, Y.; Wen, T.; et al. Metformin inhibits estrogen-dependent endometrial cancer cell growth by activating the AMPK-FOXO1 signal pathway. Cancer Sci. 2016, 107, 1806–1817. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. de Queiroz, E.A.; Akamine, E.H.; de Carvalho, M.H.; Sampaio, S.C.; Fortes, Z.B. Metformin reduces the Walker-256 tumor development in obese-MSG rats via AMPK and FOXO3a. Life Sci. 2015, 121, 78–87. [Google Scholar] [CrossRef] [PubMed]
  107. Sun, Y.; Tao, C.; Huang, X.; He, H.; Shi, H.; Zhang, Q.; Wu, H. Metformin induces apoptosis of human hepatocellular carcinoma HepG2 cells by activating an AMPK/p53/miR-23a/FOXA1 pathway. OncoTargets Ther. 2016, 9, 2845–2853. [Google Scholar]
  108. Celestini, V.; Tezil, T.; Russo, L.; Fasano, C.; Sanese, P.; Forte, G.; Peserico, A.; Lepore Signorile, M.; Longo, G.; De Rasmo, D. Uncoupling FoxO3A mitochondrial and nuclear functions in cancer cells undergoing metabolic stress and chemotherapy. Cell Death Dis. 2018, 9, 231. [Google Scholar] [CrossRef] [Green Version]
  109. Grossi, V.; Fasano, C.; Celestini, V.; Lepore Signorile, M.; Sanese, P.; Simone, C. Chasing the Foxo3: Insights into its new mitochondrial lair in colorectal cancer landscape. Cancers 2019, 11, 414. [Google Scholar] [CrossRef] [Green Version]
  110. Piccolo, S.; Dupont, S.; Cordenonsi, M. The biology of YAP/TAZ: Hippo signaling and beyond. Physiol. Rev. 2014, 94, 1287–1312. [Google Scholar] [CrossRef]
  111. Yu, F.X.; Zhao, B.; Guan, K.L. Hippo Pathway in Organ Size Control, Tissue Homeostasis, and Cancer. Cell 2015, 163, 811–828. [Google Scholar] [CrossRef] [Green Version]
  112. Hajimoradi Javarsiani, M.; Sajedianfard, J.; Haghjooy Javanmard, S. The effects of metformin on the hippo pathway in the proliferation of melanoma cancer cells: A preclinical study. Arch. Physiol. Biochem. 2022, 128, 1150–1155. [Google Scholar] [CrossRef] [PubMed]
  113. Liu, J.; Li, J.; Chen, H.; Wang, R.; Li, P.; Miao, Y.; Liu, P. Metformin suppresses proliferation and invasion of drug-resistant breast cancer cells by activation of the Hippo pathway. J. Cell. Mol. Med. 2020, 24, 5786–5796. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Jin, D.; Guo, J.; Wu, Y.; Chen, W.; Du, J.; Yang, L.; Wang, X.; Gong, K.; Dai, J.; Miao, S.; et al. Metformin-repressed miR-381-YAP-snail axis activity disrupts NSCLC growth and metastasis. J. Exp. Clin. Cancer Res. 2020, 39, 6. [Google Scholar] [CrossRef] [PubMed]
  115. Wu, Y.; Zheng, Q.; Li, Y.; Wang, G.; Gao, S.; Zhang, X.; Yan, X.; Zhang, X.; Xie, J.; Wang, Y.; et al. Metformin targets a YAP1-TEAD4 complex via AMPKα to regulate CCNE1/2 in bladder cancer cells. J. Exp. Clin. Cancer Res. 2019, 38, 376. [Google Scholar] [CrossRef] [Green Version]
  116. Yuan, X.; Wei, W.; Bao, Q.; Chen, H.; Jin, P.; Jiang, W. Metformin inhibits glioma cells stemness and epithelial-mesenchymal transition via regulating YAP activity. Biomed. Pharm. 2018, 102, 263–270. [Google Scholar] [CrossRef] [PubMed]
  117. Tian, Y.; Tang, B.; Wang, C.; Sun, D.; Zhang, R.; Luo, N.; Han, Z.; Liang, R.; Gao, Z.; Wang, L. Metformin mediates resensitivity to 5-fluorouracil in hepatocellular carcinoma via the suppression of YAP. Oncotarget 2016, 7, 46230–46241. [Google Scholar] [CrossRef] [Green Version]
  118. Zhang, H.-H.; Guo, X.-L. Combinational strategies of metformin and chemotherapy in cancers. Cancer Chemother. Pharmacol. 2016, 78, 13–26. [Google Scholar] [CrossRef]
  119. Roncolato, F.; Lindemann, K.; Willson, M.L.; Martyn, J.; Mileshkin, L. PI3K/AKT/mTOR inhibitors for advanced or recurrent endometrial cancer. Cochrane Database Syst. Rev. 2019, 10, Cd012160. [Google Scholar] [CrossRef]
  120. Liu, Y.L.; Yu, J.M.; Song, X.R.; Wang, X.W.; Xing, L.G.; Gao, B.B. Regulation of the chemokine receptor CXCR4 and metastasis by hypoxia-inducible factor in non small cell lung cancer cell lines. Cancer Biol. Ther. 2006, 5, 1320–1326. [Google Scholar] [CrossRef] [Green Version]
  121. Helbig, G.; Christopherson, K.W., 2nd; Bhat-Nakshatri, P.; Kumar, S.; Kishimoto, H.; Miller, K.D.; Broxmeyer, H.E.; Nakshatri, H. NF-kappaB promotes breast cancer cell migration and metastasis by inducing the expression of the chemokine receptor CXCR4. J. Biol. Chem. 2003, 278, 21631–21638. [Google Scholar] [CrossRef] [Green Version]
  122. Dallaglio, K.; Bruno, A.; Cantelmo, A.R.; Esposito, A.I.; Ruggiero, L.; Orecchioni, S.; Calleri, A.; Bertolini, F.; Pfeffer, U.; Noonan, D.M. Paradoxic effects of metformin on endothelial cells and angiogenesis. Carcinogenesis 2014, 35, 1055–1066. [Google Scholar] [CrossRef] [PubMed]
  123. Xia, C.; Liu, C.; He, Z.; Cai, Y.; Chen, J. Metformin inhibits cervical cancer cell proliferation by modulating PI3K/Akt-induced major histocompatibility complex class I-related chain A gene expression. J. Exp. Clin. Cancer Res. 2020, 39, 127. [Google Scholar] [CrossRef] [PubMed]
  124. Nelson, L.E.; Valentine, R.J.; Cacicedo, J.M.; Gauthier, M.-S.; Ido, Y.; Ruderman, N.B. A novel inverse relationship between metformin-triggered AMPK-SIRT1 signaling and p53 protein abundance in high glucose-exposed HepG2 cells. Am. J. Physiol.-Cell Physiol. 2012, 303, C4–C13. [Google Scholar] [CrossRef] [PubMed]
  125. Kauppinen, A.; Suuronen, T.; Ojala, J.; Kaarniranta, K.; Salminen, A. Antagonistic crosstalk between NF-κB and SIRT1 in the regulation of inflammation and metabolic disorders. Cell. Signal. 2013, 25, 1939–1948. [Google Scholar] [CrossRef] [PubMed]
  126. Jiang, L.-L.; Liu, L. Effect of metformin on stem cells: Molecular mechanism and clinical prospect. World J. Stem Cells 2020, 12, 1455. [Google Scholar] [CrossRef] [PubMed]
  127. Pierotti, M.; Berrino, F.; Gariboldi, M.; Melani, C.; Mogavero, A.; Negri, T.; Pasanisi, P.; Pilotti, S. Targeting metabolism for cancer treatment and prevention: Metformin, an old drug with multi-faceted effects. Oncogene 2013, 32, 1475–1487. [Google Scholar] [CrossRef] [PubMed]
  128. Pollak, M. Insulin and insulin-like growth factor signalling in neoplasia. Nat. Rev. Cancer 2008, 8, 915–928. [Google Scholar] [CrossRef]
  129. Li, L.; Wang, L.; Li, J.; Fan, Z.; Yang, L.; Zhang, Z.; Zhang, C.; Yue, D.; Qin, G.; Zhang, T. Metformin-induced reduction of CD39 and CD73 blocks myeloid-derived suppressor cell activity in patients with ovarian cancer. Cancer Res. 2018, 78, 1779–1791. [Google Scholar] [CrossRef] [Green Version]
  130. Wheaton, W.W.; Weinberg, S.E.; Hamanaka, R.B.; Soberanes, S.; Sullivan, L.B.; Anso, E.; Glasauer, A.; Dufour, E.; Mutlu, G.M.; Budigner, G.S. Metformin inhibits mitochondrial complex I of cancer cells to reduce tumorigenesis. elife 2014, 3, e02242. [Google Scholar] [CrossRef]
  131. Zhou, X.; Chen, J.; Yi, G.; Deng, M.; Liu, H.; Liang, M.; Shi, B.; Fu, X.; Chen, Y.; Chen, L.; et al. Metformin suppresses hypoxia-induced stabilization of HIF-1α through reprogramming of oxygen metabolism in hepatocellular carcinoma. Oncotarget 2016, 7, 873–884. [Google Scholar] [CrossRef] [Green Version]
  132. Ye, J.; Chen, K.; Qi, L.; Li, R.; Tang, H.; Zhou, C.; Zhai, W. Metformin suppresses hypoxia-induced migration via the HIF-1α/VEGF pathway in gallbladder cancer in vitro and in vivo. Oncol. Rep. 2018, 40, 3501–3510. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Baumann, T.; Dunkel, A.; Schmid, C.; Schmitt, S.; Hiltensperger, M.; Lohr, K.; Laketa, V.; Donakonda, S.; Ahting, U.; Lorenz-Depiereux, B. Regulatory myeloid cells paralyze T cells through cell–cell transfer of the metabolite methylglyoxal. Nat. Immunol. 2020, 21, 555–566. [Google Scholar] [CrossRef] [PubMed]
  134. Uehara, T.; Eikawa, S.; Nishida, M.; Kunisada, Y.; Yoshida, A.; Fujiwara, T.; Kunisada, T.; Ozaki, T.; Udono, H. Metformin induces CD11b+-cell-mediated growth inhibition of an osteosarcoma: Implications for metabolic reprogramming of myeloid cells and anti-tumor effects. Int. Immunol. 2019, 31, 187–198. [Google Scholar] [CrossRef] [PubMed]
  135. Xu, P.; Yin, K.; Tang, X.; Tian, J.; Zhang, Y.; Ma, J.; Xu, H.; Xu, Q.; Wang, S. Metformin inhibits the function of granulocytic myeloid-derived suppressor cells in tumor-bearing mice. Biomed. Pharmacother. 2019, 120, 109458. [Google Scholar] [CrossRef]
  136. Subarsky, P.; Hill, R.P. The hypoxic tumour microenvironment and metastatic progression. Clin. Exp. Metastasis 2003, 20, 237–250. [Google Scholar] [CrossRef]
  137. Mounier, C.; Bouraoui, L.; Rassart, E. Lipogenesis in cancer progression. Int. J. Oncol. 2014, 45, 485–492. [Google Scholar] [CrossRef] [Green Version]
  138. Ha, J.; Daniel, S.; Broyles, S.S.; Kim, K.-H. Critical phosphorylation sites for acetyl-CoA carboxylase activity. J. Biol. Chem. 1994, 269, 22162–22168. [Google Scholar] [CrossRef]
  139. Li, Y.; Xu, S.; Mihaylova, M.M.; Zheng, B.; Hou, X.; Jiang, B.; Park, O.; Luo, Z.; Lefai, E.; Shyy, J.Y.; et al. AMPK phosphorylates and inhibits SREBP activity to attenuate hepatic steatosis and atherosclerosis in diet-induced insulin-resistant mice. Cell Metab. 2011, 13, 376–388. [Google Scholar] [CrossRef] [Green Version]
  140. Tomic, T.; Botton, T.; Cerezo, M.; Robert, G.; Luciano, F.; Puissant, A.; Gounon, P.; Allegra, M.; Bertolotto, C.; Bereder, J.M.; et al. Metformin inhibits melanoma development through autophagy and apoptosis mechanisms. Cell Death Dis. 2011, 2, e199. [Google Scholar] [CrossRef] [Green Version]
  141. Jiang, S.; Lu, Q. A new contribution for an old drug: Prospect of metformin in colorectal oncotherapy. J. Cancer Res. Ther. 2021, 17, 1608–1617. [Google Scholar] [CrossRef]
  142. Kalender, A.; Selvaraj, A.; Kim, S.Y.; Gulati, P.; Brûlé, S.; Viollet, B.; Kemp, B.E.; Bardeesy, N.; Dennis, P.; Schlager, J.J. Metformin, independent of AMPK, inhibits mTORC1 in a rag GTPase-dependent manner. Cell Metab. 2010, 11, 390–401. [Google Scholar] [CrossRef] [Green Version]
  143. Feng, Y.; Ke, C.; Tang, Q.; Dong, H.; Zheng, X.; Lin, W.; Ke, J.; Huang, J.; Yeung, S.C.; Zhang, H. Metformin promotes autophagy and apoptosis in esophageal squamous cell carcinoma by downregulating Stat3 signaling. Cell Death Dis. 2014, 5, e1088. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Saito, S.; Furuno, A.; Sakurai, J.; Sakamoto, A.; Park, H.R.; Shin-Ya, K.; Tsuruo, T.; Tomida, A. Chemical genomics identifies the unfolded protein response as a target for selective cancer cell killing during glucose deprivation. Cancer Res. 2009, 69, 4225–4234. [Google Scholar] [CrossRef] [PubMed]
  145. Melnik, S.; Dvornikov, D.; Müller-Decker, K.; Depner, S.; Stannek, P.; Meister, M.; Warth, A.; Thomas, M.; Muley, T.; Risch, A.; et al. Cancer cell specific inhibition of Wnt/β-catenin signaling by forced intracellular acidification. Cell Discov. 2018, 4, 37. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Jang, J.H.; Song, I.H.; Sung, E.G.; Lee, T.J.; Kim, J.Y. Metformin-induced apoptosis facilitates degradation of the cellular caspase 8 (FLICE)-like inhibitory protein through a caspase-dependent pathway in human renal cell carcinoma A498 cells. Oncol. Lett. 2018, 16, 2030–2038. [Google Scholar] [CrossRef]
