Role of the tumor microenvironment in cancer hallmarks and targeted therapy (Review)

  • Authors:
    • Rayan Naser
    • Isabelle Fakhoury
    • Adam El-Fouani
    • Ralph Abi-Habib
    • Mirvat El-Sibai
  • View Affiliations

  • Published online on: December 27, 2022     https://doi.org/10.3892/ijo.2022.5471
  • Article Number: 23
Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Genetic alterations drive tumor onset and progression. However, the cross‑talk between tumor cells and the benign components of the surrounding stroma can also promote the initiation, progression and metastasis of solid tumors. These cellular and non‑cellular stromal components form the tumor microenvironment (TME), which co‑evolves with tumor cells. Their dynamic and mutualistic interactions are currently considered to be among the distinctive hallmarks of cancer. Biochemical and physical cues from the TME serve an essential role in regulating tumor onset and progression. They are also associated with resistance to treatment and poor prognosis in patients with cancer. Therefore, a deep understanding of the TME is vital for developing potent anticancer therapeutics and improving patient outcomes. The present review aims to review the biology of both cellular and non‑cellular constituents of the TME and novel findings regarding their contribution to core as well as emerging cancer hallmarks. The present review also describes key TME markers that are either targeted in interventional clinical trials or serve as promising potential anticancer therapies. Understanding TME components and their intercellular interactions is key toward identifying the mechanisms of progression and treatment resistance. Such understanding is of utmost significance for personalized and effective cancer therapy strategies.

Introduction

Cancer develops from genetically altered cells with a high proliferation rate and the ability to disseminate from a primary location to invade distant sites (1). In the past, scientists assumed that cancer progression and invasiveness were solely determined by factors within tumor cells (1,2). However, focus is now put on cancer-supporting components, which have been demonstrated to aid tumor cells in manifesting the disease (3,4). It is now widely established that the tumor microenvironment (TME) components contribute to different cancer hallmarks and are thus recognized as possible cancer therapy targets (2,57). These components include cells of the stroma [cancer-associated fibroblasts (CAFs), endothelial cells (ECs), pericytes and immune cells] in addition to non-cellular components, such as the extracellular matrix (ECM), extracellular vesicles (EVs) or exosomes, and the microbiome, collectively forming the TME (5,8,9).

Oxygen levels, metabolites, nutrients and pH have also been acknowledged as factors that may be controlled by the TME (10,11). The immunosuppressive and metabolically stressed nature of the TME serves an instrumental role in exacerbating the aggressiveness of cancer cells (11). For instance, interactions between tumor and stromal components may result in additional modifications of the TME cells, ECM remodeling and angiogenesis, thus leading to metastasis (12).

Cross-talk between cancer cells and TME components may also decrease the efficacy of antitumor treatments, contributing to drug resistance (13). Accordingly, an improved understanding of the biological and chemical nature of the TME paves the way for the development of therapeutic strategies for more efficient targeted cancer therapy. The present review aims to discuss TME components and their molecular features, and how they modulate cancer hallmarks. It also reviews key factors of the TME for targeted cancer treatment, with a focus on current TME pathways and mediators targeted in interventional clinical trials.

Composition of the TME

TME refers to all non-cancer cellular components surrounding tumor cells, and non-cellular components exerting tumor-supporting roles (2,6). Stromal cells of the TME include CAFs, ECs lining the blood vessels and immune cells (Fig. 1) that are recruited by cancer cells from neighboring tissue stroma (5). Interactions of TME stromal cells with cancer cells create a protective environment that promotes tumor growth in both cases (2,5). Tumor-associated stromal cells not only physically support cancer cells but also secrete growth factors, cytokines, chemokines and ECM proteins with tumor-promoting properties (14,15). In addition to stromal cells, scientists have identified the ECM, the microbiome and cell messengers referred to as EVs as non-cancerous constituents of the TME (10,1618).

Stromal cells
CAFs

Fibroblasts are the prevalent cell type in connective tissue stroma and the primary source for ECM and basement membrane proteins (15,19). Most fibroblasts within tumors differentiate into CAFs (15,20). CAF activation is driven by different stimuli, such as inflammation (Fig. 1), ECM stiffness and other physiological stresses (19,21). CAFs are a highly heterogeneous cell population and, to the best of our knowledge, their origin remains unclear (15). While it was earlier hypothesized that most CAFs originate from local fibroblasts that are activated and reprogrammed to support tumor growth (22), some groups have demonstrated that CAFs originate from mesenchymal stem cells, specifically bone marrow-derived stem cells located in the bones (2325). Others attribute their origin to the human adipose tissue-derived stem cells found in the adipose tissues (2628). This emphasizes the remarkable plasticity of cancer, enabling it to employ different sources to promote growth and progression (27).

CAFs differ from normal fibroblasts both functionally and phenotypically (15). Generally, CAFs express an array of different proteins, such as vimentin, α-smooth muscle actin (αSMA), platelet-derived growth factor receptor α/β (PDGFR α/β) and fibroblast activation protein (FAP) (15). However, studies have identified CAF subtypes lacking the expression of these markers (29,30). Accordingly, stromal cells that are negative for epithelial, endothelial and leukocyte markers having elongated morphology might be considered to be CAFs (19). Using transcriptomic analysis, a number of groups have identified different subpopulations of CAFs (28,31). For instance, in pancreatic ductal adenocarcinoma (PDAC), three CAF subtypes have been identified: i) Myofibrotic CAFs showing myofibroblastic features with high expression of αSMA and other contractile proteins and low IL-6 expression; ii) inflammatory CAFs (iCAFs) with low αSMA expression and high expression of cytokines involved in inflammation (such as IL-6); and iii) antigen-presenting CAFs highly expressing major histocompatibility complex II (MHC II) family genes (3033). Tumor-supporting CAFs secrete MMPs, cytokines, chemokines and angiogenic factors that can stimulate the proliferation of tumor cells and enhance angiogenesis (34). CAFs also alter ECM signaling and stiffness by upregulating ECM components, such as collagen type I and III, and fibronectin (30). Accumulation of fibrous connective tissues is referred to as desmoplasia and is associated with increased hypoxia, neoangiogenesis and drug resistance (30,35).

ECs

Given their role in angiogenesis, ECs form the inner lining of the blood vessels and remain the most extensively studied cells of the TME (36). Typically, blood vessels enable the exchange of oxygen, nutrients, wastes and immune cells between the circulatory system and body tissues (36). Due to the increased metabolic and nutritional requirements of tumor cells, ECs branch from pre-existing vasculature to form new blood vessels (37). Newly formed vessels are structurally and functionally abnormal because of their leaky nature and dissimilar chaotic branching that increase the interstitial fluid pressure rendering a hypoxic and acidic environment (38). Tumor vascularization, caused by hypoxia, involves vascular ECs and other TME cell types, including pericytes and bone marrow-derived precursor cells (39,40). The pro- and anti-angiogenic molecules secreted by cells determine the transformation of normal angiogenic processes to tumor angiogenesis (37,41). While thrombospondin-1 and Endostatin are major anti-angiogenic factors (42), VEGF-A secreted by cancer cells can stimulate the formation of new vasculature that, in turn, supplies tumor cells (Fig. 1) (37). In addition, platelet-derived growth factor β (PDGFβ) recruits pericytes to the tumor vasculature aiding blood vessel formation and maturation (43).

Infiltrating immune cells (IICs)

Tumor cells and CAFs secrete chemoattractant factors that recruit various immune cells to their niche (6). TME IICs include immunosuppressive and antitumor immune cells of myeloid lineages, such as macrophages, myeloid-derived suppressor cells (MDSCs), neutrophils, mast cells and dendritic cells (DCs), and lymphoid lineages, such as B and T lymphocytes, and natural killer cells (NKs) (14,44).

Activated macrophages can be polarized into two main subtypes: Pro-inflammatory M1 and anti-inflammatory M2 (45). Tumor-recruited macrophages infiltrating the TME constitute the tumor-associated macrophages (TAMs), the most abundant immune cells of the TME (46,47). Data indicate that TAMs are mainly of the M2 subtype, and thus, are tumor-promoting (45,48). TAMs secrete cytokines and soluble factors that contribute to tumor progression by influencing angiogenesis, cell migration, invasion and metastasis (2). The presence of TAMs is associated with poor prognosis in most cancer types (49).

MDSCs represent a unique category of immunosuppressive myeloid cells that are abundant in the TME (50,51). Chronic inflammation in cancer disturbs normal myelopoiesis, and thus, differentiation and maturation of immature myeloid cells (IMCs) are impaired (52). This disturbance drives the generation of MDSCs from IMCs (52). MDSC are subdivided in two main subsets according to their origin and phenotypical and morphological characteristics. These subsets are the granulocytic polymorphonuclear neutrophils-MDSCs and monocytic MDSCs (53).

Neutrophils, also referred to as tumor-associated neutrophils, possess immunosuppressive activity (54). Neutrophils are also polarized into two subsets: Antitumor N1 and tumor-promoting N2 (54). IICs include mast cells capable of releasing soluble factors that enhance EC proliferation and promote tumor angiogenesis (50). DCs are antigen-presenting cells (APCs) capable of producing pro-inflammatory cytokines and chemokines and promoting T cell stimulation (50,55). However, DCs in the tumor exhibit abnormal antigen-presenting capabilities, and thus, are dysfunctional (56).

Among the lymphoid lineage of IICs, different types of T cell populations infiltrate the TME (44). Cytotoxic T cells can eliminate malignant cells, and thus, are associated with a good cancer prognosis (44,57). This is the case of the antigen recognizing cytotoxic memory T cells that are positive for CD8 and produce IL-2 and IFNγ (2). CD4+ T helper (Th) cells are divided into different subtypes: The pro-inflammatory Th1 lineage, anti-inflammatory Th2 cells and the immunosuppressive regulatory T (Treg) cells (2). The ratio of Th1 to Th2 cells in cancer is associated with tumor stage and grade (50,58,59). B cells are also present at the invasive borders of tumors and in the lymph nodes and lymphoid structures neighboring the TME (2). In breast and ovarian cancer, the presence of B cells is associated with a good prognosis (60). On the other hand, the presence of immunosuppressive regulatory B cells is associated with skin cancer and may promote lung metastasis, suggesting a type-specific effect of B cells on cancer (61). Finally, NKs are cytotoxic lymphocytes capable of killing tumor cells without antigen presentation (50). NKs control tumor growth by providing innate immunity to the sites of transformed tumor cells and inducing cytotoxicity (62,63).

Overall, TME stromal cells, namely CAFs, ECs and IICs, and their secretome contribute to the growth and development of tumors (14,15). Notably, the complexity of interactions between cancer and TME cells demonstrates remarkable tumor mass heterogeneity (64). In their review, Koppensteiner et al (65) discussed how negative anticancer immune responses may result from the interactions between CAFs and T cells. On the other hand, Mun et al (66) reviewed the positive and negative relationships between immune and stromal cells of the TME. Despite the advancement in technologies capable of studying the TME at the single-cell level (67), a detailed understanding of all tumor-TME connections remains largely lacking. Therefore, anticancer strategies that only target one cell population are inadequate, and need to be fully updated in line with such rapid discoveries in TME biology.

ECM

The ECM is a dynamic network of interconnected macromolecules in which the cells reside (10,68). It comprises minerals, an array of extracellular proteins, glycosaminoglycans, and other proteoglycans (PGs) and polysaccharides (10,68). The main components include collagens, elastins, fibronectins, laminins, hyaluronic acid (HA), heparan sulfate, chondroitin sulfate and keratan sulfate (10,68). This intricately organized structure forms a supportive substrate that serves as a biological scaffold for surrounding cells and as an anchor for cell attachment to the ECM (at focal adhesions and hemidesmosomes) (6971). The ECM also regulates cell-cell and cell-matrix bidirectional signal transduction, including transport and mechano-transduction (16,72). This is partly due to the ECM being a reservoir for EVs and soluble bioactive effectors, such as cytokines, growth factors, chemokines and enzymes (7375). ECM components and ECM-associated factors collectively make up the ECM ‘matrisome’, which is responsible for regulating transport, proliferation, motility, survival, homeostasis and other fundamental cellular mechanisms (7679).

The ECM is present in all tissues and organs in the body, including tumors and the TME (68). Its organization is both cell-specific and tissue-specific (80). The ECM composition within tumors is heterogeneous and accounts for up to 60% of the tumor mass (10). Both cancer cells and stromal cells contribute to the production of the tumor ECM (77,81). However, CAFs remain the primary source of ECM in the TME (10,20,74,82). Cancer ECM differs from normal tissue ECM in composition, organization, density, and physical and biochemical properties (68). These differences are also noted across tumors of different metastatic potentials (83,84). For instance, primary and pre-metastatic cancers increase ECM production (74,8588). This TME fibrotic response, clinically termed desmoplasia, results in a substantial accumulation of collagens, fibronectins and PGs in benign and malignant tumors (89,90). Collagen and collagen-processing enzymes, laminins, integrins, MMPs and HA are among the most enriched ECM proteins in tumors (10,62,78,81,91,92).

On the other hand, the shift between low and high molecular weight PGs in different solid tumors illustrates the association between the composition of the ECM and cancer grade (93). Similarly, the tumor environment favors the increase in collagen type I, III or V at the expense of collagen type IV in breast, ovarian, lung and ductal carcinoma (9496).

Finally, the crosslinking of collagen, and other fibrillary proteins, such as elastins, renders the ECM denser and stiffer (10,74,91,9799). Changes in the ECM may be induced by proteases (MMPs and cathepsins) or nonproteolytic enzymes (heparanases and hyaluronidases) secreted by tumor and stromal cells, by oxygen free radicals produced by IICs, or as a response to hypoxia and acidosis (16,77,100).

The upsurge in the production of ECM components with altered properties, in turn, reduces the diffusion of nutrients and metabolites, and modulates cytokine secretion (79,83,101). This creates a hypoxic tumor-promoting environment capable of stimulating proliferation, tumor growth, epithelial-to-mesenchymal transition (EMT), aggressiveness, resistance to cell death, evasion from the immune system, and invasion and tumor dissemination, among others (10,97,102). Overall, this highlights the need for potent ECM-targeting therapies.

EVs

EVs are cell messengers, which mediate the signaling cross-talk between a cell and its environment (103105). Cancerous and non-cancerous cellular constituents of the TME, including the microbiome, communicate with one another by secreting soluble factors and/or releasing EVs (75,103,105110). Therefore, EVs are an integral and functional non-cellular component of the TME (75,103,111,112). Briefly, EVs are membrane-enclosed particles subdivided into exosomes, microvesicles and oncosomes depending on their size, biogenesis, function, etc. (105,113). Exosomes are intraluminal vesicles destined for exocytosis (105). They exhibit a classic dish or saucer-like morphology with diameters ranging between 30 and 100 nm (17,114,115). As the name suggests, intraluminal vesicles are formed by the inward budding of endosomal membranes inside the lumen of endosomes (17,114116). Unlike exosomes, microvesicles are the products of the outward budding and fission of the plasma membrane (105,117,118). Their diameters range between 100 and 1,000 nm (105). By contrast, oncosomes are cancer-specific large EVs with diameters ranging between 1 and 10 µm (105,118,119). They shed off the ‘non-apoptotic membrane blebs’ of amoeboid cancer cells (105,120). All aforementioned biogenesis processes simultaneously result in the packaging of various cytosolic materials inside the EVs (105). Therefore, the cargo of EVs can comprise lipids, proteins and nucleic acids, such as DNA, mRNA, microRNA (miRNA), long non-coding RNA (lncRNA) and circular RNA (circRNA) (17,114116,121,122).

Studies have demonstrated that all types of cells secrete EVs (75,103106,117,123). However, the data suggest that cancer cells secrete more EVs than normal cells and that the load of EVs can increase with cancer grade and aggressiveness (124126). For instance, hypoxic solid tumors secrete more EVs than non-hypoxic tumors and control the composition of the released EVs (126). The differential expression of wild-type or mutant p53 also impacts the secretion, size, and RNA and protein load of cancer-derived EVs (127,128). This is in line with reports demonstrating that the molecular composition of EVs and their effect on cancer hallmarks and response to chemotherapy depends primarily on the origin of the secretory cells (104,110,121,126,129133). Furthermore, emerging evidence has revealed that EVs derived from different cell types exhibit distinct content profiles (131). Researchers could even differentiate between EVs originating from various subtypes of the same lineage, namely between the exosomes of lymph node metastasis-derived LNCaP (lymph node carcinoma) and VCaP (vertebral) prostate cancer cell lines (134).

EV cargo commonly includes type-specific and or stage-specific cancer biomarkers (117). For instance, EVs isolated from patients with ovarian cancer express distinct protein and miRNA sets compared with those found in cancer-free individuals (114). Similarly, specific RNA classes are particularly abundant in EVs of patients with triple-negative breast cancer compared with those with hormone receptor-positive breast cancer (124,135). Several other non-coding RNAs have been demonstrated to serve a role in tumor development (136). The long non-coding lymph enhancer-binding factor 1-antisense RNA1 has been found to act as a tumor promoter in a number of malignant tumors; however, it acts as a tumor suppressor in myeloid cancer (137). Furthermore, circ0021205 is a non-coding circRNA that promotes cancer progression in cholangiocarcinoma and non-small cell lung cancer (NSCLC) (138). In patients with colorectal cancer (CRC), upregulated miRNA-7062-5p inhibits G protein-coupled receptor 65, thus promoting osteoclast genesis during bone metastasis (139). Long intergenic non-coding RNA (LINC)02257 is a survival-associated enhancer RNA serving important immunotherapy roles in a number of cancer types (140). In lung adenocarcinoma, the LINC00987/A2M axis acts as an effective tumor suppressor, as well as a biomarker for the evaluation of the tumor immune microenvironment or the prognostic and therapeutic potential (141).

Further observations have revealed that the expression of the polymerase I and the transcript release factor glioma biomarker in EVs is positively associated with tumor grade (142). The clinical expression of programmed cell death ligand-1 (PD-L1) in EVs is associated with diverse cancer types, including melanoma and colon cancer (143145). Finally, researchers have reported that both surgery and radiation treatments change the composition of EVs, thus demonstrating the close association between tumors and their TME (132,142). These data collectively highlight the growing interest in EVs as promising targets cancer for diagnosis, prognosis and treatment.

Microbiome

The microbiome is a component of the TME, which has become a subject of interest in cancer research recently (18,146). By definition, the microbiome represents ‘the characteristic microbial community occupying a reasonably well-defined habitat, which has distinct physico-chemical properties’ (146). In principle, only the microbial community present in or around the tumor tissues can strictly be labeled as the microbiome of the TME (18). However, microbes at distant sites from tumors (such as the gut) have also emerged as critical modulators of cancer onset and progression (147153). Therefore, this section reviews the composition and role of all microbes with direct or indirect effects on cancer to overcome the shortcomings of the TME-centric view and highlight the importance of the different layers of cancer environment beyond the immediate spatial boundaries of tumors (i.e., at the level of what is now known as the tumor organismal environment) (154). The microbiome is: i) A fingerprint for tumors; ii) a major factor in the pathology of the disease; and iii) a diagnostic tool to predict the response to treatment of patients (150,151,155164). Researchers have reported substantial differences in the microbiome composition of normal tissues compared with tumor tissues (162,163,165173). These observations have revealed tumor-type specific, tumor subtype-specific and grade-specific bacteria spanning several major phyla (162,163,165173). Furthermore, the data suggest that the majority of bacteria of the tumor microbiome of different solid tumors are intracellular (165).

The hypoxic nature of tumors contributes to the abundance of anaerobic bacteria, which are unable to survive in an oxygen-rich environment (174). The anaerobic Fusobacterium genus is particularly abundant in various carcinomas, including oral, colorectal and bladder cancer (168,171,173175). For instance, several species of Fusobacterium, namely, F. nucleatum, F. mortiferum and F. necrophorum, have been identified in metastatic colon cancer tissues, while F. nucleatum and F. periodonticum are a signature of oral cancers (174,175). Further highlighting the tumor-specificity, the Bacteroides depleted in colon cancers were, by contrast, profuse in rectal tumors (175,176). In addition, rectal tissue sample analysis revealed a distinct microbiome in cancer tissues compared with normal tissues (176,177). Phascolarctobacterium, Parabacteroides, Desulfovibrio and Odoribacter species are abundant in tumors, whereas Pseudomonas, Escherichia, Acinetobacter, Lactobacillus and Bacillus species are primarily found in healthy tissues (176). Similarly, the microbiome of the breast cancer TME indicated the enrichment in specific microbes (Actinobacteria, Listeria spp, Haemophilus influenzae, Anaerococcus, Caulobacter and Streptococcus) and the depletion of others (Propionibacterium and Staphylococcus) (164,177). Interestingly, in lung cancer, the genera Streptococcus and Staphylococcus exhibit the same expression trend as that observed in breast cancer (163). Specifically, Streptococcus and Neisseria genera thrive in cancer, unlike Staphylococcus and Dialister, which favor normal tissues (163). Different microbes, including Fusobacterium, Streptococcus and Bacteroides, are also differentially expressed in bladder cancer, prostate cancer and other cancer types (171,172,178).

