Download PDF
Review  |  Open Access  |  29 Sep 2024

Interconnections within the tumor microenvironment: extracellular vesicles as critical players of metabolic reprogramming in tumor cells

Views: 124 |  Downloads: 22 |  Cited:  0
J Cancer Metastasis Treat 2024;10:28.
10.20517/2394-4722.2024.78 |  © The Author(s) 2024.
Author Information
Article Notes
Cite This Article

Abstract

Metabolic reprogramming is an intrinsic characteristic of cancer, contributing to its establishment and progression, survival, high proliferation rates, and increased migratory and invasive potential; tumor cells establish an intimate relationship with the surrounding microenvironment, where sustained communication allows the stromal fraction of the tumor microenvironment (TME) to supply energetic substrates and facilitate the biosynthesis of macromolecules, thereby promoting tumor progression. In this context, extracellular vesicles (EVs) emerge as potential communication vehicles, carrying inside content reflecting the cellular environment of origin and thus modulating the phenotype of recipient cells. The potential of EVs as modulators in the TME has been highlighted and is now consensual; however, most available articles have focused on revealing the effect of EVs in modulating tumor phenotypes and signaling pathways in tumor cells. Regarding the metabolic modulation sustained by EVs, studies have demonstrated the role of cancer cells’ EVs as modulators of surrounding cells, like immune cells, fibroblasts, and adipocytes. Therefore, this review aims to: i. highlight the most recent studies evaluating the role of cellular vesicles released by those cells within the microenvironment in the metabolic reprogramming of cancer cells; ii. compile scientific evidence proposing how EVs could modulate the metabolic profile of tumor stem cells and lymphocytes, particularly given the lack of studies focused on such approaches; and iii. highlight possible effects of vesicles, as the metabolic modulation induced by these vesicles could have anticancer potential.

Keywords

Metabolism, cancer, extracellular vesicles, tumor microenvironment, and metabolic reprogramming

INTRODUCTION

In early 2000, Hanahan and Weinberg provided a solid foundation for understanding the complex biology of cancer in a seminal review article titled “The Hallmarks of Cancer,” which comprised six essential alterations in cell physiology - sustaining proliferative signaling, evading growth suppressors, resisting cell death, enabling replicative immortality, inducing angiogenesis, and activating invasion and metastasis - thus enabling a conceptual framework for understanding the diversity of neoplastic diseases[1]. In 2011, these six acquired characteristics were expanded, incorporating two emerging Hallmarks: energy metabolism and evading immune destruction. Furthermore, a new concept acknowledging tumors as a complex microenvironment composed of multiple distinct cell types interacting, the “Tumor Microenvironment” (TME), has emerged. The TME is constructed along the tumorigenesis and comprises cancer and Cancer stem cells (CSCs), endothelial cells, pericytes, immune-inflammatory cells, cancer-associated fibroblasts (CAFs)[2], and various additional tissue-resident cell types, such as adipocytes[3], as well as the transformed parenchyma and associated stroma[4]. Within the TME, the tumor-supportive environment depends on the heterotypic interactions between cancer cells and resident or recruited noncancerous cells. This signaling network provides intercellular communication, including cell-cell contact and paracrine signaling, through the release of cytokines, chemokines, growth factors, proteases[5], and extracellular vesicles (EVs)[6].

Furthermore, this intricate network, which favors uncontrolled cell proliferation, relies on increased energy production and the synthesis of macromolecules. Thus, malignant cells often reprogram their biochemical pathways to allow rapid absorption and degradation of nutrients to cope with increased metabolic stress, contributing to the disease’s transformation, maintenance, and progression[7]. In this scenario, it is now recognized that cancer cells undergo metabolic reprogramming, thus getting adapted to intrinsic or extrinsic cues from the microenvironment due to high flexibility in nutrient acquisition and utilization[8].

The birth of cancer metabolism research dates to the 1920s, when Otto Warburg et al. observed that tumors were taking up enormous amounts of glucose compared to what was seen in the surrounding tissue[7,9] and that even in the presence of oxygen, glucose was fermented to produce lactate, a process known as aerobic glycolysis or Warburg Effect[10].

A general view of the cancer metabolism

Metabolic reprogramming allows cancer cells to adapt to intrinsic or extrinsic cues from the microenvironment through plasticity and high flexibility in nutrient acquisition and utilization[8]. Such adaptation allows nutrient acquisition and utilization to support increased proliferation, migration, and invasiveness. This fine-tuning balance between aerobic glycolysis and oxidative phosphorylation - the so-called metabolic reprogramming - occurs differently depending on the tumor type, even in cells within the same tumor[11]. In this context, tumors can broadly be classified into two main groups regarding (adenosine triphosphate) ATP sources: those that primarily rely on glycolytic and those that undergo preferential oxidative phosphorylation[12,13].

Glycolysis

Several oncogenes and tumor suppressors drive the metabolic reprogramming underlying the Warburg effect. In this scenario, the Hypoxia-inducible factor 1-alpha (HIF-1α) is an essential protein for sustained glycolysis. In noncancerous cells undergoing hypoxia, HIF-1α stabilizes, thus promoting glycolysis and suppressing oxidative phosphorylation[14]. This sustained glycolysis is mediated by the upregulation of glucose transporter (GLUT) 1 and hexokinase (HK) 2, triggering increased uptake and retention of glucose inside the cells, respectively[15]. In cancer cells, HIF-1α is stabilized even in the presence of oxygen in response to several oncogenes like protein kinase B (PK24B/AKT), phosphoinositide 3-kinase (PI3K), Ras, and Von Hippel-Lindau (VHL)[16].

The p53 tumor suppressor protein p53 also has a critical role in limiting glycolysis, though inhibiting transcription of GLUT1, GLUT4, and GLUT12 and the glycolytic enzymes HK1, HK2, glucose-6-phosphate isomerase (GPI), phosphoglucomutase (PGM), and β-enolase; furthermore, p53 can inhibit the rate-limiting step in glycolysis through the inhibition of phosphofructokinase-1 (PFK1) activity, or by inhibiting the transcription of its regulator, the bifunctional enzyme phosphofructo-2-kinase//fructose-2,6-biphosphatase 3 (PFKFB3)[17].

Besides glycolysis, glycogenolysis provides an energy source for tumors under nutrient deprivation conditions independently of cellular replication. Several cancer types display an upregulation in glycogen metabolism, such as breast, bladder, brain, ovarian, renal, skin, and uterine[18]. Such an increase in glycogen accumulation promotes the survival of cancer cells in hypoxic conditions. It occurs via the AKT pathway[19] and the increased expression of the protein phosphatase 1 regulatory subunit 3C (PPP1R3C) in a HIF-1α-dependent manner[20]. Concurrently, suppressing glycogen synthase kinase 2 (GSK2) activity triggered a reduction in prostate tumor growth in vivo[21]. Among the glucogenic enzymes playing essential roles in cancer metabolism, both phosphoenolpyruvate carboxykinase 1 (PCK1) and phosphoenolpyruvate carboxykinase 2 (PCK2) can be highlighted; such enzymes can be inhibited by p53, a tumor suppressor[22].

One candidate in this preferential energetic status is the M2 isoform of pyruvate kinase (PKM2), a rate-limiting enzyme in glycolysis, which promotes mitochondrial fusion by interacting with mitofusin 2 (MFN2), attenuating glycolysis and mediating PFK1 degradation, triggering glycolysis inhibition[23]. The inhibition of PFK1 can channel glycolytic carbon into the pentose phosphate pathway (PPP), generating ribose-5-phosphate and Nicotinamide Adenine Dinucleotide Phosphate (NADPH), other than promoting mitochondrial oxidative phosphorylation[13].

Oxidative phosphorylation

Under increased aerobic glycolysis, the excretion of carbon as lactate diminishes glucose’s contribution as an anaplerotic repository for biosynthetic pathways. Glutamine and other intermediates can replenish tricarboxylic acid cycle (TCA), thus emerging as an essential source of biosynthetic precursors. This allows the biosynthesis of nucleic acids, amino acids, and lipids, highlighting mitochondrial importance[10,24].

Oxidative phosphorylation (OXPHOS) yields a higher ATP production per glucose unit than glycolysis. It is the preferential ATP source in many types of cancer, being upregulated compared to adjacent normal cells[25], thus contradicting the long-standing belief that OXPHOS is downregulated in cancer.

Due to their heterogeneity, tumors demonstrate a wide range of metabolic phenotypes and flexibility[26]. For instance, gliomas exhibit both OXPHOS and glycolic characteristics, depending on lactate dehydrogenase (LDH) isoforms. OXPHOS is also upregulated in leukemias, lymphomas, pancreatic ductal adenocarcinoma[27], breast cancer, and classical Hodgkin lymphoma. In breast cancer, for example, the activity of Complex I, II, and IV proteins from the electron transport chain (ETC) are upregulated compared to subjacent normal epithelial cells[28].

The tricarboxylic acid cycle

The TCA cycle is a convergence point in the cellular respiration machinery; it integrates a myriad of fuel sources such as glucose, glutamine, and fatty acids. Such cycles produce intermediates required for macromolecule biosynthesis and NADH and FADH2 coenzymes committed to ETC-reducing reactions. Regarding the source of carbons replenishing the TCA intermediates, it is widely accepted that cancer cells shunt pyruvate away from the TCA depending on glutamine and fatty acids[7,29,30].

It is now widely acknowledged that mutations in enzymes of the TCA cycle, such as isocitrate dehydrogenase (IDH), succinate dehydrogenase (SDH), fumarate hydratase (FH), and aconitase are prevalent in a wide variety of human cancers[7,25,31,32]. For instance, SDH mutations in tumors have been linked to increased aggressiveness and high proliferative rates[33]. These mutations have been associated with several types of cancer, including renal cell carcinoma (RCC), neuroblastoma, gastrointestinal stromal tumors, thyroid cancer, and testicular seminoma[34,35]. Similarly, FH mutations have been reported in uterine fibroids, hereditary leiomyomatosis, and papillary RCC[36]. The reduced expression of FH, leading to fumarate accumulation, results in SDH dysfunction in some cancers, leading to the accumulation of succinate and fumarate, ultimately stabilizing HIF-1α due to the inhibition of prolyl 4-hydroxylases (PHDs). Concurrently, mutations in the IDH, an enzyme converting isocitrate to α-KG, have been observed. The abnormal expression and activity of IDHs result in the loss of the enzyme’s ability to catalyze the conversion of isocitrate to α-KG; instead, it gains a new ability to facilitate the reduction of α-KG to D-2-hydroxyglutaric acid (D-2HG), an oncometabolite[16].

Glutamine is an essential nutrient source that feeds into the TCA cycle in numerous cancer types, including MYC-driven cancers[37]. It traverses the cell membrane through amino acid transporters, ASCT2 (alanine, serine, cysteine transporter 2), and system N transporter SN2[38]. Under high proliferative rates, glutamine is often the primary source of energy and is also committed to biosynthesis, a condition known as glutamine addiction. This amino acid is the most abundant within the bloodstream, is converted in α-KG by glutaminases and glutamate dehydrogenase (GDH), fueling the TCA cycle[39], acts as a nitrogen donor to the synthesis of purines and pyrimidines[40], and can be further converted to glutathione, by glutamate-cysteine ligase and glutathione synthetase[41]. Due to its role in neutralizing mitochondrial reactive oxygen species (ROS), it is crucial to cancer cells, which become devastated under the inhibition of the glutamine metabolism due to ROS overproduction[42].

Lipid metabolism

The role of lipid metabolism in tumorigenesis has also received increased attention in recent years[43]. The importance of lipogenesis to cancer cells is undeniable. Some tumor cells show increased levels of enzymes involved in GA synthesis. One example is ATP-citrate lyase (ACLY), whose overexpression in lung adenocarcinoma and acute myeloid leukemia (AML) is directly related to a worse prognosis[44]. The stabilization of ACLY levels is a necessary mechanism for maintaining the viability of lung tumor cells. Any disturbance in this stabilization, such as through the administration of CUL 3, a protein that regulates the ubiquitination of ACLY, leads to a reduction in lipid levels, cell proliferation, and xenograft tumor growth in mice[45].

However, the importance of fatty acid oxidation (FAO) in cancer aggressiveness remains: Are FAO enzymes as dysregulated as those committed in glycolysis and glutaminolysis? An argument against the oncogenic involvement of FAO is that lipogenesis and FAO are mutually exclusive processes coordinated by the level of malonyl-CoA. Malonyl-CoA, an intermediate of fatty acid biosynthesis, acts as an allosteric inhibitor of the FAO rate-setting enzyme carnitine palmitoyltransferase (CPT) 1, presumably preventing FAO from co-occurring with active lipogenesis[46]. Furthermore, little evidence suggests that the FAO is reprogrammed in cancer due to the activation of specific oncogenes or loss of tumor suppressors.