  147. Zhao, Y.; Luo, Q.; Mo, J.; Li, J.; Ye, D.; Ao, Z.; Chen, L.; Liu, J. Metformin in combination with JS-K inhibits growth of renal cell carcinoma cells via reactive oxygen species activation and inducing DNA breaks. J. Cancer 2020, 11, 3701–3712. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Vial, G.; Detaille, D.; Guigas, B. Role of mitochondria in the mechanism (s) of action of metformin. Front. Endocrinol. 2019, 10, 294. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Saladini, S.; Aventaggiato, M.; Barreca, F.; Morgante, E.; Sansone, L.; Russo, M.A.; Tafani, M. Metformin Impairs Glutamine Metabolism and Autophagy in Tumour Cells. Cells 2019, 8, 49. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  150. Shen, P.; Reineke, L.C.; Knutsen, E.; Chen, M.; Pichler, M.; Ling, H.; Calin, G.A. Metformin blocks MYC protein synthesis in colorectal cancer via mTOR-4EBP-eIF4E and MNK1-eIF4G-eIF4E signaling. Mol. Oncol. 2018, 12, 1856–1870. [Google Scholar] [CrossRef] [Green Version]
  151. Lei, Y.; Yi, Y.; Liu, Y.; Liu, X.; Keller, E.T.; Qian, C.N.; Zhang, J.; Lu, Y. Metformin targets multiple signaling pathways in cancer. Chin. J. Cancer 2017, 36, 17. [Google Scholar] [CrossRef] [Green Version]
  152. Blandino, G.; Valerio, M.; Cioce, M.; Mori, F.; Casadei, L.; Pulito, C.; Sacconi, A.; Biagioni, F.; Cortese, G.; Galanti, S.; et al. Metformin elicits anticancer effects through the sequential modulation of DICER and c-MYC. Nat. Commun. 2012, 3, 865. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Pulito, C.; Mori, F.; Sacconi, A.; Goeman, F.; Ferraiuolo, M.; Pasanisi, P.; Campagnoli, C.; Berrino, F.; Fanciulli, M.; Ford, R.J.; et al. Metformin-induced ablation of microRNA 21-5p releases Sestrin-1 and CAB39L antitumoral activities. Cell Discov. 2017, 3, 17022. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Cabello, P.; Pineda, B.; Tormo, E.; Lluch, A.; Eroles, P. The Antitumor Effect of Metformin Is Mediated by miR-26a in Breast Cancer. Int. J. Mol. Sci. 2016, 17, 1298. [Google Scholar] [CrossRef] [PubMed]
  155. Kato, K.; Iwama, H.; Yamashita, T.; Kobayashi, K.; Fujihara, S.; Fujimori, T.; Kamada, H.; Kobara, H.; Masaki, T. The anti-diabetic drug metformin inhibits pancreatic cancer cell proliferation in vitro and in vivo: Study of the microRNAs associated with the antitumor effect of metformin. Oncol. Rep. 2016, 35, 1582–1592. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Li, Y.; Li, L.; Zhang, G.; Wang, Y.; Chen, H.; Kong, R.; Pan, S.; Sun, B. Crucial microRNAs and genes in metformin’s anti-pancreatic cancer effect explored by microRNA-mRNA integrated analysis. Investig. New Drugs 2018, 36, 20–27. [Google Scholar] [CrossRef]
  157. Zhang, J.; Shen, C.; Wang, L.; Ma, Q.; Xia, P.; Qi, M.; Yang, M.; Han, B. Metformin inhibits epithelial-mesenchymal transition in prostate cancer cells: Involvement of the tumor suppressor miR30a and its target gene SOX4. Biochem. Biophys. Res. Commun. 2014, 452, 746–752. [Google Scholar] [CrossRef]
  158. Kalogirou, C.; Schäfer, D.; Krebs, M.; Kurz, F.; Schneider, A.; Riedmiller, H.; Kneitz, B.; Vergho, D. Metformin-Derived Growth Inhibition in Renal Cell Carcinoma Depends on miR-21-Mediated PTEN Expression. Urol. Int. 2016, 96, 106–115. [Google Scholar] [CrossRef]
  159. Fujimori, T.; Kato, K.; Fujihara, S.; Iwama, H.; Yamashita, T.; Kobayashi, K.; Kamada, H.; Morishita, A.; Kobara, H.; Mori, H.; et al. Antitumor effect of metformin on cholangiocarcinoma: In vitro and in vivo studies. Oncol. Rep. 2015, 34, 2987–2996. [Google Scholar] [CrossRef] [Green Version]
  160. Jiang, X.; Ma, N.; Wang, D.; Li, F.; He, R.; Li, D.; Zhao, R.; Zhou, Q.; Wang, Y.; Zhang, F.; et al. Metformin inhibits tumor growth by regulating multiple miRNAs in human cholangiocarcinoma. Oncotarget 2015, 6, 3178–3194. [Google Scholar] [CrossRef] [Green Version]
  161. Chiyo, T.; Kato, K.; Iwama, H.; Fujihara, S.; Fujita, K.; Tadokoro, T.; Ohura, K.; Samukawa, E.; Yamana, Y.; Kobayashi, N.; et al. Therapeutic potential of the antidiabetic drug metformin in small bowel adenocarcinoma. Int. J. Oncol. 2017, 50, 2145–2153. [Google Scholar] [CrossRef] [Green Version]
  162. Wang, F.; Xu, J.; Liu, H.; Liu, Z.; Xia, F. Metformin induces apoptosis by microRNA-26a-mediated downregulation of myeloid cell leukaemia-1 in human oral cancer cells. Mol. Med. Rep. 2016, 13, 4671–4676. [Google Scholar] [CrossRef] [Green Version]
  163. Najafi, M.; Cheki, M.; Rezapoor, S.; Geraily, G.; Motevaseli, E.; Carnovale, C.; Clementi, E.; Shirazi, A. Metformin: Prevention of genomic instability and cancer: A review. Mutat. Res. Genet. Toxicol. Environ. Mutagen 2018, 827, 1–8. [Google Scholar] [CrossRef] [PubMed]
  164. Kim, H.G.; Hien, T.T.; Han, E.H.; Hwang, Y.P.; Choi, J.H.; Kang, K.W.; Kwon, K.I.; Kim, B.H.; Kim, S.K.; Song, G.Y.; et al. Metformin inhibits P-glycoprotein expression via the NF-κB pathway and CRE transcriptional activity through AMPK activation. Br. J. Pharm. 2011, 162, 1096–1108. [Google Scholar] [CrossRef]
  165. Vander Heiden, M.G.; Cantley, L.C.; Thompson, C.B. Understanding the Warburg effect: The metabolic requirements of cell proliferation. Science 2009, 324, 1029–1033. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Warburg, O. Uber den stoffwechsel der karzinomezellen. Biochem Z 1924, 152, 309–344. [Google Scholar]
  167. Zu, X.L.; Guppy, M. Cancer metabolism: Facts, fantasy, and fiction. Biochem. Biophys. Res. Commun. 2004, 313, 459–465. [Google Scholar] [CrossRef] [PubMed]
  168. Loubiere, C.; Clavel, S.; Gilleron, J.; Harisseh, R.; Fauconnier, J.; Ben-Sahra, I.; Kaminski, L.; Laurent, K.; Herkenne, S.; Lacas-Gervais, S.; et al. The energy disruptor metformin targets mitochondrial integrity via modification of calcium flux in cancer cells. Sci. Rep. 2017, 7, 5040. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Coussens, L.M.; Werb, Z. Inflammation and cancer. Nature 2002, 420, 860–867. [Google Scholar] [CrossRef]
  170. Di Magno, L.; Di Pastena, F.; Bordone, R.; Coni, S.; Canettieri, G. The Mechanism of Action of Biguanides: New Answers to a Complex Question. Cancers 2022, 14, 3220. [Google Scholar]
  171. Huang, N.L.; Chiang, S.H.; Hsueh, C.H.; Liang, Y.J.; Chen, Y.J.; Lai, L.P. Metformin inhibits TNF-alpha-induced IkappaB kinase phosphorylation, IkappaB-alpha degradation and IL-6 production in endothelial cells through PI3K-dependent AMPK phosphorylation. Int. J. Cardiol. 2009, 134, 169–175. [Google Scholar] [CrossRef]
  172. Kelly, B.; Tannahill, G.M.; Murphy, M.P.; O’Neill, L.A. Metformin Inhibits the Production of Reactive Oxygen Species from NADH:Ubiquinone Oxidoreductase to Limit Induction of Interleukin-1β (IL-1β) and Boosts Interleukin-10 (IL-10) in Lipopolysaccharide (LPS)-activated Macrophages. J. Biol. Chem. 2015, 290, 20348–20359. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Pearce, E.L.; Walsh, M.C.; Cejas, P.J.; Harms, G.M.; Shen, H.; Wang, L.S.; Jones, R.G.; Choi, Y. Enhancing CD8 T-cell memory by modulating fatty acid metabolism. Nature 2009, 460, 103–107. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Eikawa, S.; Nishida, M.; Mizukami, S.; Yamazaki, C.; Nakayama, E.; Udono, H. Immune-mediated antitumor effect by type 2 diabetes drug, metformin. Proc. Natl. Acad. Sci. USA 2015, 112, 1809–1814. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Nishino, M.; Ramaiya, N.H.; Hatabu, H.; Hodi, F.S. Monitoring immune-checkpoint blockade: Response evaluation and biomarker development. Nat. Rev. Clin. Oncol. 2017, 14, 655–668. [Google Scholar] [CrossRef]
  176. Chen, Y.C.; Li, H.; Wang, J. Mechanisms of metformin inhibiting cancer invasion and migration. Am. J. Transl. Res. 2020, 12, 4885–4901. [Google Scholar]
  177. Hsu, S.K.; Cheng, K.C.; Mgbeahuruike, M.O.; Lin, Y.H.; Wu, C.Y.; Wang, H.D.; Yen, C.H.; Chiu, C.C.; Sheu, S.J. New Insight into the Effects of Metformin on Diabetic Retinopathy, Aging and Cancer: Nonapoptotic Cell Death, Immunosuppression, and Effects beyond the AMPK Pathway. Int. J. Mol. Sci. 2021, 22, 9453. [Google Scholar] [CrossRef]
  178. Park, J.H.; Kim, Y.H.; Park, E.H.; Lee, S.J.; Kim, H.; Kim, A.; Lee, S.B.; Shim, S.; Jang, H.; Myung, J.K.; et al. Effects of metformin and phenformin on apoptosis and epithelial-mesenchymal transition in chemoresistant rectal cancer. Cancer Sci. 2019, 110, 2834–2845. [Google Scholar] [CrossRef]
  179. Huang, D.; He, X.; Zou, J.; Guo, P.; Jiang, S.; Lv, N.; Alekseyev, Y.; Luo, L.; Luo, Z. Negative regulation of Bmi-1 by AMPK and implication in cancer progression. Oncotarget 2016, 7, 6188–6200. [Google Scholar] [CrossRef] [Green Version]
  180. Kang, S.; Kim, B.R.; Kang, M.H.; Kim, D.Y.; Lee, D.H.; Oh, S.C.; Min, B.W.; Um, J.W. Anti-metastatic effect of metformin via repression of interleukin 6-induced epithelial-mesenchymal transition in human colon cancer cells. PLoS ONE 2018, 13, e0205449. [Google Scholar] [CrossRef]
  181. Huang, H. Matrix Metalloproteinase-9 (MMP-9) as a Cancer Biomarker and MMP-9 Biosensors: Recent Advances. Sensors 2018, 18, 3249. [Google Scholar] [CrossRef] [Green Version]
  182. Teng, Y.; Ross, J.L.; Cowell, J.K. The involvement of JAK-STAT3 in cell motility, invasion, and metastasis. Jakstat 2014, 3, e28086. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Hirsch, H.A.; Iliopoulos, D.; Struhl, K. Metformin inhibits the inflammatory response associated with cellular transformation and cancer stem cell growth. Proc. Natl. Acad. Sci. USA 2013, 110, 972–977. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Jun, K.H.; Lee, J.E.; Kim, S.H.; Jung, J.H.; Choi, H.J.; Kim, Y.I.; Chin, H.M.; Yang, S.H. Clinicopathological significance of N-cadherin and VEGF in advanced gastric cancer brain metastasis and the effects of metformin in preclinical models. Oncol. Rep. 2015, 34, 2047–2053. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Wang, J.C.; Li, G.Y.; Li, P.P.; Sun, X.; Li, W.M.; Li, Y.; Lu, S.Y.; Liu, P.J. Suppression of hypoxia-induced excessive angiogenesis by metformin via elevating tumor blood perfusion. Oncotarget 2017, 8, 73892–73904. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Wang, J.C.; Li, G.Y.; Wang, B.; Han, S.X.; Sun, X.; Jiang, Y.N.; Shen, Y.W.; Zhou, C.; Feng, J.; Lu, S.Y.; et al. Metformin inhibits metastatic breast cancer progression and improves chemosensitivity by inducing vessel normalization via PDGF-B downregulation. J. Exp. Clin. Cancer Res. 2019, 38, 235. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Quesada, A.E.; Nguyen, N.D.; Rios, A.; Brown, R.E. Morphoproteomics identifies constitutive activation of the mTORC2/Akt and NF-κB pathways and expressions of IGF-1R, Sirt1, COX-2, and FASN in peripheral T-cell lymphomas: Pathogenetic implications and therapeutic options. Int. J. Clin. Exp. Pathol. 2014, 7, 8732–8739. [Google Scholar] [PubMed]
  188. Koo, Y.X.; Tan, D.S.; Tan, I.B.; Tai, D.W.; Ha, T.; Ong, W.S.; Quek, R.; Tao, M.; Lim, S.T. Effect of concomitant statin, metformin, or aspirin on rituximab treatment for diffuse large B-cell lymphoma. Leuk. Lymphoma 2011, 52, 1509–1516. [Google Scholar]
