Skip to main content

Exosomal cargos-mediated metabolic reprogramming in tumor microenvironment

Abstract

Metabolic reprogramming is one of the hallmarks of cancer. As nutrients are scarce in the tumor microenvironment (TME), tumor cells adopt multiple metabolic adaptations to meet their growth requirements. Metabolic reprogramming is not only present in tumor cells, but exosomal cargos mediates intercellular communication between tumor cells and non-tumor cells in the TME, inducing metabolic remodeling to create an outpost of microvascular enrichment and immune escape. Here, we highlight the composition and characteristics of TME, meanwhile summarize the components of exosomal cargos and their corresponding sorting mode. Functionally, these exosomal cargos-mediated metabolic reprogramming improves the "soil" for tumor growth and metastasis. Moreover, we discuss the abnormal tumor metabolism targeted by exosomal cargos and its potential antitumor therapy. In conclusion, this review updates the current role of exosomal cargos in TME metabolic reprogramming and enriches the future application scenarios of exosomes.

Background

Extracellular vesicles (EVs) are nanoscale cellular secretions that act as key mediators in many pathological/physiological processes [1, 2]. According to MISEV2018, EVs cover a variety of subtypes such as exosomes, microvesicles (MV, also known as microparticles and ectosomes) and apoptotic bodies [3,4,5,6]. In addition, some emerging subtypes of EVs have gained increasing attention, such as two non-membrane nanoparticles exomere and supermere [7, 8]. A bilayer membrane vesicle, called migrasome, generated at the end or crossover site of contractile filaments produced by the cell tail during directed cell migration, has gradually attracted interest since its discovery in 2014 [9, 10]. On this basis, the investigators further identified mitosomes, defined as migrasomes that contain mitochondria [11]. The field of EVs is currently in a developmental stage, in this review we will elaborate on the role of exosomes, a classical subtype of EVs, in the regulation of metabolic reprogramming in the tumor microenvironment around exosomes. (Table. 1).

Table 1 Major subtypes of EVs and their physiological characteristics

Exosomes are nanoscale bilayered vesicles that are actively released by cells into the extracellular fluid and carry a variety of genetic materials [13]. They are taken up by cells through autocrine or paracrine pathways, and can also be taken up by distant tissues or organs via the circulatory system, participating in a variety of physiological and pathological processes [14, 15]. Initially, exosomes were thought to be "redundant" substances released by cells, but with the progress in the field of "regulation of cellular vesicular transport", exosomes have gradually become a hot topic in basic and translational medicine [16].

Exosomes carry host cell-derived proteins, lipids, non-coding RNA (ncRNA) and metabolites. These bioactive substances can be involved in regulating intercellular communication between tumor cells and TME and mediating TME heterogeneity [17, 18]. Exosomal cargos participate in processes, such as deregulating cellular energetic, avoiding immune destruction and sustained angiogenesis [17, 19,20,21]. The presence of differentially expressed exosomal proteins, ncRNAs and metabolites profiles in many solid tumors suggests a potential clinical diagnostic and therapeutic value of exosomal cargos.

TME is a "sanctuary" for tumor cells. Previously, the primary goal of human tumor treatment was the direct elimination of tumor cells. With the introduction of the TME concept, tumor is no longer composed of aggregates of tumor cells, but of surrounding cells and non-cellular components [22]. TME not only provides the space and conditions for tumor cells to survive, but on the contrary, tumor cells can also modify TME by secreting exosomal cargos. By remodeling the metabolic pathways of non-tumor cells in TME, exosomal cargos can promote TME heterogeneity and provide a precursor to tumor recurrence or distant metastasis [23].

Main text

Main components of the tumor microenvironment

TME consist of tumor cells, resident and recruited host cells (mainly CAFs (cancer-associated fibroblasts) and immune cells), secreted products of the above cells (such as cytokines, chemokines), vasculature and ECM (extracellular matrix) [24, 25]. Specific metabolites (lactic acid, polyamines, and nitric oxide) are also present in TME [26]. These complex components contribute to the malicious behavior of TME in local resistance, immune escape and distant metastasis. (Fig. 1).

Fig. 1
figure 1

Mainly components of the TME and their biological characteristics. In the TME, NFs can differentiate into CAFs with highly expression of FAP, FGF2 and α-SMA. Adipocytes can differentiate into CAAs with low expression of PPARγ, C/EBPα, HSL, APN and FABP4. Resting stellate cells are rich in vitamin A, with highly expression of desmi and GFAP, whereas activated stellate cells lack vitamin A, with highly expression of α-SMA. CAFs, CAAs and stellate cells produce ECM by secreting components such as collagen. According to the different MHC molecules, T lymphocytes can activate into two main subtypes, CD4 + T or CD8 + T cells. M1-polarized TAMs can delay tumor progression with highly expression of IL-1β, CD80 and CD86. M2-polarized TAMs benefit to tumor progression with highly expression of IL-10, CD163 and CD206. Tumor microvascular tissue consists of ECs and pericytes. VEGFR and FGFR promote the maturation and migration of ECs. PDGFR can maintain pericyte stability the stability of pericytes. Up-regulated CCBE1, Adamts3, VEGFR-3 and its ligand VEGF-C favor lymphangiogenesis in TME

CAFs

CAFs are one of the main cell types involved in ECM remodeling in TME, with highly heterogeneity [27]. It is notable that CAFs are activated by normal fibroblasts (NFs) residing in the local microenvironment. Up-regulated FAP, FGF2 and α-SMA can promote the activation and heterogeneity of CAFs [28]. Activated CAFs contribute to the remodeling of ECM, which is conducive to tumor cell migration, invasion and treatment resistance [29]. In addition, the growth factor and chemokine secreted by CAFs are conducive to angiogenesis and the recruitment of immune cells [25, 30].

A previous report has proved the heterogeneity of CAFs. Researchers have characterized four CAF subtypes with different characteristics and activation levels in breast cancer. Costa et al. distinguished CAF into different subtypes (CAF-S1, 2, 3 and 4) based on the expression levels of CD29, FAP, α-SMA, PDGFRb, FSP1 and CAV1, among which CAF-S1 and CAF-S4 fibroblasts expressed a-SMA and were regarded as myofibroblasts [31]. In addition, CAF-S1 subset plays a crucial part in immunosuppression by increasing the survival rate of CD4 + CD25 + T lymphocytes and promoting their differentiation into CD25 + FOXP3 + cells, while CAF-S4 subset has no such activity [32]. Four CAF subtypes have also been identified in high-grade serous ovarian cancer (HGSOC). The expression of CXCL12β in CAF-S1 subset inhibits immune activity and is a reliable prognostic factor in HGSOC [33]. Further research found that CAF-S1 and CAF-S4 subsets accumulate in LN and are related to cancer cell invasion in the metastatic lymph nodes of breast cancer [34]. CAF-S4 subset promotes cancer cell invasion through NOTCH signal transduction, and patients with CAF-S4 subset are prone to distant metastasis [34]. These reports indicate that the heterogeneity of CAFs is closely related to the malignant phenotypes of TME.

TAMs

Tumor-associated macrophages (TAMs) are the major infiltrating immune cells in TME, with high heterogeneity and plasticity [35]. In the process of tumor formation, the proliferation of macrophages residing in the macrophage pool and the recruitment of monocyte-derived macrophages (MDM) to TME together represent TAM [36]. Generally, under the stimulation of different physiological and pathological factors, macrophages can differentiate into two phenotypes: classically activated (M1) and alternatively activated (M2) [37]. Among them, M1 participates in the pro-inflammatory response and also plays an anti-tumor effect, while M2 participates in the anti-inflammatory response and promotes tumor development through tumor immunosuppressive effects [38]. However, given the heterogeneity of TAMs, the M1/M2 dichotomy is no longer sufficient to explain the polarization of macrophages in TME. With the deepening of research, people are now more focused on the dynamics of the TAM polarized ecosystem [39]. Researchers have used NanoString gene expression profiles to show that TAMs in the peripheral blood of renal cell carcinoma (RCC) patients with does not completely conform to the traditional M1-M2 TAMs paradigm [40]. Some researchers have used a boolean model of macrophage polarization to simulate the activation of the corresponding phenotypes of M1, M2 and Nurse Like Cells to promote tumor activity [41]. More importantly, the heterogeneity of TAMs is not only reflected in the physical development of tumors, but there is significant heterogeneity in TAMs in autochthonous murine treated with two different treatment methods (molecular-targeted inhibitor and radiation) [42]. Therefore, the dynamics of M1 and M2 TAMs reflect the complexity of TME.

T lymphocytes

T cells recognize antigens on MHC molecules on the surface of antigen-presenting cells (APCs) to regulate tumor immune responses via the T cell receptor (TCR) [43]. T cells mainly include two types of CD4 + and CD8 + T cells. When MHC-II molecules are delivered to the surface of T cells, CD4 + T cells are activated and participate in the immune response by producing cytokines. CD4 + T cells with a high degree of plasticity in TME, and can be selectively different cytokine secreting cells in response to various signals [44]. Activated CD4 + T cells can be divided into various helper T cell subsets, such as Th1, Th2, Th9, Th17, Th22, Tfh and Treg [45, 46]. Each subset has its specific phenotype, cytokine profile and function. Previous studies believed that these phenotypes are mutually exclusive, and cross-phenotype expression will not occur. In fact, there have been reports of Th17Th1 cells that have both Th17 and Th1 characteristics [47]. CD8 + T cells detect antigens presented by MHC-I molecules through a cross-presentation mechanism, leading to cytotoxic reactions that lead to tumor cell death, thus called CD8 cytotoxic T lymphocytes (CTLs) [48]. After tumor infiltration, the initial CD8 + T cells differentiate into effector CD8 + T cells, and further differentiate and activate into cytotoxic and memory CD8 + T cells [49, 50]. Therefore, the two key factors for CD8 + T cells to exert anti-tumor effects are: induction of T cell differentiation Infiltrate with CD8 + T to the tumor site.

Generally speaking, in view of the cytotoxicity of CD8 + T cells, the increased number of CD8 + T cells in TME is often associated with a good prognosis [51,52,53]. However, CD4 + T cells have both anti- and pro-tumorigenic roles due to their different subsets and polarization [52, 54,55,56]. Reports in recent years have shown that in addition to antigen presentation signals and costimulatory molecules, epigenetic modification, metabolism, and iron death may also be involved in the differentiation and function of T cells [50, 57,58,59]. Usually, the increased number of CD8 + T cells in TME contributes to the anti-tumor response, exhaustion and dysfunction of CD8 + T cells contribute to tumor progression.

Blood vessels

Tumor growth and metastasis are inseparable from adequate oxygen and nutrient supply, which all depend on the tumor vasculature. Compared with normal tissues, the tumor vasculature exhibits tortuosity and dysfunction, which is reflected in the heterogeneous permeability [60]. Normally, abnormal vascular function in the tumor microenvironment leads to hypoxic environment, blocking of immune cell infiltration and low drug delivery efficiency [61,62,63].

Tumor vasculature mainly includes vascular sprouting, angiogenesis and vascular mimicy [64, 65]. Three important regulatory factors and their receptors are involved in the regulation of neovascularization: VEGF-VEGFR, PDGF-PDGFR and FGF-FGFR [66]. In general, VEGFR and FGFR pathways are involved in activation of endothelial cells (ECs) and maturation of blood vessels [67]. PDGFR pathway provides support for pericytes [68]. PDGFR and FGFR regulate cell migration and adhesion, and maintain the stability of blood vessel walls [66].

Lymphatic vessels

Tumor lymphatic vessels belong to the tumor vasculature, and their main function is to remove the interstitial fluid (ISF) formed by capillary filtrate and tissue immune surveillance [69]. Lymphatic vessels are composed of lymphatic endothelial cells (LECs) [70]. LEC signaling molecules VEGFR-3 and its ligand VEGF-C are the main driving factors of pathological lymphangiogenesis, and the extracellular protein CCBE1 and metalloproteinase Adamts3 are also involved [71,72,73,74,75]. Moreover, fatty acid metabolism has a regulatory effect on lymphangiogenesis [76].

The role of tumor lymphatic vessels in tumor progression has advantages and disadvantages. Certain metastatic tumors can penetrate into the lymphatic vessels and cause distant metastasis, such as breast cancer, nasopharyngeal cancer, and prostate cancer [77,78,79]. However, functional lymph nodes enhance the tumor's response to immune checkpoint inhibitors by reducing the ISF, which is helpful for immunotherapy [80, 81]. In addition, lymphatic vessel remodeling is closely related to tumor immunity. A recent study showed that melanoma-derived EVs deliver tumor antigens to LECs in lymph nodes for cross-presentation on MHC-I, resulting in apoptosis induction in antigen-specific CD8 + T cells [82, 83]. This also reflects the heterogeneity of lymphatic vessels in the TME.

Adipocytes

Adipocytes are one of the stromal cells in the TME. As a storage site for triglycerides, adipocytes not only participate in the process of energy metabolism, but also secrete various metabolites to regulate TME [84]. There are three main types of adipocytes: white adipocytes, brown adipocytes and beige adipocytes. Among them, white adipocytes are the most widely distributed adipocytes. Studies have shown that white adipose tissue (WAT) is associated with the risk of breast cancer and prostate cancer [85, 86]. In some cancer-related cachexia, the transformation of white adipocytes to brown adipocytes and atrophy may contribute to the exacerbation of cachexia [87, 88]. Brown adipocytes are cells that break down metabolic substrates such as glucose to generate heat. A new study shows that based on the dependence of cancer cells on glucose, cold exposure can activate brown fat cells to compete with cancer cells for glucose [89]. This reveals positive role of brown adipocytes in tumors. Beige adipocytes may be differentiated by a specific class of adipose progenitor cells (APCs) with plasticity [90, 91]. Beige adipocytes reduce adhesion of tumor and non-tumor mouse mammary epithelial cells, favoring tumor development [92].

Adipocytes in TME can be transformed into cancer-associated adipocytes (CAAs) after communicating with tumor cells. Compared with mature adipocytes, CAAs have smaller size, irregular shape and dispersed lipid droplets [93]. In addition to cellular morphological and structural heterogeneity, activated CAAs are associated with a reduction in markers of terminal differentiation (PPARγ, C/EBPα, HSL, APN, and FABP4) [84]. Metabolites (leptin, adiponectin, lactate, fatty acids and glutamine) secreted by CAAs can be taken up by adjacent cells in the TME to induce angiogenesis and immune escape [94, 95]. CAA-mediated secretion and processing of collagen IV induces the programming of ECM [96].

Stellate cells

Stellate cells originate from mesothelial and submesothelial cells with highly plasticity. The common stellate cells in the TME include hepatic stellate cells (HSCs) and pancreatic stellate cells (PSCs). Resting stellate cells are rich in vitamin A, accompanied by highly expression of desmi and glial fibrillary acidic protein (GFAP). Conversely, activated stellate cells lost vitamin A, accompanied by highly expression of α-SMA [97, 98]. Exosomes released by cancer cells can activate stellate cells, for example, exosomal miR-181a-5p and exosomal miR-21 derived from hepatic carcinoma cells can activate HSCs [99, 100]. Pancreatic cancer cells-derived exosomal miR-1290 and exosomal protein Lin28B can activate PSCs [101, 102]. Activated stellate cells can secrete VEGF, FGF, interleukins and matrix metalloproteinases (MMPs) to promote angiogenesis, inflammatory infiltration and ECM precipitation in TME [103].

ECM

ECM is composed of a variety of proteins and macromolecules, including collagen, glycoprotein, elastin, fibronectin and proteoglycan, which are mainly secreted by CAFs [104]. ECM has complex mechanical behaviors, and the impact of matrix stiffness on stem cells and tumor cells is a current research hotspot.

ECM has complex mechanical behavior, and the viscoelasticity and tension of the matrix can promote the stemness, metastasis and drug resistance of tumor cells [105, 106]. The increased stiffness of ECM and the remodeling of the basement membrane are conducive to tumor metastasis [107, 108]. Among them, type I collagen fibrin is very important to the stiffness of ECM. Type I collagen fibrin is a component of ECM, and its tensile strength is regulated by two enzymes: lysyl oxidases (LOXs) and lysyl hydroxylases (LHs) [109, 110]. Beyond that, the metabolism of hyaluronic acid and glucose can also lead to the remodeling of ECM components, which in turn affects cell migration [111, 112].

Hypoxia

Hypoxia is a hallmark of TME. Hypoxia or inadequate oxygenation is a key factor in the difficulty of eradicating tumors and also predisposes tumor cells to treatment resistance [113]. Causes of hypoxia in TME include the heterogeneity of the tumor vascular system and the exuberant metabolism of the tumor cells, where the homeostasis of oxygen supply and consumption is disrupted [114]. In hypoxia, tumor cells can activate a range of adaptive changes through hypoxia-inducible factors (HIFs). For example, hypoxia promotes the release of exosomes in a HIF-1α-dependent manner, where the exosomal miR-310a-p inhibits the ubiquitinated degradation of HIF-1α by targeting PHD3 [115]. This positive feedback loop promotes GC cell proliferation, invasion and EMT [115]. HIF-2α can induce stemness in breast cancer cells via the SOD2-mtROS-PDI/GRP78-UPRER pathway in hypoxia [116]. HIFs can enhance angiogenesis by regulating the expression of VEGF and MMP [117]. Hypoxia is associated with LOXs and LHs, the regulatory enzymes of collagen fibronectin, suggesting that hypoxia may induce changes in the physical characteristics of the ECM [118, 119]. This suggests that hypoxia is both a feature of TME and a regulatory factor.

Hypoxia can regulate exosome release. Tumor cells are more likely to release exosomes in the hypoxic TME, which is related to the activation of the small GTPase Rab27a, a major regulator of exosomal synthesis, regulated by HIF-1α [120]. It has been shown that hypoxia-treated natural killer cells secrete more exosomes compared to normoxic conditions [121]. Hypoxia may promote exosome release by regulating exosome biogenesis, which includes intraluminal vescicle (ILV) biogenesis and multivesicular endosome (MVE) transport [122, 123]. In addition to Rab27a, HIF-1α may also be involved in hypoxia-mediated exosome release through activation of Rab5a, Rab7, Rab22, RhoA, and ROCK [124,125,126]. Glycolysis may also be involved in the facilitation of exosomes release by hypoxia [127].