Finally, mycoplasma species, such as the Mycoplasma hyorhinis, also exhibit disparate manifestation in tumor tissues compared with normal tissues, and are associated with tumor-promoting properties, resistance to treatment and poor prognosis (179186). These reports emphasize the importance of identifying and characterizing the different genera of the tumor microbiome to determine potential biomarkers and therapeutic targets (146,154,163,176,177).

Cancer hallmark capabilities of TME stromal cells

In 2000, Hanahan and Weinberg (187) proposed a conceptual categorization to organize the distinctive key traits of cancer. They noticed that all cancer cells acquire six fundamental functional capabilities, which they named ‘The Hallmarks of Cancer’ (187). These fundamental hallmarks are: Sustaining proliferative signaling, evading growth suppressors, resisting cell death, enabling replicative immortality, inducing angiogenesis, and activating invasion and metastasis (Fig. 2). Due to the remarkable progress in cancer research, Hanahan and Weinberg (187) incorporated additional ‘emerging cancer hallmarks’, which are: Evading immune destruction and reprogramming energy metabolism (Fig. 3). In addition, ‘enabling characteristics’ that facilitate acquisition of fundamental and emerging hallmarks, were also introduced. These are: Genome instability and mutation and tumor-promoting inflammation (Fig. 3) (4,188). TME stroma markedly contribute to cancer hallmarks (6). The reciprocal communications between TME components and cancerous cells mediate cancer development (5). This section reviews the various mechanisms by which TME stromal cells influence cancer hallmarks and highlights the complexity of cancer targeting.

Sustained proliferative signaling

The mutational capabilities of cancer cells have been recognized as the core factor for sustaining cancer proliferative signaling (187). Stromal cells serve an important role in augmenting oncogenic mutations, and thus, the hyperproliferation of cancer cells by driving mitogenic signals (Fig. 2) (6,7,187).

Different CAF subtypes exhibit diverse functions and affect multiple cancer hallmark capabilities (30). CAFs secrete mitogenic epithelial growth factors (EGF), fibroblast growth factors (FGF), hepatocyte growth factors (HGF) and other signaling proteins that drive cancer cell proliferation (189191). HGF secretion results in the activation of mesenchymal epithelial transition factor (a HGF receptor), and thus, the activation of the MAPK and PI3K/AKT survival signaling pathways (192). These pathways are also activated by CAF-secreted vascular cell adhesion molecule 1 (VCAM1) and promote the proliferation of lung cancer cells (193). Furthermore, leptin, a cytokine secreted by CAFs, binds to its receptor, activates MAPK/ERK1/2 and PI3K/AKT signaling pathways, and promotes proliferation of cancer cells in NSCLC (194).

Angiogenic vascular cells (AVCs) also directly support cancer hyperproliferation. Experimentally stimulating angiogenesis results in increased proliferation of cancer cells (195,196). These ECs secrete growth-promoting factors that influence multiple hallmarks, including proliferation, invasion and metastasis (further described subsequently) (197). Similarly, IICs stimulate neoplastic cell proliferation by secreting mitogenic growth factors, such as TNF-α, ILs, chemokines, heparins and histamine, in addition to EGF, FGF and TGF-β [reviewed in (198)]. Additionally, IICs secrete metallo- and serine proteases that cleave and modify the ECM leading to chronic paracrine and juxtacrine mitogenic activity sustaining cell proliferation (199).

Evading growth suppressors

Adhesion molecules at cell-cell and cell-ECM connections transmit extrinsic growth-suppressing signals to cancer cell cycle machinery (5,187). As aforementioned, disruption of adhesion molecules is induced by IIC-secreted metallo- and serine proteases and heparinase, which cleave and modify the ECM (10). ECM modifications disrupt the transmission of antigrowth signals and the formation of growth-suppressing adhesion complexes (200202).

Notably, fibroblasts naturally exhibit extrinsic growth-suppressing capabilities to maintain epithelial homeostasis (15). In the TME, CAFs secrete high amounts TGF-β and ECM components [reviewed in (203)], thus stimulating mechanical remodeling of the ECM. Therefore, it was hypothesized that CAFs may acquire a ‘loss-of-function’ phenotype as they are reprogrammed to sustain cancer hallmarks (5). TME components can also affect tumor growth-suppressing signals by regulating cell cycle check points (204). The TME of renal cell carcinoma (RCC) exhibits FGF-dependent degradation of p27Kip1, a cyclin-dependent kinase inhibitor (204). This leads to enhanced tumor cell proliferation (Fig. 2) (204).

Resisting cell death

Cancer cells foster an intrinsic ability to resist cell death programs, mainly apoptosis (187). Stromal cells of the TME confer an additional protective mechanism for cancer cells to resist cell death and targeted cytotoxic therapy (Fig. 2) (6). CAFs mediate cancer cell survival by secreting survival factors [insulin-like growth factor 1 (IGF-1) and insulin-like growth factor 2 (IGF-2)] (15). CAFs also form neoplastic ECM that selectively transmits survival signals and promotes epithelial cell migration (199).

The prominent role of CAFs in resisting cell death is demonstrated by the contribution of CAFs to chemoresistance (15). CAFs co-cultured with NSCLC cell lines have been demonstrated to resist apoptosis and enhance chemoresistance (205,206). This is achieved by the secretion of stromal cell-derived factor-1 (SDF-1) and the expression of exosomal miRNA-103a-3p, which lead to Bcl-xL upregulation and BCL2 antagonist/killer 1 downregulation, respectively (205,206). In a separate study, Sun and Chen (207) demonstrated that CAF-secreted C-C motif chemokine ligand (CCL)5 upregulated HOX transcript antisense RNA (HOTAIR) lncRNA expression. HOTAIR, in turn, inactivated the caspase-3/BCL-2 signaling pathway in these cells conferring chemotherapy resistance in NSCLC cells (207). Additionally, the chemotherapeutic drug, cisplatin, induces CAF-secreted IL-11 in lung adenocarcinoma (208). IL-11 activates the IL-11 receptor/STAT3 anti-apoptotic signaling pathway (208). The monokine induced by IFN-γ factor is a CAFs-secreted chemokine that has also been demonstrated to upregulate the expression levels of Bcl-2 and protect Tca8113tongue squamous cell carcinoma cells from heat-induced apoptosis (209).

Tumor vascularization induced by AVCs reduces apoptosis, thus increasing cancer cell proliferation (5). This phenotype is altered by the administration of vascular disrupting agents that increase cell death in treated cancer (210). Binding of IICs to cancer cells can also inhibit detachment-induced apoptosis (5). TAMs, on the other hand, express α4-integrin that binds VCAM1 expressed on breast cancer cells (211). This interaction initiates a signaling pathway that activates anti-apoptotic PI3K/AKT signaling and resists apoptosis (211). In addition, IICs secrete cytokines, leading to cell death resistance (212). For instance, TAM-secreted IL-10 is associated with elevated Bcl-2 expression via upregulation of the IL-10/STAT3/bcl-2 anti-apoptotic signaling pathway (212). This leads to increased proliferation and is associated with drug resistance of breast cancer (212).

Enabling replicative immortality

Shortening of telomeric DNA obstructs cellular replication and triggers senescence or apoptosis (5). Cancer cells, however, need to insure limitless replication as a defense mechanisms to overcome normal senescence caused by telomere shortening (4,5). Therefore, cancer cells activate telomerases that stabilize telomere length and confer replicative immortality (4,213). This critical trait occurs in >90% of cancers (213). Telomerase activation is enhanced by the upregulation of the human telomerase reverse transcription (hTERT) gene (214). The pH of the hypoxic TME can selectively induce human telomerase (215). At present, there is little evidence for the contribution of TME stromal cells to stabilizing telomeres in cancer cells (5). For instance, bone marrow-derived mesenchymal stem cells (MSCs) of multiple myeloma secrete IL-6 and macrophage inflammatory protein (MIP)-1α (216). Li et al (216) provided evidence of a positive association between IL-6 and MIP-1α secretion and the elongation of telomere length. MSCs may thus facilitate multiple myeloma development (216). Further research is required to investigate whether other TME cellular components can regulate telomerase activity and enable replicative immortality.

Inducing angiogenesis

Angiogenesis in chronic inflammation is illustrated by constitutive activation of pro-angiogenic factors (37). In tumors, it is regulated by different components of the TME, such as CAFs, ECs, different IICs and pericytes, which secrete angiogenesis-inducing factors (Fig. 2) (2,5,6,37). Myeloid cells secrete soluble mediators that impact EC survival and new vessel remodeling (50). For instance, TAMs control tumor angiogenesis by producing VEGF-A (50,217), the bioavailability of which depends on TAM-secreted MMP-9 (218). Mast cells, on the other hand, secrete VEGF, histamine and heparin, thus regulating tumor angiogenesis (50). Mast cells also secrete proteases, such as MMP-9 and tryptase, which in turn activate pro-angiogenic signaling pathways (218220). In addition to the secretion of pro-angiogenic factors, CAFs synthesize ECM proteins that sequester angiogenic growth factors and ECM-degrading enzymes (15). For example, in hepatocellular carcinoma, CAFs secrete VEGF, regulating the enhancer of zeste homolog-2/vasohibin 1 pathway, thus promoting angiogenesis (221). CAFs also regulate tumor angiogenesis by secreting chemoattractants for myeloid cells (5,222). Additionally, pericytes promote angiogenesis in glioma by expressing periostin, which regulates both growth and branching of blood vessels (223).

Activating invasion and metastasis

A key feature of cancer cells is the ability to spread throughout the body by invasion and metastasis (187). Cancer cells and tumor stromal cells can mediate local invasive growth or seeding metastases at distant sites (2,6,7). Tumor vasculature upregulation of VEGF loosens tight junctions between ECs and reduces pericyte coverage (62,224). This impairs vascular integrity and facilitates cancer cell intravasation into circulation (62). Hypoxia, induced by hypoxia-inducible factors, then triggers tumor dissemination and metastasis (224,225). Furthermore, pericytes in the TME activate TGF-β receptors (226). The subsequent TGF-β response initiates an autocrine activation loop (226). Analysis of the secretome of these activated pericytes has revealed upregulation of IGF-binding protein-3, a key paracrine factor that has been demonstrated to promote cancer cell migration and invasion (226). Additionally, proteases secreted by TAMs and mast cells remodel ECM components, promoting tissue invasion and dissemination (227,228). Soluble factors secreted by IICs also contribute to this hallmark. For instance, TNF-α, secreted by IICs, activates downstream JNK and NF-κB signaling cascades, ultimately enhancing MMP-2 and MMP-9 activity (229). Equally importantly, IICs mediate cancer metastasis by inhibiting the expression of metastasis suppressor genes. For example, IICs inhibit maspin, a serine protease inhibitor, which normally acts as a tumor suppressor by increasing cell adhesion to extracellular matrix. Thus, maspin inhibition negatively regulates tumor migration and invasion (230,231).

Platelets also exhibit invasion and metastasis-promoting functions by physically associating with cancer cells, inducing EMT, enabling extravasation and forming secondary tumors at metastatic sites (232). Different components within the ECM might also initiate or enhance EMT-like processes (10). Collagen reorganization, PG expression and protease-mediated ECM macromolecule degradation affect cell invasion and metastasis (16). Finally, CAFs are also implicated in activating invasion and metastasis (15). CAF-derived effectors, such as HGF and TGF-β, trigger/activate c-Met signaling and EMT, respectively, mediating tumor invasion and metastasis (233). In breast cancer cells, IGF-1 and CXCL12 secreted by CAFs stimulate cancer metastasis to the bone (234). ECM proteins and remodeling enzymes produced by CAFs are also considered to support cancer invasion by modifying the structure and function of the ECM (7). One study revealed that cancer cells circulate in the blood alongside CAFs derived from the primary tumor (235). It has also been suggested that CAF-derived signals may control organ-specific metastasis of breast tumors (236). Therefore, CAFs are promising targets in pre-clinical therapeutic strategies in patients with breast cancer (236).

Evading immune destruction

Effective destruction of cancer cells necessitates an influx of immune cells, including Treg cells, NKs and NK T cells (5,187). However, tumor vasculature is considered to attenuate the influx of immune cells, rendering them incapable of killing cancer cells (5). This is partly mediated by high endothelial venules (HEVs), which typically support the homeostatic trafficking of immune cells during routine immune surveillance (43). Absence of HEVs in tumor vasculature allows cancer cells to evade immune destruction (237239).

In the TME stroma, ECs cross-present tumor antigens and stimulate the development of a tolerizing, hence immunosuppressive, environment (240). CAFs are also essential factors that allow the tumor to evade immune destruction (Fig. 3). CAF-derived interleukins (IL-4 and IL-6) and chemokines (TGF-β and CXCL8) recruit M2 macrophages and inhibit cytotoxic T lymphocytes (CTLs) and NK cells (240,241). Among the intricate tumor-promoting roles of CAFs is the ability to act as APCs to T cells (20). In addition to ECs and CAFs, IICs serve an important role in enabling cancer to avoid immune destruction (44). IICs prompt immunosuppressive activity that blocks the antitumor effect of CTLs and NK cells (14). Among these IICs are TAMs, which lack cytotoxic activity, and release immunosuppressive factors and suppress CD8+ T cell proliferation (242244). CCL22 secreted by TAMs recruits Treg cells, which enable cancer cells to evade immune destruction (245). Treg cells can also be recruited by other cytokines, such as CCL2 and TGF-β, secreted by mast cells and other immunosuppressive IICs (50,246). Treg cells can suppress B cells, NK/T cells, CD4+/CD8+ T cells, monocytes and DCs (247). Treg cells suppression can be direct through cell contact or immunosuppressive soluble mediators or indirect by suppressing APCs (14). In particular, Treg-induced inhibition of CD4+ T cells is mediated by inhibition of receptor-induced calcium, nuclear factor of activated T-cells and NFκB signaling (50,247). Tumors expressing high levels of immunosuppressive cytokines are often associated with decreased CD8+ T cell populations and poor survival (242). Reciprocal communication occurs between IICs and other TME stromal cells where M2 macrophages secrete EGF, FGF and TGF-β to support CAF survival and activation (240). Finally, TAM, MDSC and mast cell secretion of MMP9, FGF2, PDGFβ and VEGF prompts EC survival and angiogenesis (240).

Reprogramming energy metabolism

Metabolic differences between normal and cancer cells have long been observed (4,248). However, scientists have revealed that tumors are metabolically heterogeneous and can be grouped into different metabolic phenotypes (249251). This heterogeneity is likely to stem from differences in glycolytic and mitochondrial reprogramming (249253).

Advancements in physiologic magnetic resonance imaging of tumors have demonstrated that metabolic phenotypes remain flexible and can switch depending on the surrounding TME conditions and the nature of the exchanged signaling molecules (254). Therefore, the plasticity is contingent on the availability of nutrients and anaplerotic molecules (255). Consequently, interactions between cancer cells and pro-tumorigenic TMEs can reprogram tumor metabolism and fuel cancer cell proliferation (Fig. 3) (255).

Substantial evidence indicates that tumors can reshape TME metabolism and even the patients' organismal metabolism and homeostasis (256258). For instance, cancer cells can induce aerobic glycolysis in TME CAFs and stromal cells through the reverse Warburg effect (259). Tumor-derived metabolites can also exert immunosuppressive effects that block the activation and differentiation of various immune cells of the TME (260,261). In return, CAFs and other TME stromal components supply cancer cells with the nutrients required to sustain tumor metabolism and growth (260).

In summary, most findings highlight the ability of the metabolism to reshape the TME, whereby cancer cells turn the normal TME into a permissive tumor-promoting environment and a nutrient factory to be used for efficient energy production (259,262). Therefore, it is crucial to review the metabolic features of TME components and to identify critical factors behind metabolic reprogramming.

Cancer cell metabolism

Cell metabolism refers to the set of chemical reactions that sustain the normal functions of cells, including: i) Production of macromolecule building blocks; ii) energy harvesting; and iii) the elimination of metabolic wastes (248,263). Cells rely on nutrient uptake and degradation to generate ATP, the energy currency of cells (248,263). Cancer cells are fast replicating cells with high anabolic and catabolic energy requirements (4). These cells can rewire their metabolism to maintain sufficient ATP production and ensure proliferation and survival (264266). Metabolism reprogramming is achieved by a complex interplay of various signaling pathways, which can be intrinsically triggered by oncogenes or extrinsically influenced by the inhospitable TME (264,267269). One such mechanism is the significant increase in glucose uptake, typically observed in positron emission tomography scans of patients (270). Another major metabolic difference between normal and cancer cells is the fate of glucose. Normal cells harvest glucose energy through aerobic respiration, a four-stage process that combines glycolysis, pyruvate oxidation, the tricarboxylic acid (TCA) cycle and oxidative phosphorylation (OXPHOS) (248,263). In this route, glucose is oxidized into carbon dioxide and water.

On the other hand, cancer cells switch to aerobic glycolysis, which converts glucose into pyruvate and then lactate via the lactate dehydrogenase enzyme (248,265). This phenomenon, also known as the Warburg effect or the Warburg phenotype, is commonly stimulated by energy demands of rapid proliferation and/or hypoxia (251,270,271). To the best of our knowledge, it is still unclear whether the Warburg effect is a cause or consequence of carcinogenesis (272). However, this phenotype allows faster ATP production than OXPHOS and may provide a selective advantage for cells, specifically in the hypoxic TME (273,274).

Another important benefit of aerobic glycolysis is that the generated glycolytic metabolites facilitate the biosynthesis of sugars, amino acids, nucleotides and fatty acids critical for rapid cell proliferation (248). These advantages explain why high amounts of aerobic glycolysis have been detected in proliferating cells and progressive cancer types (270,275). The TCA cycle is also active in these cells, with the resulting substrates rerouted for use in de novo synthesis pathways, particularly lipogenesis (276).

Cancer cells heavily rely on the uptake and metabolism of glutamine as an alternative carbon source to glucose (276). Glutamine, the amide derivative of glutamate, is one of the most abundant nutrients in the plasma (255,276). Glutaminolysis converts glutamine to lactate and NADPH (255,276). In normal cells, glutamine is used as a nitrogen source to synthesize nucleotides and other non-essential amino acids (264). However, in cancer cells, glutamine metabolism exceeds the needs of cells for de novo proteins and nucleotides production (276). Instead, the data suggest that glutamine metabolism allows the cells to use glucose-derived carbon and TCA cycle intermediates as precursors for fatty acid synthesis (276). This is achieved by continuous replenishment of the TCA cycle intermediates (mainly oxaloacetate) through a set of five chemical reactions, which combined, form the anaplerosis process (276).

Drivers of metabolic reprogramming

Metabolic reprogramming in the TME is driven by oncogenic alterations in cancer cells, as well as by changes in the signaling of normal cells (269). Typically, these modifications impact the dynamics underlying nutrient uptake and bioenergetic gene expression (269). For instance, constitutive activation of the PI3K-AKT-mTOR signaling pathway has been directly linked to glycolysis stimulation in cancer cells and in CAFs (269,277,278).

Similarly, there is a clear association between Myc transcription factor and the expression of various metabolic genes, glycolytic enzymes, and glucose and glutamine transporters (279). Specifically, Myc activates glucose and glutamine metabolism, as well as purine, pyrimidine, fatty acid and cholesterol synthesis (269,280282). Oncogenic KRAS also triggers TME metabolic reprogramming by upregulating glycolytic enzyme activity, glutamine metabolism, and glucose and lactate transporter expression (283285). In addition, KRAS stimulates nucleotide biosynthesis by channeling glycolytic metabolite intermediates to the pentose phosphate pathway (284286). Furthermore, KRAS sustains autocrine and paracrine signaling by inducing the expression of cell surface receptors responsible for upregulating type I cytokine receptors (269,287,288). KRAS promotes micropinocytosis and autophagy processes for nutrient scavenging by cancer cells (289,290). On the other hand, loss of p53 function in cancer cells increases glycolysis, glucose transporters and lipid metabolism, among others (291294). Overall, cells of the TME have several distinct metabolic signatures that are directly associated with tumor growth and represent promising targets for cancer therapy.

Genome instability and mutation

Genome maintenance in normal cells results in a low rate of spontaneous mutations (4). Compromised check points and sensitivity to mutagenic agents increase the rate of mutation (4). The TME is characterized by hypoxia, low pH and nutrient deprivation (295). These conditions contribute to genome instability and mutations (Fig. 3) (295). For instance, hypoxia stimulates the downregulation of mutL homolog 1 and PMS1 homolog 2, mismatch repair system component, which are protein members of the mismatch repair pathway and required to rectify DNA mismatch errors (296,297). Therefore, hypoxia is a major factor inducing substantial DNA damage leading to genetic instability of solid tumors (295).

In addition to hypoxia, it has been recently shown that a novel competing network of competing endogenous RNA can regulate genomic integrity (298). Therefore, these genome instability-related lncRNAs may act as biomarkers for genetic instability, immunetherapy prognosis and therapeutic sensitivity in colon adenocarcinoma and colon cancer (298,299).