  189. Chang, S.H.; Luo, S.; O’Brian, K.K.; Thomas, T.S.; Colditz, G.A.; Carlsson, N.P.; Carson, K.R. Association between metformin use and progression of monoclonal gammopathy of undetermined significance to multiple myeloma in US veterans with diabetes mellitus: A population-based retrospective cohort study. Lancet Haematol. 2015, 2, e30–e36. [Google Scholar] [CrossRef] [Green Version]
  190. Boursi, B.; Mamtani, R.; Yang, Y.X.; Weiss, B.M. Impact of metformin on the progression of MGUS to multiple myeloma. Leuk. Lymphoma 2017, 58, 1265–1267. [Google Scholar] [CrossRef] [Green Version]
  191. Zi, F.M.; He, J.S.; Li, Y.; Wu, C.; Yang, L.; Yang, Y.; Wang, L.J.; He, D.H.; Zhao, Y.; Wu, W.J.; et al. Metformin displays anti-myeloma activity and synergistic effect with dexamethasone in in vitro and in vivo xenograft models. Cancer Lett. 2015, 356 Pt B, 443–453. [Google Scholar] [CrossRef]
  192. Wu, W.; Merriman, K.; Nabaah, A.; Seval, N.; Seval, D.; Lin, H.; Wang, M.; Qazilbash, M.H.; Baladandayuthapani, V.; Berry, D.; et al. The association of diabetes and anti-diabetic medications with clinical outcomes in multiple myeloma. Br. J. Cancer 2014, 111, 628–636. [Google Scholar] [CrossRef] [PubMed]
  193. 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] [PubMed] [Green Version]
  194. Jagannathan, S.; Abdel-Malek, M.A.; Malek, E.; Vad, N.; Latif, T.; Anderson, K.C.; Driscoll, J.J. Pharmacologic screens reveal metformin that suppresses GRP78-dependent autophagy to enhance the anti-myeloma effect of bortezomib. Leukemia 2015, 29, 2184–2191. [Google Scholar] [CrossRef] [PubMed]
  195. You, R.; Wang, B.; Chen, P.; Zheng, X.; Hou, D.; Wang, X.; Zhang, B.; Chen, L.; Li, D.; Lin, X.; et al. Metformin sensitizes AML cells to chemotherapy through blocking mitochondrial transfer from stromal cells to AML cells. Cancer Lett. 2022, 532, 215582. [Google Scholar] [CrossRef]
  196. Zhou, F.J.; Zeng, C.X.; Kuang, W.; Cheng, C.; Liu, H.C.; Yan, X.Y.; Chen, X.P.; Zhou, G.; Cao, S. Metformin exerts a synergistic effect with venetoclax by downregulating Mcl-1 protein in acute myeloid leukemia. J. Cancer 2021, 12, 6727–6739. [Google Scholar] [CrossRef]
  197. Gámez, B.; Morris, E.V.; Olechnowicz, S.W.Z.; Webb, S.; Edwards, J.R.; Sowman, A.; Turner, C.J.; Edwards, C.M. The antidiabetic drug metformin acts on the bone microenvironment to promote myeloma cell adhesion to preosteoblasts and increase myeloma tumour burden in vivo. Transl. Oncol. 2022, 15, 101301. [Google Scholar] [CrossRef]
  198. Arnason, T.G.; MacDonald-Dickinson, V.; Gaunt, M.C.; Davies, G.F.; Lobanova, L.; Trost, B.; Gillespie, Z.E.; Waldner, M.; Baldwin, P.; Borrowman, D.; et al. Activation of the Anaphase Promoting Complex Reverses Multiple Drug Resistant Cancer in a Canine Model of Multiple Drug Resistant Lymphoma. Cancers 2022, 14, 4215. [Google Scholar] [CrossRef]
  199. Yuan, F.; Cheng, C.; Xiao, F.; Liu, H.; Cao, S.; Zhou, G. Inhibition of mTORC1/P70S6K pathway by Metformin synergistically sensitizes Acute Myeloid Leukemia to Ara-C. Life Sci. 2020, 243, 117276. [Google Scholar] [CrossRef]
  200. 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]
  201. Wang, Y.; Xu, W.; Yan, Z.; Zhao, W.; Mi, J.; Li, J.; Yan, H. Metformin induces autophagy and G0/G1 phase cell cycle arrest in myeloma by targeting the AMPK/mTORC1 and mTORC2 pathways. J. Exp. Clin. Cancer Res. 2018, 37, 63. [Google Scholar] [CrossRef] [PubMed]
  202. Dalva-Aydemir, S.; Bajpai, R.; Martinez, M.; Adekola, K.U.; Kandela, I.; Wei, C.; Singhal, S.; Koblinski, J.E.; Raje, N.S.; Rosen, S.T.; et al. Targeting the metabolic plasticity of multiple myeloma with FDA-approved ritonavir and metformin. Clin. Cancer Res. 2015, 21, 1161–1171. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. 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] [PubMed]
  204. Glamoclija, U.; Mahmutovic, L.; Bilajac, E.; Soljic, V.; Vukojevic, K.; Suljagic, M. Metformin and Thymoquinone Synergistically Inhibit Proliferation of Imatinib-Resistant Human Leukemic Cells. Front. Pharm. 2022, 13, 867133. [Google Scholar] [CrossRef]
  205. Liu, L.; Patnana, P.K.; Xie, X.; Frank, D.; Nimmagadda, S.C.; Rosemann, A.; Liebmann, M.; Klotz, L.; Opalka, B.; Khandanpour, C. High Metabolic Dependence on Oxidative Phosphorylation Drives Sensitivity to Metformin Treatment in MLL/AF9 Acute Myeloid Leukemia. Cancers 2022, 14, 486. [Google Scholar] [CrossRef] [PubMed]
  206. Saber, M.M.; Al-Mahallawi, A.M.; Stork, B. Metformin dampens cisplatin cytotoxicity on leukemia cells after incorporation into cubosomal nanoformulation. Biomed. Pharm. 2021, 143, 112140. [Google Scholar] [CrossRef] [PubMed]
  207. Valiulienė, G.; Vitkevičienė, A.; Skliutė, G.; Borutinskaitė, V.; Navakauskienė, R. Pharmaceutical Drug Metformin and MCL1 Inhibitor S63845 Exhibit Anticancer Activity in Myeloid Leukemia Cells via Redox Remodeling. Molecules 2021, 26, 2303. [Google Scholar] [CrossRef] [PubMed]
  208. Na, Y.J.; Yu, E.S.; Kim, D.S.; Lee, D.H.; Oh, S.C.; Choi, C.W. Metformin enhances the cytotoxic effect of nilotinib and overcomes nilotinib resistance in chronic myeloid leukemia cells. Korean J. Intern. Med. 2021, 36 (Suppl. S1), S196–S206. [Google Scholar] [CrossRef] [PubMed]
  209. Saito, T.; Itoh, M.; Tohda, S. Metformin suppresses the growth of leukemia cells partly through downregulation of AXL receptor tyrosine kinase. Leuk. Res. 2020, 94, 106383. [Google Scholar] [CrossRef]
  210. Zhou, X.; Kuang, Y.; Liang, S.; Wang, L. Metformin inhibits cell proliferation in SKM-1 cells via AMPK-mediated cell cycle arrest. J. Pharm. Sci. 2019, 141, 146–152. [Google Scholar] [CrossRef]
  211. Vitkevičienė, A.; Janulis, V.; Žučenka, A.; Borutinskaitė, V.; Kaupinis, A.; Valius, M.; Griškevičius, L.; Navakauskienė, R. Oxidative phosphorylation inhibition induces anticancerous changes in therapy-resistant-acute myeloid leukemia patient cells. Mol. Carcinog. 2019, 58, 2008–2016. [Google Scholar] [CrossRef]
  212. Chen, H.L.; Ma, P.; Chen, Y.L.; Sun, L.; Xing, Y.; Wang, F.; Wang, F.; Cao, W.J.; Huang, Y.M.; Zhang, R.H. Effect of Metformin on Proliferation Capacity, Apoptosis and Glycolysis in K562 Cells. Zhongguo Shi Yan Xue Ye Xue Za Zhi 2019, 27, 1387–1394. [Google Scholar]
  213. Renner, K.; Seilbeck, A.; Kauer, N.; Ugele, I.; Siska, P.J.; Brummer, C.; Bruss, C.; Decking, S.M.; Fante, M.; Schmidt, A.; et al. Combined Metabolic Targeting With Metformin and the NSAIDs Diflunisal and Diclofenac Induces Apoptosis in Acute Myeloid Leukemia Cells. Front. Pharm. 2018, 9, 1258. [Google Scholar] [CrossRef] [PubMed]
  214. Asik, A.; Kayabasi, C.; Ozmen Yelken, B.; Yılmaz Susluer, S.; Dogan Sigva, Z.O.; Balcı Okcanoglu, T.; Saydam, G.; Biray Avci, C.; Gunduz, C. Antileukemic effect of paclitaxel in combination with metformin in HL-60 cell line. Gene 2018, 647, 213–220. [Google Scholar] [CrossRef] [PubMed]
  215. Ravera, S.; Cossu, V.; Tappino, B.; Nicchia, E.; Dufour, C.; Cavani, S.; Sciutto, A.; Bolognesi, C.; Columbaro, M.; Degan, P.; et al. Concentration-dependent metabolic effects of metformin in healthy and Fanconi anemia lymphoblast cells. J. Cell. Physiol. 2018, 233, 1736–1751. [Google Scholar] [CrossRef] [PubMed]
  216. Liang, X.; Kong, P.; Wang, J.; Xu, Y.; Gao, C.; Guo, G. Effects of metformin on proliferation and apoptosis of human megakaryoblastic Dami and MEG-01 cells. J. Pharm. Sci. 2017, 135, 14–21. [Google Scholar] [CrossRef]
  217. Yi, Y.; Gao, L.; Wu, M.; Ao, J.; Zhang, C.; Wang, X.; Lin, M.; Bergholz, J.; Zhang, Y.; Xiao, Z.J. Metformin Sensitizes Leukemia Cells to Vincristine via Activation of AMP-activated Protein Kinase. J. Cancer 2017, 8, 2636–2642. [Google Scholar] [CrossRef] [Green Version]
  218. Sabnis, H.S.; Bradley, H.L.; Tripathi, S.; Yu, W.M.; Tse, W.; Qu, C.K.; Bunting, K.D. Synergistic cell death in FLT3-ITD positive acute myeloid leukemia by combined treatment with metformin and 6-benzylthioinosine. Leuk. Res. 2016, 50, 132–140. [Google Scholar] [CrossRef] [Green Version]
  219. Velez, J.; Pan, R.; Lee, J.T.; Enciso, L.; Suarez, M.; Duque, J.E.; Jaramillo, D.; Lopez, C.; Morales, L.; Bornmann, W.; et al. Biguanides sensitize leukemia cells to ABT-737-induced apoptosis by inhibiting mitochondrial electron transport. Oncotarget 2016, 7, 51435–51449. [Google Scholar] [CrossRef]
  220. Voltan, R.; Rimondi, E.; Melloni, E.; Gilli, P.; Bertolasi, V.; Casciano, F.; Rigolin, G.M.; Zauli, G.; Secchiero, P. Metformin combined with sodium dichloroacetate promotes B leukemic cell death by suppressing anti-apoptotic protein Mcl-1. Oncotarget 2016, 7, 18965–18977. [Google Scholar] [CrossRef] [Green Version]
  221. Wang, F.; Liu, Z.; Zeng, J.; Zhu, H.; Li, J.; Cheng, X.; Jiang, T.; Zhang, L.; Zhang, C.; Chen, T.; et al. Metformin synergistically sensitizes FLT3-ITD-positive acute myeloid leukemia to sorafenib by promoting mTOR-mediated apoptosis and autophagy. Leuk. Res. 2015, 39, 1421–1427. [Google Scholar] [CrossRef]
  222. Shi, R.; Lin, J.; Gong, Y.; Yan, T.; Shi, F.; Yang, X.; Liu, X.; Naren, D. The antileukemia effect of metformin in the Philadelphia chromosome-positive leukemia cell line and patient primary leukemia cell. Anti-Cancer Drugs 2015, 26, 913–922. [Google Scholar] [CrossRef] [PubMed]
  223. Rodríguez-Lirio, A.; Pérez-Yarza, G.; Fernández-Suárez, M.R.; Alonso-Tejerina, E.; Boyano, M.D.; Asumendi, A. Metformin Induces Cell Cycle Arrest and Apoptosis in Drug-Resistant Leukemia Cells. Leuk. Res. Treat. 2015, 2015, 516460. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Zhang, B.; Liu, L.L.; Mao, X.; Zhang, D.H. Effects of metformin on FOXM1 expression and on the biological behavior of acute leukemia cell lines. Mol. Med. Rep. 2014, 10, 3193–3198. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Leclerc, G.M.; Leclerc, G.J.; Kuznetsov, J.N.; DeSalvo, J.; Barredo, J.C. Metformin induces apoptosis through AMPK-dependent inhibition of UPR signaling in ALL lymphoblasts. PLoS ONE 2013, 8, e74420. [Google Scholar] [CrossRef]
  226. Huai, L.; Wang, C.; Zhang, C.; Li, Q.; Chen, Y.; Jia, Y.; Li, Y.; Xing, H.; Tian, Z.; Rao, Q.; et al. Metformin induces differentiation in acute promyelocytic leukemia by activating the MEK/ERK signaling pathway. Biochem. Biophys. Res. Commun. 2012, 422, 398–404. [Google Scholar] [CrossRef]
  227. Liu, M.K.; Cheng, L.L.; Yi, H.M.; He, Y.; Li, X.; Fu, D.; Dai, Y.T.; Fang, H.; Cheng, S.; Xu, P.P.; et al. Enhanced lipid metabolism confers the immunosuppressive tumor microenvironment in CD5-positive non-MYC/BCL2 double expressor lymphoma. Front. Oncol. 2022, 12, 885011. [Google Scholar] [CrossRef]
  228. Bagaloni, I.; Visani, A.; Biagiotti, S.; Ruzzo, A.; Navari, M.; Etebari, M.; Mundo, L.; Granai, M.; Lazzi, S.; Isidori, A.; et al. Metabolic Switch and Cytotoxic Effect of Metformin on Burkitt Lymphoma. Front. Oncol. 2021, 11, 661102. [Google Scholar] [CrossRef]