Origin and hallmarks of exosomes

Exosomes are small extracellular vesicles (sEVs) with a diameter between 40-160 nm [12]. It originates from the endosome and is widely present in blood, urine, ascites, and cerebrospinal fluid[35]. Exosomes contain many components of cells, such as DNA, RNA, proteins, lipids and cellular metabolites [12]. Even a new study found that mitochondria from fat cells can be transported between organs through small extracellular vesicles [128]. Exosomes are internalized by receptor cells through receptor-mediated endocytosis, pinocytosis, phagocytosis, or fusion with cell membranes, resulting in the direct release of their cargo into the cytoplasm [129]. This cell-to-cell communication has been shown to change the function of recipient cells and is widespread in TME [130].

Origin of exosomes

The specific mechanism of exosomes formation is still unclear, and the endosomal sorting complex (ESCRT) required for transportation is currently recognized as a classic pathway. The exosomes biogenesis pass through the stages of early sorting endosome, late sorting endosome, ILV and multivesicular body (MVB). Finally MVB is subsequently degraded by lysosomes or fused with the plasma membrane to release its contents, including exosomes [131]. The sorting of MVB into exosomes may have a specific regulatory mechanism, but it is currently poorly understood. It is reported that Rab27A and Rab27A, members of the GTPase Rab family, play a role in special types of secretion (such as exosome secretion and mast cell secretion) [132,133,134]. With the deepening of research, Rab35, Rab11 and Rab7 are also involved in the process of MVB fusion with the plasma membrane and release of exosomes [135,136,137,138]. In addition, studies have shown that actin cytoskeleton regulatory protein (cortactin) can bind to the branched actin nucleation Arp2/3 complex and further control the fusion of MVB with the plasma membrane [139].

Characteristics and purification of exosomes

Several proteins are related to the biogenesis of exosomes, such as Rab GTPase and ESCRT protein. Exosome surface proteins including transmembrane 4 superfamily (CD9, CD63, CD81), lipid raft protein (flotillin-1), and Ceramide are often used as biomarkers for exosomes [12]. Some proteins enriched in exosomes are also commonly used as biomarkers for exosomes, such as HSP70, TSG101, and ALIX [140, 141]. Researchers can identify exosomes and their biogenesis based on these biomarkers, but there is still a lack of exosomes-directed tracking systems.

Based on the potentials of exosomes in the diagnosis and treatment of tumors, there is an urgent need for high-efficiency isolation of exosomes [142]. Density gradient differential ultracentrifugation (DGUC) has always been the most classic exosome purification method. On the basis of DGUC, an improvised one-step sucrose cushion ultracentrifugation method for exosome isolation is beneficial to maintain the integrity of exosomes and remove protein contamination [143]. In addition, many exosome separation technologies based on different principles have also been applied, such as size exclusion chromatography (SEC), ultrafiltration (UF), Anion exchange chromatography (AIEX) polymer-based precipitation, and immunoaffinity capture [144, 145]. In recent years, microfluidic-based exosome isolation techniques has been developing rapidly. Compared with traditional separation technology, microfluidic device can separate exosomes in various samples with high selectivity and high yield, while reducing processing time, cost and sample consumption [145]. Another method that uses electricity and acoustic forces to manipulate biological particles and submicron particles for deterministic sorting has been applied to the purification of exosomes. The purity of the exosomes purified by this method is more than 95% and the recovery rate is 81% [146].

Cargos in exosomes

Exosome cargos are the core components that confer biological effects on exosomes. Nearly 100,000 proteins and over 1,000 lipids have been reported to be associated with exosomes [147]. Nucleic acids mainly mRNAs and ncRNAs were enriched in exosomes, including more than 27,000 mRNAs and more than 10,000 ncRNAs were identified in sEV [147]. Genomic DNA (gDNA) and mitochondria DNA are present in exosomes in the form of s single-stranded or double-stranded [148, 149]. Together, these biologically active substances make up the exosomal cargos. (Fig. 2).

Fig. 2
figure 2

Biomarkers of exosomes and main components of exosomal cargos. Several exosome surface proteins are considered to be biomarkers of exosomes, including transmembrane 4 superfamily (CD9, CD63, CD81), lipid raft protein (flotillin-1), and Ceramide. Exosome content proteins HSP70, TSG101 and ALIX are also biomarkers of exosomes. Exosomes carry a variety of cargos, such as nucleic acids, proteins, enzymes and metabolites (mainly lipids)

Nucleic acids

The role of exosomal RNAs in tumors has been widely reported, mainly including mRNAs and ncRNAs. Regarding functional exosomal mRNAs, an early study reported that exosomal mRNAs can complete translation in receptor cells [150]. Recently, exosomal mRNAs CUL9, KMT2D, PBRM1, PREX2 and SETD2 were found to be possible novel potential biomarkers for clear cell renal cell carcinoma (ccRCC) [151]. Studies on exosomal ncRNAs have focused on miRNAs, lncRNAs & circRNAs. Usually exosomal ncRNAs are transported to the recipient cells as molecular sponges.

Exosomal DNA mainly includes gDNA and mtDNA [149]. The TME of "hibernating" cancer cells secretes EVs containing mtDNA, leading to endocrine therapy resistance in breast cancer cells [152]. T cells secrete EVs containing gDNA and mtDNA, which activate the cGAS/STING signaling pathway and induce antiviral responses in DCs [153].

Lipids

The lipid cargos of exosomes mainly include sphingolipids, cholesterol, phosphatidylserine, saturated fatty acids, and ceramides, which are mainly associated with exosome biogenesis [154]. The bilayer lipid membrane structure of exosomes determines the enrichment of membrane lipid components (phosphoglycerolipids, sphingolipids, and sterols) [147]. The fact that neutral sphingomyelinase inhibitors can reduce exosome secretion further illustrates the importance of membrane lipids for exosomes [155]. Cholesterol enrichment in exosomes is associated with MVB. Different subcellular organelles have different cholesterol concentrations. Oxysterol-binding protein-related proteins (ORPs) are involved in cholesterol transport and are able to maintain the proper cholesterol concentration required for MVB biogenesis [156]. In addition, when low-density lipoprotein-cholesterol is low in endosomes, endoplasmic reticulum stress-derived cholesterol can be transferred to MVB [157].

Some biologically active lipids are important cargos for exosomes to perform their biological functions. LTB4 is packaged and released in exosomes [158]. This LTB-enriched exosome biogenesis originates from the nuclear envelope of centrocytes and is an unconventional pathway of exosome production [159]. Granulocyte myeloid-derived suppressor cells can secrete exosomal PGE2 to ameliorate collagen-induced arthritis [160]. Ubiquitination of 15-LO2 in hypoxia promotes 15-LO2 sorting to exosomes, which are involved in the regulation of pulmonary vascular homeostasis [161].

Proteins

Proteins are important cargos of exosomes. Protein cargos from different sources of exosomes are heterogeneous, but the proteins involved in exosome biogenesis, exosome content sorting and exosome release are invariant. These protein cargos are as described in Sect. 2.2.2. Exosomal heterogeneity is mainly reflected in protein cargos with signal transduction function and enzyme cargos with biological activity [162]. The main exosomal signaling proteins include EGFR, VEGF, TGF-β, PTEN and STAT [163,164,165,166]. Enzymes mainly include metabolic enzymes (such as ATPase, pyruvate kinase, fatty acid synthases), RNA editing enzymes, proteases, glycosyl transferases, glycosidases [147, 167, 168]. In addition, some special exosomal proteins deserve attention. Exosomal PD-L1 is a biomarker for tumor diagnosis and immunotherapy efficacy prediction. There is glycosylation heterogeneity in exosomal PD-L1, and the sensitivity and specificity of glycosylated exosomal PD-L1 is superior compared to exosomal PD-L1 [169]. Exosomal BECN1 participated in the regulation of ferroptosis [170].

Cargos sorting to exosomes

In general, nucleic acids can be sorted to exosomes through the interaction of RNA-binding proteins or lipid rafts in MVB [171]. Some exosomal biomarkers may regulate the process of protein sorting into exosomes [172]. The sorting of cellular metabolites may be related to the process of exosome formation [159]. In addition, there are some special mechanisms involved in the sorting of exosomal cargos. (Table 2).

Table 2 Different ways of various cargos sorting to exosomes

Nucleic acids sorting to exosomes

Initially, four mechanisms of miRNA sorting to exosomes were proposed. The neural sphingomyelinase 2 (nSMase2), heterogeneous nuclear ribonucleoprotein (hnRNP), uridylation at 3’ends and argonaute 2 (Ago2) are involved in this process [173,174,175,176]. Recent studies have shown that two RNA binding proteins, Alyref and Fus, mediate miRNA sorting into sEVs, enriching the understanding of exosomal miRNA biogenesis [99, 177]. In addition, activation of the NLRP3 inflammasome and cleavage of RILP increase exosome production, and the cleaved form of RILP interacts with FMR1 to regulate exosomal miR-155 content [178].

Some lncRNAs are also found in exosomes, probably by forming lncRNA–RBP complexes [179]. Although the specific mechanism of lncRNA sorting to exosomes is still unclear. Interestingly, lncRNA-encoded microproteins were identified in glioma-derived exosomes, indicating the biological functional diversity of exosomal lncRNAs [188].

CircRNA is a component of exosomal ncRNAs, and different circRNAs were found in exosomes originating from different cells, indicating that the sorting process of exosomal circRNAs is selective [189]. SNF8, a key component of the ESCRT-II complex, sorts circRHOBTB3 into exosomes by binding to specific sequences (141-240nt) on circRHOBTB3 [180]. After ectopic expression of miR-7 in HEK293T and MCF-7 cells, the level of circRNA CDR1as was significantly downregulated in exosomes but slightly increased in cells. This result partially suggests that sorting of circRNAs to exosomes was regulated by changes of associated miRNA levels [181].

The sorting mechanism of exosomal DNA is still unclear. Certain physiological pathways may be involved in this sorting process. MtDNA sorting to exosomes may be related to the endosomal pathway [153], and gDNA sorting to exosomes may involve micronuclei (MN) [182].

Protein sorting to exosomes

Typically, proteins can enter cells together with cell surface proteins by endocytosis and invagination of the plasma membrane [12]. The protein sorting process can be done in organelles such as mitochondria, endoplasmic reticulum and Golgi apparatus [190]. During the budding stage of exosome biogenesis, early-sorting endosome (ESE) may fuse and communicate with mitochondria, endoplasmic reticulum and trans-Golgi network, indicating the reason why the protein can be detected in exosomes of host cells [12]. In addition, some regulatory proteins in the formation, transport and secretion of exosomes may be involved in the sorting process of exosomal proteins, such as Rab GTPase, ESCRT proteins, tetraspanins and SNARE protein complexes [172, 191,192,193].

Some non-canonical mechanisms are also involved in the regulation of exosomal protein sorting. A recent study shows that proteins containing the KFERQ pentapeptide sequence can be sorted to exosomes by a process dependent on the membrane protein LAMP2A, a novel mechanism independent of ESCRT [183, 184]. Epigenetic modifications can be involved in the sorting of exosomal proteins. Cav1 can be sorted to MVB and ILV in a phosphorylation and ubiquitination-dependent manner, regulates exosome biogenesis by regulating MVB cholesterol contents, and delivers specific ECM-associated proteins (Tenascin-C, fibronectin, nidogen, elilin, EDIL3 and heparan sulfate proteoglycans) to exosomes [185]. Some proteins have UBL domains (ubiquitin-like sequences), among which UBL3/MUB proteins, as one of the conserved UBLs, can act as post-translational modification (PTM) factors to regulate the process of protein sorting to sEVs [186]. In addition, endosomal microautophagy is also involved in the sorting of exosomal proteins, and the chaperone HSC70-mediated proteins are sorted into exosomes under the electrostatic interaction between the cationic domain of HSC70 and the MVB membrane [187, 194].

Metabolites sorting to exosomes

Exosomes contain intact metabolites (mainly lipid metabolites) with characteristic of the host cells [195]. For example, exosomes secreted by granulocyte-myeloid derived suppressor cells (G-MDSCs) are enriched in PGE2, exosomes secreted by neutrophils are enriched in LTB4 [158, 160]. Triglycerides (TG) and sphingomyelin were found in T cell-derived exosomes isolated from human plasma [196]. Metabolomics of cancer stem cell-derived exosomes from melanoma revealed that multiple lipid metabolites, such as glycerophosphoglycerol (PG), glycerophosphatidylserine (PS), TG, and glycerophosphorylcholine (PC) in exosomes [197]. Currently, the mechanism by which these metabolites are sorted to exosomes remains unclear, but may be related to the interaction of lyn and fotillin-1 through the lipid domains of exosomal lipid membranes [159].

Biology of tumor metabolic reprogramming

Metabolic abnormalities are one of the hallmarkers of tumor cells, which are metabolically reprogrammed to meet their rapid proliferation requirements [198]. Given the specific physicochemical characteristics of high pressure, high pH and hypoxia within the TME, as well as the heterogeneity of the tumor vasculature, local tumor cells often have limited metabolic resources, which accelerates the digestion of nutrients and the accumulation of metabolites. In this particular environment, tumor cells adjust their metabolism in order to maintain their growth, which not only allows them to re-meet their energy supply needs, but also regulates their gene expression and protein modifications, facilitating the spread of tumor cells [199].

Glucose metabolism

Glucose is the main source of energy for cellular metabolism and biosynthesis. Glucose metabolism includes the glycolytic pathway, the pentose phosphate pathway (PPP), the serine synthesis pathway (SSP) and the tricarboxylic acid (TCA) cycle pathway [200]. Tumor cells are able to re-edit these pathways to obtain ATP with various biological macromolecules. In contrast to normal cells, tumor cells produce large amounts of lactate via the aerobic glycolytic pathway, resulting in an acidic TME that contributes to the proliferation [201]. In addition, increased uptake of glucose leads to the accumulation of intermediate metabolites in the glycolytic pathway, activating the cellular PPP while inhibiting the intracellular TCA[202]. Activation of the PPP provids NADPH to tumor cells, while inhibition of the TCA lead to a decrease in intracellular ROS and promoted tumor cell proliferation [203].

Heterogeneity of TME is associated with glucose metabolism. It was shown that hypoxic conditions in TME promoted HIF-1-induced glycolysis [204]. HIF-1 promotes glycolysis by upregulating hexokinase and glucose transporters, while inhibiting mitochondrial biosynthesis [204]. HIF-1 stimulates pyruvate kinase 2 (PKM2) into the nucleus to drive transcription of glycolysis-related genes [205]. HIF-1 not only regulates pH through nanohydrogen pump, but also regulates glycolysis by up-regulating hexokinase, aldolase, pyruvate kinase and downregulating pyruvate dehydrogenase to promote the conversion of pyruvate to acetyl coenzyme A, which enters the citric acid cycle [206, 207]. Abnormalities in glucose metabolism can directly affect tumor cells and non-tumor cells in TME. Up-regulated PPP in most tumor cells correlates with resistance to radiotherapy [208]. Up-regulated TCA in breast cancer cells promotes α-ketoglutarate production and facilitates tumor metastasis [209, 210]. In addition, up-regulated glycolysis in GC and NSCLC cells can regulate PD-1 expression in Treg [211]. Increased glycolysis in Treg may affect the therapeutic effect of CTLA-4 blockade, which is associated with immune infiltration in the TME [212].

Fatty acid metabolism

Abnormal fatty acid (FA) metabolism is particularly noteworthy, as it is structural components of the membrane matrix and important secondary messengers that can serve as a fuel source for energy production [213]. In mammalian cells, FA can either be taken up directly from the surrounding environment or synthesized from scratch through nutrients. The uptake of exogenous FA requires specialized transporter proteins to achieve transmembrane, including CD36, fatty acid transporter protein (FATP) family and plasma membrane fatty acid binding protein (FABPpm), all of which are highly expressed in tumors [214]. Fat from scratch synthesis is a process that uses carbon from glucose and amino acids to convert to FA. Normally, fat from scratch synthesis is restricted to hepatocytes and adipocytes, however, tumor cells can reactivate this metabolic pathway [198].

Glucose or glutamine in tumor cells is oxidized or reversibly carboxylated by pyruvate in the TCA cycle to generate citrate, respectively [215]. Citrate is converted to acetyl coenzyme A by ATP-citrate lyase (ACLY), followed by irreversible carboxylation to malonyl coenzyme A. Finally, condensation of seven malonyl coenzyme A with one acetyl coenzyme A is catalyzed by fatty acid synthase (FASN) to produce saturated palmitic acid [213]. Palmitic acid can be converted to other FA species (like phospholipids and triglycerides) through different pathways, contributing to transmembrane signaling in tumor cells, while regulating the structure and fluidity of cell membranes and promoting epithelial-mesenchymal transition (EMT) [216].

FA contributes to the remodeling of TME. Arachidonate is an important class of bioactive lipid molecules, such as prostaglandins, leukotrienes and ω-hydroxylase [213, 217, 218]. The main enzymes involved in prostaglandin production are prostaglandin G/H synthases COX1 and COX2, are closely related to inflammation [219,220,221]. In addition, accumulation of FA in TME can lead to CD8 + T cell dysfunction in the pancreas [222, 223]. FA oxidation can promote IL-1β secretion by M2-type mononuclear macrophages and remodel tumor metabolism with TME [224].

Amino acid metabolism

Amino acids are one of the raw materials for cellular synthesis of biomolecules, such as proteins, lipids, and nucleic acids. In contrast to normal cells, tumor cells require large uptake of amino acids for malignant development, as well as other amino acids to provide a source of carbon and nitrogen [225]. Reprogramming of amino acid metabolism plays an important role in tumor.

Glutamine (Gln) is one of the most abundant non-essential amino acids in the body, a recent study has shown that it is Gln that is the highest nutrient intake by cancer cells [226]. Glycolysis is the main way for tumor cells to obtain energy, and the metabolism of glucose is regulated by Gln, which is able to replace glucose as the main energy source of TCA under hypoxic microenvironment [226]. The synergistic effect between Gln and leucine can promote the breakdown of Gln to produce α-ketoglutarate and activate mTORC1,which promotes the proliferation of tumor cells [227]. Serine (Ser) is an important one-carbon unit raw material. It also serves as a methyl donor and is involved in the methylation modification of biological macromolecules [228]. Under hypoxic microenvironment, the expression of serine hydroxymethyltransferase 2 (SHMT2) was upregulated to promote the production of NADPH and glutathione to maintain redox homeostasis [229]. Tryptophan (Trp) is one of the essential amino acids and can be metabolized through three pathways: kynurenine (Kyn), 5-hydroxytryptamine (5-HT) and indole [230]. Trp may be a tumor biomarker. Trp metabolites may be biomarkers of esophageal cancer susceptibility, metastasis and prognosis [231]. Trp metabolite Kyn is overexpressed in advanced colorectal cancer and induces CD8 + T cell exhaustion [232]. Phenylalanine-tryptophan may be a combination biomarker for early diagnosis of hepatocellular carcinoma [233].

