Next Article in Journal
Isolated German Sign Language Recognition for Classifying Polar Answers Using Landmarks and Lightweight Transformers
Next Article in Special Issue
Injectable Hydrogel Systems for Targeted Drug Delivery: From Site-Specific Application to Design Strategy
Previous Article in Journal
Sodium-Reduced Canned Dog Pâtés Enriched with Collagen Hydrolysate and Salicornia perennans: A Sustainable Strategy to Enhance Technological Quality and Oxidative Stability
Previous Article in Special Issue
In Silico Identification of Six Mushroom-Derived Sterol and Triterpenoid Compounds as Potential P-Glycoprotein Modulators in Multidrug Resistance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Dual Role of Metformin: Repurposing an Antidiabetic Drug for Cancer Therapy

1
UNIPRO—Oral Pathology and Rehabilitation Research Unit, University Institute of Health Sciences (IUCS)—CESPU, 4585-116 Gandra, Portugal
2
Associate Laboratory i4HB—Institute for Health and Bioeconomy, University Institute of Health Sciences (IUCS)—CESPU, 4585-116 Gandra, Portugal
3
UCIBIO—Applied Molecular Biosciences Unit, Translational Toxicology Research Laboratory, University Institute of Health Sciences (1H-TOXRUN, IUCS-CESPU), 4585-116 Gandra, Portugal
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Appl. Sci. 2025, 15(21), 11576; https://doi.org/10.3390/app152111576
Submission received: 3 October 2025 / Revised: 19 October 2025 / Accepted: 27 October 2025 / Published: 29 October 2025
(This article belongs to the Special Issue Anticancer Drugs: New Developments and Discoveries)

Abstract

Maintaining glucose homeostasis is vital for normal physiological function, and any disturbance in this balance is associated with the development of degenerative and chronic diseases, like Type 2 Diabetes (T2D) and certain types of cancer, where altered glucose metabolism plays a central role. Epidemiological evidence indicates a positive association between diabetes and an increased risk of developing certain types of cancer. Such a correlation may be driven by shared risk factors, namely obesity, inflammation, and insulin resistance. The observed association between diabetes and an increased risk of certain cancers, along with the rising incidence of both diseases, has in recent years raised interest in treatments that may target both conditions. Among them, the biguanide metformin, the first-line drug prescribed for T2D, has attracted significant attention as a repurposed drug due to its potential role in cancer treatment. Metformin is a glucose-lowering drug that reduces hepatic glucose production and improves insulin sensitivity, promoting glucose uptake by the skeletal muscle, contributing to better glycemic control in individuals with T2D and prediabetic syndromes. However, beyond its metabolic effects, metformin also influences key signaling pathways involved in cell growth and survival, such as the AMP-activated protein kinase (AMPK)/mTOR axis, raising interest in its potential application as an anticancer agent. Furthermore, metformin inhibits mitochondrial complex I, disrupting cellular energy production, which is essential for cancer proliferation. This review aims to explore and clarify the multifunctional role of metformin in both T2D and cancer, focusing on the metabolic alterations observed in these diseases. It highlights how glucose metabolism dysregulation contributes to disease progression in both contexts and explores the molecular targets of metformin in each condition and its potential for dual therapeutic benefit. Finally, selected clinical trials concerning metformin use in cancer therapy, alone or in combination, will be presented, highlighting its potential to enhance treatment response, reduce resistance, and improve overall patient outcomes.

1. Introduction

Type 2 diabetes mellitus (T2D) is a chronic global health problem, characterized by decreased insulin response and persistently elevated blood glucose levels, which is associated with several long-term comorbidities such as retinopathy, nephropathy, neuropathy, cardiovascular disease, and non-alcoholic fatty liver disease (NAFLD) [1]. The prevalence of the disease has dramatically increased in recent years worldwide, mainly due to unhealthy lifestyles (increased consumption of processed and sugary foods coupled with reduced physical activity), growing obesity prevalence, and aging of the population. Nevertheless, T2D is a multifactorial disease, and several factors can be involved in its development, with environmental factors interacting with genetic predisposition. Many of the patients who develop T2D are obese, and the increase in sedentary lifestyle and obesity appears as risk factors for this type of disease, associated with heredity, being a strong link between diabetes mellitus and obesity consistently observed across different racial and ethnic groups [2]. The pathogenesis of T2D involves resistance to the action of insulin and may also sometimes involve failures in insulin secretion. Fat accumulation in adipose tissue contributes to insulin resistance, whereas the accumulation in the pancreas is related to pancreatic beta cells’ dysfunction [3]. Such insulin resistance is closely associated with low-grade chronic inflammation, which contributes to the pathogenesis of the disease [4]. White adipose tissue, especially the visceral, produces several inflammatory molecules, like cytokines and chemokines, triggering inflammatory processes [3]. Such low-grade chronic inflammation, assessed by raised levels of C-reactive protein (CRP), can also contribute to other complications of diabetes, including the increased risk for other chronic degenerative diseases, like cancer, cardiovascular (CVD), and neurodegenerative diseases [5,6,7]. In fact, diabetes is associated with an increased risk of developing certain types of cancer and an increased cancer mortality, namely for colon, pancreatic, liver, kidney, bladder, endometrium, and breast cancers, although the risk varies between women and men [8,9,10]. T2D and cancer share some risk factors, including overweight and obesity, unhealthy diet, lack of physical activity, and aging. Moreover, several patients are diagnosed with both those cancers and diabetes more often than would happen just by chance, although no correlation was found with other cancers, or is inconclusive for lung cancer or non-Hodgkin lymphoma, and in some cases, like prostate cancer, the risk is actually lower [9]. Thus, T2D might impact the development of some types of cancer through various mechanisms, including low-grade chronic inflammation. Chronic inflammation is directly associated with cancer, creating conditions that support tumor growth and development through the release of reactive oxygen species (ROS), growth factors, and cytokines that promote cell growth and proliferation, altering the immune response and leading to changes in the tumor microenvironment [11,12]. Likewise, in the 2011 paper “Hallmarks of Cancer: The Next Generation”, revising the original identification of cancer hallmarks, authored by Douglas Hanahan and Robert A. Weinberg, tumor-promoting inflammation has been identified as a key enabling characteristic fostering multiple hallmark functions that lead to tumor proliferation and progression [13]. This inflammatory state can be found in either T2D or cancer, worsening outcomes in both diseases. Furthermore, other associations can be found in both diseases, like dysregulated glucose metabolism and signaling pathways, such as insulin/IGF or phosphoinositide 3-kinase (PI3K)/protein kinase B (Akt)/mechanistic target of rapamycin (mTOR) [14]. Some well-known antidiabetic drugs belonging to the biguanide class, like metformin (the most prescribed) or phenformin, have these pathways as targets [15,16,17]. Although originally developed for diabetes treatment, these drugs have demonstrated interesting antitumor properties, and several reports exploit these properties by repurposing them for cancer therapy. Since these drugs are well known in terms of their effects and side effects, their use in oncology could be safer and more predictable. Accordingly, evidence from various studies suggests that the use of these compounds, namely metformin, was consistently linked to a lower risk of cancer development in diabetic patients, particularly pancreatic, colorectal, and liver cancers [18,19]. Furthermore, their use has also been associated with increased survival rates in cancer patients, decreased resistance, and a lower probability of cancer recurrence [19]. In this review, we address these multiple roles of metformin therapeutics, both in T2D and cancer, exploiting its effect on glucose metabolism and homeostasis, its cellular targets and mechanisms of action, as an antidiabetic or anticancer drug. Finally, we will review the clinical mechanisms of metformin decreasing cancer resistance, as well as clinical trials evaluating metformin in cancer therapy, combined with standard therapies.

2. Disturbance of Glucose Homeostasis in Diabetes and Cancer

As previously mentioned, a common feature of both diabetes and cancer is the disturbance of glucose homeostasis. However, the causes and consequences of this disturbance differ between the two diseases and have distinct implications. Glucose is a fundamental and universal fuel in the human body, particularly vital for tissues and cells like the brain, retina, erythrocytes, and renal medulla [20,21,22], being regulated through a complex network involving hormones, transporters, and metabolic and signaling pathways. The first step of glucose utilization consists of its uptake into the cell, a process mediated by glucose transporters (GLUTs). Different GLUT isoforms are expressed differently in specific tissues, ensuring that glucose delivery is adapted to the metabolic needs of each tissue. After being transported into the cell, glucose can be directed into distinct metabolic pathways according to the cell’s energy status and functional requirements. It may be directed towards glycolysis to produce ATP, stored as glycogen, or enter the pentose phosphate pathway (PPP) to generate NADPH and ribose-5-phosphate, providing cells with energy and biosynthetic precursors. Together with glucagon, insulin is essential for glucose homeostasis. They play opposing roles in maintaining glucose homeostasis by promoting the storage or mobilization of glucose, respectively. Additionally, hormones such as cortisol, adrenaline, and growth hormone are also players in this regulation, particularly during stress or fasting, to help maintain metabolic balance [23,24]. Intracellular signaling pathways, including the PI3K/Akt and AMP-activated protein kinase (AMPK) pathways, also participate in glucose homeostasis [25]. In healthy individuals, such regulation ensures a balance between glucose supply and demand across different tissues, according to their needs and physiological status. However, in some diseases, like diabetes and cancer, this accurate regulation becomes disrupted (Figure 1). Although both disorders involve altered glucose metabolism, the inherent mechanisms and adaptive responses show marked differences, reflecting the distinct pathophysiological features of each disease.
GLUTs belong to the solute transporter (SLC2A) family of proteins and are present in many tissues/cells of the body, e.g., brain, erythrocytes, adipocytes, and liver, where they mediate glucose uptake [26]. The fourteen different isoforms of GLUTs are subdivided into three distinct protein classes, according to their sequence homology. Each GLUT isoform has a unique tissue distribution, a substrate specificity and a specific physiological function [27]. All GLUT proteins were originally assumed to catalyze the transport of hexoses into and out of cells. This is clearly the case for the class 1 GLUT proteins (GLUTs 1–4 and 14). However, class 2 (GLUTs 5, 7, 9 and 11) and class 3 (GLUTs 6, 8, 10, 12 and 13) GLUT proteins do not necessarily have a primary role in catalyzing glucose transport [28]. GLUT-1 is expressed in tissues with a high glycolytic rate, such as erythrocytes, being responsible for glucose uptake in high-need cells [26,28]. However, this transporter also plays a central role in tumorigenesis as it delivers glucose in hypoxic environments. In fact, GLUT-1, a target gene of HIF-1, is highly expressed in hypoxic cancer cells, allowing the maintenance of a high metabolic rate in these cells [26].

2.1. Diabetes

Diabetes mellitus is a metabolic disease characterized by a deficiency in insulin production, increased destruction of the hormone, or its inefficiency [29,30,31,32]. Its main consequence is the loss of cells’ ability to capture glucose from the blood, making hyperglycemia the most prominent symptom of the disease [29,30,31,32]. As previously mentioned, T2D results from a relative resistance to endogenous insulin—a condition promoted by environmental factors, namely obesity –, and from a decreased secretion of this hormone [29,30,31,32]. It has been associated with several defects in the early and intermediate steps of the insulin signaling pathway [32,33,34,35,36,37] and accounts for about 90% of diabetics [30,31,38].
Insulin is produced in the β cells of the pancreas in response to rising blood glucose levels [29]. Among insulin’s functions are storing fuel and facilitating glucose uptake by cells, particularly in adipose tissue and muscle [29,30]. Insulin binds to receptors on target cells, triggering a cellular response that causes vesicles containing glucose transporters to fuse with the plasma membrane. This increases the number of glucose transporters on the membrane, promoting glucose uptake from the blood [29,39].

2.1.1. Insulin Signaling Pathway, PI3K/Akt Signaling, and Glucose Metabolism

The insulin signaling pathway starts when insulin binds to its specific receptor on the cell surface. This binding is both necessary and sufficient to produce the biological effects of insulin, serving as the main trigger for the next steps in signaling. The insulin receptor is a heterotetrameric membrane glycoprotein, made up of two α- and two β-subunits linked by disulfide bonds [30,31,32,34,39,40,41]. The extracellular α-subunits contain the insulin-binding sites. When insulin binds, these subunits are drawn closer together, causing a conformational change in the receptor. This conformational change activates the tyrosine kinase activity of the β-subunits, promoting ATP binding and receptor autophosphorylation on key tyrosine residues [30,32,33,34,37,39,40,41]. This step initiates downstream signaling [32,33,34,39,40,41]. It is also important to note that signal transduction through the insulin receptor is not limited to its activation at the cell surface since the ligand-receptor complexes move from the microvilli of the cell surface toward clathrin-coated pits and are then internalized into endosomes [40,42]. This process allows the tyrosine kinases to access substrates that might not otherwise be reachable. After the endosomal lumen is acidified by proton pumps, the complex dissociates, which, in hepatocytes, is followed by degradation of insulin by endosomal acidic insulinase [32,42]. This transient process contributes to the acute nature of insulin’s metabolic effects [32,40,42].
Once insulin binds to the α chains of its receptor, the tyrosine kinase activity of the β chains is activated, and each αβ monomer transfers a phosphoryl group from ATP to the hydroxyl group of three critical tyrosine residues located near the carboxyl terminus of the β chain of its partner in the dimer [34]. Autophosphorylation enables phosphorylation of downstream target proteins, including the insulin receptor substrate (IRS) family [32,34,43]. The IRS family is composed of four closely related members (IRS-1, IRS-2, IRS-3, and IRS-4)—associated with, to a greater or lesser extent, β-cell growth and function and insulin sensing—and a more distantly related homolog, Gab-1—more likely involved in hepatic growth factor signaling [32,33,40]. When phosphorylated, insulin receptor substrate-1 (IRS-1) becomes the nucleation point for a complex of proteins that trigger two primary pathways—the mitogen-activated protein kinase (MAPK) pathway and the PI3K/Akt pathway [33,40].
To activate the MAPK pathway, IRS-1 binds growth factor receptor binding protein 2 (Grb2). Grb2 binds to Son-of-sevenless (Sos) and then to Ras, causing GDP-GTP exchange and Ras activation [29,33,34,39,40,43]. Activated Ras recruits c-Raf, which phosphorylates and activates MAPK/Erk kinase (MEK). MEK then phosphorylates extracellular signal-regulated kinase (Erk). Once activated, Erk moves into the nucleus, where it undergoes further phosphorylation and activates transcription factors such as ELK1, which modulates the transcription of about 100 insulin-regulated genes. This sequence ultimately stimulates cell division, protein synthesis, and cell growth [29,32,33,34,39,40,43].
In turn, to activate the PI3K/Akt pathway, PI3K binds IRS-1 through its SH2 domain [32,39,40,43,44]. Upon activation, its catalytic subunit phosphorylates proteins at serine residues or converts the membrane lipid phosphatidylinositol 4,5-bisphosphate (PIP2) to phosphatidylinositol 3,4,5-trisphosphate (PIP3). In turn, PIP3 functions as an intracellular messenger, being responsible for growth stimulation, changes in intracellular trafficking, and activation of phosphoinositide-dependent kinases [29,30,32,33]. In effect, PIP3 binds to the pleckstrin homology domain of Akt and induces its translocation to the plasma membrane [29,30]; there, it also functions as an allosteric regulator of phosphoinositide-dependent protein kinase 1 (PDK1), which phosphorylates Akt and protein kinase A (PKA) [30].
Akt is a serine/threonine kinase that acts on more than 100 target proteins, one of them being glycogen synthase kinase 3 (GSK3) [29,30,40,43,45]. When phosphorylated by Akt, GSK3 loses its catalytic activity and is no longer able to phosphorylate and inactivate glycogen synthase (GS). By preventing GS inactivation, insulin signaling stimulates glycogen synthesis in the liver and muscle [29,30,43,45,46,47]. In addition, Akt promotes the translocation of glucose transporters (of which GLUT4 appears to be the predominant isoform in muscle and adipose tissues in vivo) from internal vesicles to the plasma membrane, stimulating glucose uptake from the blood by a facilitative process [29,30,33,40,43,48,49]. Yet, it is not crucial to this function, since its inhibition does not completely abolish insulin’s effects on glucose transport [50]. By phosphorylating and inhibiting tuberous sclerosis complex 2 (TSC2), Akt is equally responsible for mechanistic target of rapamycin complex 1 (mTORC1)-mediated activation of protein biosynthesis and cell proliferation [30,32,35,36,40,45,46].
Besides these activities, an increasing number of studies have assigned Akt a role in promoting cell survival and regulating cell growth and differentiation [45,46,51]. Since it can phosphorylate proteins that regulate lipid, protein, and glycogen synthesis and cell survival, Akt provides a direct link between insulin receptor signaling and insulin’s biological effects [32,40,45,51].

2.1.2. Molecular Mechanisms of T2D Pathogenesis

The pathogenesis of T2D is complex and multifactorial, primarily involving insulin resistance—a reduced cellular response to insulin—and, in many cases, impaired insulin secretion and disrupted insulin signaling. Although insulin secretion is not completely absent, pancreatic β-cells often fail to compensate adequately for the increased demand, especially as the disease progresses [29,30,31,32]. These alterations are closely associated with obesity and chronic low-grade inflammation, as discussed below. Such factors collectively disrupt key metabolic pathways that regulate glucose homeostasis.
Insulin resistance may arise from abnormalities in insulin receptor function or defects in post-receptor signaling pathways, resulting in a marked reduction in glucose uptake by peripheral tissues, particularly muscle and adipose tissue, which become less responsive to insulin [29,30,31,32]. Moreover, there is impaired suppression of hepatic glucose production and increased lipolysis in adipose tissue. These disturbances contribute to persistent hyperglycemia, which in turn exacerbates β-cell stress and gradually impairs insulin secretion [29,30,31,32].
Regarding impaired insulin signaling, defects can occur at several points along the pathway. For example, serine phosphorylation of IRS-1, often triggered by pro-inflammatory cytokines such as tumor necrosis factor (TNF-α) and activation of stress-related kinases like c-Jun N-terminal kinase (JNK) and IKKβ, impairs insulin signaling, contributing to insulin resistance [52,53]. Reduced expression of GLUT4 in muscle and adipose tissue also limits glucose uptake, while the upregulation of suppressors of cytokine signaling proteins promotes IRS degradation, further weakening insulin action [52,54].
Among the key risk factors, obesity constitutes a major contributor to T2D and plays a pivotal role in the development of insulin resistance through multiple interconnected mechanisms. In obesity, an enlargement of adipose tissue is observed, along with increased levels of free fatty acids and adipokines, namely resistin, that interfere with insulin signaling [52]. Furthermore, dysfunctional adipose tissue exhibits increased infiltration of immune cells, which secrete pro-inflammatory cytokines that exacerbate insulin resistance [52,54]. Elevated circulating levels of cytokines such as TNF-α, IL-6, and IL-1β are commonly observed in individuals with metabolic syndrome and T2D, which impair insulin signaling via activation of inflammatory pathways [52,53,54,55]. These inflammatory molecules also promote insulin resistance in the liver and contribute to T2D-related complications like NAFLD and non-alcoholic steatohepatitis (NASH) [52]. In fact, in addition to disrupted glucose metabolism, lipid metabolism is also impaired, with increased lipolysis in adipose tissue and ectopic lipid accumulation in organs such as the liver and muscle, increasing insulin resistance [52,56]. Additionally, mitochondrial dysfunction is frequently observed in insulin-resistant tissues, impairing fatty acid oxidation and increasing ROS production [57]. This oxidative stress leads to the accumulation of lipotoxic intermediates and activation of stress and inflammatory pathways, further disrupting insulin signaling and reducing β-cell viability [57,58]. As the disease progresses, pancreatic β-cell function worsens due to glucotoxicity, lipotoxicity, and cytokine-induced stress, ultimately resulting in insufficient insulin secretion relative to metabolic demand [52,56].

2.1.3. T2D Treatment

Lifestyle modifications, particularly diet and physical activity, are key components in T2D management. Weight loss often improves insulin sensitivity, as excess adipose tissue contributes to insulin resistance. One such mechanism involves adipokines, including resistin [59]. Exercise is especially beneficial, as increased muscle activity enhances glucose uptake independently of insulin, thereby helping to reduce blood glucose levels and alleviate hyperglycemia. In this context, both aerobic and anaerobic exercises have been shown to significantly improve glycemic control, leading to reductions in % hemoglobin A1c (HbA1c) levels, enhanced lipid profiles, and decreased blood pressure [60,61,62].
Concerning nutritional therapy, it should be focused on a calorie-controlled diet low in simple sugars and rich in dietary fiber to slow carbohydrate absorption and reduce postprandial glucose peaks. Diets rich in whole grains, legumes, fruits, vegetables, lean proteins, and healthy fats have been shown to improve glycemic control, significantly lower HbA1c, and promote weight loss [63,64].
Nevertheless, in addition to lifestyle changes, many patients require oral pharmacological treatment. These oral hypoglycemic agents include insulin secretagogues, such as sulfonylureas, and insulin sensitizers, primarily biguanides like metformin, which remains the first-line treatment for T2D [65,66,67]. Sulfonylureas act by stimulating insulin secretion through closure of ATP-sensitive potassium channels in pancreatic β-cells and are typically used when metformin is contraindicated [67]. However, they are associated with a higher risk of hypoglycemia, weight gain, and progressive loss of efficacy over time [66,67]. In contrast, metformin reduces hepatic glucose production, improves insulin sensitivity, is weight-neutral or may induce minor weight loss, carries a low risk of hypoglycemia, and is generally well-tolerated [67,68].