  229. Cristofoletti, C.; Bresin, A.; Picozza, M.; Picchio, M.C.; Monzo, F.; Helmer Citterich, M.; Passarelli, F.; Frezzolini, A.; Scala, E.; Monopoli, A.; et al. Blood and skin-derived Sezary cells: Differences in proliferation-index, activation of PI3K/AKT/mTORC1 pathway and its prognostic relevance. Leukemia 2019, 33, 1231–1242. [Google Scholar] [CrossRef]
  230. Chukkapalli, V.; Gordon, L.I.; Venugopal, P.; Borgia, J.A.; Karmali, R. Metabolic changes associated with metformin potentiates Bcl-2 inhibitor, Venetoclax, and CDK9 inhibitor, BAY1143572 and reduces viability of lymphoma cells. Oncotarget 2018, 9, 21166–21181. [Google Scholar] [CrossRef] [Green Version]
  231. Granato, M.; Gilardini Montani, M.S.; Romeo, M.A.; Santarelli, R.; Gonnella, R.; D’Orazi, G.; Faggioni, A.; Cirone, M. Metformin triggers apoptosis in PEL cells and alters bortezomib-induced Unfolded Protein Response increasing its cytotoxicity and inhibiting KSHV lytic cycle activation. Cell. Signal. 2017, 40, 239–247. [Google Scholar] [CrossRef]
  232. Schlesser, C.; Meul, T.; Stathopoulos, G.; Meiners, S. Metformin Induces Resistance of Cancer Cells to the Proteasome Inhibitor Bortezomib. Biomolecules 2022, 12, 756. [Google Scholar] [CrossRef] [PubMed]
  233. Song, Y.; Chen, S.; Xiang, W.; Xiao, M.; Xiao, H. The mechanism of treatment of multiple myeloma with metformin by way of metabolism. Arch. Med. Sci. 2021, 17, 1056–1063. [Google Scholar] [CrossRef]
  234. Kocemba-Pilarczyk, K.A.; Trojan, S.; Ostrowska, B.; Lasota, M.; Dudzik, P.; Kusior, D.; Kot, M. Influence of metformin on HIF-1 pathway in multiple myeloma. Pharm. Rep. 2020, 72, 1407–1417. [Google Scholar] [CrossRef] [PubMed]
  235. Liu, X.; Zhao, Y.; Zhang, E.; Yan, H.; Lv, N.; Cai, Z. The synergistic effect of PFK15 with metformin exerts anti-myeloma activity via PFKFB3. Biochem. Biophys. Res. Commun. 2019, 515, 332–338. [Google Scholar] [CrossRef] [PubMed]
  236. Tong, J.Y.; Wang, C.; Liu, Y.F.; Wang, W.Q.; Hao, Q.Q.; Ma, J.; Sun, L.; Sun, H. Metformin Induces Apoptosis of Human Multiple Myeloma Cell U266 through the Mitochondrial Apoptotic Pathway. Zhongguo Shi Yan Xue Ye Xue Za Zhi 2018, 26, 489–492. [Google Scholar] [PubMed]
  237. Zhao, Y.; Zhang, E.; Lv, N.; Ma, L.; Yao, S.; Yan, M.; Zi, F.; Deng, G.; Liu, X.; He, J.; et al. Metformin and FTY720 Synergistically Induce Apoptosis in Multiple Myeloma Cells. Cell. Physiol. Biochem. 2018, 48, 785–800. [Google Scholar] [CrossRef]
  238. Lu, B.; Gui, S.Y.; Zhou, H.H.; Liu, Z.L. Effect of Metformin on Proliferation of Multiple Myeloma Cells. Zhongguo Shi Yan Xue Ye Xue Za Zhi 2017, 25, 1097–1100. [Google Scholar]
  239. Short, N.J.; Rytting, M.E.; Cortes, J.E. Acute myeloid leukaemia. Lancet 2018, 392, 593–606. [Google Scholar] [CrossRef]
  240. Green, A.S.; Chapuis, N.; Maciel, T.T.; Willems, L.; Lambert, M.; Arnoult, C.; Boyer, O.; Bardet, V.; Park, S.; Foretz, M.; et al. The LKB1/AMPK signaling pathway has tumor suppressor activity in acute myeloid leukemia through the repression of mTOR-dependent oncogenic mRNA translation. Blood 2010, 116, 4262–4273. [Google Scholar] [CrossRef]
  241. Martelli, A.M.; Chiarini, F.; Evangelisti, C.; Ognibene, A.; Bressanin, D.; Billi, A.M.; Manzoli, L.; Cappellini, A.; McCubrey, J.A. Targeting the liver kinase B1/AMP-activated protein kinase pathway as a therapeutic strategy for hematological malignancies. Expert. Opin. Ther. Targets 2012, 16, 729–742. [Google Scholar] [CrossRef] [PubMed]
  242. Rosilio, C.; Ben-Sahra, I.; Bost, F.; Peyron, J.F. Metformin: A metabolic disruptor and anti-diabetic drug to target human leukemia. Cancer Lett. 2014, 346, 188–196. [Google Scholar] [CrossRef] [PubMed]
  243. Huai, L.; Wang, C.C.; Zhang, C.P.; Li, Q.H.; Chen, Y.R.; Jia, Y.J.; Wang, M.; Wang, J.X. Effect of metformin on acute promyelocytic leukemia cell line NB4 and its mechanism. Zhongguo Shi Yan Xue Ye Xue Za Zhi 2012, 20, 1322–1326. [Google Scholar] [PubMed]
  244. Ceacareanu, A.C.; Nimako, G.K.; Wintrob, Z.A. Missing the benefit of metformin in acute myeloid leukemia: A problem of contrast? J. Res. Pharm. Pract. 2017, 6, 145. [Google Scholar] [CrossRef]
  245. Groffen, J.; Heisterkamp, N.; Stam, K. Oncogene activation by chromosomal translocation in chronic myelocytic leukemia. Cold Spring Harb. Symp. Quant. Biol. 1986, 51 Pt 2, 911–921. [Google Scholar] [CrossRef] [PubMed]
  246. Deininger, M.W.; Goldman, J.M.; Melo, J.V. The molecular biology of chronic myeloid leukemia. BLOOD-NEW YORK 2000, 96, 3343–3356. [Google Scholar]
  247. Vakana, E.; Altman, J.K.; Glaser, H.; Donato, N.J.; Platanias, L.C. Antileukemic effects of AMPK activators on BCR-ABL-expressing cells. Blood 2011, 118, 6399–6402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  248. Pokorny, R.; Stenehjem, D.D.; Gilreath, J.A. Impact of metformin on tyrosine kinase inhibitor response in chronic myeloid leukemia. J. Oncol. Pharm. Pr. 2022, 28, 916–923. [Google Scholar] [CrossRef] [PubMed]
  249. Paganin, M.; Ferrando, A. Molecular pathogenesis and targeted therapies for NOTCH1-induced T-cell acute lymphoblastic leukemia. Blood Rev. 2011, 25, 83–90. [Google Scholar] [CrossRef] [Green Version]
  250. Aifantis, I.; Raetz, E.; Buonamici, S. Molecular pathogenesis of T-cell leukaemia and lymphoma. Nat. Rev. Immunol. 2008, 8, 380–390. [Google Scholar] [CrossRef]
  251. Grimaldi, C.; Chiarini, F.; Tabellini, G.; Ricci, F.; Tazzari, P.L.; Battistelli, M.; Falcieri, E.; Bortul, R.; Melchionda, F.; Iacobucci, I.; et al. AMP-dependent kinase/mammalian target of rapamycin complex 1 signaling in T-cell acute lymphoblastic leukemia: Therapeutic implications. Leukemia 2012, 26, 91–100. [Google Scholar] [CrossRef]
  252. Pan, J.; Chen, C.; Jin, Y.; Fuentes-Mattei, E.; Velazquez-Tores, G.; Benito, J.M.; Konopleva, M.; Andreeff, M.; Lee, M.H.; Yeung, S.C. Differential impact of structurally different anti-diabetic drugs on proliferation and chemosensitivity of acute lymphoblastic leukemia cells. Cell Cycle 2012, 11, 2314–2326. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  253. Mavrakis, K.J.; Wolfe, A.L.; Oricchio, E.; Palomero, T.; de Keersmaecker, K.; McJunkin, K.; Zuber, J.; James, T.; Khan, A.A.; Leslie, C.S.; et al. Genome-wide RNA-mediated interference screen identifies miR-19 targets in Notch-induced T-cell acute lymphoblastic leukaemia. Nat. Cell. Biol. 2010, 12, 372–379. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  254. Mayr, M.; Liem, D.; Zhang, J.; Li, X.; Avliyakulov, N.K.; Yang, J.I.; Young, G.; Vondriska, T.M.; Ladroue, C.; Madhu, B.; et al. Proteomic and metabolomic analysis of cardioprotection: Interplay between protein kinase C epsilon and delta in regulating glucose metabolism of murine hearts. J. Mol. Cell. Cardiol. 2009, 46, 268–277. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  255. Niture, S.K.; Gnatt, A.; Jaiswal, A.K. Retraction: Oncogene PKCε controls INrf2-Nrf2 interaction in normal and cancer cells through phosphorylation of INrf2. J. Cell Sci. 2017, 130, 815. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  256. Tang, H.M.; Gao, W.W.; Chan, C.P.; Siu, Y.T.; Wong, C.M.; Kok, K.H.; Ching, Y.P.; Takemori, H.; Jin, D.Y. LKB1 tumor suppressor and salt-inducible kinases negatively regulate human T-cell leukemia virus type 1 transcription. Retrovirology 2013, 10, 40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  257. Huntly, B.J.; Gilliland, D.G. Leukaemia stem cells and the evolution of cancer-stem-cell research. Nat. Rev. Cancer 2005, 5, 311–321. [Google Scholar] [CrossRef] [PubMed]
  258. Iliopoulos, D.; Hirsch, H.A.; Struhl, K. Metformin decreases the dose of chemotherapy for prolonging tumor remission in mouse xenografts involving multiple cancer cell types. Cancer Res. 2011, 71, 3196–3201. [Google Scholar] [CrossRef] [Green Version]
  259. Ramos-Peñafiel, C.; Olarte-Carrillo, I.; Cerón-Maldonado, R.; Rozen-Fuller, E.; Kassack-Ipiña, J.J.; Meléndez-Mier, G.; Collazo-Jaloma, J.; Martínez-Tovar, A. Effect of metformin on the survival of patients with ALL who express high levels of the ABCB1 drug resistance gene. J. Transl. Med. 2018, 16, 245. [Google Scholar] [CrossRef] [PubMed]
  260. Trucco, M.; Barredo, J.C.; Goldberg, J.; Leclerc, G.M.; Hale, G.A.; Gill, J.; Setty, B.; Smith, T.; Lush, R.; Lee, J.K.; et al. A phase I window, dose escalating and safety trial of metformin in combination with induction chemotherapy in relapsed refractory acute lymphoblastic leukemia: Metformin with induction chemotherapy of vincristine, dexamethasone, PEG-asparaginase, and doxorubicin. Pediatr. Blood Cancer 2018, 65, e27224. [Google Scholar]
  261. Vu, K.; Busaidy, N.; Cabanillas, M.E.; Konopleva, M.; Faderl, S.; Thomas, D.A.; O’Brien, S.; Broglio, K.; Ensor, J.; Escalante, C.; et al. A randomized controlled trial of an intensive insulin regimen in patients with hyperglycemic acute lymphoblastic leukemia. Clin. Lymphoma Myeloma Leuk. 2012, 12, 355–362. [Google Scholar] [CrossRef] [Green Version]
  262. Ramos-Peñafiel, C.O.; Martínez-Murillo, C.; Santoyo-Sánchez, A.; Jiménez-Ponce, F.; Rozen-Fuller, E.; Collazo-Jaloma, J.; Olarte-Carrillo, I.; Martínez-Tovar, A. Effect of metformin addition to an acute lymphoblastic leukemia chemotherapy treatment. Rev. Médica Inst. Mex. Seguro Soc. 2014, 52, 270–275. [Google Scholar]
  263. Bostrom, B.; Uppal, P.; Chu, J.; Messinger, Y.; Gandrud, L.; McEvoy, R. Safety and efficacy of metformin for therapy-induced hyperglycemia in children with acute lymphoblastic leukemia. J. Pediatr. Hematol. Oncol. 2013, 35, 504–508. [Google Scholar] [CrossRef] [PubMed]
  264. Bruno, S.; Ledda, B.; Tenca, C.; Ravera, S.; Orengo, A.M.; Mazzarello, A.N.; Pesenti, E.; Casciaro, S.; Racchi, O.; Ghiotto, F.; et al. Metformin inhibits cell cycle progression of B-cell chronic lymphocytic leukemia cells. Oncotarget 2015, 6, 22624–22640. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  265. Herishanu, Y.; Pérez-Galán, P.; Liu, D.; Biancotto, A.; Pittaluga, S.; Vire, B.; Gibellini, F.; Njuguna, N.; Lee, E.; Stennett, L.; et al. The lymph node microenvironment promotes B-cell receptor signaling, NF-kappaB activation, and tumor proliferation in chronic lymphocytic leukemia. Blood 2011, 117, 563–574. [Google Scholar] [CrossRef] [PubMed]
  266. Yaktapour, N.; Übelhart, R.; Schüler, J.; Aumann, K.; Dierks, C.; Burger, M.; Pfeifer, D.; Jumaa, H.; Veelken, H.; Brummer, T.; et al. Insulin-like growth factor-1 receptor (IGF1R) as a novel target in chronic lymphocytic leukemia. Blood 2013, 122, 1621–1633. [Google Scholar] [CrossRef]
  267. Hofbauer, S.W.; Piñón, J.D.; Brachtl, G.; Haginger, L.; Wang, W.; Jöhrer, K.; Tinhofer, I.; Hartmann, T.N.; Greil, R. Modifying akt signaling in B-cell chronic lymphocytic leukemia cells. Cancer Res. 2010, 70, 7336–7344. [Google Scholar] [CrossRef] [Green Version]
  268. Zhuang, J.; Hawkins, S.F.; Glenn, M.A.; Lin, K.; Johnson, G.G.; Carter, A.; Cawley, J.C.; Pettitt, A.R. Akt is activated in chronic lymphocytic leukemia cells and delivers a pro-survival signal: The therapeutic potential of Akt inhibition. Haematologica 2010, 95, 110–118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  269. Liu, Z.; Hazan-Halevy, I.; Harris, D.M.; Li, P.; Ferrajoli, A.; Faderl, S.; Keating, M.J.; Estrov, Z. STAT-3 activates NF-kappaB in chronic lymphocytic leukemia cells. Mol. Cancer Res. 2011, 9, 507–515. [Google Scholar] [CrossRef] [Green Version]