Disturbance in amino acid metabolism can remodel the immune microenvironment of tumors. It has been shown that Glutamine metabolism can regulate the immunosuppressive function of myeloid-derived suppressor cells (MDSCs) [234]. Glutamine small-molecule inhibitor not only inhibits tumor growth, but also suppresses MDSCs production and recruitment. Targeting tumor glutamine metabolism leads to a decrease in CSF3, resulting in an increase in inflammatory TAM [235, 236]. In a study of renal cancer, CAFs were found to up-regulate tryptophan 2, 3-dioxygenase (TDO) expression, resulting in enhanced secretion of Kyn, which ultimately activate AKT and STAT3 signaling pathways and induce chemoresistance [237]. (Table 3).

Table 3 Effects of metabolic changing on the TME

Exosomal cargos-mediated metabolic reprogramming in TME

Tumor occurrence not only requires the metabolic reprogramming of cancer cells, but also the metabolic reprogramming of non-cancer cells in TME also participates in tumor progression. Exosomal cargos, as a type of intercellular communication messenger, mediate the metabolic regulation of different types of cells in TME [269, 270]. After the exosomal cargos secreted by cancer cells are accepted by recipient cells, the recipient cells undergo changes in various metabolic pathways, in which energy metabolism is reprogrammed to meet energy supply and biosynthesis. The secretion of various metabolites emphasizes the heterogeneity of TME [26, 271]. (Table 2).

Exosomal cargos induce metabolic reprogramming of CAFs

CAFs are the most common cell type in TME and the major cells producing ECM. The metabolic reprogramming of CAFs is beneficial to the growth and metastasis of tumor cells. Studies have reported that exosomal miRNAs are involved in the regulation of glucose and lipid metabolism of CAFs [100, 238]. The exosomal miR-105 secreted by breast cancer (BC) targets MXI1, activates the MYC pathway in CAFs, enhances the glycolysis and glutamine decomposition of CAFs, and detoxifies the metabolites (lactate and NH4 +) to fuel adjacent cancer cells [238]. ITGB4 is highly expressed in various cancers and contributes to tumor progression [272]. Studies on triple-negative breast cancer (TNBC) found that cancer cell-derived exosomal ITGB4 could be delivered to CAFs to induce BNIP3L-dependent mitophagy and lactate production in CAFs. Inhibition of exosomal ITGB4 delivery can inhibit mitophagy and glycolysis in CAFs [239]. Colorectal cancer (CRC) cells-derived exosomes can regulate metabolic reprogramming of CAFs, upregulation of glycogen metabolism (GAA), amino acid biosynthesis (SHMT2, IDH2) and membrane transporters of glucose (GLUT-1), lactate (MCT4), and amino acids (SLC1A5/3A5) [240]. HCC-derived exosomal HSPC111 induces differentiation of HSCs into CAFs, and exosomal HSPC111 reshapes lipid metabolism of CAFs by regulating ACLY, up-regulating acetyl-CoA levels and down-regulating citrate levels [241]. These exosome-mediated metabolic reprogramming can induce the secretory function of CAFs and enhance the heterogeneity of the TME [273]. (Fig. 3).

Fig. 3
figure 3

The mechanism of tumor-serected exosomal cargos regulate metabolic reprogramming of CAFs. TEXs can be delivered to NFs and stellate cells. Breast cancer cells-secreted exosomal miR-105 can target MXI1 to activate the MYC pathway and enhance the glycolysis, glutamine breakdown, lactate secretion and NH4 + clearance of CAFs. HCC cells-secreted exosomal HSPC111 can promote ACLY expression to alter lipid metabolism of CAFs. TNBC cells-secreted exosomal ITGB4 can induce BNIP3L-dependent lactate production in CAFs. CRC-derived exosomes can promote GAA of CAFs and up-regulate SHMT2, IDH2, GLUT-1, MCT4 and SLC1A5/3A5 expression

Exosomal cargos induce angiogenesis through metabolic reprogramming

Exosomal cargos are involved in the abnormal metabolism of ECs to promote tumor angiogenesis. A study has shown that endothelial progenitor cell-derived (EPC-derived) exosomal miR-210 can reduce ROS production and promote ATP production in ECs, contributing to the maintenance of ECs stability [242]. Acute myeloid leukemia (AML) cells-derived exosomes containing VEGF and VEGFR messenger RNA induce VEGFR expression in human umbilical vein endothelial cells (HUVECs). This modulation enhances glycolysis, leading to vascular remodeling and chemoresistance [243]. A study about head and neck squamous cell carcinoma (HNSCC) found that tumor cell-derived exosomes (TEXs) carry enzymatically active CD39, CD73 and adenosine (ADO), which directly induce ECs growth [244, 274]. In addition, these exosomal cargos can also induce macrophages to an angiogenic phenotype (mainly M2 polarization), the M2-polarized macrophages further secrete several pro-angiogenic factors (Angiopoietin-1, Endothelin-1, Platelet Factor 4 and Serpin E1) to stimulate the function of ECs. This dual effect (direct and indirect) promotes angiogenesis. (Fig. 4).

Fig. 4
figure 4

The mechanism of tumor-serected exosomal cargos induce angiogenesis through metabolic reprogramming. TEXs can be delivered to various types of cells to induce angiogenesis. AML cells-derived exosomal VEGF and VEGFR mRNA can enhance ECs glycolysis and induce angiogenesis. HNSCC cells-derived exosomal CD39, CD73 and metabolite ADO can induce A2BR-mediated M2-polarized TAMs and promote their secretion of angiogenic factors to induce angiogenesis. Exosomal CD39, CD73 and ADO can also directly promote ECs growth. EPC-derived exosomal miR-210 maintains vascular stability by adjusting ROS and ATP levels

Exosomal cargos induce metabolic reprogramming of immune cells

TEX-mediated metabolic reprogramming is associated with tumor immunity. It was found that purine metabolites (adenosine and inosine) enriched in TEXs may promote immune escape, and the levels of purine metabolites in circulating exosomes may suggest clinical stage and lymph node metastasis in HNSCC patients [275,276,277]. T lymphocytes play a key role in the tumor immune response. MiR-451 is a cell metabolism-related miRNA. Exosomal miR-451 can redistribute from gastric cancer cells with low glucose concentration to T cells and promote Th17 polarized differentiation of T cells by decreasing AMPK activity and increasing mTOR activity [245]. The cervical squamous carcinoma-derived exosomal miR-142-5p targets ARID2 to inhibit DNMT1 recruitment to the INF-γ promoter, leading to upregulation of IDO expression, thereby suppressing and exhausting CD8 + T cells [246]. TEXs can deliver a sustained signal to Treg, resulting in the conversion of extracellular ATP to inosine and inhibition of Treg function. This regulatory mechanism is dependent on surface signaling and does not require internalization of TEXs by the recipient cells [247].

Tumor antigen presentation by dendritic cells (DCs) is the initiating step of the immune response in vivo [278]. A study on prostate cancer exosomes and tumor antigen presentation found that exosomal Rab27a could induce the expression of CD73 on DCs. CD73 hydrolyzes AMP to adenosine and inhibits the production of TNFα and IL-1L in an ATP-dependent manner, which inhibits the function of DCs[248]. Exosome-mediated metabolic reprogramming induces M2 polarization in macrophages and contributes to tumor progression. Hypoxia-induced TEX can carry let-7a, and exosomal let-7a enhances OXPHOS in macrophages and inhibits the insulin-AKT-mTOR signaling pathway. This leads to M2 polarization of macrophages [249, 279]. TEX carry enzymatically active CD39, CD73 and ADO induce M2 polarization of macrophages through A2BR-dependent signal transduction [244]. Hypoxia is one of the features of TME, and glucose metabolism is associated with hypoxia [210]. PKM2 is one of the key enzymes of glycolysis, and studies have shown that under hypoxic conditions, lung cancer cells-derived exosomal PKM2 induces M2-polarized macrophages by activating the AMPK signaling pathway, in which exosome-mediated remodeling of glucose metabolism may play an important role [250]. Considering that the infiltration of immune cells is related to the therapeutic effect of tumors, exosome-mediated metabolic reprogramming may be an entry point for improving immunotherapy. (Fig. 5).

Fig. 5
figure 5

The mechanism of tumor-serected exosomal cargos regulate metabolic reprogramming of immune cells. Tumor cells-secreted exosomal miR-451 can target AMPK to activate the mTOR pathway and promote Th17 polarized differentiation of T cells through glucose deprivation. Cervical squamous carcinoma-derived exosomal miR-142-5p can target ARID2 to inhibit DNMT1 recruitment to the INF-γ promoter, leading to up-regulation of IDO, thereby suppressing and exhausting CD8 + T cells. Prostate cancer-derived exosomal Rab27a can promote CD73 expression in DCs, which hydrolyzes AMP to adenosine and inhibits the production of TNFα and IL-1L in an ATP-dependent manner, resulting in functional inhibition of DCs. Hypoxia-induced exosomal let-7a can inhibit the insulin-AKT-mTOR pathway and induce M2-polarized TAMs by enhancing OXPHOS. Lung cancer cell-secreted exosomal PKM2 can activate the AMPK pathway to induce M2-polarized TAMs, in which exosome-mediated glycolytic remodeling may play a role. HNSCC cells-derived exosomal metabolite adenosine can induce A2BR-mediated M2-polarized TAMs

Exosomal cargos induce metabolic reprogramming of adipocytes

Adipocytes in the TME can participate in tumor progression through their secretory functions, and exosomal cargos-mediated metabolic reprogramming of adipocytes can induce adipocytes to assume the CAAs phenotype. It has been shown that pancreatic cancer-derived exosomes can modulate lipid metabolism in adipocytes, in which TG is significantly down-regulated. This may be related to the increased expression of IL-6 and the promotion of lipolysis [251]. Breast cancer secreted exosomal miRNAs induce adipocyte differentiation by regulating metabolism. Exosomal miR-155 targets PPARγ and increases catabolism characterized by the release of metabolites, promoting beige/brown differentiation of adipocytes [252, 253]. Exosomal miR-126 inhibits lipid droplet accumulation and glucose uptake in adipocytes by disrupting IRS/Glut-4 signaling, and exosomal miR-144 promotes beige/brown differentiation by down-regulating the MAP3K8/ERK1/2/PPARγ axis [254]. In addition, lipolysis-inducing factors may be present in exosomes and alter the metabolism of adipocytes. Adrenomedullin (AM), which is abundant in pancreatic cancer exosomes, can promote intracellular lipolysis through the p38/ERK1/2 signaling axis and promote increased free fatty acid content in conditioned media [255].

Exosomal cargos induce metabolic reprogramming of stellate cells

Activation of stellate cells is commonly seen in liver metastasis of colorectal cancer (CRLM) and pancreatic ductal adenocarcinoma (PDAC). As part of the stroma in TME, activated stellate cells tend to have the characteristics of CAFs that facilitate the composition of the pre-metastatic niche [280]. A number of studies have shown that exosomal cargos can induce the activation of HSCs and PSCs, presenting a profibrogenic phenotype [99, 101]. However, the mechanisms of TME stellate cell activation remains unclear, and exosomal cargos-mediated metabolic reprogramming may explain this phenomenon. IL-17B secreted by pancreatic cancer can be delivered to stromal PSCs by EVs and induce the expression of IL-17RB. Up-regulation of IL-17RB in PSCs enhanced OXPHOS while reducing mitochondrial turnover to activate PSCs [256]. CRC-derived exosomal HSPC111 activates HSCs by altering the acetyl-CoA levels, citrate content and phosphorylation of ATP-citrate lyase (ACLY), causing them to exhibit a profibrogenic phenotype similar to CAFs [241]. (Fig. 6).

Fig. 6
figure 6

The mechanism of tumor-serected exosomal cargos regulate metabolic reprogramming of stromal cells. In addition to CAFs, adipocytes and stellate cells are the stromal cell components in TME. Pancreatic cancer-derived exosomes can reduce TG production by promoting IL-6 expression and lipolysis. AM in pancreatic cancer exosomes can activate the p38/ERK1/2 signaling axis, promote intracellular lipolysis, and increase extracellular free fatty acids. Breast cancer-secreted exosomal miR-155 can target PPARγ to promote CAAs beige/brown differentiation. Breast cancer-secreted exosomal miR-126 can inhibit the IRS/Glut-4 axis to reduce CAAs lipid droplet accumulation and glucose uptake, and exosomal miR-144 can inhibit the MAP3K8/ERK1/2/PPARγ axis to promote CAAs beige/brown differentiation. Pancreatic cancer-secreted exosomal IL-17B can activate PSCs by promotingIL-17RB expression and enhancing OXPHOS. CRC-secreted exosomal HSPC111 can activate HSCs by regulating acetyl-CoA expression, ACLY phosphorylation, and increase citrate content

Exosomal cargos remodel ECM through metabolic reprogramming

It is well known that CAFs and CAAs are the main cells that produce ECM. Exosome-mediated metabolic reprogramming can activate CAFs and CAAs in TME, which inevitably leads to an impact on ECM. Colorectal cancer (CRC) cell-derived exosomes enhance the secretion of ECM (COL1A1, Tenascin-C/X) by CAFs based on the regulation of metabolic reprogramming [240]. In addition to increased secretion, imbalance in physicochemical properties is an important manifestation of ECM heterogeneity. A study has found that human melanoma-derived exosomes are rich in exosomal miR-155 and miR-210. These exosomal miRNAs reprogram human adult dermal fibroblasts by promoting glycolysis and inhibiting oxidative phosphorylation (OXPHOS), leading to the acidification of ECM [257, 281]. The local acidification of ECM is conducive to the formation of tumor pre-metastasis niches.

TME-derived exosomal cargos-mediated metabolic reprogramming

Exosomal cargos-mediated metabolic reprogramming is bidirectional between tumor cells and TME. Tex induces non-tumor cells in TME to exhibit a malignant phenotype, and these cells can secrete exosomal cargos to regulate metabolic reprogramming of tumor cells and accelerate tumor progression. In this way, a malignant positive feedback regulation pattern is formed between TME and tumor cells.

CAFs-derived exosomal cargos induce metabolic reprogramming

Fibroblasts in the TME are activated by stimulatory signals into CAFs and remodel the TME through their secretory functions (such as CCL2, VEGF, and IL-6) [282]. The exosomal cargos secreted by CAFs can act on tumor cells and induce metabolic reprogramming, which is beneficial to tumor progression. These exosomal cargos are mainly composed of nucleic acids. (Fig. 7).

Fig. 7
figure 7

The mechanism of CAFs-derived exosomal cargos regulate metabolic reprogramming of tumor cells. CAFs regulate metabolic reprogramming of tumor cells mainly by exosomal nucleic acids. Exosomal miR-522 activates the USP7/hnRNPA1 pathway by targeting ALOX15, blocking the accumulation of lipid-ROS. Exosomal SNHG3 inhibits OXPHOS and promotes glycolysis through the miR-330-5p/PKM axis. Exosomal TUG1 promotes glycolysis through the miR-524-5p/SIX1 axis. Exosomal LINC01614 enhances tumor cells glutamine uptake by up-regulating glutamine transporters SLC38A2 and SLC7A5. In addition, exosomal mtDNA can lead to endocrine therapy resistance in OXPHOS-dependent breast cancer

Exosomal miR-522 secreted by CAFs targets ALOX15 to activate the USP7/hnRNPA1 pathway, blocking lipid-ROS accumulation and inhibiting the ferroptosis of gastric cancer cells [258, 283]. Exosomal lncRNA SNHG3 secreted by CAFs serves as a molecular sponge for miR-330-5p in breast cancer cells. MiR-330-5p further targets PKM to inhibit OXPHOS and promote glycolysis, favoring breast cancer cells proliferation [259]. Exosomal lncRNA TUG1 secreted by CAFs promotes glycolysis in HCC cells via the miR-524-5p/SIX1 axis [260]. CAFs can regulate amino acid metabolism in lung adenocarcinoma (LUAD) cells via exosomes. CAFs-derived exosomal LINC01614 can interact with ANXA2, p65 and activate NF-κB pathway to promote the expression of SLC38A2 and SLC7A5, thus enhancing tumor cells glutamine uptake [261]. Furthermore, CAFs-derived exosomal mtDNA may regulate hormonal therapy-resistant (HTR) in breast cancer patients. A study identified mitochondrial genomes in exosomes isolated from plasma from HTR breast cancer patients. CAFs-derived exosomal mtDNA can lead to fulvestrant resistance in OXPHOS-dependent breast cancer cells [262].

TAMs-derived exosomal cargos induce metabolic reprogramming

Macrophages are the most numerous white blood cells in TME, and TAMs play a key regulatory role in the occurrence and development of tumors. Exosomal ncRNAs secreted by TAMs can affect the metabolic state of tumor cells [284]. In recent studies, macrophages-derived exosomal miR-503-3p targets DACT2, activates the Wnt/β-catenin signaling pathway, promotes glycolysis and reduces mitochondrial OXPHOS in BC cells [263]. M2 macrophages-derived exosomal miR-222-3p targets PDLIM2 to reduce ubiquitination of PFKL. The stabilization of PFKL promotes glycolysis in laryngeal cancer cells [264]. TAMs-derived exosomal lncRNA HISLA inhibits the degradation of HIF-1α by inhibiting the binding of HIF-1α, a key transcription factor of aerobic glycolysis with its hydroxylase PHD2. Conversely, the lactic acid secreted by tumor cells in aerobic glycolysis state can promote the sorting and loading of exosomes HISLA in macrophages [265]. This positive feedback metabolic regulation enhances the apoptosis resistance of BC cells. (Fig. 8).