Tumor-promoting inflammation

Chronic inflammation is a major contributor to cancer, and the inflammatory response can be triggered by various factors, including pathogens, carcinogen exposure and imbalanced immune regulation (Fig. 3) (300). Immune cells can either exert an antitumor or protumor activity depending on the polarization state. For example, Th1 cells act as antitumor agents, whereas Th17 subsets of CD4+ T cells act as tumor-promoters (300,301). On the other hand, anti-inflammatory M2 macrophages and N2 neutrophils are both tumor-promoting cells that secrete cytokines, proteases and growth factors, contributing to tissue remodeling and angiogenesis, eventually leading to the conversion of cells into malignant cells and cancer formation (300,301). TAMs, an M2 subtype, produce VEGF-C and VEGF-D, which leads to peritumoral inflammation and lymphangiogenesis in human cervical cancer (302).

CAFs exert pleiotropic functions in immunomodulation mainly by secreting a range of pro-inflammatory factors, which recruit and activate IICs (191,303,304). In 2010, Erez et al (191) revealed that the CAF secretome causes tumor-promoting inflammation in a NF-κB-dependent manner. iCAFs secrete various chemokines and cytokines, such as CXC-chemokine ligand (CXCL)1, cyclooxygenase-2, IL-1, IL-6 and SDF-1, and receptors, such as IL 6 receptor α and IL-6 cytokine family signal transducer, which add to the tumor-promoting inflammatory milieu of the TME (305308). In addition to TGF-β production, CAFs secrete thymic stromal lymphopoietin, favoring Th2 cell polarization, which is associated with poor prognosis (309). Overall, TME-mediated inflammation influences tumor development, invasion, angiogenesis, metastasis and immunosuppression (191,305,306). Therefore, targeting inflammation may be a promising tool for cancer treatment.

Therefore, cancer development is mediated by TME components that contribute to major cancer hallmarks, including tumor proliferation, survival, angiogenesis, invasion and metastasis (2,66,240). Overall, these findings have prompted researchers to target TME cells for cancer treatment alone or in combination with conventional therapeutic modalities (9,67,310).

Targeting the TME for cancer therapy

Clinical anticancer therapeutic efficiency is limited due to several factors, including tumor heterogeneity and the ability of cancer cells to develop multidrug resistance (11). Another reason for this observation is that cancer cells exhibit different responses when moving from bench to bedside translational medicine (311). The cross-talk among tumor cells, stromal cells and other TME components adds to the complexity of efficient treatment (8). Increasing awareness of the role of the TME in tumor development brought about novel cancer therapy strategies targeting TME components (310). Additionally, combined therapies targeting more than one cell type or signaling pathway are also being investigated (11). The present review highlights the TME cells, signaling pathways and soluble factors that are targeted for cancer treatment. It also reviews anticancer drugs that are currently in clinical trials or show promising results for drug development.

Targeting CAFs

Membrane-bound serine protease FAP expression distinguishes tumor tissues from healthy tissues (30). Inhibitors selectively targeting FAP (FAPi) are currently in phase I and II clinical trials (http://clinicaltrials.gov). For instance, 68Ga-DOTA-FAPI is being studied for FAP-based imaging and therapy using gallium-68 (312,313). FAP is also targeted by CD40 agonist (RO7300490) or 4-1BB agonist (RO7122290; phase I/II; NCT04826003; Table I). The latter resulted in activation of T and NK cells in the first-in-human phase I study, suggesting potential antitumor activity (314). An early phase I clinical trial is investigating the effect of combining FAPi with anti-neoplastic monoclonal antibodies (NCT01722149). More studies are warranted to determine the efficiency of targeting FAP-positive CAFs and tumor cells.

Table I.

Therapeutic agents targeting cancer-associated fibroblasts and vascular endothelial cells in interventional clinical trials.

Table I.

Therapeutic agents targeting cancer-associated fibroblasts and vascular endothelial cells in interventional clinical trials.

Targeting strategyTherapeutic agentTherapeutic targetTherapeutic strategyCancer typePhaseClinical trial
Targeting fibroblastsRO7122290FAP targeted 4-1BB ligandRO7122290 in combination with immunotherapeutic cibisatamab after pretreatment with chemotherapeutic obinutuzumabMetastatic colorectal cancerPhase I/IINCT04826003
68Ga-DOTA-FAPIFAPExploring the application value of positron emission tomography molecular imaging targeting FAPOral squamous cell carcinomaEarly recruitingNCT05030597
FAP-specific re-directed T cellsFAPTransferred FAP-specific re-directed T cells are given directly in the pleural effusionMalignant pleural mesotheliomaEarly phase 1NCT01722149
Targeting vascular endothelial cellsCabozantinibVEGFR2, RET and METCabozantinib vs. placeboDifferentiated thyroid cancerPhase IIINCT03690388
BevacizumabVEGFAApatinib vs. placeboAdvanced/metastatic gastric cancerPhase III NCT01512745/NCT00970138
Oxaliplatin chemotherapy associated or not associated with the targeted therapies (anti-EGFR panitumumab and bevacizumab)Metastatic colorectal cancer restricted to the liver (OSCAR)Phase IIINCT02885753
Tamoxifen citrate or letrozole with or without bevacizumabStage IIIB or stage IV breast cancerPhase IIINCT00601900
Bevacizumab in combination with paclitaxel compared with paclitaxel plus placeboHer2-negative metastatic breast cancerPhase IIINCT01663727
Bevacizumab, pemetrexed, or a combination of bevacizumab and pemetrexed following carboplatin, paclitaxel chemotherapy and bevacizumabNon-squamous non-small cell lung cancerPhase IIINCT01107626
Chemotherapeutic paclitaxel, nab-paclitaxel, or ixabepilone with or without bevacizumabStage IIIC or stage IV breast cancerPhase IIINCT00785291
Chemotherapeutic imatinib mesylate with or without bevacizumabGastrointestinal stromal tumorPhase IIINCT00324987
Bevacizumab, everolimus (also referred to as RAD001) and lapatinibVEGFA, mTOR, and Her-1 and Her-2 receptor tyrosine kinase, respectivelyBevacizumab, everolimus (RAD001) and lapatinib as neoadjuvant docetaxel chemotherapyPrimary breast cancerPhase IIINCT00567554
QL1101 and Avastin® (also referred to as bevacizumab)VEGFQL1101 and Avastin, respectively, combined with paclitaxel and carboplatinNon-squamous non-small cell lung cancerPhase IIINCT03169335
Bevacizumab and LY01008VEGFBevacizumab and LY01008 combined with paclitaxel and carboplatinNon-small cell lung cancerPhase IIINCT03533127
Everolimus (RAD001)mTOR, HIF and VEGFAloneMetastatic renal cell carcinomaPhase IV NCT01266837/NCT01206764
Everolimus vs. placeboNeuroendocrine tumorsPhase IIINCT00510068
Belzutifan (anti-HIF-2α) vs. everolimusAdvanced renal cell carcinomaPhase IIINCT04195750
SorafenibVEGFR, Raf, PDGFR and KitDovitinib (inhibitor of FGFRs) vs. sorafenibMetastatic renal cell carcinomaPhase IIINCT01223027
Vandetanib (also referred to as ZACTIMA)VEGFR, PDGFR and EGFRZACTIMA combined with docetaxel chemotherapy compared with docetaxelNon-small cell lung cancerPhase IIINCT00312377
CediranibVEGFRsOlaparib (PARP inhibitor) with cediranib or olaparib aloneOvarian cancerPhase IIINCT03278717
BD0801VEGFBD0801 combined with paclitaxel chemotherapy vs. placebo combined with chemotherapyEpithelial ovarian cancerPhase IIINCT04908787
EndostatinVEGFR2Endostatin combined with vinorelbine and cisplatin (NP) as neoadjuvant therapyNon-small cell lung cancerPhase IVNCT02497118

[i] Only interventional phase III and IV clinical trials are listed here for therapeutic agents targeting vascular endothelial cells. Data were acquired from the U.S. National Library of Medicine (http://clinicaltrials.gov; date accessed, January 6, 2022). FAP, fibroblast activation protein; FGFR, fibroblast growth factor receptor; HIF, hypoxia-inducible factor; MET, mesenchymal-epithelial transition factor; nab-paclitaxel, paclitaxel albumin-stabilized nanoparticle formulation; NP, vinorelbine and cisplatin; OSCAR, OlmeSartan and Calcium Antagonists Randomized; PARP, poly (ADP-ribose) polymerase; PDGFR, platelet-derived growth factor receptor; RET, rearranged during transfection.

Targeting ECs

One main mechanism for inhibiting angiogenesis is targeting VEGF or VEGFR, alone or in combination with chemotherapeutic drugs (315). More than 400 interventional clinical trials are investigating the anticancer potential of VEGF targeting (based on http://clinicaltrials.gov; accessed, January 6, 2022). Promising results have been reported with the administration of the U.S. Food and Drug Administration (FDA)-approved bevacizumab (Avastin), an anti-angiogenic antibody that targets VEGF, and Bevacizumab-IRDye800CW, its fluorescent form (316,317). Combined administration with chemotherapeutic agents resulted in increased overall survival or progression-free survival (PFS) in CRC, NSCLC and breast cancer (315). Additionally, therapeutic strategies are currently considering the administration of two anti-angiogenic agents. For instance, a phase III trial carried out on patients with NSCLC is comparing the efficacy of two anti-VEGF antibodies, LY1008 and bevacizumab (Avastin), combined with the chemotherapeutic drugs paclitaxel and carboplatin (NCT03533127). In addition to VEGF or VEGFR targeting, anti-angiogenic strategies include using multi-receptor tyrosine kinase inhibitors that stimulate the inhibition of VEGF, VGEFR, PDGFR or c-Kit (NCT03533127). The FDA-approved pazopanib (Votrient) is one example that inhibits VEGFR1/2/3 and c-Kit for the treatment of patients with soft tissue sarcoma and RCC (318,319). There are currently three phase IV clinical trials that target the VEGF pathway. Two of these trials studied the effect of everolimus (RAD001), an mTOR inhibitor, for the treatment of patients with advanced or metastatic RCC and (NCT01206764 and NCT01266837). Another study assessed endostatin, an inhibitor of tyrosine phosphorylation of VEGFR2 (320) in combination with the anti-mitotic vinorelbine and cisplatin for the treatment of patients with NSCLC (NCT02497118). Therapeutic agents targeting vascular ECs in interventional clinical trials (in phases III and IV) are listed in Table I.

Targeting IICs

Given the pivotal role of the immune system in cancer, several anti-inflammatory drugs have been designed to inhibit tumor-promoting inflammation (50). TME immune cells within the tumor are targets of several clinical phase trials (8,51,315). One approach is the inhibition of macrophage recruitment and activation in tumors (9). This involves targeting and inhibiting colony-stimulating factor-1 (CSF-1), a macrophage-recruiting mediator, and its receptor (CSF-1R) (51). This approach is associated with reduced infiltration of TAMs and MDSCs, and inhibiting tumor progression and metastasis (51). At present, there are >50 clinical trials targeting CSF-1 and 16 clinical trials targeting CSF-1R, according to https://clinicaltrials.gov (date accessed, January 6, 2022). One promising drug, vimseltinib, a CSF-1R inhibitor also referred to as DCC-3014, has reached a phase III clinical trial and is being assessed for its efficacy in treating patients with tenosynovial giant cell tumor (NCT05059262). TAMs are characterized by the high expression of CD204 (macrophage scavenger receptor class A) and folate receptor β (FRβ) on their surface (321). TAMs were successfully eliminated using an anti-CD204 antibody and a targeted FRβ-immunotoxin in mice and rat models, respectively (322,323).

Secondly, a promising approach is to target pro-tumorigenic factors secreted by IICs (310). The use of TGF-β inhibitors and immune checkpoint inhibitors (ICIs), such as PD-L1 antibodies, and cytotoxic T-lymphocyte-associated protein 4 (CTLA-4) antibodies has been reported in a number of clinical phase trials (310,324). These strategies increase the infiltration of T cells into the tumor vicinity and the inhibition of Treg cells (8,51,325,326). Furthermore, signal transducers and transcription factors that mediate tumor growth and survival, such as STAT3 and NF-κB, are targeted (50,51). Prolonged inhibition of NFκB may lead to immune deficiency and enhanced acute inflammation (315). Consequently, the progress of NF-κB inhibitor development is obstructed in clinical trials (315,327). Pro-inflammatory chemokines and cytokines are also targeted. In in vivo studies, receptor antagonists are used to inhibit C-C chemokine receptor 2 and CXC chemokine receptor 4 (229). Clinical trials are also evaluating inhibitors targeting other cytokines, such as IL-1, IL-6 and TNFα (51,315). One important example is the FDA-approved anakinra, an IL-1 receptor antagonist used to treat patients with pancreatic cancer and metastatic breast, colon and prostate cancer (NCT02550327/NCT03233776).

Enhancing the antitumor activity by increasing the infiltration of pro-inflammatory cells is also a promising approach (51). For instance, embelin, a small-molecule inhibitor of X-linked inhibitor of apoptosis protein, induces apoptosis and suppresses gastric carcinoma and pancreatic cancer in vivo (328,329). Mechanistically, this is achieved by increasing the infiltration of pro-inflammatory immune cells, such as Th1 cells, NKs and NK T cells, while decreasing the infiltration of immunosuppressive MDSCs and IL-8- and IL-6-positive immune cells (8,328,329). Therefore, it would be interesting to move this research forward in clinical trials.

Another approach is the use of pro-inflammatory cytokines for tumor treatment. One example is the cytokine granulocyte-macrophage colony-stimulating factor (GM-CSF), which stimulates antigen presentation on macrophages and DCs, thus enhancing antibody-dependent cellular cytotoxicity (330). GM-CSF has been evaluated in a number of clinical trials, both as a monotherapy or adjuvant (NCT02451488 and NCT03686683 for example). A phase IV clinical study is testing the neoadjuvant effect of GM-CSF in cutaneous stage L-III melanoma (NCT02451488; Table III). In addition, one active non-recruiting phase III clinical trial is investigating the therapeutic potential of sipuleucel-T in patients with prostate adenocarcinoma (NCT03686683). Sipuleucel-T is an autologous cell product comprising APCs loaded with a recombinant fusion protein, PA2024, composed of prostatic acid phosphatase linked to GM-CSF (NCT03686683). Drugs targeting CSFs in phase III and IV clinical trials are presented in Table II.

Table III.

Molecularly cloned myeloid growth factors used in interventional clinical trials.

Table III.

Molecularly cloned myeloid growth factors used in interventional clinical trials.

Therapeutic agentTherapeutic strategyCancer typePhaseClinical trial
GM-CSFAloneAcute myeloid leukemiaPhase IIINCT00880243
Neoadjuvant GM-CSF treatmentCutaneous stage L-III melanomaPhase IVNCT02451488
Administration of Sipuleucel-T (an autologous cell product consisting of APCs loaded with PA2024, a recombinant fusion protein composed of prostatic acid phosphatase, linked to GM-CSF.Adenocarcinoma of the ProstatePhase IIINCT03686683
PEG-G-CSFCombination of cabazitaxel (microtubule inhibitor) with prednisolone (corticosteroid hormone) with primary prophylaxis with PEG-G-CSFProstate cancerPhase IVNCT02441894
PEG-rhG-CSFPEG-rhG-CSF for preventing neutropenia after intensive chemotherapyBreast cancerPhase IVNCT04009941 and NCT02944604
Lenograstim (G-CSF)G-CSF as primary prophylaxis for chemotherapy-induced neutropeniaSolid tumorsPhase IVNCT01107756
Neupogen (G-CSF)Neupogen vs. ciprofloxacin antibioticBreast cancerPhase IVNCT02816112
Filgrastim (G-CSF)Weight-based plerixafor (antagonist of CXCR4) compared with a fixed dose of plerixafor in combination with filgrastimNon-Hodgkin's lymphomaPhase IVNCT01164475

[i] Data were acquired from the U.S. National Library of Medicine (http://clinicaltrials.gov; date accessed, January 6, 2022). APC, antigen presenting cell; CXCR4, C-X-C chemokine receptor type 4; G-CSF, granulocyte colony stimulating factor; GM-CSF, granulocyte-macrophage colony-stimulating factor; PEG-G-CSF, PEGylated human granulocyte colony-stimulating factor; PEG-rhG-CSF, pegylated recombinant human granulocyte colony stimulating factor.

Table II.

Therapeutic agents targeting colony stimulating factors in interventional clinical trials.

Table II.

Therapeutic agents targeting colony stimulating factors in interventional clinical trials.

Therapeutic agentTherapeutic agent descriptionTherapeutic strategyCancer typePhaseClinical trial
Pexidartinib (PLX3397)CSF-1R inhibitorPexidartinib vs. placeboGiant cell tumor of the tendon sheathPhase IIINCT02371369
Durvalumab and pexidartinibAnti-PD-L1 antibody and CSF-1R inhibitorDurvalumab combined with pexidartinibPancreatic or colorectal cancerPhase INCT02777710
LY3022855 (IMC-CS4)CSF-1R inhibitorLY3022855 in combination with durvalumab (checkpoint inhibitor, targeting PD-L1/PD-1 interaction) or tremelimumab (anti-CTLA-4)Advanced solid tumorsPhase INCT02718911
Combining cyclophosphamide, pembrolizumab (an antibody that blocks negative signals to T cells), GVAX (GM-CSF pancreatic cancer vaccine) and LY3022855Pancreatic adenocarcinomaEarly phase INCT03153410
PDR001 and MCS110Anti-PD-1 and CSF-1 antibody, respectivelyPDR001 in combination with MCS110Gastric cancerPhase IINCT03694977
DCC-3014 (vimseltinib)CSF-1R inhibitorAloneTenosynovial giant cell tumorPhase I/IINCT03069469
Vimseltinib vs. placeboTenosynovial giant cell tumorPhase IIINCT05059262
DCC-3014 administered concurrently with avelumab (anti-PD-L1 antibody)SarcomasPhase INCT04242238
SNDX-6352 (UCB6352)CSF-1R antibodySNDX-6352 alone or in combination with durvalumabSolid tumors/intrahepatic cholangiocarcinomaPhase I and phase IINCT03238027 and NCT04301778
TPX-0022MET/CSF-1R/SRC inhibitorAloneSolid tumorsPhase INCT03993873
NMS-03592088FLT3, KIT and CSF-1R inhibitorAloneAcute myeloid leukemia or chronic myelomonocytic leukemiaPhase I/IINCT03922100
Q702Axl/Mer/CSF-1R selective tyrosine kinase inhibitorAloneSolid tumorPhase INCT04648254
ChiauranibTyrosine kinase inhibitorChiauranib vs. placeboSmall cell lung cancerPhase IIINCT04830813

[i] Data were acquired from the U.S. National Library of Medicine (http://clinicaltrials.gov; date accessed, January 6, 2022). CSF-1, colony-stimulating factor-1; CSF-1R, colony-stimulating factor-1 receptor; CTLA-4, cytotoxic T-lymphocyte-associated protein 4; FLT3, ms-like tyrosine kinase 3; GM-CSF, granulocyte-macrophage colony-stimulating factor; GVAX, GM-CSF pancreatic cancer vaccine; MET, mesenchymal-epithelial transition factor; PD-1, programmed cell death 1; PD-L1, programmed cell death ligand-1.

In addition to cytokine therapies and ICIs, immunity of the TME can be also triggered by adoptive cell therapy (ACT) and cancer vaccines (50). During ACT, autologous T lymphocytes with antitumor activity are isolated from a patient, expanded ex vivo, and then amplified tumor-resident or engineered T cells are transferred back to patients (331,332). One promising ACT approach is the chimeric antigen receptor (CAR) gene therapy where CAR modified T cells recognize various types of antigens regardless of their presentation on MHC molecules (333). T cells then mediate tumor killing via: i) The perforin and granzyme axis; ii) cytokine secretion; or iii) Fas-Fas ligand axis (334). Currently, there are 48 completed clinical trials that used CAR-T cell therapy on different malignancies (based on http://clinicaltrials.gov; accessed, January 6, 2022). One study evaluated CAR-engineered autologous primary human CD8+ T lymphocytes against IL13 receptor α2 in 3 patients with recurrent glioblastoma (NCT00730613), and reported promising anti-glioma activity (335). CAR-T cell immunotherapy has shown promise in terms of efficacy, while causing minimal toxicity (334,336). However, limitations such as tumor heterogeneity and antigen heterogeneous expression, as well as the function of T lymphocytes at tumor sites, make tumor eradication difficult (336).

In addition to ACT, cancer vaccines are currently intensively studied as a promising therapeutic approach that activates the humoral and cellular immunity of patients with cancer (337339). An efficient cancer vaccine design depends on a good antigen selection, where an ideal antigen should be specifically expressed and presented on all cancer cells but not normal cells, highly immunogenic and essential for the survival of cancer cells (340). After antigen delivery, DCs will uptake these antigens and present relevant antigens on MHC I and MHC II to CD8+ and CD4+ T cells, respectively (337). Effector T cells then migrate to the TME, recognize and kill cancer cells by releasing cytotoxic particles, including perforin, granzymes, IFN-γ or TNF-α, or by directly inducing apoptosis (341). In addition to T cells, B lymphocytes, NK cells and macrophages promote tumor eradication (341). Personalized vaccines are also gaining interest. There are currently three clinical trials that evaluated personalized cancer vaccination in patients with glioblastoma (active, not recruiting, NCT00045968; and completed, NCT01280552 and NCT00643097) (324). These studies reported increased numbers of infiltrating T cells with improved PFS (338,339,342). These studies show that immunization with vaccines has a promising effect in patients with cancer.