On the other hand, although FAO enzyme mutations are not frequent in cancer, studies revealing its overexpression highlight its role in malignancy. The overexpression of CD36, CPT1A, CPT1B, CPT1C, CPT-2, carnitine transporter CT2, and Acyl-CoA synthetase long-chain 3 has been reported. Additionally, the high activity of FAO has been observed in K-Ras mutant lung cancer, triple-negative breast cancer (TNBC), AML, hepatitis B-induced hepatocellular carcinoma, glioma, and low-grade astrocytoma[11].

Camarda et al. revealed that triple-negative breast cancer cells overexpressing the oncogenic transcription factor c-Myc presented an increase in FAO enzymes and metabolic intermediates. In the presence of etomoxir, an FAO inhibitor, the tumorigenic potential was inhibited, thus highlighting the crucial role of FAO for triple-negative breast cancer cells[47]. In the same line, Wang et al. revealed that FAO inhibition selectively eliminated breast CSCs and increased the cells’ sensitivity to paclitaxel chemotherapy, being an essential mechanism for chemoresistance[48].

Some tumor cells show increased levels of enzymes involved in fatty acid synthesis. One example is ACLY, whose overexpression in lung adenocarcinoma and AML is directly related to a worse prognosis[44]. It has already been demonstrated that the stabilization of ACLY is a necessary mechanism for maintaining the viability of lung tumor cells. The administration of CUL 3, which disturbs the stabilization of ACLY through its ubiquitination, reduces lipid levels, cell proliferation, and xenograft tumor growth in mice[45].

EVs within the TME

Previously, studies focusing on the communication between cancer cells’ interaction with surrounding cells were limited to soluble mediators. However, EVs have become crucial to tumor formation, progression, and metastasis[49]. Evidence has shown that EVs can facilitate and sustain the bidirectional communication between tumor cells and their microenvironment, which is crucial for tumor occurrence and progression. EVs can be secreted by virtually all cells. Initially, such an effect was adopted to refer to ways cells undergo to eliminate unneeded components[50]. However, it is currently known that EVs sustain cell-to-cell communication by exchanging DNA, mRNAs, miRNAs, proteins, and lipids in a paracrine or systemic manner. The EVs cargo reflects the physiological or pathological microenvironment of its donor cell, allowing cell-to-cell communication through the exchange of its content, thus harboring the potential to regulate proliferation, survival, and immune effector status in recipient cells, through changes in intracellular signaling[49,51,52]. This potential of EVs to regulate critical aspects of cell behavior offers hope for novel therapeutic strategies in cancer.

The term “EVs” is a broad category that includes membrane vesicles of various sizes and origins. These vesicles are classified into main categories based on their biogenesis and size. By origin, exosomes are derived from the endosomal system, while microvesicles originate from the plasma membrane. By size, EVs are further categorized as small EVs (sEVs, ranging from 40 to 150 nm) and medium/large EVs (m/lEVs, ranging from 150 to 1,000 nm) according to the guidelines of the International Society for Extracellular Vesicle (ISEV)[53].

Receptors and ligands on the surface of the EV facilitate specific targeting for biodistribution and can trigger signaling changes within the recipient cell[54]. For example, a breakthrough study conducted by Peinado et al. demonstrated the role of EVs released by high metastatic melanomas under the reprogramming of the bone marrow progenitors-derived cells through the Met tyrosine kinase receptor, thus underscoring EV’s role upon metastatic dissemination[55].

EVs are essential in communicating cancer cells with surrounding cells within the TME. In this scenario, it is crucial to highlight the bidirectional nature of such communication, which requires active engagement from researchers. For example, when cancer cells undergo a metabolic shift toward an OXPHOS profile, TME cells release lactate - as a result of enhanced glycolysis - thus replenishing carbons to the TCA cycle[56]. Other than that, TME cells also release amino acids and lipids, thus replenishing the TCA and donating build blocks to biosynthesis, favoring cancer progression[57]. This bidirectional communication, where cancer cells influence the TME and vice versa, presents a dynamic and challenging research field for cancer biologists and therapeutics professionals, stimulating their eagerness to contribute to this evolving field.

While it is widely accepted that EVs can modulate tumor progression, our understanding of how EVs and their cargo influence metabolic reprogramming in the TME remains limited. In this field, most studies have focused on the role of cancer cell-derived EVs in altering the metabolic profile of TME surrounding cells. In this paper, we highlight cancer metabolism focusing on EVs; furthermore, we have gathered recent evidence on the role of EVs released by TME stromal cells as metabolic modulators in cancer cells, thereby revealing another important communication pathway. Furthermore, considering recent studies, we discuss how EVs could act as modulators of tumor progression and therapeutics.

EVs and the metabolic reprogramming within the TME

EVs play a bidirectional role in the communication between cancer and stromal cells within the TME. EVs are now recognized as critical signal transmitters committed to fine-tuning glucose and lipid metabolism processes. In this scenario, several studies have highlighted how EVs from cancer cells impact the metabolism of non-cancer cells within TME. It was already reported that exosomes from human primary and metastatic colorectal cancer (CRC) cells upregulate GLUT1 expression in fibroblasts in a caveolin-1-dependent manner[58]. Furthermore, miR-155- and miR-210-enriched exosomes from melanoma cells (Mel 1) increased aerobic glycolysis while decreasing OXPHOS in human dermal fibroblasts, thus leading to extracellular acidification[59]. Increased glucose and glutamine metabolism was already observed in fibroblasts exposed to EVs from breast cancer cells in a miR-105-dependent manner[60]; concurrently, both aerobic glycolysis and autophagia are increased in fibroblasts turned into CAFs upon exposure to EVs from nasopharyngeal carcinoma via latent membrane protein 1 (LMP1) nuclear factor kappa B (NF-κB)/p65 pathway[61].

Exosomes from hypoxia-induced tumor cells transfer let-7a miRNA to bone marrow-derived macrophages, enhancing OXPHOS by inhibiting the insulin-Akt-mTOR signaling pathway and thus inducing an M2-like macrophage polarization[62]. Sagar et al. demonstrated that endothelial cells exposed to exosomes from AML cells - enriched in vascular endothelial growth factor (VEGF)/VEGFR- displayed a glycolytic profile and undergo glycolysis. Adipocytes exposed to exosomes from Lewis Lung Cancer (LLC) exhibited a lower content of lipid droplets and high levels of glycerol release. Concurrently, exosomes from pancreatic cancer cells induced lipolysis in adipocytes via the ERK1/2 and MAPK p38 signaling pathways[63].

Furthermore, it was already observed that ARG1-containing EVs from ovarian carcinoma suppress T cell proliferation[64]; EVs from cervical squamous cell carcinoma (CSCC) containing miR-142-5p induce the expression of indoleamine 2,3-dioxygenase (IDO) by lymphatic endothelial cells, which in turn trigger CD8+ T cell suppression and exhaustion[65].

While most existing studies have concentrated on the influence of EVs from cancer cells on the metabolism of non-cancer cells in the TME, this review takes a fresh perspective. We delve into the less-explored territory of how EVs from non-cancer cells in the TME stimulate metabolic and phenotypic alterations in cancer cells. We underscore the pivotal role of EV bioactive charges in these processes, offering a novel angle for your consideration.

EMERGING ROLES OF EVS FROM SURROUNDING TME IN MODULATING CANCER CELL METABOLISM

Macrophage-derived EVs

As phagocytic cells, macrophages play a crucial role in maintaining tissue homeostasis and act as immunological sentinels during infections and in tissue abnormalities such as tumor growth. These cells, distributed in all organs and tumors, exhibit a polarization spectrum dependent on the microenvironment. The extremes of this spectrum are the M1, or classically activated, and M2, or alternatively activated. Tumor-associated macrophages (TAMs), the most abundant leukocytes in the TME, are classified into tumor-killing (M1) or tumor-promoting (M2). Although most TAMs have the M2-like phenotype, they can acquire a range of profiles in response to stimuli, constituting a heterogeneous and plastic population[66].

Given that TAMs are the most prevalent leukocytes in the TME[67], significant efforts have been directed toward understanding their role in tumor progression. Research has consistently shown that TAMs are associated with a poor prognosis in various cancer types, including colorectal[68], bladder[69], prostate[70], pancreas[71], and breast cancer[72]. This is primarily due to their ability to induce epithelial-mesenchymal transition (EMT)[73], angiogenesis[71], invasion[68], and, most importantly, alter the metabolism of tumor cells[74].

Recent studies have reported that metabolic changes in tumor cells promoted by macrophages are EVs-mediated. Chen et al. demonstrated that EVs derived from TAMs increase the aerobic glycolysis of breast cancer cells. Peripheral blood monocytes (PBDMs) were isolated from breast cancer patients or healthy donors and treated with a human breast adenocarcinoma cell line (MDA-MB-231) conditioned medium (CM) to obtain TAMs. Then, EVs derived from TAMs were isolated and used to treat MDA-MB-231 cells. TAM EVs enhanced glucose consumption, lactate production, and apoptosis resistance of tumor cells. These effects occurred due to the delivery of HISLA, a long non-coding RNA (lncRNA), which, upon binding to prolyl hydroxylase domain protein 2 (PDH2), decreases the hydroxylation and degradation of HIF-1α, stabilizing it. Furthermore, MDA-MB-231 cells were inoculated into the mammary fat of NOD/SCID mice. This model observed that the intratumoral injection of TAM EVs reduced the response to chemotherapy, enhanced 18F-fluorodeoxyglucose (18FDG) accumulation, and increased lung metastasis. Besides that, GLUT1, GLUT3, HK2, and HIF-1α expression were increased in the xenografts, indicating that TAM-EVs play an essential role in the metabolic reprogramming of breast tumor cells, in vitro and in vivo [Figure 1 and Table 1][75].

Interconnections within the tumor microenvironment: extracellular vesicles as critical players of metabolic reprogramming in tumor cells

Figure 1. Overview of the metabolic alterations on cancer cell lines and tumors promoted by EVs derived from macrophages, fibroblasts, adipocytes, and AT. Fibroblasts liberate exosomes-containing lncRNA SNHG3, leading to positive regulation of PKM, resulting in the augmentation of glycolysis carboxylation and reduction of mitochondrial OXPHOS in breast cancer cell lines. Prostate cancer cells stimulated with CDEs demonstrated increased glucose uptake, glycolysis, lactate secretion, reductive carboxylation of glutamine, and reduced OXPHOS. These effects were possibly achieved by miR-22, miR-let7a, and miR-125b interaction through CDE delivery. CDEs transport LINC01614 to lung cancer cell lines, upregulating OXPHOS, glutamine transporters, and ATP synthesis. CDEs carrying lower concentrations of miR-7641 promote HIF-1α stabilization, resulting in an upregulation of HK2, GLUT1, LDHA, and increased glycolysis rate and cell proliferation in breast cancer cell lines. TAM-EVs deliver HISLA to breast cancer cells, generating enhanced glucose consumption, lactate production, 18FDG accumulation, and elevated GLUT1, GLUT3, HK2, and HIF-1α expression. DOCK 7 was found to be carried on the surface of TAM-EVs, leading to increased cholesterol content and altered cholesterol metabolism in CRC cells. TAM-derived exosomes demonstrated a positive regulation of glucose consumption, lactate production, and GLUT1, LDHA, and HK2 expression in HCC. This effect was due to the transport of lncMMPA by TAM-derived exosomes. Moreover, TAM-derived exosomes were observed to carry miR-193-3b-3p, culminating in enhanced glutamine uptake in the human pancreatic adenocarcinoma cell line. AD-EXOs promote ferroptosis resistance mediated by the transference of MTTP to colorectal cancer cell lines, affecting the response to chemotherapy. The miRNAs miR-155-5p, miR-10a-3p, and miR-30a-3p were proposed to be enriched in O-EVs and delivered to human breast cancer cell lines, stimulating increased mitochondrial density, oxygen consumption rate, and ATP production. AD EVs carry enzymes responsible for FAO, upregulating this pathway, leading to an increased invasion and migration of the human melanoma cell line. AD EVs increase glucose uptake, lactate production, and ATP generation in prostate cancer cell lines, although the EVs cargo is unknown. EVs: Extracellular vesicles; AT: adipose tissue; SNHG3: small nucleolar RNA host gene 3; HIF-1α: hypoxia-inducible factor 1-alpha; DOCK 7: dedicator of cytokinesis 7; TAMs:tumor-associated macrophages; CRC:colorectal cancer; HCC:hepatocellular carcinoma; lncRNA: long non-coding RNA; lncMMPA: lncMMPA: RP-11-1100L38; ATP: adenosine triphosphate; FAO: fatty acid oxidation; 18FDG accum.: 18F-fluorodeoxyglucose accumulation; Glutamine carb.: Glutamine carboxylation; Glucose cons.: Glucose consumption; GLUT: glucose transporter; Glucose up.: Glucose uptake; HK2: hexokinase 2; LDHA exp.: lactate dehydrogenase expression; Lactate prod.: Lactate production; Lactate sec.: Lactate secretion; lnc-RNA: long non-coding ribonucleic acid; miR: micro-RNA; MTTP: microsomal triglyceride transfer protein; OXPHOS: oxidative phosphorylation; Oxygen cons. rate: Oxygen consumption rate; PKM exp.: pyruvate kinase M expression (Created with BioRender).