Fig. 8
figure 8

The mechanism of TAMs-derived exosomal cargos regulate metabolic reprogramming of tumor cells. TAMs-derived exosomal cargos mainly affect tumor cells glycolysis and OXPHOS. Exosomal HISLA promotes glycolysis by inhibiting the binding of HIF-1α to its hydroxylase PHD2. Exosomal miR-222-3p promotes glycolysis by targeting PDLIM2 to inhibit the ubiquitinated degradation of PFKL. Exosomal miR-503-3p activates the Wnt/β-catenin signaling pathway by targeting DACT2 to promote glycolysis and inhibit OXPHOS

Adipocytes-derived exosomal cargos induce metabolic reprogramming

Previous studies on exosomes and stromal cells in TME have mostly focused on CAF, but little attention has been paid to CAA, especially the exosomes secreted by CAAs and their functions. In fact, adipocytes, as a kind of secretory cells, can not only secrete metabolites such as leptin and fatty acids directly, but also secrete bioactive cargos in the form of exosomes, which can modify the metabolism of tumor cells. Advanced melanoma cells are in direct contact with subcutaneous adipocytes. Adipocytes can directly transfer fatty acids into melanocytes, resulting in increased lipid content and abnormal lipid droplet formation. [285]. In addition to direct contact, exosomes act as a bridge between melanoma and adipocytes. Mass spectrometry showed that adipocytes-derived exosomes were rich in fatty acid oxidation (FAO) related proteins, which promoted FAO and migration of melanoma cells. [266]. This suggests that adipocytes-derived exosomes may induce metabolic reprogramming of tumor cells, the mechanism remains to be elucidated [286]. Subsequently, it was found that the adipocyte-secreted exosomal miR-23a/b regulates GLUT-1 expression by targeting the VHL/HIF-1α axis, leading to 5-FU resistance in HCC cells [267]. A recent study found that the adipocyte-derived exosomal miR-433-3p in a hypoxic environment can target SCD1 (a key regulatory gene for the synthesis of monounsaturated fatty acids) to promote lipid accumulation in NPC cells and facilitate proliferation and migration [268]. (Fig. 9).

Fig. 9
figure 9

The mechanism of adipocytes-derived exosomal cargos regulate metabolic reprogramming of tumor cells. The effect of adipocytes-derived exosomal cargos on tumor cells metabolism is reflected in lipid. Some FAO-associated proteins are enriched in adipocytes-derived exosomes and promote FAO in tumor cells via exosomes. Exosomal miR-433-3p targets SCD1 to promote lipid accumulation in tumor cells. Exosomal miR-23a/b targets VHL to promote HIF-1α expression, which enhances GLUT-1 expression and 5-FU resistance in tumor cells

Exosome-based tumor metabolic therapies

As oncology researches have become more detailed, the understanding of tumors has transitioned from genetic diseases to chronic metabolic diseases. The aim of anti-tumor therapy has also gradually evolved from targeting molecular biomarkers to targeting tumor metabolic pathways. The development of single-cell sequencing technology has led to a more comprehensive understanding of TME and the discovery that metabolic reprogramming exists not only in tumor cells, but also in other components of TME (stromal cells, immune cells and endothelial cells) [287, 288]. These non-tumor cells have adapted to TME by altering metabolic pathways, creating a pre-metastatic niche that favors tumor progression [289]. It is hopeful to target specific exosomal cargos-mediated metabolic pathways and develop exosome-based vehicles for anti-tumor therapy. (Table 4).

Table 4 Therapeutics targeting exosome-mediated metabolism

The most concise way to target exosomal cargos-mediated reprogramming of TME metabolism for tumor treatment is to inhibit exosomal secretion. PKM2 is a key enzyme in glycolysis and is involved in the secretion of exosomes in addition to directly reshaping cellular metabolism through the OXPHOS and Warburg effects [297]. During the release of exosomes, phosphorylated PKM2 acts as a protein kinase to promote the formation of SNARE complexes by enhancing the phosphorylation of SNAP23 [298]. During the release of exosomes, phosphorylated PKM2 acts as a protein kinase to promote the formation of SNARE complexes by enhancing the phosphorylation of SNAP23. PKM2 combines metabolic regulation with non-metabolic regulation of exosome secretion, is an ideal target for exosome and tumor metabolic therapy. Shikonin is the active ingredient of Comfrey, a naphthoquinone compound [290]. Shikonin is a specific PKM2 inhibitor that not only inhibits glucose uptake and lactate production in tumor cells, but also inhibits glycolysis by reducing extracellular secretion of exosomal PKM2 and enhances cisplatin sensitivity in NSCLC cells [290]. A study on bladder cancer found that highly-expression of PKM2 was associated with cisplatin resistance. Shikonin can promote cisplatin sensitivity of bladder cancer cells by reducing the release of exosomes, but the specific mechanism remains to be explored [291].

Exosomal ncRNAs, as common cargos in TME, play an important role in the metabolic reprogramming of TME. It was found that oxaliplatin-resistant CRC cells-derived exosomal circRNA ciRS-122 was delivered to sensitive cells, which enhanced glycolysis and chemoresistance in sensitive cells via miR-122/PKM2 signaling axis [292]. Development of exosome-transported si-ciRS-122 can reverse the ciRS-122/miR-122/PKM2 signaling axis to inhibit glycolysis and enhance chemosensitivity in CRC cells. In addition to tumor cells, targeting exosomal circRNAs derived from CAFs in TME has antitumor effects. It was found that exosomal circCCT3 secreted by CAFs could enhance glucose metabolism by regulating the expression of HK2. It was found that exosomal circCCT3 secreted by CAFs could enhance glucose metabolism by regulating the expression of HK2. Treatment of CAFs with coptisine inhibited the secretion of exosomal circCCT3 and suppressed HCC cell proliferation and invasion [293]. In addition, docosahexaenoic acid (DHA) as an omega 3 free fatty acid has been reported to exert anti-angiogenesis effects. DHA can alter the expression of angiogenesis-related exosomal miRNAs in breast cancer cells, inhibits angiogenesis by up-regulating exosomal miR-101, miR-199, and miR-342, and down-regulating exosomal mir-382 and miR-21 to exert anti-tumor effects [294, 299].

Benefiting from the targeting and biocompatibility of exosomes, exploitation of exosomes as carriers for drug delivery targeting tumor metabolism has a bright future. Although there are currently no engineered exosomes to directly target various metabolic pathways in the TME, exosomes can be combined with classical drugs or modalities as an adjuvant therapy to improve the efficacy of anti-tumor therapy. It has been shown that combination of exosome-mediated cPLA2 siRNA and metformin reduces the growth of glioblastoma xenografts by impairing the energy metabolism of mitochondria [295]. Photodynamic therapy (PDT) is a novel method of treating tumors with photosensitizing drugs and laser activation [300]. Aggregation-induced emission luminogens (AIEgens) are photosensitizers for PDT whose efficacy is limited by GSH. A recent work developed TEX for the co-delivery of AIEgens and proton pump inhibitor (PPI) for tumor combination therapy. TEX can specifically deliver AIEgens and PPI to tumor sites, and PPI inhibits GSH and ATP produced by glutamine metabolism in tumor cells, which contribute to the efficacy of AIEgens [296]. The combination of exosomes, glutamine metabolism and PDT may be a new option for future tumor treatment, but treatments that inhibit glutamine metabolism still need to be approached with caution. Glutamine depletion may stimulate release from Rab11a compartments of exosomes with pro-tumorigenic functions [301]. Therefore, exosome-based tumor metabolic therapy still needs further refinement to find the balance between pro-tumorigenesis and anti-tumorigenesis.

Conclusion

This review highlights the multiple roles and molecular mechanisms of exosome-mediated metabolic reprogramming in TME reprogramming. The field of exosomes (or EVs) has made great progress in recent years benefiting from technological breakthroughs in isolation, purification, in vivo tracking and content analysis [2]. This has led to the identification of other types of EVs besides exosomes and their functions becoming a novel hotspot in the field of EVs. In the future, the understanding of exosomes will be enriched by how to precisely distinguish exosomes from other EVs subtypes and exclude contaminants to further obtain high purity exosomes. This will also help to improve the targeting and biosafety of antitumor therapies developed with exosomes as vectors.

We describe the cell–cell communication mediated by exosomal cargos in TME and how these cargos are sorted to exosomes. Along with technological advances, the way of sorting various types of cargos into exosomes (or specific subtypes of EVs) is the bottleneck for further development in the field of exosomes. The bioactive cargos are the key to the function of exosomes. In addition to the cell-derived exosomal cargos in human TME, milk exosomes have a bright future as an oral drug delivery system, due to the biocompatibility of milk exosomes with exogenous cargos [302].

The heterogeneity of TME promotes tumor proliferation, metastasis, stemness and drug resistance. We summarized the main components and characteristics of TME, and highlighted the role and mechanism of exosomal cargos-mediated metabolic reprogramming in the heterogeneity of TME. Improving TME becomes an emerging strategy for anti-tumor treatment. The plasticity of tumor metabolism is both promising and challenging. Given the complex composition of TME, targeting one component for metabolic remodeling is difficult, and we need to consider more whether altered metabolism has the same therapeutic effects on multiple components of TME. Application of tumor organoid platforms to exosomes may be used to simulate the effect of exosomes on TME.

Exosomal cargos-mediated abnormalities metabolism in TME remains to be extensively studied. Considering the widespread of exosomal cargos, exploring the molecular mechanisms of exosomal cargos-induced metabolic reprogramming is beneficial for tumor precision treatment. As more and more biologic companies are entering the exosome field, the development of exosome-based drug delivery modalities to reshape metabolism in TME is promising. Combining chemotherapy, radiotherapy or targeted therapy with novel metabolic therapies may be the future trend in tumor treatment.

Availability of data and materials

Not applicable.

Abbreviations

TME:

Tumor microenvironment

EVs:

Extracellular vesicles

MVs:

Microvesicles

ncRNA:

Non-coding RNA

CAFs:

Cancer-associated fibroblasts

ECM:

Extracellular matrix

NFs:

Normal fibroblasts

HGSOC:

High-grade serous ovarian cancer

TAMs:

Tumor-associated macrophages

MDMs:

Monocyte-derived macrophages

RCC:

Renal cell carcinoma

APCs:

Antigen-presenting cells

TCR:

T cell receptor

CTLs:

Cytotoxic T lymphocytes

ECs:

Endothelial cells

ISF:

Interstitial fluid

LECs:

Lymphatic endothelial cells

WAT:

White adipose tissue

CAAs:

Cancer-associated adipocytes

HSCs:

Hepatic stellate cells

PSCs:

Pancreatic stellate cells

LOXs:

Lysyl oxidases

LHs:

Lysyl hydroxylases

HIFs:

Hypoxia-inducible factors

ILV:

Intraluminal vescicles

MVE:

Multivesicular endosome

sEVs:

Small extracellular vesicles

ESCRT:

Endosomal sorting complex

MVB:

Multivesicular body

MMPs:

Matrix metalloproteinases

DGUC:

Density gradient differential ultracentrifugatio

SEC:

Size exclusion chromatography

UF:

Ultrafiltration

AIEX:

Anion exchange chromatography

gDNA:

Genomic DNA

ccRCC:

Clear cell renal cell carcinoma

OPRs:

Oxysterol-binding protein-related proteins

nSMase2:

Neural sphingomyelinase 2

hnRNP:

Heterogeneous nuclear ribonucleoprotein

Ago2:

Argonaute 2

MN:

Micronuclei

ESE:

Early-sorting endosome

PTM:

Post-translational modification

G-MDSCs:

Granulocyte-myeloid derived suppressor cells

TG:

Triglycerides

PG:

Glycerophosphoglycerol

PS:

Glycerophosphatidylserine

PC:

Glycerophosphorylcholine

PPP:

Pentose phosphate pathway

SSP:

Serine synthesis pathway

TCA:

Tricarboxylic acid

PKM2:

Pyruvate kinase 2

FA:

Fatty acid

FATP:

Fatty acid transporter protein

FABPpm:

Plasma membrane fatty acid binding protein

ACLY:

ATP-citrate lyase

FASN:

Fatty acid synthase

EMT:

Epithelial-mesenchymal transition

Gln:

Glutamine

Ser:

Serine

Trp:

Tryptophan

Kyn:

Kynurenine

5-HT:

5-Hydroxytryptamine

MDSCs:

Myeloid-derived suppressor cells

TDO:

Tryptophan 2, 3-dioxygenase

BC:

Breast cancer

HCC:

Hepatocellular carcinoma

TNBC:

Triple-negative breast cancer

CRC:

Colorectal cancer

AML:

Acute myeloid leukemia

EPC:

Endothelial progenitor cell

HUVECs:

Human umbilical vein endothelial cells

HNSCC:

Head and neck squamous cell carcinoma

TEXs:

Tumor cell-derived exosomes

ADO:

Adenosine

DCs:

Dendritic cells

AM:

Adrenomedullin

CRLM:

Metastasis of colorectal cancer

PDAC:

Pancreatic ductal adenocarcinoma

OXPHOS:

Oxidative phosphorylation

LUAD:

Lung adenocarcinoma

HTR:

Hormonal therapy-resistant

FAO:

Fatty acid oxidation

NSCLC:

Non-small cell lung cancer

DHA:

Docosahexaenoic acid

PDT:

Photodynamic therapy

PPI:

Proton pump inhibitor

References

  1. Qian F, Huang Z, Zhong H, Lei Q, Ai Y, Xie Z, et al. Analysis and biomedical applications of functional cargo in extracellular vesicles. ACS Nano. 2022;16(12):19980–20001.

    Article  CAS  PubMed  Google Scholar 

  2. Lucotti S, Kenific CM, Zhang H, Lyden D. Extracellular vesicles and particles impact the systemic landscape of cancer. EMBO J. 2022;41(18):e109288.

    Article  CAS  PubMed  Google Scholar 

  3. Thery C, Witwer KW, Aikawa E, Alcaraz MJ, Anderson JD, Andriantsitohaina R, 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(1):1535750.

    Article  PubMed  PubMed Central  Google Scholar 

  4. Xie F, Zhou X, Fang M, Li H, Su P, Tu Y, et al. Extracellular vesicles in cancer immune microenvironment and cancer immunotherapy. Adv Sci (Weinh). 2019;6(24):1901779.

    Article  CAS  PubMed  Google Scholar 

  5. Zhou M, Li YJ, Tang YC, Hao XY, Xu WJ, Xiang DX, et al. Apoptotic bodies for advanced drug delivery and therapy. J Control Release. 2022;351:394–406.

    Article  CAS  PubMed  Google Scholar 

  6. Bian X, Xiao YT, Wu T, Yao M, Du L, Ren S, et al. Microvesicles and chemokines in tumor microenvironment: mediators of intercellular communications in tumor progression. Mol Cancer. 2019;18(1):50.

    Article  PubMed  PubMed Central  Google Scholar 

  7. Zhang H, Freitas D, Kim HS, Fabijanic K, Li Z, Chen H, et al. Identification of distinct nanoparticles and subsets of extracellular vesicles by asymmetric flow field-flow fractionation. Nat Cell Biol. 2018;20(3):332–43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Zhang Q, Jeppesen DK, Higginbotham JN, Graves-Deal R, Trinh VQ, Ramirez MA, et al. Supermeres are functional extracellular nanoparticles replete with disease biomarkers and therapeutic targets. Nat Cell Biol. 2021;23(12):1240–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Ma L, Li Y, Peng J, Wu D, Zhao X, Cui Y, et al. Discovery of the migrasome, an organelle mediating release of cytoplasmic contents during cell migration. Cell Res. 2015;25(1):24–38.

    Article  CAS  PubMed  Google Scholar 

  10. Zhao X, Lei Y, Zheng J, Peng J, Li Y, Yu L, et al. Identification of markers for migrasome detection. Cell Discov. 2019;5:27.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Jiao H, Jiang D, Hu X, Du W, Ji L, Yang Y, et al. Mitocytosis, a migrasome-mediated mitochondrial quality-control process. Cell. 2021;184(11):2896-910 e13.

    Article  CAS  PubMed  Google Scholar 

  12. Kalluri R, LeBleu V. The biology function and biomedical applications of exosomes. Science (New York, NY). 2020;367(6478):eaau6977.

    Article  CAS  Google Scholar 

  13. Shao H, Im H, Castro CM, Breakefield X, Weissleder R, Lee H. New Technologies for Analysis of Extracellular Vesicles. Chem Rev. 2018;118(4):1917–50.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Srivastava A, Rathore S, Munshi A, Ramesh R. Organically derived exosomes as carriers of anticancer drugs and imaging agents for cancer treatment. Semin Cancer Biol. 2022;86:80.

    Article  CAS  PubMed  Google Scholar 

  15. Ma YS, Yang XL, Xin R, Liu JB, Fu D. Power and promise of exosomes as clinical biomarkers and therapeutic vectors for liquid biopsy and cancer control. Biochim Biophys Acta Rev Cancer. 2021;1875(1):188497.

    Article  CAS  PubMed  Google Scholar 

  16. Cappello F, Fais S. Extracellular vesicles in cancer pros and cons: The importance of the evidence-based medicine. Semin Cancer Biol. 2022;86:4.

    Article  CAS  PubMed  Google Scholar 

  17. Yang K, Zhou Q, Qiao B, Shao B, Hu S, Wang G, et al. Exosome-derived noncoding RNAs: Function, mechanism, and application in tumor angiogenesis. Mol Ther Nucleic Acids. 2022;27:983–97.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Banik A, Sharma R, Chauhan A, Singh S. Cutting the umbilical cord: Cancer stem cell-targeted therapeutics. Life Sci. 2022;299:120502.

    Article  CAS  PubMed  Google Scholar 

  19. Wu Y, Niu D, Deng S, Lei X, Xie Z, Yang X. Tumor-derived or non-tumor-derived exosomal noncodingRNAs and signaling pathways in tumor microenvironment. Int Immunopharmacol. 2022;106:108626.

    Article  CAS  PubMed  Google Scholar 

  20. Lampropoulou D, Pliakou E, Aravantinos G, Filippou D, Gazouli M. The Role of Exosomal Non-Coding RNAs in Colorectal Cancer Drug Resistance. Int J Mol Sci. 2022;23(3):1473.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Arora S, Khan S, Zaki A, Tabassum G, Mohsin M, Bhutto H, et al. Integration of chemokine signaling with non-coding RNAs in tumor microenvironment and heterogeneity in different cancers. Semin Cancer Biol. 2022;86:720.

    Article  CAS  PubMed  Google Scholar 

  22. Aghanejad A, Bonab S, Sepehri M, Haghighi F, Tarighatnia A, Kreiter C, et al. A review on targeting tumor microenvironment: The main paradigm shift in the mAb-based immunotherapy of solid tumors. Int J Biol Macromol. 2022;207:592–610.

    Article  CAS  PubMed  Google Scholar 

  23. Petroni G, Buqué A, Coussens L, Galluzzi L. Targeting oncogene and non-oncogene addiction to inflame the tumour microenvironment. Nat Rev Drug Discovery. 2022;21(6):440–62.