Targeting the ECM and exosomes

In addition to targeting of the cellular components of the TME and their soluble factors, TME non-cellular features are also targeted and evaluated in clinical trials. ECM remodeling and increased stiffness (desmoplasia), for instance, are targeted to reduce mortality in different cancer types (310). FDA-approved angiotensin II receptor antagonists, such as losartan and candesartan, increased the survival of patients with gastro-esophageal cancer by inhibiting the TGF-β signaling pathway and consequently reducing collagen I secretion and desmoplasia (343). Losartan has also shown clinical benefits in pancreatic cancer phase II trials (NCT01821729) when combined with FOLFIRINOX and chemoradiation with fluorouracil or capecitabine (344). The ECM may be alternatively modified by targeting integrins or focal adhesion kinase (FAK) proteins using the FAK inhibitor defactinib (NCT01870609) (345). MMP inhibitors target MMPs. However, trials failed clinically mainly because these inhibitors exhibit a broad-spectrum activity that may result in secondary side effects (346), and ECM degradation may boost cancer progression instead of inhibiting it (310,347,348).

Exosomes are: i) Targeted for reducing vesicle trafficking in cancer cells; ii) used as biomarkers for cancer diagnosis; or iii) used as vehicles of small interfering RNA for targeted therapy (51). Furthermore, the association between non-coding RNA and the TME is gaining interest, especially with respect to the TME immune environment (136,349). In this regard, Huang et al (350) developed a novel TME-related lncRNA risk model that could be used as a predictor of ICIs and a prognostic biomarker in patients with hepatocellular carcinoma.

Targeting the microbiome

Published research has linked the gut and intratumoral microbiota to response and toxicity in a variety of treatments, including chemotherapy (351). For instance, commensal microbes interact with chemotherapeutics primarily by modulating drug metabolism and host immunity (351,352). Drug activity can either be directly driven by microbes or indirectly driven by microbe-derived metabolites (352). Therefore, targeting the microbiome may hold promise for improving chemotherapeutic efficacy and lowering toxicity (18). Retrospective clinical studies on patients with PDAC demonstrated that administration of antibiotics to target bacteria that produce a long isoform of cytidine deaminase resulted in improved gemcitabine response, and thus, overcame the intratumoral bacterial-induced chemoresistance (353355).

The microbiome has also been recognized for its intricate interaction with host immunity, and thus, is considered a potential therapeutic target to optimize immunotherapy responses (310). Gut microbiota, in particular, serve a role in modulating immune checkpoint blockade responses in multiple cancer types (356359). A recent recruiting observational study aims to evaluate the effect of the microbiome in terms of efficacy and toxicity of ICIs in patients with advanced cancer (NCT04107168). The search for biomarkers in the gut microbiome has resulted in the identification of microbiome signatures that aid in determining when ICIs are effective (360,361). According to a study that looked at phase II neoadjuvant trials of anti-programmed cell death 1 (PD-1)/anti-CTLA-4 antibodies for melanoma, NSCLC and sarcoma, patients with high abundance of Ruminococcus were reported as responders with a marked increase in B cell signatures (362). Given that the favorable microbiota signatures result in enhanced intratumoral immune infiltrates (357359), creating an ideal combination of bacteria is a potential therapeutic approach to be administered in combination with checkpoint blockade.

In addition to targeting the gut microbiome, efforts are now being made to target the tumor microbiome in order to slow cancer progression and improve the response to cancer therapy. For instance, targeting the tumoral microbiome with antibiotics results in enhanced response to both chemotherapy and ICIs in CRC and pancreatic cancer (181,363,364). The intratumoral microbiome can also be targeted by bioengineered bacteria that can either kill tumor cells directly, or create an immune microenvironment that encourages antitumor immune responses (18). In mice for example, attenuated Salmonella strains expressing Vibrio-derived toll-like receptor 5 ligand flagellin elicited an immune response that recruited an antitumor immune responses against orthotopic human CRC87 lines (365).

A growing body of evidence suggests that the microbiome serves a role in determining cancer therapeutic efficacy and toxicity (18,351). Laboratory research and clinical trials have also shown that microbiota modulation can help with cancer treatment (18,366368). Therefore, understanding the microbiome and its interactions with cancer is critical in personalized medicine. Manipulation of the gut microbiota may yield novel cancer treatment insights for enhanced cancer therapeutic responses. Since the microbiome exhibits complex interactions with both the host immunity and cancer cells, it would be challenging to identify an optimal bacterial consortia and metabolites to affect the TME, as well as to introduce it for cancer treatment (18).

In addition to the aforementioned targeted therapeutic strategies, combination therapies have gained popularity as they result in enhanced efficacy, reduced drug resistance and lowered toxicity compared with monotherapy (369). Studies are investigating combination regimens that simultaneously affect several targets, thus achieving cooperative and synergistic effects (369,370). In this context, therapeutic agents targeting multiple stromal cells of the TME have been evaluated in interventional clinical trials (Table IV) (51,371). For example, the PD-1/PD-L1 signaling pathway is targeted with anti-angiogenic interventions targeting VEGFR1/2/3, PDGFR or c-Kit (235), or with tyrosine kinase inhibitors (372374). Additionally, anti-stromal interventions are also combined with chemotherapeutics and radiation agents (51,315,369,370). Furthermore, to better decide on a combination therapy, researchers should fully understand the stage of the tumor, as well as the specific state of the TME and its immune markers.

Table IV.

Therapeutic agents for combinatorial therapy targeting multiple stromal cells of the tumor microenvironment.

Table IV.

Therapeutic agents for combinatorial therapy targeting multiple stromal cells of the tumor microenvironment.

Therapeutic strategyCancer typeStatusClinical trial
Simlukafusp alfa (IL-2 variant targeting FAP-A) in combination with atezolizumab (anti-PD-L1)Head and neck, oesophageal, and cervical CancerPhase IINCT03386721
RO7300490 (4-1BB agonist targeting FAP) and atezolizumab (anti-PD-L1 antibody)Solid tumorsPhase INCT04857138
Sintilimab (anti-PD-1) with IBI305 (anti-VEGF) with chemotherapeutic pemetrexed and cisplatinNon-squamous non-small cell lung cancerPhase IIINCT03802240
Chemotherapeutic PLD with atezolizumab (PD-L1 inhibitor) vs. PLD with bevacizumab (anti-VEGF) and atezolizumab vs. PLD with bevacizumabOvarian, fallopian tube and peritoneal carcinomaPhase II/IIINCT02839707
Bevacizumab (anti-VEGF) with chemotherapeutic carboplatin and pemetrexed vs. bevacizumab with atezolizumab (anti-PD-1), carboplatin and pemetrexedMalignant pleural mesotheliomaPhase IIINCT03762018

[i] Data were acquired from the U.S. National Library of Medicine (http://clinicaltrials.gov; date accessed, January 6, 2022). FAP, fibroblast activation protein; PD-1, programmed cell death 1; PD-L1, programmed cell death ligand-1; PLD, pegylated liposomal doxorubicin.

Discussion

Cancer is a complex disease caused by malignant cells and a supporting TME. The cross-talk between these two main entities embodied by bidirectional mediators governs tumor progression. Low levels of both oxygen and pH create local stress within the TME, triggering a response from, thereby activated, stromal cells and the infiltration of more immune cells (136). Such communications are not only governed by cytokines, chemokines and metabolic products secreted by TME stromal cells but other factors such as epigenetic factors (such as miRNA), methylation DNA and histone modification are also critical (375). Furthermore, an increasing number of studies have highlighted the important effects of metabolism on the activities of immune cells, and thus, their effect on cancer progression (257,260,376). The orchestration of autocrine and paracrine communications within the tumor environment may expedite ECM stiffness, inflammation and angiogenesis, and possibly cancer cell dissemination and metastasis (4). These results warrant examining the effect of TME components on the outcome of the disease. Such findings urged research to investigate the relevance of TME targeting for more efficient therapeutic methods. While most studies focus mainly on the stromal composition of the TME (64,66), the present review provides a comprehensive examination of not only the stromal components but also the non-cellular TME components, including the ECM, exosomes and microbiome. The present discusses the contribution of cellular and non-cellular TME components to fundamental cancer hallmarks as well as emerging hallmarks and enabling characteristics. The present review also provides a detailed report on TME cells, signaling pathways and soluble factors that can be targeted for cancer therapy, highlighting TME components that are currently targeted in interventional clinical trials.

TME targeting provides promising strategies to overcome the chemotherapeutic resistance of tumor cells. Research efforts have resulted in the development of FDA-approved or newly developed TME-targeted drugs, including anti-angiogenic and anti-inflammatory agents. Clinical implementation of these drugs also shows promising successful clinical results (9,51,315). In addition, and with the help of large-scale data mining and bioinformatics analysis, several immune-related gene signatures serving as predictors for therapeutic outcomes or biomarkers for prognosis in several cancer types have been constructed (136,377380).

TME-targeted strategies may soon become mainstream for cancer therapy and can be used in combination with conventional antitumor methods. However, further research is required to address the time of TME-targeted drug administration and the treatment strategies, since certain studies have indicated that TME components may augment tumor resistance to cancer therapy (97,181,200,381). For instance, CAFs promote resistance to chemotherapy primarily by mediating EMT, maintaining the stemness of cancer stem cells and promoting metabolic reprogramming (382). The augmented ECM deposition and increased cytokine secretions mediated by CAFs may aid tumor cells in resisting cancer-therapies and, in particular, chemotherapy (383). Furthermore, a growing body of evidence suggests that hypoxia-driven residual VEGF and other pro-angiogenic factors cause resistance to VEGF receptor inhibition (381,384). Therefore, combinations of medicines targeting these factors may enhance treatment outcomes compared with single VEGF pathway blocking alone (385).

In addition to the aforementioned concerns, the pharmacokinetics and biodistribution of TME-targeted drugs is not yet well investigated due to the difficulty of detecting the exact state of the TME (386). One way to overcome this limitation is using 3D cell culture systems to recapitulate the complexity of tumor architecture and simulate the TME (387). Another method is using animal models that facilitate the recreation of a developed tumor in an improved pathophysiologic environment (324,388390). Tumor tissues obtained from a patient are processed into patient-derived organoids or patient-derived xenografts, which are then functionally and quantitively analyzed after treatment [reviewed in (391)]. These methods may help identify clinically relevant immune checkpoints and predict treatment efficacy (391). Such efforts allow for an enhanced pre-clinical validation of novel cancer methodologies towards full integration of immunotherapeutic prediction tools (391,392). In addition to combination therapy, nanotechnology also promises good therapeutic prospects (324,393). Regarding all discussed interventions and their limitations, and arriving at the era of the comprehensive cancer model treatment, it is important to treat tumors as a multifactorial disease in a stage, tissue/organ and patient-specific manner.

Acknowledgements

Not applicable.

Funding

Funding: No funding was received.

Availability of data and materials

Not applicable.

Authors' contributions

RN and IF equally designed the review article and wrote the majority of the article. AEF researched references and contributed to the writing. RAH and MES wrote the final draft and edited the manuscript. Data authentication is not applicable. All authors read and approved the final manuscript.

Ethics approval and consent to participate

Not applicable.

Patient consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Glossary

Abbreviations

Abbreviations:

ACT

adoptive cell therapy

αSMA

α-smooth muscle actin

APCs

antigen-presenting cells

AVCs

angiogenic vascular cells

CAFs

cancer-associated fibroblasts

CAR

chimeric antigen receptor

CCL

C-C motif chemokine ligand

circRNA

circular RNA

CRC

colorectal cancer

CSF-1

colony-stimulating factor-1

CSF-1R

colony-stimulating factor-1 receptor

CTLs

cytotoxic T lymphocytes

CTLA-4

cytotoxic T-lymphocyte-associated protein 4

CXCL

CXC-chemokine ligand

DCs

dendritic cells

ECs

endothelial cells

ECM

extracellular matrix

EGF

epithelial growth factors

EMT

epithelial-to-mesenchymal transition

EVs

extracellular vesicles

FAK

focal adhesion kinase

FAP

fibroblast activation protein

FGF

fibroblast growth factors

GM-CSF

granulocyte-macrophage colony-stimulating factor

HA

hyaluronic acid

HEVs

high endothelial venules

HGF

hepatocyte growth factor

HOTAIR

HOX transcript antisense RNA

ICIs

immune checkpoint inhibitors

IFN

interferon

IGF

insulin-like growth factor

IICs

infiltrating immune cells

IMCs

immature myeloid cells

LINC

long intergenic non-coding RNA

lncRNA

long non-coding RNA

MDSCs

myeloid-derived suppressor cells

MHC

major histocompatibility complex

MIP

macrophage inflammatory protein

miRNA

microRNA

MSCs

mesenchymal stem cells

NKs

natural killer cells

NSCLC

non-small cell lung cancer

OXPHOS

oxidative phosphorylation

PD-1

programmed cell death 1

PD-L1

programmed cell death ligand-1

PDAC

pancreatic ductal adenocarcinoma

PDGFR

platelet-derived growth factor receptor

PFS

progression-free survival

PGs

proteoglycans

RCC

renal cell carcinoma

SDF

stromal cell-derived factor

TAMs

tumor-associated macrophages

TCA

tricarboxylic acid

TME

tumor microenvironment

VCAM1

secreted vascular cell adhesion molecule 1

References

1 

Cooper GM: The development and causes of cancer, The cell: A molecular approach. 2nd edition. Sunderland (MA): Sinauer Associates; 2000

2 

Balkwill FR, Capasso M and Hagemann T: The tumor microenvironment at a glance. J Cell Sci. 125:5591–5596. 2012. View Article : Google Scholar : PubMed/NCBI

3 

Lu P, Weaver VM and Werb Z: The extracellular matrix: A dynamic niche in cancer progression. J Cell Biol. 196:395–406. 2012. View Article : Google Scholar : PubMed/NCBI

4 

Hanahan D and Weinberg RA: Hallmarks of cancer: The next generation. Cell. 144:646–674. 2011. View Article : Google Scholar : PubMed/NCBI

5 

Hanahan D and Coussens LM: Accessories to the crime: Functions of cells recruited to the tumor microenvironment. Cancer Cell. 21:309–322. 2012. View Article : Google Scholar : PubMed/NCBI

6 

Li H, Fan X and Houghton J: Tumor microenvironment: The role of the tumor stroma in cancer. J Cell Biochem. 101:805–815. 2007. View Article : Google Scholar : PubMed/NCBI

7 

Pietras K and Ostman A: Hallmarks of cancer: Interactions with the tumor stroma. Exp Cell Res. 316:1324–1331. 2010. View Article : Google Scholar : PubMed/NCBI

8 

Li H, Zhou L, Zhou J, Li Q and Ji Q: Underlying mechanisms and drug intervention strategies for the tumour microenvironment. J Exp Clin Cancer Res. 40:972021. View Article : Google Scholar : PubMed/NCBI

9 

Joyce JA: Therapeutic targeting of the tumor microenvironment. Cancer Cell. 7:513–520. 2005. View Article : Google Scholar : PubMed/NCBI

10 

Henke E, Nandigama R and Ergün S: Extracellular matrix in the tumor microenvironment and its impact on cancer therapy. Front Mol Biosci. 6:1602020. View Article : Google Scholar : PubMed/NCBI

11 

Tsai MJ, Chang WA, Huang MS and Kuo PL: Tumor microenvironment: A new treatment target for cancer. ISRN Biochem. 2014:e3519592014. View Article : Google Scholar : PubMed/NCBI

12 

Willumsen N, Thomsen LB, Bager CL, Jensen C and Karsdal MA: Quantification of altered tissue turnover in a liquid biopsy: A proposed precision medicine tool to assess chronic inflammation and desmoplasia associated with a pro-cancerous niche and response to immuno-therapeutic anti-tumor modalities. Cancer Immunol Immunother. 67:1–12. 2018. View Article : Google Scholar : PubMed/NCBI

13 

Chen F, Zhuang X, Lin L, Yu P, Wang Y, Shi Y, Hu G and Sun Y: New horizons in tumor microenvironment biology: Challenges and opportunities. BMC Med. 13:452015. View Article : Google Scholar : PubMed/NCBI

14 

Binnewies M, Roberts EW, Kersten K, Chan V, Fearon DF, Merad M, Coussens LM, Gabrilovich DI, Ostrand-Rosenberg S, Hedrick CC, et al: Understanding the tumor immune microenvironment (TIME) for effective therapy. Nat Med. 24:541–550. 2018. View Article : Google Scholar : PubMed/NCBI

15 

Chen X and Song E: Turning foes to friends: Targeting cancer-associated fibroblasts. Nat Rev Drug Discov. 18:99–115. 2019. View Article : Google Scholar : PubMed/NCBI

16 

Brassart-Pasco S, Brézillon S, Brassart B, Ramont L, Oudart JB and Monboisse JC: Tumor microenvironment: Extracellular matrix alterations influence tumor progression. Front Oncol. 10:3972020. View Article : Google Scholar : PubMed/NCBI

17 

Wang Z, Chen JQ, Liu JL and Tian L: Exosomes in tumor microenvironment: Novel transporters and biomarkers. J Transl Med. 14:2972016. View Article : Google Scholar : PubMed/NCBI

18 

Helmink BA, Khan MAW, Hermann A, Gopalakrishnan V and Wargo JA: The microbiome, cancer, and cancer therapy. Nat Med. 25:377–388. 2019. View Article : Google Scholar : PubMed/NCBI

19 

Sahai E, Astsaturov I, Cukierman E, DeNardo DG, Egeblad M, Evans RM, Fearon D, Greten FR, Hingorani SR, Hunter T, et al: A framework for advancing our understanding of cancer-associated fibroblasts. Nat Rev Cancer. 20:174–186. 2020. View Article : Google Scholar : PubMed/NCBI

20 

Ganguly D, Chandra R, Karalis J, Teke M, Aguilera T, Maddipati R, Wachsmann MB, Ghersi D, Siravegna G, Zeh HJ III, et al: Cancer-associated fibroblasts: Versatile players in the tumor microenvironment. Cancers (Basel). 12:26522020. View Article : Google Scholar : PubMed/NCBI

21 

Liao Z, Tan ZW, Zhu P and Tan NS: Cancer-associated fibroblasts in tumor microenvironment-accomplices in tumor malignancy. Cell Immunol. 343:1037292019. View Article : Google Scholar : PubMed/NCBI

22 

Arina A, Idel C, Hyjek EM, Alegre ML, Wang Y, Bindokas VP, Weichselbaum RR and Schreiber H: Tumor-associated fibroblasts predominantly come from local and not circulating precursors. Proc Natl Acad Sci USA. 113:7551–7556. 2016. View Article : Google Scholar : PubMed/NCBI

23 

Jung Y, Kim JK, Shiozawa Y, Wang J, Mishra A, Joseph J, Berry JE, McGee S, Lee E, Sun H, et al: Recruitment of mesenchymal stem cells into prostate tumours promotes metastasis. Nat Commun. 4:17952013. View Article : Google Scholar : PubMed/NCBI

24 

Mishra PJ, Mishra PJ, Humeniuk R, Medina DJ, Alexe G, Mesirov JP, Ganesan S, Glod JW and Banerjee D: Carcinoma-associated fibroblast-like differentiation of human mesenchymal stem cells. Cancer Res. 68:4331–4339. 2008. View Article : Google Scholar : PubMed/NCBI

25 

Zhu Q, Zhang X, Zhang L, Li W, Wu H, Yuan X, Mao F, Wang M, Zhu W, Qian H and Xu W: The IL-6-STAT3 axis mediates a reciprocal crosstalk between cancer-derived mesenchymal stem cells and neutrophils to synergistically prompt gastric cancer progression. Cell Death Dis. 5:e12952014. View Article : Google Scholar : PubMed/NCBI

26 

Jotzu C, Alt E, Welte G, Li J, Hennessy BT, Devarajan E, Krishnappa S, Pinilla S, Droll L and Song YH: Adipose tissue-derived stem cells differentiate into carcinoma-associated fibroblast-like cells under the influence of tumor-derived factors. Anal Cell Pathol (Amst). 33:61–79. 2010. View Article : Google Scholar : PubMed/NCBI

27 

Kidd S, Spaeth E, Watson K, Burks J, Lu H, Klopp A, Andreeff M and Marini FC: Origins of the tumor microenvironment: Quantitative assessment of adipose-derived and bone marrow-derived stroma. PLoS One. 7:e305632012. View Article : Google Scholar : PubMed/NCBI

28 

Miyazaki Y, Oda T, Mori N and Kida YS: Adipose-derived mesenchymal stem cells differentiate into pancreatic cancer-associated fibroblasts in vitro. FEBS Open Bio. 10:2268–2281. 2020. View Article : Google Scholar : PubMed/NCBI