Table 1

Overview of EVs derived from macrophages, fibroblasts, adipocytes, and AT cargoes and effects in cancer cell biology and metabolism

Study typeDonor cellEVs contentRecipient cellAnimal modelEffectsRef.
Macrophages
In vitro
and in vivo
PBDMs from breast cancer patients or healthy donors treated with MDA-MB-231 cells conditioned mediumHISLAMDA-MB-231 (Breast cancer cell line)NOD/SCID mice inoculated with MDA-MB-231Increased glucose consumption, lactate production, and apoptosis resistancethrough HIF-1α stabilization in vitro. 18FDG accumulation, lung metastasis, chemioresistance, increased GLUT1, GLUT3, HK2, and HIF-1α, in vivo.[75]
In vitro
and in vivo
Macrophages isolated from hepatocellular carcinoma samples and PBDMs from healthy donors treated with PMA and IL-4lncMMPAHep3b (Hepatocellular carcinoma cel line)BALB/c nude mice inoculated with Hep3B cellsIncreased glucose consumption, lactate production, extracellular acidification rate, proliferation, GLUT 1, HK2, LDHA expression in vitro, and tumor growth in vivo.[76]
In vitro
and in vivo
THP-1 (Leukemia monocytic cell line) treated with PMA and IL-4miR- 193b-3pSW1990 (Pancreatic cancer cell line)Lung metastasis model inoculated with SW1990 cellsIncreased glutamine uptake, proliferation, migration, and invasion in vitro and lung metastasis in vivo through TRIM62/cMyc signaling.[77]
In vitro
and in vivo
Bone marrow-derived macrophages from C57CL/7 mice and RAW264.7 (mice macrophage cell line) treated with MC-38 conditioned mediumDOCK 7MC-38 and CT26 (colorectal cancer cell lines)C57BL/6 and BALB-C mice inoculated with MC-38 and CT-26, pretreated with TAM-EVsIncreased migration and invasion, altered EMT markers, decreased membrane and intracellular cholesterol content, and enhanced membrane fluidity in vitro through the RAC1/ABCA1/AKT/FOXO1 pathway. It also increases metastatic burden in vivo.
[81]

Fibroblasts
In vitro
and in vivo
Fibroblasts from breast cancer patientsSNHG3MCF7 and MDA-MB-453 (Breast cancer cell lines)BALB/c nude mice inoculated with MDA-MB-453 cellsSNHG3 functions as a miR-330-5p sponge to positively regulate PKM expression, inhibit mitochondrial oxidative phosphorylation, and increase glycolysis carboxylation in vitro. SNHG3 knockout in CAF-derived exosomes inhibits breast tumor growth in vivo.[91]
In vitroFibroblasts from prostate cancer patientspossible miR-22, let7a, miR-125bPC3 (Prostate cancer cell line)No longer detectedReduced OXPHOS; glycolysis increase; increased glucose uptake and lactate secretion; upregulated reductive carboxylation of glutamine.[51]
In vitroFibroblasts from pancreatic cancer patientsAmino acids, lactate,
acetate,
TCA intermediates and lipids
BxPC3 (WildType Kras) and MiaPaCa-2 (Homozygous Kras) (Pancreatic cancer cell lines)No longer detected Facilitate cancer cell proliferation under nutrient deprivation conditions.[51]
In vitro
and in vivo
Fibroblasts from lung adenocarcinoma patientslncRNA-LINC01614A549 (Lung cancer cell line)BALB/c nude and NCG mice, and zebrafish inoculated with A549Upregulation of glutamine transporters in vitro and in vivo. Higher mitochondrial oxidative phosphorylation and ATP synthesis in vitro. Increase in proliferation, migration, and invasion, in vitro and in vivo.[92]
In vitro
and in vivo
Fibroblasts from breast cancer patientsmiRNA-7641 (Lower concentrations)MDA-MB-231 and SKBR3 (Breast cancer cell lines)BALB/c nude mice inoculated with MDA-MB-231Higher glycolysis rate and proliferation in breast cancer cells by targeting HIF-1α regulation and upregulating HK2, GLUT1, and LDHA in vitro. Promotion of breast cancer growth, in vivo
[90]
Adipocytes and the AT
In vitro
and in vivo
Breast AT(obese and overweight)miR-155-5p, miR-30a-3p, and miR-10a-3pMCF7 (Breast cancer cell line)C57BL/6 mice treated with EO771 cells previously educated with O-EVIncreased cell proliferation, mitochondrial density, oxygen consumption rate, and ATP production in vitro through miR-155-5p, miR-10a-3p e miR-30a-3p transfer, and Akt/mTOR/P70S6K activation. Promotion of breast cancer growth, in vivo.[82]
In vitro
and in vivo
Adipocytes differentiated from 3T3-L1MTTPHCT116 and SW480 (Colorectal cancer cell lines)Ob/ob mice previously treated with or without KD-MTTP lentivirus and MC-38 cells.Promotes ferroptosis resistance through MTTP transfers, which increases GPX4 and xCT levels and worsens response to oxaliplatin.[83]
In vitro
and in vivo
Adipocytes differentiated from 3T3-F442A and Adipocytes isolated from HFD and LFD C57BL/6J miceFAO Enzymes (ECHA, HCDH) and fatty acids (FA)SKMEL28 and 1205L (Melanoma cancer cell lines)Nude athymic mice treated with GFP-SKMEL28 and HFD AD-EXOIncreased FAO, mitochondrial dynamics rearrangement, and melanoma cell migration support through FAO enzymes and FA transport.Increased metastatic colonization in mice lungs.[84]

[85]
In vitroAdipocytes differentiated from 3T3-L1No longer detectedPC3 and DU145 (Prostate cancer cell lines)No longer
detected
Increased glucose uptake, lactate production, and ATP generation.[86]

Similarly, TAM-derived exosomes increase aerobic glycolysis in hepatocellular carcinoma (HCC) cells. A human HCC cell line (Hep3b) was treated with exosomes obtained by the CM released by macrophages isolated from HCC tumor samples or PBDMs obtained from healthy donors polarized into M2 phenotype, with phorbol 12-myristate 13-acetate (PMA) and interleukin 4 (IL-4). It was observed that both exosomes induced an increase in glucose consumption, lactate production, extracellular acidification rate, and proliferation in Hep3b cells. Furthermore, it also increased GLUT 1, HK2, and LDHA expression. These changes were attributable to the delivery, through exosomes, of RP-11-1100L38 - an M2 macrophage polarization-associated lncRNA (lncMMPA). This acts as a competing endogenous RNA for miR-548, which targets aldehyde dehydrogenase 1 family member A3 (ALDH1A3). Reducing ALDH1A3 expression, in turn, promotes lactate production and cell proliferation. Concurrently, it was observed that exosomes obtained from PBDMs polarized into M2 phenotype increased tumor growth in BALB/c nude mice inoculated with Hep3B [Figure 1 and Table 1][76].

Concerning the role of EVs in promoting metabolic shifts within the TME, Zhang et al. have demonstrated an increase in glutamine uptake in a human pancreatic adenocarcinoma cell line (SW1990 cells) when co-cultured with M2 macrophages (differentiated in vitro from THP-1, a human leukemia monocytic cell line). Interestingly, this effect was prevented in the presence of GW4869, an inhibitor of exosome biogenesis. This metabolic alteration occurs due to miR-193-3b-3p delivery through exosomes, which targets tripartite Motif Containing 62 (TRIM62), thus reducing c-Myc degradation. The exact mechanism also promoted increased proliferation, migration, and invasion of SW1990 cells and lung metastasis in mice inoculated with SW1990 cells previously treated with M2 exosomes [Figure 1 and Table 1][77]. Those effects may be due to the increased glutamine uptake, which, as a carbon and nitrogen source, has a role in amino acid, nucleotide, and lipid biosynthesis and performs anaplerotic replenishment of the TCA[78], essential elements sustaining cancer cell proliferation[79], migration, and invasion[80].

It has already been demonstrated that TAM-EVs also alter cholesterol metabolism in CRC cells. A study conducted by Chen et al. showed that EVs obtained from C57CL/7 mice bone marrow-derived macrophages and a mice macrophage cell line (RAW264.7) previously educated with the CM from a mice CRC cell line (MC-38) promoted migration and invasion in mice CRC cell lines (MC-38 and CT26). Besides that, it also altered EMT markers and deregulated cholesterol metabolism by decreasing membrane and intracellular cholesterol content in a dedicator Of Cytokinesis 7 (DOCK 7)-dependent manner. Their results demonstrated that DOCK 7 was anchored on the TAM-EVs surface and delivered to the tumor cells. This promoted the upregulation of RAC1, a classic small GTPase, triggering the ATP-binding cassette transporter (ABCA1) increase via AKT/FOXO1. These alterations increased membrane fluidity, resulting in cell motility and invasiveness. Those results were corroborated in vivo; the injection of C57BL/6 and BALB-C mice with MC-38 and CT-26 cells previously treated with TAM-EVs increased metastatic burden. Concurrently, simvastatin, an ABCA1 inhibitor, decreased hepatic metastatic nodules. These results demonstrate the relevance of cholesterol metabolism in CRC, both in vitro and in vivo [Figure 1 and Table 1][81].

Adipocyte and adipose tissue-derived EVs

While the role of adipose tissue (AT) and cancer-associated adipocytes (CAAs) in cancer is well-established, with their secretion of adipokines, growth factors, and EVs stimulating tumor cell proliferation, invasion, and chemoresistance, the mechanisms by which EVs released by adipocytes and AT modulate tumor metabolism remain largely unexplored.

In a study by Liu et al., the influence of EVs derived from obese breast AT on the metabolism of breast tumor cells was investigated. The results demonstrated that a human breast adenocarcinoma cell line (MCF-7 cells) educated with EVs derived from obese and overweight women AT (O-EVs) displayed increased proliferation, mitochondrial density, oxygen consumption rate, and ATP production. The increase in proliferation by O-EVs was shown to be OXPHOS-dependent as metformin, an inhibitor of mitochondrial complex I, reversed the proliferative effects. Concurrently, the previously educated MCF-7 cells had increased their proliferative capacity. This was abolished when the stimuli with O-EV occurred in the presence of metformin - an inhibitor of complex I and ATP synthesis. One proposed mechanism was the transfer of specific miRNAs, such as miR-155-5p, miR-10a-3p, and miR-30a-3p, which are enriched in O-EVs compared to EVs from lean women [Figure 1 and Table 1]. Silencing these miRNAs partially reduced the increase in proliferation and OXPHOS induced by O-EVs. Furthermore, activation of the AKT/mTOR/P70S6K pathway has also been shown to be a crucial mechanism for the metabolic and functional effects mediated by O-EVs. In vivo assays with female C57BL6 mice fed low-fat (LFD) or high-fat diets (HFD) showed that murine mammary cancer cells (EO771) previously educated with EVs isolated from HFD mammary AT increased tumor growth compared to EO771 educated with EVs from LFD-fed mice[82].

Zhang et al. demonstrated that exosomes derived from adipocytes (AD-EXOs) 3T3-L1 could reduce susceptibility to ferroptosis in human CRC cell lines (HCT116 and SW480), thus promoting chemoresistance to oxaliplatin. One proposed mechanism was the transfer of microsomal triglyceride transfer protein (MTTP), which increases AD-EXOs [Figure 1 and Table 1]. They showed that the MTTP/PRAP1 complex inhibits the expression of ZEB1 and increases the levels of glutathione peroxidase 4 (GPX4) and the cystine transporter (xCT). Together, these alterations triggered lower lipid peroxidation and a reduced proportion of polyunsaturated fatty acids (PUFA), decreasing the ferroptosis induced by oxaliplatin. The in vivo assays involved genetically obese mice (ob/ob), local adipose KD-MTTP using lentivirus transfection, and tumor implantation (MC-38 cells). The obese mice treated with KD-MTTP showed a more significant response to oxaliplatin, confirming that MTTP activity inversely regulates sensitivity to chemotherapy[83].