    Article  CAS  PubMed  Google Scholar 

  24. Arora S, Khan S, Zaki A, Tabassum G, Mohsin M, Bhutto HN, et al. Integration of chemokine signaling with non-coding RNAs in tumor microenvironment and heterogeneity in different cancers. Semin Cancer Biol. 2022;86(Pt 2):720–36.

    Article  CAS  PubMed  Google Scholar 

  25. Maman S, Witz IP. A history of exploring cancer in context. Nat Rev Cancer. 2018;18(6):359–76.

    Article  CAS  PubMed  Google Scholar 

  26. Xia L, Oyang L, Lin J, Tan S, Han Y, Wu N, et al. The cancer metabolic reprogramming and immune response. Mol Cancer. 2021;20(1):28.

    Article  PubMed  PubMed Central  Google Scholar 

  27. Ping Q, Yan R, Cheng X, Wang W, Zhong Y, Hou Z, et al. Cancer-associated fibroblasts: overview, progress, challenges, and directions. Cancer Gene Ther. 2021;28(9):984–99.

    Article  CAS  PubMed  Google Scholar 

  28. Czekay R, Cheon D, Samarakoon R, Kutz S, Higgins P. Cancer-associated fibroblasts: mechanisms of tumor progression and novel therapeutic targets. Cancers. 2022;14(5):1231.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Poon S, Ailles L. Modeling the role of cancer-associated fibroblasts in tumor cell invasion. Cancers. 2022;14(4):962.

    Article  PubMed  PubMed Central  Google Scholar 

  30. Jia W, Liang S, Cheng B, Ling C. The role of cancer-associated fibroblasts in hepatocellular carcinoma and the value of traditional Chinese medicine treatment. Front Oncol. 2021;11:763519.

    Article  PubMed  PubMed Central  Google Scholar 

  31. Bejarano L, Jordāo M, Joyce J. Therapeutic targeting of the tumor microenvironment. Cancer Discov. 2021;11(4):933–59.

    Article  CAS  PubMed  Google Scholar 

  32. Costa A, Kieffer Y, Scholer-Dahirel A, Pelon F, Bourachot B, Cardon M, et al. Fibroblast heterogeneity and immunosuppressive environment in human breast cancer. Cancer Cell. 2018;33(3):463-79.e10.

    Article  CAS  PubMed  Google Scholar 

  33. Givel A, Kieffer Y, Scholer-Dahirel A, Sirven P, Cardon M, Pelon F, et al. miR200-regulated CXCL12β promotes fibroblast heterogeneity and immunosuppression in ovarian cancers. Nat Commun. 2018;9(1):1056.

    Article  PubMed  PubMed Central  Google Scholar 

  34. Pelon F, Bourachot B, Kieffer Y, Magagna I, Mermet-Meillon F, Bonnet I, et al. Cancer-associated fibroblast heterogeneity in axillary lymph nodes drives metastases in breast cancer through complementary mechanisms. Nat Commun. 2020;11(1):404.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Han C, Zhang C, Wang H, Zhao L. Exosome-mediated communication between tumor cells and tumor-associated macrophages: implications for tumor microenvironment. Oncoimmunology. 2021;10(1):1887552.

    Article  PubMed  PubMed Central  Google Scholar 

  36. DeNardo D, Ruffell B. Macrophages as regulators of tumour immunity and immunotherapy. Nat Rev Immunol. 2019;19(6):369–82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Inagaki K, Kunisho S, Takigawa H, Yuge R, Oka S, Tanaka S, et al. Role of tumor-associated macrophages at the invasive front in human colorectal cancer progression. Cancer Sci. 2021;112(7):2692–704.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Tan S, Xia L, Yi P, Han Y, Tang L, Pan Q, et al. Exosomal miRNAs in tumor microenvironment. J Exp Clin Cancer Res: CR. 2020;39(1):67.

    Article  PubMed  PubMed Central  Google Scholar 

  39. Sica A, Mantovani A. Macrophage plasticity and polarization: in vivo veritas. J Clin Investig. 2012;122(3):787–95.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Nirschl T, El Asmar M, Ludwig W, Ganguly S, Gorin M, Johnson M, et al. Transcriptional profiling of tumor associated macrophages in human renal cell carcinoma reveals significant heterogeneity and opportunity for immunomodulation. Am J Clin Exp Urol. 2020;8(1):48–58.

    PubMed  PubMed Central  Google Scholar 

  41. Marku M, Verstraete N, Raynal F, Madrid-Mencía M, Domagala M, Fournié J, et al. Insights on TAM formation from a Boolean model of macrophage polarization based on in vitro studies. Cancers. 2020;12(12):3664.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Dang M, Gonzalez M, Gaonkar K, Rathi K, Young P, Arif S, et al. Macrophages in SHH subgroup medulloblastoma display dynamic heterogeneity that varies with treatment modality. Cell Rep. 2021;34(13):108917.

    Article  CAS  PubMed  Google Scholar 

  43. Xie F, Zhou X, Fang M, Li H, Su P, Tu Y, et al. Extracellular vesicles in cancer immune microenvironment and cancer immunotherapy. Adv Sci (Weinheim, Baden-Wurttemberg, Germany). 2019;6(24):1901779.

    CAS  Google Scholar 

  44. Saillard M, Cenerenti M, Romero P, Jandus C. Impact of immunotherapy on CD4 T cell phenotypes and function in cancer. Vaccines. 2021;9(5):454.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Miggelbrink A, Jackson J, Lorrey S, Srinivasan E, Waibl-Polania J, Wilkinson D, et al. CD4 T-cell exhaustion: does it exist and what are its roles in cancer? Clin Cancer Res J Am Asso Cancer Res. 2021;27(21):5742–52.

    Article  CAS  Google Scholar 

  46. Yang P, Peng Y, Feng Y, Xu Z, Feng P, Cao J, et al. Immune cell-derived extracellular vesicles - new strategies in cancer immunotherapy. Front Immunol. 2021;12:771551.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Nanke Y, Kobashigawa T, Yago T, Kawamoto M, Yamanaka H, Kotake S. Detection of IFN-γ+IL-17+ cells in salivary glands of patients with Sjögren’s syndrome and Mikulicz’s disease: Potential role of Th17 Th1 in the pathogenesis of autoimmune diseases Nihon Rinsho Men’eki Gakkai kaishi. Jap J Clin Immunol. 2016;39(5):473–7.

    Article  CAS  Google Scholar 

  48. Farhood B, Najafi M, Mortezaee K. CD8 cytotoxic T lymphocytes in cancer immunotherapy: A review. J Cell Physiol. 2019;234(6):8509–21.

    Article  CAS  PubMed  Google Scholar 

  49. Maimela N, Liu S, Zhang Y. Fates of CD8+ T cells in Tumor Microenvironment. Comput Struct Biotechnol J. 2019;17:1–13.

    Article  CAS  PubMed  Google Scholar 

  50. Henning A, Roychoudhuri R, Restifo N. Epigenetic control of CD8 T cell differentiation. Nat Rev Immunol. 2018;18(5):340–56.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Dong D, Fan P, Feng Y, Liu G, Peng Y, Dong T, et al. Association between circulating CD39+CD8+ T cells pre-chemoradiotherapy and prognosis in patients with nasopharyngeal carcinoma. Chin Med J. 2021;134(17):2066–72.

    Article  PubMed  PubMed Central  Google Scholar 

  52. Jørgensen N, Hviid T, Nielsen L, Sønderstrup I, Eriksen J, Ejlertsen B, et al. Tumour-infiltrating CD4-, CD8- and FOXP3-positive immune cells as predictive markers of mortality in BRCA1- and BRCA2-associated breast cancer. Br J Cancer. 2021;125(10):1388–98.

    Article  PubMed  PubMed Central  Google Scholar 

  53. Giatromanolaki A, Anestopoulos I, Panayiotidis M, Mitrakas A, Pappa A, Koukourakis M. Prognostic relevance of the relative presence of CD4, CD8 and CD20 expressing tumor infiltrating lymphocytes in operable non-small cell lung cancer patients. Anticancer Res. 2021;41(8):3989–95.

    Article  CAS  PubMed  Google Scholar 

  54. Toor S, Sasidharan Nair V, Saleh R, Taha R, Murshed K, Al-Dhaheri M, et al. Transcriptome of tumor-infiltrating T cells in colorectal cancer patients uncovered a unique gene signature in CD4 T cells associated with poor disease-specific survival. Vaccines. 2021;9(4):334.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Borsetto D, Tomasoni M, Payne K, Polesel J, Deganello A, Bossi P, et al. Prognostic significance of CD4+ and CD8+ tumor-infiltrating lymphocytes in head and neck squamous cell carcinoma: a meta-analysis. Cancers. 2021;13(4):781.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Gu Y, Chen Y, Jin K, Cao Y, Liu X, Lv K, et al. Intratumoral CD103CD4 T cell infiltration defines immunoevasive contexture and poor clinical outcomes in gastric cancer patients. Oncoimmunology. 2020;9(1):1844402.

    Article  PubMed  PubMed Central  Google Scholar 

  57. Wang W, Green M, Choi J, Gijón M, Kennedy P, Johnson J, et al. CD8 T cells regulate tumour ferroptosis during cancer immunotherapy. Nature. 2019;569(7755):270–4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Ma X, Bi E, Lu Y, Su P, Huang C, Liu L, et al. Cholesterol induces CD8 T cell exhaustion in the tumor microenvironment. Cell Metab. 2019;30(1):143-56.e5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Knox MC, Ni J, Bece A, Bucci J, Chin Y, Graham PH, et al. A clinician’s guide to cancer-derived exosomes: immune interactions and therapeutic implications. Front Immunol. 2020;11:1612.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Taskaeva I, Bgatova N. Microvasculature in hepatocellular carcinoma: an ultrastructural study. Microvasc Res. 2021;133:104094.

    Article  CAS  PubMed  Google Scholar 

  61. Schito L, Rey S. Hypoxia: turning vessels into vassals of cancer immunotolerance. Cancer Lett. 2020;487:74–84.

    Article  CAS  PubMed  Google Scholar 

  62. Schaaf M, Garg A, Agostinis P. Defining the role of the tumor vasculature in antitumor immunity and immunotherapy. Cell Death Dis. 2018;9(2):115.

    Article  PubMed  PubMed Central  Google Scholar 

  63. Peri S, Biagioni A, Versienti G, Andreucci E, Staderini F, Barbato G, et al. Enhanced vasculogenic capacity induced by 5-fluorouracil chemoresistance in a gastric cancer cell line. Int J Mol Sci. 2021;22(14):7698.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Morales-Guadarrama G, García-Becerra R, Méndez-Pérez E, García-Quiroz J, Avila E, Díaz L. Vasculogenic mimicry in breast cancer: clinical relevance and drivers. Cells. 2021;10(7):1758.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Jia T, Jacquet T, Dalonneau F, Coudert P, Vaganay E, Exbrayat-Héritier C, et al. FGF-2 promotes angiogenesis through a SRSF1/SRSF3/SRPK1-dependent axis that controls VEGFR1 splicing in endothelial cells. BMC Biol. 2021;19(1):173.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Haider T, Sandha K, Soni V, Gupta P. Recent advances in tumor microenvironment associated therapeutic strategies and evaluation models. Mater Sci Eng, C Mater Biol Appl. 2020;116:111229.

    Article  CAS  PubMed  Google Scholar 

  67. Ichikawa K, Watanabe Miyano S, Minoshima Y, Matsui J, Funahashi Y. Activated FGF2 signaling pathway in tumor vasculature is essential for acquired resistance to anti-VEGF therapy. Sci Rep. 2020;10(1):2939.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Tsioumpekou M, Cunha S, Ma H, Åhgren A, Cedervall J, Olsson A, et al. Specific targeting of PDGFRβ in the stroma inhibits growth and angiogenesis in tumors with high PDGF-BB expression. Theranostics. 2020;10(3):1122–35.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Petrova T, Koh G. Biological functions of lymphatic vessels. Sci (New York, NY). 2020;369(6500):eaax4063.

    Article  CAS  Google Scholar 

  70. das Neves SP, Delivanoglou N, Da Mesquita S. CNS-draining meningeal lymphatic vasculature: roles conundrums and future challenges. Front Pharmacol. 2021;12:655052.

    Article  PubMed  PubMed Central  Google Scholar 

  71. Garnier L, Gkountidi A, Hugues S. Tumor-associated lymphatic vessel features and immunomodulatory functions. Front Immunol. 2019;10:720.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Brouillard P, Dupont L, Helaers R, Coulie R, Tiller G, Peeden J, et al. Loss of ADAMTS3 activity causes Hennekam lymphangiectasia-lymphedema syndrome 3. Hum Mol Genet. 2017;26(21):4095–104.

    Article  CAS  PubMed  Google Scholar 

  73. Jha S, Rauniyar K, Chronowska E, Mattonet K, Maina E, Koistinen H, et al. KLK3/PSA and cathepsin D activate VEGF-C and VEGF-D. eLife. 2019;8:e44478.

    Article  PubMed  PubMed Central  Google Scholar 

  74. Jha S, Rauniyar K, Jeltsch M. Key molecules in lymphatic development, function, and identification. Ann Anatomy Anatomischer Anzeiger Organ Anatomische Gesellschaft. 2018;219:25–34.

    Article  Google Scholar 

  75. Petrova TV, Koh GY. Biological functions of lymphatic vessels. Science. 2020;369(6500):eaax063.

    Article  Google Scholar 

  76. Yu P, Wu G, Lee H, Simons M. Endothelial metabolic control of lymphangiogenesis. BioEssays. 2018;40(6):e1700245.

    Article  PubMed  PubMed Central  Google Scholar 

  77. Wilczak W, Wittmer C, Clauditz T, Minner S, Steurer S, Büscheck F, et al. marked prognostic impact of minimal lymphatic tumor spread in prostate cancer. Eur Urol. 2018;74(3):376–86.

    Article  PubMed  Google Scholar 

  78. Chen C, Zheng H, Luo Y, Kong Y, An M, Li Y, et al. SUMOylation promotes extracellular vesicle-mediated transmission of lncRNA ELNAT1 and lymph node metastasis in bladder cancer. J Clin Invest. 2021;131(8):146431.

    Article  PubMed  Google Scholar 

  79. Qu W, Li S, Zhang M, Qiao Q. Pattern and prognosis of distant metastases in nasopharyngeal carcinoma: A large-population retrospective analysis. Cancer Med. 2020;9(17):6147–58.

    Article  PubMed  PubMed Central  Google Scholar 

  80. Jain R, Martin J, Stylianopoulos T. The role of mechanical forces in tumor growth and therapy. Annu Rev Biomed Eng. 2014;16:321–46.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Li Z, Zhang X, Liu C, Ma J. Non-immune cell components in the gastrointestinal tumor microenvironment influencing tumor immunotherapy. Front Cell Dev Biol. 2021;9:729941.

    Article  PubMed  PubMed Central  Google Scholar 

  82. Leary N, Walser S, He Y, Cousin N, Pereira P, Gallo A, et al. Melanoma-derived extracellular vesicles mediate lymphatic remodelling and impair tumour immunity in draining lymph nodes. J Extracell Vesicles. 2022;11(2):e12197.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Liu C, Wang Y, Li L, He D, Chi J, Li Q, et al. Engineered extracellular vesicles and their mimetics for cancer immunotherapy. J Control Release. 2022;349:679–98.

    Article  CAS  PubMed  Google Scholar 

  84. Tang Y, Zhang W, Sheng T, He X, Xiong X. Overview of the molecular mechanisms contributing to the formation of cancer-associated adipocytes (Review). Mol Med Rep. 2021;24(5):768.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Iyengar N, Zhou X, Mendieta H, Giri D, El-Hely O, Winston L, et al. Effects of adiposity and exercise on breast tissue and systemic metabo-inflammatory factors in women at high risk or diagnosed with breast cancer. Cancer Prev Res (Phila). 2021;14(5):541–50.

    Article  CAS  PubMed  Google Scholar 

  86. Saha A, Hamilton-Reeves J, DiGiovanni J. White adipose tissue-derived factors and prostate cancer progression: mechanisms and targets for interventions. Cancer Metastasis Rev. 2022;41:649.

    Article  PubMed  Google Scholar 

  87. Liu A, Pan W, Zhuang S, Tang Y, Zhang H. Cancer cell-derived exosomal miR-425-3p induces white adipocyte atrophy. Adipocyte. 2022;11(1):487–500.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Wu Y, Li X, Li Q, Cheng C, Zheng L. Adipose tissue-to-breast cancer crosstalk: Comprehensive insights. Biochim Biophys Acta Rev Cancer. 2022;1877(5):188800.

    Article  CAS  PubMed  Google Scholar 

  89. Seki T, Yang Y, Sun X, Lim S, Xie S, Guo Z, et al. Brown-fat-mediated tumour suppression by cold-altered global metabolism. Nature. 2022;608(7922):421–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Oguri Y, Shinoda K, Kim H, Alba D, Bolus W, Wang Q, et al. CD81 controls beige fat progenitor cell growth and energy balance via FAK signaling. Cell. 2020;182(3):563-77.e20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Han J, Jeon Y, Oh M, Lee G, Nahmgoong H, Han S, et al. Adipocyte HIF2α functions as a thermostat via PKA Cα regulation in beige adipocytes. Nat Commun. 2022;13(1):3268.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Gantov M, Pagnotta P, Lotufo C, Rindone G, Riera M, Calvo J, et al. Beige adipocytes contribute to breast cancer progression. Oncol Rep. 2021;45(1):317–28.

    Article  PubMed  Google Scholar 

  93. Suárez-Nájera L, Chanona-Pérez J, Valdivia-Flores A, Marrero-Rodríguez D, Salcedo-Vargas M, García-Ruiz D, et al. Morphometric study of adipocytes on breast cancer by means of photonic microscopy and image analysis. Microsc Res Tech. 2018;81(2):240–9.

    Article  PubMed  Google Scholar 

  94. Wu Q, Li B, Li Z, Li J, Sun S, Sun S. Cancer-associated adipocytes: key players in breast cancer progression. J Hematol Oncol. 2019;12(1):95.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Munteanu R, Onaciu A, Moldovan C, Zimta A, Gulei D, Paradiso A, et al. Adipocyte-based cell therapy in oncology: the role of cancer-associated adipocytes and their reinterpretation as delivery platforms. Pharmaceutics. 2020;12(5):402.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Zhao C, Wu M, Zeng N, Xiong M, Hu W, Lv W, et al. Cancer-associated adipocytes: emerging supporters in breast cancer. J Exp Clin Cancer Res CR. 2020;39(1):156.