2.2. Cancer

In healthy, non-transformed cells, glucose serves as a primary source of energy. Under normoxic conditions, glucose metabolism proceeds through a tightly regulated sequence of biochemical pathways, including glycolysis, the tricarboxylic acid (TCA) cycle, and oxidative phosphorylation (OXPHOS) [69]. While normal cells predominantly rely on mitochondrial OXPHOS under aerobic conditions, many cancer cells exhibit a fundamentally different metabolic behavior. In what is known as the Warburg effect, first described by Otto Warburg in the 1920s, tumor cells preferentially convert glucose into lactate via aerobic glycolysis, even when oxygen is available to support mitochondrial respiration [26,70,71]. This phenomenon is now understood as a strategic adaptation rather than a consequence of mitochondrial dysfunction. Despite its lower energetic efficiency, the Warburg effect offers several advantages to rapidly proliferating tumor cells. First, glycolytic intermediates are redirected into anabolic pathways such as the PPP and serine biosynthesis, supplying nucleotides, amino acids, and lipids [70]. Second, lactate production contributes to the acidification of the tumor microenvironment (TME), which facilitates tissue invasion, suppresses immune responses, and promotes angiogenesis [70,72]. Third, glycolysis can proceed at a much faster rate than OXPHOS, allowing for rapid energy generation to meet acute demands [72]. Collectively, the Warburg effect represents a hallmark of cancer metabolism and is closely associated with tumor progression, therapy resistance, and immune evasion.
To meet the elevated metabolic demands imposed by aerobic glycolysis, cancer cells increase their capacity to import glucose. This is primarily achieved through the upregulation of GLUTs [26]. Among these, GLUT1 and GLUT3 are the most frequently overexpressed isoforms across various tumor types [73]. GLUT1, a high-affinity glucose transporter, is commonly upregulated in lung, breast, prostate, and brain cancers [74,75,76]. GLUT3, which has even higher glucose affinity than GLUT1, is particularly elevated in highly proliferative and aggressive tumors, including gastric cancers [77]. The overexpression of these transporters not only correlates with tumor grade and poor prognosis but also reflects the metabolic shift toward increased glycolytic flux. The expression of GLUTs in tumor cells is tightly regulated by oncogenic signaling pathways. For example, the MYC oncogene directly enhances the transcription of GLUT1 and several glycolytic enzymes. Similarly, activation of the PI3K/Akt/mTOR signaling axis promotes GLUT surface localization and stability, while also stimulating glycolytic enzyme activity. A central mediator in this regulation is hypoxia-inducible factor 1-alpha (HIF-1α), a transcription factor stabilized under hypoxic conditions, common in solid tumors. HIF-1α induces the expression of GLUT1 and glycolytic enzymes, facilitating adaptation to oxygen-poor microenvironments [26,78]. Together, these mechanisms ensure that tumor cells maintain continuous and elevated glucose uptake, supporting both energy production and the generation of biosynthetic precursors required for rapid proliferation. The dependency on GLUT-mediated glucose transport has emerged as a potential vulnerability in cancer, with therapeutic strategies under investigation to inhibit GLUT function or expression.
In addition to the upregulation of glucose transporters, cancer cells exhibit profound alterations in the expression and activity of key glycolytic enzymes, further enhancing the glycolytic flux that supports rapid growth and survival under adverse conditions. These enzymatic changes are not essential for maintaining high metabolic throughput but also enable the redirection of intermediates into biosynthetic pathways, contributing to the metabolic flexibility characteristic of cancer cells [70]. One of the most frequently upregulated enzymes in cancer is hexokinase II (HK2), which catalyzes the first committed step of glycolysis [79]. Similarly, phosphofructokinase-1 and the tumor-specific pyruvate kinase M2 isoform (PKM2) play key roles in regulating glycolytic flux and facilitating the diversion of intermediates into biosynthetic pathways. Together, these enzymes are tightly regulated by oncogenic signaling and contribute to tumor progression, proliferation, and survival under metabolic stress [80,81]. Among them, lactate dehydrogenase A (LDHA), which converts pyruvate to lactate, is typically overexpressed in tumor cells and strongly regulated by HIF-1α. Shu et al. demonstrated that elevated LDHA activity supports the Warburg effect by regenerating NAD+ required for sustained glycolysis and by producing lactate, which contributes to microenvironmental acidification. This acidification promotes tumor invasion and immune evasion, reinforcing the malignant phenotype [82].
While the Warburg effect is a hallmark of cancer metabolism, mitochondrial function remains essential. Pyruvate and acetyl-CoA link glycolysis to the TCA cycle, fueling biosynthesis and maintaining redox homeostasis. Although the Warburg effect is the most recognized metabolic alteration in cancer, glycolysis is not the sole source of ATP in most tumors. In fact, solid tumors are composed of metabolically heterogeneous cell populations that may rely on aerobic glycolysis, anaerobic glycolysis, or OXPHOS, often coexisting within the same microenvironment [83,84]. This dual metabolic strategy, combining rapid but inefficient ATP production via glycolysis with mitochondrial oxidative metabolism for biosynthetic precursor generation, confers a metabolic advantage, particularly under fluctuating environmental conditions. Notably, OXPHOS plays a critical role in certain tumor subpopulations, such as cancer stem cells, and in tumors that develop resistance to glycolytic inhibition [85,86]. Several tumor types, including melanoma, ovarian cancer, and acute myeloid leukemia, have been shown to rely heavily on OXPHOS. In these contexts, OXPHOS supports not only ATP production but also redox balance and survival under metabolic stress, making it a promising therapeutic target [87].
The TME is profoundly shaped by these metabolic adaptations. Hypoxia, a common feature of solid tumors, stabilizes HIF-1α, which upregulates GLUT1, HK2, LDHA, and enzymes of the PPP, reinforcing the glycolytic phenotype and promoting survival under low oxygen tension [26,78,88]. Lactate accumulation and export by monocarboxylate transporters (MCTs) acidifies the extracellular environment, facilitating invasion, immune evasion, and angiogenesis [88,89,90]. Moreover, the high glucose consumption by tumor cells creates a competitive metabolic landscape. Immune cells, particularly activated T lymphocytes, require glucose for their effector functions. In glucose-depleted environments, immune cells become metabolically exhausted, which contributes to immune suppression and tumor immune escape [90,91].
These metabolic interactions within the TME underscore the dual role of glucose metabolism in supporting tumor cell fitness and modulating immune responses. Beyond the Warburg effect, mitochondrial metabolism also plays a central role in tumor progression and immune regulation. As such, targeting oxidative phosphorylation (OXPHOS) has emerged as a promising therapeutic approach. In this context, the antidiabetic drug metformin has gained attention for its ability to inhibit mitochondrial complex I, modulate tumor metabolism, and restore immune function.

3. Metformin Targets and Mechanism of Action

3.1. Metformin Targets and Mechanism of Action as an Antidiabetic Drug

Metformin is widely recognized as the first-line therapy for T2D, exerting its glucose-lowering effects through several mechanisms. A central mechanism of its antidiabetic action is the inhibition of hepatic gluconeogenesis, which involves both redox-dependent and mitochondrial pathways. Disruption of the cytosolic redox balance plays a central role in metformin’s mechanism of action. At clinically relevant concentrations (50–100 μM), metformin inhibits mitochondrial glycerol-3-phosphate dehydrogenase (GPD2), thereby altering the cytosolic NADH/NAD+ ratio and selectively suppressing hepatic gluconeogenesis from substrates such as glycerol and lactate [68,92,93]. In fact, the inhibition of GPD2 and the resulting redox imbalance are considered the predominant pathways at therapeutic concentrations [68]. Concerning the mitochondrial pathway, it involves the inhibition of mitochondrial complex I, resulting in decreased cellular ATP levels and thereby limiting the energy supply required for gluconeogenesis [94]. Another important component of metformin’s antidiabetic action is the activation of AMPK in hepatocytes and other metabolically active tissues, which has significant effects on metabolic pathways. AMPK activation is associated with metformin’s partial inhibition of mitochondrial complex I, leading to subtle changes in the cellular energy charge, specifically an increased AMP/ATP ratio, which promotes AMPK activation (Figure 2) [68,95,96]. The metabolic consequences of AMPK activation include: (1) reduced hepatic gluconeogenesis via downregulation of key gluconeogenic enzymes, leading to decreased hepatic glucose output; (2) enhanced fatty acid oxidation and inhibition of lipogenesis; (3) increased glucose uptake in skeletal muscle mediated by GLUT4 translocation to the plasma membrane; and (4) inhibition of mTOR signaling, leading to decreased protein synthesis and reduced cell growth [96,97]. However, it is important to note that complex I inhibition leading to AMPK activation has been observed at supra-pharmacological concentrations of metformin [68].

3.1.1. Metformin and Gut Microbiome

Since patients with metabolic diseases, such as T2D, often exhibit dysbiosis, the gut microbiome has acquired heightened importance in the context of metabolic disease treatment [98]. When administered orally, metformin levels in the gut are 100- to 300-fold higher than in serum, making the gut a major reservoir for the drug in humans. Furthermore, there is accumulating evidence that it also acts through gut-related pathways [99,100,101]. Its hypoglycemic effect is more pronounced when administered orally, and the lower bowel appears to be a major site of action [102]. Metformin’s impact on gut microbiome comes as no surprise when considering the historic use of biguanides as antimicrobials, with folate metabolism impairment possibly being one of the mechanisms underlying such effects [103]. The bacterial reshaping observed after treatment may result from taxon-specific resistance/sensitivity to the bacteriostatic or bactericidal activity of the drug [104].
Although it is difficult to disentangle the effects of pharmacological treatment from those deriving from metabolic disease itself [98,105], different studies have successfully associated metformin use with alterations in the relative abundance of several bacterial strains, most of which are Proteobacteria and Firmicutes [106,107,108]. At the genus level, treatment consistently results in an increase in Escherichia and a decrease in Intestinibacter [103,104,105,109,110]. Microbial changes in individuals initially not taking metformin but later switched to receive the drug correlate with those observed upon longer-term treatment. Additionally, these individuals also experience an increase in Bifidobacterium [106]. Various studies also report treatment-induced abundance of Akkermansia muciniphila. This bacterial strain had already been associated with improved metabolic features in mice and humans [101,103,105,106,111]. Metformin is also associated with an increase in mucin-producing cells, and A. muciniphila occurs predominantly in the mucus layer of the colon, where it enhances mucus secretion, promotes intestinal integrity, and decreases epithelial permeability and inflammation; in fact, it modulates the immune system and delays the onset of type 1 diabetes mellitus (Figure 3) [98,99,107,111,112]. Whole-genome shotgun analysis revealed that Bifidobacterium adolescentis was increased by metformin treatment, and a negative correlation was found between its abundance and HbA1c, suggesting a contribution to the drug’s antidiabetic effect [106]. Interestingly, transferring microbiota from metformin-treated individuals to germ-free mice improves glucose tolerance, indicating causality [103,106].
In vivo and in vitro metagenomics and metatranscriptomics evidenced consistent shifts in lipopolysaccharide (LPS) biosynthesis, incretin, and short-chain fatty acid (SCFA) metabolism; the latter was further confirmed by an increase in acetate, propionate, and butyrate [98,102,104,105,106,107,109,110,113]. In vivo data further support that metformin enhances the growth of SCFA-producing bacterial taxa in general, including Allobaculum, Bacteroides, Bifidobacterium, Blautia, Butyricicoccus, Butyrivibrio, Lactobacillus, Akkermansia, Phascolarctobacterium, Butyricimonas, Coprococcus, and Ruminococcus [98,99,103,111]. In rodents, augmented SCFA production was shown to increase intestinal gluconeogenesis and reduce hepatic gluconeogenesis, appetite, and body weight, thereby promoting beneficial metabolic effects [101,103,104]. A recent integrative perspective suggests the gut-kidney-brain axis is implicated in the concerted regulation of the anorexigenic signals behind metformin action [102]. The interaction between metformin and the gut microbiota modulates epithelial cell activity, which produces metabolites, gut hormones, anorexigenic agents, SCFAs, and secondary bile acids that circulate in the bloodstream, targeting multiple organs and regulating glucose metabolism [102,107]. SCFAs, through G-protein-coupled receptors GPR41/43, lead to the intestinal secretion of glucagon-like peptide 1 (GLP-1) and peptide YY (PYY), regulating glucose homeostasis and food intake (Figure 3) [107]. In addition, butyrate has been shown to inhibit histone deacetylases, thereby having anticancer properties [109].
In parallel, bile acid reabsorption in the distal ileum decreases, whereas the plasma levels of conjugated bile acids, such as glycoursodeoxycholic acid (GUDCA) and tauroursodeoxycholic acid (TUDCA), increase with metformin treatment. Higher bile acid metabolism is negatively associated with HbA1c and serum cholesterol levels, suggesting a potential link between bile acid modulation and improved metabolic health (Figure 3) [98,106,107,109,111].
On the other hand, some microbiome alterations, namely increased abundance of Escherichia and associated virulence factors, also underlie metformin’s side effects, such as bloating, diarrhea, nausea, and intestinal discomfort (Figure 3) [102,104,105].
Interindividual differences in gut microbiota, resulting from heredity, ethnicity, and lifestyle, among others, may account, at least in part, for the interindividual differences in metformin therapeutic success, tolerance or intolerance, and even for the development of side effects or toxicity. This opens potential for personalized medicine based on microbiome profiling [99,102]. It should also be noted that shifts in the microbiome are closely followed by metabolome changes, with metformin causing enrichment in microbial pathways involved in the TCA cycle, butanoate, arginine, and proline metabolism; this has been suggested to have implications in various aspects of the host’s physiology and health, including their cognition, mood and behavior [108,112].

3.1.2. Metformin and Obesity

As previously mentioned, obesity, hyperinsulinemia, and insulin resistance are typical features of T2D. Since obesity and T2D share similarities in terms of insulin-receptor binding, tyrosine kinase signaling, and glucose transport, metformin is efficacious in both conditions [114,115]. Accordingly, insulin resistance is the underlying cause of obesity in T2D diabetic patients, but also in non-diabetic conditions, such as polycystic ovary syndrome. Metformin restores sodium-glucose cotransporter 1 (SGLT1) expression in the upper small intestine, improving glucose sensing [111]. By increasing insulin sensitivity, it also affects body weight, being useful in obesity treatment and prevention [114,115,116]. In line with this, metformin treatment has been shown to lead to lower caloric intake, as well as moderate weight loss in insulin-sensitive and insulin-resistant obese patients, with the percentage of weight loss being more pronounced in the latter [100,114,115,117].
Improved insulin sensitivity reduces delayed postprandial peaks of the hormone, which frequently cause hypoglycemia and consequent carbohydrate and compensatory food cravings. Furthermore, as stated earlier, metformin improves insulin-stimulated glucose distribution to skeletal muscle, inhibits hepatic gluconeogenesis and glucose output, and reduces glucose intestinal absorption [100,114,116,118]. Even when skeletal muscle and liver cells are resistant to insulin, adipose tissue remains relatively sensitive to the hormone, which justifies the accumulation of abdominal fat. In this context, as metformin restores peripheral insulin sensitivity, it prevents pancreatic hypersecretion of the hormone, attenuating insulin signaling in adipose tissue and accumulation of abdominal fat [114]. Collectively, these effects contribute to lower glucose availability and consequent fat storage in adipose tissue, explaining why metformin diminishes adipose tissue, but not lean body mass [114,117,118].
The fact that metformin treatment also significantly reduces body weight in non-diabetic patients suggests that additional mechanisms, besides decreasing insulin resistance, are involved in metformin-mediated weight loss. In this sense, metformin has been shown to suppress appetite in a dose-dependent manner, possibly because it acts by decreasing leptin levels, thereby improving leptin resistance [114,115,117]. The rise in GLP-1 and PYY levels under treatment with the biguanide, as well as fewer hyperinsulinemia events, are possible factors accountable for diminished hunger and increased satiety [100,114,115]. Lactate accumulation, due to the inhibition of the electron transport chain and increased anaerobic glycolysis, leads to mild metabolic acidosis, which is also appetite-suppressing [100]. On the other hand, metformin-induced weight loss seems to be independent of changes in energy expenditure [100]. There is also evidence that metformin modulates the activity of the central nervous system, namely the parahippocampal gyrus, ventromedial prefrontal cortex, fusiform gyrus, and hypothalamus. These areas are involved in semantic memory, reward formation, and orexigenic pathways; such interactions may then underlie altered food-reward relationships and decreased hunger [100,116].
As explored in the previous section, metformin’s interaction with the gut microbiome may also promote appetite reduction and weight loss. Increased abundance of SCFA-producing species decreases hepatic gluconeogenesis, limits fatty acid release from adipocytes, and suppresses appetite via the incretin system [100,101].
Currently, the use of metformin with the sole purpose of weight reduction remains off-label; inconsistencies in weight loss, side effects, and the emergence of newer, more effective drugs have prevented its approval as a first-line treatment for obesity [117]. Still, it represents an option for patients at high risk of metabolic complications and who do not tolerate or comply with other interventions, either pharmacological or non-pharmacological [100]. It is also being investigated as a potential tool for counteracting sarcopenic obesity, a condition characterized by increased fat mass and decreased lean mass that is typical of aging. Interestingly, ongoing studies focus on this biguanide as a strategy to mitigate obesity’s sequelae, such as NAFLD, obstructive sleep apnea, metabolic bone disease, and osteoarthritis [100,116].

3.2. Metformin Targets and Mechanism of Action as an Anticancer Drug

Although the Warburg effect or aerobic glycolysis is widely recognized as a cancer hallmark, it represents only one aspect of the metabolic reprogramming observed in malignant cells. In fact, and in opposition to Warburg’s first postulate that cancer cells have defective mitochondria, it is now well established that cancer cells can also rely on mitochondrial metabolism to support their growth and survival [119,120,121,122]. It has already been reported that OXPHOS contributes more to total ATP production than aerobic glycolysis in many tumors [123]. Tumors are very heterogeneous and both pathways can contribute to ATP production, providing cells with metabolic flexibility to adapt and proliferate in response to environmental conditions [121]. Although many cancers primarily rely on aerobic glycolysis for proliferation, the contribution of this pathway to total ATP production can vary widely, ranging from as little as 1% to as much as 64% [123]. Recent studies have demonstrated that OXPHOS can be upregulated in different types of cancer, like leukemias, lymphomas, pancreatic ductal adenocarcinoma, and this can happen even in the context of elevated glycolytic activity, highlighting the ability of cancer cells to simultaneously activate diverse metabolic pathways for energy production [87]. Several agents that inhibit the glycolytic pathway—such as 3-bromopyruvate, dichloroacetate (DCA), and 2-deoxyglucose (2-DG)—have shown promising results as anticancer agents [124]. Nevertheless, others like phenformin, which targets mitochondrial complex I and thereby inhibits oxidative phosphorylation, have also been effective in reducing cell viability, ATP production, and resistance to conventional anticancer drugs [125]. Consequently, in predominantly glycolytic tumors, targeting the Warburg effect may offer greater therapeutic benefit, whereas in cancers with a more oxidative metabolic profile, OXPHOS inhibitors such as metformin or phenformin may provide a more effective treatment strategy. In an in vitro study using glioma—a highly aggressive cancer known for its marked metabolic plasticity—as a model, it was observed that the glycolytic U251 cell line was more sensitive to the antiglycolytic agents DCA and 2-DG. However, treatment with phenformin did not significantly reduce intracellular ATP levels in these cells. In contrast, the more oxidative glioma cell line SW1088 showed reduced ATP levels and cell proliferation in response to all tested compounds, indicating a higher degree of metabolic plasticity [125]. Nevertheless, a stronger impact of OXPHOS inhibition on ATP production was observed, in alignment with the predominantly oxidative metabolism of this cell line. In both cell lines, all the tested agents increased the sensitivity to temozolomide, the gold standard treatment for glioma, suggesting their potential to enhance therapeutic efficacy when used in combination strategies [125]. However, despite the previous use of phenformin in lowering blood glucose levels, it has been withdrawn from clinical use in most countries due to its association with severe, and sometimes fatal, lactic acidosis [126]. Although metformin has been reported in some studies to be less potent in inhibiting cancer cell growth compared to phenformin [127], its well-established safety profile and extensive clinical experience have made it a standard treatment in medical practice. Therefore, several studies investigating metformin’s potential as an anticancer agent are currently underway, and encouraging findings indicate that diabetic patients treated with metformin exhibit improved survival rates in lung, colorectal, and prostate cancers, along with a decreased risk of developing pancreatic, breast, colorectal, and liver cancers [128]. There are some well-established metformin targets, but its role in inhibiting the mitochondrial respiratory chain complex I, therefore impairing OXPHOS, causing bioenergetic stress, and altering metabolism in different cancer types, has been consistently reported (Table 1). Metformin is transported into the cell via an organic cation transporter (OCT) and accumulates primarily in the mitochondria, where it exerts its action by inhibiting complex I of the electron transport chain (Figure 2). Due to this inhibition, NADH accumulates and mitochondrial ATP production decreases, leading to an increase in the AMP/ATP ratio. This shift activates AMPK, a key energy sensor that helps restore energy balance by turning on catabolic energy-producing pathways and turning off anabolic energy-consuming ones [129]. Through these mechanisms, metformin influences cellular metabolism and may contribute to its therapeutic effects, not only in its primary role improving insulin sensitivity but also showing potential anticancer activity.
As said before, the Warburg effect is not the only metabolic strategy in cancer, and this model does not capture the full metabolic complexity, heterogeneity and plasticity of tumors. Moreover, emerging evidence supports the concept of the reverse Warburg effect, in which cancer-associated fibroblasts or other stromal cells perform aerobic glycolysis, exporting lactate that can be used by nearby cancer cells to fuel their oxidative metabolism [147]. Consequently, mitochondrial metabolism is not only preserved but often enhanced in several types of cancer, providing a metabolic plasticity that supports tumor growth, survival, and therapy resistance beyond the traditional Warburg paradigm. In this context, both antiglycolytic agents and electron transport chain inhibitors, such as metformin, may play a valuable role as anticancer therapies. Although the exact mechanism of action of metformin is not yet fully understood, one of the primary targets is believed to be the already mentioned mitochondrial complex I of the electron transport chain, leading to a reduction in ATP production [148,149]. Inhibition of complex I also increases the leakage of electrons and the production of ROS, contributing to oxidative stress, which further disrupts mitochondrial function [150]. Similar effects have been observed with phenformin, which, despite its discontinued clinical use due to the risk of lactic acidosis at high blood glucose levels, is a more potent inhibitor of mitochondrial complex I compared to metformin [87]. As a result of impaired mitochondrial respiration, cells treated with metformin (or phenformin) increase their dependence on glycolysis for ATP generation. This metabolic reprogramming, unlike the Warburg effect, results in a moderate increase in lactate production, especially under conditions of high glucose availability or hypoxia [151,152,153]. Metformin can be particularly effective in a context of low glucose availability due to high glucose uptake by cancer cells. In these conditions, metformin enhances the metabolic stress, inhibiting mitochondrial ATP production. At the same time, the limited extracellular glucose restricts ATP generation through glycolysis. The combined impairment of both major energy-producing pathways leads to a critical energy crisis within the cell, ultimately triggering cell death [154].
Besides the respiratory chain (mostly complex I, though some studies suggest it may also inhibit complexes II and IV [155]), it was reported that metformin also inhibits HIF-1 activity, [156]. HIF-1 plays a central role in activating the transcription of GLUTs, glycolytic enzymes, MCTs, or vascular endothelial growth factor (VEGF), promoting angiogenesis, metabolic reprogramming, and cell survival under hypoxic conditions—processes that support tumor growth and progression [157,158]. By interfering with HIF-1 activity, metformin may also disrupt these adaptive responses, contributing to its potential anticancer effects. Furthermore, metformin has also been shown to inhibit glutaminase (GLS), an enzyme often overexpressed in cancer cells, essential to feed the TCA with α-ketoglutarate, supporting energy production and biosynthesis. Furthermore, blocking GLS activity, metformin may reduce ammonia production and limit ammonia-induced autophagy—a survival mechanism used by tumor cells under stress [159]. These effects on glutamine metabolism could contribute to metformin’s antitumor action by disrupting a key energy and biosynthesis source for cancer cells. Besides, the bioenergetic stress induced by metformin can make tumor cells more vulnerable, thereby increasing the efficacy of conventional anticancer treatments.