Concurrently, AD-EXOs from 3T3-F442A were reported to increase both the migratory and invasive capacity of two lineages of human melanoma (SKMEL28 and 1205L) in an FAO-dependent manner [Figure 1 and Table 1]. The functional effect of these EVs was confirmed in an in vivo model, in which the injection of GFP-SKMEL28 treated with AD-EXOs increased the metastatic colonization in the lungs of female nude athymic mice[84]. Proteomic and western blotting analysis of these EVs revealed the presence of enzymes committed to FAO, such as ECHA (a subunit of the trifunctional protein) and HCDH (hydroxy acyl-coenzyme A dehydrogenase). Those findings aligned with the increased rate of FAO in the cells treated with AD-EXOs, sustaining the rise in migration and invasion once reverted in etomoxir’s presence[84]. As expected, those effects were even more pronounced within the obesity context; in an HFD model using C57BL/6J mice, the results showed an increase in exosomes released from adipocytes derived from obese AT, which further promoted the invasive capacity of SKMEL28 and 1205L cells[84].

To further comprehend the mechanisms by which obesity increased the invasive capacity of Mel 1 in an FAO-dependent manner, Clement et al. monitored the transference of proteins from exosomes derived from 3T3-F442A adipocytes to SKMEL28 cells using SILAC (Stable Isotope Labeling of Amino Acids in Cell Culture) followed by mass spectrometry, and identified that transferred proteins from exosome to SKMEL28 cells were selectively sorted. Even though obesity was not related to an increment in transferred proteins, it was related to the increased transfer of fatty acids to the Mel 1. Such fatty acids were then mobilized by lipophagy, in conjunction with OXPHOS-transferred enzymes, thus increasing FAO. The increase in FAO, in turn, was related to a modulation in mitochondrial dynamics changes, which were co-localized with the cytoskeleton protrusions to support the migration of Mel 1 [Figure 1 and Table 1][85].

In addition to the effects of EVs derived from adipocytes upon changes in FAO, it was observed that EVs originating from differentiated 3T3-L1 adipocytes (AD-EVs) promoted a metabolic switch toward a glycolytic state in two human prostate cancer tumor lines (PC3 and DU145). Upon treatment with AD-EVs, both cell lineages displayed increased glucose uptake and lactate production, along with ATP generation sustained by HIF-1α stabilization in an AKT-dependent manner [Figure 1 and Table 1]. Such metabolic reprogramming promoted by AD EVs sustained the aggressiveness by increasing the rates of proliferation, migration, and invasion, other than resistance to the chemotherapy drug docetaxel[86].

Although the importance of EVs released by adipocytes and AT is recognized in cancer, few studies thoroughly explore their role as a metabolic modulator of tumor cells. In addition to the scarcity in the literature, another limitation concerns the model for obtaining adipocyte EVs; most studies isolate these vesicles from differentiated immortalized murine lines such as 3T3-L1 preadipocytes, which may not reliably reflect the conditions of human preadipocytes. Finally, many articles associate the findings of EVs derived from adipocytes and extrapolate their interpretation to obesity. The interest in this condition can be understood by the fact that an increase in AT characterizes obesity and is directly associated with several types of cancer. However, few studies detail the metabolic mechanisms induced by EVs released by adipocytes or AT in this condition.

Fibroblast-derived EVs

Fibroblasts, integral to the stroma, are key players in maintaining the extracellular matrix (ECM) and engaging in signal exchange with the microenvironment, particularly in wound healing. In the context of cancer, the TME is densely populated by fibroblasts, which undergo an activation process mediated by neoplastic cells, transforming them into CAFs. These CAFs, as subverted cells, overexpress different types of molecules, such as transforming growth factor beta (TGF-β), stromal cell-derived factor 1 (SDF-1), matrix metalloproteinase-2 (MMP-2), and so on[87], thus significantly contributing to tumor initiation, progression, and metastasis through a variety of interactions[88]. Therefore, the role of CAFs in the TME is paramount, and understanding their mechanisms is crucial in cancer biology and TME.

EVs released from neoplastic cells also contribute to converting fibroblasts to CAFs. These EVs trigger the metabolic reprogramming of CAFs toward glycolysis, producing lactate, pyruvate, and other metabolites that cancer cells absorb to fuel their metabolic pathway and survive. Conversely, these “educated” fibroblasts can also secrete EVs that may interact with cancer cells, inducing metabolic reprogramming and potentially leading to cancer cell progression[89].

A study performed by Zhao et al. provided compelling evidence of the role of exosomes released by CAFs (CDEs) derived from prostate cancer biopsies. These CDEs drive the oxygen consumption rate (OCR) inhibition in PC3 cells. The effects of CDEs were partly due to miRNAs, including miR-22, let7a, and miR-125b. In addition to the dysregulation in OXPHOS, CDEs also upregulated glucose uptake and glycolysis [Figure 1 and Table 1]. Oppositely, CDEs increased the reductive carboxylation of glutamine-driven glutamine’s contribution to lipogenic acetyl-CoA for enabling membrane synthesis, which probably sustained the enhanced proliferative state of the prostate cancer cells[51].

Corroborating the role of EVs in the metabolic reprogramming of cancer cells, CDEs from biopsies of pancreatic cancers, enriched in amino acids (glutamine, threonine, phenylalanine, valine, isoleucine, glycine, arginine, and serine), lactate, acetate, lipids, and TCA intermediates, displayed a role on supplying amino acids to oncogenic pancreatic cancer cells. The results demonstrated that this supply was independent of Kras since CDEs could rescue the loss of proliferation in the two human pancreatic cancer cell lines studied: BxPC3 (wild-type Kras) and MiaPaCa-2 (homozygous Kras) under nutrient deprivation [Figure 1 and Table 1][51].

In line with those findings, CAFs-derived exosomes obtained from breast cancer biopsies containing lower levels of miR-7641 drove glycolytic profiles in human breast cancer cell lines (MDA-MB-231 and SKBR3) and increased proliferative rates compared to exosomes derived from normal fibroblasts. This mechanism was attributable to the tumor suppressive properties of miR-7641 in targeting the transcription factor HIF-1α; when in low concentrations of miR-7641, upregulation of HIF-1α was reported, thus triggering increased expression of HK2, GLUT1, and LDHA [Figure 1 and Table 1][90].

Exosomes from CAFs derived from breast cancer biopsies were reported to be enriched in SNH3G, a lncRNA that acts as a molecular sponge for miRNA-330-5p. Theoretically, miR-330-5p decreases pyruvate kinase M (PKM) expression, suppressing glycolysis metabolism and proliferation in tumor cells. However, the uptake of those exosomes induced a PKM-positive regulation in a human breast cancer cell line (MDA-MB-453), inhibiting the OXPHOS [Figure 1 and Table 1]. This effect was due to the suppressive SNHG3/miR-330 signaling axis regulating the proliferation and metabolism of breast tumor cells by modulating PKM at the post-transcription level[91].

Liu et al. identified that exosomes derived from CAFs from lung adenocarcinoma (LUAD) biopsies were enriched in the long intergenic non-protein coding RNA 1614 (LINC01614). LINC01614, in turn, enhanced the glutamine metabolism of a human LUAD cell lineage (A549), thus promoting glutamine “cell addiction.” The results demonstrated that LINC01614 was transported via exosomes from CAFs and subsequently absorbed by LUAD cells, modulating the tumor metabolism by upregulating the expression of glutamine transporters and consequently inducing a preferential uptake of glutamine, driving higher proliferation, migration, and invasion abilities. Both glycolysis and OXPHOS rates did not differ by the treatment with the exosomes from CAFs, although a higher OXPHOS and ATP synthesis were observed. Interestingly, it was shown that LINC01614 transported within CAF exosomes enhanced Annexin A2 and p65 (Rela) interactions, promoting NF-κB activation and the release of interleukin-6 (IL-6) and CXCL10 by A549 cells, thus upregulating LINC01614 by CAFs when in co-culture, revealing a feedforward loop [Figure 1 and Table 1][92].

A growing landscape of studies corroborates the idea that noncancerous cells associated with tumors secrete a variety of cargo delivered by EVs that can modulate some of the well-known hallmarks of cancer. Regarding cancer cell metabolism, we have compiled evidence highlighting the role of CAF-derived EVs in metabolic reprogramming. In this scenario, CAFs-derived EVs fuel tumor cells with amino acids and ready-to-use metabolic intermediates, modifying the target cells’ balance between glycolysis and OXPHOS. In addition, miRNAs and lncRNA carried by CAF EVs induce metabolic reprogramming, ultimately sustaining enhanced cancer cell proliferation, migration, and invasiveness.

Figure 1 and Table 1 show the recognized modulatory roles of macrophage, fibroblast, adipocyte, preadipocyte, and AT-derived EVs within the TME upon cancer cell metabolism.

POSSIBLE ROLES OF EVS FROM SURROUNDING TME IN MODULATING CANCER CELL METABOLISM: A PERSPECTIVE VIEW

Tumor-infiltrating lymphocytes-derived EVs

Tumor-infiltrating lymphocytes (TILs) are a crucial component of the TME. They are lymphoid cells comprising T cells, natural killers (NKs), and, more recently, B cells, and are easily found within TME. Initially committed to antitumor defense, TILs can lose their tumor-killing ability in response to immunosuppressive factors secreted in TME, thereby leading to disease progression[93].

NK cells act in the initial defense against the tumor because they do not need prior stimulation and have a non-specific killing ability[94]. It has been demonstrated by several studies that NK-derived EVs (NK-EVs) contain miRNAs that are known to regulate the glycolytic metabolism of tumor cells. Among them, miR-186[95] reduces aerobic glycolysis through the downregulation of HIF-1α in the human osteosarcoma (OS) (U2 and HOS) [111] and in the human gastric[96] cell lines (MKN45 and SGC7901) [Figure 2A], thus functioning as a tumor suppressor. Concurrently, miR-125b-5p and miR-199a-5p, whose levels are increased in NK-EVs[97], also downregulate HK2 activity [Figure 2A], decreasing glycolysis in human laryngeal squamous cell carcinoma cell lines (AMC-HN-8 and Tu-177)[98], and human hepatocarcinoma cell lines (Huh-7, HepG2, and Hep3b)[99] [Figure 2A].

Interconnections within the tumor microenvironment: extracellular vesicles as critical players of metabolic reprogramming in tumor cells

Figure 2. Proposed tumor metabolic modulation exerted by EVs derived from TILs and MSCs. Small circles represent EVs and exosomes from the origin cells. Suggested tumor metabolic targets of the microparticles cargo are depicted and represented by dotted arrows (red: negative regulation; green: positive regulation). A: Reduced tumor glycolysis could be achieved by miRs within NK-derived EVs by downregulating HIF-1α (miR-186) and HK (miR-125b-5p and miR-199a-5p, as detailed). Conversely, increased glycolysis may be epigenetically induced by miR-21 and miR-1180 within hBM-MSC exosomes at specific regulatory checkpoints. Expected glycolytic targets of miRs within NK and h-BM-MSC-derived EVs and exosomes. Depending on the EV-secreting cell and specific miR cargo, both glycolysis downregulation (at HK level, by miR-186, miR-125b-5p, and miR-199a-5p from NK-derived EV, and at FBP1 and PDH levels, by the action of miR-21 contained in h-BM-MSC EVs) and upregulation (at HK, PK, PDK, and LDH levels through a miR-1180 modulation mediated by h-BM-MSC EVs) can be observed; B: cytotoxic T lymphocytes (CD8+) can release miR-765 within exosomes. This miRNA could decrease the lipid content of tumor cells through PLP2 downregulation. Inversely, the onco-miR-21 contained in the exosomes of h-BM-MSC may increase the expression of the fatty acid translocase (CD36) in tumor cells, leading to enhanced exogenous uptake and intracellular levels of the energetic supply of phospholipids, neutral lipids, and TAG. Additionally, increased intracellular levels of the enzymes ACC and FASN related to lipid synthesis and FABP5 related to tumor progression in many cancers (reviewed in 26) may also be observed due to the miR-21 within h-BM-MSC exosomes. C: helper T lymphocytes (CD4+) release exosomes enriched in miR-155-5p, which could negatively regulate the expression of the GATM gene that encodes glycine amidinotransferase, an enzyme involved in creatine synthesis; D: EVs derived from B lymphocytes contain CD39 and CD73, which are proteins involved in the phosphohydrolysis of ATP and ADP to AMP, producing ADO. Inversely, pro-tumoral roles of lactate and miR-1180 within h-MSC microparticles could be induced by increased ATP production and migratory capacity as a result of a modulation of the oxidative (E) or glycolytic (A) metabolism, respectively; E: hMSLC-EVs may provide Glutamine and further converted in α-ketoglutarate, replenishing the citric acid cycle (TCA) and fueling the metabolism of cancer cells. Similarly, a possible conversion of the exosomal lactate from hAT-MSC exosome to pyruvate may also replenish the TCA and enhance the ATP production by the mitochondrial complexes, thus boosting the tumor cell migration in an uptake mediated by the MCT-1 transporter. EVs: Extracellular vesicles; TILs: tumor-infiltrating lymphocytes; ACC: acetyl-CoA carboxylase; ADO: adenosine; ADP: adenosine diphosphate; AMP: adenosine monophosphate; ATP: adenosine triphosphate; CD4: T-cell surface glycoprotein CD4; CD8: a cluster of differentiation 8; CD36: fatty acid translocase; CD39: ectonucleoside triphosphate diphosphohydrolase 1; CD73: ecto-5′-nucleotidase; FABP5: fatty acid binding protein 5; FASN: fatty acid synthase; FBP1: fructose-1,6-bisphosphatase; GATM: glycine amidinotransferase; hAT-MSC: human adipose tissue mesenchymal stem cell; hBM-MSC: human bone marrow mesenchymal stem cell; HIF-1𝛼: hypoxia-inducible factor 1-alpha; HK: hexokinase; hMSLC: human mesenchymal stem-like cell; LDH: lactate dehydrogenase; MCT-1: monocarboxylate transporter 1; miR: micro-RNA; NK: natural killer; PDH: pyruvate dehydrogenase; PDK: pyruvate dehydrogenase kinase; PK: pyruvate kinase; PLP2: proteolipid protein 2; TAG: triacylglycerol; TCA: tricarboxylic acid cycle; I-IV: mitochondrial respiratory complexes. (Created with BioRender).