    Article  CAS  PubMed  Google Scholar 

  97. Trivedi P, Wang S, Friedman S. The power of plasticity-metabolic regulation of hepatic stellate cells. Cell Metab. 2021;33(2):242–57.

    Article  CAS  PubMed  Google Scholar 

  98. Yin C, Evason K, Asahina K, Stainier D. Hepatic stellate cells in liver development, regeneration, and cancer. J Clin Investig. 2013;123(5):1902–10.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Zhao S, Mi Y, Zheng B, Wei P, Gu Y, Zhang Z, et al. Highly-metastatic colorectal cancer cell released miR-181a-5p-rich extracellular vesicles promote liver metastasis by activating hepatic stellate cells and remodelling the tumour microenvironment. J Extracell Vesicles. 2022;11(1):e12186.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Zhou Y, Ren H, Dai B, Li J, Shang L, Huang J, et al. Hepatocellular carcinoma-derived exosomal miRNA-21 contributes to tumor progression by converting hepatocyte stellate cells to cancer-associated fibroblasts. J Exp Clin Cancer Res CR. 2018;37(1):324.

    Article  CAS  PubMed  Google Scholar 

  101. Masamune A, Yoshida N, Hamada S, Takikawa T, Nabeshima T, Shimosegawa T. Exosomes derived from pancreatic cancer cells induce activation and profibrogenic activities in pancreatic stellate cells. Biochem Biophys Res Commun. 2018;495(1):71–7.

    Article  CAS  PubMed  Google Scholar 

  102. Zhang Y, Zhou Y, Zhang B, Huang S, Li P, He X, et al. Pancreatic cancer-derived exosomes promoted pancreatic stellate cells recruitment by pancreatic cancer. J Cancer. 2019;10(18):4397–407.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Schnittert J, Bansal R, Prakash J. Targeting pancreatic stellate cells in cancer. Trends Cancer. 2019;5(2):128–42.

    Article  CAS  PubMed  Google Scholar 

  104. Abyaneh H, Regenold M, McKee T, Allen C, Gauthier M. Towards extracellular matrix normalization for improved treatment of solid tumors. Theranostics. 2020;10(4):1960–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  105. Henke E, Nandigama R, Ergün S. Extracellular matrix in the tumor microenvironment and its impact on cancer therapy. Front Mol Biosci. 2019;6:160.

    Article  CAS  PubMed  Google Scholar 

  106. Ayad N, Weaver V. Tension in tumour cells keeps metabolism high. Nature. 2020;578(7796):517–8.

    Article  CAS  PubMed  Google Scholar 

  107. Chaudhuri O, Cooper-White J, Janmey P, Mooney D, Shenoy V. Effects of extracellular matrix viscoelasticity on cellular behaviour. Nature. 2020;584(7822):535–46.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Fiore V, Krajnc M, Quiroz F, Levorse J, Pasolli H, Shvartsman S, et al. Publisher correction: mechanics of a multilayer epithelium instruct tumour architecture and function. Nature. 2020;586(7827):E9.

    Article  CAS  PubMed  Google Scholar 

  109. Mendonça D, Miguez P, Mendonça G, Yamauchi M, Aragão F, Cooper L. Titanium surface topography affects collagen biosynthesis of adherent cells. Bone. 2011;49(3):463–72.

    Article  PubMed  Google Scholar 

  110. Smithen D, Leung L, Challinor M, Lawrence R, Tang H, Niculescu-Duvaz D, et al. 2-Aminomethylene-5-sulfonylthiazole Inhibitors of Lysyl Oxidase (LOX) and LOXL2 show significant efficacy in delaying tumor growth. J Med Chem. 2020;63(5):2308–24.

    Article  CAS  PubMed  Google Scholar 

  111. Sullivan W, Mullen P, Schmid E, Flores A, Momcilovic M, Sharpley M, et al. Extracellular matrix remodeling regulates glucose metabolism through TXNIP destabilization. Cell. 2018;175(1):117-32.e21.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  112. Torrino S, Grasset E, Audebert S, Belhadj I, Lacoux C, Haynes M, et al. Mechano-induced cell metabolism promotes microtubule glutamylation to force metastasis. Cell Metab. 2021;33(7):1342-57.e10.

    Article  CAS  PubMed  Google Scholar 

  113. Lequeux A, Noman MZ, Xiao M, Van Moer K, Hasmim M, Benoit A, et al. Targeting HIF-1 alpha transcriptional activity drives cytotoxic immune effector cells into melanoma and improves combination immunotherapy. Oncogene. 2021;40(28):4725–35.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Bai R, Li Y, Jian L, Yang Y, Zhao L, Wei M. The hypoxia-driven crosstalk between tumor and tumor-associated macrophages: mechanisms and clinical treatment strategies. Mol Cancer. 2022;21(1):177.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Xia X, Wang S, Ni B, Xing S, Cao H, Zhang Z, et al. Hypoxic gastric cancer-derived exosomes promote progression and metastasis via MiR-301a-3p/PHD3/HIF-1alpha positive feedback loop. Oncogene. 2020;39(39):6231–44.

    Article  CAS  PubMed  Google Scholar 

  116. Yan Y, He M, Zhao L, Wu H, Zhao Y, Han L, et al. A novel HIF-2alpha targeted inhibitor suppresses hypoxia-induced breast cancer stemness via SOD2-mtROS-PDI/GPR78-UPR(ER) axis. Cell Death Differ. 2022;29(9):1769–89.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Bui BP, Nguyen PL, Lee K, Cho J. Hypoxia-inducible factor-1: a novel therapeutic target for the management of cancer, drug resistance, and cancer-related pain. Cancers (Basel). 2022;14(24):6054.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  118. Anvari G, Bellas E. Hypoxia induces stress fiber formation in adipocytes in the early stage of obesity. Sci Rep. 2021;11(1):21473.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. Rosell-Garcia T, Palomo-Alvarez O, Rodriguez-Pascual F. A hierarchical network of hypoxia-inducible factor and SMAD proteins governs procollagen lysyl hydroxylase 2 induction by hypoxia and transforming growth factor beta1. J Biol Chem. 2019;294(39):14308–18.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  120. Maruyama T, Shimoda M, Sako A, Ueda K, Hakoda H, Sakata A, et al. Predictive effectiveness of the glasgow prognostic score for gastrointestinal stromal tumors. Nutr Cancer. 2021;73(8):1333–9.

    Article  PubMed  Google Scholar 

  121. Jiang Y, Jiang H, Wang K, Liu C, Man X, Fu Q. Hypoxia enhances the production and antitumor effect of exosomes derived from natural killer cells. Ann Transl Med. 2021;9(6):473.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  122. Qian D, Xie Y, Huang M, Gu J. Tumor-derived exosomes in hypoxic microenvironment: release mechanism, biological function and clinical application. J Cancer. 2022;13(5):1685–94.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. He G, Peng X, Wei S, Yang S, Li X, Huang M, et al. Exosomes in the hypoxic TME: from release, uptake and biofunctions to clinical applications. Mol Cancer. 2022;21(1):19.

    Article  PubMed  PubMed Central  Google Scholar 

  124. Dorayappan KDP, Wanner R, Wallbillich JJ, Saini U, Zingarelli R, Suarez AA, et al. Hypoxia-induced exosomes contribute to a more aggressive and chemoresistant ovarian cancer phenotype: a novel mechanism linking STAT3/Rab proteins. Oncogene. 2018;37(28):3806–21.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Panigrahi GK, Praharaj PP, Peak TC, Long J, Singh R, Rhim JS, et al. Hypoxia-induced exosome secretion promotes survival of African-American and Caucasian prostate cancer cells. Sci Rep. 2018;8(1):3853.

    Article  PubMed  PubMed Central  Google Scholar 

  126. Li B, Antonyak MA, Zhang J, Cerione RA. RhoA triggers a specific signaling pathway that generates transforming microvesicles in cancer cells. Oncogene. 2012;31(45):4740–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  127. Kumar A, Deep G. Hypoxia in tumor microenvironment regulates exosome biogenesis: molecular mechanisms and translational opportunities. Cancer Lett. 2020;479:23–30.

    Article  CAS  PubMed  Google Scholar 

  128. Crewe C, Funcke J, Li S, Joffin N, Gliniak C, Ghaben A, et al. Extracellular vesicle-based interorgan transport of mitochondria from energetically stressed adipocytes. Cell Metab. 2021;33(9):1853-68.e11.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Tang X, Guo T, Gao X, Wu X, Xing X, Ji J, et al. Exosome-derived noncoding RNAs in gastric cancer: functions and clinical applications. Mol Cancer. 2021;20(1):99.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  130. Book A. ISEV2017. J Extracell Vesicles. 2017;6(sup1):1310414.

    Article  Google Scholar 

  131. Fan Q, Yang L, Zhang X, Peng X, Wei S, Su D, et al. The emerging role of exosome-derived non-coding RNAs in cancer biology. Cancer Lett. 2018;414:107–15.

    Article  CAS  PubMed  Google Scholar 

  132. Fukuda M. Rab27 effectors, pleiotropic regulators in secretory pathways. Traffic (Copenhagen, Denmark). 2013;14(9):949–63.

    Article  CAS  PubMed  Google Scholar 

  133. Wang M, Yu F, Li P, Wang K. Emerging function and clinical significance of exosomal circRNAs in cancer. Mol Ther Nucleic Acids. 2020;21:367–83.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  134. Fukuda M. Rab27 effectors, pleiotropic regulators in secretory pathways. Traffic. 2013;14(9):949–63.

    Article  CAS  PubMed  Google Scholar 

  135. Klinkert K, Echard A. Rab35 GTPase: a central regulator of phosphoinositides and F-actin in endocytic recycling and beyond. Traffic (Copenhagen, Denmark). 2016;17(10):1063–77.

    Article  CAS  PubMed  Google Scholar 

  136. Arbo B, Cechinel L, Palazzo R, Siqueira I. Endosomal dysfunction impacts extracellular vesicle release: Central role in Aβ pathology. Ageing Res Rev. 2020;58:101006.

    Article  CAS  PubMed  Google Scholar 

  137. Corbeil D, Santos M, Karbanová J, Kurth T, Rappa G, Lorico A. Uptake and fate of extracellular membrane vesicles: nucleoplasmic reticulum-associated late endosomes as a new gate to intercellular communication. Cells. 2020;9(9):1931.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  138. Li X, Wang Y, Wang Q, Liu Y, Bao W, Wu S. Exosomes in cancer: Small transporters with big functions. Cancer Lett. 2018;435:55–65.

    Article  CAS  PubMed  Google Scholar 

  139. Simonetti B, Cullen P. Actin-dependent endosomal receptor recycling. Curr Opin Cell Biol. 2019;56:22–33.

    Article  CAS  PubMed  Google Scholar 

  140. Nam G, Choi Y, Kim G, Kim S, Kim S, Kim I. Emerging prospects of exosomes for cancer treatment: from conventional therapy to immunotherapy. Adv Materials (Deerfield Beach, Fla). 2020;32(51):e2002440.

    Article  Google Scholar 

  141. Fu M, Gu J, Jiang P, Qian H, Xu W, Zhang X. Exosomes in gastric cancer: roles, mechanisms, and applications. Mol Cancer. 2019;18(1):41.

    Article  PubMed  PubMed Central  Google Scholar 

  142. Qiu Y, Chen T, Hu R, Zhu R, Li C, Ruan Y, et al. Next frontier in tumor immunotherapy: macrophage-mediated immune evasion. Biomark Res. 2021;9(1):72.

    Article  PubMed  PubMed Central  Google Scholar 

  143. Gupta S, Rawat S, Arora V, Kottarath S, Dinda A, Vaishnav P, et al. An improvised one-step sucrose cushion ultracentrifugation method for exosome isolation from culture supernatants of mesenchymal stem cells. Stem Cell Res Ther. 2018;9(1):180.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  144. Tian Y, Gong M, Hu Y, Liu H, Zhang W, Zhang M, et al. Quality and efficiency assessment of six extracellular vesicle isolation methods by nano-flow cytometry. J Extracell Vesicles. 2020;9(1):1697028.

    Article  CAS  PubMed  Google Scholar 

  145. Shirejini S, Inci F. The Yin and Yang of exosome isolation methods: conventional practice, microfluidics, and commercial kits. Biotechnol Adv. 2022;54:107814.

    Article  CAS  PubMed  Google Scholar 

  146. Tayebi M, Yang D, Collins D, Ai Y. Deterministic sorting of submicrometer particles and extracellular vesicles using a combined electric and acoustic field. Nano Lett. 2021;21(16):6835–42.

    Article  CAS  PubMed  Google Scholar 

  147. Tenchov R, Sasso JM, Wang X, Liaw WS, Chen CA, Zhou QA. Exosomes horizontal line nature’s lipid nanoparticles, a rising star in drug delivery and diagnostics. ACS Nano. 2022;16(11):17802–46.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  148. Balaj L, Lessard R, Dai L, Cho YJ, Pomeroy SL, Breakefield XO, et al. Tumour microvesicles contain retrotransposon elements and amplified oncogene sequences. Nat Commun. 2011;2:180.

    Article  PubMed  Google Scholar 

  149. Isaac R, Reis FCG, Ying W, Olefsky JM. Exosomes as mediators of intercellular crosstalk in metabolism. Cell Metab. 2021;33(9):1744–62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  150. Valadi H, Ekstrom K, Bossios A, Sjostrand M, Lee JJ, Lotvall JO. Exosome-mediated transfer of mRNAs and microRNAs is a novel mechanism of genetic exchange between cells. Nat Cell Biol. 2007;9(6):654–9.

    Article  CAS  PubMed  Google Scholar 

  151. He X, Tian F, Guo F, Zhang F, Zhang H, Ji J, et al. Circulating exosomal mRNA signatures for the early diagnosis of clear cell renal cell carcinoma. BMC Med. 2022;20(1):270.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  152. Sansone P, Savini C, Kurelac I, Chang Q, Amato LB, Strillacci A, et al. Packaging and transfer of mitochondrial DNA via exosomes regulate escape from dormancy in hormonal therapy-resistant breast cancer. Proc Natl Acad Sci U S A. 2017;114(43):E9066–75.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. Torralba D, Baixauli F, Villarroya-Beltri C, Fernandez-Delgado I, Latorre-Pellicer A, Acin-Perez R, et al. Priming of dendritic cells by DNA-containing extracellular vesicles from activated T cells through antigen-driven contacts. Nat Commun. 2018;9(1):2658.

    Article  PubMed  PubMed Central  Google Scholar 

  154. Skotland T, Sagini K, Sandvig K, Llorente A. An emerging focus on lipids in extracellular vesicles. Adv Drug Deliv Rev. 2020;159:308–21.

    Article  CAS  PubMed  Google Scholar 

  155. Menck K, Sonmezer C, Worst TS, Schulz M, Dihazi GH, Streit F, et al. Neutral sphingomyelinases control extracellular vesicles budding from the plasma membrane. J Extracell Vesicles. 2017;6(1):1378056.

    Article  PubMed  PubMed Central  Google Scholar 

  156. Kobuna H, Inoue T, Shibata M, Gengyo-Ando K, Yamamoto A, Mitani S, et al. Multivesicular body formation requires OSBP-related proteins and cholesterol. PLoS Genet. 2010;6(8):e1001055.

    Article  PubMed  PubMed Central  Google Scholar 

  157. Eden ER, Sanchez-Heras E, Tsapara A, Sobota A, Levine TP, Futter CE. Annexin A1 tethers membrane contact sites that mediate ER to endosome cholesterol transport. Dev Cell. 2016;37(5):473–83.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  158. Majumdar R, Tavakoli Tameh A, Arya SB, Parent CA. Exosomes mediate LTB4 release during neutrophil chemotaxis. PLoS Biol. 2021;19(7):e3001271.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  159. Arya SB, Chen S, Jordan-Javed F, Parent CA. Ceramide-rich microdomains facilitate nuclear envelope budding for non-conventional exosome formation. Nat Cell Biol. 2022;24(7):1019–28.

    Article  CAS  PubMed  Google Scholar 

  160. Wu X, Zhu D, Tian J, Tang X, Guo H, Ma J, et al. Granulocytic myeloid-derived suppressor cell exosomal prostaglandin E2 ameliorates collagen-induced arthritis by enhancing IL-10(+) B cells. Front Immunol. 2020;11:588500.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  161. Zhang M, Xin W, Ma C, Zhang H, Mao M, Liu Y, et al. Exosomal 15-LO2 mediates hypoxia-induced pulmonary artery hypertension in vivo and in vitro. Cell Death Dis. 2018;9(10):1022.

    Article  PubMed  PubMed Central  Google Scholar 

  162. Eguchi T, Sheta M, Fujii M, Calderwood SK. Cancer extracellular vesicles, tumoroid models, and tumor microenvironment. Semin Cancer Biol. 2022;86(Pt 1):112–26.

    Article  CAS  PubMed  Google Scholar 

  163. Shi F, Deng Z, Zhou Z, Jiang B, Jiang CY, Zhao RZ, et al. Heat injured stromal cells-derived exosomal EGFR enhances prostatic wound healing after thulium laser resection through EMT and NF-kappaB signaling. Prostate. 2019;79(11):1238–55.

    Article  CAS  PubMed  Google Scholar 

  164. Zhang Z, Zhou Y, Jia Y, Wang C, Zhang M, Xu Z. PRR34-AS1 promotes exosome secretion of VEGF and TGF-beta via recruiting DDX3X to stabilize Rab27a mRNA in hepatocellular carcinoma. J Transl Med. 2022;20(1):491.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  165. Patnam S, Samal R, Koyyada R, Joshi P, Singh AD, Nagalla B, et al. Exosomal PTEN as a predictive marker of aggressive gliomas. Neurol India. 2022;70(1):215–22.

    PubMed  Google Scholar 

  166. Qin Z, Li Y, Li J, Jiang L, Zhang Z, Chang K, et al. Exosomal STAT1 derived from high phosphorus-stimulated vascular endothelial cells induces vascular smooth muscle cell calcification via the Wnt/beta-catenin signaling pathway. Int J Mol Med. 2022;50(6):139.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  167. Clement E, Lazar I, Attane C, Carrie L, Dauvillier S, Ducoux-Petit M, et al. Adipocyte extracellular vesicles carry enzymes and fatty acids that stimulate mitochondrial metabolism and remodeling in tumor cells. EMBO J. 2020;39(3):e102525.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  168. Wang D, Zhao C, Xu F, Zhang A, Jin M, Zhang K, et al. Cisplatin-resistant NSCLC cells induced by hypoxia transmit resistance to sensitive cells through exosomal PKM2. Theranostics. 2021;11(6):2860–75.