3.2.1. Metformin-Induced Alterations in ROS Metabolism in Cancer Cells

Metformin has shown the ability to interfere with key cellular pathways, potentially inducing cell death, particularly in tumor cells. In addition to decreasing hepatic gluconeogenesis, metformin also modulates ROS levels, contributing to oxidative stress and impacting tumor cell viability [160]. Furthermore, metformin promotes autophagy via AMPK/mTOR signaling pathways, which suppresses the proliferation, migration, and invasion of cancer cells [161]. These diverse effects enhance metformin’s potential for cancer prevention and treatment, in addition to its advantages in managing complications related to diabetes and processes associated with aging [162].
ROS are acknowledged for their essential function in maintaining homeostasis, as they regulate oxidative stress, intracellular signaling, and cell survival. In the context of cancer, ROS exhibit a dual function, facilitating both pro-tumorigenic and anti-tumorigenic signaling [163]. Cancer cells demonstrate increased levels of ROS because of metabolic changes and heightened proliferation rates. To preserve redox homeostasis and avoid apoptosis, cancer cells enhance their antioxidant mechanisms [164,165]. Tumor cells are recognized for exhibiting increased basal levels of ROS, which arise from metabolic changes, rapid proliferation rates, and enhanced survival abilities [166,167]. Nevertheless, metformin has demonstrated the potential to mitigate this characteristic, which contributes to the tumor cells’ heightened resistance to apoptosis and improved proliferation capacity [167,168]. By interfering with the mitochondrial complex I, metformin leads to a decrease in ATP levels, which subsequently raises the AMP/ATP ratio, disrupts electron transport, and ultimately increases the production of ROS, such as superoxide anion and hydrogen peroxide, resulting in harmful oxidative stress [169,170]. The accumulation of ROS has various cellular consequences, particularly causing damage to DNA, lipids, and essential proteins, and can also initiate cell cycle arrest, senescence, and apoptosis. It activates stress response pathways including p53, MAPK, or JNK, and may enhance the sensitivity of tumor cells to chemotherapy and radiotherapy, thereby intensifying the effects of genotoxic agents [171,172]. Additionally, metformin inhibits the expression of NFE2L1, which further worsens ROS accumulation [173]. There is evidence that metformin can selectively modulate ROS levels in tumor cells. In one study using AsPC-1 pancreatic cancer cells, metformin increased ROS, triggering apoptosis, autophagy, and necroptosis, with minimal effects on normal cells. These effects involved the AMPK–FOXO3a–MnSOD pathway and were enhanced by co-treatment with apigenin [150]. In contrast, another study using MIA PaCa and Panc1 cells under physiological glucose conditions found that metformin reduced ROS by increasing MnSOD and decreasing NOX2/NOX4 expression, also leading to reduced tumor cell viability [174]. In hepatocellular carcinoma (HCC), metformin-loaded hyaluronic acid-derived carbon dots enhanced intracellular ROS generation by inhibiting glutaminase-1 and GLUT1, leading to AMPK activation, Akt inhibition, and selective induction of apoptosis [175]. As stated in Table 1, in cases of ovarian cancer, metformin promotes apoptosis and ferroptosis during energy stress by directly affecting the NDUFB4 subunit of mitochondrial complex I. This interaction results in mitochondrial dysfunction and a rise in ROS production [130]. In non-small cell lung cancer (NSCLC) cells, the cytotoxic effects of metformin and mitomet were enhanced through both apoptosis and ferroptosis, likely due to an increase in mitochondrial reactive oxygen species (mROS) production and a reduction in glutathione (GSH) levels. These effects were particularly evident in A549 cells deficient in LKB1. Furthermore, in vivo studies demonstrated that both compounds significantly reduced tumor burden, with mitomet showing a potency that was 100 times greater than that of metformin [133]. Overall, these findings suggest that metformin selectively disrupts redox homeostasis in tumor cells, with effects depending on the metabolic context and tumor type.
Unlike rotenone, metformin inhibits mitochondrial complex I without increasing ROS production in certain contexts, such as colorectal cancer. This results in reduced oxidative stress and the activation of HIF-1α triggered by hypoxia, highlighting the context-dependent effect of metformin on ROS modulation [146]. Moreover, mitochondria are essential in the regulation of ferroptosis by affecting iron metabolism, lipid peroxidation, and the production of ROS in ovarian cancer cells [130]. Furthermore, metformin has been associated with a decrease in the expression of antioxidant enzymes, such as glutathione peroxidase (GPx) and superoxide dismutase (SOD), which exacerbates the accumulation of ROS [176,177]. The exact mechanism through which metformin interacts with complex I of the mitochondrial respiratory chain is not fully understood [152].
The effect of metformin on the metabolism of ROS varies according to the tumor type, the intrinsic redox state of the cells, and the tumor microenvironment. However, the increase in ROS induced by metformin has been demonstrated to be selective for tumor cells, sparing normal cells, which have superior redox regulation [150,174]. Thus, the modulation of ROS metabolism by metformin constitutes one of the central mechanisms of its anticancer action, especially when combined with therapies that exploit oxidative stress as a pathway for cancer cell death.

3.2.2. The Role of PI3K/AKT/mTOR Pathway in Mediating the Anticancer Effects of Metformin

The effect of metformin on mitochondrial respiratory chain complex I is also associated with activation of the AMPK pathway, a central kinase in detecting the cellular energy status [178,179]. In tumor cells, the activation of AMPK and inhibition of the mTOR pathway, which results from AMPK activation, have been linked to metformin’s antitumor potential, although the exact mechanism is not yet fully understood [132,143] (Figure 4).
While AMPK activation has been linked to metformin’s antitumor potential, recent studies suggest that AMPK may actually attenuate metformin’s toxicity by enhancing glycolysis [180]. Metformin inhibits hepatic mTORC1 signaling through dose-dependent mechanisms involving AMPK and the TSC complex [181]. It also suppresses mitochondrial-dependent biosynthesis, reducing cancer cell proliferation [182]. The AMPK pathway is one of the main sensors of cellular energy status and is involved in regulating multiple metabolic processes, including glucose homeostasis, lipid metabolism, autophagy, and cell proliferation. Under conditions of energy stress, such as low ATP levels, AMPK is activated to restore metabolic balance [183,184]. It also blocks lipogenesis by inhibiting acetyl-CoA carboxylase (ACC), reduces nucleotide and cholesterol biosynthesis, limiting precursors for highly proliferative cells, and promotes autophagy as a cellular recycling mechanism and, in some contexts, induces cell death [185].
Metformin activates this pathway indirectly through inhibition of mitochondrial complex I. This leads to decreased ATP production and increased levels of AMP and ADP. This imbalance in the AMP/ATP ratio promotes phosphorylation and activation of AMPK by the LKB1 kinase (liver kinase B1), particularly at the Thr172 residue of the catalytic subunit [136,184]. The AMPK has a multifaceted role in cancer, functioning as both a tumor suppressor and a promoter, based on the specific context [186]. AMPK inhibits mTORC1, which leads to a decrease in cell proliferation and tumor growth [187]. Conversely, mTORC1 can directly inhibit AMPK, thereby encouraging cell proliferation during periods of nutrient stress [188]. Before the onset of cancer, AMPK serves as a tumor suppressor; however, once cancer is established, it may transition to promoting tumor survival [189]. Metformin plays a key role in modulating this pathway. For instance in HCC, metformin binds to the mitochondrial voltage-dependent anion channel 1 (VDAC1), disrupting the IP3R-GRP75-VDAC1 complex and leading to AMPK activation, which contributes to autophagy-mediated cell death and antiproliferative effects [136]. In breast cancer cells, metformin induces G0/G1 cell cycle arrest via AMPK activation and cyclin D1 downregulation. However, full growth inhibition also requires the presence of CDK inhibitors, highlighting the role of AMPK in modulating cell cycle machinery [138]. In pancreatic cancer, metformin activates AMPK at Thr172 and inhibits insulin/IGF-I signaling, promoting apoptosis under normoglycaemic conditions. Hyperglycaemia, however, impairs this effect by increasing Ser485 phosphorylation, which antagonizes Thr172 activation [142].
Besides metabolic effects associated with AMPK activation, it may also lead to reduced expression of cyclins and growth factors, causing cell cycle arrest in G1 phase, reducing invasion and metastasis, possibly by inhibiting epithelial–mesenchymal transition (EMT), and modifying the inflammatory and immune profile of the tumor microenvironment [190,191]. This effect is especially relevant in HNSCC, where metformin impairs tumor growth by AMPK-mediated mTOR inhibition. Notably, when HNSCC cells express the metformin-insensitive mitochondrial protein NDI1, metformin fails to activate AMPK or reduce tumor burden, suggesting its action depends on complex I inhibition and metabolic reprogramming [132]. Moreover, as previously mentioned, AMPK activation can inhibit aerobic glycolysis (the Warburg effect), redirecting cellular metabolism and creating an unfavorable environment for tumor growth. Many studies have linked AMPK activation by metformin directly to reduced tumor cell viability and clonogenicity, as well as increased sensitivity to chemotherapy [192,193].
In summary, by activating AMPK, metformin promotes an intracellular environment that is hostile to tumor survival, thereby limiting the growth, proliferation, and adaptive capacity of cancer cells [193,194]. The PI3K/Akt/mTOR pathway is a major pro-survival and pro-proliferative signaling route in tumor cells, frequently hyperactivated by oncogene mutations or loss of tumor suppressor genes such as PTEN [195,196]. This pathway regulates fundamental processes like growth, metabolism, cell survival, and therapy resistance, and is one of the most frequently deregulated pathways in cancer. Aberrant activation contributes to malignant transformation by promoting uncontrolled proliferation, apoptosis evasion, angiogenesis, and metastasis [196]. It stimulates increased protein synthesis, glycolysis, and lipid biosynthesis, inhibits apoptosis, and activates mTORC1 [197]. Metformin interferes with this pathway through two main mechanisms. First, it can indirectly inhibit it via AMPK activation. By activating AMPK, metformin inhibits the mTORC1 complex, a central regulator of cell growth. This inhibition primarily occurs through phosphorylation and activation of TSC2 and the protein raptor, both direct inhibitors of mTORC1 [181,194,198]. Consequently, protein synthesis is interrupted by inhibiting phosphorylation of key proteins like 4E-BP1 and S6K1, essential for protein translation [199,200]. This blockade reduces cell cycle progression and compromises tumor cell survival. This mechanism is especially relevant because it acts downstream of the PI3K/Akt pathway, which is often activated in many cancer types. Thus, metformin may exert antitumor effects even in the presence of mutations that activate Akt or PI3K, bypassing excessive activation and inhibiting tumor growth [198,201].
On the other hand, metformin may have direct effects on PI3K and Akt. Some studies have shown that metformin can also reduce Akt phosphorylation at key residues such as Ser473 and Thr308, interfere with PI3K binding to membrane receptors, reducing activation of pro-survival signaling, increase tumor cell sensitivity to apoptosis, and to therapies targeting the PI3K/Akt pathway [202,203,204]. However, these direct effects are highly dependent on tumor type, presence of mutations in PTEN, LKB1, or p53, and the metabolic conditions of the cell [205]. For example, in metformin-sensitized cells such as HL60 and MOLM14, metformin treatment increased and sustained phosphorylation of Akt at Ser473, while in PTEN-null cells like U937, which exhibit high basal glycolysis and Akt phosphorylation, no further increase in Akt phosphorylation was observed upon metformin treatment. These metabolic differences contribute to these cells’ resistance to metformin-induced cell death, likely due to their reduced mitochondrial dependence [144].
Metformin’s impact also depends on nutrient conditions. Under low glucose, it is more cytotoxic by inhibiting Akt/mTOR and reducing ATP, without significantly changing AMPK phosphorylation, indicating AMPK-independent effects. Conversely, high glucose maintains glycolytic compensation, reducing metformin’s efficacy [206]. For instance, in breast cancer cells, metformin induces apoptosis mainly under nutrient-poor conditions through AMPK activation and downregulation of PKM2. High glucose and amino acid levels abolish this effect, suggesting that nutrient availability critically modulates metformin’s anticancer activity [207].
Metformin’s interference with the PI3K/AKT/mTOR pathway reinforces its potential as an adjuvant agent in oncology, especially in combination with specific pathway inhibitors (e.g., rapamycin, everolimus), in tumors resistant to conventional chemotherapy, and in cancer models metabolically dependent on PI3K/Akt signaling [198]. Figure 4 summarizes the main metformin targets and the drug’s effect on oncogenic characteristics.

3.2.3. Metformin-Based Strategies to Enhance Antitumor Therapy

To overcome the pharmacokinetic limitations of metformin and maximize its therapeutic potential in oncology, a range of innovative strategies has been developed. These include nanoparticle-based delivery systems and combination regimens with chemotherapeutic agents, immunotherapies, and molecularly targeted drugs [208]. These approaches aim to improve drug bioavailability, enable tumor-specific delivery, and exploit synergistic mechanisms of action. This section reviews the most promising advances in metformin-based antitumor strategies, with a particular focus on nanotechnology-driven delivery systems and combination therapies, highlighting their mechanistic rationale and translational relevance.

3.2.4. Metformin Encapsulation and Nanoparticle Delivery

Despite its well-documented antitumor properties, the clinical application of metformin in oncology has been hindered by its low bioavailability, rapid systemic clearance, and limited tumor tissue penetration [209,210]. To address these challenges, nanotechnology-based delivery systems have emerged as a powerful tool to enhance the pharmacokinetic and pharmacodynamic profiles of metformin. Encapsulation into nanocarriers, such as liposomes, polymeric nanoparticles, solid lipid nanoparticles, and functionalized hybrid systems, has demonstrated improved drug stability, controlled release, and preferential accumulation in tumor tissues via the enhanced permeability and retention effect [209,210,211,212]. Furthermore, surface functionalization with ligands such as folate or hyaluronic acid enables active targeting of tumor-specific receptors, thereby increasing therapeutic efficacy while minimizing off-target toxicity [211,213,214].
Table 2 summarizes the most relevant nanoparticle-based systems developed for metformin delivery in cancer therapy, highlighting their composition, target indications, therapeutic goals, and supporting studies.

3.2.5. Metformin in Combination Therapies

The repositioning of metformin as an adjuvant in cancer therapy has garnered significant attention due to its ability to modulate key oncogenic pathways, enhance treatment sensitivity, and counteract resistance mechanisms. When combined with chemotherapeutic agents, metformin has demonstrated synergistic effects that extend beyond its intrinsic antitumor activity [221,222,223]. These benefits are largely attributed to its capacity to activate AMPK, inhibit the mTOR pathway, reduce insulin/IGF-1 signaling, and alter the TME. Moreover, metformin may sensitize cancer cells to cytotoxic agents by inducing metabolic stress, impairing DNA repair, and modulating immune responses [221,224,225,226]. This section reviews the most relevant preclinical and clinical studies exploring metformin-based combination regimens, highlighting their mechanistic underpinnings and therapeutic implications across different cancer types.
The ability of metformin to enhance the cytotoxicity of chemotherapeutic agents has been demonstrated in multiple cancer models. These effects are often mediated by metabolic reprogramming, increased oxidative stress, and inhibition of survival pathways. Table 3 provides an overview of key metformin-based combinations and their mechanistic implications.

4. Clinical Trials

Metformin is a well-established antidiabetic drug widely used as a first-line treatment for T2D. Its primary effect is the reduction in blood glucose levels, achieved mainly through the inhibition of hepatic gluconeogenesis and improvement of insulin sensitivity in peripheral tissues, particularly skeletal muscle. The use of already clinically approved and well-tolerated drugs in cancer treatment—an approach known as drug repurposing—is gaining increasing attention. This strategy offers several advantages, including reduced development time, cost effectiveness, and established safety profiles [250]. In particular, interest has grown around repurposing metabolic drugs, such as metformin, due to their potential to target cancer-related pathways beyond their original indications. The prolonged use of metformin has been associated with a decrease in tumor incidence and a lower rate of cancer-related mortality [251]. Based on these premises, several clinical trials have been conducted using metformin to assess its efficacy and safety profile in the combinatorial therapy of various types of cancer. The search of clinical trials clinicaltrials.gov, using “cancer” and “metformin” as keywords, gave 484 registered studies, and 54 at the EU Clinical Trials Register (https://www.clinicaltrialsregister.eu) (accessed on 25 July 2025). Table S1 of Supplementary Materials summarizes some selected clinical trials in different cancer types and with different drugs, highlighting cancer types, treatment protocols, outcomes, and safety profiles, based on the clinical trials that presented outcomes in their description and with published results [252,253,254,255,256,257,258,259,260,261,262,263,264,265,266]. Although numerous large randomized clinical trials across different tumor types have been conducted, the overall results have been disappointing, concerning the use of metformin in combination with anticancer therapies. Nevertheless, a few studies have yielded promising findings, like the NCT05351021 clinical trial. Furthermore, for many clinical trials, several limitations were found. Detailed trial results with clinical efficacy data are not yet publicly disclosed and, in many cases, the lack of adequate sample sizes prevents conclusive interpretation, and prospective studies are thus needed.
Breast cancer is among the cancer types with the highest number of clinical trials exploring the combination of metformin with chemotherapy, targeted therapies, or immunotherapies, with several phase II and III studies completed or currently ongoing. Nevertheless, for many cases, metformin, when added to standard breast cancer therapy, failed to show significant benefits in terms of disease–free survival or overall survival, compared to placebo. Still, different reports demonstrated that metformin improves metabolic factors that can be linked to breast cancer, namely in body weight, glucose, insulin, leptin or C-reactive protein [267]. Furthermore, several clinical and preclinical studies have demonstrated that metformin decreases ATP production, which is essential for the function of efflux pumps, lowers blood insulin levels, and can reverse the EMT [268]. Additionally, besides its potential use in combined anticancer therapies, there are clinical assays reporting the reduction in the risk of growing cancer after using metformin [269]. The lack of promising results in clinical trials may also be linked to glucose levels. Indeed, some studies have reported that the effects of metformin are particularly pronounced under conditions of glucose deprivation—a common feature of the tumor microenvironment due to the Warburg effect–, where its use has been associated with increased apoptosis. In contrast, under high-glucose conditions, this pro-apoptotic effect is reduced [268]. These potential benefits of metformin are further supported by in vivo assays. In different breast cancer mouse models, the combination of metformin with other agents, such as propranolol, curcumin, or glycolytic inhibitors, has shown promising therapeutic effects without associated toxicity [270,271,272]. In other clinical trials (NCT01929811, NCT00897884, NCT02472353), although metformin did not enhance pathological response or survival outcomes, it showed a beneficial effect in preventing increases in cholesterol levels during treatment (NCT01929811), decreasing insulin and insulin receptors (NCT00897884), and reducing the tendency of significant change in left ventricle ejection fraction (NCT02472353). Additionally, a meta-analysis review of clinical trials indicated that metformin proliferation markers like Ki-67 decreased in three trials [273].
In other types of cancer, metformin has also shown encouraging signs in the fight against cancer. For instance, clinical trials in endometrial cancer patients reported significant reductions in proliferation markers following metformin monotherapy [274]. In ovarian cancer, patients treated with metformin demonstrated significantly longer overall survival compared to those who did not receive the drug [275]. The combination of metformin and irinotecan in patients with refractory colorectal cancer was positive concerning disease control, but further studies are necessary to validate these findings [276]. In oral cancer, a phase II trial of metformin provided evidence that metformin treatment leads to promising histological improvements and modulation of the mTOR signaling pathway [277], whereas in pancreatic cancer, metformin use decreased the activity of pro-tumoral M2 macrophages while enhancing the recruitment and function of tumor-resolving dendritic cells (DCs), thereby promoting antitumor immune responses and survival outcomes [278].

5. Conclusions

Drug repurposing refers to the strategy of using existing, approved medications for new therapeutic indications beyond their original clinical use. This approach can significantly reduce development time, costs, and adverse effects, as the safety profiles of these drugs are already well-established [279]. In the global emergency of the COVID-19 pandemic, drug repurposing emerged as one of the key strategies employed to rapidly identify effective treatments, but this approach is gaining particular importance in other diseases, particularly in cancer [280]. Recently, Khzem et al. described several classes of drugs being investigated for such effects, namely anti-platelet, anti-helminthic, anti-viral, cardiovascular, antibiotic, anti-malarial, antipsychotic, non-steroidal anti-inflammatory drug (NSAID), antirheumatic, anti-epileptic, anesthetic, or antidiabetic medications [281]. From this panacea of drugs, numerous preclinical and clinical studies have reported positive effects of anti-diabetic medication in the context of cancer treatment, namely the ones mentioned in this review. Diabetes and cancer share several risk factors, like chronic inflammation and obesity, as well as metabolic features, including altered glucose metabolism, all of which can influence tumor development and progression. These shared mechanisms have prompted growing interest in exploring anti-diabetic drugs, such as metformin, with the potential to improve cancer therapeutic outcomes. Metformin targets critical signaling pathways that regulate cell growth and survival, namely the AMPK/mTOR axis, involved in key biological functions such as metabolism, energy homeostasis, and cellular proliferation. All these processes are compromised in both diseases, making this signaling pathway a promising therapeutic target. Metformin exerts other beneficial effects on diabetes that can also be relevant to cancer, including modulation of the gut microbiome and the regulation of obesity, improving metabolic health, and reducing inflammation. Taking these premises, metformin is increasingly being explored as an adjuvant in cancer treatment, often in combination with different therapies, to enhance treatment effectiveness and overcome drug resistance. To improve its bioavailability and targeted delivery, novel approaches such as metformin encapsulation and nanoparticle-based delivery systems are being developed, aiming to maximize its anticancer potential while minimizing side effects. However, although preclinical and in vivo studies have demonstrated promising anticancer effects of metformin, the translation of these findings into clinical settings has been less consistent. Several clinical trials have yielded disappointing or inconclusive outcomes, particularly regarding survival benefits. These discrepancies may be due to differences in patient populations, tumor types, dosing regimens, or metabolic conditions [282]. As such, further research is needed to better define the context in which metformin can be most effective in cancer therapy.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/app152111576/s1, Table S1: Summary of the clinical trials regarding metformin use in anticancer therapeutic strategies and their main findings.