The infiltration of CD8+ cells, also known as cytotoxic T lymphocytes, into the TME indicates a better prognosis for several types of cancer; hence, CD8+ T cells are recognized as critical drivers of antitumor activity[100]. In this context, emerging evidence suggests that one of the primary mechanisms by which CD8+ cells exert tumor-killing mechanisms involves the delivery of cargo such as mRNA, miRNA, proteins, and lipids to tumoral cells within EVs[101,102]. It has already been reported that miR-765 exists within exosomes from healthy CD45RO-CD8+ T cells[102]. Interestingly, miR-765 has been recognized as a tumor suppressor, once promoting decreased lipid content in two human lineages of clear cell renal cell carcinoma (ccRCC) cell lines (A498 and Caki-1). This effect was mediated by the downregulation of proteolipid protein 2 (PLP2), a direct target gene of miR-765[103]. Thus, such evidence can highlight a possible mechanism by which EVs derived from CD8+ T cells affect lipid metabolism in ccRCC [Figure 2B].

CD4+ T cells, also known as helper T lymphocytes, assist the CD8+ T and B cell response[104]. They can also produce pro-inflammatory cytokines and directly kill tumor cells[105]. In the presence of interleukin-2 (IL-2), which promotes its activation, CD4+ T cells release exosomes enriched in miRNAs, such as miR-155-5p, among others[106]. The role of miR-155-5p in mediating the metabolic reprogramming in renal cancer has been highlighted by Bogusławska et al., which demonstrated the commitment of this miRNA under the downregulation of glycine amidinotransferase (GATM) gene expression in human ccRCC cell lines (Caki-2 and KIJ265T). GATM encodes glycine amidinotransferase, an enzyme committed to creatine synthesis[107] [Figure 2C]. It is worth mentioning that creatine is predicted to inhibit the growth of tumor cells by inhibiting glycolysis and attenuation of acidosis. However, the exact mechanism remains unknown[107,108].

B lymphocytes also play an essential role in tumor immunity due to their ability to produce antibodies[105]. An analysis performed by Zhang et al. revealed that EVs derived from B lymphocytes obtained from C57BL/6J mice contained CD39 and CD73[109]. These proteins are involved in the phosphohydrolysis of ATP and ADP to AMP[110], producing adenosine, a metabolite produced at high concentrations in TME, playing a role in tumor-mediated immune evasion[111] [Figure 2D]. Concurrently, it was already reported that CD39 and CD73 are committed to the EMT of cancer cells[112].

This compiled evidence shows that EVs derived from TILs probably play an essential role in modulating tumor metabolism, increasing malignancy, or providing antitumoral properties. Therefore, studies are necessary to elucidate the role of EVs derived from TILs upon metabolic reprogramming in cancer cells.

Stem cell and mesenchymal cell-derived EVs

Mesenchymal stromal cells (MSCs) consist of a heterogeneous group of multipotent progenitor cells found in distinct adult tissues, harboring the capacity for self-renewal and differentiation into different cell types, such as adipocytes, chondrocytes, and osteoblasts[113]. Due to metabolic plasticity, MSCs acquire specific phenotypes in response to changes in the microenvironment. Likewise, in hypoxic conditions, glycolysis and PPP maintain the MSC stemness profile[114,115]. CSCs, which can be raised from mutated MSCs, also present self-renewal ability and differentiate into multiple malignant cell types, strongly related to tumorigenesis, treatment resistance, tumor metastasis, and recurrence[116].

Exosomes (MSC-Exo) and EVs (MSC-EVs) released by MSCs can exert both pro- or antitumoral effects, depending on tumor type and the source of the MSCs[117-124]. However, details about the underlying molecular mechanisms by which these microparticles modulate cancer cell metabolism remain scarce. To the best of our knowledge, very few studies captured their content[118,124-126] and, to a lesser extent, the EVs released by CSCs (CSC-EVs)[127,128]. Biological activities, both in vitro[117-121,123-126,128-130] and in vivo[117-119,124,125,129], including internalization or cargo transference to tumor cells[118,119,121,124,125,130] were also identified. Based on captured knowledge and considering the known effects of representative individual cargo[131-135], the proposed modulatory mechanisms for cancer cell metabolism are depicted in Figure 2.

A large-scale study conducted in hypoxic and nutrient-deprived conditions to investigate the tumor-supporting role of human bone marrow MSC (hBM-MSC)-derived EVs (hBM-MSC-EVs) revealed the presence of critical components such as lactic and glutamic acids, lysosome-associated membrane glycoprotein 2 (LAMP2), tissue inhibitors of metalloproteinases-1 (TIMP-1) and -2 (TIMP-2), CD9 (tetraspanin-29), platelet-derived growth factor receptor (PDGFR), sphingomyelin, diacylglycerol, miR-21, and miR-34a. These hBM-MSC-EVs were then used as a stimulus in MCF-7 cells and human OS (KHOS) cells, leading to increased cell survival. Importantly, xenograft assays of MCF-7 cells injected with hBM-MSC-EVs provided in vivo evidence of their potential to support breast tumor growth, underscoring the practical implications of the study’s findings[125].

Regarding the cargo present in MSC-Exo, studies have already demonstrated an enriched content of FDFT1 (farnesyl-diphosphate farnesyltransferase 1) involved in cholesterol metabolism, stearoyl CoA desaturase (SCD) involved in lipid metabolism, aldolase B and enolase 3 involved in glycolysis, apoptosis signal-regulating kinase 1 (ASK1, also known as mitogen-activated protein kinase 5 - MAP3K5), MKK3/6 (dual specificity mitogen-activated protein kinases 3 and 6), and ferritin heavy chain 1 (FTH1) involved in NRF2/glutathione pathway, as well as epidermal growth factor (EGF), fibroblast growth factor (FGF), platelet-derived growth factor (PDGF), and proteins related to the canonical angiogenesis pathway[126]. Five miRs (miR-21, miR-92, miR-100, miR-143, miR-148) were identified as highly expressed within exosomes derived from human umbilical cord MSCs (hUC-MSC-Exo), of which miR-100-5p promoted both proliferation and migration of human pancreatic cancer (Panc-1) cells and tumor growth in a xenograft tumor model in which nude mice were injected with panc-1 cells in the presence of hUC-MSC-Exo[124]. Bioinformatic analysis revealed that ascorbic acid and aldehyde metabolism are the most significantly enriched pathways of the predicted targeted genes of these miRs. Besides that, (i) pentose and glucuronate interconversion; (ii) porphyrin metabolism; (iii) starch and sucrose metabolism; (iv) alanine, aspartate, and glutamate metabolism; and (v) regulation of lipid metabolism were also highlighted, among others[124]. A study by Vallabhaneni et al. demonstrated changes in the expression of twenty-nine miRs related to metastasis and cellular response to nutrients in KHOS cells treated with hBM-MSC-EVs. Through bioinformatic analysis, the results regarding the upregulated genes predicted interactions between the monocarboxylate transporter (MCT-1), ATPB5 (ATP-synthase), and oxidative stress players (Nrf2)[123]. Such findings highlight the hypothesis that MSC-EVs may affect oxidative metabolism in cancer cells.

In this context, an in vitro model demonstrated that treating the human OS cell line (Saos-2) with human MSCs derived from AT (hAT-MSC) revealed the influence of extracellular lactate in this tumor cell line. Indeed, the addition of known lactate concentrations to the culture of the OS cells triggered increased mitochondrial biogenesis, increased expression of I, II, IV, and V complexes of the respiratory chain, together with an increase in ATP production in an MCT-1 transporter-dependent manner, thus driving increased migratory capacity. Hence, lactate from hAT-MSC exosome to pyruvate may also replenish the TCA and enhance ATP production[129] [Figure 2D and E].

Another study demonstrated the influence of the glutamine released by human primary mesenchymal stem-like cells (hMSLC) on A549 and MCF-7 cell lines and primary cells isolated from malignant pleural effusion and ascites samples from cancer subjects. The increased glutamine levels within the CM from hMSLC triggered tumor-repopulating cells’ proliferation and colony formation in 3D soft fibrin gels. Such effects were blocked by the addition of l-γ-glutamyl-p-nitroanilide (GPNA), a glutamine transporter inhibitor, thus suggesting that glutamine may provide energy supply once glutamine can be converted in α-ketoglutarate, therefore replenishing the TCA[130] [Figure 2E].

miR-21 presents as a remarkable onco-miR in many human cancers[136], thus modulating the cancer cell metabolism at distinct targets. Additionally, the transference of miR-21 within hBM-MSC exosomes to MCF-7 and MDA-MB-231 cells was demonstrated, being related to chemoresistance and tumor progression, both from in vitro and in vivo tumor xenograft models, due to the miR-25-induced S100A6 protein expression[135]. Indeed, increased glycolysis is observed in the human gastric cell lines (MGC803 and SGC‑7901) by the direct effect of the miR-21 upon the reduction of mRNA and protein levels of pyruvate dehydrogenase A1 (PDHA1)[132]. Similar findings were observed in an in vitro study performed in A549, evidencing the suppression of the expression of glycolytic enzyme fructose-1,6-bis phosphatase (FBP1) as a direct target of miR-21[133] [Figure 2A]. Additionally, miR-21 promoted the growth and migration of A549 and H1703 cells by increasing the expression of the CD36 (fatty acid translocase), intracellular levels of phospholipids, neutral lipids, and triacylglycerides, together with an increase in both fatty acid synthase (FASN), acetyl-CoA carboxylase 1 (ACC1), and fatty acid binding protein 5 (FABP5) levels[131] [Figure 2B]. Similarly, increased pro-tumoral activity and chemoresistance were features mediated by the miR-1180, the most abundant miRNA within CM from hBM-MSC. This was demonstrated by evaluation based on in vitro and in silico models with the human ovarian cancer cell lines (SKOV3 and COC1). The mechanism involved: i. the modulation of the secreted frizzled-related protein 1 (SFRP1); ii. activation of the Wnt signaling pathway; iii. increased glycolysis togheter with the increased expression of glycolitic enzymes [Figure 2A] and; iv. increased ATP production[134] [Figure 2D].

The metabolomic profile of hCSC-EVs obtained from human primary metastatic Mel 1 highlights the differentially expressed metabolites glycerophosphocholine (GPC), triacylglycerol (TAG), glycerophosphoglycerol (PG), glycerophosphoserine (PS)[127]. Such findings are compatible with lipid metabolism’s role in the metastatic process[137]. A multi-omic study of the small EVs (sEVs) from distinct subpopulations of human glioblastoma stem-like cells (GSC) (proneural: NCH421k, NCH644, NCH441; mesenchymal: NCH705, NCH711d) has demonstrated their role as critical mediators of proteins and metabolites related to the metabolism of amino acids and fatty acids both related to glioblastoma progression. Furthermore, the role of such EVs in the phenotype modulation of GSC subpopulations highlights their relevance in supporting the metabolic heterogeneity in glioblastoma[128].

The findings captured by distinct experimental approaches support the modulatory role of microparticles shed by MSCs and CSCs upon the metabolism of different cancer cells. Despite the scarcity of studies in this field, the outcomes of biological relevance reinforce the need for additional investigations to highlight the underlying molecular mechanisms that may suggest possible targets for metabolic cancer control.

Therapeutic role of EVs in cancer metabolism

The applicability of EVs in cancer treatment has become an interesting strategy. Some studies investigate the role of non-tumor EVs in altering the tumor metabolism, leading to an antitumoral effect. Another interesting approach is to treat tumor cells with EVs that have had their content previously modified.

An interesting study showed that EVs from Lactobacillus plantarum can affect the metabolism of CRC. The study observed that the EVs can decrease glycolysis in 5-fluorouracil-resistant colorectal cells, and this metabolic shift can inhibit cancer cell proliferation and lead to apoptosis. The authors also observed that these effects rely on tumoral PDK2 expression decrease in p53/p21-dependent metabolic signaling[138].