    Article  PubMed  PubMed Central  Google Scholar 

  169. Zhu L, Xu Y, Kang S, Lin B, Zhang C, You Z, et al. Quantification-promoted discovery of glycosylated exosomal PD-L1 as a potential tumor biomarker. Small Methods. 2022;6(9):e2200549.

    Article  PubMed  Google Scholar 

  170. Tan Y, Huang Y, Mei R, Mao F, Yang D, Liu J, et al. HucMSC-derived exosomes delivered BECN1 induces ferroptosis of hepatic stellate cells via regulating the xCT/GPX4 axis. Cell Death Dis. 2022;13(4):319.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  171. Mańka R, Janas P, Sapoń K, Janas T, Janas T. Role of RNA motifs in RNA interaction with membrane lipid rafts: implications for therapeutic applications of exosomal RNAs. Int J Mol Sci. 2021;22(17):9416.

    Article  PubMed  PubMed Central  Google Scholar 

  172. Matsui T, Sakamaki Y, Nakashima S, Fukuda M. Rab39 and its effector UACA regulate basolateral exosome release from polarized epithelial cells. Cell Rep. 2022;39(9):110875.

    Article  CAS  PubMed  Google Scholar 

  173. Kosaka N, Iguchi H, Hagiwara K, Yoshioka Y, Takeshita F, Ochiya T. Neutral sphingomyelinase 2 (nSMase2)-dependent exosomal transfer of angiogenic microRNAs regulate cancer cell metastasis. J Biol Chem. 2013;288(15):10849–59.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  174. Villarroya-Beltri C, Gutierrez-Vazquez C, Sanchez-Cabo F, Perez-Hernandez D, Vazquez J, Martin-Cofreces N, et al. Sumoylated hnRNPA2B1 controls the sorting of miRNAs into exosomes through binding to specific motifs. Nat Commun. 2013;4:2980.

    Article  PubMed  Google Scholar 

  175. Koppers-Lalic D, Hackenberg M, Bijnsdorp IV, van Eijndhoven MAJ, Sadek P, Sie D, et al. Nontemplated nucleotide additions distinguish the small RNA composition in cells from exosomes. Cell Rep. 2014;8(6):1649–58.

    Article  CAS  PubMed  Google Scholar 

  176. Guduric-Fuchs J, O’Connor A, Camp B, O’Neill CL, Medina RJ, Simpson DA. Selective extracellular vesicle-mediated export of an overlapping set of microRNAs from multiple cell types. BMC Genomics. 2012;13:357.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  177. Garcia-Martin R, Wang G, Brandão B, Zanotto T, Shah S, Kumar Patel S, et al. MicroRNA sequence codes for small extracellular vesicle release and cellular retention. Nature. 2022;601(7893):446–51.

    Article  CAS  PubMed  Google Scholar 

  178. Willson J. RILP gets cleaved and exosomes leave. Nat Rev Mol Cell Biol. 2020;21(11):658–9.

    Article  CAS  PubMed  Google Scholar 

  179. Statello L, Maugeri M, Garre E, Nawaz M, Wahlgren J, Papadimitriou A, et al. Identification of RNA-binding proteins in exosomes capable of interacting with different types of RNA: RBP-facilitated transport of RNAs into exosomes. PLoS ONE. 2018;13(4):e0195969.

    Article  PubMed  PubMed Central  Google Scholar 

  180. Chen C, Yu H, Han F, Lai X, Ye K, Lei S, et al. Tumor-suppressive circRHOBTB3 is excreted out of cells via exosome to sustain colorectal cancer cell fitness. Mol Cancer. 2022;21(1):46.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  181. Li Y, Zheng Q, Bao C, Li S, Guo W, Zhao J, et al. Circular RNA is enriched and stable in exosomes: a promising biomarker for cancer diagnosis. Cell Res. 2015;25(8):981–4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  182. Yokoi A, Villar-Prados A, Oliphint PA, Zhang J, Song X, De Hoff P, et al. Mechanisms of nuclear content loading to exosomes. Sci Adv. 2019;5(11):eaax8849.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  183. Ferreira J, da Rosa SA, Pereira P. LAMP2A mediates the loading of proteins into endosomes and selects exosomal cargo. Autophagy. 2022;18(9):2263–5.

    Article  CAS  PubMed  Google Scholar 

  184. Ferreira J, da Rosa Soares A, Ramalho J, Máximo Carvalho C, Cardoso M, Pintado P, et al. LAMP2A regulates the loading of proteins into exosomes. Sci Adv. 2022;8(12):eabm1140.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  185. Albacete-Albacete L, Navarro-Lérida I, López J, Martín-Padura I, Astudillo A, Ferrarini A, et al. ECM deposition is driven by caveolin-1-dependent regulation of exosomal biogenesis and cargo sorting. J Cell Biol. 2020;219(11):e202006178.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  186. Ageta H, Ageta-Ishihara N, Hitachi K, Karayel O, Onouchi T, Yamaguchi H, et al. UBL3 modification influences protein sorting to small extracellular vesicles. Nat Commun. 2018;9(1):3936.

    Article  PubMed  PubMed Central  Google Scholar 

  187. Sahu R, Kaushik S, Clement C, Cannizzo E, Scharf B, Follenzi A, et al. Microautophagy of cytosolic proteins by late endosomes. Dev Cell. 2011;20(1):131–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  188. Cai T, Zhang Q, Wu B, Wang J, Li N, Zhang T, et al. LncRNA-encoded microproteins: a new form of cargo in cell culture-derived and circulating extracellular vesicles. J Extracell Vesicles. 2021;10(9):e12123.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  189. Hinger S, Cha D, Franklin J, Higginbotham J, Dou Y, Ping J, et al. Diverse long RNAs are differentially sorted into extracellular vesicles secreted by colorectal cancer cells. Cell Rep. 2018;25(3):715-25.e4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  190. Vats S, Galli T. Role of SNAREs in unconventional secretion-focus on the VAMP7-dependent secretion. Front Cell Dev Biol. 2022;10:884020.

    Article  PubMed  PubMed Central  Google Scholar 

  191. Vietri M, Radulovic M, Stenmark H. The many functions of ESCRTs. Nat Rev Mol Cell Biol. 2020;21(1):25–42.

    Article  CAS  PubMed  Google Scholar 

  192. Guo L, Zhang Y, Wei R, Zhang X, Wang C, Feng M. viaProinflammatory macrophage-derived microvesicles exhibit tumor tropism dependent on CCL2/CCR2 signaling axis and promote drug delivery SNARE-mediated membrane fusion. Theranostics. 2020;10(15):6581–98.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  193. An M, Lohse I, Tan Z, Zhu J, Wu J, Kurapati H, et al. Quantitative proteomic analysis of serum exosomes from patients with locally advanced pancreatic cancer undergoing chemoradiotherapy. J Proteome Res. 2017;16(4):1763–72.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  194. De Lellis L, Florio R, Di Bella MC, Brocco D, Guidotti F, Tinari N, et al. Exosomes as pleiotropic players in pancreatic cancer. Biomedicines. 2021;9(3):275.

    Article  PubMed  PubMed Central  Google Scholar 

  195. Zhao H, Yang L, Baddour J, Achreja A, Bernard V, Moss T, et al. Tumor microenvironment derived exosomes pleiotropically modulate cancer cell metabolism. Elife. 2016;5:e10250.

    Article  PubMed  PubMed Central  Google Scholar 

  196. Zebrowska A, Jelonek K, Mondal S, Gawin M, Mrowiec K, Widłak P, et al. Proteomic and metabolomic profiles of T cell-derived exosomes isolated from human plasma. Cells. 2022;11(12):1965.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  197. Palacios-Ferrer J, García-Ortega M, Gallardo-Gómez M, García M, Díaz C, Boulaiz H, et al. Metabolomic profile of cancer stem cell-derived exosomes from patients with malignant melanoma. Mol Oncol. 2021;15(2):407–28.

    Article  CAS  PubMed  Google Scholar 

  198. Pavlova N, Zhu J, Thompson C. The hallmarks of cancer metabolism: Still emerging. Cell Metab. 2022;34(3):355–77.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  199. Stine Z, Schug Z, Salvino J, Dang C. Targeting cancer metabolism in the era of precision oncology. Nat Rev Drug Discovery. 2022;21(2):141–62.

    Article  CAS  PubMed  Google Scholar 

  200. Brunner J, Finley L. SnapShot: Cancer metabolism. Mol cell. 2021;81(18):3878-e1.

    Article  CAS  PubMed  Google Scholar 

  201. Noe J, Rendon B, Geller A, Conroy L, Morrissey S, Young L, et al. Lactate supports a metabolic-epigenetic link in macrophage polarization. Sci Adv. 2021;7(46):eabi8602.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  202. Du D, Liu C, Qin M, Zhang X, Xi T, Yuan S, et al. Metabolic dysregulation and emerging therapeutical targets for hepatocellular carcinoma. Acta pharmaceutica Sinica B. 2022;12(2):558–80.

    Article  CAS  PubMed  Google Scholar 

  203. Xu I, Lai R, Lin S, Tse A, Chiu D, Koh H, et al. Transketolase counteracts oxidative stress to drive cancer development. Proc Natl Acad Sci USA. 2016;113(6):E725–34.

    Article  PubMed  PubMed Central  Google Scholar 

  204. Tseng H, Zeng Y, Lin Y, Huang J, Lin C, Lee M, et al. A novel AMPK activator shows therapeutic potential in hepatocellular carcinoma by suppressing HIF1α-mediated aerobic glycolysis. Mol Oncol. 2022;16(11):2274–94.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  205. Huang Y, Chen Z, Lu T, Bi G, Li M, Liang J, et al. HIF-1α switches the functionality of TGF-β signaling via changing the partners of smads to drive glucose metabolic reprogramming in non-small cell lung cancer. J Exp Clin Cancer Res: CR. 2021;40(1):398.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  206. Li Y, Li Y, Liu X, Zhang L, Chen Y, Zhao Q, et al. Blockage of citrate export prevents TCA cycle fragmentation via Irg1 inactivation. Cell Rep. 2022;38(7):110391.

    Article  CAS  PubMed  Google Scholar 

  207. Qi X, Zhang Y, Ji H, Wu X, Wang F, Xie M, et al. Knockdown of prohibitin expression promotes glucose metabolism in eutopic endometrial stromal cells from women with endometriosis. Reprod Biomed Online. 2014;29(6):761–70.

    Article  CAS  PubMed  Google Scholar 

  208. Mittal A, Nenwani M, Sarangi I, Achreja A, Lawrence T, Nagrath D. Radiotherapy-induced metabolic hallmarks in the tumor microenvironment. Trends Cancer. 2022;8:855.

    Article  CAS  PubMed  Google Scholar 

  209. Liu L, Yuan L, Huang D, Han Q, Cai J, Wang S, et al. miR126 regulates the progression of epithelial ovarian cancer in vitro and in vivo by targeting VEGFA. Int J Oncol. 2020;57(3):825–34.

    Article  CAS  PubMed  Google Scholar 

  210. Castillo-Sanchez R, Churruca-Schuind A, Martinez-Ival M, Salazar EP. Cancer-associated fibroblasts communicate with breast tumor cells through extracellular vesicles in tumor development. Technol Cancer Res Treat. 2022;21:15330338221131648.

    Article  PubMed  PubMed Central  Google Scholar 

  211. Kumagai S, Koyama S, Itahashi K, Tanegashima T, Lin Y, Togashi Y, et al. Lactic acid promotes PD-1 expression in regulatory T cells in highly glycolytic tumor microenvironments. Cancer Cell. 2022;40(2):201-18.e9.

    Article  CAS  PubMed  Google Scholar 

  212. Zappasodi R, Serganova I, Cohen I, Maeda M, Shindo M, Senbabaoglu Y, et al. CTLA-4 blockade drives loss of T stability in glycolysis-low tumours. Nature. 2021;591(7851):652–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  213. Koundouros N, Poulogiannis G. Reprogramming of fatty acid metabolism in cancer. Br J Cancer. 2020;122(1):4–22.

    Article  CAS  PubMed  Google Scholar 

  214. Chassen S, Ferchaud-Roucher V, Gupta M, Jansson T, Powell T. Alterations in placental long chain polyunsaturated fatty acid metabolism in human intrauterine growth restriction. Clin Sci (London, England : 1979). 2018;132(5):595–607.

    Article  CAS  Google Scholar 

  215. Zhao H, Li Y. Cancer metabolism and intervention therapy. Mol Biomed. 2021;2(1):5.

    Article  PubMed  PubMed Central  Google Scholar 

  216. de Araujo JR, Eich C, Jorquera C, Schomann T, Baldazzi F, Chan A, et al. Ceramide and palmitic acid inhibit macrophage-mediated epithelial-mesenchymal transition in colorectal cancer. Mol Cell Biochem. 2020;468:153–68.

    Article  Google Scholar 

  217. Das U. Essential fatty acids and their metabolites in the pathobiology of inflammation and its resolution. Biomolecules. 2021;11(12):1873.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  218. Wang B, Wu L, Chen J, Dong L, Chen C, Wen Z, et al. Metabolism pathways of arachidonic acids: mechanisms and potential therapeutic targets. Signal Transduct Target Ther. 2021;6(1):94.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  219. Kennedy B, Harris R. Cyclooxygenase and lipoxygenase gene expression in the inflammogenesis of breast cancer. Inflammopharmacology. 2018;26:909.

    Article  CAS  Google Scholar 

  220. Vecchi L, Araújo T, Azevedo F, Mota S, Ávila V, Ribeiro M, et al. Phospholipase a drives tumorigenesis and cancer aggressiveness through its interaction with annexin A1. Cells. 2021;10(6):1472.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  221. Osman W, Youssef N. Combined use of COX-1 and VEGF immunohistochemistry refines the histopathologic prognosis of renal cell carcinoma. Int J Clin Exp Pathol. 2015;8(7):8165–77.

    PubMed  PubMed Central  Google Scholar 

  222. O’Sullivan D, van der Windt G, Huang S, Curtis J, Chang C, Buck M, et al. Memory CD8(+) T cells use cell-intrinsic lipolysis to support the metabolic programming necessary for development. Immunity. 2014;41(1):75–88.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  223. Liu QP, Lin JY, An P, Chen YY, Luan X, Zhang H. LncRNAs in tumor microenvironment: The potential target for cancer treatment with natural compounds and chemical drugs. Biochem Pharmacol. 2021;193:114802.

    Article  CAS  PubMed  Google Scholar 

  224. Machado M, Patente T, Rouillé Y, Heumel S, Melo E, Deruyter L, et al. Streptococcus pneumoniaeAcetate improves the killing of by alveolar macrophages NLRP3 inflammasome and glycolysis-HIF-1α Axis. Front Immunol. 2022;13:773261.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  225. Pranzini E, Pardella E, Paoli P, Fendt S, Taddei M. Metabolic reprogramming in anticancer drug resistance: a focus on amino acids. Trends Cancer. 2021;7(8):682–99.

    Article  CAS  PubMed  Google Scholar 

  226. Reinfeld B, Madden M, Wolf M, Chytil A, Bader J, Patterson A, et al. Cell-programmed nutrient partitioning in the tumour microenvironment. Nature. 2021;593(7858):282–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  227. Durán R, Oppliger W, Robitaille A, Heiserich L, Skendaj R, Gottlieb E, et al. Glutaminolysis activates Rag-mTORC1 signaling. Mol Cell. 2012;47(3):349–58.

    Article  PubMed  Google Scholar 

  228. Mattaini K, Sullivan M, Vander HM. The importance of serine metabolism in cancer. J Cell Biol. 2016;214(3):249–57.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  229. Xie M, Pei D. Serine hydroxymethyltransferase 2: a novel target for human cancer therapy. Invest New Drugs. 2021;39(6):1671–81.

    Article  CAS  PubMed  Google Scholar 

  230. Peyraud F, Guegan J, Bodet D, Cousin S, Bessede A, Italiano A. Targeting tryptophan catabolism in cancer immunotherapy era: challenges and perspectives. Front Immunol. 2022;13:807271.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  231. Chen Y, Chen J, Guo D, Yang P, Chen S, Zhao C, et al. Tryptophan metabolites as biomarkers for esophageal cancer susceptibility, metastasis, and prognosis. Front Oncol. 2022;12:800291.

    Article  PubMed  PubMed Central  Google Scholar 

  232. Wu D, Zhu Y. Role of kynurenine in promoting the generation of exhausted CD8 T cells in colorectal cancer. Am J Transl Res. 2021;13(3):1535–47.

    CAS  PubMed  PubMed Central  Google Scholar 

  233. Luo P, Yin P, Hua R, Tan Y, Li Z, Qiu G, et al. A Large-scale, multicenter serum metabolite biomarker identification study for the early detection of hepatocellular carcinoma. Hepatology (Baltimore, MD). 2018;67(2):662–75.

    Article  CAS  PubMed  Google Scholar 

  234. Udumula M, Sakr S, Dar S, Alvero A, Ali-Fehmi R, Abdulfatah E, et al. Ovarian cancer modulates the immunosuppressive function of CD11bGr1 myeloid cells via glutamine metabolism. Mol Metab. 2021;53:101272.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  235. Oh M, Sun I, Zhao L, Leone R, Sun I, Xu W, et al. Targeting glutamine metabolism enhances tumor-specific immunity by modulating suppressive myeloid cells. J Clin Investig. 2020;130(7):3865–84.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  236. Oh MH, Sun IH, Zhao L, Leone RD, Sun IM, Xu W, et al. Targeting glutamine metabolism enhances tumor-specific immunity by modulating suppressive myeloid cells. J Clin Invest. 2020;130(7):3865–84.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  237. Chen L, Zhu S, Liu T, Zhao H, Chen P, Duan Y, et al. Cancer associated fibroblasts promote renal cancer progression through a TDO/Kyn/AhR dependent signaling pathway. Front Oncol. 2021;11:628821.

    Article  PubMed  PubMed Central  Google Scholar 

  238. Yan W, Wu X, Zhou W, Fong M, Cao M, Liu J, 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(5):597–609.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  239. Sung J, Kang C, Kang S, Jang Y, Chae Y, Kim B, et al. ITGB4-mediated metabolic reprogramming of cancer-associated fibroblasts. Oncogene. 2020;39(3):664–76.