Author Contributions

Conceptualization, study design, writing—original draft preparation, writing—review and editing, F.B., A.C., J.B., J.F. and O.Q. Supervision, O.Q. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by CESPU through the project FLAV4CANCER-GI2-CESPU-2025.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study.

Acknowledgments

F.B. gratefully acknowledges CESPU for a Ph.D. Grant BD/DCB/CESPU/01/2024.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Rohm, T.V.; Meier, D.T.; Olefsky, J.M.; Donath, M.Y. Inflammation in obesity, diabetes, and related disorders. Immunity 2022, 55, 31–55. [Google Scholar] [CrossRef]
  2. Expert Panel on the Identification, Evaluation, and Treatment of Overweight and Obesity in Adults (U.S.). Clinical Guidelines on the Identification, Evaluation, and Treatment of Overweight and Obesity in Adults; National Institutes of Health, National Heart, Lung, and Blood Institute: Bethesda, MD, USA, 1998; Volume 97, pp. 20–21+24–25+27–37. [Google Scholar]
  3. Tsalamandris, S.; Antonopoulos, A.S.; Oikonomou, E.; Papamikroulis, G.-A.; Vogiatzi, G.; Papaioannou, S.; Deftereos, S.; Tousoulis, D. The Role of Inflammation in Diabetes: Current Concepts and Future Perspectives. Eur. Cardiol. Rev. 2019, 14, 50–59. [Google Scholar] [CrossRef]
  4. Shoelson, S.E.; Lee, J.; Goldfine, A.B. Inflammation and insulin resistance. J. Clin. Investig. 2006, 116, 1793–1801. [Google Scholar] [CrossRef] [PubMed]
  5. Sharif, S.; Cramer, M.J.; de Borst, G.J.; Visseren, F.L.J.; Westerink, J.; The SMART Study Group; van Petersen, R.; Dinther, B.G.F.; Algra, A.; van der Graaf, Y.; et al. Low-grade inflammation as a risk factor for cardiovascular events and all-cause mortality in patients with type 2 diabetes. Cardiovasc. Diabetol. 2021, 20, 220. [Google Scholar] [CrossRef]
  6. Bonaccio, M.; Di Castelnuovo, A.; Pounis, G.; De Curtis, A.; Costanzo, S.; Persichillo, M.; Cerletti, C.; Donati, M.B.; de Gaetano, G.; Iacoviello, L. A score of low-grade inflammation and risk of mortality: Prospective findings from the Moli-sani study. Haematologica 2016, 101, 1434–1441. [Google Scholar] [CrossRef] [PubMed]
  7. Peng, Y.; Wang, P.; Gong, J.; Liu, F.; Qiao, Y.; Si, C.; Wang, X.; Zhou, H.; Song, F. Association between the Finnish Diabetes Risk Score and cancer in middle-aged and older adults: Involvement of inflammation. Metabolism 2023, 144, 155586. [Google Scholar] [CrossRef]
  8. Coughlin, S.S.; Calle, E.E.; Teras, L.R.; Petrelli, J.; Thun, M.J. Diabetes mellitus as a predictor of cancer mortality in a large cohort of US adults. Am. J. Epidemiol. 2004, 159, 1160–1167. [Google Scholar] [CrossRef]
  9. Giovannucci, E.; Harlan, D.M.; Archer, M.C.; Bergenstal, R.M.; Gapstur, S.M.; Habel, L.A.; Pollak, M.; Regensteiner, J.G.; Yee, D. Diabetes and cancer: A consensus report. Diabetes Care 2010, 33, 1674–1685. [Google Scholar] [CrossRef] [PubMed]
  10. Grote, V.; Becker, S.; Kaaks, R. Diabetes mellitus type 2—An independent risk factor for cancer? Exp. Clin. Endocrinol. Diabetes 2010, 118, 4–8. [Google Scholar] [CrossRef]
  11. Grivennikov, S.I.; Greten, F.R.; Karin, M. Immunity, Inflammation, and Cancer. Cell 2010, 140, 883–899. [Google Scholar] [CrossRef]
  12. Coussens, L.M.; Werb, Z. Inflammation and cancer. Nature 2002, 420, 860–867. [Google Scholar] [CrossRef]
  13. Hanahan, D.; Weinberg, R.A. Hallmarks of Cancer: The Next Generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef]
  14. Zhang, Y.-Y.; Li, Y.-J.; Xue, C.-D.; Li, S.; Gao, Z.-N.; Qin, K.-R. Effects of T2DM on cancer progression: Pivotal precipitating factors and underlying mechanisms. Front. Endocrinol. 2024, 15, 1396022. [Google Scholar] [CrossRef]
  15. Ahmadabad, H.A.; Panah, S.M.; Ghasemnejad-Berenji, H.; Ghojavand, S.; Ghasemnejad-Berenji, M.; Khezri, M.R. Metformin and the PI3K/AKT signaling pathway: Implications for cancer, cardiovascular, and central nervous system diseases. Naunyn Schmiedebergs Arch. Pharmacol. 2025, 398, 1035–1055. [Google Scholar] [CrossRef]
  16. Sarfstein, R.; Friedman, Y.; Attias-Geva, Z.; Fishman, A.; Bruchim, I.; Werner, H. Metformin downregulates the insulin/IGF-I signaling pathway and inhibits different uterine serous carcinoma (USC) cells proliferation and migration in p53-dependent or -independent manners. PLoS ONE 2013, 8, e61537. [Google Scholar] [CrossRef] [PubMed]
  17. ScienceDirect. Phenformin—Medicine & Dentistry Topic Page. 2024. Available online: https://www.sciencedirect.com/topics/medicine-and-dentistry/phenformin (accessed on 3 June 2025).
  18. Kasznicki, J.; Sliwinska, A.; Drzewoski, J. Metformin in cancer prevention and therapy. Ann. Transl. Med. 2014, 2, 57. [Google Scholar]
  19. 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, 3295–3313. [Google Scholar] [CrossRef]
  20. Kumagai, A.K. Glucose transport in brain and retina: Implications in the management and complications of diabetes. Diabetes Metab. Res. Rev. 1999, 15, 261–273. [Google Scholar] [CrossRef]
  21. Ross, B.D.; Espinal, J.; Silva, P. Glucose metabolism in renal tubular function. Kidney Int. 1986, 29, 54–67. [Google Scholar] [CrossRef] [PubMed]
  22. van Wijk, R.; van Solinge, W.W. The energy-less red blood cell is lost: Erythrocyte enzyme abnormalities of glycolysis. Blood 2005, 106, 4034–4042. [Google Scholar] [CrossRef]
  23. Thau, L.; Sharma, S.G.J. Physiology, Cortisol. In StatPearls [Internet]; StatPearls, Ed.; StatPearls Publishing: Treasure Island, FL, USA, 2023. [Google Scholar]
  24. Gerich, J.E. Glucose Counterregulation and Its Impact on Diabetes Mellitus. Diabetes 1988, 37, 1608–1617. [Google Scholar] [CrossRef]
  25. Schultze, S.M.; Hemmings, B.A.; Niessen, M.; Tschopp, O. PI3K/AKT, MAPK and AMPK signalling: Protein kinases in glucose homeostasis. Expert Rev. Mol. Med. 2012, 14, e1. [Google Scholar] [CrossRef]
  26. Cameron, M.E.; Yakovenko, A.; Trevino, J.G. Glucose and Lactate Transport in Pancreatic Cancer: Glycolytic Metabolism Revisited. J. Oncol. 2018, 2018, 6214838. [Google Scholar] [CrossRef]
  27. Reckzeh, E.S.; Waldmann, H. Small-Molecule Inhibition of Glucose Transporters GLUT-1-4. Chembiochem 2020, 21, 45–52. [Google Scholar] [CrossRef]
  28. Holman, G.D. Structure, function and regulation of mammalian glucose transporters of the SLC2 family. Pflugers Arch. 2020, 472, 1155–1175. [Google Scholar] [CrossRef]
  29. Rahman, M.S.; Hossain, K.S.; Das, S.; Kundu, S.; Adegoke, E.O.; Rahman, M.A.; Uddin, M.J.; Pang, M.-G. Role of Insulin in Health and Disease: An Update. Int. J. Mol. Sci. 2021, 22, 6403. [Google Scholar] [CrossRef]
  30. Batista, T.M.; Haider, N.; Kahn, C.R. Defining the underlying defect in insulin action in type 2 diabetes. Diabetologia 2021, 64, 994–1006. [Google Scholar] [CrossRef] [PubMed]
  31. Janssen, J. The Impact of Westernization on the Insulin/IGF-I Signaling Pathway and the Metabolic Syndrome: It Is Time for Change. Int. J. Mol. Sci. 2023, 24, 4551. [Google Scholar] [CrossRef] [PubMed]
  32. Saltiel, A.R. Insulin signaling in health and disease. J. Clin. Investig. 2021, 131, e142241. [Google Scholar] [CrossRef] [PubMed]
  33. Le, T.K.C.; Dao, X.D.; Nguyen, D.V.; Luu, D.H.; Bui, T.M.H.; Le, T.H.; Nguyen, H.T.; Le, T.N.; Hosaka, T.; Nguyen, T.T.T. Insulin signaling and its application. Front. Endocrinol. 2023, 14, 1226655. [Google Scholar] [CrossRef]
  34. Perz, M.; Torlinska, T. Insulin receptor--structural and functional characteristics. Med. Sci. Monit. 2001, 7, 169–177. [Google Scholar]
  35. Savova, M.S.; Mihaylova, L.V.; Tews, D.; Wabitsch, M.; Georgiev, M.I. Targeting PI3K/AKT signaling pathway in obesity. Biomed. Pharmacother. 2023, 159, 114244. [Google Scholar] [CrossRef]
  36. Sylow, L.; Tokarz, V.L.; Richter, E.A.; Klip, A. The many actions of insulin in skeletal muscle, the paramount tissue determining glycemia. Cell Metab. 2021, 33, 758–780. [Google Scholar] [CrossRef]
  37. Youngren, J.F. Regulation of insulin receptor function. Cell Mol. Life Sci. 2007, 64, 873–891. [Google Scholar] [CrossRef]
  38. Ahmad, E.; Lim, S.; Lamptey, R.; Webb, D.R.; Davies, M.J. Type 2 diabetes. Lancet 2022, 400, 1803–1820. [Google Scholar] [CrossRef]
  39. Lee, J.; Pilch, P.F. The insulin receptor: Structure, function, and signaling. Am. J. Physiol. 1994, 266, C319–C334. [Google Scholar] [CrossRef] [PubMed]
  40. Choi, E.; Bai, X.C. The Activation Mechanism of the Insulin Receptor: A Structural Perspective. Annu. Rev. Biochem. 2023, 92, 247–272. [Google Scholar] [CrossRef] [PubMed]
  41. Ye, L.; Maji, S.; Sanghera, N.; Gopalasingam, P.; Gorbunov, E.; Tarasov, S.; Epstein, O.; Klein-Seetharaman, J. Structure and dynamics of the insulin receptor: Implications for receptor activation and drug discovery. Drug Discov. Today 2017, 22, 1092–1102. [Google Scholar] [CrossRef]
  42. Posner, B.I. Insulin Signalling: The Inside Story. Can. J. Diabetes 2017, 41, 108–113. [Google Scholar] [CrossRef] [PubMed]
  43. Summers, S.A.; Birnbaum, M.J. A role for the serine/threonine kinase, Akt, in insulin-stimulated glucose uptake. Biochem. Soc. Trans. 1997, 25, 981–988. [Google Scholar] [CrossRef]
  44. Holman, G.D.; Kasuga, M. From receptor to transporter: Insulin signalling to glucose transport. Diabetologia 1997, 40, 991–1003. [Google Scholar] [CrossRef]
  45. Manning, B.D.; Toker, A. AKT/PKB Signaling: Navigating the Network. Cell 2017, 169, 381–405. [Google Scholar] [CrossRef] [PubMed]
  46. Taha, C.; Klip, A. The insulin signaling pathway. J. Membr. Biol. 1999, 169, 1–12. [Google Scholar] [CrossRef]
  47. Yudhani, R.D.; Sari, Y.; Nugrahaningsih, D.A.A.; Sholikhah, E.N.; Rochmanti, M.; Purba, A.K.R.; Khotimah, H.; Nugrahenny, D.; Mustofa, M. In Vitro Insulin Resistance Model: A Recent Update. J. Obes. 2023, 2023, 1964732. [Google Scholar] [CrossRef]
  48. Calejman, C.M.; Doxsey, W.; Fazakerley, D.; Guertin, D. Integrating adipocyte insulin signaling and metabolism in the multi-omics era. Trends Biochem. Sci. 2022, 47, 531–546. [Google Scholar] [CrossRef]
  49. Querio, G.; Antoniotti, S.; Levi, R.; Fleischmann, B.K.; Gallo, M.P.; Malan, D. Insulin-Activated Signaling Pathway and GLUT4 Membrane Translocation in hiPSC-Derived Cardiomyocytes. Int. J. Mol. Sci. 2024, 25, 8197. [Google Scholar] [CrossRef]
  50. Kido, Y.; Nakae, J.; Accili, D. Clinical review 125: The insulin receptor and its cellular targets. J. Clin. Endocrinol. Metab. 2001, 86, 972–979. [Google Scholar] [PubMed]
  51. Gray, C.W.; Coster, A.C.F. From insulin to Akt: Time delays and dominant processes. J. Theor. Biol. 2020, 507, 110454. [Google Scholar] [CrossRef]
  52. Scheithauer, T.P.M.; Rampanelli, E.; Nieuwdorp, M.; Vallance, B.A.; Verchere, C.B.; van Raalte, D.H.; Herrema, H. Gut Microbiota as a Trigger for Metabolic Inflammation in Obesity and Type 2 Diabetes. Front. Immunol. 2020, 11, 571731. [Google Scholar] [CrossRef]
  53. Sabio, G.; Das, M.; Mora, A.; Zhang, Z.; Jun, J.Y.; Ko, H.J.; Barrett, T.; Kim, J.K.; Davis, R.J. A stress signaling pathway in adipose tissue regulates hepatic insulin resistance. Science 2008, 322, 1539–1543. [Google Scholar] [CrossRef] [PubMed]
  54. Yaribeygi, H.; Maleki, M.; Sathyapalan, T.; Jamialahmadi, T.; Sahebkar, A. Pathophysiology of Physical Inactivity-Dependent Insulin Resistance: A Theoretical Mechanistic Review Emphasizing Clinical Evidence. J. Diabetes Res. 2021, 2021, 7796727. [Google Scholar] [CrossRef]
  55. Acosta-Martinez, M.; Cabail, M.Z. The PI3K/Akt Pathway in Meta-Inflammation. Int. J. Mol. Sci. 2022, 23, 15330. [Google Scholar] [CrossRef]
  56. Sanches, J.M.; Na Zhao, L.; Salehi, A.; Wollheim, C.B.; Kaldis, P. Pathophysiology of type 2 diabetes and the impact of altered metabolic interorgan crosstalk. FEBS J. 2023, 290, 620–648. [Google Scholar] [CrossRef] [PubMed]
  57. Kim, J.A.; Wei, Y.; Sowers, J.R. Role of mitochondrial dysfunction in insulin resistance. Circ. Res. 2008, 102, 401–414. [Google Scholar] [CrossRef] [PubMed]
  58. Sergi, D.; Naumovski, N.; Heilbronn, L.K.; Abeywardena, M.; O’callaghan, N.; Lionetti, L.; Luscombe-Marsh, N. Mitochondrial (Dys)function and Insulin Resistance: From Pathophysiological Molecular Mechanisms to the Impact of Diet. Front. Physiol. 2019, 10, 532. [Google Scholar] [CrossRef] [PubMed]
  59. Su, K.-Z.; Li, Y.-R.; Zhang, D.; Yuan, J.-H.; Zhang, C.-S.; Liu, Y.; Song, L.-M.; Lin, Q.; Li, M.-W.; Dong, J. Relation of Circulating Resistin to Insulin Resistance in Type 2 Diabetes and Obesity: A Systematic Review and Meta-Analysis. Front. Physiol. 2019, 10, 1399. [Google Scholar] [CrossRef]
  60. Bassin, S.R.; Srinath, R. The Impact of Physical Activity in Patients With Type 2 Diabetes. Am. J. Lifestyle Med. 2025, 19, 147–161. [Google Scholar] [CrossRef]
  61. Colberg, S.R.; Sigal, R.J.; Yardley, J.E.; Riddell, M.C.; Dunstan, D.W.; Dempsey, P.C.; Horton, E.S.; Castorino, K.; Tate, D.F. Physical Activity/Exercise and Diabetes: A Position Statement of the American Diabetes Association. Diabetes Care 2016, 39, 2065–2079. [Google Scholar] [CrossRef]
  62. Umpierre, D.; Ribeiro, P.A.B.; Kramer, C.K.; Leitão, C.B.; Zucatti, A.T.N.; Azevedo, M.J.; Gross, J.L.; Ribeiro, J.P.; Schaan, B.D. Physical activity advice only or structured exercise training and association with HbA1c levels in type 2 diabetes: A systematic review and meta-analysis. JAMA 2011, 305, 1790–1799. [Google Scholar] [CrossRef]
  63. Hamdy, O.; Goodyear, L.J.; Horton, E.S. Diet and exercise in type 2 diabetes mellitus. Endocrinol. Metab. Clin. N. Am. 2001, 30, 883–907. [Google Scholar] [CrossRef]
  64. García-Molina, L.; Lewis-Mikhael, A.-M.; Riquelme-Gallego, B.; Cano-Ibáñez, N.; Oliveras-López, M.-J.; Bueno-Cavanillas, A. Improving type 2 diabetes mellitus glycaemic control through lifestyle modification implementing diet intervention: A systematic review and meta-analysis. Eur. J. Nutr. 2020, 59, 1313–1328. [Google Scholar] [CrossRef]
  65. Campbell, D.J.T.; Campbell, D.B.; Ogundeji, Y.; Au, F.; Beall, R.; Ronksley, P.E.; Quinn, A.E.; Manns, B.J.; Hemmelgarn, B.R.; Tonelli, M.; et al. First-line pharmacotherapy for incident type 2 diabetes: Prescription patterns, adherence and associated costs. Diabet. Med. 2021, 38, e14622. [Google Scholar] [CrossRef]
  66. Ashcroft, F.M. Mechanisms of the glycaemic effects of sulfonylureas. Horm. Metab. Res. 1996, 28, 456–463. [Google Scholar] [CrossRef]
  67. Sibony, R.W.; Segev, O.; Dor, S.; Raz, I. Drug Therapies for Diabetes. Int. J. Mol. Sci. 2023, 24, 17147. [Google Scholar] [CrossRef] [PubMed]
  68. LaMoia, T.E.; Shulman, G.I. Cellular and Molecular Mechanisms of Metformin Action. Endocr. Rev. 2021, 42, 77–96. [Google Scholar] [CrossRef] [PubMed]
  69. Ortega, Á.D.; Sánchez-Aragó, M.; Giner-Sánchez, D.; Sánchez-Cenizo, L.; Willers, I.; Cuezva, J.M. Glucose avidity of carcinomas. Cancer Lett. 2009, 276, 125–135. [Google Scholar] [CrossRef]
  70. Zaal, E.A.; Berkers, C.R. The Influence of Metabolism on Drug Response in Cancer. Front. Oncol. 2018, 8, 500. [Google Scholar] [CrossRef]
  71. Sun, H.; Zhu, A.; Zhou, X.; Wang, F. Suppression of pyruvate dehydrogenase kinase-2 re-sensitizes paclitaxel-resistant human lung cancer cells to paclitaxel. Oncotarget 2017, 8, 52642–52650. [Google Scholar] [CrossRef] [PubMed]
  72. Vanhove, K.; Graulus, G.-J.; Mesotten, L.; Thomeer, M.; Derveaux, E.; Noben, J.-P.; Guedens, W.; Adriaensens, P. The Metabolic Landscape of Lung Cancer: New Insights in a Disturbed Glucose Metabolism. Front. Oncol. 2019, 9, 1215. [Google Scholar] [CrossRef]
  73. Kawatani, M.; Osada, H. Small-molecule inhibitors of glucose transporters. Vitam. Horm. 2025, 128, 213–242. [Google Scholar]
  74. Zhou, Z.; Li, Y.; Chen, S.; Xie, Z.; Du, Y.; Liu, Y.; Shi, Y.; Lin, X.; Zeng, X.; Zhao, H.; et al. GLUT1 promotes cell proliferation via binds and stabilizes phosphorylated EGFR in lung adenocarcinoma. Cell Commun. Signal 2024, 22, 303. [Google Scholar] [CrossRef] [PubMed]
  75. Wang, K.; Li, Q.; Fan, Y.; Fang, P.; Zhou, H.; Huang, J. OBHS Drives Abnormal Glycometabolis Reprogramming via GLUT1 in Breast Cancer. Int. J. Mol. Sci. 2023, 24, 7136. [Google Scholar] [CrossRef] [PubMed]
  76. Meziou, S.; Goulet, C.R.; Hovington, H.; Lefebvre, V.; Lavallée, E.; Bergeron, M.; Brisson, H.; Champagne, A.; Neveu, B.; Lacombe, D.; et al. GLUT1 expression in high-risk prostate cancer: Correlation with (18)F-FDG-PET/CT and clinical outcome. Prostate Cancer Prostatic Dis. 2020, 23, 441–448. [Google Scholar] [CrossRef]
  77. Yang, H.; Yang, S.; He, J.; Li, W.; Zhang, A.; Li, N.; Zhou, G.; Sun, B. Glucose transporter 3 (GLUT3) promotes lactylation modifications by regulating lactate dehydrogenase A (LDHA) in gastric cancer. Cancer Cell Int. 2023, 23, 303. [Google Scholar] [CrossRef]
  78. Al Tameemi, W.; Dale, T.P.; Al-Jumaily, R.M.K.; Forsyth, N.R. Hypoxia-Modified Cancer Cell Metabolism. Front. Cell Dev. Biol. 2019, 7, 4. [Google Scholar] [CrossRef]
  79. Panpan, S.I.; Wei, G.E.; Kaiming, W.U.; Zhang, R. O-GlcNAcylation of hexokinase 2 modulates mitochondrial dynamics and enhances the progression of lung cancer. Mol. Cell Biochem. 2025, 480, 2633–2643. [Google Scholar] [CrossRef]
  80. Chen, C.; Zhang, X. Glycolysis regulator PFKP induces human melanoma cell proliferation and tumor growth. Clin. Transl. Oncol. 2023, 25, 2183–2191. [Google Scholar] [CrossRef]
  81. Zheng, S.; Liu, Q.; Liu, T.; Lu, X. Posttranslational modification of pyruvate kinase type M2 (PKM2): Novel regulation of its biological roles to be further discovered. J. Physiol. Biochem. 2021, 77, 355–363. [Google Scholar] [CrossRef]
  82. Shu, Y.; Yue, J.; Li, Y.; Yin, Y.; Wang, J.; Li, T.; He, X.; Liang, S.; Zhang, G.; Liu, Z.; et al. Development of human lactate dehydrogenase a inhibitors: High-throughput screening, molecular dynamics simulation and enzyme activity assay. J. Comput. Aided Mol. Des. 2024, 38, 28. [Google Scholar] [CrossRef]
  83. Vander Heiden, M.G.; DeBerardinis, R.J. Understanding the Intersections between Metabolism and Cancer Biology. Cell 2017, 168, 657–669. [Google Scholar] [CrossRef] [PubMed]
  84. Zong, W.X.; Rabinowitz, J.D.; White, E. Mitochondria and Cancer. Mol. Cell 2016, 61, 667–676. [Google Scholar] [CrossRef] [PubMed]
  85. Sancho, P.; Barneda, D.; Heeschen, C. Hallmarks of cancer stem cell metabolism. Br. J. Cancer 2016, 114, 1305–1312. [Google Scholar] [CrossRef]