Alcaya et al. investigated the role of EVs derived from menstrual stem cells affecting prostate tumor-induced angiogenesis once it became known that endometrial cells induce an angiostatic condition associated with the end of the menstrual cycle. So, the authors observed that the prostate tumor cells treated with the EVs had decreased VEGF levels, NF-κB activity, and tubulogenesis in a ROS-dependent manner. Furthermore, the authors showed that EVs derived only from endometrial mesenchymal cells had antitumoral properties, affecting cancer cell metabolism and inhibiting VEGF and HIF-1α expression. Finally, they observed in a xenograft model that the exosomes can inhibit angiogenesis in vivo[139]. It is important to note that HIF-1α is a pivotal modulator of the metabolic reprogramming that takes place in cancer cells, inducing glycolysis and glycogenesis, favoring PPP, and reducing oxidative metabolism through inhibition of pyruvate conversion to acetyl-CoA via PDK[140].

It is well described that miR-100 may act as a tumor suppressor in breast cancer since it has been found that this miRNA inhibits tumor growth and is downregulated in clinical breast cancers[141,142]. Thus, Pakravam et al. observed that EVs derived from human bone marrow mesenchymal cells are rich in miR-100 and investigated their role in breast cancer cells. They showed that the EVs can transfer miR-100 to MDA-MB-231, inhibiting the mTOR-HIF-1α axis. Finally, they observed that this effect could lead to decreased VEGF production by breast cancer cells, culminating in a decrease in tubulogenesis by endothelial cells[143].

Interestingly, Bruno et al. showed that EVs derived from human bone marrow mesenchymal cells inhibited cell cycle progression in all cell lines, induced apoptosis in HepG2 and Kaposi’s cells, and necrosis in the ovarian tumor cells (Skov-3). The authors observed that the EVs’ in vitro effects on cell metabolism rely on the activation of negative regulators of the cell cycle. Then, this antitumoral in vitro property was confirmed through in vivo analysis when EVs inhibited tumor growth of all cancer cells (HepG2, Kaposi´s, and Skov-3)[144].

Another interesting approach is to modify EVs in vitro and investigate their antitumor effects by overexpressing miR or adding a drug delivery system. Considering this, Lou et al. isolated human mesenchymal cells from subcutaneous AT and constructed cells expressing miR-199a. In this work, they isolated EVs enriched in miR-199a from these mesenchymal cells, and they observed its ability to deliver this miR to HCC cells, which were able to inhibit the mTOR pathway, reducing chemoresistance to doxorubicin in these cancer cells[145]. It is essential to the hallmark that mTOR plays a pivotal role in cell metabolism: 1) promoting the uptake of nutrients by tumor cells to allow rapid growth of tumors; 2) activating sterol regulatory element-binding protein (SREBP) pathway, which in turn promotes new lipid synthesis, inducing breast cancer proliferation[146]; 3) mTOR is implicated in chemotherapy resistance activating Fanconi anemia DNA repair pathway, and suppressing autophagy[147]. In the same way, Lin et al. suppressed oxaliplatin chemoresistance in colon cancer cells utilizing a drug-delivered approach. They generated EVs from modified HEK293T cells containing iRGD, which can bind to αvβ3, αvβ5, and NRP-1 in different tumor cells. Once the authors observed that CPT1A inhibition could reverse the chemoresistance of colon cancer cells to oxaliplatin, they loaded the iRGD EVs with siRNA CPT1A. They observed the ability of these EVs to reverse oxaliplatin resistance and inhibit tumor growth in vivo[148].

Figure 3 shows all the effects of EVs or previously modified EVs on different tumor types. It identifies the mechanisms of action and how the vesicles can affect tumor metabolism.

Interconnections within the tumor microenvironment: extracellular vesicles as critical players of metabolic reprogramming in tumor cells

Figure 3. The therapeutic role of EVs in cancer metabolism. Different approaches involving modified or unmodified EVs point to their role in inhibiting tumorigenesis in various types of cancer. EVs: Extracellular vesicles; AT: adipose tissue; BM: bone marrow; CAF: cancer-associated fibroblast; MSC: mesenchymal stem cell.

CONCLUSION

The pivotal role of vesicles as metabolic modulators in the microenvironment cannot be overstated. Vesicles derived from non-cancer cells in the TME play a significant role in modulating metabolic pathways in various cancer cell types. Research has revealed that EVs induce an aggressive phenotype in tumor cells through metabolic reprogramming. However, the number of studies exploring this area is limited, particularly in understanding how EVs precisely alter metabolic pathways, the specific role of the cargo within vesicles, and their impact on tumor phenotype. Understanding EVs’ communication mechanisms and systems is crucial in unraveling the intricacies of metabolic rearrangement in cancer cells. It can provide valuable insights for developing new therapeutic approaches.

DECLARATIONS

Acknowledgments

Figures were created with BioRender.com.

Authors’ contributions

Made substantial contributions to the design of the study, writing, and compiled evidence: Costa Encarnação C, Renovato-Martins M

Made substantial contributions to the conception of the study, writing, and critical review: Renovato- Martins M

Contributed to writing: Marianne Faria G, Aguiar Franco V, Xavier Botelho LG, Moraes JA

Made substantial contributions to the study design and review: Costa Encarnação C, Moraes JA, Renovato -Martins M

Availability of data and materials

Not applicable.

Financial support and sponsorship

None.

Conflicts of interest

All authors declared that there are no conflicts of interest.

Ethical approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Copyright

© The Author(s) 2024.

REFERENCES

1. Hanahan D, Weinberg RA. The hallmarks of cancer. Cell 2000;100:57-70.

2. Hanahan D, Weinberg RA. Hallmarks of cancer: the next generation. Cell 2011;144:646-74.

3. Visser KE, Joyce JA. The evolving tumor microenvironment: from cancer initiation to metastatic outgrowth. Cancer Cell 2023;41:374-403.

4. Hanahan D. Hallmarks of cancer: new dimensions. Cancer Discov 2022;12:31-46.

5. Dominiak A, Chełstowska B, Olejarz W, Nowicka G. Communication in the cancer microenvironment as a target for therapeutic interventions. Cancers (Basel) 2020;12:1232.

6. Niel G, Carter DRF, Clayton A, Lambert DW, Raposo G, Vader P. Challenges and directions in studying cell-cell communication by extracellular vesicles. Nat Rev Mol Cell Biol 2022;23:369-82.

7. Anderson NM, Mucka P, Kern JG, Feng H. The emerging role and targetability of the TCA cycle in cancer metabolism. Protein Cell 2018;9:216-37.

8. Faubert B, Solmonson A, DeBerardinis RJ. Metabolic reprogramming and cancer progression. Science 2020;368:eaaw5473.

9. Warburg O. The metabolism of carcinoma cells. J Cancer Res 1925;9:148-63.

10. Liberti MV, Locasale JW. The warburg effect: how does it benefit cancer cells? Trends Biochem Sci 2016;41:211-8.

11. Ma Y, Temkin SM, Hawkridge AM, et al. Fatty acid oxidation: an emerging facet of metabolic transformation in cancer. Cancer Lett 2018;435:92-100.

12. Moreno-Sánchez R, Marín-Hernández A, Saavedra E, Pardo JP, Ralph SJ, Rodríguez-Enríquez S. Who controls the ATP supply in cancer cells? Int J Biochem Cell Biol 2014;50:10-23.

13. Rodrigues MF, Carvalho É, Pezzuto P, Rumjanek FD, Amoêdo ND. Reciprocal modulation of histone deacetylase inhibitors sodium butyrate and trichostatin a on the energy metabolism of breast cancer cells. J Cell Biochem 2015;116:797-808.

14. Semenza GL. HIF-1: upstream and downstream of cancer metabolism. Curr Opin Genet Dev 2010;20:51-6.

15. Christofk HR, Vander Heiden MG, Harris MH, et al. The M2 splice isoform of pyruvate kinase is important for cancer metabolism and tumour growth. Nature 2008;452:230-3.

16. Bose S, Zhang C, Le A. Glucose metabolism in cancer: the warburg effect and beyond. In: Le A, editor. The heterogeneity of cancer metabolism. Cham: Springer International Publishing; 2021. pp. 3-15.

17. Liu Y, Gu W. The complexity of p53-mediated metabolic regulation in tumor suppression. Semin Cancer Biol 2022;85:4-32.

18. Rousset M, Zweibaum A, Fogh J. Presence of glycogen and growth-related variations in 58 cultured human tumor cell lines of various tissue origins. Cancer Res 1981;41:1165-70. Available from: https://aacrjournals.org/cancerres/article/41/3/1165/485476/Presence-of-Glycogen-and-Growth-related-Variations [Last accessed on 27 Sep 2024].

19. Cheng KW, Agarwal R, Mitra S, et al. Rab25 increases cellular ATP and glycogen stores protecting cancer cells from bioenergetic stress. EMBO Mol Med 2012;4:125-41.

20. Shen GM, Zhang FL, Liu XL, Zhang JW. Hypoxia-inducible factor 1-mediated regulation of PPP1R3C promotes glycogen accumulation in human MCF-7 cells under hypoxia. FEBS Lett 2010;584:4366-72.

21. Zhu Q, Yang J, Han S, et al. Suppression of glycogen synthase kinase 3 activity reduces tumor growth of prostate cancer in vivo. Prostate 2011;71:835-45.

22. Khan MW, Biswas D, Ghosh M, Mandloi S, Chakrabarti S, Chakrabarti P. mTORC2 controls cancer cell survival by modulating gluconeogenesis. Cell Death Discov 2015;1:15016.

23. Li T, Han J, Jia L, Hu X, Chen L, Wang Y. PKM2 coordinates glycolysis with mitochondrial fusion and oxidative phosphorylation. Protein Cell 2019;10:583-94.

24. Lunt SY, Vander Heiden MG. Aerobic glycolysis: meeting the metabolic requirements of cell proliferation. Annu Rev Cell Dev Biol 2011;27:441-64.

25. Liu Y, Sun Y, Guo Y, et al. An Overview: The diversified role of mitochondria in cancer metabolism. Int J Biol Sci 2023;19:897-915.

26. Martinez-Outschoorn UE, Peiris-Pagés M, Pestell RG, Sotgia F, Lisanti MP. Cancer metabolism: a therapeutic perspective. Nat Rev Clin Oncol 2017;14:11-31.

27. Ashton TM, McKenna WG, Kunz-Schughart LA, Higgins GS. Oxidative phosphorylation as an emerging target in cancer therapy. Clin Cancer Res 2018;24:2482-90.

28. Whitaker-Menezes D, Martinez-Outschoorn UE, Flomenberg N, et al. Hyperactivation of oxidative mitochondrial metabolism in epithelial cancer cells in situ: visualizing the therapeutic effects of metformin in tumor tissue. Cell Cycle 2011;10:4047-64.

29. Chen JQ, Russo J. Dysregulation of glucose transport, glycolysis, TCA cycle and glutaminolysis by oncogenes and tumor suppressors in cancer cells. Biochim Biophys Acta 2012;1826:370-84.

30. Pavlova NN, Thompson CB. The emerging hallmarks of cancer metabolism. Cell Metab 2016;23:27-47.

31. Cairns RA, Harris IS, Mak TW. Regulation of cancer cell metabolism. Nat Rev Cancer 2011;11:85-95.

32. Laurenti G, Tennant DA. Isocitrate dehydrogenase (IDH), succinate dehydrogenase (SDH), fumarate hydratase (FH): three players for one phenotype in cancer? Biochem Soc Trans 2016;44:1111-6.

33. Bardella C, Pollard PJ, Tomlinson I. SDH mutations in cancer. Biochim Biophys Acta 2011;1807:1432-43.

34. Cascón A, Landa I, López-Jiménez E, et al. Molecular characterisation of a common SDHB deletion in paraganglioma patients. J Med Genet 2008;45:233-8.

35. Ricketts C, Woodward ER, Killick P, Morris MR, Astuti D, Latif F, Maher ER. Germline SDHB mutations and familial renal cell carcinoma. J Natl Cancer Inst 2008;100:1260-2.

36. Zantour B, Guilhaume B, Tissier F, et al. A thyroid nodule revealing a paraganglioma in a patient with a new germline mutation in the succinate dehydrogenase B gene. Eur J Endocrinol 2004;151:433-8.

37. DeBerardinis RJ, Mancuso A, Daikhin E, et al. Beyond aerobic glycolysis: transformed cells can engage in glutamine metabolism that exceeds the requirement for protein and nucleotide synthesis. Proc Natl Acad Sci U S A 2007;104:19345-50.

38. Umapathy NS, Dun Y, Martin PM, et al. Expression and function of system N glutamine transporters (SN1/SN2 or SNAT3/SNAT5) in retinal ganglion cells. Invest Ophthalmol Vis Sci 2008;49:5151-60.