    Article  CAS  PubMed  Google Scholar 

  240. Rai A, Greening D, Chen M, Xu R, Ji H, Simpson R. Exosomes derived from human primary and metastatic colorectal cancer cells contribute to functional heterogeneity of activated fibroblasts by reprogramming their proteome. Proteomics. 2019;19(8):e1800148.

    Article  PubMed  Google Scholar 

  241. Zhang C, Wang XY, Zhang P, He TC, Han JH, Zhang R, et al. Cancer-derived exosomal HSPC111 promotes colorectal cancer liver metastasis by reprogramming lipid metabolism in cancer-associated fibroblasts. Cell Death Dis. 2022;13(1):57.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  242. Ma X, Wang J, Li J, Ma C, Chen S, Lei W, et al. Loading MiR-210 in endothelial progenitor cells derived exosomes boosts their beneficial effects on hypoxia/Reoxygeneation-injured human endothelial cells via protecting mitochondrial function. Cell Physiol Biochem: Int J Exp Cell Physiol, Biochem, Pharmacol. 2018;46(2):664–75.

    Article  CAS  Google Scholar 

  243. Wang B, Wang X, Hou D, Huang Q, Zhan W, Chen C, et al. Exosomes derived from acute myeloid leukemia cells promote chemoresistance by enhancing glycolysis-mediated vascular remodeling. J Cell Physiol. 2019;234(7):10602–14.

    Article  CAS  PubMed  Google Scholar 

  244. Ludwig N, Yerneni S, Azambuja J, Gillespie D, Menshikova E, Jackson E, et al. Tumor-derived exosomes promote angiogenesis via adenosine a receptor signaling. Angiogenesis. 2020;23(4):599–610.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  245. Liu F, Bu Z, Zhao F, Xiao D. Increased T-helper 17 cell differentiation mediated by exosome-mediated microRNA-451 redistribution in gastric cancer infiltrated T cells. Cancer Sci. 2018;109(1):65–73.

    Article  CAS  PubMed  Google Scholar 

  246. Zhou C, Zhang Y, Yan R, Huang L, Mellor A, Yang Y, 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(2):715–29.

    Article  CAS  PubMed  Google Scholar 

  247. Muller L, Simms P, Hong C, Nishimura M, Jackson E, Watkins S, et al. Human tumor-derived exosomes (TEX) regulate Treg functions via cell surface signaling rather than uptake mechanisms. Oncoimmunology. 2017;6(8):e1261243.

    Article  PubMed  PubMed Central  Google Scholar 

  248. Salimu J, Webber J, Gurney M, Al-Taei S, Clayton A, Tabi Z. Dominant immunosuppression of dendritic cell function by prostate-cancer-derived exosomes. J Extracell Vesicles. 2017;6(1):1368823.

    Article  PubMed  PubMed Central  Google Scholar 

  249. Park J, Dutta B, Tse S, Gupta N, Tan C, Low J, et al. Hypoxia-induced tumor exosomes promote M2-like macrophage polarization of infiltrating myeloid cells and microRNA-mediated metabolic shift. Oncogene. 2019;38(26):5158–73.

    Article  CAS  PubMed  Google Scholar 

  250. Zhou S, Lan Y, Li Y, Li Z, Pu J, Wei L. Hypoxic tumor-derived exosomes induce M2 macrophage polarization via PKM2/AMPK to promote lung cancer progression. Cell Transplant. 2022;31:9636897221106998.

    Article  PubMed  Google Scholar 

  251. Wang S, Xu M, Xiao X, Wang L, Sun Z, Guan M, et al. Pancreatic cancer cell exosomes induce lipidomics changes in adipocytes. Adipocyte. 2022;11(1):346–55.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  252. Wu Q, Sun S, Li Z, Yang Q, Li B, Zhu S, et al. Tumour-originated exosomal miR-155 triggers cancer-associated cachexia to promote tumour progression. Mol Cancer. 2018;17(1):155.

    Article  PubMed  PubMed Central  Google Scholar 

  253. Wu Q, Li J, Li Z, Sun S, Zhu S, Wang L, et al. Exosomes from the tumour-adipocyte interplay stimulate beige/brown differentiation and reprogram metabolism in stromal adipocytes to promote tumour progression. J Exp Clin Cancer Res. 2019;38(1):223.

    Article  PubMed  PubMed Central  Google Scholar 

  254. Wu Q, Li J, Li Z, Sun S, Zhu S, Wang L, et al. Exosomes from the tumour-adipocyte interplay stimulate beige/brown differentiation and reprogram metabolism in stromal adipocytes to promote tumour progression. J Exp Clin Cancer Res : CR. 2019;38(1):223.

    Article  PubMed  PubMed Central  Google Scholar 

  255. Sagar G, Sah R, Javeed N, Dutta S, Smyrk T, Lau J, et al. Pathogenesis of pancreatic cancer exosome-induced lipolysis in adipose tissue. Gut. 2016;65(7):1165–74.

    Article  CAS  PubMed  Google Scholar 

  256. Li J, Wu X, Schiffmann L, MacVicar T, Zhou C, Wang Z, et al. IL-17B/RB activation in pancreatic stellate cells promotes pancreatic cancer metabolism and growth. Cancers. 2021;13(21):5338.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  257. Shu S, Yang Y, Allen C, Maguire O, Minderman H, Sen A, et al. Publisher Correction: Metabolic reprogramming of stromal fibroblasts by melanoma exosome microRNA favours a pre-metastatic microenvironment. Sci Rep. 2019;9(1):4959.

    Article  PubMed  PubMed Central  Google Scholar 

  258. Zhang H, Deng T, Liu R, Ning T, Yang H, Liu D, et al. CAF secreted miR-522 suppresses ferroptosis and promotes acquired chemo-resistance in gastric cancer. Mol Cancer. 2020;19(1):43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  259. 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(3):1084–99.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  260. Lu L, Huang J, Mo J, Da X, Li Q, Fan M, et al. Exosomal lncRNA TUG1 from cancer-associated fibroblasts promotes liver cancer cell migration, invasion, and glycolysis by regulating the miR-524-5p/SIX1 axis. Cell Mol Biol Lett. 2022;27(1):17.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  261. Liu T, Han C, Fang P, Ma Z, Wang X, Chen H, et al. Cancer-associated fibroblast-specific lncRNA LINC01614 enhances glutamine uptake in lung adenocarcinoma. J Hematol Oncol. 2022;15(1):141.

    Article  PubMed  PubMed Central  Google Scholar 

  262. Sansone P, Savini C, Kurelac I, Chang Q, Amato L, Strillacci A, et al. Packaging and transfer of mitochondrial DNA via exosomes regulate escape from dormancy in hormonal therapy-resistant breast cancer. Proc Natl Acad Sci USA. 2017;114(43):E9066–75.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  263. Huang S, Fan P, Zhang C, Xie J, Gu X, Lei S, et al. Exosomal microRNA-503-3p derived from macrophages represses glycolysis and promotes mitochondrial oxidative phosphorylation in breast cancer cells by elevating DACT2. Cell Death Discov. 2021;7(1):119.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  264. Wang P, Li GY, Zhou L, Jiang HL, Yang Y, Wu HT. Exosomes from M2 macrophages promoted glycolysis in FaDu cells by inhibiting PDLIM2 expression to stabilize PFKL. Neoplasma. 2022;69(5):1041–53.

    Article  PubMed  Google Scholar 

  265. Chen F, Chen J, Yang L, Liu J, Zhang X, Zhang Y, 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(4):498–510.

    Article  CAS  PubMed  Google Scholar 

  266. Lazar I, Clement E, Dauvillier S, Milhas D, Ducoux-Petit M, LeGonidec S, et al. Adipocyte exosomes promote melanoma aggressiveness through fatty acid oxidation: a novel mechanism linking obesity and cancer. Can Res. 2016;76(14):4051–7.

    Article  CAS  Google Scholar 

  267. Liu Y, Tan J, Ou S, Chen J, Chen L. Adipose-derived exosomes deliver miR-23a/b to regulate tumor growth in hepatocellular cancer by targeting the VHL/HIF axis. J Physiol Biochem. 2019;75(3):391–401.

    Article  CAS  PubMed  Google Scholar 

  268. Yin H, Qiu X, Shan Y, You B, Xie L, Zhang P, et al. HIF-1α downregulation of miR-433-3p in adipocyte-derived exosomes contributes to NPC progression via targeting SCD1. Cancer Sci. 2021;112(4):1457–70.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  269. Alamoudi A, Alnoury A, Gad H. miRNA in tumour metabolism and why could it be the preferred pathway for energy reprograming. Brief Funct Genomics. 2018;17(3):157–69.

    Article  CAS  PubMed  Google Scholar 

  270. Sung JY, Cheong JH. New immunometabolic strategy based on cell type-specific metabolic reprogramming in the tumor immune microenvironment. Cells. 2022;11(5):768.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  271. Martínez-Reyes I, Chandel N. Cancer metabolism: looking forward. Nat Rev Cancer. 2021;21(10):669–80.

    Article  PubMed  Google Scholar 

  272. Huang W, Fan L, Tang Y, Chi Y, Li J. A pan-cancer analysis of the oncogenic role of integrin Beta4 (ITGB4) in human tumors. Int J Gen Med. 2021;14:9629–45.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  273. Wang WT, Jin WL, Li X. Intercellular communication in the tumour microecosystem: Mediators and therapeutic approaches for hepatocellular carcinoma. Biochim Biophys Acta Mol Basis Dis. 2022;1868(12):166528.

    Article  CAS  PubMed  Google Scholar 

  274. Bai S, Wei Y, Liu R, Xu R, Xiang L, Du J. Role of tumour-derived exosomes in metastasis. Biomed Pharmacother. 2022;147:112657.

    Article  CAS  PubMed  Google Scholar 

  275. Ludwig N, Gillespie D, Reichert T, Jackson E, Whiteside T. Purine metabolites in tumor-derived exosomes may facilitate immune escape of head and neck squamous cell carcinoma. Cancers. 2020;12(6):1602.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  276. Guo W, Qiao T, Dong B, Li T, Liu Q, Xu X. The effect of hypoxia-induced exosomes on anti-tumor immunity and its implication for immunotherapy. Front Immunol. 2022;13:915985.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  277. Zhang Y, Mao Q, Xia Q, Cheng J, Huang Z, Li Y, et al. Noncoding RNAs link metabolic reprogramming to immune microenvironment in cancers. J Hematol Oncol. 2021;14(1):169.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  278. Jhunjhunwala S, Hammer C, Delamarre L. Antigen presentation in cancer: insights into tumour immunogenicity and immune evasion. Nat Rev Cancer. 2021;21(5):298–312.

    Article  CAS  PubMed  Google Scholar 

  279. Nazari N, Jafari F, Ghalamfarsa G, Hadinia A, Atapour A, Ahmadi M, et al. The emerging role of microRNA in regulating the mTOR signaling pathway in immune and inflammatory responses. Immunol Cell Biol. 2021;99(8):814–32.

    Article  CAS  PubMed  Google Scholar 

  280. Sun L, Wang Y, Wang X, Navarro-Corcuera A, Ilyas S, Jalan-Sakrikar N, et al. PD-L1 promotes myofibroblastic activation of hepatic stellate cells by distinct mechanisms selective for TGF-β receptor I versus II. Cell Rep. 2022;38(6):110349.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  281. Zhao L, Ma X, Yu J. Exosomes and organ-specific metastasis. Mol Ther Methods Clin Dev. 2021;22:133–47.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  282. Chen Y, McAndrews K, Kalluri R. Clinical and therapeutic relevance of cancer-associated fibroblasts. Nat Rev Clin Oncol. 2021;18(12):792–804.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  283. Miki Y, Yashiro M, Moyano-Galceran L, Sugimoto A, Ohira M, Lehti K. Crosstalk between cancer associated fibroblasts and cancer cells in scirrhous type gastric cancer. Front Oncol. 2020;10:568557.

    Article  PubMed  PubMed Central  Google Scholar 

  284. Chen L, Huang L, Gu Y, Cang W, Sun P, Xiang Y. Lactate-lactylation hands between metabolic reprogramming and immunosuppression. Int J Mol Sci. 2022;23(19):11943.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  285. Zhang M, Di Martino JS, Bowman RL, Campbell NR, Baksh SC, Simon-Vermot T, et al. Adipocyte-derived lipids mediate melanoma progression via FATP proteins. Cancer Discov. 2018;8(8):1006–25.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  286. Dogra S, Hannafon BN. Breast cancer microenvironment cross talk through extracellular vesicle RNAs. Am J Pathol. 2021;191(8):1330–41.

    Article  CAS  PubMed  Google Scholar 

  287. Li X, Shen Y, Zhang L, Guo X, Wu J. Understanding initiation and progression of hepatocellular carcinoma through single cell sequencing. Biochim Biophys Acta. 2022;1877(3):188720.

    CAS  Google Scholar 

  288. Han W, Wang S, Qi Y, Wu F, Tian N, Qiang B, et al. Targeting HOTAIRM1 ameliorates glioblastoma by disrupting mitochondrial oxidative phosphorylation and serine metabolism. iScience. 2022;25(8):104823.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  289. Yuan Y, Li H, Pu W, Chen L, Guo D, Jiang H, et al. Cancer metabolism and tumor microenvironment: fostering each other? Sci China Life Sci. 2022;65(2):236–79.

    Article  CAS  PubMed  Google Scholar 

  290. Dai Y, Liu Y, Li J, Jin M, Yang H, Huang G. Shikonin inhibited glycolysis and sensitized cisplatin treatment in non-small cell lung cancer cells via the exosomal pyruvate kinase M2 pathway. Bioengineered. 2022;13(5):13906–18.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  291. Wang Y, Hao F, Nan Y, Qu L, Na W, Jia C, et al. PKM2 inhibitor Shikonin overcomes the cisplatin resistance in bladder cancer by inducing necroptosis. Int J Biol Sci. 2018;14(13):1883–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  292. Wang X, Zhang H, Yang H, Bai M, Ning T, Deng T, et al. Exosome-delivered circRNA promotes glycolysis to induce chemoresistance through the miR-122-PKM2 axis in colorectal cancer. Mol Oncol. 2020;14(3):539–55.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  293. Lv B, Zhu W, Feng C. circCCT3Coptisine blocks secretion of exosomal from cancer-associated fibroblasts to reprogram glucose metabolism in hepatocellular carcinoma. DNA and Cell Biol. 2020;39:2281.

    Article  CAS  Google Scholar 

  294. Aslan C, Maralbashi S, Kahroba H, Asadi M, Soltani-Zangbar M, Javadian M, et al. Docosahexaenoic acid (DHA) inhibits pro-angiogenic effects of breast cancer cells via down-regulating cellular and exosomal expression of angiogenic genes and microRNAs. Life Sci. 2020;258:118094.

    Article  CAS  PubMed  Google Scholar 

  295. Zhan Q, Yi K, Cui X, Li X, Yang S, Wang Q, et al. Blood exosomes-based targeted delivery of cPLA2 siRNA and metformin to modulate glioblastoma energy metabolism for tailoring personalized therapy. Neuro-Oncology. 1871;2022:24.

    Google Scholar 

  296. Zhu D, Zhang T, Li Y, Huang C, Suo M, Xia L, et al. Tumor-derived exosomes co-delivering aggregation-induced emission luminogens and proton pump inhibitors for tumor glutamine starvation therapy and enhanced type-I photodynamic therapy. Biomaterials. 2022;283:121462.

    Article  CAS  PubMed  Google Scholar 

  297. Chen M, Liu H, Li Z, Ming A, Chen H. Mechanism of PKM2 affecting cancer immunity and metabolism in tumor microenvironment. J Cancer. 2021;12(12):3566–74.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  298. Wei Y, Wang D, Jin F, Bian Z, Li L, Liang H, et al. Pyruvate kinase type M2 promotes tumour cell exosome release via phosphorylating synaptosome-associated protein 23. Nat Commun. 2017;8:14041.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  299. Tan Y, Luo X, Lv W, Hu W, Zhao C, Xiong M, et al. Tumor-derived exosomal components: the multifaceted roles and mechanisms in breast cancer metastasis. Cell Death Dis. 2021;12(6):547.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  300. Yang Z, Wang D, Zhang C, Liu H, Hao M, Kan S, et al. The applications of gold nanoparticles in the diagnosis and treatment of gastrointestinal cancer. Front Oncol. 2021;11:819329.

    Article  PubMed  Google Scholar 

  301. Fan S, Kroeger B, Marie P, Bridges E, Mason J, McCormick K, et al. Glutamine deprivation alters the origin and function of cancer cell exosomes. EMBO J. 2020;39(16):e103009.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  302. Zhong J, Xia B, Shan S, Zheng A, Zhang S, Chen J, et al. High-quality milk exosomes as oral drug delivery system. Biomaterials. 2021;277:121126.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We wish to thank lab members for their valuable and enthusiastic work and scientific discussion.

Funding

This work was supported in part by grants from the following sources: the National Natural Science Foundation of China (82,203,233, 82,202,966, 82,173,142, 81,972,636, 81,872,281), the Natural Science Foundation of Hunan Province (2022JJ80078, 2020JJ5336), the Research Project of Health Commission of Hunan Province (202,203,034,978, 202,202,055,318, 202,109,031,837, 202,109,032,010, 20,201,020), Key Research and Development Program of Hunan Province (2022SK2051), Hunan Provincial Science and Technology Department (2020TP1018), the Changsha Science and Technology Board (kh2201054), Ascend Foundation of National cancer center (NCC201909B06), and by Hunan Cancer Hospital Climb Plan (ZX2020001-3, YF2020002).

Author information

Authors and Affiliations

Authors

Contributions

ST, YY and YW drafted the manuscript and prepared the figures. YH, LH, RY, ZH, YT, LL, YL, LO, JL, QP, XJ, XX, LX, MP, NW and YT helped in collecting the related literatures and participated in discussion. DC, QL and YZ designed the review and revised the manuscript. All authors read and approved the final manuscript.

Corresponding authors

Correspondence to Deliang Cao, Qianjin Liao or Yujuan Zhou.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

Not applicable.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Tan, S., Yang, Y., Yang, W. et al. Exosomal cargos-mediated metabolic reprogramming in tumor microenvironment. J Exp Clin Cancer Res 42, 59 (2023). https://doi.org/10.1186/s13046-023-02634-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13046-023-02634-z

Keywords