  86. Martinez-Reyes, I.; Chandel, N.S. Mitochondrial TCA cycle metabolites control physiology and disease. Nat. Commun. 2020, 11, 102. [Google Scholar] [CrossRef]
  87. Ashton, T.M.; McKenna, W.G.; Kunz-Schughart, L.A.; Higgins, G.S. Oxidative Phosphorylation as an Emerging Target in Cancer Therapy. Clin. Cancer Res. 2018, 24, 2482–2490. [Google Scholar] [CrossRef] [PubMed]
  88. Paredes, F.; Williams, H.C.; San Martin, A. Metabolic adaptation in hypoxia and cancer. Cancer Lett. 2021, 502, 133–142. [Google Scholar] [CrossRef]
  89. Romero-Garcia, S.; Moreno-Altamirano, M.M.B.; Prado-Garcia, H.; Sánchez-García, F.J. Lactate Contribution to the Tumor Microenvironment: Mechanisms, Effects on Immune Cells and Therapeutic Relevance. Front. Immunol. 2016, 7, 52. [Google Scholar] [CrossRef]
  90. Li, J.; Eu, J.Q.; Kong, L.R.; Wang, L.; Lim, Y.C.; Goh, B.C.; Wong, A.L.A. Targeting Metabolism in Cancer Cells and the Tumour Microenvironment for Cancer Therapy. Molecules 2020, 25, 4831. [Google Scholar] [CrossRef]
  91. Xia, L.; Oyang, L.; Lin, J.; Tan, S.; Han, Y.; Wu, N.; Yi, P.; Tang, L.; Pan, Q.; Rao, S.; et al. The cancer metabolic reprogramming and immune response. Mol. Cancer. 2021, 20, 28. [Google Scholar] [CrossRef] [PubMed]
  92. Madiraju, A.K.; Qiu, Y.; Perry, R.J.; Rahimi, Y.; Zhang, X.-M.; Zhang, D.; Camporez, J.-P.G.; Cline, G.W.; Butrico, G.M.; Kemp, B.E.; et al. Metformin inhibits gluconeogenesis via a redox-dependent mechanism in vivo. Nat. Med. 2018, 24, 1384–1394. [Google Scholar] [CrossRef]
  93. 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]
  94. 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]
  95. Kim, J.; Yang, G.; Kim, Y.; Kim, J.; Ha, J. AMPK activators: Mechanisms of action and physiological activities. Exp. Mol. Med. 2016, 48, e224. [Google Scholar] [CrossRef]
  96. Zhou, G.; Mayers, R.; Li, Y.; Chen, Y.; Shen, X.; Fenyk-Melody, J.; Wu, M.; Ventre, J.; Doebber, T.; Fujii, N.; et al. Role of AMP-activated protein kinase in mechanism of metformin action. J. Clin. Investig. 2001, 108, 1167–1174. [Google Scholar] [CrossRef] [PubMed]
  97. Goel, S.; Singh, R.; Singh, V.; Singh, H.; Kumari, P.; Chopra, H.; Sharma, R.; Nepovimova, E.; Valis, M.; Kuca, K.; et al. Metformin: Activation of 5’ AMP-activated protein kinase and its emerging potential beyond anti-hyperglycemic action. Front. Genet. 2022, 13, 1022739. [Google Scholar] [CrossRef]
  98. Lee, C.B.; Chae, S.U.; Jo, S.J.; Jerng, U.M.; Bae, S.K. The Relationship between the Gut Microbiome and Metformin as a Key for Treating Type 2 Diabetes Mellitus. Int. J. Mol. Sci. 2021, 22, 3566. [Google Scholar] [CrossRef] [PubMed]
  99. Induri, S.N.R.; Kansara, P.; Thomas, S.C.; Xu, F.; Saxena, D.; Li, X. The Gut Microbiome, Metformin, and Aging. Annu. Rev. Pharmacol. Toxicol. 2022, 62, 85–108. [Google Scholar] [CrossRef]
  100. Yerevanian, A.; Soukas, A.A. Metformin: Mechanisms in Human Obesity and Weight Loss. Curr. Obes. Rep. 2019, 8, 156–164. [Google Scholar] [CrossRef]
  101. Maniar, K.; Moideen, A.; Bhattacharyya, R.; Banerjee, D. Metformin exerts anti-obesity effect via gut microbiome modulation in prediabetics: A hypothesis. Med. Hypotheses 2017, 104, 117–120. [Google Scholar] [CrossRef] [PubMed]
  102. Chan, J.C.N.; Yang, A.; Chu, N.; Chow, E. Current type 2 diabetes guidelines: Individualized treatment and how to make the most of metformin. Diabetes Obes. Metab. 2024, 26 (Suppl. S3), 55–74. [Google Scholar] [CrossRef]
  103. Whang, A.; Nagpal, R.; Yadav, H. Bi-directional drug-microbiome interactions of anti-diabetics. eBioMedicine 2019, 39, 591–602. [Google Scholar] [CrossRef]
  104. Forslund, K.; Hildebrand, F.; Nielsen, T.; Falony, G.; Le Chatelier, E.; Sunagawa, S.; Prifti, E.; Vieira-Silva, S.; Gudmundsdottir, V.; Krogh Pedersen, H.; et al. Disentangling type 2 diabetes and metformin treatment signatures in the human gut microbiota. Nature 2015, 528, 262–266. [Google Scholar] [CrossRef]
  105. Weersma, R.K.; Zhernakova, A.; Fu, J. Interaction between drugs and the gut microbiome. Gut 2020, 69, 1510–1519. [Google Scholar] [CrossRef] [PubMed]
  106. Wu, H.; Esteve, E.; Tremaroli, V.; Khan, M.T.; Caesar, R.; Mannerås-Holm, L.; Ståhlman, M.; Olsson, L.M.; Serino, M.; Planas-Fèlix, M.; et al. Metformin alters the gut microbiome of individuals with treatment-naive type 2 diabetes, contributing to the therapeutic effects of the drug. Nat. Med. 2017, 23, 850–858. [Google Scholar] [CrossRef] [PubMed]
  107. Pascale, A.; Marchesi, N.; Govoni, S.; Coppola, A.; Gazzaruso, C. The role of gut microbiota in obesity, diabetes mellitus, and effect of metformin: New insights into old diseases. Curr. Opin. Pharmacol. 2019, 49, 1–5. [Google Scholar] [CrossRef]
  108. Rosell-Díaz, M.; Petit-Gay, A.; Molas-Prat, C.; Gallardo-Nuell, L.; Ramió-Torrentà, L.; Garre-Olmo, J.; Pérez-Brocal, V.; Moya, A.; Jové, M.; Pamplona, R.; et al. Metformin-induced changes in the gut microbiome and plasma metabolome are associated with cognition in men. Metabolism 2024, 157, 155941. [Google Scholar] [CrossRef] [PubMed]
  109. Jones, G.R.; Molloy, M.P. Metformin, Microbiome and Protection Against Colorectal Cancer. Dig. Dis. Sci. 2021, 66, 1409–1414. [Google Scholar] [CrossRef]
  110. Mueller, N.T.; Differding, M.K.; Zhang, M.; Maruthur, N.M.; Juraschek, S.P.; Miller, E.R., 3rd; Appel, L.J.; Yeh, H.-C. Metformin Affects Gut Microbiome Composition and Function and Circulating Short-Chain Fatty Acids: A Randomized Trial. Diabetes Care 2021, 44, 1462–1471. [Google Scholar] [CrossRef]
  111. Zhang, Q.; Hu, N. Effects of Metformin on the Gut Microbiota in Obesity and Type 2 Diabetes Mellitus. Diabetes Metab. Syndr. Obes. 2020, 13, 5003–5014. [Google Scholar] [CrossRef]
  112. Xie, X.; Li, W.; Xiong, Z.; Xu, J.; Liao, T.; Sun, L.; Xu, H.; Zhang, M.; Zhou, J.; Xiong, W.; et al. Metformin reprograms tryptophan metabolism via gut microbiome-derived bile acid metabolites to ameliorate depression-Like behaviors in mice. Brain Behav. Immun. 2025, 123, 442–455. [Google Scholar] [CrossRef]
  113. Broadfield, L.A.; Saigal, A.; Szamosi, J.C.; Hammill, J.A.; Bezverbnaya, K.; Wang, D.; Gautam, J.; Tsakiridis, E.E.; Di Pastena, F.; McNicol, J.; et al. Metformin-induced reductions in tumor growth involves modulation of the gut microbiome. Mol. Metab. 2022, 61, 101498. [Google Scholar] [CrossRef]
  114. Seifarth, C.; Schehler, B.; Schneider, H.J. Effectiveness of metformin on weight loss in non-diabetic individuals with obesity. Exp. Clin. Endocrinol. Diabetes 2013, 121, 27–31. [Google Scholar] [CrossRef]
  115. Lee, A.; Morley, J.E. Metformin decreases food consumption and induces weight loss in subjects with obesity with type II non-insulin-dependent diabetes. Obes. Res. 1998, 6, 47–53. [Google Scholar] [CrossRef]
  116. Haber, R.; Zarzour, F.; Ghezzawi, M.; Saadeh, N.; Bacha, D.S.; Al Jebbawi, L.; Chakhtoura, M.; Mantzoros, C.S. The impact of metformin on weight and metabolic parameters in patients with obesity: A systematic review and meta-analysis of randomized controlled trials. Diabetes Obes. Metab. 2024, 26, 1850–1867. [Google Scholar] [CrossRef]
  117. Pu, R.; Shi, D.; Gan, T.; Ren, X.; Ba, Y.; Huo, Y.; Bai, Y.; Zheng, T.; Cheng, N. Effects of metformin in obesity treatment in different populations: A meta-analysis. Ther. Adv. Endocrinol. Metab. 2020, 11, 2042018820926000. [Google Scholar] [CrossRef] [PubMed]
  118. 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]
  119. Warburg, O. On the origin of cancer cells. Science 1956, 123, 309–314. [Google Scholar] [CrossRef]
  120. 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]
  121. Zheng, J. Energy metabolism of cancer: Glycolysis versus oxidative phosphorylation (Review). Oncol. Lett. 2012, 4, 1151–1157. [Google Scholar] [CrossRef]
  122. Koppenol, W.H.; Bounds, P.L.; Dang, C.V. Otto Warburg’s contributions to current concepts of cancer metabolism. Nat. Rev. Cancer 2011, 11, 325–337. [Google Scholar] [CrossRef] [PubMed]
  123. Zu, X.L.; Guppy, M. Cancer metabolism: Facts, fantasy, and fiction. Biochem. Biophys. Res. Commun. 2004, 313, 459–465. [Google Scholar] [CrossRef] [PubMed]
  124. Cunha, A.; Silva, P.M.A.; Sarmento, B.; Queirós, O. Targeting Glucose Metabolism in Cancer Cells as an Approach to Overcoming Drug Resistance. Pharmaceutics 2023, 15, 2610. [Google Scholar] [CrossRef]
  125. Tavares-Valente, D.; Granja, S.; Baltazar, F.; Queirós, O. Bioenergetic modulators hamper cancer cell viability and enhance response to chemotherapy. J. Cell Mol. Med. 2018, 22, 3782–3794. [Google Scholar] [CrossRef]
  126. Yendapally, R.; Sikazwe, D.; Kim, S.S.; Ramsinghani, S.; Fraser-Spears, R.; Witte, A.P.; La-Viola, B. A review of phenformin, metformin, and imeglimin. Drug Dev. Res. 2020, 81, 390–401. [Google Scholar] [CrossRef]
  127. Bailey, C.J. Metformin: Historical overview. Diabetologia 2017, 60, 1566–1576. [Google Scholar] [CrossRef]
  128. Andrzejewski, S.; Siegel, P.M.; St-Pierre, J. Metabolic Profiles Associated with Metformin Efficacy in Cancer. Front. Endocrinol. 2018, 9, 372. [Google Scholar] [CrossRef]
  129. Hardie, D.G. AMP-activated protein kinase: An energy sensor that regulates all aspects of cell function. Genes Dev. 2011, 25, 1895–1908. [Google Scholar] [CrossRef]
  130. Wu, Y.; Zhang, Z.; Ren, M.; Chen, Y.; Zhang, J.; Li, J.; Gao, F.; Bao, Y.; Huang, Y.; Yang, X.; et al. Metformin Induces Apoptosis and Ferroptosis of Ovarian Cancer Cells Under Energy Stress Conditions. Cells 2025, 14, 213. [Google Scholar] [CrossRef] [PubMed]
  131. Sikka, A.; Kaur, M.; Agarwal, C.; Deep, G.; Agarwal, R. Metformin suppresses growth of human head and neck squamous cell carcinoma via global inhibition of protein translation. Cell Cycle 2012, 11, 1374–1382. [Google Scholar] [CrossRef] [PubMed]
  132. Wu, X.; Yeerna, H.; Goto, Y.; Ando, T.; Wu, V.H.; Zhang, X.; Wang, Z.; Amornphimoltham, P.; Murphy, A.N.; Tamayo, P.; et al. Metformin Inhibits Progression of Head and Neck Squamous Cell Carcinoma by Acting Directly on Carcinoma-Initiating Cells. Cancer Res. 2019, 79, 4360–4370. [Google Scholar] [CrossRef] [PubMed]
  133. Shameem, M.; Bagherpoor, A.J.; Nakhi, A.; Dosa, P.; Georg, G.; Kassie, F. Mitochondria-targeted metformin (mitomet) inhibits lung cancer in cellular models and in mice by enhancing the generation of reactive oxygen species. Mol. Carcinog. 2023, 62, 1619–1629. [Google Scholar] [CrossRef]
  134. Teng, M.; Li, Z.; Gu, Y.; Fan, Y.; Wang, D.; Liu, M.; Li, Y.; Wei, G.; Huang, Y. Real-time monitoring of glucose metabolism and effects of metformin on HepG2 cells using (13)C in-cell NMR spectroscopy. Biochem. Biophys. Res. Commun. 2024, 694, 149383. [Google Scholar] [CrossRef] [PubMed]
  135. Dykens, J.A.; Jamieson, J.; Marroquin, L.; Nadanaciva, S.; Billis, P.A.; Will, Y. Biguanide-induced mitochondrial dysfunction yields increased lactate production and cytotoxicity of aerobically-poised HepG2 cells and human hepatocytes in vitro. Toxicol. Appl. Pharmacol. 2008, 233, 203–210. [Google Scholar] [CrossRef] [PubMed]
  136. Ko, M.; Kim, J.; Lazim, R.; Lee, J.Y.; Kim, J.Y.; Gosu, V.; Lee, Y.; Choi, S.; Kwon, H.J. The anticancer effect of metformin targets VDAC1 via ER-mitochondria interactions-mediated autophagy in HCC. Exp. Mol. Med. 2024, 56, 2714–2725. [Google Scholar] [CrossRef]
  137. Lee, J.O.; Kang, M.J.; Byun, W.S.; Kim, S.A.; Seo, I.H.; Han, J.A.; Moon, J.W.; Kim, J.H.; Kim, S.J.; Lee, E.J.; et al. Metformin overcomes resistance to cisplatin in triple-negative breast cancer (TNBC) cells by targeting RAD51. Breast Cancer Res. 2019, 21, 115. [Google Scholar] [CrossRef]
  138. Zhuang, Y.; Miskimins, W.K. Cell cycle arrest in Metformin treated breast cancer cells involves activation of AMPK, downregulation of cyclin D1, and requires p27Kip1 or p21Cip1. J. Mol. Signal 2008, 3, 18. [Google Scholar] [CrossRef]
  139. Alimova, I.N.; Liu, B.; Fan, Z.; Edgerton, S.M.; Dillon, T.; Lind, S.E.; Thor, A.D. Metformin inhibits breast cancer cell growth, colony formation and induces cell cycle arrest in vitro. Cell Cycle 2009, 8, 909–915. [Google Scholar] [CrossRef]
  140. Amaral, I.; Silva, C.; Correia-Branco, A.; Martel, F. Effect of metformin on estrogen and progesterone receptor-positive (MCF-7) and triple-negative (MDA-MB-231) breast cancer cells. Biomed. Pharmacother. 2018, 102, 94–101. [Google Scholar] [CrossRef] [PubMed]
  141. Lonardo, E.; Cioffi, M.; Sancho, P.; Sanchez-Ripoll, Y.; Trabulo, S.M.; Dorado, J.; Balic, A.; Hidalgo, M.; Heeschen, C. Metformin targets the metabolic achilles heel of human pancreatic cancer stem cells. PLoS ONE 2013, 8, e76518. [Google Scholar] [CrossRef]
  142. Karnevi, E.; Said, K.; Andersson, R.; Rosendahl, A.H. Metformin-mediated growth inhibition involves suppression of the IGF-I receptor signalling pathway in human pancreatic cancer cells. BMC Cancer 2013, 13, 235. [Google Scholar] [CrossRef]
  143. 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]
  144. Scotland, S.; Saland, E.; Skuli, N.; de Toni, F.; Boutzen, H.; Micklow, E.; Sénégas, I.; Peyraud, R.; Peyriga, L.; Théodoro, F.; et al. Mitochondrial energetic and AKT status mediate metabolic effects and apoptosis of metformin in human leukemic cells. Leukemia 2013, 27, 2129–2138. [Google Scholar] [CrossRef]
  145. Ben Sahra, I.; Laurent, K.; Giuliano, S.; Larbret, F.; Ponzio, G.; Gounon, P.; Le Marchand-Brustel, Y.; Giorgetti-Peraldi, S.; Cormont, M.; Bertolotto, C.; et al. Targeting cancer cell metabolism: The combination of metformin and 2-deoxyglucose induces p53-dependent apoptosis in prostate cancer cells. Cancer Res. 2010, 70, 2465–2475. [Google Scholar] [CrossRef]
  146. 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.; et al. Metformin inhibits mitochondrial complex I of cancer cells to reduce tumorigenesis. eLife 2014, 3, e02242. [Google Scholar] [CrossRef] [PubMed]
  147. Pavlides, S.; Whitaker-Menezes, D.; Castello-Cros, R.; Flomenberg, N.; Witkiewicz, A.K.; Frank, P.G.; Casimiro, M.C.; Wang, C.; Fortina, P.; Addya, S.; et al. The reverse Warburg effect: Aerobic glycolysis in cancer associated fibroblasts and the tumor stroma. Cell Cycle 2009, 8, 3984–4001. [Google Scholar] [CrossRef] [PubMed]
  148. 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]
  149. Owen, M.R.; Doran, E.; Halestrap, A.P. Evidence that metformin exerts its anti-diabetic effects through inhibition of complex 1 of the mitochondrial respiratory chain. Biochem. J. 2000, 348 Pt 3, 607–614. [Google Scholar] [CrossRef]
  150. Warkad, M.S.; Kim, C.-H.; Kang, B.-G.; Park, S.-H.; Jung, J.-S.; Feng, J.-H.; Inci, G.; Kim, S.-C.; Suh, H.-W.; Lim, S.S.; et al. Metformin-induced ROS upregulation as amplified by apigenin causes profound anticancer activity while sparing normal cells. Sci. Rep. 2021, 11, 14002. [Google Scholar] [CrossRef] [PubMed]
  151. Westhaus, A.; Blumrich, E.M.; Dringen, R. The Antidiabetic Drug Metformin Stimulates Glycolytic Lactate Production in Cultured Primary Rat Astrocytes. Neurochem. Res. 2017, 42, 294–305. [Google Scholar] [CrossRef]
  152. Feng, J.; Wang, X.; Ye, X.; Ares, I.; Lopez-Torres, B.; Martínez, M.; Martínez-Larrañaga, M.-R.; Wang, X.; Anadón, A.; Martínez, M.-A. Mitochondria as an important target of metformin: The mechanism of action, toxic and side effects, and new therapeutic applications. Pharmacol. Res. 2022, 177, 106114. [Google Scholar] [CrossRef]
  153. Liu, S.; Washio, J.; Sato, S.; Abiko, Y.; Shinohara, Y.; Kobayashi, Y.; Otani, H.; Sasaki, S.; Wang, X.; Takahashi, N. Rewired Cellular Metabolic Profiles in Response to Metformin under Different Oxygen and Nutrient Conditions. Int. J. Mol. Sci. 2022, 23, 989. [Google Scholar] [CrossRef]
  154. Weinberg, S.E.; Chandel, N.S. Targeting mitochondria metabolism for cancer therapy. Nat. Chem. Biol. 2015, 11, 9–15. [Google Scholar] [CrossRef]
  155. 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]
  156. Wilson, W.R.; Hay, M.P. Targeting hypoxia in cancer therapy. Nat. Rev. Cancer 2011, 11, 393–410. [Google Scholar] [CrossRef]
  157. Semenza, G.L. HIF-1: Upstream and downstream of cancer metabolism. Curr. Opin. Genet. Dev. 2010, 20, 51–56. [Google Scholar] [CrossRef]
  158. Semenza, G.L. Evaluation of HIF-1 inhibitors as anticancer agents. Drug Discov. Today 2007, 12, 853–859. [Google Scholar] [CrossRef] [PubMed]
  159. 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]
  160. Apostolova, N.; Iannantuoni, F.; Gruevska, A.; Muntane, J.; Rocha, M.; Victor, V.M. Mechanisms of action of metformin in type 2 diabetes: Effects on mitochondria and leukocyte-endothelium interactions. Redox Biol. 2020, 34, 101517. [Google Scholar] [CrossRef]
  161. Zamanian, M.Y.; Golmohammadi, M.; Yumashev, A.; Hjazi, A.; Toama, M.A.; AbdRabou, M.A.; Gehlot, A.; Alwaily, E.R.; Shirsalimi, N.; Yadav, P.K.; et al. Effects of metformin on cancers in experimental and clinical studies: Focusing on autophagy and AMPK/mTOR signaling pathways. Cell Biochem. Funct. 2024, 42, e4071. [Google Scholar] [CrossRef]
  162. Viollet, B.; Guigas, B.; Garcia, N.S.; Leclerc, J.; Foretz, M.; Andreelli, F. Cellular and molecular mechanisms of metformin: An overview. Clin. Sci. 2011, 122, 253–270. [Google Scholar] [CrossRef] [PubMed]
  163. Reczek, C.R.; Chandel, N.S. The Two Faces of Reactive Oxygen Species in Cancer. Annu. Rev. Cancer Biol. 2017, 1, 79–98. [Google Scholar] [CrossRef] [PubMed]
  164. Marengo, B.; Nitti, M.; Furfaro, A.L.; Colla, R.; De Ciucis, C.; Marinari, U.M.; Pronzato, M.A.; Traverso, N.; Domenicotti, C. Redox Homeostasis and Cellular Antioxidant Systems: Crucial Players in Cancer Growth and Therapy. Oxidative Med. Cell. Longev. 2016, 2016, 6235641. [Google Scholar] [CrossRef]
  165. Harris, I.S.; DeNicola, G.M. The Complex Interplay between Antioxidants and ROS in Cancer. Trends Cell Biol. 2020, 30, 440–451. [Google Scholar] [CrossRef] [PubMed]
  166. Nakamura, H.; Takada, K. Reactive oxygen species in cancer: Current findings and future directions. Cancer Sci. 2021, 112, 3945–3952. [Google Scholar] [CrossRef]
  167. Zhao, Y.; Ye, X.; Xiong, Z.; Ihsan, A.; Ares, I.; Martínez, M.; Lopez-Torres, B.; Martínez-Larrañaga, M.-R.; Anadón, A.; Wang, X.; et al. Cancer Metabolism: The Role of ROS in DNA Damage and Induction of Apoptosis in Cancer Cells. Metabolites 2023, 13, 796. [Google Scholar] [CrossRef] [PubMed]