39. Xiao D, Zeng L, Yao K, Kong X, Wu G, Yin Y. The glutamine-alpha-ketoglutarate (AKG) metabolism and its nutritional implications. Amino Acids 2016;48:2067-80.

40. DeBerardinis RJ, Cheng T. Q’s next: the diverse functions of glutamine in metabolism, cell biology and cancer. Oncogene 2010;29:313-24.

41. Wu G, Fang YZ, Yang S, Lupton JR, Turner ND. Glutathione metabolism and its implications for health. J Nutr 2004;134:489-92.

42. Le A, Lane AN, Hamaker M, et al. Glucose-independent glutamine metabolism via TCA cycling for proliferation and survival in B cells. Cell Metab 2012;15:110-21.

43. Sajnani K, Islam F, Smith RA, Gopalan V, Lam AK. Genetic alterations in Krebs cycle and its impact on cancer pathogenesis. Biochimie 2017;135:164-72.

44. Icard P, Wu Z, Fournel L, Coquerel A, Lincet H, Alifano M. ATP citrate lyase: a central metabolic enzyme in cancer. Cancer Lett 2020;471:125-34.

45. Zhang C, Liu J, Huang G, et al. Cullin3-KLHL25 ubiquitin ligase targets ACLY for degradation to inhibit lipid synthesis and tumor progression. Genes Dev 2016;30:1956-70.

46. Carracedo A, Cantley LC, Pandolfi PP. Cancer metabolism: fatty acid oxidation in the limelight. Nat Rev Cancer 2013;13:227-32.

47. Camarda R, Zhou AY, Kohnz RA, et al. Inhibition of fatty acid oxidation as a therapy for MYC-overexpressing triple-negative breast cancer. Nat Med 2016;22:427-32.

48. Wang T, Fahrmann JF, Lee H, et al. JAK/STAT3-regulated fatty acid β-oxidation is critical for breast cancer stem cell self-renewal and chemoresistance. Cell Metab 2018;27:136-150.e5.

49. Moraes JA, Encarnação C, Franco VA, et al. Adipose tissue-derived extracellular vesicles and the tumor microenvironment: revisiting the hallmarks of cancer. Cancers (Basel) 2021;13:3328.

50. Niel G, D'Angelo G, Raposo G. Shedding light on the cell biology of extracellular vesicles. Nat Rev Mol Cell Biol 2018;19:213-28.

51. Zhao H, Yang L, Baddour J, et al. Tumor microenvironment derived exosomes pleiotropically modulate cancer cell metabolism. Elife 2016;5:e10250.

52. Colombo M, Raposo G, Théry C. Biogenesis, secretion, and intercellular interactions of exosomes and other extracellular vesicles. Annu Rev Cell Dev Biol 2014;30:255-89.

53. Théry C, Witwer KW, Aikawa E, et al. Minimal information for studies of extracellular vesicles 2018 (MISEV2018): a position statement of the international society for extracellular vesicles and update of the MISEV2014 guidelines. J Extracell Vesicles 2018;7:1535750.

54. Lucchetti D, Ricciardi Tenore C, Colella F, Sgambato A. Extracellular vesicles and cancer: a focus on metabolism, cytokines, and immunity. Cancers (Basel) 2020;12:171.

55. Peinado H, Alečković M, Lavotshkin S, et al. Melanoma exosomes educate bone marrow progenitor cells toward a pro-metastatic phenotype through MET. Nat Med 2012;18:883-91.

56. Pavlides S, Whitaker-Menezes D, Castello-Cros R, et al. The reverse warburg effect: aerobic glycolysis in cancer associated fibroblasts and the tumor stroma. Cell Cycle 2009;8:3984-4001.

57. Fridman ES, Ginini L, Gil Z. The role of extracellular vesicles in metabolic reprogramming of the tumor microenvironment. Cells 2022;11:1433.

58. Rai A, Greening DW, Chen M, Xu R, Ji H, Simpson RJ. Exosomes derived from human primary and metastatic colorectal cancer cells contribute to functional heterogeneity of activated fibroblasts by reprogramming their proteome. Proteomics 2019;19:e1800148.

59. Shu S, Yang Y, Allen CL, et al. Metabolic reprogramming of stromal fibroblasts by melanoma exosome microRNA favours a pre-metastatic microenvironment. Sci Rep 2018;8:12905.

60. Yan W, Wu X, Zhou W, et al. Cancer-cell-secreted exosomal miR-105 promotes tumour growth through the MYC-dependent metabolic reprogramming of stromal cells. Nat Cell Biol 2018;20:597-609.

61. Wu X, Zhou Z, Xu S, et al. Extracellular vesicle packaged LMP1-activated fibroblasts promote tumor progression via autophagy and stroma-tumor metabolism coupling. Cancer Lett 2020;478:93-106.

62. Park JE, Dutta B, Tse SW, et al. Hypoxia-induced tumor exosomes promote M2-like macrophage polarization of infiltrating myeloid cells and microRNA-mediated metabolic shift. Oncogene 2019;38:5158-73.

63. Sagar G, Sah RP, Javeed N, et al. Pathogenesis of pancreatic cancer exosome-induced lipolysis in adipose tissue. Gut 2016;65:1165-74.

64. Czystowska-Kuzmicz M, Sosnowska A, Nowis D, et al. Small extracellular vesicles containing arginase-1 suppress T-cellresponses and promote tumor growth in ovarian carcinoma. Nat Commun 2019;10:3000.

65. Zhou C, Zhang Y, Yan R, et al. Exosome-derived miR-142-5p remodels lymphatic vessels and induces IDO to promote immune privilege in the tumour microenvironment. Cell Death Differ 2021;28:715-29.

66. Mantovani A, Schioppa T, Porta C, Allavena P, Sica A. Role of tumor-associated macrophages in tumor progression and invasion. Cancer Metastasis Rev 2006;25:315-22.

67. Cassetta L, Pollard JW. A timeline of tumour-associated macrophage biology. Nat Rev Cancer 2023;23:238-57.

68. Yang C, Wei C, Wang S, et al. Elevated CD163+/CD68+ Ratio at Tumor Invasive Front is Closely Associated with Aggressive Phenotype and Poor Prognosis in Colorectal Cancer. Int J Biol Sci 2019;15:984-98.

69. Hu WM, Li M, Ning JZ, et al. FAM171B stabilizes vimentin and enhances CCL2-mediated TAM infiltration to promote bladder cancer progression. J Exp Clin Cancer Res 2023;42:290.

70. Chen S, Lu K, Hou Y, et al. YY1 complex in M2 macrophage promotes prostate cancer progression by upregulating IL-6. J Immunother Cancer 2023;11:e006020.

71. Atanasov G, Pötner C, Aust G, et al. TIE2-expressing monocytes and M2-polarized macrophages impact survival and correlate with angiogenesis in adenocarcinoma of the pancreas. Oncotarget 2018;9:29715-26.

72. Chen X, Yang M, Yin J, et al. Tumor-associated macrophages promote epithelial-mesenchymal transition and the cancer stem cell properties in triple-negative breast cancer through CCL2/AKT/β-catenin signaling. Cell Commun Signal 2022;20:92.

73. Kuwada K, Kagawa S, Yoshida R, et al. The epithelial-to-mesenchymal transition induced by tumor-associated macrophages confers chemoresistance in peritoneally disseminated pancreatic cancer. J Exp Clin Cancer Res 2018;37:307.

74. Jeong H, Kim S, Hong BJ, et al. Tumor-associated macrophages enhance tumor hypoxia and aerobic glycolysis. Cancer Res 2019;79:795-806.

75. Chen F, Chen J, Yang L, et al. Extracellular vesicle-packaged HIF-1α-stabilizing lncRNA from tumour-associated macrophages regulates aerobic glycolysis of breast cancer cells. Nat Cell Biol 2019;21:498-510.

76. Xu M, Zhou C, Weng J, et al. Tumor associated macrophages-derived exosomes facilitate hepatocellular carcinoma malignance by transferring lncMMPA to tumor cells and activating glycolysis pathway. J Exp Clin Cancer Res 2022;41:253.

77. Zhang K, Li YJ, Peng LJ, Gao HF, Liu LM, Chen H. M2 macrophage-derived exosomal miR-193b-3p promotes progression and glutamine uptake of pancreatic cancer by targeting TRIM62. Biol Direct 2023;18:1.

78. Jin J, Byun JK, Choi YK, Park KG. Targeting glutamine metabolism as a therapeutic strategy for cancer. Exp Mol Med 2023;55:706-15.

79. Demas DM, Demo S, Fallah Y, et al. Glutamine metabolism drives growth in advanced hormone receptor positive breast cancer. Front Oncol 2019;9:686.

80. Prasad P, Roy SS. Glutamine regulates ovarian cancer cell migration and invasion through ETS1. Heliyon 2021;7:e07064.

81. Chen W, Zhou M, Guan B, et al. Tumour-associated macrophage-derived DOCK7-enriched extracellular vesicles drive tumour metastasis in colorectal cancer via the RAC1/ABCA1 axis. Clin Transl Med 2024;14:e1591.

82. Liu S, Benito-Martin A, Pelissier Vatter FA, et al. Breast adipose tissue-derived extracellular vesicles from obese women alter tumor cell metabolism. EMBO Rep 2023;24:e57339.

83. Zhang Q, Deng T, Zhang H, et al. Adipocyte-derived exosomal MTTP suppresses ferroptosis and promotes chemoresistance in colorectal cancer. Adv Sci (Weinh) 2022;9:e2203357.

84. Lazar I, Clement E, Dauvillier S, et al. Adipocyte exosomes promote melanoma aggressiveness through fatty acid oxidation: a novel mechanism linking obesity and cancer. Cancer Res 2016;76:4051-7.

85. Clement E, Lazar I, Attané C, et al. Adipocyte extracellular vesicles carry enzymes and fatty acids that stimulate mitochondrial metabolism and remodeling in tumor cells. EMBO J 2020;39:e102525.

86. Fontana F, Anselmi M, Carollo E, et al. Adipocyte-derived extracellular vesicles promote prostate cancer cell aggressiveness by enabling multiple phenotypic and metabolic changes. Cells 2022;11:2388.

87. Chen X, Song E. Turning foes to friends: targeting cancer-associated fibroblasts. Nat Rev Drug Discov 2019;18:99-115.

88. Sazeides C, Le A. Metabolic relationship between cancer-associated fibroblasts and cancer cells. In: Le A, editor. The heterogeneity of cancer metabolism. Cham: Springer International Publishing; 2018. pp. 149-65.

89. Li Z, Sun C, Qin Z. Metabolic reprogramming of cancer-associated fibroblasts and its effect on cancer cell reprogramming. Theranostics 2021;11:8322-36.

90. Liu Y, Hua F, Zhan Y, et al. Carcinoma associated fibroblasts small extracellular vesicles with low miR-7641 promotes breast cancer stemness and glycolysis by HIF-1α. Cell Death Discov 2021;7:176.

91. Li Y, Zhao Z, Liu W, Li X. SNHG3 functions as miRNA sponge to promote breast cancer cells growth through the metabolic reprogramming. Appl Biochem Biotechnol 2020;191:1084-99.

92. Liu T, Han C, Fang P, et al. Cancer-associated fibroblast-specific lncRNA LINC01614 enhances glutamine uptake in lung adenocarcinoma. J Hematol Oncol 2022;15:141.

93. Lin B, Du L, Li H, Zhu X, Cui L, Li X. Tumor-infiltrating lymphocytes: warriors fight against tumors powerfully. Biomed Pharmacother 2020;132:110873.

94. Wang S, Shi Y. Exosomes derived from immune cells: the new role of tumor immune microenvironment and tumor therapy. Int J Nanomedicine 2022;17:6527-50.

95. Neviani P, Wise PM, Murtadha M, et al. Natural killer-derived exosomal miR-186 inhibits neuroblastoma growth and immune escape mechanisms. Cancer Res 2019;79:1151-64.

96. Liu L, Wang Y, Bai R, Yang K, Tian Z. MiR-186 inhibited aerobic glycolysis in gastric cancer via HIF-1α regulation. Oncogenesis 2016;5:e224.

97. Dosil SG, Lopez-Cobo S, Rodriguez-Galan A, et al. Natural killer (NK) cell-derived extracellular-vesicle shuttled microRNAs control T cell responses. Elife 2022:11.

98. Hui L, Zhang J, Guo X. MiR-125b-5p suppressed the glycolysis of laryngeal squamous cell carcinoma by down-regulating hexokinase-2. Biomed Pharmacother 2018;103:1194-201.

99. Guo W, Qiu Z, Wang Z, et al. MiR-199a-5p is negatively associated with malignancies and regulates glycolysis and lactate production by targeting hexokinase 2 in liver cancer. Hepatology 2015;62:1132-44.