29 

Nurmik M, Ullmann P, Rodriguez F, Haan S and Letellier E: In search of definitions: Cancer-associated fibroblasts and their markers. Int J Cancer. 146:895–905. 2020. View Article : Google Scholar : PubMed/NCBI

30 

Simon T and Salhia B: Cancer-Associated fibroblast subpopulations with diverse and dynamic roles in the tumor microenvironment. Mol Cancer Res. 20:183–192. 2022. View Article : Google Scholar : PubMed/NCBI

31 

Sebastian A, Hum NR, Martin KA, Gilmore SF, Peran I, Byers SW, Wheeler EK, Coleman MA and Loots GG: Single-cell transcriptomic analysis of tumor-derived fibroblasts and normal tissue-resident fibroblasts reveals fibroblast heterogeneity in breast cancer. Cancers (Basel). 12:13072020. View Article : Google Scholar : PubMed/NCBI

32 

Öhlund D, Handly-Santana A, Biffi G, Elyada E, Almeida AS, Ponz-Sarvise M, Corbo V, Oni TE, Hearn SA, Lee EJ, et al: Distinct populations of inflammatory fibroblasts and myofibroblasts in pancreatic cancer. J Exp Med. 214:579–596. 2017. View Article : Google Scholar : PubMed/NCBI

33 

Elyada E, Bolisetty M, Laise P, Flynn WF, Courtois ET, Burkhart RA, Teinor JA, Belleau P, Biffi G, Lucito MS, et al: Cross-species single-cell analysis of pancreatic ductal adenocarcinoma reveals antigen-presenting cancer-associated fibroblasts. Cancer Discov. 9:1102–1123. 2019. View Article : Google Scholar : PubMed/NCBI

34 

Cohen N, Shani O, Raz Y, Sharon Y, Hoffman D, Abramovitz L and Erez N: Fibroblasts drive an immunosuppressive and growth-promoting microenvironment in breast cancer via secretion of Chitinase 3-like 1. Oncogene. 36:4457–4468. 2017. View Article : Google Scholar : PubMed/NCBI

35 

Amatangelo MD, Bassi DE, Klein-Szanto AJP and Cukierman E: Stroma-derived three-dimensional matrices are necessary and sufficient to promote desmoplastic differentiation of normal fibroblasts. Am J Pathol. 167:475–488. 2005. View Article : Google Scholar : PubMed/NCBI

36 

Carmeliet P and Jain RK: Principles and mechanisms of vessel normalization for cancer and other angiogenic diseases. Nat Rev Drug Discov. 10:417–427. 2011. View Article : Google Scholar : PubMed/NCBI

37 

Nishida N, Yano H, Nishida T, Kamura T and Kojiro M: Angiogenesis in cancer. Vasc Health Risk Manag. 2:213–219. 2006. View Article : Google Scholar : PubMed/NCBI

38 

Jain RK: Normalization of tumor vasculature: An emerging concept in antiangiogenic therapy. Science. 307:58–62. 2005. View Article : Google Scholar : PubMed/NCBI

39 

Semenza GL: Cancer-stromal cell interactions mediated by hypoxia-inducible factors promote angiogenesis, lymphangiogenesis, and metastasis. Oncogene. 32:4057–4063. 2013. View Article : Google Scholar : PubMed/NCBI

40 

Weis SM and Cheresh DA: Tumor angiogenesis: Molecular pathways and therapeutic targets. Nat Med. 17:1359–1370. 2011. View Article : Google Scholar : PubMed/NCBI

41 

Hillen F and Griffioen AW: Tumour vascularization: Sprouting angiogenesis and beyond. Cancer Metastasis Rev. 26:489–502. 2007. View Article : Google Scholar : PubMed/NCBI

42 

Italiano JE Jr, Richardson JL, Patel-Hett S, Battinelli E, Zaslavsky A, Short S, Ryeom S, Folkman J and Klement GL: Angiogenesis is regulated by a novel mechanism: Pro- and antiangiogenic proteins are organized into separate platelet alpha granules and differentially released. Blood. 111:1227–1233. 2008. View Article : Google Scholar : PubMed/NCBI

43 

Armulik A, Genové G and Betsholtz C: Pericytes: Developmental, physiological, and pathological perspectives, problems, and promises. Dev Cell. 21:193–215. 2011. View Article : Google Scholar : PubMed/NCBI

44 

Balta E, Wabnitz GH and Samstag Y: Hijacked immune cells in the tumor microenvironment: Molecular mechanisms of immunosuppression and cues to improve T cell-based immunotherapy of solid tumors. Int J Mol Sci. 22:57362021. View Article : Google Scholar : PubMed/NCBI

45 

Biswas SK and Mantovani A: Macrophage plasticity and interaction with lymphocyte subsets: Cancer as a paradigm. Nat Immunol. 11:889–896. 2010. View Article : Google Scholar : PubMed/NCBI

46 

Franklin RA, Liao W, Sarkar A, Kim MV, Bivona MR, Liu K, Pamer EG and Li MO: The cellular and molecular origin of tumor-associated macrophages. Science. 344:921–925. 2014. View Article : Google Scholar : PubMed/NCBI

47 

Wu K, Lin K, Li X, Yuan X, Xu P, Ni P and Xu D: Redefining tumor-associated macrophage subpopulations and functions in the tumor microenvironment. Front Immunol. 11:17312020. View Article : Google Scholar : PubMed/NCBI

48 

Dehne N, Mora J, Namgaladze D, Weigert A and Brüne B: Cancer cell and macrophage cross-talk in the tumor microenvironment. Curr Opin Pharmacol. 35:12–19. 2017. View Article : Google Scholar : PubMed/NCBI

49 

Cassetta L, Fragkogianni S, Sims AH, Swierczak A, Forrester LM, Zhang H, Soong DYH, Cotechini T, Anur P, Lin EY, et al: Human tumor-associated macrophage and monocyte transcriptional landscapes reveal cancer-specific reprogramming, biomarkers, and therapeutic targets. Cancer Cell. 35:588–602.e10. 2019. View Article : Google Scholar : PubMed/NCBI

50 

Costa AC, Santos JMO, Gil da Costa RM and Medeiros R: Impact of immune cells on the hallmarks of cancer: A literature review. Crit Rev Oncol Hematol. 168:1035412021. View Article : Google Scholar : PubMed/NCBI

51 

Roma-Rodrigues C, Mendes R, Baptista PV and Fernandes AR: Targeting tumor microenvironment for cancer therapy. Int J Mol Sci. 20:8402019. View Article : Google Scholar : PubMed/NCBI

52 

Ostrand-Rosenberg S and Sinha P: Myeloid-derived suppressor cells: Linking inflammation and cancer. J Immunol. 182:4499–4506. 2009. View Article : Google Scholar : PubMed/NCBI

53 

Veglia F, Sanseviero E and Gabrilovich DI: Myeloid-derived suppressor cells in the era of increasing myeloid cell diversity. Nat Rev Immunol. 21:485–498. 2021. View Article : Google Scholar : PubMed/NCBI

54 

Masucci MT, Minopoli M and Carriero MV: Tumor associated neutrophils. Their role in tumorigenesis, metastasis, prognosis and therapy. Front Oncol. 9:11462019. View Article : Google Scholar : PubMed/NCBI

55 

Ye Y, Gaugler B, Mohty M and Malard F: Plasmacytoid dendritic cell biology and its role in immune-mediated diseases. Clin Transl Immunology. 9:e11392020. View Article : Google Scholar : PubMed/NCBI

56 

Karthaus N, Torensma R and Tel J: Deciphering the message broadcast by tumor-infiltrating dendritic cells. Am J Pathol. 181:733–742. 2012. View Article : Google Scholar : PubMed/NCBI

57 

Benavente S, Sánchez-García A, Naches S, LLeonart ME and Lorente J: Therapy-induced modulation of the tumor microenvironment: New opportunities for cancer therapies. Front Oncol. 10:5828842020. View Article : Google Scholar : PubMed/NCBI

58 

Campbell DJ and Koch MA: Treg cells: Patrolling a dangerous neighborhood. Nat Med. 17:929–930. 2011. View Article : Google Scholar : PubMed/NCBI

59 

Hsieh CS, Lee HM and Lio CWJ: Selection of regulatory T cells in the thymus. Nat Rev Immunol. 12:157–167. 2012. View Article : Google Scholar : PubMed/NCBI

60 

Coronella JA, Telleman P, Kingsbury GA, Truong TD, Hays S and Junghans RP: Evidence for an antigen-driven humoral immune response in medullary ductal breast cancer. Cancer Res. 61:7889–7899. 2001.PubMed/NCBI

61 

Schioppa T, Moore R, Thompson RG, Rosser EC, Kulbe H, Nedospasov S, Mauri C, Coussens LM and Balkwill FR: B regulatory cells and the tumor-promoting actions of TNF-α during squamous carcinogenesis. Proc Natl Acad Sci USA. 108:10662–10667. 2011. View Article : Google Scholar : PubMed/NCBI

62 

Quail DF and Joyce JA: Microenvironmental regulation of tumor progression and metastasis. Nat Med. 19:1423–1437. 2013. View Article : Google Scholar : PubMed/NCBI

63 

Wu J and Lanier LL: Natural killer cells and cancer. Adv Cancer Res. 90:127–156. 2003. View Article : Google Scholar : PubMed/NCBI

64 

Hong J, Jin JO, Chen WY, Poggi A and Cheong JH: Editorial: Emerging roles and mechanisms of stromal cells in carcinomas at the molecular level. Front Immunol. 13:10258382022. View Article : Google Scholar : PubMed/NCBI

65 

Koppensteiner L, Mathieson L, O'Connor RA and Akram AR: Cancer associated fibroblasts-an impediment to effective anti-cancer T cell immunity. Front Immunol. 13:8873802022. View Article : Google Scholar : PubMed/NCBI

66 

Mun JY, Leem SH, Lee JH and Kim HS: Dual relationship between stromal cells and immune cells in the tumor microenvironment. Front Immunol. 13:8647392022. View Article : Google Scholar : PubMed/NCBI

67 

Belli C, Antonarelli G, Repetto M, Boscolo Bielo L, Crimini E and Curigliano G: Targeting cellular components of the tumor microenvironment in solid malignancies. Cancers (Basel). 14:42782022. View Article : Google Scholar : PubMed/NCBI

68 

Theocharis AD, Skandalis SS, Gialeli C and Karamanos NK: Extracellular matrix structure. Adv Drug Deliv Rev. 97:4–27. 2016. View Article : Google Scholar : PubMed/NCBI

69 

Badylak SF: The extracellular matrix as a biologic scaffold material. Biomaterials. 28:3587–3593. 2007. View Article : Google Scholar : PubMed/NCBI

70 

Fuhrmann A and Engler AJ: The Cytoskeleton regulates cell attachment strength. Biophys J. 109:57–65. 2015. View Article : Google Scholar : PubMed/NCBI

71 

Kechagia JZ, Ivaska J and Roca-Cusachs P: Integrins as biomechanical sensors of the microenvironment. Nat Rev Mol Cell Biol. 20:457–473. 2019. View Article : Google Scholar : PubMed/NCBI

72 

Romani P, Valcarcel-Jimenez L, Frezza C and Dupont S: Crosstalk between mechanotransduction and metabolism. Nat Rev Mol Cell Biol. 22:22–38. 2021. View Article : Google Scholar : PubMed/NCBI

73 

Cichon MA and Radisky DC: Extracellular matrix as a contextual determinant of transforming growth factor-β signaling in epithelial-mesenchymal transition and in cancer. Cell Adhes Migr. 8:588–594. 2014. View Article : Google Scholar : PubMed/NCBI

74 

Karamanos NK, Piperigkou Z, Passi A, Götte M, Rousselle P and Vlodavsky I: Extracellular matrix-based cancer targeting. Trends Mol Med. 27:1000–1013. 2021. View Article : Google Scholar : PubMed/NCBI

75 

Rilla K, Mustonen AM, Arasu UT, Härkönen K, Matilainen J and Nieminen P: Extracellular vesicles are integral and functional components of the extracellular matrix. Matrix Biol. 75–76. 201–219. 2019.PubMed/NCBI

76 

Apte MV, Yang L, Phillips PA, Xu Z, Kaplan W, Cowley M, Pirola RC and Wilson JS: Extracellular matrix composition significantly influences pancreatic stellate cell gene expression pattern: Role of transgelin in PSC function. Am J Physiol Gastrointest Liver Physiol. 305:G408–G417. 2013. View Article : Google Scholar : PubMed/NCBI

77 

Naba A, Clauser KR, Hoersch S, Liu H, Carr SA and Hynes RO: The matrisome: In silico definition and in vivo characterization by proteomics of normal and tumor extracellular matrices. Mol Cell Proteomics. 11:M111.014647. 2012. View Article : Google Scholar : PubMed/NCBI

78 

Musiime M, Chang J, Hansen U, Kadler KE, Zeltz C and Gullberg D: Collagen assembly at the cell surface: Dogmas revisited. Cells. 10:6622021. View Article : Google Scholar : PubMed/NCBI

79 

Eble JA and Niland S: The extracellular matrix in tumor progression and metastasis. Clin Exp Metastasis. 36:171–198. 2019. View Article : Google Scholar : PubMed/NCBI

80 

Yue B: Biology of the extracellular matrix: An overview. J Glaucoma. 23 (8 Suppl 1):S20–S23. 2014. View Article : Google Scholar : PubMed/NCBI

81 

Naba A, Pearce OMT, Del Rosario A, Ma D, Ding H, Rajeeve V, Cutillas PR, Balkwill FR and Hynes RO: Characterization of the extracellular matrix of normal and diseased tissues using proteomics. J Proteome Res. 16:3083–3091. 2017. View Article : Google Scholar : PubMed/NCBI

82 

Erdogan B and Webb DJ: Cancer-associated fibroblasts modulate growth factor signaling and extracellular matrix remodeling to regulate tumor metastasis. Biochem Soc Trans. 45:229–236. 2017. View Article : Google Scholar : PubMed/NCBI

83 

Najafi M, Farhood B and Mortezaee K: Extracellular matrix (ECM) stiffness and degradation as cancer drivers. J Cell Biochem. 120:2782–2790. 2019. View Article : Google Scholar : PubMed/NCBI

84 

Muncie JM and Weaver VM: The physical and biochemical properties of the extracellular matrix regulate cell fate. Curr Top Dev Biol. 130:1–37. 2018. View Article : Google Scholar : PubMed/NCBI

85 

Provenzano PP, Inman DR, Eliceiri KW, Knittel JG, Yan L, Rueden CT, White JG and Keely PJ: Collagen density promotes mammary tumor initiation and progression. BMC Med. 6:112008. View Article : Google Scholar : PubMed/NCBI

86 

Mammoto T, Jiang A, Jiang E, Panigrahy D, Kieran MW and Mammoto A: Role of collagen matrix in tumor angiogenesis and glioblastoma multiforme progression. Am J Pathol. 183:1293–1305. 2013. View Article : Google Scholar : PubMed/NCBI

87 

Peinado H, Zhang H, Matei IR, Costa-Silva B, Hoshino A, Rodrigues G, Psaila B, Kaplan RN, Bromberg JF, Kang Y, et al: Pre-metastatic niches: Organ-specific homes for metastases. Nat Rev Cancer. 17:302–317. 2017. View Article : Google Scholar : PubMed/NCBI

88 

Moreira AM, Pereira J, Melo S, Fernandes MS, Carneiro P, Seruca R and Figueiredo J: The extracellular matrix: An accomplice in gastric cancer development and progression. Cells. 9:3942020. View Article : Google Scholar : PubMed/NCBI

89 

Høgdall D, Lewinska M and Andersen JB: Desmoplastic tumor microenvironment and immunotherapy in cholangiocarcinoma. Trends Cancer. 4:239–255. 2018. View Article : Google Scholar : PubMed/NCBI

90 

Ho WJ, Jaffee EM and Zheng L: The tumour microenvironment in pancreatic cancer-clinical challenges and opportunities. Nat Rev Clin Oncol. 17:527–540. 2020. View Article : Google Scholar : PubMed/NCBI

91 

Levental KR, Yu H, Kass L, Lakins JN, Egeblad M, Erler JT, Fong SF, Csiszar K, Giaccia A, Weninger W, et al: Matrix crosslinking forces tumor progression by enhancing integrin signaling. Cell. 139:891–906. 2009. View Article : Google Scholar : PubMed/NCBI

92 

Damodarasamy M, Vernon RB, Chan CK, Plymate SR, Wight TN and Reed MJ: Hyaluronan in aged collagen matrix increases prostate epithelial cell proliferation. In Vitro Cell Dev Biol Anim. 51:50–58. 2015. View Article : Google Scholar : PubMed/NCBI

93 

Suhovskih AV, Mostovich LA, Kunin IS, Boboev MM, Nepomnyashchikh GI, Aidagulova SV and Grigorieva EV: Proteoglycan expression in normal human prostate tissue and prostate cancer. ISRN Oncol. 2013:6801362013.PubMed/NCBI

94 

Ajeti V, Nadiarnykh O, Ponik SM, Keely PJ, Eliceiri KW and Campagnola PJ: Structural changes in mixed Col I/Col V collagen gels probed by SHG microscopy: Implications for probing stromal alterations in human breast cancer. Biomed Opt Express. 2:2307–2316. 2011. View Article : Google Scholar : PubMed/NCBI

95 

Fang S, Dai Y, Mei Y, Yang M, Hu L, Yang H, Guan X and Li J: Clinical significance and biological role of cancer-derived type I collagen in lung and esophageal cancers. Thorac Cancer. 10:277–288. 2019. View Article : Google Scholar : PubMed/NCBI

96 

Miskolczi Z, Smith MP, Rowling EJ, Ferguson J, Barriuso J and Wellbrock C: Collagen abundance controls melanoma phenotypes through lineage-specific microenvironment sensing. Oncogene. 37:3166–3182. 2018. View Article : Google Scholar : PubMed/NCBI

97 

Rossow L, Veitl S, Vorlová S, Wax JK, Kuhn AE, Maltzahn V, Upcin B, Karl F, Hoffmann H, Gätzner S, et al: LOX-catalyzed collagen stabilization is a proximal cause for intrinsic resistance to chemotherapy. Oncogene. 37:4921–4940. 2018. View Article : Google Scholar : PubMed/NCBI

98 

Guzman A, Ziperstein MJ and Kaufman LJ: The effect of fibrillar matrix architecture on tumor cell invasion of physically challenging environments. Biomaterials. 35:6954–6963. 2014. View Article : Google Scholar : PubMed/NCBI

99 

Xu S, Xu H, Wang W, Li S, Li H, Li T, Zhang W, Yu X and Liu L: The role of collagen in cancer: From bench to bedside. J Transl Med. 17:3092019. View Article : Google Scholar : PubMed/NCBI

100 

Gilkes DM, Semenza GL and Wirtz D: Hypoxia and the extracellular matrix: Drivers of tumour metastasis. Nat Rev Cancer. 14:430–439. 2014. View Article : Google Scholar : PubMed/NCBI

101 

Allen SC, Widman JA, Datta A and Suggs LJ: Dynamic extracellular matrix stiffening induces a phenotypic transformation and a migratory shift in epithelial cells. Integr Biol (Camb). 12:161–174. 2020. View Article : Google Scholar : PubMed/NCBI

102 

Fischer KR, Durrans A, Lee S, Sheng J, Li F, Wong ST, Choi H, El Rayes T, Ryu S, Troeger J, et al: Epithelial-to-mesenchymal transition is not required for lung metastasis but contributes to chemoresistance. Nature. 527:472–476. 2015. View Article : Google Scholar : PubMed/NCBI

103 

Han L, Lam EWF and Sun Y: Extracellular vesicles in the tumor microenvironment: Old stories, but new tales. Mol Cancer. 18:592019. View Article : Google Scholar : PubMed/NCBI

104 

Kanada M, Bachmann MH and Contag CH: Signaling by extracellular vesicles advances cancer hallmarks. Trends Cancer. 2:84–94. 2016. View Article : Google Scholar : PubMed/NCBI

105 

Minciacchi VR, Freeman MR and Di Vizio D: Extracellular vesicles in cancer: Exosomes, microvesicles and the emerging role of large oncosomes. Semin Cell Dev Biol. 40:41–51. 2015. View Article : Google Scholar : PubMed/NCBI

106 

Chronopoulos A and Kalluri R: Emerging role of bacterial extracellular vesicles in cancer. Oncogene. 39:6951–6960. 2020. View Article : Google Scholar : PubMed/NCBI

107 

Morad G and Moses MA: Brainwashed by extracellular vesicles: The role of extracellular vesicles in primary and metastatic brain tumour microenvironment. J Extracell Vesicles. 8:16271642019. View Article : Google Scholar : PubMed/NCBI

108 

Parayath NN, Padmakumar S and Amiji MM: Extracellular vesicle-mediated nucleic acid transfer and reprogramming in the tumor microenvironment. Cancer Lett. 482:33–43. 2020. View Article : Google Scholar : PubMed/NCBI

109 

Patras L and Banciu M: Intercellular crosstalk via extracellular vesicles in tumor milieu as emerging therapies for cancer progression. Curr Pharm Des. 25:1980–2006. 2019. View Article : Google Scholar : PubMed/NCBI

110 

Xie C, Ji N, Tang Z, Li J and Chen Q: The role of extracellular vesicles from different origin in the microenvironment of head and neck cancers. Mol Cancer. 18:832019. View Article : Google Scholar : PubMed/NCBI

111 

Caicedo-Carvajal CE, Liu Q and Goy A: Three-dimensional cell culture models for biomarker discoveries and cancer research. Transl Med. 1:1–8. 2012.