  168. Okon, I.S.; Zou, M.-H. Mitochondrial ROS and cancer drug resistance: Implications for therapy. Pharmacol. Res. 2015, 100, 170–174. [Google Scholar] [CrossRef]
  169. 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]
  170. Chowdhury, A.R.; Zielonka, J.; Kalyanaraman, B.; Hartley, R.C.; Murphy, M.P.; Avadhani, N.G. Mitochondria-targeted paraquat and metformin mediate ROS production to induce multiple pathways of retrograde signaling: A dose-dependent phenomenon. Redox Biol. 2020, 36, 101606. [Google Scholar] [CrossRef]
  171. Mu, W.; Jiang, Y.; Liang, G.; Feng, Y.; Qu, F. Metformin: A Promising Antidiabetic Medication for Cancer Treatment. Curr. Drug Targets 2023, 24, 41–54. [Google Scholar] [CrossRef]
  172. Jara, J.A.; López-Muñoz, R. Metformin and cancer: Between the bioenergetic disturbances and the antifolate activity. Pharmacol. Res. 2015, 101, 102–108. [Google Scholar] [CrossRef]
  173. Gou, S.; Qiu, L.; Yang, Q.; Li, P.; Zhou, X.; Sun, Y.; Zhou, X.; Zhao, W.; Zhai, W.; Li, G.; et al. Metformin leads to accumulation of reactive oxygen species by inhibiting the NFE2L1 expression in human hepatocellular carcinoma cells. Toxicol. Appl. Pharmacol. 2021, 420, 115523. [Google Scholar] [CrossRef]
  174. Cheng, G.; Lanza-Jacoby, S. Metformin decreases growth of pancreatic cancer cells by decreasing reactive oxygen species: Role of NOX4. Biochem. Biophys. Res. Commun. 2015, 465, 41–46. [Google Scholar] [CrossRef]
  175. Ghosh, A.; Ghosh, A.K.; Zaman, A.; Das, P.K. Metformin-Loaded Hyaluronic Acid-Derived Carbon Dots for Targeted Therapy against Hepatocellular Carcinoma by Glutamine Metabolic Reprogramming. Mol. Pharm. 2023, 20, 6391–6406. [Google Scholar] [CrossRef] [PubMed]
  176. Diaz-Morales, N.; Rovira-Llopis, S.; Bañuls, C.; Lopez-Domenech, S.; Escribano-Lopez, I.; Veses, S.; Jover, A.; Rocha, M.; Hernandez-Mijares, A.; Victor, V.M. Does Metformin Protect Diabetic Patients from Oxidative Stress and Leukocyte-Endothelium Interactions? Antioxid. Redox Signal 2017, 27, 1439–1445. [Google Scholar] [CrossRef]
  177. Wang, X.; Yang, L.; Kang, L.; Li, J.; Yang, L.; Zhang, J.; Liu, J.; Zhu, M.; Zhang, Q.; Shen, Y.; et al. Metformin attenuates myocardial ischemia-reperfusion injury via up-regulation of antioxidant enzymes. PLoS ONE 2017, 12, e0182777. [Google Scholar] [CrossRef]
  178. Vancura, A.; Bu, P.; Bhagwat, M.; Zeng, J.; Bhagwat, I. Metformin as an Anticancer Agent. Trends Pharmacol. Sci. 2018, 39, 867–878. [Google Scholar] [CrossRef] [PubMed]
  179. Ikhlas, S.; Ahmad, M. Metformin: Insights into its anticancer potential with special reference to AMPK dependent and independent pathways. Life Sci. 2017, 185, 53–62. [Google Scholar] [CrossRef]
  180. Palma, F.R.; Ratti, B.A.; Paviani, V.; Coelho, D.R.; Miguel, R.; Danes, J.M.; Zaichik, S.V.; de Abreu, A.L.; Silva, S.O.; Chen, Y.; et al. AMPK-deficiency forces metformin-challenged cancer cells to switch from carbohydrate metabolism to ketogenesis to support energy metabolism. Oncogene 2021, 40, 5455–5467. [Google Scholar] [CrossRef]
  181. Howell, J.J.; Hellberg, K.; Turner, M.; Talbott, G.; Kolar, M.J.; Ross, D.S.; Hoxhaj, G.; Saghatelian, A.; Shaw, R.J.; Manning, B.D. Metformin Inhibits Hepatic mTORC1 Signaling via Dose-Dependent Mechanisms Involving AMPK and the TSC Complex. Cell Metab. 2017, 25, 463–471. [Google Scholar] [CrossRef] [PubMed]
  182. Griss, T.; Vincent, E.E.; Egnatchik, R.; Chen, J.; Ma, E.H.; Faubert, B.; Viollet, B.; DeBerardinis, R.J.; Jones, R.G. Metformin Antagonizes Cancer Cell Proliferation by Suppressing Mitochondrial-Dependent Biosynthesis. PLoS Biol. 2015, 13, e1002309. [Google Scholar] [CrossRef]
  183. Herzig, S.; Shaw, R.J. AMPK: Guardian of metabolism and mitochondrial homeostasis. Nat. Rev. Mol. Cell Biol. 2018, 19, 121–135. [Google Scholar] [CrossRef]
  184. Hardie, D.G.; Schaffer, B.E.; Brunet, A. AMPK: An Energy-Sensing Pathway with Multiple Inputs and Outputs. Trends Cell Biol. 2016, 26, 190–201. [Google Scholar] [CrossRef]
  185. Hardie, D.G. Molecular Pathways: Is AMPK a Friend or a Foe in Cancer? Clin. Cancer Res. 2015, 21, 3836–3840. [Google Scholar] [CrossRef]
  186. Vara-Ciruelos, D.; Russell, F.M.; Hardie, D.G. The strange case of AMPK and cancer: Dr Jekyll or Mr Hyde? Open Biol. 2019, 9, 190099. [Google Scholar] [CrossRef]
  187. Sadria, M.; Seo, D.; Layton, A.T. The mixed blessing of AMPK signaling in Cancer treatments. BMC Cancer 2022, 22, 105. [Google Scholar] [CrossRef]
  188. Ling, N.X.Y.; Kaczmarek, A.; Hoque, A.; Davie, E.; Ngoei, K.R.W.; Morrison, K.R.; Smiles, W.J.; Forte, G.M.; Wang, T.; Lie, S.; et al. mTORC1 directly inhibits AMPK to promote cell proliferation under nutrient stress. Nat. Metab. 2020, 2, 41–49. [Google Scholar] [CrossRef]
  189. Vara-Ciruelos, D.; Dandapani, M.; Hardie, D.G. AMP-Activated Protein Kinase: Friend or Foe in Cancer? Annu. Rev. Cancer Biol. 2020, 4, 1–16. [Google Scholar] [CrossRef]
  190. Cai, X.; Hu, X.; Tan, X.; Cheng, W.; Wang, Q.; Chen, X.; Guan, Y.; Chen, C.; Jing, X. Metformin Induced AMPK Activation, G0/G1 Phase Cell Cycle Arrest and the Inhibition of Growth of Esophageal Squamous Cell Carcinomas In Vitro and In Vivo. PLoS ONE 2015, 10, e0133349. [Google Scholar] [CrossRef]
  191. Keerthana, C.K.; Rayginia, T.P.; Shifana, S.C.; Anto, N.P.; Kalimuthu, K.; Isakov, N.; Anto, R.J. The role of AMPK in cancer metabolism and its impact on the immunomodulation of the tumor microenvironment. Front. Immunol. 2023, 14, 1114582. [Google Scholar] [CrossRef]
  192. Faubert, B.; Boily, G.; Izreig, S.; Griss, T.; Samborska, B.; Dong, Z.; Dupuy, F.; Chambers, C.; Fuerth, B.J.; Viollet, B.; et al. AMPK Is a Negative Regulator of the Warburg Effect and Suppresses Tumor Growth In Vivo. Cell Metab. 2013, 17, 113–124. [Google Scholar] [CrossRef]
  193. Li, W.; Saud, S.M.; Young, M.R.; Chen, G.; Hua, B. Targeting AMPK for cancer prevention and treatment. Oncotarget 2015, 6, 7365–7378. [Google Scholar] [CrossRef]
  194. Chomanicova, N.; Gazova, A.; Adamickova, A.; Valaskova, S.; Kyselovic, J. The role of AMPK/mTOR signaling pathway in anticancer activity of metformin. Physiol. Res. 2021, 70, 501–508. [Google Scholar] [CrossRef]
  195. Glaviano, A.; Foo, A.S.C.; Lam, H.Y.; Yap, K.C.H.; Jacot, W.; Jones, R.H.; Eng, H.; Nair, M.G.; Makvandi, P.; Geoerger, B.; et al. PI3K/AKT/mTOR signaling transduction pathway and targeted therapies in cancer. Mol. Cancer 2023, 22, 138. [Google Scholar] [CrossRef]
  196. Yu, L.; Wei, J.; Liu, P. Attacking the PI3K/Akt/mTOR signaling pathway for targeted therapeutic treatment in human cancer. Semin. Cancer Biol. 2022, 85, 69–94. [Google Scholar] [CrossRef]
  197. Popova, N.V.; Jücker, M. The Role of mTOR Signaling as a Therapeutic Target in Cancer. Int. J. Mol. Sci. 2021, 22, 1743. [Google Scholar] [CrossRef]
  198. Daugan, M.; Wojcicki, A.D.; D’hAyer, B.; Boudy, V. Metformin: An anti-diabetic drug to fight cancer. Pharmacol. Res. 2016, 113, 675–685. [Google Scholar] [CrossRef]
  199. Sathe, A.; Chalaud, G.; Oppolzer, I.; Wong, K.Y.; von Busch, M.; Schmid, S.C.; Tong, Z.; Retz, M.; Gschwend, J.E.; Schulz, W.A.; et al. Parallel PI3K, AKT and mTOR inhibition is required to control feedback loops that limit tumor therapy. PLoS ONE 2018, 13, e0190854. [Google Scholar] [CrossRef]
  200. Hsieh, A.C.; Nguyen, H.G.; Wen, L.; Edlind, M.P.; Carroll, P.R.; Kim, W.; Ruggero, D. Cell type-specific abundance of 4EBP1 primes prostate cancer sensitivity or resistance to PI3K pathway inhibitors. Sci. Signal. 2015, 8, ra116. [Google Scholar] [CrossRef]
  201. Hervieu, A.; Kermorgant, S. The Role of PI3K in Met Driven Cancer: A Recap. Front. Mol. Biosci. 2018, 5, 86. [Google Scholar] [CrossRef]
  202. Ferreira, G.D.; Germeyer, A.; Machado, A.d.B.; Nascimento, T.L.D.; Strowitzki, T.; Brum, I.S.; Corleta, H.v.E.; Capp, E. Metformin modulates PI3K and GLUT4 expression and Akt/PKB phosphorylation in human endometrial stromal cells after stimulation with androgen and insulin. Eur. J. Obstet. Gynecol. Reprod. Biol. 2014, 175, 157–162. [Google Scholar] [CrossRef]
  203. Tzatsos, A.; Kandror, K.V. Nutrients Suppress Phosphatidylinositol 3-Kinase/Akt Signaling via Raptor-Dependent mTOR-Mediated Insulin Receptor Substrate 1 Phosphorylation. Mol. Cell. Biol. 2006, 26, 63–76. [Google Scholar] [CrossRef]
  204. Jang, S.-K.; Hong, S.-E.; Lee, D.-H.; Kim, J.-Y.; Hong, J.; Park, I.-C.; Jin, H.-O. Inhibition of AKT Enhances the Sensitivity of NSCLC Cells to Metformin. Anticancer. Res. 2021, 41, 3481–3487. [Google Scholar] [CrossRef]
  205. Papa, A.; Pandolfi, P.P. The PTEN–PI3K Axis in Cancer. Biomolecules 2019, 9, 153. [Google Scholar] [CrossRef]
  206. Zhuang, Y.; Chan, D.K.; Haugrud, A.B.; Miskimins, W.K. Mechanisms by which low glucose enhances the cytotoxicity of metformin to cancer cells both in vitro and in vivo. PLoS ONE 2014, 9, e108444. [Google Scholar] [CrossRef]
  207. Silvestri, A.; Palumbo, F.; Rasi, I.; Posca, D.; Pavlidou, T.; Paoluzi, S.; Castagnoli, L.; Cesareni, G. Metformin Induces Apoptosis and Downregulates Pyruvate Kinase M2 in Breast Cancer Cells Only When Grown in Nutrient-Poor Conditions. PLoS ONE 2015, 10, e0136250. [Google Scholar] [CrossRef]
  208. Yeşildağ, A.; Kızıloğlu, H.T.; Dirican, E.; Erbaş, E.; Gelen, V.; Kara, A. Anticarcinogenic Effects of Gold Nanoparticles and Metformin Against MCF-7 and A549 Cells. Biol. Trace. Elem. Res. 2024, 202, 4494–4507. [Google Scholar] [CrossRef]
  209. De, A.; Wadhwani, A.; Sauraj; Roychowdhury, P.; Kang, J.H.; Ko, Y.T.; Kuppusamy, G. WZB117 Decorated Metformin-Carboxymethyl Chitosan Nanoparticles for Targeting Breast Cancer Metabolism. Polymers 2023, 15, 976. [Google Scholar] [CrossRef]
  210. Abdolahinia, E.D.; Nadri, S.; Rahbarghazi, R.; Barar, J.; Aghanejad, A.; Omidi, Y. Enhanced penetration and cytotoxicity of metformin and collagenase conjugated gold nanoparticles in breast cancer spheroids. Life Sci. 2019, 231, 116545. [Google Scholar] [CrossRef]
  211. Jafari-Gharabaghlou, D.; Dadashpour, M.; Khanghah, O.J.; Salmani-Javan, E.; Zarghami, N. Potentiation of Folate-Functionalized PLGA-PEG nanoparticles loaded with metformin for the treatment of breast Cancer: Possible clinical application. Mol. Biol. Rep. 2023, 50, 3023–3033. [Google Scholar] [CrossRef]
  212. Chen, C.; Yang, L.; Peng, Y.; Zhang, W.J.; Yang, X.X.; Zhou, W. Autophagic blockage by metformin-loaded PLGA nanoparticles causes cell cycle arrest of HepG2 cells. Nanomedicine 2024, 19, 43–58. [Google Scholar] [CrossRef]
  213. Ebrahimnejad, P.; Rezaeiroshan, A.; Babaei, A.; Khanali, A.; Aghajanshakeri, S.; Farmoudeh, A.; Nokhodchi, A. Hyaluronic Acid-Coated Chitosan/Gelatin Nanoparticles as a New Strategy for Topical Delivery of Metformin in Melanoma. Biomed. Res. Int. 2023, 2023, 3304105. [Google Scholar] [CrossRef]
  214. Zhang, F.; Liu, W.; Long, Y.; Peng, H. Targeted Delivery of Metformin Against Lung Cancer Cells Via Hyaluronan-Modified Mesoporous Silica Nanoparticles. Appl. Biochem. Biotechnol. 2023, 195, 4067–4083. [Google Scholar] [CrossRef]
  215. Snima, K.; Jayakumar, R.; Unnikrishnan, A.; Nair, S.V.; Lakshmanan, V.-K. O-carboxymethyl chitosan nanoparticles for metformin delivery to pancreatic cancer cells. Carbohydr. Polym. 2012, 89, 1003–1007. [Google Scholar] [CrossRef]
  216. Shiridokht, F.; Dadashi, H.; Vandghanooni, S.; Eskandani, M.; Farajollahi, A. Metformin-loaded chitosan nanoparticles augment silver nanoparticle-induced radiosensitization in breast cancer cells during radiation therapy. Colloids Surf. B Biointerfaces 2025, 245, 114220. [Google Scholar] [CrossRef]
  217. Sun, L.; Yao, H.-J.; Li, J.-C.; Zhao, B.-Q.; Wang, Y.-A.; Zhang, Y.-G. Activated Carbon nanoparticles Loaded with Metformin for Effective Against Hepatocellular Cancer Stem Cells. Int. J. Nanomed. 2023, 18, 2891–2910. [Google Scholar] [CrossRef]
  218. Abd-Rabou, A.A.; Abdelaziz, A.M.; Shaker, O.G.; Ayeldeen, G. Metformin-loaded lecithin nanoparticles induce colorectal cancer cytotoxicity via epigenetic modulation of noncoding RNAs. Mol. Biol. Rep. 2021, 48, 6805–6820. [Google Scholar] [CrossRef]
  219. Ngwuluka, N.C.; Kotak, D.J.; Devarajan, P.V. Design and Characterization of Metformin-Loaded Solid Lipid Nanoparticles for Colon Cancer. AAPS PharmSciTech 2017, 18, 358–368. [Google Scholar] [CrossRef]
  220. Yang, Y.; Zhang, Z.; Chen, Q.; You, Y.; Li&Ast, X.; Chen, T. Functionalized Selenium Nanoparticles Synergizes With Metformin to Treat Breast Cancer Cells Through Regulation of Selenoproteins. Front. Bioeng. Biotechnol. 2021, 9, 758482. [Google Scholar] [CrossRef]
  221. Parikh, A.B.; Kozuch, P.; Rohs, N.; Becker, D.J.; Levy, B.P. Metformin as a repurposed therapy in advanced non-small cell lung cancer (NSCLC): Results of a phase II trial. Investig. New Drugs 2017, 35, 813–819. [Google Scholar] [CrossRef]
  222. Fatehi-Agdam, M.; Vatankhah, M.A.; Panahizadeh, R.; Jeddi, F.; Najafzadeh, N. Efficacy of Metformin and Chemotherapeutic Agents on the Inhibition of Colony Formation and Shh/Gli1 Pathway: Metformin/Docetaxel Versus Metformin/5-Fluorouracil. Drug Res. 2021, 71, 17–25. [Google Scholar] [CrossRef]
  223. Mlicka, A.; Mlicki, P.; Niewiadomski, P.; Zielińska, W.; Hałas-Wiśniewska, M.; Izdebska, M. Synergistic effect of metformin and doxorubicin on the metastatic potential of T24 cells. Acta Histochem. 2023, 125, 151975. [Google Scholar] [CrossRef]
  224. Patel, P.J.; Shah, J.S. Metformin pretreatment potentiates the antiproliferative action of doxorubicin against breast cancer. Ann. Pharm. Fr. 2023, 81, 636–652. [Google Scholar] [CrossRef]
  225. Sahu, P.; Camarillo, I.G.; Sundararajan, R. Enhanced Antiproliferation Potency of Electrical Pulse-Mediated Metformin and Cisplatin Combination Therapy on MDA-MB-231 Cells. Appl. Biochem. Biotechnol. 2022, 194, 18–36. [Google Scholar] [CrossRef] [PubMed]
  226. Shanchun, H.; You, P.; Sujuan, N.; Xuebing, Z.; Yijie, B.; Xiaohui, X.; Jianming, H.; La, N.; Zhehui, B.; Qi, L.; et al. Integrative analyses of biomarkers and pathways for metformin reversing cisplatin resistance in head and neck squamous cell carcinoma cells. Arch. Oral. Biol. 2023, 147, 105637. [Google Scholar] [CrossRef] [PubMed]
  227. Tortelli, T.C.; Tamura, R.E.; Junqueira, M.d.S.; Mororó, J.d.S.; Bustos, S.O.; Natalino, R.J.M.; Russell, S.; Désaubry, L.; Strauss, B.E.; Chammas, R. Metformin-induced chemosensitization to cisplatin depends on P53 status and is inhibited by Jarid1b overexpression in non-small cell lung cancer cells. Aging 2021, 13, 21914–21940. [Google Scholar] [CrossRef] [PubMed]
  228. Morelli, A.P.; Jr, T.C.T.; Pavan, I.C.B.; Silva, F.R.; Granato, D.C.; Peruca, G.F.; Pauletti, B.A.; Domingues, R.R.; Bezerra, R.M.N.; De Moura, L.P.; et al. Metformin impairs cisplatin resistance effects in A549 lung cancer cells through mTOR signaling and other metabolic pathways. Int. J. Oncol. 2021, 58, 28. [Google Scholar] [CrossRef] [PubMed]
  229. Pradhan, P.M.; Lee, Y.-H.; Jang, S.; Yi, H.-K. Synergistic anti-cancer effects of metformin and cisplatin on YD-9 oral squamous carcinoma cells via AMPK pathway. J. Appl. Oral. Sci. 2025, 33, e20240385. [Google Scholar] [CrossRef]
  230. Zhou, T.; Zhang, X.; Yang, D.; Wei, W.; Gan, J.; Xia, X.; Chen, Q.; Jiang, J.; Feng, X. Metformin overcomes chemoresistance by regulating stemness via KLF4 in oral squamous cell carcinoma. Oral Dis. 2025, 31, 468–481. [Google Scholar] [CrossRef]
  231. Chen, Z.; Zeng, Z.; Zhu, S.; Zeng, Y.; Lin, Q.; Luo, L.; Hong, X. Effects of different combined regimens of cisplatin, metformin, and quercetin on nasopharyngeal carcinoma cells and subcutaneous xenografts. Sci. Rep. 2021, 11, 1040. [Google Scholar] [CrossRef]
  232. Zhang, J.; Zhou, P.; Wu, T.; Zhang, L.; Kang, J.; Liao, J.; Jiang, D.; Hu, Z.; Han, Z.; Zhou, B. Metformin combined with cisplatin reduces anticancer activity via ATM/CHK2-dependent upregulation of Rad51 pathway in ovarian cancer. Neoplasia 2024, 57, 101037. [Google Scholar] [CrossRef]
  233. Jafarzadeh, E.; Montazeri, V.; Aliebrahimi, S.; Sezavar, A.H.; Ghahremani, M.H.; Ostad, S.N. Targeting Cancer Stem Cells and Hedgehog Pathway: Enhancing Cisplatin Efficacy in Ovarian Cancer with Metformin. J. Cell Mol. Med. 2025, 29, e70508. [Google Scholar] [CrossRef]
  234. Xiao, Y.-Y.; Xiao, J.-X.; Wang, X.-Y.; Wang, T.; Qu, X.-H.; Jiang, L.-P.; Tou, F.-F.; Chen, Z.-P.; Han, X.-J. Metformin-induced AMPK activation promotes cisplatin resistance through PINK1/Parkin dependent mitophagy in gastric cancer. Front. Oncol. 2022, 12, 956190. [Google Scholar] [CrossRef]
  235. Fang, C.-W.; Yang, J.-S.; Chiang, J.-H.; Shieh, P.-C.; Tsai, F.-J.; Tsai, C.-W.; Chang, W.-S. Metformin induces autophagy of cisplatin-resistant human gastric cancer cells in addition to apoptosis. Biomedicine 2023, 13, 14–23. [Google Scholar] [CrossRef]
  236. Duan, G.; Qi, M.; Xun, L.; An, Y.; Zuo, Z.; Luo, Y.; Song, Z. Metformin Enhances the Chemosensitivity of Gastric Cancer to Cisplatin by Downregulating Nrf2 Level. Anal. Cell Pathol. 2025, 2025, 5714423. [Google Scholar] [CrossRef] [PubMed]
  237. Hassani, E.; Mozzendizaji, S.; Shafiei-Irannejad, V.; Mohammadzadeh, A. Metformin boosts doxorubicin efficacy and increases CD8+ T cell frequency in mouse breast cancer. Clin. Transl. Oncol. 2025, 27, 2060–2066. [Google Scholar] [CrossRef] [PubMed]
  238. Vázquez-Ibarra, K.C.; López, J.Y.S.; Razo, T.D.P.; Lozano, J.R.C.; Ortiz-Tamayo, B.G.; Palafox-Mariscal, L.A.; Arreola, R.M.G.; González-García, J.R.; Ortiz-Lazareno, P.C. Metformin in combination with chemotherapy increases apoptosis in gastric cancer cells and counteracts senescence induced by chemotherapy. Oncol. Lett. 2024, 28, 457. [Google Scholar] [CrossRef] [PubMed]
  239. Asghariazar, V.; Ojarood, M.V.; Vatankhah, M.A.; Panahizadeh, R.; Haris, H.M.; Najafzadeh, N.; Khakbaz, P.; Soozangar, N.; Jeddi, F. Metformin reverses 5-FU chemoresistance by downregulating DKK1, WNT5A, and ABCB1 expressions in gastric cancer: An experimental and bioinformatic study. Naunyn Schmiedebergs Arch. Pharmacol. 2025, 398, 10155–10167. [Google Scholar] [CrossRef]
  240. Sang, J.; Tang, R.; Yang, M.; Sun, Q. Metformin Inhibited Proliferation and Metastasis of Colorectal Cancer and presented a Synergistic Effect on 5-FU. Biomed. Res. Int. 2020, 2020, 9312149. [Google Scholar] [CrossRef]