  270. Kesanakurti, D.; Chetty, C.; Rajasekhar Maddirela, D.; Gujrati, M.; Rao, J.S. Essential role of cooperative NF-κB and Stat3 recruitment to ICAM-1 intronic consensus elements in the regulation of radiation-induced invasion and migration in glioma. Oncogene 2013, 32, 5144–5155. [Google Scholar] [CrossRef] [Green Version]
  271. Del Gaizo Moore, V.; Brown, J.R.; Certo, M.; Love, T.M.; Novina, C.D.; Letai, A. Chronic lymphocytic leukemia requires BCL2 to sequester prodeath BIM, explaining sensitivity to BCL2 antagonist ABT-737. J. Clin. Investig. 2007, 117, 112–121. [Google Scholar] [CrossRef] [Green Version]
  272. van Delft, M.F.; Wei, A.H.; Mason, K.D.; Vandenberg, C.J.; Chen, L.; Czabotar, P.E.; Willis, S.N.; Scott, C.L.; Day, C.L.; Cory, S.; et al. The BH3 mimetic ABT-737 targets selective Bcl-2 proteins and efficiently induces apoptosis via Bak/Bax if Mcl-1 is neutralized. Cancer Cell 2006, 10, 389–399. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  273. Rajkumar, S.V. Evolving diagnostic criteria for multiple myeloma. Hematol. Am. Soc. Hematol. Educ. Program 2015, 2015, 272–278. [Google Scholar] [CrossRef] [PubMed]
  274. White-Al Habeeb, N.M.; Garcia, J.; Fleshner, N.; Bapat, B. Metformin Elicits Antitumor Effects and Downregulates the Histone Methyltransferase Multiple Myeloma SET Domain (MMSET) in Prostate Cancer Cells. Prostate 2016, 76, 1507–1518. [Google Scholar] [CrossRef] [PubMed]
  275. Birmann, B.M.; Neuhouser, M.L.; Rosner, B.; Albanes, D.; Buring, J.E.; Giles, G.G.; Lan, Q.; Lee, I.M.; Purdue, M.P.; Rothman, N.; et al. Prediagnosis biomarkers of insulin-like growth factor-1, insulin, and interleukin-6 dysregulation and multiple myeloma risk in the Multiple Myeloma Cohort Consortium. Blood 2012, 120, 4929–4937. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  276. Verdura, S.; Cuyàs, E.; Martin-Castillo, B.; Menendez, J.A. Metformin as an archetype immuno-metabolic adjuvant for cancer immunotherapy. Oncoimmunology 2019, 8, e1633235. [Google Scholar] [CrossRef] [PubMed]
  277. Azab, A.K.; Hu, J.; Quang, P.; Azab, F.; Pitsillides, C.; Awwad, R.; Thompson, B.; Maiso, P.; Sun, J.D.; Hart, C.P.; et al. Hypoxia promotes dissemination of multiple myeloma through acquisition of epithelial to mesenchymal transition-like features. Blood 2012, 119, 5782–5794. [Google Scholar] [CrossRef] [Green Version]
  278. Kocemba, K.A.; van Andel, H.; de Haan-Kramer, A.; Mahtouk, K.; Versteeg, R.; Kersten, M.J.; Spaargaren, M.; Pals, S.T. The hypoxia target adrenomedullin is aberrantly expressed in multiple myeloma and promotes angiogenesis. Leukemia 2013, 27, 1729–1737. [Google Scholar] [CrossRef] [Green Version]
  279. Colla, S.; Storti, P.; Donofrio, G.; Todoerti, K.; Bolzoni, M.; Lazzaretti, M.; Abeltino, M.; Ippolito, L.; Neri, A.; Ribatti, D.; et al. Low bone marrow oxygen tension and hypoxia-inducible factor-1α overexpression characterize patients with multiple myeloma: Role on the transcriptional and proangiogenic profiles of CD138(+) cells. Leukemia 2010, 24, 1967–1970. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  280. Storti, P.; Bolzoni, M.; Donofrio, G.; Airoldi, I.; Guasco, D.; Toscani, D.; Martella, E.; Lazzaretti, M.; Mancini, C.; Agnelli, L.; et al. Hypoxia-inducible factor (HIF)-1α suppression in myeloma cells blocks tumoral growth in vivo inhibiting angiogenesis and bone destruction. Leukemia 2013, 27, 1697–1706. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  281. Tyszka-Czochara, M.; Bukowska-Strakova, K.; Kocemba-Pilarczyk, K.A.; Majka, M. Caffeic Acid Targets AMPK Signaling and Regulates Tricarboxylic Acid Cycle Anaplerosis while Metformin Downregulates HIF-1α-Induced Glycolytic Enzymes in Human Cervical Squamous Cell Carcinoma Lines. Nutrients 2018, 10, 841. [Google Scholar] [CrossRef] [Green Version]
  282. Guimarães, T.A.; Farias, L.C.; Santos, E.S.; de Carvalho Fraga, C.A.; Orsini, L.A.; de Freitas Teles, L.; Feltenberger, J.D.; de Jesus, S.F.; de Souza, M.G.; Santos, S.H.; et al. Metformin increases PDH and suppresses HIF-1α under hypoxic conditions and induces cell death in oral squamous cell carcinoma. Oncotarget 2016, 7, 55057–55068. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  283. Gao, L.; Li, L.; Hu, J.; Li, G.; Zhang, Y.; Dai, X.; De, Z.; Xu, F. Metformin Inhibits Multiple Myeloma Serum-induced Endothelial Cell Thrombosis by Down-Regulating miR-532. Ann. Vasc. Surg. 2022, 85, 347–357.e2. [Google Scholar] [CrossRef] [PubMed]
  284. Ragbourne, S.C.; Maghsoodi, N.; Streetly, M.; Crook, M.A. The Association between Metabolic Syndrome and Multiple Myeloma. Acta Haematol. 2021, 144, 24–33. [Google Scholar] [CrossRef] [PubMed]
  285. Boursi, B.; Weiss, B.M.; Haynes, K.; Mamtani, R.; Yang, Y.X. Reappraisal of risk factors for monoclonal gammopathy of undetermined significance. Am. J. Hematol. 2016, 91, 581–584. [Google Scholar] [CrossRef] [Green Version]
  286. Siegel, R.L.; Miller, K.D.; Fuchs, H.E.; Jemal, A. Cancer statistics, 2022. CA Cancer J. Clin. 2022, 72, 7–33. [Google Scholar] [CrossRef]
  287. Naserian, M.; Ramazani, E.; Iranshahi, M.; Tayarani-Najaran, Z. The Role of SAPK/JNK Pathway in the Synergistic Effects of Metformin and Dacarbazine on Apoptosis in Raji and Ramos Lymphoma Cells. Curr. Mol. Pharm. 2018, 11, 336–342. [Google Scholar] [CrossRef]
  288. Anderson, J.R.; Armitage, J.O.; Weisenburger, D.D. Epidemiology of the non-Hodgkin’s lymphomas: Distributions of the major subtypes differ by geographic locations. Non-Hodgkin’s Lymphoma Classification Project. Ann. Oncol. 1998, 9, 717–720. [Google Scholar] [CrossRef]
  289. Wright, G.W.; Huang, D.W.; Phelan, J.D.; Coulibaly, Z.A.; Roulland, S.; Young, R.M.; Wang, J.Q.; Schmitz, R.; Morin, R.D.; Tang, J.; et al. A Probabilistic Classification Tool for Genetic Subtypes of Diffuse Large B Cell Lymphoma with Therapeutic Implications. Cancer Cell 2020, 37, 551–568.e14. [Google Scholar] [CrossRef]
  290. Schmitz, R.; Wright, G.W.; Huang, D.W.; Johnson, C.A.; Phelan, J.D.; Wang, J.Q.; Roulland, S.; Kasbekar, M.; Young, R.M.; Shaffer, A.L.; et al. Genetics and Pathogenesis of Diffuse Large B-Cell Lymphoma. N. Engl. J. Med. 2018, 378, 1396–1407. [Google Scholar] [CrossRef]
  291. Chapuy, B.; Stewart, C.; Dunford, A.J.; Kim, J.; Kamburov, A.; Redd, R.A.; Lawrence, M.S.; Roemer, M.G.M.; Li, A.J.; Ziepert, M.; et al. Molecular subtypes of diffuse large B cell lymphoma are associated with distinct pathogenic mechanisms and outcomes. Nat. Med. 2018, 24, 679–690. [Google Scholar] [CrossRef]
  292. Jiang, X.N.; Zhang, Y.; Wang, W.G.; Sheng, D.; Zhou, X.Y.; Li, X.Q. Alteration of Cholesterol Metabolism by Metformin Is Associated With Improved Outcome in Type II Diabetic Patients With Diffuse Large B-Cell Lymphoma. Front. Oncol. 2021, 11, 608238. [Google Scholar] [CrossRef]
  293. Casulo, C.; Friedberg, J.W. Burkitt lymphoma- a rare but challenging lymphoma. Best Pr. Res. Clin. Haematol. 2018, 31, 279–284. [Google Scholar] [CrossRef] [PubMed]
  294. Wang, Y.; Maurer, M.J.; Larson, M.C.; Allmer, C.; Feldman, A.L.; Bennani, N.N.; Thompson, C.A.; Porrata, L.F.; Habermann, T.M.; Witzig, T.E.; et al. Impact of metformin use on the outcomes of newly diagnosed diffuse large B-cell lymphoma and follicular lymphoma. Br. J. Haematol. 2019, 186, 820–828. [Google Scholar] [CrossRef] [PubMed]
  295. Ye, X.; Zhang, G.; Righolt, C.; Johnston, J.B.; Banerji, V.; Gibson, S.B.; Mahmud, S.M. Metformin Is Not Associated with Incidence Risk of Non-Hodgkin Lymphomas among Diabetic Patients. Cancer Epidemiol. Biomark. Prev. 2018, 27, 610–612. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  296. Smyth, L.; Blunt, D.N.; Gatov, E.; Nagamuthu, C.; Croxford, R.; Mozessohn, L.; Cheung, M.C. Statin and cyclooxygenase-2 inhibitors improve survival in newly diagnosed diffuse large B-cell lymphoma: A large population-based study of 4913 subjects. Br. J. Haematol. 2020, 191, 396–404. [Google Scholar] [CrossRef] [PubMed]
  297. Tseng, C.H. Metformin is associated with a lower risk of non-Hodgkin lymphoma in patients with type 2 diabetes. Diabetes Metab. 2019, 45, 458–464. [Google Scholar] [CrossRef]
  298. Wynn, A.; Vacheron, A.; Zuber, J.; Solomon, S.S. Metformin Associated With Increased Survival in Type 2 Diabetes Patients With Pancreatic Cancer and Lymphoma. Am. J. Med. Sci. 2019, 358, 200–203. [Google Scholar] [CrossRef]
  299. Singh, A.; Gu, J.; Yanamadala, V.; Czuczman, M.S.; Hernandez-Ilizaliturri, F.J. Metformin lowers the mitochondrial potential of lymphoma cells and its use during front-line rituximab-based chemo-immunotherapy improves the clinical outcome of diffuse Large B-cell lymphoma. Blood 2013, 122, 1825. [Google Scholar] [CrossRef]
  300. Fan, X.; Zhao, B.; Yang, B.; Zhao, Y.; Ye, J.; Lu, Y.; Wang, C.; Xiong, H.; Chen, S.; Janin, A. Metformin prolonged the survival of diffuse large B-cell lymphoma and grade 3b follicular lymphoma patients responding to first-line treatment with rituximab plus cyclophosphamide, doxorubicin, vincristine, and prednisone: A prospective phase II clinical trial. Transl. Cancer Res. 2018, 2018, 4. [Google Scholar]
  301. Bolen, S.; Feldman, L.; Vassy, J.; Wilson, L.; Yeh, H.C.; Marinopoulos, S.; Wiley, C.; Selvin, E.; Wilson, R.; Bass, E.B.; et al. Systematic review: Comparative effectiveness and safety of oral medications for type 2 diabetes mellitus. Ann. Intern. Med. 2007, 147, 386–399. [Google Scholar] [CrossRef]
  302. Dawed, A.Y.; Zhou, K.; van Leeuwen, N.; Mahajan, A.; Robertson, N.; Koivula, R.; Elders, P.J.M.; Rauh, S.P.; Jones, A.G.; Holl, R.W.; et al. Variation in the Plasma Membrane Monoamine Transporter (PMAT) (Encoded by SLC29A4) and Organic Cation Transporter 1 (OCT1) (Encoded by SLC22A1) and Gastrointestinal Intolerance to Metformin in Type 2 Diabetes: An IMI DIRECT Study. Diabetes Care 2019, 42, 1027–1033. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  303. Song, Y.; Wu, Z.; Zhao, P. The effects of metformin in the treatment of osteoarthritis: Current perspectives. Front. Pharm. 2022, 13, 952560. [Google Scholar] [CrossRef] [PubMed]
  304. Marshall, S.M. 60 years of metformin use: A glance at the past and a look to the future. Diabetologia 2017, 60, 1561–1565. [Google Scholar] [CrossRef] [PubMed]
  305. Aroda, V.R.; Edelstein, S.L.; Goldberg, R.B.; Knowler, W.C.; Marcovina, S.M.; Orchard, T.J.; Bray, G.A.; Schade, D.S.; Temprosa, M.G.; White, N.H.; et al. Long-term Metformin Use and Vitamin B12 Deficiency in the Diabetes Prevention Program Outcomes Study. J. Clin. Endocrinol. Metab. 2016, 101, 1754–1761. [Google Scholar] [CrossRef] [PubMed]
  306. Kim, J.; Ahn, C.W.; Fang, S.; Lee, H.S.; Park, J.S. Association between metformin dose and vitamin B12 deficiency in patients with type 2 diabetes. Medicine 2019, 98, e17918. [Google Scholar] [CrossRef]