112 

Wang HX and Gires O: Tumor-derived extracellular vesicles in breast cancer: From bench to bedside. Cancer Lett. 460:54–64. 2019. View Article : Google Scholar : PubMed/NCBI

113 

Gould SJ and Raposo G: As we wait: Coping with an imperfect nomenclature for extracellular vesicles. J Extracell Vesicles. 2:2013. View Article : Google Scholar : PubMed/NCBI

114 

Beach A, Zhang HG, Ratajczak MZ and Kakar SS: Exosomes: An overview of biogenesis, composition and role in ovarian cancer. J Ovarian Res. 7:142014. View Article : Google Scholar : PubMed/NCBI

115 

Mashouri L, Yousefi H, Aref AR, Ahadi A mohammad, Molaei F and Alahari SK: Exosomes: Composition, biogenesis, and mechanisms in cancer metastasis and drug resistance. Mol Cancer. 18:752019. View Article : Google Scholar : PubMed/NCBI

116 

Théry C, Zitvogel L and Amigorena S: Exosomes: Composition, biogenesis and function. Nat Rev Immunol. 2:569–579. 2002. View Article : Google Scholar : PubMed/NCBI

117 

Chang WH, Cerione RA and Antonyak MA: Extracellular Vesicles and their roles in cancer progression. Methods Mol Biol. 2174:143–170. 2021. View Article : Google Scholar : PubMed/NCBI

118 

Menck K, Sivaloganathan S, Bleckmann A and Binder C: Microvesicles in cancer: Small size, large potential. Int J Mol Sci. 21:53732020. View Article : Google Scholar : PubMed/NCBI

119 

Al-Nedawi K, Meehan B, Micallef J, Lhotak V, May L, Guha A and Rak J: Intercellular transfer of the oncogenic receptor EGFRvIII by microvesicles derived from tumour cells. Nat Cell Biol. 10:619–624. 2008. View Article : Google Scholar : PubMed/NCBI

120 

Di Vizio D, Kim J, Hager MH, Morello M, Yang W, Lafargue CJ, True LD, Rubin MA, Adam RM, Beroukhim R, et al: Oncosome formation in prostate cancer: Association with a region of frequent chromosomal deletion in metastatic disease. Cancer Res. 69:5601–5609. 2009. View Article : Google Scholar : PubMed/NCBI

121 

Gurunathan S, Kang MH, Jeyaraj M, Qasim M and Kim JH: Review of the isolation, characterization, biological function, and multifarious therapeutic approaches of exosomes. Cells. 8:3072019. View Article : Google Scholar : PubMed/NCBI

122 

Hessvik NP and Llorente A: Current knowledge on exosome biogenesis and release. Cell Mol Life Sci. 75:193–208. 2018. View Article : Google Scholar : PubMed/NCBI

123 

Baig AM, Khaleeq A, Ali U and Syeda H: Evidence of the COVID-19 virus targeting the CNS: Tissue distribution, host-virus interaction, and proposed neurotropic mechanisms. ACS Chem Neurosci. 11:995–998. 2020. View Article : Google Scholar : PubMed/NCBI

124 

Li I and Nabet BY: Exosomes in the tumor microenvironment as mediators of cancer therapy resistance. Mol Cancer. 18:322019. View Article : Google Scholar : PubMed/NCBI

125 

Peinado H, Alečković M, Lavotshkin S, Matei I, Costa-Silva B, Moreno-Bueno G, Hergueta-Redondo M, Williams C, García-Santos G, Ghajar C, et al: Melanoma exosomes educate bone marrow progenitor cells toward a pro-metastatic phenotype through MET. Nat Med. 18:883–891. 2012. View Article : Google Scholar : PubMed/NCBI

126 

Walbrecq G, Margue C, Behrmann I and Kreis S: Distinct cargos of small extracellular vesicles derived from hypoxic cells and their effect on cancer cells. Int J Mol Sci. 21:50712020. View Article : Google Scholar : PubMed/NCBI

127 

Azulay EE, Cooks T and Elkabets M: Potential oncogenic roles of mutant-p53-derived exosomes in the tumor-host interaction of head and neck cancers. Cancer Immunol Immunother. 69:285–292. 2020. View Article : Google Scholar : PubMed/NCBI

128 

Pavlakis E, Neumann M and Stiewe T: Extracellular vesicles: Messengers of p53 in tumor-stroma communication and cancer metastasis. Int J Mol Sci. 21:96482020. View Article : Google Scholar : PubMed/NCBI

129 

de Jong OG, Verhaar MC, Chen Y, Vader P, Gremmels H, Posthuma G, Schiffelers RM, Gucek M and van Balkom BW: Cellular stress conditions are reflected in the protein and RNA content of endothelial cell-derived exosomes. J Extracell Vesicles. 1:2012. View Article : Google Scholar : PubMed/NCBI

130 

Drake RR and Kislinger T: The proteomics of prostate cancer exosomes. Expert Rev Proteomics. 11:167–177. 2014. View Article : Google Scholar : PubMed/NCBI

131 

Hamzah RN, Alghazali KM, Biris AS and Griffin RJ: Exosome traceability and cell source dependence on composition and cell-cell cross talk. Int J Mol Sci. 22:53462021. View Article : Google Scholar : PubMed/NCBI

132 

Jelonek K, Widlak P and Pietrowska M: The influence of ionizing radiation on exosome composition, secretion and intercellular communication. Protein Pept Lett. 23:656–663. 2016. View Article : Google Scholar : PubMed/NCBI

133 

Jiao YJ, Jin DD, Jiang F, Liu JX, Qu LS, Ni WK, Liu ZX, Lu CH, Ni RZ, Zhu J and Xiao MB: Characterization and proteomic profiling of pancreatic cancer-derived serum exosomes. J Cell Biochem. 120:988–999. 2019. View Article : Google Scholar : PubMed/NCBI

134 

Hornung T, O'Neill HA, Logie SC, Fowler KM, Duncan JE, Rosenow M, Bondre AS, Tinder T, Maher V, Zarkovic J, et al: ADAPT identifies an ESCRT complex composition that discriminates VCaP from LNCaP prostate cancer cell exosomes. Nucleic Acids Res. 48:4013–4027. 2020. View Article : Google Scholar : PubMed/NCBI

135 

Nabet BY, Qiu Y, Shabason JE, Wu TJ, Yoon T, Kim BC, Benci JL, DeMichele AM, Tchou J, Marcotrigiano J and Minn AJ: Exosome RNA unshielding couples stromal activation to pattern recognition receptor signaling in cancer. Cell. 170:352–366.e13. 2017. View Article : Google Scholar : PubMed/NCBI

136 

Luo N: Editorial: Tumor microenvironment in cancer hallmarks and therapeutics. Front Mol Biosci. 9:10198302022. View Article : Google Scholar : PubMed/NCBI

137 

Zheng Q, Yu X, Zhang M, Zhang S, Guo W and He Y: Current research progress of the role of LncRNA LEF1-AS1 in a variety of tumors. Front Cell Dev Biol. 9:7500842021. View Article : Google Scholar : PubMed/NCBI

138 

Tu J, Chen W, Zheng L, Fang S, Zhang D, Kong C, Yang Y, Qiu R, Zhao Z, Lu C, et al: Circular RNA Circ0021205 promotes cholangiocarcinoma progression through MiR-204-5p/RAB22A axis. Front Cell Dev Biol. 9:6532072021. View Article : Google Scholar : PubMed/NCBI

139 

Chen L, Wang Y, Lu X, Zhang L and Wang Z: miRNA-7062-5p promoting bone resorption after bone metastasis of colorectal cancer through inhibiting GPR65. Front Cell Dev Biol. 9:6819682021. View Article : Google Scholar : PubMed/NCBI

140 

Xiao J, Liu Y, Yi J and Liu X: LINC02257, an enhancer RNA of prognostic value in colon adenocarcinoma, correlates with multi-omics immunotherapy-related analysis in 33 cancers. Front Mol Biosci. 8:6467862021. View Article : Google Scholar : PubMed/NCBI

141 

Ma J, Lin X, Wang X, Min Q, Wang T and Tang C: Reconstruction and analysis of the immune-related LINC00987/A2M axis in lung adenocarcinoma. Front Mol Biosci. 8:6445572021. View Article : Google Scholar : PubMed/NCBI

142 

Huang K, Fang C, Yi K, Liu X, Qi H, Tan Y, Zhou J, Li Y, Liu M, Zhang Y, et al: The role of PTRF/Cavin1 as a biomarker in both glioma and serum exosomes. Theranostics. 8:1540–1557. 2018. View Article : Google Scholar : PubMed/NCBI

143 

Ayala-Mar S, Donoso-Quezada J and González-Valdez J: Clinical implications of exosomal PD-L1 in cancer immunotherapy. J Immunol Res. 2021:88399782021. View Article : Google Scholar : PubMed/NCBI

144 

Chen G, Huang AC, Zhang W, Zhang G, Wu M, Xu W, Yu Z, Yang J, Wang B, Sun H, et al: Exosomal PD-L1 contributes to immunosuppression and is associated with anti-PD-1 response. Nature. 560:382–386. 2018. View Article : Google Scholar : PubMed/NCBI

145 

Rai A, Greening DW, Chen M, Xu R, Ji H and Simpson RJ: Exosomes derived from human primary and metastatic colorectal cancer cells contribute to functional heterogeneity of activated fibroblasts by reprogramming their proteome. Proteomics. 19:e18001482019. View Article : Google Scholar : PubMed/NCBI

146 

Berg G, Rybakova D, Fischer D, Cernava T, Vergès MC, Charles T, Chen X, Cocolin L, Eversole K, Corral GH, et al: Microbiome definition re-visited: Old concepts and new challenges. Microbiome. 8:1032020. View Article : Google Scholar : PubMed/NCBI

147 

AlHilli MM and Bae-Jump V: Diet and gut microbiome interactions in gynecologic cancer. Gynecol Oncol. 159:299–308. 2020. View Article : Google Scholar : PubMed/NCBI

148 

Anfossi S and Calin GA: Gut microbiota: A new player in regulating immune- and chemo-therapy efficacy. Cancer Drug Resist. 3:356–370. 2020.PubMed/NCBI

149 

De Almeida CV, de Camargo MR, Russo E and Amedei A: Role of diet and gut microbiota on colorectal cancer immunomodulation. World J Gastroenterol. 25:151–162. 2019. View Article : Google Scholar : PubMed/NCBI

150 

Di Modica M, Gargari G, Regondi V, Bonizzi A, Arioli S, Belmonte B, De Cecco L, Fasano E, Bianchi F, Bertolotti A, et al: Gut microbiota condition the therapeutic efficacy of trastuzumab in HER2-positive breast cancer. Cancer Res. 81:2195–2206. 2021. View Article : Google Scholar : PubMed/NCBI

151 

Liu L, Tabung FK, Zhang X, Nowak JA, Qian ZR, Hamada T, Nevo D, Bullman S, Mima K, Kosumi K, et al: Diets that promote colon inflammation associate with risk of colorectal carcinomas that contain Fusobacterium nucleatum. Clin Gastroenterol Hepatol. 16:1622–1631.e3. 2018. View Article : Google Scholar : PubMed/NCBI

152 

Ponziani FR, Nicoletti A, Gasbarrini A and Pompili M: Diagnostic and therapeutic potential of the gut microbiota in patients with early hepatocellular carcinoma. Ther Adv Med Oncol. 11:17588359198481842019. View Article : Google Scholar : PubMed/NCBI

153 

Rodríguez-García C, Sánchez-Quesada C, Algarra I and Gaforio JJ: The high-fat diet based on extra-virgin olive oil causes dysbiosis linked to colorectal cancer prevention. Nutrients. 12:17052020. View Article : Google Scholar : PubMed/NCBI

154 

Laplane L, Duluc D, Bikfalvi A, Larmonier N and Pradeu T: Beyond the tumour microenvironment. Int J Cancer. 145:2611–2618. 2019. View Article : Google Scholar : PubMed/NCBI

155 

Chakladar J, Kuo SZ, Castaneda G, Li WT, Gnanasekar A, Yu MA, Chang EY, Wang XQ and Ongkeko WM: The pancreatic microbiome is associated with carcinogenesis and worse prognosis in males and smokers. Cancers (Basel). 12:26722020. View Article : Google Scholar : PubMed/NCBI

156 

Chandel D, Sharma M, Chawla V, Sachdeva N and Shukla G: Isolation, characterization and identification of antigenotoxic and anticancerous indigenous probiotics and their prophylactic potential in experimental colon carcinogenesis. Sci Rep. 9:147692019. View Article : Google Scholar : PubMed/NCBI

157 

Clanton R, Saucier D, Ford J and Akabani G: Microbial influences on hormesis, oncogenesis, and therapy: A review of the literature. Environ Res. 142:239–256. 2015. View Article : Google Scholar : PubMed/NCBI

158 

Guo W, Zhang Y, Guo S, Mei Z, Liao H, Dong H, Wu K, Ye H, Zhang Y, Zhu Y, et al: Tumor microbiome contributes to an aggressive phenotype in the basal-like subtype of pancreatic cancer. Commun Biol. 4:10192021. View Article : Google Scholar : PubMed/NCBI

159 

Huang H, Ren Z, Gao X, Hu X, Zhou Y, Jiang J, Lu H, Yin S, Ji J, Zhou L and Zheng S: Integrated analysis of microbiome and host transcriptome reveals correlations between gut microbiota and clinical outcomes in HBV-related hepatocellular carcinoma. Genome Med. 12:1022020. View Article : Google Scholar : PubMed/NCBI

160 

Ingman WV: The gut microbiome: A new player in breast cancer metastasis. Cancer Res. 79:3539–3541. 2019. View Article : Google Scholar : PubMed/NCBI

161 

Jenkins SV, Robeson MS II, Griffin RJ, Quick CM, Siegel ER, Cannon MJ, Vang KB and Dings RPM: Gastrointestinal tract dysbiosis enhances distal tumor progression through suppression of leukocyte trafficking. Cancer Res. 79:5999–6009. 2019. View Article : Google Scholar : PubMed/NCBI

162 

Li L, Deng X, Zou Y, Lv X and Guo Y: Characterization of the nasopharynx microbiota in patients with nasopharyngeal carcinoma vs healthy controls. Braz J Microbiol. 52:1873–1880. 2021. View Article : Google Scholar : PubMed/NCBI

163 

Liu HX, Tao LL, Zhang J, Zhu YG, Zheng Y, Liu D, Zhou M, Ke H, Shi MM and Qu JM: Difference of lower airway microbiome in bilateral protected specimen brush between lung cancer patients with unilateral lobar masses and control subjects. Int J Cancer. 142:769–778. 2018. View Article : Google Scholar : PubMed/NCBI

164 

Tzeng A, Sangwan N, Jia M, Liu CC, Keslar KS, Downs-Kelly E, Fairchild RL, Al-Hilli Z, Grobmyer SR and Eng C: Human breast microbiome correlates with prognostic features and immunological signatures in breast cancer. Genome Med. 13:602021. View Article : Google Scholar : PubMed/NCBI

165 

Livyatan I, Nejman D, Shental N and Straussman R: Characterization of the human tumor microbiome reveals tumor-type specific intra-cellular bacteria. Oncoimmunology. 9:18009572020. View Article : Google Scholar : PubMed/NCBI

166 

Burns MB, Lynch J, Starr TK, Knights D and Blekhman R: Virulence genes are a signature of the microbiome in the colorectal tumor microenvironment. Genome Med. 7:552015. View Article : Google Scholar : PubMed/NCBI

167 

Burns MB, Montassier E, Abrahante J, Priya S, Niccum DE, Khoruts A, Starr TK, Knights D and Blekhman R: Colorectal cancer mutational profiles correlate with defined microbial communities in the tumor microenvironment. PLoS Genet. 14:e10073762018. View Article : Google Scholar : PubMed/NCBI

168 

Lee JA, Yoo SY, Oh HJ, Jeong S, Cho NY, Kang GH and Kim JH: Differential immune microenvironmental features of microsatellite-unstable colorectal cancers according to Fusobacterium nucleatum status. Cancer Immunol Immunother. 70:47–59. 2021. View Article : Google Scholar : PubMed/NCBI

169 

Liu W, Zhang R, Shu R, Yu J, Li H, Long H, Jin S, Li S, Hu Q, Yao F, et al: Study of the relationship between microbiome and colorectal cancer susceptibility using 16SrRNA sequencing. Biomed Res Int. 2020:78283922020.PubMed/NCBI

170 

Liu X, Shao L, Liu X, Ji F, Mei Y, Cheng Y, Liu F, Yan C, Li L and Ling Z: Alterations of gastric mucosal microbiota across different stomach microhabitats in a cohort of 276 patients with gastric cancer. EBioMedicine. 40:336–348. 2019. View Article : Google Scholar : PubMed/NCBI

171 

Oresta B, Braga D, Lazzeri M, Frego N, Saita A, Faccani C, Fasulo V, Colombo P, Guazzoni G, Hurle R and Rescigno M: The microbiome of catheter collected urine in males with bladder cancer according to disease stage. J Urol. 205:86–93. 2021. View Article : Google Scholar : PubMed/NCBI

172 

Wu P, Zhang G, Zhao J, Chen J, Chen Y, Huang W, Zhong J and Zeng J: Profiling the urinary microbiota in male patients with bladder cancer in China. Front Cell Infect Microbiol. 8:1672018. View Article : Google Scholar : PubMed/NCBI

173 

Yin J, Dong L, Zhao J, Wang H, Li J, Yu A, Chen W and Wei W: Composition and consistence of the bacterial microbiome in upper, middle and lower esophagus before and after Lugol's iodine staining in the esophagus cancer screening. Scand J Gastroenterol. 55:1467–1474. 2020. View Article : Google Scholar : PubMed/NCBI

174 

Minarovits J: Anaerobic bacterial communities associated with oral carcinoma: Intratumoral, surface-biofilm and salivary microbiota. Anaerobe. 68:1023002021. View Article : Google Scholar : PubMed/NCBI

175 

Kostic AD, Gevers D, Pedamallu CS, Michaud M, Duke F, Earl AM, Ojesina AI, Jung J, Bass AJ, Tabernero J, et al: Genomic analysis identifies association of Fusobacterium with colorectal carcinoma. Genome Res. 22:292–298. 2012. View Article : Google Scholar : PubMed/NCBI

176 

Thomas AM, Jesus EC, Lopes A, Aguiar S Jr, Begnami MD, Rocha RM, Carpinetti PA, Camargo AA, Hoffmann C, Freitas HC, et al: Tissue-associated bacterial alterations in rectal carcinoma patients revealed by 16S rRNA community profiling. Front Cell Infect Microbiol. 6:1792016. View Article : Google Scholar : PubMed/NCBI

177 

Thompson KJ, Ingle JN, Tang X, Chia N, Jeraldo PR, Walther-Antonio MR, Kandimalla KK, Johnson S, Yao JZ, Harrington SC, et al: A comprehensive analysis of breast cancer microbiota and host gene expression. PLoS One. 12:e01888732017. View Article : Google Scholar : PubMed/NCBI

178 

Cavarretta I, Ferrarese R, Cazzaniga W, Saita D, Lucianò R, Ceresola ER, Locatelli I, Visconti L, Lavorgna G, Briganti A, et al: The microbiome of the prostate tumor microenvironment. Eur Urol. 72:625–631. 2017. View Article : Google Scholar : PubMed/NCBI

179 

Duan H, Chen L, Qu L, Yang H, Song SW, Han Y, Ye M, Chen W, He X and Shou C: Mycoplasma hyorhinis infection promotes NF-κB-dependent migration of gastric cancer cells. Cancer Res. 74:5782–5794. 2014. View Article : Google Scholar : PubMed/NCBI

180 

Duan H, Qu L and Shou C: Mycoplasma hyorhinis induces epithelial-mesenchymal transition in gastric cancer cell MGC803 via TLR4-NF-κB signaling. Cancer Lett. 354:447–454. 2014. View Article : Google Scholar : PubMed/NCBI

181 

Geller LT, Barzily-Rokni M, Danino T, Jonas OH, Shental N, Nejman D, Gavert N, Zwang Y, Cooper ZA, Shee K, et al: Potential role of intratumor bacteria in mediating tumor resistance to the chemotherapeutic drug gemcitabine. Science. 357:1156–1160. 2017. View Article : Google Scholar : PubMed/NCBI