  241. Wang, S.; Lin, Y.; Zhao, Q.; Chen, H.; Du, S.; Zeng, Z. Metformin reverses 5-FU resistance induced by radiotherapy through mediating folate metabolism in colorectal cancer. Mol. Med. 2025, 31, 199. [Google Scholar] [CrossRef]
  242. Ling, S.; Tian, Y.; Zhang, H.; Jia, K.; Feng, T.; Sun, D.; Gao, Z.; Xu, F.; Hou, Z.; Li, Y.; et al. [Corrigendum] Metformin reverses multidrug resistance in human hepatocellular carcinoma Bel-7402/5-fluorouracil cells. Mol. Med. Rep. 2022, 25, 175. [Google Scholar] [CrossRef]
  243. Vakili-Ghartavol, R.; Mehrabian, A.; Mirzavi, F.; Rezayat, S.M.; Mashreghi, M.; Farhoudi, L.; Kharrazi, S.; Sadri, K.; Jaafari, M.R. Docetaxel in combination with metformin enhances antitumour efficacy in metastatic breast carcinoma models: A promising cancer targeting based on PEGylated liposomes. J. Pharm. Pharmacol. 2022, 74, 1307–1319. [Google Scholar] [CrossRef]
  244. Catapano, J.; Luty, M.; Wróbel, T.; Pudełek, M.; Piwowarczyk, K.; Kędracka-Krok, S.; Siedlar, M.; Madeja, Z.; Czyż, J. Acquired drug resistance interferes with the susceptibility of prostate cancer cells to metabolic stress. Cell Mol. Biol. Lett. 2022, 27, 100. [Google Scholar] [CrossRef] [PubMed]
  245. Yu, Z.; Wang, Y.; Wang, B.; Zhai, J. Metformin Affects Paclitaxel Sensitivity of Ovarian Cancer Cells Through Autophagy Mediated by Long Noncoding RNASNHG7/miR-3127–5p Axis. Cancer Biother. Radiopharm. 2022, 37, 792–801. [Google Scholar] [CrossRef] [PubMed]
  246. Tang, Z.; Zhang, Y.; Yu, Z.; Luo, Z. Metformin Suppresses Stemness of Non-Small-Cell Lung Cancer Induced by Paclitaxel through FOXO3a. Int. J. Mol. Sci. 2023, 24, 16611. [Google Scholar] [CrossRef] [PubMed]
  247. Al-Kofahi, T.; Altrad, B.; Amawi, H.; A Aljabali, A.; Abul-Haija, Y.M.; A Obeid, M. Paclitaxel-loaded niosomes in combination with metformin: Development, characterization and anticancer potentials. Ther. Deliv. 2024, 15, 109–118. [Google Scholar] [CrossRef]
  248. Feng, S.-W.; Chang, P.-C.; Chen, H.-Y.; Hueng, D.-Y.; Li, Y.-F.; Huang, S.-M. Exploring the Mechanism of Adjuvant Treatment of Glioblastoma Using Temozolomide and Metformin. Int. J. Mol. Sci. 2022, 23, 8171. [Google Scholar] [CrossRef]
  249. Ohno, M.; Kitanaka, C.; Miyakita, Y.; Tanaka, S.; Sonoda, Y.; Mishima, K.; Ishikawa, E.; Takahashi, M.; Yanagisawa, S.; Ohashi, K.; et al. Metformin with Temozolomide for Newly Diagnosed Glioblastoma: Results of Phase I Study and a Brief Review of Relevant Studies. Cancers 2022, 14, 4222. [Google Scholar] [CrossRef]
  250. Xia, Y.; Sun, M.; Huang, H.; Jin, W.L. Drug repurposing for cancer therapy. Signal Transduct. Target Ther. 2024, 9, 92. [Google Scholar] [CrossRef]
  251. Zhang, K.; Bai, P.; Dai, H.; Deng, Z. Metformin and risk of cancer among patients with type 2 diabetes mellitus: A systematic review and meta-analysis. Prim. Care Diabetes 2021, 15, 52–58. [Google Scholar] [CrossRef]
  252. Bakry, H.M.; Mansour, N.O.; ElKhodary, T.R.; Soliman, M.M. Efficacy of metformin in prevention of paclitaxel-induced peripheral neuropathy in breast cancer patients: A randomized controlled trial. Front. Pharmacol. 2023, 14, 1181312. [Google Scholar] [CrossRef]
  253. Martin-Castillo, B.; Pernas, S.; Dorca, J.; Álvarez, I.; Martínez, S.; Pérez-Garcia, J.M.; Batista-López, N.; Rodríguez-Sánchez, C.A.; Amillano, K.; Domínguez, S.; et al. A phase 2 trial of neoadjuvant metformin in combination with trastuzumab and chemotherapy in women with early HER2-positive breast cancer: The METTEN study. Oncotarget 2018, 9, 35687–35704. [Google Scholar] [CrossRef]
  254. Zhao, Y.; Gong, C.; Wang, Z.; Zhang, J.; Wang, L.; Zhang, S.; Cao, J.; Tao, Z.; Li, T.; Wang, B.; et al. A randomized phase II study of aromatase inhibitors plus metformin in pre-treated postmenopausal patients with hormone receptor positive metastatic breast cancer. Oncotarget 2017, 8, 84224–84236. [Google Scholar] [CrossRef]
  255. Fenn, K.; Maurer, M.; Lee, S.M.; Crew, K.D.; Trivedi, M.S.; Accordino, M.K.; Hershman, D.L.; Kalinsky, K. Phase 1 Study of Erlotinib and Metformin in Metastatic Triple-Negative Breast Cancer. Clin. Breast Cancer 2020, 20, 80–86. [Google Scholar] [CrossRef]
  256. Rocca, A.; Cortesi, P.; Cortesi, L.; Gianni, L.; Matteucci, F.; Fantini, L.; Maestri, A.; Giunchi, D.C.; Cavanna, L.; Ciani, R.; et al. Phase II study of liposomal doxorubicin, docetaxel and trastuzumab in combination with metformin as neoadjuvant therapy for HER2-positive breast cancer. Ther. Adv. Med. Oncol. 2021, 13, 1758835920985632. [Google Scholar] [CrossRef]
  257. Gennari, A.; Foca, F.; Zamarchi, R.; Rocca, A.; Amadori, D.; De Censi, A.; Bologna, A.; Cavanna, L.; Gianni, L.; Scaltriti, L.; et al. Insulin-like growth factor-1 receptor (IGF-1R) expression on circulating tumor cells (CTCs) and metastatic breast cancer outcome: Results from the TransMYME trial. Breast Cancer Res. Treat. 2020, 181, 61–68. [Google Scholar] [CrossRef]
  258. Akce, M.; Farran, B.; Switchenko, J.M.; Rupji, M.; Kang, S.; Khalil, L.; Ruggieri-Joyce, A.; Olson, B.; Shaib, W.L.; Wu, C.; et al. Phase II trial of nivolumab and metformin in patients with treatment-refractory microsatellite stable metastatic colorectal cancer. J. Immunother. Cancer 2023, 11, e007235. [Google Scholar] [CrossRef]
  259. Ahmed, J.; Stephen, B.; Khawaja, M.R.; Yang, Y.; Salih, I.; Barrientos-Toro, E.; Raso, M.G.; Karp, D.D.; Piha-Paul, S.A.; Sood, A.K.; et al. A phase I study of temsirolimus in combination with metformin in patients with advanced or recurrent endometrial cancer. Gynecol. Oncol. 2025, 193, 73–80. [Google Scholar] [CrossRef] [PubMed]
  260. Gulati, S.; Desai, J.; Palackdharry, S.M.; Morris, J.C.; Zhu, Z.; Jandarov, R.; Riaz, M.K.; Takiar, V.; Mierzwa, M.; Gutkind, J.S.; et al. Phase 1 dose-finding study of metformin in combination with concurrent cisplatin and radiotherapy in patients with locally advanced head and neck squamous cell cancer. Cancer 2020, 126, 354–362. [Google Scholar] [CrossRef] [PubMed]
  261. Arrieta, O.; Barrón, F.; Padilla, M.Á.S.; Avilés-Salas, A.; Ramírez-Tirado, L.A.; Jiménez, M.J.A.; Vergara, E.; Zatarain-Barrón, Z.L.; Hernández-Pedro, N.; Cardona, A.F.; et al. Effect of Metformin Plus Tyrosine Kinase Inhibitors Compared With Tyrosine Kinase Inhibitors Alone in Patients with Epidermal Growth Factor Receptor-Mutated Lung Adenocarcinoma: A Phase 2 Randomized Clinical Trial. JAMA Oncol. 2019, 5, e192553. [Google Scholar] [CrossRef] [PubMed]
  262. Marrone, K.A.; Zhou, X.; Forde, P.M.; Purtell, M.; Brahmer, J.R.; Hann, C.L.; Kelly, R.J.; Coleman, B.; Gabrielson, E.; Rosner, G.L.; et al. A Randomized Phase II Study of Metformin plus Paclitaxel/Carboplatin/Bevacizumab in Patients with Chemotherapy-Naive Advanced or Metastatic Nonsquamous Non-Small Cell Lung Cancer. Oncologist 2018, 23, 859–865. [Google Scholar] [CrossRef]
  263. Skinner, H.; Hu, C.; Tsakiridis, T.; Santana-Davila, R.; Lu, B.; Erasmus, J.J.; Doemer, A.J.; Videtic, G.M.; Coster, J.; Yang, A.X.; et al. Addition of Metformin to Concurrent Chemoradiation in Patients with Locally Advanced Non-Small Cell Lung Cancer: The NRG-LU001 Phase 2 Randomized Clinical Trial. JAMA Oncol. 2021, 7, 1324–1332. [Google Scholar] [CrossRef] [PubMed]
  264. IRCCS San Raffaele. Combination Chemotherapy with or Without Metformin Hydrochloride in Treating Patients with Metastatic Pancreatic Cancer, NCT Number NCT01167738. 2015. Available online: https://clinicaltrials.gov/study/NCT01167738 (accessed on 26 October 2025).
  265. Kordes, S.; Pollak, M.N.; Zwinderman, A.H.; Mathôt, R.A.; Weterman, M.J.; Beeker, A.; Punt, C.J.; Richel, D.J.; Wilmink, J.W. Metformin in patients with advanced pancreatic cancer: A double-blind, randomised, placebo-controlled phase 2 trial. Lancet Oncol. 2015, 16, 839–847. [Google Scholar] [CrossRef]
  266. Martin, M.P.; Borchiellini, D.; Thamphya, B.; Guillot, A.; Paoli, J.B.; Besson, D.; Hilgers, W.; Priou, F.; El Kouri, C.; Hoch, B.; et al. TAXOMET: A French Prospective Multicentric Randomized Phase II Study of Docetaxel Plus Metformin Versus Docetaxel Plus Placebo in Metastatic Castration-Resistant Prostate Cancer. Clin. Genitourin. Cancer 2021, 19, 501–509. [Google Scholar] [CrossRef]
  267. Goodwin, P.J.; Chen, B.E.; Gelmon, K.A.; Whelan, T.J.; Ennis, M.; Lemieux, J.; Ligibel, J.A.; Hershman, D.L.; Mayer, I.A.; Hobday, T.J.; et al. Effect of Metformin vs Placebo on Invasive Disease-Free Survival in Patients with Breast Cancer: The MA.32 Randomized Clinical Trial. JAMA 2022, 327, 1963–1973. [Google Scholar] [CrossRef]
  268. De, A.; Kuppusamy, G. Metformin in breast cancer: Preclinical and clinical evidence. Curr. Probl. Cancer. 2020, 44, 100488. [Google Scholar] [CrossRef]
  269. Jiralerspong, S.; Palla, S.L.; Giordano, S.H.; Meric-Bernstam, F.; Liedtke, C.; Barnett, C.M.; Hsu, L.; Hung, M.-C.; Hortobagyi, G.N.; Gonzalez-Angulo, A.M. Metformin and pathologic complete responses to neoadjuvant chemotherapy in diabetic patients with breast cancer. J. Clin. Oncol. 2009, 27, 3297–3302. [Google Scholar] [CrossRef] [PubMed]
  270. Rico, M.; Baglioni, M.; Bondarenko, M.; Laluce, N.C.; Rozados, V.; André, N.; Carré, M.; Scharovsky, O.G.; Márquez, M.M. Metformin and propranolol combination prevents cancer progression and metastasis in different breast cancer models. Oncotarget 2017, 8, 2874–2889. [Google Scholar] [CrossRef] [PubMed]
  271. Falah, R.R.; Talib, W.H.; Shbailat, S.J. Combination of metformin and curcumin targets breast cancer in mice by angiogenesis inhibition, immune system modulation and induction of p53 independent apoptosis. Ther. Adv. Med. Oncol. 2017, 9, 235–252. [Google Scholar] [CrossRef]
  272. Xintaropoulou, C.; Ward, C.; Wise, A.; Marston, H.; Turnbull, A.; Langdon, S.P. A comparative analysis of inhibitors of the glycolysis pathway in breast and ovarian cancer cell line models. Oncotarget 2015, 6, 25677–25695. [Google Scholar] [CrossRef] [PubMed]
  273. Farkhondeh, T.; Amirabadizadeh, A.; Aramjoo, H.; Llorens, S.; Roshanravan, B.; Saeedi, F.; Talebi, M.; Shakibaei, M.; Samarghandian, S. Impact of Metformin on Cancer Biomarkers in Non-Diabetic Cancer Patients: A Systematic Review and Meta-Analysis of Clinical Trials. Curr. Oncol. 2021, 28, 1412–1423. [Google Scholar] [CrossRef]
  274. Laskov, I.; Drudi, L.; Beauchamp, M.-C.; Yasmeen, A.; Ferenczy, A.; Pollak, M.; Gotlieb, W.H. Anti-diabetic doses of metformin decrease proliferation markers in tumors of patients with endometrial cancer. Gynecol. Oncol. 2014, 134, 607–614. [Google Scholar] [CrossRef]
  275. Wen, K.-C.; Sung, P.-L.; Wu, A.T.H.; Chou, P.-C.; Lin, J.-H.; Huang, C.-Y.F.; Yeung, S.-C.J.; Lee, M.-H. Neoadjuvant metformin added to conventional chemotherapy synergizes anti-proliferative effects in ovarian cancer. J. Ovarian Res. 2020, 13, 95. [Google Scholar] [CrossRef]
  276. Bragagnoli, A.C.; Araujo, R.L.C.; Ferraz, M.W.; dos Santos, L.V.; Abdalla, K.C.; Comar, F.; Santos, F.A.; Oliveira, M.A.; Carvalheira, J.B.C.; Cárcano, F.M.; et al. Metformin plus lrinotecan in patients with refractory colorectal cancer: A phase 2 clinical trial. Br. J. Cancer 2021, 124, 1072–1078. [Google Scholar] [CrossRef]
  277. Gutkind, J.S.; Molinolo, A.A.; Wu, X.; Wang, Z.; Nachmanson, D.; Harismendy, O.; Alexandrov, L.B.; Wuertz, B.R.; Ondrey, F.G.; Laronde, D.; et al. Inhibition of mTOR signaling and clinical activity of metformin in oral premalignant lesions. JCI Insight 2021, 6, e147096. [Google Scholar] [CrossRef] [PubMed]
  278. van Eijck, C.W.F.; Vadgama, D.; van Eijck, C.H.J.; Wilmink, J.W.; for The Dutch Pancreatic Cancer Group (DPCG). Metformin boosts antitumor immunity and improves prognosis in upfront resected pancreatic cancer: An observational study. J. Natl. Cancer Inst. 2024, 116, 1374–1383. [Google Scholar] [CrossRef] [PubMed]
  279. Kulkarni, V.S.; Alagarsamy, V.; Solomon, V.R.; Jose, P.A.; Murugesan, S. Drug Repurposing: An Effective Tool in Modern Drug Discovery. Russ. J. Bioorg. Chem. 2023, 49, 157–166. [Google Scholar] [CrossRef] [PubMed]
  280. Usha, T.; Middha, S.K.; Kukanur, A.A.; Shravani, R.V.; Anupama, M.N.; Harshitha, N.; Rahamath, A.; Kulkarni, S.S.; Goyal, A.K. Drug Repurposing Approaches: Existing Leads for Novel Threats and Drug Targets. Curr. Protein Pept. Sci. 2021, 22, 251–271. [Google Scholar] [CrossRef]
  281. Al Khzem, A.H.; Gomaa, M.S.; Alturki, M.S.; Tawfeeq, N.; Sarafroz, M.; Alonaizi, S.M.; Al Faran, A.; Alrumaihi, L.A.; Alansari, F.A.; Alghamdi, A.A. Drug Repurposing for Cancer Treatment: A Comprehensive Review. Int. J. Mol. Sci. 2024, 25, 12441. [Google Scholar] [CrossRef]
  282. Lord, S.R.; Harris, A.L. Is it still worth pursuing the repurposing of metformin as a cancer therapeutic? Br. J. Cancer 2023, 128, 958–966. [Google Scholar] [CrossRef]
Figure 1. Glucose utilization in healthy individuals, in type 2 diabetes, and cancer cells. This diagram illustrates the differences in glucose metabolism under three different physiological and pathological conditions. In healthy individuals, glucose homeostasis is tightly regulated, ensuring balanced glucose supply, demand, and energy production. In T2D, insulin resistance impairs glucose uptake in peripheral tissues, leading to disturbed regulation and hyperglycemia. In cancer cells, a reprogrammed use of glucose occurs with a preference for aerobic glycolysis (the Warburg effect), supporting rapid proliferation and biosynthetic demands.
Figure 1. Glucose utilization in healthy individuals, in type 2 diabetes, and cancer cells. This diagram illustrates the differences in glucose metabolism under three different physiological and pathological conditions. In healthy individuals, glucose homeostasis is tightly regulated, ensuring balanced glucose supply, demand, and energy production. In T2D, insulin resistance impairs glucose uptake in peripheral tissues, leading to disturbed regulation and hyperglycemia. In cancer cells, a reprogrammed use of glucose occurs with a preference for aerobic glycolysis (the Warburg effect), supporting rapid proliferation and biosynthetic demands.
Applsci 15 11576 g001
Figure 2. Inhibition of hepatic gluconeogenesis by metformin. The effects involve both redox-dependent and mitochondrial pathways. Mitochondrial glycerol-3-phosphate dehydrogenase (GPD2) is inhibited, leading to a decreased cytosolic NADH/NAD+ ratio and reduced availability of substrates such as lactate. Mitochondrial complex I is also inhibited, resulting in lower cellular ATP levels. Subsequent activation of AMPK leads to downregulation of key gluconeogenic enzymes and decreased hepatic glucose output. ↑: increased; ↓: decreased.
Figure 2. Inhibition of hepatic gluconeogenesis by metformin. The effects involve both redox-dependent and mitochondrial pathways. Mitochondrial glycerol-3-phosphate dehydrogenase (GPD2) is inhibited, leading to a decreased cytosolic NADH/NAD+ ratio and reduced availability of substrates such as lactate. Mitochondrial complex I is also inhibited, resulting in lower cellular ATP levels. Subsequent activation of AMPK leads to downregulation of key gluconeogenic enzymes and decreased hepatic glucose output. ↑: increased; ↓: decreased.
Applsci 15 11576 g002
Figure 3. Metformin interferes with metabolism, glucose homeostasis and weight control by affecting the gut microbiota, namely by increasing the relative abundance of Akkermansia, Bifidobacterium, Escherichia, and SCFA-producing bacteria, and decreasing the proportion of Intestinibacter. For more details on each mechanism, please refer to the main text. GI: gastrointestinal; GLP-1: glucagon-like peptide 1; GPR41/43: G-protein-coupled receptors 41 and 43; GUDCA: glycoursodeoxycholic acid; HbA1c: % hemoglobin A1c; PYY: peptide YY; SCFA: short-chain fatty acids; SGLT1: sodium-glucose cotransporter 1; TUDCA: tauroursodeoxycholic acid; ↑: increased; ↓: decreased.
Figure 3. Metformin interferes with metabolism, glucose homeostasis and weight control by affecting the gut microbiota, namely by increasing the relative abundance of Akkermansia, Bifidobacterium, Escherichia, and SCFA-producing bacteria, and decreasing the proportion of Intestinibacter. For more details on each mechanism, please refer to the main text. GI: gastrointestinal; GLP-1: glucagon-like peptide 1; GPR41/43: G-protein-coupled receptors 41 and 43; GUDCA: glycoursodeoxycholic acid; HbA1c: % hemoglobin A1c; PYY: peptide YY; SCFA: short-chain fatty acids; SGLT1: sodium-glucose cotransporter 1; TUDCA: tauroursodeoxycholic acid; ↑: increased; ↓: decreased.
Applsci 15 11576 g003
Figure 4. Schematic representation of the effects of metformin in cancer cells. Metformin inhibits mitochondrial complex I, increasing the AMP/ATP ratio and leading to activation of AMPK. AMPK activation suppresses mTOR, thereby promoting apoptosis, autophagy, cell cycle arrest and reducing protein synthesis, cell proliferation, migration and invasion. Metformin also inhibits MAPK, leading to increased apoptosis and decreased migration. Additionally, metformin suppresses NF-κB, resulting in reduced expression of drug resistance genes such as MDR1 (P-gp). Finally, metformin decreases HIF-1α stability/activation, inhibiting adaptive hypoxic responses (e.g., anaerobic glycolysis and angiogenesis). Black lines indicate activation and red lines indicate inhibition; arrows with ↑/↓ indicate increased/decreased processes.
Figure 4. Schematic representation of the effects of metformin in cancer cells. Metformin inhibits mitochondrial complex I, increasing the AMP/ATP ratio and leading to activation of AMPK. AMPK activation suppresses mTOR, thereby promoting apoptosis, autophagy, cell cycle arrest and reducing protein synthesis, cell proliferation, migration and invasion. Metformin also inhibits MAPK, leading to increased apoptosis and decreased migration. Additionally, metformin suppresses NF-κB, resulting in reduced expression of drug resistance genes such as MDR1 (P-gp). Finally, metformin decreases HIF-1α stability/activation, inhibiting adaptive hypoxic responses (e.g., anaerobic glycolysis and angiogenesis). Black lines indicate activation and red lines indicate inhibition; arrows with ↑/↓ indicate increased/decreased processes.
Applsci 15 11576 g004