  307. Ting, R.Z.; Szeto, C.C.; Chan, M.H.; Ma, K.K.; Chow, K.M. Risk factors of vitamin B(12) deficiency in patients receiving metformin. Arch. Intern. Med. 2006, 166, 1975–1979. [Google Scholar] [CrossRef] [Green Version]
  308. Bauman, W.A.; Shaw, S.; Jayatilleke, E.; Spungen, A.M.; Herbert, V. Increased intake of calcium reverses vitamin B12 malabsorption induced by metformin. Diabetes Care 2000, 23, 1227–1231. [Google Scholar] [CrossRef] [Green Version]
  309. Wang, Y.W.; He, S.J.; Feng, X.; Cheng, J.; Luo, Y.T.; Tian, L.; Huang, Q. Metformin: A review of its potential indications. Drug Des. Devel. Ther. 2017, 11, 2421–2429. [Google Scholar] [CrossRef] [Green Version]
  310. Boucaud-Maitre, D.; Ropers, J.; Porokhov, B.; Altman, J.J.; Bouhanick, B.; Doucet, J.; Girardin, E.; Kaloustian, E.; Lassmann Vague, V.; Emmerich, J. Lactic acidosis: Relationship between metformin levels, lactate concentration and mortality. Diabet. Med. 2016, 33, 1536–1543. [Google Scholar] [CrossRef]
  311. Al-Abri, S.A.; Hayashi, S.; Thoren, K.L.; Olson, K.R. Metformin overdose-induced hypoglycemia in the absence of other antidiabetic drugs. Clin. Toxicol. 2013, 51, 444–447. [Google Scholar] [CrossRef]
  312. Salpeter, S.R.; Greyber, E.; Pasternak, G.A.; Salpeter, E.E. Risk of fatal and nonfatal lactic acidosis with metformin use in type 2 diabetes mellitus. Cochrane Database Syst. Rev. 2010, 2010, Cd002967. [Google Scholar]
  313. Donnan, K.; Segar, L. SGLT2 inhibitors and metformin: Dual antihyperglycemic therapy and the risk of metabolic acidosis in type 2 diabetes. Eur. J. Pharm. 2019, 846, 23–29. [Google Scholar] [CrossRef] [PubMed]
  314. Ahmad, E.; Sargeant, J.A.; Zaccardi, F.; Khunti, K.; Webb, D.R.; Davies, M.J. Where Does Metformin Stand in Modern Day Management of Type 2 Diabetes? Pharmaceuticals 2020, 13, 427. [Google Scholar] [CrossRef] [PubMed]
  315. Adler, A.I.; Shaw, E.J.; Stokes, T.; Ruiz, F. Newer agents for blood glucose control in type 2 diabetes: Summary of NICE guidance. BMJ 2009, 338, b1668. [Google Scholar] [CrossRef]
  316. Tanner, C.; Wang, G.; Liu, N.; Andrikopoulos, S.; Zajac, J.D.; Ekinci, E.I. Metformin: Time to review its role and safety in chronic kidney disease. Med. J. Aust. 2019, 211, 37–42. [Google Scholar] [CrossRef] [PubMed]
  317. Wensink, M.J.; Lu, Y.; Tian, L.; Shaw, G.M.; Rizzi, S.; Jensen, T.K.; Mathiesen, E.R.; Skakkebæk, N.E.; Lindahl-Jacobsen, R.; Eisenberg, M.L. Preconception Antidiabetic Drugs in Men and Birth Defects in Offspring: A Nationwide Cohort Study. Ann. Intern. Med. 2022, 175, 665–673. [Google Scholar] [CrossRef] [PubMed]
  318. Alhowail, A. Potential mechanisms of metformin-induced memory impairment. Eur. Rev. Med. Pharm. Sci. 2021, 25, 4757–4761. [Google Scholar]
  319. Wang, G.S.; Hoyte, C. Review of Biguanide (Metformin) Toxicity. J. Intensive Care Med. 2019, 34, 863–876. [Google Scholar] [CrossRef] [PubMed]
  320. Update on metformin safety in pregnancy. Drug Ther. Bull. 2022, 60, 99. [CrossRef] [PubMed]
  321. Notaro, A.L.G.; Neto, F.T.L. The use of metformin in women with polycystic ovary syndrome: An updated review. J Assist Reprod. Genet. 2022, 39, 573–579. [Google Scholar] [CrossRef]
  322. Bailey, C.J.; Gwilt, M. Diabetes, Metformin and the Clinical Course of Covid-19: Outcomes, Mechanisms and Suggestions on the Therapeutic Use of Metformin. Front. Pharm. 2022, 13, 784459. [Google Scholar] [CrossRef] [PubMed]
  323. Syngelaki, A.; Nicolaides, K.H.; Balani, J.; Hyer, S.; Akolekar, R.; Kotecha, R.; Pastides, A.; Shehata, H. Metformin versus Placebo in Obese Pregnant Women without Diabetes Mellitus. N. Engl. J. Med. 2016, 374, 434–443. [Google Scholar] [CrossRef] [PubMed]
  324. Ferro, A.; Goyal, S.; Kim, S.; Wu, H.; Taunk, N.K.; Schiff, D.; Pirlamarla, A.; Haffty, B.G. Evaluation of Diabetic Patients with Breast Cancer Treated with Metformin during Adjuvant Radiotherapy. Int. J. Breast Cancer 2013, 2013, 659723. [Google Scholar] [CrossRef] [Green Version]
  325. Mortezaee, K.; Shabeeb, D.; Musa, A.E.; Najafi, M.; Farhood, B. Metformin as a Radiation Modifier; Implications to Normal Tissue Protection and Tumor Sensitization. Curr. Clin. Pharm. 2019, 14, 41–53. [Google Scholar] [CrossRef] [PubMed]
  326. Brown, S.L.; Kolozsvary, A.; Isrow, D.M.; Al Feghali, K.; Lapanowski, K.; Jenrow, K.A.; Kim, J.H. A Novel Mechanism of High Dose Radiation Sensitization by Metformin. Front. Oncol. 2019, 9, 247. [Google Scholar] [CrossRef] [PubMed]
  327. Suissa, S.; Azoulay, L. Metformin and cancer: Mounting evidence against an association. Diabetes Care 2014, 37, 1786–1788. [Google Scholar] [CrossRef] [Green Version]
  328. Mamtani, R.; Pfanzelter, N.; Haynes, K.; Finkelman, B.S.; Wang, X.; Keefe, S.M.; Haas, N.B.; Vaughn, D.J.; Lewis, J.D. Incidence of bladder cancer in patients with type 2 diabetes treated with metformin or sulfonylureas. Diabetes Care 2014, 37, 1910–1917. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  329. Pollak, M.N. Investigating metformin for cancer prevention and treatment: The end of the beginning. Cancer Discov. 2012, 2, 778–790. [Google Scholar] [CrossRef] [Green Version]
  330. Samuel, S.M.; Varghese, E.; Varghese, S.; Büsselberg, D. Challenges and perspectives in the treatment of diabetes associated breast cancer. Cancer Treat. Rev. 2018, 70, 98–111. [Google Scholar] [CrossRef] [Green Version]
  331. Hatoum, D.; McGowan, E.M. Recent advances in the use of metformin: Can treating diabetes prevent breast cancer? Biomed. Res. Int 2015, 2015, 548436. [Google Scholar] [CrossRef] [Green Version]
  332. Samuel, S.M.; Varghese, E.; Kubatka, P.; Triggle, C.R.; Büsselberg, D. Metformin: The Answer to Cancer in a Flower? Current Knowledge and Future Prospects of Metformin as an Anti-Cancer Agent in Breast Cancer. Biomolecules 2019, 9, 846. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  333. Scherbakov, A.M.; Sorokin, D.V.; Tatarskiy, V.V., Jr.; Prokhorov, N.S.; Semina, S.E.; Berstein, L.M.; Krasil’nikov, M.A. The phenomenon of acquired resistance to metformin in breast cancer cells: The interaction of growth pathways and estrogen receptor signaling. IUBMB Life 2016, 68, 281–292. [Google Scholar] [CrossRef] [PubMed]
  334. Farmer, R.E.; Ford, D.; Forbes, H.J.; Chaturvedi, N.; Kaplan, R.; Smeeth, L.; Bhaskaran, K. Metformin and cancer in type 2 diabetes: A systematic review and comprehensive bias evaluation. Int. J. Epidemiol. 2017, 46, 728–744. [Google Scholar] [CrossRef] [PubMed]
  335. Elgendy, M.; Cirò, M.; Hosseini, A.; Weiszmann, J.; Mazzarella, L.; Ferrari, E.; Cazzoli, R.; Curigliano, G.; DeCensi, A.; Bonanni, B.; et al. Combination of Hypoglycemia and Metformin Impairs Tumor Metabolic Plasticity and Growth by Modulating the PP2A-GSK3β-MCL-1 Axis. Cancer Cell 2019, 35, 798–815.e15. [Google Scholar] [CrossRef]
Figure 1. The main systemic effects of metformin to lower blood sugar. GLP1: glucagon-like peptide-1; GDF15: growth differentiation factor 15.
Figure 1. The main systemic effects of metformin to lower blood sugar. GLP1: glucagon-like peptide-1; GDF15: growth differentiation factor 15.
Biomolecules 13 00250 g001
Figure 2. The main cellular actions of metformin to lower blood sugar. OCT1: organic cation trans porter 1; FBP1: fructose-1,6-bisphosphatase; IRS1: insulin receptor substrate 1; GLUT1: glucose transporter protein 1; mGPD: mitochondrial glycerophosphate dehydrogenase; OXPHOS: oxidative phosphorylation; ATP: adenosine triphosphate; AMP: adenosine monophosphate; cAMP: cyclic adenosine monophosphate; IR: insulin receptor; PKA: protein kinase A; ACC: acetyl coenzyme A carboxylase; mTOR: mammalian target of rapamycin; AMPK: adenosine monophosphate-activated protein kinase; CaMKKβ: calcium/calmodulin-dependent protein kinaseβ; LKB1: liver kinase B1; AXIN: axis inhibitor; SIRT1: Sirtuin1; v-ATPase: vacuolar ATP hydrolase.
Figure 2. The main cellular actions of metformin to lower blood sugar. OCT1: organic cation trans porter 1; FBP1: fructose-1,6-bisphosphatase; IRS1: insulin receptor substrate 1; GLUT1: glucose transporter protein 1; mGPD: mitochondrial glycerophosphate dehydrogenase; OXPHOS: oxidative phosphorylation; ATP: adenosine triphosphate; AMP: adenosine monophosphate; cAMP: cyclic adenosine monophosphate; IR: insulin receptor; PKA: protein kinase A; ACC: acetyl coenzyme A carboxylase; mTOR: mammalian target of rapamycin; AMPK: adenosine monophosphate-activated protein kinase; CaMKKβ: calcium/calmodulin-dependent protein kinaseβ; LKB1: liver kinase B1; AXIN: axis inhibitor; SIRT1: Sirtuin1; v-ATPase: vacuolar ATP hydrolase.
Biomolecules 13 00250 g002
Figure 3. The main cellular anti-cancer actions of metformin. OCT1: organic cation transporter 1; AMP: adenosine monophosphate; ATP: adenosine triphosphate; OXPHOS: oxidative phosphorylation; REDD1: deoxyribonucleic acid (DNA) damage response 1; TNFα: tumor necrosis factor α; PTEN: phosphatase and tensin homolog; STAT3: transcription 3; IGF-1: insulin growth factor-1; YAP: yes-associated protein; TAZ: tea domain transcription factor; PI3K: phosphoinositide 3-kinase; AKT: protein kinase B; mTOR: mammalian target of rapamycin; COX-2: cyclooxygenase-2; Bcl-2: B-cell lymphoma-2; Bax: Bcl-2 associated x; NF-κB: nuclear factor kappa B; AMPK: adenosine monophosphate-activated protein kinase; ROS: reactive oxygen species; LKB1: liver kinase B1; FOXO3: forkhead box O3; HATs: histone acetyl-transferases; HDACs: histone deacetylases; DNMT: deoxyribonucleic acid (DNA) methyltransferase; ACC: acetyl coenzyme A carboxylase; HIF-1α: hypoxia-inducible factor-1α; Cyclin D1: cell cycle protein D1; EMT: mesenchymal transition; FASN: fatty acid synthase; UPR: unfolded protein response; MEK: mitogen-activated protein kinase; MRK: extracellular-related kinase; FasL: Fas ligand.
Figure 3. The main cellular anti-cancer actions of metformin. OCT1: organic cation transporter 1; AMP: adenosine monophosphate; ATP: adenosine triphosphate; OXPHOS: oxidative phosphorylation; REDD1: deoxyribonucleic acid (DNA) damage response 1; TNFα: tumor necrosis factor α; PTEN: phosphatase and tensin homolog; STAT3: transcription 3; IGF-1: insulin growth factor-1; YAP: yes-associated protein; TAZ: tea domain transcription factor; PI3K: phosphoinositide 3-kinase; AKT: protein kinase B; mTOR: mammalian target of rapamycin; COX-2: cyclooxygenase-2; Bcl-2: B-cell lymphoma-2; Bax: Bcl-2 associated x; NF-κB: nuclear factor kappa B; AMPK: adenosine monophosphate-activated protein kinase; ROS: reactive oxygen species; LKB1: liver kinase B1; FOXO3: forkhead box O3; HATs: histone acetyl-transferases; HDACs: histone deacetylases; DNMT: deoxyribonucleic acid (DNA) methyltransferase; ACC: acetyl coenzyme A carboxylase; HIF-1α: hypoxia-inducible factor-1α; Cyclin D1: cell cycle protein D1; EMT: mesenchymal transition; FASN: fatty acid synthase; UPR: unfolded protein response; MEK: mitogen-activated protein kinase; MRK: extracellular-related kinase; FasL: Fas ligand.