182 

Huang S, Li JY, Wu J, Meng L and Shou CC: Mycoplasma infections and different human carcinomas. World J Gastroenterol. 7:266–269. 2001. View Article : Google Scholar : PubMed/NCBI

183 

Liu D, Hu Y, Guo Y, Zhu Z, Lu B, Wang X and Huang Y: Mycoplasma-associated multidrug resistance of hepatocarcinoma cells requires the interaction of P37 and Annexin A2. PLoS One. 12:e01845782017. View Article : Google Scholar : PubMed/NCBI

184 

Liu X, Rong Z and Shou C: Mycoplasma hyorhinis infection promotes gastric cancer cell motility via β-catenin signaling. Cancer Med. 8:5301–5312. 2019. View Article : Google Scholar : PubMed/NCBI

185 

Xu Y, Li H, Chen W, Yao X, Xing Y, Wang X, Zhong J and Meng G: Mycoplasma hyorhinis activates the NLRP3 inflammasome and promotes migration and invasion of gastric cancer cells. PLoS One. 8:e779552013. View Article : Google Scholar : PubMed/NCBI

186 

Gedye C, Cardwell T, Dimopoulos N, Tan BS, Jackson H, Svobodová S, Anaka M, Behren A, Maher C, Hofmann O, et al: Mycoplasma infection alters cancer stem cell properties in vitro. Stem Cell Rev Rep. 12:156–161. 2016. View Article : Google Scholar : PubMed/NCBI

187 

Hanahan D and Weinberg RA: The hallmarks of cancer. Cell. 100:57–70. 2000. View Article : Google Scholar : PubMed/NCBI

188 

Hanahan D: Hallmarks of cancer: New dimensions. Cancer Discov. 12:31–46. 2022. View Article : Google Scholar : PubMed/NCBI

189 

Franco OE, Shaw AK, Strand DW and Hayward SW: Cancer associated fibroblasts in cancer pathogenesis. Semin Cell Dev Biol. 21:33–39. 2010. View Article : Google Scholar : PubMed/NCBI

190 

Spaeth EL, Dembinski JL, Sasser AK, Watson K, Klopp A, Hall B, Andreeff M and Marini F: Mesenchymal stem cell transition to tumor-associated fibroblasts contributes to fibrovascular network expansion and tumor progression. PLoS One. 4:e49922009. View Article : Google Scholar : PubMed/NCBI

191 

Erez N, Truitt M, Olson P, Arron ST and Hanahan D: Cancer-associated fibroblasts are activated in incipient neoplasia to orchestrate tumor-promoting inflammation in an NF-kappaB-dependent manner. Cancer Cell. 17:135–147. 2010. View Article : Google Scholar : PubMed/NCBI

192 

Straussman R, Morikawa T, Shee K, Barzily-Rokni M, Qian ZR, Du J, Davis A, Mongare MM, Gould J, Frederick DT, et al: Tumour micro-environment elicits innate resistance to RAF inhibitors through HGF secretion. Nature. 487:500–504. 2012. View Article : Google Scholar : PubMed/NCBI

193 

Zhou Z, Zhou Q, Wu X, Xu S, Hu X, Tao X, Li B, Peng J, Li D, Shen L, et al: VCAM-1 secreted from cancer-associated fibroblasts enhances the growth and invasion of lung cancer cells through AKT and MAPK signaling. Cancer Lett. 473:62–73. 2020. View Article : Google Scholar : PubMed/NCBI

194 

Li F, Zhao S, Guo T, Li J and Gu C: The nutritional cytokine leptin promotes NSCLC by activating the PI3K/AKT and MAPK/ERK pathways in NSCLC cells in a paracrine manner. Biomed Res Int. 2019:25857432019.PubMed/NCBI

195 

Folkman J, Watson K, Ingber D and Hanahan D: Induction of angiogenesis during the transition from hyperplasia to neoplasia. Nature. 339:58–61. 1989. View Article : Google Scholar : PubMed/NCBI

196 

Hanahan D and Folkman J: Patterns and emerging mechanisms of the angiogenic switch during tumorigenesis. Cell. 86:353–364. 1996. View Article : Google Scholar : PubMed/NCBI

197 

Butler JM, Kobayashi H and Rafii S: Instructive role of the vascular niche in promoting tumour growth and tissue repair by angiocrine factors. Nat Rev Cancer. 10:138–146. 2010. View Article : Google Scholar : PubMed/NCBI

198 

Balkwill F, Charles KA and Mantovani A: Smoldering and polarized inflammation in the initiation and promotion of malignant disease. Cancer Cell. 7:211–217. 2005. View Article : Google Scholar : PubMed/NCBI

199 

Lu P, Takai K, Weaver VM and Werb Z: Extracellular matrix degradation and remodeling in development and disease. Cold Spring Harb Perspect Biol. 3:a0050582011. View Article : Google Scholar : PubMed/NCBI

200 

Pontiggia O, Sampayo R, Raffo D, Motter A, Xu R, Bissell MJ, Joffé EB and Simian M: The tumor microenvironment modulates tamoxifen resistance in breast cancer: A role for soluble stromal factors and fibronectin through β1 integrin. Breast Cancer Res Treat. 133:459–471. 2012. View Article : Google Scholar : PubMed/NCBI

201 

Mohamed MM and Sloane BF: Cysteine cathepsins: Multifunctional enzymes in cancer. Nat Rev Cancer. 6:764–775. 2006. View Article : Google Scholar : PubMed/NCBI

202 

Xu R, Boudreau A and Bissell MJ: Tissue architecture and function: Dynamic reciprocity via extra- and intra-cellular matrices. Cancer Metastasis Rev. 28:167–176. 2009. View Article : Google Scholar : PubMed/NCBI

203 

Goulet CR and Pouliot F: TGFβ signaling in the tumor microenvironment. Adv Exp Med Biol. 1270:89–105. 2021. View Article : Google Scholar : PubMed/NCBI

204 

Hou W, Kaczorowski A, Lantwin P, Kippenberger M, Schütz V, Franke D, Dieffenbacher SC, Hohenfellner M and Duensing S: Microenvironment-derived FGF-2 stimulates renal cell carcinoma cell proliferation through modulation of p27Kip1: Implications for spatial niche formation and functional intratumoral heterogeneity. Pathobiology. 87:114–124. 2020. View Article : Google Scholar : PubMed/NCBI

205 

Zou F, Zhang ZH, Zhang YT, Zhao JQ, Zhang XL, Wen CL, Song XY and Zhou WM: Cancer-associated-fibroblasts regulate the chemoresistance of lung cancer cell line A549 via SDF-1 secretion. Zhonghua Zhong Liu Za Zhi. 39:339–343. 2017.(In Chinese). PubMed/NCBI

206 

Wang H, Huang H, Wang L, Liu Y, Wang M, Zhao S, Lu G and Kang X: Cancer-associated fibroblasts secreted miR-103a-3p suppresses apoptosis and promotes cisplatin resistance in non-small cell lung cancer. Aging (Albany NY). 13:14456–14468. 2021. View Article : Google Scholar : PubMed/NCBI

207 

Sun X and Chen Z: Cancer-associated fibroblast-derived CCL5 contributes to cisplatin resistance in A549 NSCLC cells partially through upregulation of lncRNA HOTAIR expression. Oncol Lett. 22:6962021. View Article : Google Scholar : PubMed/NCBI

208 

Tao L, Huang G, Wang R, Pan Y, He Z, Chu X, Song H and Chen L: Cancer-associated fibroblasts treated with cisplatin facilitates chemoresistance of lung adenocarcinoma through IL-11/IL-11R/STAT3 signaling pathway. Sci Rep. 6:384082016. View Article : Google Scholar : PubMed/NCBI

209 

Bian L, Sun X, Jin K and He Y: Oral cancer-associated fibroblasts inhibit heat-induced apoptosis in Tca8113 cells through upregulated expression of Bcl-2 through the Mig/CXCR3 axis. Oncol Rep. 28:2063–2068. 2012. View Article : Google Scholar : PubMed/NCBI

210 

Daenen LG, Shaked Y, Man S, Xu P, Voest EE, Hoffman RM, Chaplin DJ and Kerbel RS: Low-dose metronomic cyclophosphamide combined with vascular disrupting therapy induces potent antitumor activity in preclinical human tumor xenograft models. Mol Cancer Ther. 8:2872–2881. 2009. View Article : Google Scholar : PubMed/NCBI

211 

Chen Q, Zhang XHF and Massagué J: Macrophage binding to receptor VCAM-1 transmits survival signals in breast cancer cells that invade the lungs. Cancer Cell. 20:538–549. 2011. View Article : Google Scholar : PubMed/NCBI

212 

Yang C, He L, He P, Liu Y, Wang W, He Y, Du Y and Gao F: Increased drug resistance in breast cancer by tumor-associated macrophages through IL-10/STAT3/bcl-2 signaling pathway. Med Oncol. 32:3522015. View Article : Google Scholar : PubMed/NCBI

213 

Harley CB, Kim NW, Prowse KR, Weinrich SL, Hirsch KS, West MD, Bacchetti S, Hirte HW, Counter CM, Greider CW, et al: Telomerase, cell immortality, and cancer. Cold Spring Harb Symp Quant Biol. 59:307–315. 1994. View Article : Google Scholar : PubMed/NCBI

214 

Jäger K and Walter M: Therapeutic targeting of telomerase. Genes (Basel). 7:392016. View Article : Google Scholar : PubMed/NCBI

215 

Ge Y, Wu S, Xue Y, Tao J, Li F, Chen Y, Liu H, Ma W, Huang J and Zhao Y: Preferential extension of short telomeres induced by low extracellular pH. Nucleic Acids Res. 44:8086–8096. 2016. View Article : Google Scholar : PubMed/NCBI

216 

Li S, Jiang Y, Li A, Liu X, Xing X, Guo Y, Xu Y, Hao Y and Zheng C: Telomere length is positively associated with the expression of IL-6 and MIP-1α in bone marrow mesenchymal stem cells of multiple myeloma. Mol Med Rep. 16:2497–2504. 2017. View Article : Google Scholar : PubMed/NCBI

217 

Lin EY, Li JF, Bricard G, Wang W, Deng Y, Sellers R, Porcelli SA and Pollard JW: Vascular endothelial growth factor restores delayed tumor progression in tumors depleted of macrophages. Mol Oncol. 1:288–302. 2007. View Article : Google Scholar : PubMed/NCBI

218 

Bergers G, Brekken R, McMahon G, Vu TH, Itoh T, Tamaki K, Tanzawa K, Thorpe P, Itohara S, Werb Z and Hanahan D: Matrix metalloproteinase-9 triggers the angiogenic switch during carcinogenesis. Nat Cell Biol. 2:737–744. 2000. View Article : Google Scholar : PubMed/NCBI

219 

Coussens LM, Raymond WW, Bergers G, Laig-Webster M, Behrendtsen O, Werb Z, Caughey GH and Hanahan D: Inflammatory mast cells up-regulate angiogenesis during squamous epithelial carcinogenesis. Genes Dev. 13:1382–1397. 1999. View Article : Google Scholar : PubMed/NCBI

220 

Khazaie K, Blatner NR, Khan MW, Gounari F, Gounaris E, Dennis K, Bonertz A, Tsai FN, Strouch MJ, Cheon E, et al: The significant role of mast cells in cancer. Cancer Metastasis Rev. 30:45–60. 2011. View Article : Google Scholar : PubMed/NCBI

221 

Huang B, Huang M and Li Q: Cancer-associated fibroblasts promote angiogenesis of hepatocellular carcinoma by VEGF-mediated EZH2/VASH1 pathway. Technol Cancer Res Treat. 18:15330338198799052019. View Article : Google Scholar : PubMed/NCBI

222 

Räsänen K and Vaheri A: Activation of fibroblasts in cancer stroma. Exp Cell Res. 316:2713–2722. 2010. View Article : Google Scholar : PubMed/NCBI

223 

Huizer K, Zhu C, Chirifi I, Krist B, Zorgman D, van der Weiden M, van den Bosch TPP, Dumas J, Cheng C, Kros JM and Mustafa DA: Periostin is expressed by pericytes and is crucial for angiogenesis in glioma. J Neuropathol Exp Neurol. 79:863–872. 2020. View Article : Google Scholar : PubMed/NCBI

224 

Xian X, Håkansson J, Ståhlberg A, Lindblom P, Betsholtz C, Gerhardt H and Semb H: Pericytes limit tumor cell metastasis. J Clin Invest. 116:642–651. 2006. View Article : Google Scholar : PubMed/NCBI

225 

Branco-Price C, Zhang N, Schnelle M, Evans C, Katschinski DM, Liao D, Ellies L and Johnson RS: Endothelial cell HIF-1α and HIF-2α differentially regulate metastatic success. Cancer Cell. 21:52–65. 2012. View Article : Google Scholar : PubMed/NCBI

226 

Navarro R, Tapia-Galisteo A, Martín-García L, Tarín C, Corbacho C, Gómez-López G, Sánchez-Tirado E, Campuzano S, González-Cortés A, Yáñez-Sedeño P, et al: TGF-β-induced IGFBP-3 is a key paracrine factor from activated pericytes that promotes colorectal cancer cell migration and invasion. Mol Oncol. 14:2609–2628. 2020. View Article : Google Scholar : PubMed/NCBI

227 

Pirilä E, Ramamurthy NS, Sorsa T, Salo T, Hietanen J and Maisi P: Gelatinase A (MMP-2), collagenase-2 (MMP-8), and laminin-5 gamma2-chain expression in murine inflammatory bowel disease (ulcerative colitis). Dig Dis Sci. 48:93–98. 2003. View Article : Google Scholar : PubMed/NCBI

228 

Vasiljeva O, Papazoglou A, Krüger A, Brodoefel H, Korovin M, Deussing J, Augustin N, Nielsen BS, Almholt K, Bogyo M, et al: Tumor cell-derived and macrophage-derived cathepsin B promotes progression and lung metastasis of mammary cancer. Cancer Res. 66:5242–5250. 2006. View Article : Google Scholar : PubMed/NCBI

229 

Balkwill F: Tumour necrosis factor and cancer. Nat Rev Cancer. 9:361–371. 2009. View Article : Google Scholar : PubMed/NCBI

230 

Abraham S, Zhang W, Greenberg N and Zhang M: Maspin functions as tumor suppressor by increasing cell adhesion to extracellular matrix in prostate tumor cells. J Urol. 169:1157–1161. 2003. View Article : Google Scholar : PubMed/NCBI

231 

Gorden DL, Fingleton B, Crawford HC, Jansen DE, Lepage M and Matrisian LM: Resident stromal cell-derived MMP-9 promotes the growth of colorectal metastases in the liver microenvironment. Int J Cancer. 121:495–500. 2007. View Article : Google Scholar : PubMed/NCBI

232 

Labelle M, Begum S and Hynes RO: Direct signaling between platelets and cancer cells induces an epithelial-mesenchymal-like transition and promotes metastasis. Cancer Cell. 20:576–590. 2011. View Article : Google Scholar : PubMed/NCBI

233 

Chaffer CL and Weinberg RA: A perspective on cancer cell metastasis. Science. 331:1559–1564. 2011. View Article : Google Scholar : PubMed/NCBI

234 

Zhang XHF, Jin X, Malladi S, Zou Y, Wen YH, Brogi E, Smid M, Foekens JA and Massagué J: Selection of bone metastasis seeds by mesenchymal signals in the primary tumor stroma. Cell. 154:1060–1073. 2013. View Article : Google Scholar : PubMed/NCBI

235 

Duda DG, Duyverman AMMJ, Kohno M, Snuderl M, Steller EJA, Fukumura D and Jain RK: Malignant cells facilitate lung metastasis by bringing their own soil. Proc Natl Acad Sci USA. 107:21677–21682. 2010. View Article : Google Scholar : PubMed/NCBI

236 

Lappano R, Rigiracciolo DC, Belfiore A, Maggiolini M and De Francesco EM: Cancer associated fibroblasts: Role in breast cancer and potential as therapeutic targets. Expert Opin Ther Targets. 24:559–572. 2020. View Article : Google Scholar : PubMed/NCBI

237 

Onrust SV, Hartl PM, Rosen SD and Hanahan D: Modulation of L-selectin ligand expression during an immune response accompanying tumorigenesis in transgenic mice. J Clin Invest. 97:54–64. 1996. View Article : Google Scholar : PubMed/NCBI

238 

Fisher DT, Chen Q, Skitzki JJ, Muhitch JB, Zhou L, Appenheimer MM, Vardam TD, Weis EL, Passanese J, Wang WC, et al: IL-6 trans-signaling licenses mouse and human tumor microvascular gateways for trafficking of cytotoxic T cells. J Clin Invest. 121:3846–3859. 2011. View Article : Google Scholar : PubMed/NCBI

239 

Manzur M, Hamzah J and Ganss R: Modulation of the ‘blood-tumor’ barrier improves immunotherapy. Cell Cycle. 7:2452–2455. 2008. View Article : Google Scholar : PubMed/NCBI

240 

Turley SJ, Cremasco V and Astarita JL: Immunological hallmarks of stromal cells in the tumour microenvironment. Nat Rev Immunol. 15:669–682. 2015. View Article : Google Scholar : PubMed/NCBI

241 

Stover DG, Bierie B and Moses HL: A delicate balance: TGF-beta and the tumor microenvironment. J Cell Biochem. 101:851–861. 2007. View Article : Google Scholar : PubMed/NCBI

242 

DeNardo DG, Brennan DJ, Rexhepaj E, Ruffell B, Shiao SL, Madden SF, Gallagher WM, Wadhwani N, Keil SD, Junaid SA, et al: Leukocyte complexity predicts breast cancer survival and functionally regulates response to chemotherapy. Cancer Discov. 1:54–67. 2011. View Article : Google Scholar : PubMed/NCBI

243 

Qian BZ and Pollard JW: Macrophage diversity enhances tumor progression and metastasis. Cell. 141:39–51. 2010. View Article : Google Scholar : PubMed/NCBI

244 

Topalian SL, Drake CG and Pardoll DM: Targeting the PD-1/B7-H1(PD-L1) pathway to activate anti-tumor immunity. Curr Opin Immunol. 24:207–212. 2012. View Article : Google Scholar : PubMed/NCBI

245 

Curiel TJ, Coukos G, Zou L, Alvarez X, Cheng P, Mottram P, Evdemon-Hogan M, Conejo-Garcia JR, Zhang L, Burow M, et al: Specific recruitment of regulatory T cells in ovarian carcinoma fosters immune privilege and predicts reduced survival. Nat Med. 10:942–949. 2004. View Article : Google Scholar : PubMed/NCBI

246 

van der Vliet HJJ, Koon HB, Atkins MB, Balk SP and Exley MA: Exploiting regulatory T-cell populations for the immunotherapy of cancer. J Immunother. 30:591–595. 2007. View Article : Google Scholar : PubMed/NCBI

247 

Schmidt A, Oberle N and Krammer P: Molecular mechanisms of treg-mediated T cell suppression. Front Immunol. 3:512012. View Article : Google Scholar : PubMed/NCBI

248 

Kalyanaraman B: Teaching the basics of cancer metabolism: Developing antitumor strategies by exploiting the differences between normal and cancer cell metabolism. Redox Biol. 12:833–842. 2017. View Article : Google Scholar : PubMed/NCBI

249 

Hensley CT, Faubert B, Yuan Q, Lev-Cohain N, Jin E, Kim J, Jiang L, Ko B, Skelton R, Loudat L, et al: Metabolic heterogeneity in human lung tumors. Cell. 164:681–694. 2016. View Article : Google Scholar : PubMed/NCBI

250 

Yu TJ, Ma D, Liu YY, Xiao Y, Gong Y, Jiang YZ, Shao ZM, Hu X and Di GH: Bulk and single-cell transcriptome profiling reveal the metabolic heterogeneity in human breast cancers. Mol Ther. 29:2350–2365. 2021. View Article : Google Scholar : PubMed/NCBI

251 

Hao X, Ren Y, Feng M, Wang Q and Wang Y: Metabolic reprogramming due to hypoxia in pancreatic cancer: Implications for tumor formation, immunity, and more. Biomed Pharmacother. 141:1117982021. View Article : Google Scholar : PubMed/NCBI

252 

Kim J and DeBerardinis RJ: Mechanisms and implications of metabolic heterogeneity in cancer. Cell Metab. 30:434–446. 2019. View Article : Google Scholar : PubMed/NCBI

253 

Xiao Z, Dai Z and Locasale JW: Metabolic landscape of the tumor microenvironment at single cell resolution. Nat Commun. 10:37632019. View Article : Google Scholar : PubMed/NCBI

254 

Stadlbauer A, Oberndorfer S, Zimmermann M, Renner B, Buchfelder M, Heinz G, Doerfler A, Kleindienst A and Roessler K: Physiologic MR imaging of the tumor microenvironment revealed switching of metabolic phenotype upon recurrence of glioblastoma in humans. J Cereb Blood Flow Metab. 40:528–538. 2020. View Article : Google Scholar : PubMed/NCBI

255 

Gupta S, Roy A and Dwarakanath BS: Metabolic cooperation and competition in the tumor microenvironment: Implications for therapy. Front Oncol. 7:682017. View Article : Google Scholar : PubMed/NCBI