Table 1. Effects of metformin on different cancer types in in vitro and in vivo models, focusing on its action through mitochondrial complex I inhibition and metabolic effects.
Table 1. Effects of metformin on different cancer types in in vitro and in vivo models, focusing on its action through mitochondrial complex I inhibition and metabolic effects.
Cancer TypeIn Vitro Cell Lines/In Vivo Model and Dose RangeEffects on Cancer CharacteristicsMechanismMetabolic EffectsReferences
OvarianIn vitro studies with SKOV3 and A2780 cell linesReduces cell viability and proliferation; induces apoptosis and ferroptosisInhibition of mitochondrial complex I; energy depletion; ferroptosis under stressMitochondrial dysfunction and decreased ATP production; altered glucose metabolism[130]
Head and Neck Squamous Cell Carcinoma (HNSCC)In vitro studies with FaDU, Detroit 562 and HNSCC cell lines Reduces growth and colony formation; cell cycle arrestComplex I inhibition; loss of NAD+/NADH homeostasisReduced ATP levels; increased metabolic stress markers (increased AMP/ATP ratio)[131,132]
LungIn vitro studies with BEAS-2B, 1170 and NSCLC H522, H2030, H1299, H2009, H838, A549 and H1975 cell lines
In vivo studies with A/J mice (200 mg metformin /kg by intraperitoneal injection)
Decreases proliferation and colony formation; induces apoptosis through oxidative stressComplex I inhibition causing energy depletion; increased mitochondrial ROSLower ATP production; altered lactate levels[133]
LiverIn vitro studies with LO2, HepG2, Huh-7, LX-2 cell lines and with primary human hepatocytesReduces proliferation; induces autophagy and mitochondrial dysfunction; inhibition of cell cycle progressionMitochondrial dysfunction; inhibition of mitochondrial complex I; disruption of mitochondrial voltage-dependent anion channel 1 (VDAC1)Decreased ATP production; increased lactate production; influence on metabolite levels of lactate, alanine, glycerol-3-phosphate, glycerol, glycine, and glutamate, derived from glucose[134,135,136]
BreastIn vitro studies with MCF-7, BT-474, SKBR-3 and TNBC Hs 578T and MDA-MB-231 cell linesSuppresses proliferation; activates apoptosis; reduces cell growth and colony formation; potentiates the effect of conventional anticancer drugs; cell cycle arrestDirect inhibition of complex I leading to metabolic stress;
decreased mitochondrial respiration and increased aerobic glycolysis
Increased glucose consumption and lactate production; short-term exposure to metformin reduces cellular glucose uptake, but long-term exposure leads to the opposite[137,138,139,140]
PancreaticIn vitro studies with AsPC-1, BxPC-3, PANC-1 and MIAPaCa-2 cell lines and with primary pancreatic CSC
In vivo studies with mice xenografted with pancreatic carcinomas (150 mg metformin/kg by intraperitoneal injection)
Limits tumor progression and proliferation; specifically eliminates pancreatic cancer stem cellsComplex I inhibition, reducing oxidative phosphorylationDecreased ATP production; impaired mitochondrial metabolism;
hyperglycemia protects against the metformin-induced growth inhibition
[141,142]
LeukemiaIn vitro studies with REL NALM-6, HL60, MOLM14 and U937 cell lines Induces apoptosis and energy stress; induces superoxide generation and oxidative stress; blockage of cell cycle and inhibition of cell proliferation and colony formation of leukemic cellsInhibition of mitochondrial electron transportReduced oxygen consumption and mitochondrial ATP synthesis; increased glycolysis and lactate production, pentose phosphate pathway and fatty acid metabolism[143,144]
ProstateIn vitro studies with LNCaP, P69, PC-3 and DU145 cell linesDecreases proliferation and migration; induces energetic crisis;
synergic effect with 2DG
Complex I inhibition induces energetic crisisDecreased ATP levels; metabolic crisis indicators[145]
ColorectalIn vitro studies with HCT116 p53−/− cell line
In vivo studies with mice xenografted with this cell line (drinking water containing 1–5 mg/mL of metformin)
Sensitizes cells to chemotherapy; reduces proliferation, migration, and invasionComplex I inhibition causes energy depletion and enhances chemosensitivityReduced ATP; altered glucose metabolism[146]
HNSCC, Head and neck squamous cell carcinoma NSCLC, non-small cell lung cancer; VDAC1, voltage-dependent anion channel 1 CSC; Cancer Stem Cells.
Table 2. Overview of in vitro studies on nanoparticle-based systems for metformin delivery in cancer therapy, including the type of nanoparticle used, composition/carrier, target/indication, purpose, and mechanism/outcome.
Table 2. Overview of in vitro studies on nanoparticle-based systems for metformin delivery in cancer therapy, including the type of nanoparticle used, composition/carrier, target/indication, purpose, and mechanism/outcome.
Nanoparticle TypeComposition/CarrierTarget/IndicationPurposeMechanism/OutcomeReferences
Functionalized polymeric NPsFolate-PLGA-PEGBreast cancer (MDA-MB-231)Targeted delivery, enhanced apoptosis and gene modulationThe effect of metformin was enhanced through increased cellular internalization, activation of the AMPK pathway, and inhibition of the mTOR pathway, promoting apoptosis and cell cycle arrest[211]
Functionalized chitosan NPsWZB117-OCMCBreast cancerTarget GLUT1-overexpressing breast cancer cells and enhance metabolic disruption via dual deliveryDual inhibition of GLUT1 and mTOR pathways enhances metformin uptake, reduces glycolysis, and promotes apoptosis and cell cycle arrest[209]
O-CMC (O-carboxymethyl chitosan)Pancreatic cancerImprove intracellular retention and controlled release of metforminNPs enhance metformin’s cellular uptake and retention, enabling sustained drug release and increased apoptosis in pancreatic cancer cells[215]
Polymeric NPsPLGALiver cancer (HepG2 cells)Deliver metformin efficiently to hepatocellular carcinoma cells and enhance radiosensitizationNPs enhance metformin’s antitumor effect by blocking autophagy, inhibiting mTOR/p53/HIF1A signaling, and inducing cell cycle arrest in HepG2 cells[212]
Dual gold NP systemMET-GNPs + COL-GNPsBreast cancer spheroids Facilitate deep tumor penetration and selective delivery of metformin in dense tumor ECMNPs enhance tumor penetration by degrading ECM and improve metformin delivery, leading to increased apoptosis in breast cancer spheroids[210]
Hybrid NPsChitosan + Silver NPsBreast cancer (radiation therapy)Radiosensitization enhancement during radiotherapyNPs enhance radiosensitization by increasing oxidative stress, impairing DNA repair, and promoting apoptosis in breast cancer cells[216]
Carbon-based NPsActivated Carbon NPsHepatocellular CSCs Target hepatocellular CSCs with sustained metformin delivery and selective cytotoxicity Activated carbon nanoparticles improve metformin delivery to hepatocellular CSCs, enhancing apoptosis and suppressing proliferation via AMPK activation and mTOR inhibition[217]
Topical polymeric NPsHA-coated chitosan/gelatinMelanoma (topical)Enable topical delivery of metformin to melanoma cells with enhanced skin penetration and retentionHA-coated chitosan/gelatin NPs enhance topical metformin delivery by improving skin penetration, cellular uptake, and cytotoxicity against melanoma cells[213]
Functionalized mesoporous silica NPsHA-coated MSNsLung cancer (A549)Target CD44-positive lung cancer cells via HA-modified silica NPs for selective metformin deliveryHA-modified mesoporous silica NPs enhance metformin delivery to A549 cells via CD44 targeting, promoting AMPK activation, mTOR inhibition, and apoptosis[214]
Lipid-based NPsLecithinColorectal cancerImprove metformin bioavailability and enable epigenetic modulation Lecithin NPs enhance metformin’s cytotoxicity in colorectal cancer cells via epigenetic modulation of noncoding RNAs, leading to apoptosis and reduced cell viability[218]
Colon cancerEnhance metformin stability and sustained release for colorectal cancer therapyNPs improve metformin’s stability and sustained release, enhancing cellular uptake and cytotoxicity against colorectal cancer cells.[219]
Functionalized selenium NPsTW80-SeNPs + MetforminBreast cancer (MCF-7)Combine selenium NPs with metformin to synergistically target breast cancer cellsNPs synergize with metformin by modulating selenoproteins, increasing oxidative stress, and promoting apoptosis in breast cancer cell[220]
Gold NPsAuNPs + MetforminBreast and lung cancer (MCF-7, A549)Evaluate synergistic antitumor effects of metformin and gold NPs Nps synergize with metformin to enhance apoptosis, modulate BAX/BCL2 expression, and PI3K/Akt/mTOR pathway inhibition, reducing viability in breast and lung cancer cells[208]
PLGA, Poly(lactic-co-glycolic acid); PEG, Polyethylene glycol; OCMC, O-carboxymethyl chitosan; NPs, nanoparticles; GNPS, Gold nanoparticles; MET-GNPS, Metformin-loaded gold nanoparticles; COL-GNPs, Collagen-coated gold nanoparticles; HA, Hyaluronic acid; MSNs, Mesoporous silica nanoparticles; CSCs, Cancer Stem Cells; TW80, Tween 80 (Polysorbate 80); SeNPs, Selenium nanoparticles; AuNPs, Gold nanoparticles; AMPK, AMP-activated protein kinase; mTOR, Mechanistic target of rapamycin; PI3K, Phosphoinositide 3-kinase; Akt, Protein kinase B; ROS, Reactive oxygen species; ECM, Extracellular matrix.
Table 3. Effect of metformin combinations with chemotherapeutic agents in cancer therapy and respective mechanism of synergy.
Table 3. Effect of metformin combinations with chemotherapeutic agents in cancer therapy and respective mechanism of synergy.
Chemotherapeutic AgentCancer TypeObserved EffectMechanism of SynergyReferences
Carboplatin + PemetrexedAdvanced NSCLC, non-squamousORR of 23%, PFS of 3.9 months, OS of 11.7 months. No significant improvement over historical controlsPotential metabolic modulation via AMPK activation and mTOR inhibition. No LKB1/STK11 mutations identified in patients[221]
CisplatinNSCLCMetformin sensitized p53 wild-type NSCLC cells (A549, HCC827) to cisplatin; no effect in p53-null cells (H1299, H358)Chemosensitization is p53-dependent; inhibited by Jarid1b overexpression. Metformin alters p53 localization to mitochondria and reverses cisplatin-induced resistance[227]
Metformin restored cisplatin sensitivity in resistant cells (A549); reduced cell viability and colony formation; enhanced apoptosisInhibition of mTOR signaling; modulation of apoptosis, oxidative stress, and G2/M transition; regulation of transcription and RNA processing pathways[228]
Triple-negative breast cancer (MDA-MB-231 cells)Cell viability reduced to ~25.9% with 30 µM cisplatin + 5 mM metformin after 24 h; enhanced antiproliferative effect compared to drugs aloneElectrical pulses enhanced drug uptake; metformin suppressed glucose metabolism and increased ROS, promoting apoptosis [225]
HNSCCMetformin reversed cisplatin resistance; enhanced apoptosis and reduced proliferation in resistant cellsTranscriptomic analysis revealed modulation of base excision repair pathway and other DNA damage response genes; reduced self-repair capacity after chemotherapy[226]
Oral squamous cell carcinoma Synergistic reduction in cell viability and proliferation; increased apoptosis and ROS; inhibition of EMT and migrationActivation of AMPK pathway; inhibition of EMT; confirmed by reversal with AMPK inhibitor (Compound C)[229]
Synergistic inhibition of cell proliferation and increased apoptosis in chemoresistant cellsMetformin downregulates cancer stemness via suppression of KLF4 expression, enhancing cisplatin sensitivity[230]
Nasopharyngeal carcinoma (Sune-1 cells and xenografts)The combinations significantly inhibited cell proliferation and tumor growthSynergistic inhibition via distinct anticancer mechanisms; metformin contributed to metabolic stress and enhanced cisplatin efficacy[231]
Ovarian cancerCombination reduced apoptosis and DNA damage; induced resistance to cisplatinMetformin activated the ATM/CHK2 pathway, leading to upregulation of Rad51, which enhanced DNA repair and reduced cisplatin efficacy[232]
Synergistic inhibition of cell viability, proliferation, and colony formation; enhanced apoptosis and S-phase arrestDownregulation of pluripotency factors (Oct-4, Sox2, Nanog); inhibition of Hedgehog signaling; suppression of MDR1 and ERCC1; modulation of autophagy and DNA damage response[233]
Gastric cancer Metformin reduced cisplatin cytotoxicity; increased cell survival and resistanceAMPK activation induced PINK1/Parkin-dependent mitophagy, promoting mitochondrial quality control and reducing apoptosis[234]
Metformin reduced viability and confluence of resistant cells; induced both autophagy and apoptosisInduction of autophagy markers (Atg5, Atg12, LC3-II); increased caspase-3 and -7 activity; effect reversed by 3-MA, confirming autophagy involvement[235]
Increased chemosensitivity to cisplatin; reduced cell viability and metabolic activity; enhanced apoptosis and oxidative stressDownregulation of Nrf2; activation of p53 and AMPK pathways; reversal of effects by Nrf2 overexpression or AMPK/p53 inhibition[236]
DoxorubicinBladder cancer (T24 cells)Synergistic inhibition of metastatic potential; reduced migration and invasionLikely via modulation of EMT and metabolic stress pathways[223]
Breast cancer (4T1 mouse model)Enhanced antitumor efficacy; increased CD8+ T cell frequency in tumor microenvironment; reduced toxicityImmunomodulation via increased CD8+ T cell infiltration; modulation of HIF-1α and STAT3 expression; reduced doxorubicin toxicity[237]
Breast cancer (DMBA-induced in rats)Reduced tumor incidence and volume; improved survival; reduced organ toxicityEnhanced antioxidant and anti-inflammatory activity; reduced Ki67 expression; decreased IL-6, IL-1β, and NF-κB levels[224]
5-FluorouracilGastric cancerInhibition of tumor colony formation; reduced expression of Gli1 and TWIST1Suppression of the Shh/Gli1 signaling pathway; EMT modulation[222]
Increased apoptosis; reduced chemotherapy-induced senescenceLoss of mitochondrial membrane potential; activation of caspase-dependent apoptosis pathways[238]
Reversal of 5-FU resistance; increased apoptosis and cytotoxicityDownregulation of DKK1, WNT5A, ABCB1 (MDR1), P-gp, and CD44; inhibition of drug resistance pathways[239]
Colorectal cancerInhibition of proliferation and metastasis; enhanced anti-tumor effect in vivoSynergistic suppression of tumor growth and invasion; possibly via modulation of cell cycle and apoptosis pathways[240]
Reversal of 5-FU resistance induced by radiotherapy; enhanced chemosensitivityRestoration of folate metabolism via HM13-GGH-5,10-CH2-THF axis; modulation of gene expression linked to drug resistance[241]
Hepatocellular carcinomaReversal of multidrug resistance; enhanced anti-proliferative and pro-apoptotic effectsActivation of AMPK/mTOR pathway; suppression of HIF-1α; downregulation of P-gp and MRP1[242]
DocetaxelGastric cancerInhibition of tumor colony formation; reduced expression of Gli1 and TWIST1Suppression of the Shh/Gli1 signaling pathway; EMT modulation[222]
Breast cancer (metastatic)Enhanced antitumor efficacy; prolonged survival in vivoDual targeting of tumor bulk and cancer stem-like cells; improved delivery via PEGylated liposomes[243]
Prostate cancerReduced proliferation, motility, and viability in wild-type cells; partial reversal of resistanceInactivation of ABC drug transporters; metabolic imbalance; Warburg effect induction; lineage-specific EMT responses[244]
PaclitaxelOvarian cancerReversal of paclitaxel resistance; reduced cell viability, migration, and invasionInhibition of SNHG7/miR-3127-5p axis; suppression of autophagy; promotion of apoptosis[245]
NSCLCSuppression of cancer stemness; reduced self-renewal and tumorigenicity in resistant cellsActivation of AMPK; upregulation of FOXO3a; inhibition of Akt and MEK pathways; downregulation of stemness markers (c-MYC, Oct4, Nanog, Notch)[246]
Breast cancer (T47D cells)Significant reduction in IC50 of paclitaxel when combined with metformin; enhanced cytotoxicityEnhanced cellular uptake via niosomes; metformin-induced sensitization; potential reduction in required paclitaxel dose[247]
TemozolomideGlioblastomaEnhanced inhibition of proliferation and migration; increased apoptosisSuppression of mitochondrial biogenesis, EMT, and MGMT expression; increased
ROS production and mitochondrial dysfunction
[248]
Well-tolerated combination; promising progression-free survival and overall survivalActivation of AMPK-FOXO3a pathway; differentiation of glioma stem-like cells; suppression of tumor formation[249]
NSCLC, Non-small cell lung cancer; ORR; overall response rate; PFS, progression-free survival; OS, overall survival; AMPK, AMP-activated protein kinase; NPC, nasopharyngeal carcinoma; ROS, reactive oxygen species; HNSCC, head and neck squamous cell carcinoma; EMT, epithelial–mesenchymal transition; ATM/CHK2, ataxia telangiectasia mutated/checkpoint kinase 2; MDR1, multidrug resistance protein 1; P-gp, P-glycoprotein; MRP1, multidrug resistance-associated protein 1; STAT3, signal transducer and activator of transcription 3; HIF-1a, hypoxia-inducible factor 1-alpha; shh, sonic hedgehog; Gli1, GLI family zinc finger 1; FOXO3a, forkhead box O3; MEK, mitogen-activated protein kinase.
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

Barbosa, F.; Cunha, A.; Barbosa, J.; Faria, J.; Queirós, O. The Dual Role of Metformin: Repurposing an Antidiabetic Drug for Cancer Therapy. Appl. Sci. 2025, 15, 11576. https://doi.org/10.3390/app152111576

AMA Style

Barbosa F, Cunha A, Barbosa J, Faria J, Queirós O. The Dual Role of Metformin: Repurposing an Antidiabetic Drug for Cancer Therapy. Applied Sciences. 2025; 15(21):11576. https://doi.org/10.3390/app152111576

Chicago/Turabian Style

Barbosa, Flávia, Andrea Cunha, Joana Barbosa, Juliana Faria, and Odília Queirós. 2025. "The Dual Role of Metformin: Repurposing an Antidiabetic Drug for Cancer Therapy" Applied Sciences 15, no. 21: 11576. https://doi.org/10.3390/app152111576

APA Style

Barbosa, F., Cunha, A., Barbosa, J., Faria, J., & Queirós, O. (2025). The Dual Role of Metformin: Repurposing an Antidiabetic Drug for Cancer Therapy. Applied Sciences, 15(21), 11576. https://doi.org/10.3390/app152111576

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