100. der Leun AM, Thommen DS, Schumacher TN. CD8+ T cell states in human cancer: insights from single-cell analysis. Nat Rev Cancer 2020;20:218-32.

101. Cocucci E, Racchetti G, Meldolesi J. Shedding microvesicles: artefacts no more. Trends Cell Biol 2009;19:43-51.

102. Zhou WJ, Zhang J, Xie F, et al. CD45RO-CD8+ T cell-derived exosomes restrict estrogen-driven endometrial cancer development via the ERβ/miR-765/PLP2/Notch axis. Theranostics 2021;11:5330-45.

103. Xiao W, Wang C, Chen K, et al. MiR-765 functions as a tumour suppressor and eliminates lipids in clear cell renal cell carcinoma by downregulating PLP2. EBioMedicine 2020;51:102622.

104. Borst J, Ahrends T, Bąbała N, Melief CJM, Kastenmüller W. CD4+ T cell help in cancer immunology and immunotherapy. Nat Rev Immunol 2018;18:635-47.

105. Maibach F, Sadozai H, Seyed Jafari SM, Hunger RE, Schenk M. Tumor-infiltrating lymphocytes and their prognostic value in cutaneous melanoma. Front Immunol 2020;11:2105.

106. Shin S, Jung I, Jung D, et al. Novel antitumor therapeutic strategy using CD4+ T cell-derived extracellular vesicles. Biomaterials 2022;289:121765.

107. Bogusławska J, Popławski P, Alseekh S, et al. MicroRNA-mediated metabolic reprograming in renal cancer. Cancers (Basel) 2019;11:1825.

108. Campos-Ferraz PL, Gualano B, das Neves W, et al. Exploratory studies of the potential anti-cancer effects of creatine. Amino Acids 2016;48:1993-2001.

109. Zhang F, Li R, Yang Y, et al. Specific decrease in B-cell-derived extracellular vesicles enhances post-chemotherapeutic CD8+ T cell responses. Immunity 2019;50:738-750.e7.

110. Chen S, Wainwright DA, Wu JD, et al. CD73: an emerging checkpoint for cancer immunotherapy. Immunotherapy 2019;11:983-97.

111. Xia C, Yin S, To KKW, Fu L. CD39/CD73/A2AR pathway and cancer immunotherapy. Mol Cancer 2023;22:44.

112. Iser IC, Vedovatto S, Oliveira FD, Beckenkamp LR, Lenz G, Wink MR. The crossroads of adenosinergic pathway and epithelial-mesenchymal plasticity in cancer. Semin Cancer Biol 2022;86:202-13.

113. Tavakoli S, Ghaderi Jafarbeigloo HR, Shariati A, et al. Mesenchymal stromal cells; a new horizon in regenerative medicine. J Cell Physiol 2020;235:9185-210.

114. Folmes CD, Dzeja PP, Nelson TJ, Terzic A. Metabolic plasticity in stem cell homeostasis and differentiation. Cell Stem Cell 2012;11:596-606.

115. Liu Y, Muñoz N, Tsai AC, Logan TM, Ma T. Metabolic reconfiguration supports reacquisition of primitive phenotype in human mesenchymal stem cell aggregates. Stem Cells 2017;35:398-410.

116. Atashzar MR, Baharlou R, Karami J, et al. Cancer stem cells: a review from origin to therapeutic implications. J Cell Physiol 2020;235:790-803.

117. Zhu W, Huang L, Li Y, et al. Exosomes derived from human bone marrow mesenchymal stem cells promote tumor growth in vivo. Cancer Lett 2012;315:28-37.

118. Roccaro AM, Sacco A, Maiso P, et al. BM mesenchymal stromal cell-derived exosomes facilitate multiple myeloma progression. J Clin Invest 2013;123:1542-55.

119. Ono M, Kosaka N, Tominaga N, et al. Exosomes from bone marrow mesenchymal stem cells contain a microRNA that promotes dormancy in metastatic breast cancer cells. Sci Signal 2014;7:ra63.

120. Rezaeian A, Khatami F, Heidari Keshel S, et al. The effect of mesenchymal stem cells-derived exosomes on the prostate, bladder, and renal cancer cell lines. Sci Rep 2022;12:20924.

121. Jahangiri B, Khalaj-Kondori M, Asadollahi E, Purrafee Dizaj L, Sadeghizadeh M. MSC-derived exosomes suppress colorectal cancer cell proliferation and metastasis via miR-100/mTOR/miR-143 pathway. Int J Pharm 2022;627:122214.

122. Lee JK, Park SR, Jung BK, et al. Exosomes derived from mesenchymal stem cells suppress angiogenesis by down-regulating VEGF expression in breast cancer cells. PLoS One 2013;8:e84256.

123. Vallabhaneni KC, Hassler MY, Abraham A, et al. Mesenchymal stem/stromal cells under stress increase osteosarcoma migration and apoptosis resistance via extracellular vesicle mediated communication. PLoS One 2016;11:e0166027.

124. Ding Y, Mei W, Zheng Z, et al. Exosomes secreted from human umbilical cord mesenchymal stem cells promote pancreatic ductal adenocarcinoma growth by transferring miR-100-5p. Tissue Cell 2021;73:101623.

125. Vallabhaneni KC, Penfornis P, Dhule S, et al. Extracellular vesicles from bone marrow mesenchymal stem/stromal cells transport tumor regulatory microRNA, proteins, and metabolites. Oncotarget 2015;6:4953-67.

126. Anderson JD, Johansson HJ, Graham CS, et al. Comprehensive proteomic analysis of mesenchymal stem cell exosomes reveals modulation of angiogenesis via nuclear factor-kappab signaling. Stem Cells 2016;34:601-13.

127. Palacios-Ferrer JL, García-Ortega MB, Gallardo-Gómez M, et al. Metabolomic profile of cancer stem cell-derived exosomes from patients with malignant melanoma. Mol Oncol 2021;15:407-28.

128. Lokumcu T, Iskar M, Schneider M, et al. Proteomic, metabolomic, and fatty acid profiling of small extracellular vesicles from glioblastoma stem-like cells and their role in tumor heterogeneity. ACS Nano 2024;18:2500-19.

129. Bonuccelli G, Avnet S, Grisendi G, et al. Role of mesenchymal stem cells in osteosarcoma and metabolic reprogramming of tumor cells. Oncotarget 2014;5:7575-88.

130. Tang K, Hu L, Ma J, et al. Brief report: human mesenchymal stem-like cells facilitate floating tumorigenic cell growth via glutamine-ammonium cycle. Stem Cells 2015;33:2877-84.

131. Ni K, Wang D, Xu H, et al. miR-21 promotes non-small cell lung cancer cells growth by regulating fatty acid metabolism. Cancer Cell Int 2019;19:219.

132. Liu Z, Yu M, Fei B, Fang X, Ma T, Wang D. miR‑21‑5p targets PDHA1 to regulate glycolysis and cancer progression in gastric cancer. Oncol Rep 2018;40:2955-63.

133. Dai Q, Li N, Zhou X. Increased miR-21a provides metabolic advantages through suppression of FBP1 expression in non-small cell lung cancer cells. Am J Cancer Res 2017;7:2121-30.

134. Gu ZW, He YF, Wang WJ, Tian Q, Di W. MiR-1180 from bone marrow-derived mesenchymal stem cells induces glycolysis and chemoresistance in ovarian cancer cells by upregulating the Wnt signaling pathway. J Zhejiang Univ Sci B 2019;20:219-37.

135. Luo T, Liu Q, Tan A, et al. Mesenchymal stem cell-secreted exosome promotes chemoresistance in breast cancer via enhancing miR-21-5p-mediated S100A6 expression. Mol Ther Oncolytics 2020;19:283-93.

136. Giordo R, Ahmadi FAM, Husaini NA, et al. microRNA 21 and long non-coding RNAs interplays underlie cancer pathophysiology: a narrative review. Noncoding RNA Res 2024;9:831-52.

137. Raynor A, Jantscheff P, Ross T, et al. Saturated and mono-unsaturated lysophosphatidylcholine metabolism in tumour cells: a potential therapeutic target for preventing metastases. Lipids Health Dis 2015;14:69.

138. An J, Ha EM. Extracellular vesicles derived from Lactobacillus plantarum restore chemosensitivity through the PDK2-mediated glucose metabolic pathway in 5-FU-resistant colorectal cancer cells. J Microbiol 2022;60:735-45.

139. Alcayaga-Miranda F, González PL, Lopez-Verrilli A, et al. Prostate tumor-induced angiogenesis is blocked by exosomes derived from menstrual stem cells through the inhibition of reactive oxygen species. Oncotarget 2016;7:44462-77.

140. Infantino V, Santarsiero A, Convertini P, Todisco S, Iacobazzi V. Cancer cell metabolism in hypoxia: role of HIF-1 as key regulator and therapeutic target. Int J Mol Sci 2021;22:5703.

141. Chen D, Sun Y, Yuan Y, et al. miR-100 induces epithelial-mesenchymal transition but suppresses tumorigenesis, migration and invasion. PLoS Genet 2014;10:e1004177.

142. Deng L, Shang L, Bai S, et al. MicroRNA100 inhibits self-renewal of breast cancer stem-like cells and breast tumor development. Cancer Res 2014;74:6648-60.

143. Pakravan K, Babashah S, Sadeghizadeh M, et al. MicroRNA-100 shuttled by mesenchymal stem cell-derived exosomes suppresses in vitro angiogenesis through modulating the mTOR/HIF-1α/VEGF signaling axis in breast cancer cells. Cell Oncol (Dordr) 2017;40:457-70.

144. Bruno S, Collino F, Deregibus MC, Grange C, Tetta C, Camussi G. Microvesicles derived from human bone marrow mesenchymal stem cells inhibit tumor growth. Stem Cells Dev 2013;22:758-71.

145. Lou G, Chen L, Xia C, et al. MiR-199a-modified exosomes from adipose tissue-derived mesenchymal stem cells improve hepatocellular carcinoma chemosensitivity through mTOR pathway. J Exp Clin Cancer Res 2020;39:4.

146. Zou Z, Tao T, Li H, Zhu X. mTOR signaling pathway and mTOR inhibitors in cancer: progress and challenges. Cell Biosci 2020;10:31.

147. Gremke N, Polo P, Dort A, et al. mTOR-mediated cancer drug resistance suppresses autophagy and generates a druggable metabolic vulnerability. Nat Commun 2020;11:4684.

148. Lin D, Zhang H, Liu R, et al. iRGD-modified exosomes effectively deliver CPT1A siRNA to colon cancer cells, reversing oxaliplatin resistance by regulating fatty acid oxidation. Mol Oncol 2021;15:3430-46.

Cite This Article

Review
Open Access
Interconnections within the tumor microenvironment: extracellular vesicles as critical players of metabolic reprogramming in tumor cells
Carol Costa Encarnação, ... Mariana Renovato-MartinsMariana Renovato-Martins

How to Cite

Encarnação, C. C.; Faria G. M.; Franco V. A.; Botelho L. G. X.; Moraes J. A.; Renovato-Martins M. Interconnections within the tumor microenvironment: extracellular vesicles as critical players of metabolic reprogramming in tumor cells. J. Cancer. Metastasis. Treat. 2024, 10, 28. http://dx.doi.org/10.20517/2394-4722.2024.78

Download Citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click on download.

Export Citation File:

Type of Import

Tips on Downloading Citation

This feature enables you to download the bibliographic information (also called citation data, header data, or metadata) for the articles on our site.

Citation Manager File Format

Use the radio buttons to choose how to format the bibliographic data you're harvesting. Several citation manager formats are available, including EndNote and BibTex.

Type of Import

If you have citation management software installed on your computer your Web browser should be able to import metadata directly into your reference database.

Direct Import: When the Direct Import option is selected (the default state), a dialogue box will give you the option to Save or Open the downloaded citation data. Choosing Open will either launch your citation manager or give you a choice of applications with which to use the metadata. The Save option saves the file locally for later use.

Indirect Import: When the Indirect Import option is selected, the metadata is displayed and may be copied and pasted as needed.

About This Article

© The Author(s) 2024. Open Access This article is licensed under a Creative Commons Attribution 4.0 International License (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, sharing, adaptation, distribution and reproduction in any medium or format, for any purpose, even commercially, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Data & Comments

Data

Views
124
Downloads
22
Citations
0
Comments
0
2

Comments

Comments must be written in English. Spam, offensive content, impersonation, and private information will not be permitted. If any comment is reported and identified as inappropriate content by OAE staff, the comment will be removed without notice. If you have any queries or need any help, please contact us at support@oaepublish.com.

0
Download PDF
Share This Article
Scan the QR code for reading!
See Updates
Contents
Figures
Related
Journal of Cancer Metastasis and Treatment
ISSN 2454-2857 (Online) 2394-4722 (Print)

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/