Biomolecules 13 00250 g003
Table 1. Summary of preclinical (in vivo) use of metformin in hematological malignancies models.
Table 1. Summary of preclinical (in vivo) use of metformin in hematological malignancies models.
Research SubjectsModelsMetformin DosageYearJoint Effect of Other DrugsEffectsRef.
NCG miceAcute Myelocytic
Leukemia
40 mg/kg/day i.v., 4 days2022Ara-CMetformin potentiated the anti-tumor efficacy of Ara-C in vivo in an NCG immunodeficient mouse xenograft model by inhibiting the mitochondrial transfer and OXPHOS activity in the engrafted human AML cells.[195]
Female nude mice (aged 5 weeks; average weight, 16 g)Acute Myelocytic
Leukemia
200 mg/kg i.p., 14 days2022VenetoclaxMetformin downregulated the expression of anti-apoptotic proteins Mcl-1 and Bcl-xl by inhibiting protein production, and shows a synergistic anti-tumor effect with ABT-199 in acute myeloid leukemia.[196]
Female C57BL/KaLwRij mice (aged 5, 6 weeks)Multiple Myeloma2.5 mg/mL p.o., 4 weeks2022-Metformin increased OPN expression in preosteoblasts, increasing myeloma cell adherence.[197]
MDR canines were male/neutered with recurrent B-cell lymphoma.Lymphoma250 mg OD or 500 mg BID p.o., 0–184 days2022CHOP or DoxorubicinMetformin reduced MDR protein markers in all canines in the study.[198]
BALC female nude mice (SPF level)Acute Myelocytic
Leukemia
125 mg/kg i.p., 21 days2020Ara-CThe synergistic anti-tumor effect of Ara-C/metformin in AML was via inhibiting the mTORC1/P70S6K pathway.[199]
SCID mice (aged 6–8 weeks)Burkitt’s Lymphoma2 μg/mL p.o., 3 months2020RituximabMetformin in combination with rituximab showed improved survival compared with rituximab monotherapy.[200]
NOD/SCID mice (aged 5 weeks)Multiple Myeloma250 mg/kg/day p.o., 21 days2018-Metformin inhibited the proliferation of myeloma cells by inducing autophagy and cell-cycle arrest. The molecular mechanism involved the dual repression of mTORC1 and mTORC2 pathways via AMPK activation.[201]
miceMultiple Myeloma600 μg/mL i.v., 18 or 27 days2015BortezomibMetformin suppressed GRP78, and supported the pharmacologic repositioning of metformin to enhance the anti-myeloma benefit of bortezomib.[194]
NOD/SCID CB17-strain 394 (white) mice (aged 5, 6 weeks)Multiple Myeloma125 mg/kg i.p., 1 week2015RitonavirRitonavir and metformin effectively suppressed AKT and mTORC1 phosphorylation and pro-survival BCL-2 family member MCL-1 expression in multiple myeloma cell lines.[202]
male CB17/SCID mice (aged 4 weeks)Multiple Myeloma200 mg/kg/day i.p., 21 days2015DexamethasoneMetformin inhibited multiple myeloma cell proliferation via the IGF-1R/PI3K/AKT/mTOR signaling pathway.[191]
Pten-deficient mice (tPTEN−/−)Lymphoma2 mg/mouse/day i.p., 18 days2013-Metformin strongly decreased the growth of luciferase-expressing tPTEN−/− cells xenografted in nude mice.[203]
nude mice (aged 5–6 weeks)Lymphoma4 mg/kg/day (i.p.) or 3 mg/kg/day (p.o.) for 21 days2012Doxorubicin or TemsirolimusMetformin induced AMPK activation, mTOR inhibition and remarkably blocked tumor growth in murine lymphoma xenografts.[193]
i.p.: intraperitoneal injection; i.v.: intravenous injection; p.o.: preoral; OD: once daily; BID: twice daily; CHOP: four-drug chemotherapy cocktail given over 19 weeks including Cyclophosphamide, Doxorubicin (Adriamycin), Vincristine, Prednisone; Ara-C: cytarabine; AMPK: adenosine monophosphate-activated protein kinase; mTOR: mammalian target of rapamycin; AKT: protein kinase B; PI3K: phosphoinositide 3-kinase; OXPHOS: oxidative phosphorylation; BCL-2: B-cell lymphoma-2; MDR: multidrug resistance; GRP78: glucose-regulated protein 78; AML: acute myelocytic leukemia; OPN: osteopontin; MCL-1: myeloid cell leukemia-1; IGF-1R: insulin-like growth factor-1 receptor; PTEN: phosphatase and tensin homolog deleted on chromosome 10; mTORC1: mechanistic target of rapamycin complex 1; P70S6K: 70 kDa ribosomal protein S6 kinase; NCG: NOD/SCID/IL-2rynull; Bcl-xl: B-cell lymphoma xl; ABT-199: venetoclax; SCID: severe compromised immunodeficiency; NOD: non-obese diabetes.
Table 2. Summary of preclinical (in vitro) use of metformin in leukemia model.
Table 2. Summary of preclinical (in vitro) use of metformin in leukemia model.
Research SubjectsModelsMetformin DosageYearJoint Effect of Other DrugsEffectsRef.
U937 and HL-60 cellsAML5 mM, 24 h2022Ara-CMetformin inhibited mitochondrial transfer and significantly enhanced the chemosensitivity of AML cells co-cultured with BMSCs.[195]
LAMA-84s, LAMA-84r and K562 cellsCML2.5–80 mM, 48 h2022thymoquinoneMetformin and thymoquinone monotherapies possessed a significant anti-leukemic effect that was more pronounced when combinatorial therapies are applied.[204]
K562, K562TRBSR, MOLM13, THP1, HEL, HL60, HL60TRBSR, OCI/AML3 cells, KG1 cells and HEK293T.AML5 mM, 72 h2022-AML cells with an MLL/AF9 genotype have a high dependency on OXPHOS and could be therapeutically targeted by metformin.[205]
697 cellsALL0–15 mM, 48 h2022cisplatinMetformin and cisplatin exerted a cytotoxic effect on 697 cells. When both drugs were combined they demonstrated antagonistic effects.[206]
KG-1, Kasumi-1 and THP-1 cellsAML0–2 mM, 48 h2022venetoclaxMetformin downregulated the expression of anti-apoptotic proteins Mcl-1 and Bcl-xl by inhibiting protein production, and showed a synergistic anti-tumor effect with ABT-199 in acute myeloid leukemia.[196]
NB4, KG1 and KG1A cell linesLeukemia10 mM, 24 h or 72 h2021MCL1 Inhibitor S63845A combined treatment with metformin and S63845 had a stronger inhibitory effect on AML cell oxidative phosphorylation and glycolysis rate and, consequently, on cellular ATP levels. The apoptosis induced by treatment was related to the change of ROS level.[207]
K562 and KU812 cellsCML0–10 mM, 48 h2021nilotinibMetformin was effective in decreasing phosphorylated JNK levels, resulting in the restoration of nilotinib sensitivity.[208]
OCI/AML2, K562, OCI/AML3, and THP-1 cellsLeukemia0–10 mM, 24–72 h2020TP-0903Metformin inhibited phosphorylation-dependent activation of TAM RTKs, which regulate molecular pathways associated with leukemia.[209]
HL-60 and THP-1 cellsAML0–12 mM, 24–72 h2020Ara-CThe synergistic anti-tumor effect of Ara-C and metformin in AML was achieved via inhibition of the mTORC1/P70S6K pathway.[199]
SKM-1 cellsAML-MDS0–20 mM, 24–72 h2019-Metformin inhibited proliferation of SKM-1 cells, potentially via an AMPK-mediated cell cycle arrest.[210]
chemoresistant AML patientsAML10 mM, 24 h2019Cytarabine, VenetoclaxMetformin decreased therapy-resistant-AML cell oxidative phosphorylation in vitro, while co-treatment with cytarabine and venetoclax slightly increased the effect.[211]
K562 cellsCML0–30 mM, 48 h2019-Metformin can inhibit the growth and proliferation of K562 cells and promote the apoptosis of K562 cells by inhibiting glycolysis energy metabolism. The PI3K/Akt/mTOR signaling pathway may be one of the molecular mechanisms of metformin on k562 cells.[212]
U937 and THP-1 cellsAML0–10 mM, 24 h or 48 h2018Diflunisal, DiclofenacLow concentrations of metformin and the two NSAIDs diclofenac and diflunisal exerted a synergistic inhibitory effect on AML proliferation and induced apoptosis, most likely by blocking tumor-cell metabolism.[213]
HL-60 cellsAPL0–1 μM, 24 h, 48 h, and 72 h2018PaclitaxelThe combination of paclitaxel and metformin triggered differentiation and apoptosis according to gene expression changes.[214]
HL60 cellsAPL15 μM, 150 μM, and 1.5 mM, 4 h2018-Low concentrations (15 and 150 µM) increased both oxidative phosphorylation and the oxidative stress response, acting on the AMPK/Sirt1 pathway, while high concentration (1.5 mM) inhibited the respiratory chain and altered cell morphology, becoming toxic to the cells.[215]
Dami and MEG-01 cellsAMKL4 mM, 0–72 h2017-Metformin inhibited the proliferation and induced the apoptosis of human megakaryoblastic cell lines.[216]
K562 and A301 cellsCML
ALL
0–25 mM, 24 h2017VincristineAMPK activation was critical to metformin’s effects on vincristine-induced apoptosis.[217]
MV4-11, MOLM-14, OCI-AML3, Nomo-1, THP-1 and HL-60 cellsAML10 mM, 24 h or 48 h20166-BenzylthioinosineThe combination of 6-BT with metformin resulted in significant cytotoxicity (60–70%) in monocytic AML cell lines and was associated with the inhibition of FLT3-ITD activated STAT5 and reduced c-Myc and GLUT-1 expression.[218]
OCI-AML3, REH, NALM-6-6, and KBM5 cellsAML
ALL
CML
0–10 mM, 0–16 h2016ABT-737Inhibition of mitochondrial metabolism by metformin or phenformin was associated with increased leukemia cell susceptibility to induction of intrinsic apoptosis.[219]
EHEB, JVM-2 and MEC-2 cellsLeukemic0.1–10 mM, 0–48 h2016Sodium DichloroacetateThe combination of metformin and DCA increased the cytotoxicity in the B-type leukemia cell line and the cell culture derived from B-CLL patients, which was the result of cell inhibition and apoptosis promotion.[220]
MV4-11(FLT3-ITD positive) and THP-1(FLT3-ITD negative) cellsAML0.2–16 mM, 0–72 h2015SorafenibIn the presence of metformin, the anti-cancer potential of sorafenib, accompanied by increased LC3 levels, was found to be synergistically enhanced with the remarkably reduced protein expression of the mTOR/p70S6K/4EBP1 pathway, while not appreciably altering the cell cycle.[221]
SUP-B15, K562 and K562R cellsALL
CML
0–50 mM, 0–72 h2015-Metformin mediated anti-leukemia effects by proapoptosis and inhibition of mTORC1 signaling, potentiating the anti-cancer efficacy of imatinib in Ph+ ALL cells. Metformin-induced autophagy was associated with the activation of the ERK pathway.[222]
10E1-CEM cellsALL10 mM, 0–72 h2015-Metformin induced cell-cycle arrest and apoptosis in drug-resistant leukemia cells.[223]
Kassumi, NB-4, THP-1, ML-2, K562, Jurkat, Raji and HUT-78 cellsAML
ALL
0–8 μM, 0–48 h2014-Metformin may be involved in the downregulation of FOXM1.
Metformin promoted the apoptosis of ML-2 cells, induced cell-cycle arrest at the G0/G1 and G2/M phases, and inhibited proliferation.
[224]
CCRF-CEM, Jurkat, REH and NALM6 cellsALL0–10 mM, 48 h2013-Metformin induced ALL cell death by triggering ER and proteotoxic stress and simultaneously down-regulating the physiologic UPR response responsible for effectively buffering proteotoxic stress.[225]
Kasumi-1, SKNO-1, HL-60, KG-1a and NB4 cellsAML
APL
0–5 mM, 0–72 h2012ATRAThe synergism between metformin and ATRA triggered the maturation pathway in APL cells.[226]
ALL: acute lymphocytic leukemia; AML: acute myeloid leukemia; APL: acute promyelocytic leukemia; CML: chronic myeloid leukemia; CLL: chronic lymphocytic leukemia; AMKL: acute megakaryocytic leukemia; MDS: myelodysplastic syndrome; Ara-C: cytarabine; ATRA: all-trans-retinoic acid; BMSCs: bone marrow mesenchymal stem cells; OXPHOS: oxidative phosphorylation; UPR: unfolded protein response; ER: endoplasmic reticulum; FOXM1: forkhead box protein M1; 4EBP1: 4e-binding protein 1; ATP: adenosine triphosphate; ROS: reactive oxygen species; mTORC1: mammalian target of rapamycin complex 1; P70S6K: p70S6 kinase; AMPK: AMP-activated protein kinase; PI3K: phosphatidylinositol 3 kinase; AKT: protein kinase B; mTOR: mammalian target of rapamycin; NSAIDs: non-steroidal anti-inflammatory drugs; STAT5: transcription 5; ABT-199: venetoclax; MLL/AF9: mixed-lineage leukaemia-AF9; FLT3-ITD: flt3-internal tandem duplication; LC3: light chain 3; DCA: deoxycholic; GLUT-1: glucose transporter-1; Bcl-xl: B-cell lymphoma-xl; Mcl-1: myeloid cell leukemia-1; TAM RTKs: TAM receptor tyrosine kinases.
Table 3. Summary of preclinical (in vitro) use of metformin in lymphoma model.
Table 3. Summary of preclinical (in vitro) use of metformin in lymphoma model.
Research SubjectsModelsMetformin DosageYearJoint Effect of Other DrugsEffectsRef.
OCI-LY-10, DB and THP-1 cellsLymphoma200 μM, 72 h2022-Metformin could target altered lipid metabolism and decrease M2 macrophages in DLBCL, especially in CD5+ non-DE lymphoma.[227]
DAUDI cellsBurkitt‘s lymphoma10 mM, 0–72 h2021-Metformin could induce cell death in BL cells by stressing cellular metabolism via the induction of GLUT1, PKM2, and LDHA.[228]
Raji, U2392 and RL cell linesLymphoma4, 8, or 16 mM, 0–72 h2020RituximabMetformin caused cell-cycle arrest in the G1 phase. Metformin induced apoptosis, ROS production, and increased mitochondrial membrane permeability. Metformin exhibited additive/synergistic effects when combined with traditional chemotherapy or rituximab in vitro.[200]
Hut78, H9 and HH cellsLymphoma1 or 10 mM, 2 h2019-Metformin inhibited mTORC1 signaling and migration of SS cells induced by SDF-1.[229]
Daudi, SUDHL-4 and Jeko-1 cellsLymphoma0–10 mM, 48 h or 7 days2018Venetoclax, and BAY-1143572Metformin inhibited oxidative phosphorylation in lymphoma cells.
Metformin increased caspase 3/7 activity in venetoclax and BAY-1143572-treated lymphoma cells.
Metformin showed potentiation with venetoclax and BAY-1143572 in a cell-type-dependent manner.
[230]
BC3 and BCBL1 cellsPrimary exudative lymphoma15, 20 and 30 mM, 24 h2017BortezomibThe cytotoxic effect of metformin was correlated with intracellular reactive oxygen species reduction, activation of AMPK and the inhibition of pro-survival pathways such as mTOR and STAT3.
Metformin altered UPR activated by bortezomib, leading to a reduced expression of BiP, upregulation of CHOP and downregulation of Bcl-2.
[231]
Jurkat, Sil-ALL, SupT1 and Ke37 cellsLymphoma
0–10 mM, 0–48 h20132-deoxyglucoseMetformin synergized with 2DG to impair tumour cell survival.[203]
SU-DHL-4, Nalmawa, DB, SU-DHL-5, Daudi, Jurkat, 6T-CEM, Kappas, H9 and HUT78 cellsLymphoma0–40 mM, 0–72 h2012-Metformin-induced AMPK activation was associated with the inhibition of mTOR signaling without involving AKT.[193]
2DG: 2-deoxyglucose; AKT: protein kinase B; mTOR: mammalian target of rapamycin; AMPK: AMP-activated protein kinase; STAT3: transcription 3; UPR: unfolded protein response; DLBCL: diffuse large B-cell lymphoma; mTORC1: mammalian target of rapamycin complex 1; Bcl-2: B-cell lymphoma-2; DE: MYC/BCL2 double expressor; GLUT1: glucose transporter 1; PKM2: Pyruvate kinase M2; ROS: reactive oxygen species; SDF-1: stromal cell-derived factor-1; LDHA: Lactate Dehydrogenase A; CHOP: C/EBP homologous protein.
Table 4. Summary of preclinical (in vitro) use of metformin in myeloma model.
Table 4. Summary of preclinical (in vitro) use of metformin in myeloma model.
Research SubjectsModelsMetformin DosageYearJoint Effect of Other DrugsEffectsRef.
Jeko-1 cellsMCL10 mM, 0–72 h2022BortezomibMetformin treatment induced the resistance of cancer cells to the proteasome inhibitor Bortezomib by impairing the activity and assembly of the 26S proteasome complexes.[232]
5TGM1, MM.1S cellsMM0–10 mM, 48 h2022-Metformin increased OPN expression in preosteoblasts, increasing myeloma cell adherence.[197]
RPMI-8226 cellsMM0–40 mM, 48 h2020MelphalanMetformin could promote DNA damage induced by melphalan and decreased the concentration of ATP in the process of repairing DNA damage to hinder the anti-apoptotic process of tumor cells.[233]
L363 and RPMI-8226 cellsMM0–10 mM, 0–72 h2020-Metformin inhibited HIF-1 signaling in MM cells, and the effect of metformin was mainly oxygen dependent.
Metformin triggered the growth arrest without inducing apoptosis in either normoxic or hypoxic conditions.
[234]
I-8266, U266-
B1, MM.1S, OPM1, OPM2, ANBL6, OCI-MY5, JJN3,
KP6, DP6, KAS-6/1, KMS12PE, K562 and NALM6 cells
MM
CML
ALL
0–80 mM, 0–72 h2019-Metformin specifically decreased IL-6R expression which is mediated via AMPK, mTOR, and miR34a.[92]
RPMI8226, ARP-1 and OPM2 cellsMM0–40 mM, 24 h2019PFK15Metformin was found to inhibit PFKFB3 protein expression. PFK15 also demonstrated a synergistic effect with metformin to eliminate MM cells.[235]
U266 cellsMM5–50 mM, 0–72 h2018-Metformin could inhibit cell proliferation and induced U266 cell apoptosis via the mitochondrial apoptotic pathway.[236]
RPMI-8226 and U266 cellsMM0–80 mM, 0–72 h2018-Metformin inhibited the proliferation of myeloma cells by inducing autophagy and cell-cycle arrest. The molecular mechanism involved the dual repression of mTORC1 and mTORC2 pathways via AMPK activation.[201]
U266, RPMI8226, LP-1 and NCI-H929 cellsMM20 mM, 24 h2018FTY720Exposure to metformin in combination with FTY720 potently induced apoptosis in MM cells in a ROS-dependent manner.[237]
RPMI-8226 and U266 cellsMM0–80 mM, 0–72 h2017-Metformin can inhibit the proliferation and induce apoptosis of RPMI8226 and U266 cell lines, which may be related to downregulation of the STAT3 signal transduction pathway.[238]
RPMI-8226 cellsMM500 μM, 24 h2015BortezomibMetformin inhibited GRP78 to enhance the anti-myeloma effect of bortezomib.[194]
KMS11, L363, and JJN3 cellsMM0–5 mM, 0–72 h2015RitonavirRitonavir and metformin effectively suppressed AKT and mTORC1 phosphorylation and prosurvival BCL-2 family member MCL-1 expression in multiple myeloma cell lines.[202]
RPMI8226, MM.1S, MM.1R, and U266 cellsMM0–80 mM, 0–72 h2015DexamethasoneMetformin inhibitedMM cell proliferation via the IGF-1R/PI3K/AKT/mTOR signaling pathway.[191]
MCL: myeloid cell leukemia; MM: multiple myeloma; CML: chronic myeloid leukemia; ALL: acute lymphocytic leukemia; PI3K: phosphatidylinositol 3 kinase; AKT: protein kinase B; PFK15: (1-(Pyridin-4-yl)-3-(quinolin-2-yl)prop-2-en-1-one); IGF-1R: insulin-like growth factor-1 receptor; Mcl-1: myeloid cell leukemia-1; mTOR: mammalian target of rapamycin; mTORC1: mammalian target of rapamycin complex 1; ATP: adenosine triphosphate; AMPK: AMP-activated protein kinase; STAT3: transcription 3; BCL-2: B-cell lymphoma-2; GRP78: glucose-regulated protein 78; FTY720: fingolimod; OPN: osteopontin; DNA: deoxyribonucleic acid; HIF-1: hypoxia-inducible factor-1; IL-6R: IL-6 receptor; PFKFB3: 6-phosphofructo-2-kinase/fructose-2,6-biphosphatase; ROS: reactive oxygen species.
Table 5. Metformin-based combinatorial therapy in the treatment of HM.
Table 5. Metformin-based combinatorial therapy in the treatment of HM.
NCT NumberOfficial TitleActual EnrollmentStatusType of DiseaseCombination TreatmentPhase
03118128Effect of the Addition of Metformin Hydrochloride on the Prognosis of Patients With B-cell Precursor (Ph+ Negative) Acute Lymphoblastic Leukemia with High Expression of ABCB1 Gene102CompletedALLConventional ChemotherapyNA
01750567A Phase II Pilot Study of Metformin Therapy in Patients with Relapsed Chronic Lymphocytic Leukemia and Un-treated CLL Patients with Genomic Deletion 11q40RecruitingRelapsed CLLNonePhase 2
05326984Effect of Metformin on ABCB1 and AMPK Expression in Adolescents with Newly Diagnosed Acute Lymphoblastic Leukemia20RecruitingALLConventional ChemotherapyNA
04741945STOP-LEUKEMIA: Repurposing Metformin as a Leukemia-preventive Drug in CCUS and LR-MDS24RecruitingPreleukemia,
MDS,
Cytopenia
NonePhase 2
01324180A Phase I Window, Dose Escalating and Safety Trial of Metformin in Combination with Induction Chemotherapy in Relapsed Refractory Acute Lymphoblastic Leukemia: Metformin with Induction Chemotherapy of Vincristine, Dexamethasone, Doxorubicin, and PEG-asparaginase (VPLD)14CompletedALLVLPDPhase 1
02948283A Pilot Feasibility Study of Metformin/Ritonavir Combination Treatment in Patients with Relapsed/Refractory Multiple Myeloma or Chronic Lymphocytic Leukemia3CompletedR/R MM,
R/R CLL
RitonavirPhase 1
01486043Metformin as an Adjunctive Therapy for Transient Hyperglycemia in Patients with Acute Lymphoblastic Leukemia During Induction Chemotherapy4TerminatedALLInsulinNA
01849276A Phase I Study of Metformin and Cytarabine for the Treatment of Relapsed/Refractory Acute Myeloid Leukemia2TerminatedR/R AMLCytarabinePhase 1
00659568A Phase I Study of Temsirolimus in Combination with Metformin in Advanced Solid Tumours28CompletedLymphomaTemsirolimusPhase 1
02967276Phase II Trial, Open Label, Clinical Activity of Metformin in Combination with High-dose of Dexamethasone (HDdexa) in Patients with Relapsed/Refractory Multiple Myeloma28UnknownMMHigh doses of dexamethasonePhase 2
04850846A Randomized Placebo-Controlled Phase 2 Study of Metformin for the Prevention of Progression of Monoclonal Gammopathy of Undetermined Significance and Smoldering Multiple Myeloma80RecruitingMGUS,
SMM
NonePhase 2
03829020An Open-Label Phase 1 Study of Metformin and Nelfinavir in Combination with Bortezomib in Patients with Relapsed and/or Refractory Multiple Myeloma9Active, not recruitingR/R PCMBortezomib, Nelfinavir MesylatePhase 1
03200015Effect of Metformin in Combination with R-CHOP for the First Line Treatment of Patients with Diffuse Large B-cell Lymphoma15UnknownDLBCLRCHOPPhase 2
02815397A Phase II Study Evaluating the Efficacy and Safety of Metformin in Combination with Standard Induction Therapy (DA-EPOCH-R) for Previously Untreated C-myc+ Diffuse Large B-Cell Lymphoma2TerminatedDLBCLDA-EPOCHRPhase 2
03600363A Prospective Randomized Controlled Phase II Clinical Trial of Metformin in the Maintenance Therapy of High Risk Diffuse Large B Lymphoma/Stage III Follicular Lymphoma Patients with Complete Remission250UnknownDLBCL,
Stage III FL
RCHOPPhase 2
02531308A Phase ll Study Evaluating the Efficacy and Safety of Metformin in Combination with Standard Induction Therapy (RM-CHOP) for Previously Untreated Aggressive Diffuse Large B-cell Lymphoma5TerminatedDLBCLRCHOPPhase 2
02145559A Pharmacodynamic Study of Sirolimus and Metformin in Patients with Advanced Solid Tumors24CompletedLymphomaSirolimusPhase 1
NA: Not Applicable; CCUS: clonal cytopenia of undetermined significance; MDS: myelodysplastic syndrome; SMM: smoldering multiple myeloma; Refractory PCM: refractory plasma cell myeloma; RCHOP: rituximab, cyclophosphamide, doxorubicin, vincristine and prednisone; DA-EPOCHR: dose-adjusted etoposide, prednisone, vincristine, cyclophosphamide, doxorubicin, rituximab; Stage III FL: Stage III follicular Lymphoma; VPLD: vincristine, dexamethasone, doxorubicin and PEG asparaginase; Conventional Chemotherapy: the remission induction chemotherapy included a steroid pre-phase of 7 days of prednisone 60 mg/m2. The proper remission induction phase consisted of prednisone 60 mg/m2 daily from day 0 to 28; Vincristine 1.5 mg/m2 on days 0, 7, 14 and 21; Doxorubicin 25 mg/m2 on days 0, 7, 21; L-asparaginase 10,000 U/m2 on days 2, 4, 6, 8, 10 and 12. Etoposide 300 mg/m2 and cytarabine 300 mg/m2 on days 22, 25 and 29. Intrathecal chemotherapy was administered on days 0, 7, 14 and 21.
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

Zhang, Y.; Zhou, F.; Guan, J.; Zhou, L.; Chen, B. Action Mechanism of Metformin and Its Application in Hematological Malignancy Treatments: A Review. Biomolecules 2023, 13, 250. https://doi.org/10.3390/biom13020250

AMA Style

Zhang Y, Zhou F, Guan J, Zhou L, Chen B. Action Mechanism of Metformin and Its Application in Hematological Malignancy Treatments: A Review. Biomolecules. 2023; 13(2):250. https://doi.org/10.3390/biom13020250

Chicago/Turabian Style

Zhang, Yi, Fang Zhou, Jiaheng Guan, Lukun Zhou, and Baoan Chen. 2023. "Action Mechanism of Metformin and Its Application in Hematological Malignancy Treatments: A Review" Biomolecules 13, no. 2: 250. https://doi.org/10.3390/biom13020250

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