256 

Martinez-Outschoorn U, Sotgia F and Lisanti MP: Tumor microenvironment and metabolic synergy in breast cancers: Critical importance of mitochondrial fuels and function. Semin Oncol. 41:1952014. View Article : Google Scholar : PubMed/NCBI

257 

Ocaña MC, Martínez-Poveda B, Quesada AR and Medina MÁ: Metabolism within the tumor microenvironment and its implication on cancer progression: An ongoing therapeutic target. Med Res Rev. 39:70–113. 2019. View Article : Google Scholar : PubMed/NCBI

258 

Wang YA, Li XL, Mo YZ, Fan CM, Tang L, Xiong F, Guo C, Xiang B, Zhou M, Ma J, et al: Effects of tumor metabolic microenvironment on regulatory T cells. Mol Cancer. 17:1682018. View Article : Google Scholar : PubMed/NCBI

259 

Pavlides S, Whitaker-Menezes D, Castello-Cros R, Flomenberg N, Witkiewicz AK, Frank PG, Casimiro MC, Wang C, Fortina P, Addya S, et al: The reverse Warburg effect: Aerobic glycolysis in cancer associated fibroblasts and the tumor stroma. Cell Cycle. 8:3984–4001. 2009. View Article : Google Scholar : PubMed/NCBI

260 

Reina-Campos M, Moscat J and Diaz-Meco M: Metabolism shapes the tumor microenvironment. Curr Opin Cell Biol. 48:47–53. 2017. View Article : Google Scholar : PubMed/NCBI

261 

Dietl K, Renner K, Dettmer K, Timischl B, Eberhart K, Dorn C, Hellerbrand C, Kastenberger M, Kunz-Schughart LA, Oefner PJ, et al: Lactic acid and acidification inhibit TNF secretion and glycolysis of human monocytes. J Immunol. 184:1200–1209. 2010. View Article : Google Scholar : PubMed/NCBI

262 

Whitaker-Menezes D, Martinez-Outschoorn UE, Lin Z, Ertel A, Flomenberg N, Witkiewicz AK, Birbe RC, Howell A, Pavlides S, Gandara R, et al: Evidence for a stromal-epithelial ‘lactate shuttle’ in human tumors: MCT4 is a marker of oxidative stress in cancer-associated fibroblasts. Cell Cycle. 10:1772–1783. 2011. View Article : Google Scholar : PubMed/NCBI

263 

Alberts B, Johnson A, Lewis J, Raff M, Roberts K and Walter P: Molecular biology of the cell. 4th edition. New York: Garland Science; 2002

264 

Deberardinis RJ, Sayed N, Ditsworth D and Thompson CB: Brick by brick: Metabolism and tumor cell growth. Curr Opin Genet Dev. 18:54–61. 2008. View Article : Google Scholar : PubMed/NCBI

265 

Yoshida GJ: Metabolic reprogramming: the emerging concept and associated therapeutic strategies. J Exp Clin Cancer Res. 34:1112015. View Article : Google Scholar : PubMed/NCBI

266 

Koundouros N and Poulogiannis G: Reprogramming of fatty acid metabolism in cancer. Br J Cancer. 122:4–22. 2020. View Article : Google Scholar : PubMed/NCBI

267 

Michalopoulou E, Bulusu V and Kamphorst JJ: Metabolic scavenging by cancer cells: When the going gets tough, the tough keep eating. Br J Cancer. 115:635–640. 2016. View Article : Google Scholar : PubMed/NCBI

268 

Pavlova NN and Thompson CB: The emerging hallmarks of cancer metabolism. Cell Metab. 23:27–47. 2016. View Article : Google Scholar : PubMed/NCBI

269 

Dey P, Kimmelman AC and DePinho RA: Metabolic codependencies in the tumor microenvironment. Cancer Discov. 11:1067–1081. 2021. View Article : Google Scholar : PubMed/NCBI

270 

Bailey KM, Wojtkowiak JW, Hashim AI and Gillies RJ: Targeting the metabolic microenvironment of tumors. Adv Pharmacol. 65:63–107. 2012. View Article : Google Scholar : PubMed/NCBI

271 

Sormendi S and Wielockx B: Hypoxia pathway proteins as central mediators of metabolism in the tumor cells and their microenvironment. Front Immunol. 9:402018. View Article : Google Scholar : PubMed/NCBI

272 

Devic S: Warburg effect-a consequence or the cause of carcinogenesis? J Cancer. 7:817–822. 2016. View Article : Google Scholar : PubMed/NCBI

273 

Cairns RA: Drivers of the Warburg phenotype. Cancer J. 21:56–61. 2015. View Article : Google Scholar : PubMed/NCBI

274 

El Hassouni B, Granchi C, Vallés-Martí A, Supadmanaba IGP, Bononi G, Tuccinardi T, Funel N, Jimenez CR, Peters GJ, Giovannetti E and Minutolo F: The dichotomous role of the glycolytic metabolism pathway in cancer metastasis: Interplay with the complex tumor microenvironment and novel therapeutic strategies. Semin Cancer Biol. 60:238–248. 2020. View Article : Google Scholar : PubMed/NCBI

275 

Chang CH, Qiu J, O'Sullivan D, Buck MD, Noguchi T, Curtis JD, Chen Q, Gindin M, Gubin MM, van der Windt GJ, et al: Metabolic competition in the tumor microenvironment is a driver of cancer progression. Cell. 162:1229–1241. 2015. View Article : Google Scholar : PubMed/NCBI

276 

DeBerardinis RJ, Mancuso A, Daikhin E, Nissim I, Yudkoff M, Wehrli S and Thompson CB: Beyond aerobic glycolysis: Transformed cells can engage in glutamine metabolism that exceeds the requirement for protein and nucleotide synthesis. Proc Natl Acad Sci USA. 104:19345–19350. 2007. View Article : Google Scholar : PubMed/NCBI

277 

Carito V, Bonuccelli G, Martinez-Outschoorn UE, Whitaker-Menezes D, Caroleo MC, Cione E, Howell A, Pestell RG, Lisanti MP and Sotgia F: Metabolic remodeling of the tumor microenvironment: Migration factor (MSF) reprograms myofibroblasts toward lactate production, fueling anabolic tumor growthstimulating. Cell Cycle. 11:3403–3414. 2012. View Article : Google Scholar : PubMed/NCBI

278 

Manning BD and Toker A: AKT/PKB signaling: Navigating the network. Cell. 169:381–405. 2017. View Article : Google Scholar : PubMed/NCBI

279 

Dang CV: MYC on the path to cancer. Cell. 149:22–35. 2012. View Article : Google Scholar : PubMed/NCBI

280 

Stine ZE, Walton ZE, Altman BJ, Hsieh AL and Dang CV: MYC, metabolism, and cancer. Cancer Discov. 5:1024–1039. 2015. View Article : Google Scholar : PubMed/NCBI

281 

Osthus RC, Shim H, Kim S, Li Q, Reddy R, Mukherjee M, Xu Y, Wonsey D, Lee LA and Dang CV: Deregulation of glucose transporter 1 and glycolytic gene expression by c-Myc. J Biol Chem. 275:21797–21800. 2000. View Article : Google Scholar : PubMed/NCBI

282 

Zhao X, Petrashen AP, Sanders JA, Peterson AL and Sedivy JM: SLC1A5 glutamine transporter is a target of MYC and mediates reduced mTORC1 signaling and increased fatty acid oxidation in long-lived Myc hypomorphic mice. Aging Cell. 18:e129472019. View Article : Google Scholar : PubMed/NCBI

283 

Sasaki H, Shitara M, Yokota K, Hikosaka Y, Moriyama S, Yano M and Fujii Y: Overexpression of GLUT1 correlates with Kras mutations in lung carcinomas. Mol Med Rep. 5:599–602. 2012.PubMed/NCBI

284 

Ying H, Kimmelman AC, Lyssiotis CA, Hua S, Chu GC, Fletcher-Sananikone E, Locasale JW, Son J, Zhang H, Coloff JL, et al: Oncogenic Kras maintains pancreatic tumors through regulation of anabolic glucose metabolism. Cell. 149:656–670. 2012. View Article : Google Scholar : PubMed/NCBI

285 

Son J, Lyssiotis CA, Ying H, Wang X, Hua S, Ligorio M, Perera RM, Ferrone CR, Mullarky E, Shyh-Chang N, et al: Glutamine supports pancreatic cancer growth through a KRAS-regulated metabolic pathway. Nature. 496:101–105. 2013. View Article : Google Scholar : PubMed/NCBI

286 

Baek G, Tse YF, Hu Z, Cox D, Buboltz N, McCue P, Yeo CJ, White MA, DeBerardinis RJ, Knudsen ES and Witkiewicz AK: MCT4 defines a glycolytic subtype of pancreatic cancer with poor prognosis and unique metabolic dependencies. Cell Rep. 9:2233–2249. 2014. View Article : Google Scholar : PubMed/NCBI

287 

Dey P, Li J, Zhang J, Chaurasiya S, Strom A, Wang H, Liao WT, Cavallaro F, Denz P, Bernard V, et al: Oncogenic KRAS-driven metabolic reprogramming in pancreatic cancer cells utilizes cytokines from the tumor microenvironment. Cancer Discov. 10:608–625. 2020. View Article : Google Scholar : PubMed/NCBI

288 

Amendola CR, Mahaffey JP, Parker SJ, Ahearn IM, Chen WC, Zhou M, Court H, Shi J, Mendoza SL, Morten MJ, et al: KRAS4A directly regulates hexokinase 1. Nature. 576:482–486. 2019. View Article : Google Scholar : PubMed/NCBI

289 

Zhang Y and Commisso C: Macropinocytosis in cancer: A complex signaling network. Trends Cancer. 5:332–334. 2019. View Article : Google Scholar : PubMed/NCBI

290 

Commisso C, Davidson SM, Soydaner-Azeloglu RG, Parker SJ, Kamphorst JJ, Hackett S, Grabocka E, Nofal M, Drebin JA, Thompson CB, et al: Macropinocytosis of protein is an amino acid supply route in Ras-transformed cells. Nature. 497:633–637. 2013. View Article : Google Scholar : PubMed/NCBI

291 

Schwartzenberg-Bar-Yoseph F, Armoni M and Karnieli E: The tumor suppressor p53 down-regulates glucose transporters GLUT1 and GLUT4 gene expression. Cancer Res. 64:2627–2633. 2004. View Article : Google Scholar : PubMed/NCBI

292 

Wang L, Xiong H, Wu F, Zhang Y, Wang J, Zhao L, Guo X, Chang LJ, Zhang Y, You MJ, et al: Hexokinase 2-mediated Warburg effect is required for PTEN- and p53-deficiency-driven prostate cancer growth. Cell Rep. 8:1461–1474. 2014. View Article : Google Scholar : PubMed/NCBI

293 

Yahagi N, Shimano H, Matsuzaka T, Najima Y, Sekiya M, Nakagawa Y, Ide T, Tomita S, Okazaki H, Tamura Y, et al: p53 activation in adipocytes of obese mice. J Biol Chem. 278:25395–25400. 2003. View Article : Google Scholar : PubMed/NCBI

294 

Bensaad K, Tsuruta A, Selak MA, Vidal MN, Nakano K, Bartrons R, Gottlieb E and Vousden KH: TIGAR, a p53-inducible regulator of glycolysis and apoptosis. Cell. 126:107–120. 2006. View Article : Google Scholar : PubMed/NCBI

295 

Sonugür FG and Akbulut H: The role of tumor microenvironment in genomic instability of malignant tumors. Front Genet. 10:10632019. View Article : Google Scholar : PubMed/NCBI

296 

Nakamura H, Tanimoto K, Hiyama K, Yunokawa M, Kawamoto T, Kato Y, Yoshiga K, Poellinger L, Hiyama E and Nishiyama M: Human mismatch repair gene, MLH1, is transcriptionally repressed by the hypoxia-inducible transcription factors, DEC1 and DEC2. Oncogene. 27:4200–4209. 2008. View Article : Google Scholar : PubMed/NCBI

297 

Rodríguez-Jiménez FJ, Moreno-Manzano V, Lucas-Dominguez R and Sánchez-Puelles JM: Hypoxia causes downregulation of mismatch repair system and genomic instability in stem cells. Stem Cells. 26:2052–2062. 2008. View Article : Google Scholar : PubMed/NCBI

298 

Ren Z, Wang Z, Gu D, Ma H, Zhu Y, Cai M and Zhang J: Genome instability and long noncoding RNA reveal biomarkers for immunotherapy and prognosis and novel competing endogenous RNA mechanism in colon adenocarcinoma. Front Cell Dev Biol. 9:7404552021. View Article : Google Scholar : PubMed/NCBI

299 

Guo JN, Xia TY, Deng SH, Xue WN, Cui BB and Liu YL: Prognostic immunity and therapeutic sensitivity analyses based on differential genomic instability-associated LncRNAs in left- and right-sided colon adenocarcinoma. Front Mol Biosci. 8:6688882021. View Article : Google Scholar : PubMed/NCBI

300 

Fridlender ZG, Sun J, Kim S, Kapoor V, Cheng G, Ling L, Worthen GS and Albelda SM: Polarization of tumor-associated neutrophil phenotype by TGF-beta: ‘N1’ versus ‘N2’ TAN. Cancer Cell. 16:183–194. 2009. View Article : Google Scholar : PubMed/NCBI

301 

DeNardo DG, Barreto JB, Andreu P, Vasquez L, Tawfik D, Kolhatkar N and Coussens LM: CD4(+) T cells regulate pulmonary metastasis of mammary carcinomas by enhancing protumor properties of macrophages. Cancer Cell. 16:91–102. 2009. View Article : Google Scholar : PubMed/NCBI

302 

Schoppmann SF, Birner P, Stöckl J, Kalt R, Ullrich R, Caucig C, Kriehuber E, Nagy K, Alitalo K and Kerjaschki D: Tumor-associated macrophages express lymphatic endothelial growth factors and are related to peritumoral lymphangiogenesis. Am J Pathol. 161:947–956. 2002. View Article : Google Scholar : PubMed/NCBI

303 

Celis JE, Gromov P, Cabezón T, Moreira JM, Ambartsumian N, Sandelin K, Rank F and Gromova I: Proteomic characterization of the interstitial fluid perfusing the breast tumor microenvironment: A novel resource for biomarker and therapeutic target discovery. Mol Cell Proteomics. 3:327–344. 2004. View Article : Google Scholar : PubMed/NCBI

304 

Dirat B, Bochet L, Dabek M, Daviaud D, Dauvillier S, Majed B, Wang YY, Meulle A, Salles B, Le Gonidec S, et al: Cancer-associated adipocytes exhibit an activated phenotype and contribute to breast cancer invasion. Cancer Res. 71:2455–2465. 2011. View Article : Google Scholar : PubMed/NCBI

305 

Ershaid N, Sharon Y, Doron H, Raz Y, Shani O, Cohen N, Monteran L, Leider-Trejo L, Ben-Shmuel A, Yassin M, et al: NLRP3 inflammasome in fibroblasts links tissue damage with inflammation in breast cancer progression and metastasis. Nat Commun. 10:43752019. View Article : Google Scholar : PubMed/NCBI

306 

Wu X, Tao P, Zhou Q, Li J, Yu Z, Wang X, Li J, Li C, Yan M, Zhu Z, et al: IL-6 secreted by cancer-associated fibroblasts promotes epithelial-mesenchymal transition and metastasis of gastric cancer via JAK2/STAT3 signaling pathway. Oncotarget. 8:20741–20750. 2017. View Article : Google Scholar : PubMed/NCBI

307 

Davidson S, Efremova M, Riedel A, Mahata B, Pramanik J, Huuhtanen J, Kar G, Vento-Tormo R, Hagai T, Chen X, et al: Single-cell RNA sequencing reveals a dynamic stromal niche that supports tumor growth. Cell Rep. 31:1076282020. View Article : Google Scholar : PubMed/NCBI

308 

Erez N, Glanz S, Raz Y, Avivi C and Barshack I: Cancer associated fibroblasts express pro-inflammatory factors in human breast and ovarian tumors. Biochem Biophys Res Commun. 437:397–402. 2013. View Article : Google Scholar : PubMed/NCBI

309 

De Monte L, Reni M, Tassi E, Clavenna D, Papa I, Recalde H, Braga M, Di Carlo V, Doglioni C and Protti MP: Intratumor T helper type 2 cell infiltrate correlates with cancer-associated fibroblast thymic stromal lymphopoietin production and reduced survival in pancreatic cancer. J Exp Med. 208:469–478. 2011. View Article : Google Scholar : PubMed/NCBI

310 

Bejarano L, Jordāo MJC and Joyce JA: Therapeutic targeting of the tumor microenvironment. Cancer Discov. 11:933–959. 2021. View Article : Google Scholar : PubMed/NCBI

311 

Lieu CH, Tan AC, Leong S, Diamond JR and Eckhardt SG: From bench to bedside: Lessons learned in translating preclinical studies in cancer drug development. J Natl Cancer Inst. 105:1441–1456. 2013. View Article : Google Scholar : PubMed/NCBI

312 

Syed M, Flechsig P, Liermann J, Windisch P, Staudinger F, Akbaba S, Koerber SA, Freudlsperger C, Plinkert PK, Debus J, et al: Fibroblast activation protein inhibitor (FAPI) PET for diagnostics and advanced targeted radiotherapy in head and neck cancers. Eur J Nucl Med Mol Imaging. 47:2836–2845. 2020. View Article : Google Scholar : PubMed/NCBI

313 

Chen H, Pang Y, Wu J, Zhao L, Hao B, Wu J, Wei J, Wu S, Zhao L, Luo Z, et al: Comparison of [68Ga]Ga-DOTA-FAPI-04 and [18F] FDG PET/CT for the diagnosis of primary and metastatic lesions in patients with various types of cancer. Eur J Nucl Med Mol Imaging. 47:1820–1832. 2020. View Article : Google Scholar : PubMed/NCBI

314 

Melero I, Sanmamed MF, Calvo E, Moreno I, Moreno V, Hernandez Guerrero TC, Martinez-Garcia M, Rodriguez-Vida A, Tabernero J, Azaro Pedrazzoli AB, et al: 1025MO first-in-human (FIH) phase I study of RO7122290 (RO), a novel FAP-targeted 4-1BB agonist, administered as single agent and in combination with atezolizumab (ATZ) to patients with advanced solid tumours. Ann Oncol. 31 (Suppl 4):S7072020. View Article : Google Scholar

315 

Sounni NE and Noel A: Targeting the tumor microenvironment for cancer therapy. Clin Chem. 59:85–93. 2013. View Article : Google Scholar : PubMed/NCBI

316 

Lamberts LE, Koch M, de Jong JS, Adams ALL, Glatz J, Kranendonk MEG, Terwisscha van Scheltinga AGT, Jansen L, de Vries J, Lub-de Hooge MN, et al: Tumor-specific uptake of fluorescent bevacizumab-IRDye800CW microdosing in patients with primary breast cancer: A phase I feasibility study. Clin Cancer Res. 23:2730–2741. 2017. View Article : Google Scholar : PubMed/NCBI

317 

Harlaar NJ, Koller M, de Jongh SJ, van Leeuwen BL, Hemmer PH, Kruijff S, van Ginkel RJ, Been LB, de Jong JS, Kats-Ugurlu G, et al: Molecular fluorescence-guided surgery of peritoneal carcinomatosis of colorectal origin: A single-centre feasibility study. Lancet Gastroenterol Hepatol. 1:283–290. 2016. View Article : Google Scholar : PubMed/NCBI

318 

Nakano K, Funauchi Y, Hayakawa K, Tanizawa T, Ae K, Matsumoto S and Takahashi S: Relative dose intensity of induction-phase pazopanib treatment of soft tissue sarcoma: Its relationship with prognoses of pazopanib responders. J Clin Med. 8:602019. View Article : Google Scholar : PubMed/NCBI

319 

Noda S, Yoshida T, Hira D, Murai R, Tomita K, Tsuru T, Kageyama S, Kawauchi A, Ikeda Y, Morita SY and Terada T: Exploratory investigation of target pazopanib concentration range for patients with renal cell carcinoma. Clin Genitourin Cancer. 17:e306–e313. 2019. View Article : Google Scholar : PubMed/NCBI

320 

Wang S, Lu J, You Q, Huang H, Chen Y and Liu K: The mTOR/AP-1/VEGF signaling pathway regulates vascular endothelial cell growth. Oncotarget. 7:53269–53276. 2016. View Article : Google Scholar : PubMed/NCBI

321 

Komohara Y, Fujiwara Y, Ohnishi K and Takeya M: Tumor-associated macrophages: Potential therapeutic targets for anti-cancer therapy. Adv Drug Deliv Rev. 99:180–185. 2016. View Article : Google Scholar : PubMed/NCBI

322 

Bak SP, Walters JJ, Takeya M, Conejo-Garcia JR and Berwin BL: Scavenger receptor-A-targeted leukocyte depletion inhibits peritoneal ovarian tumor progression. Cancer Res. 67:4783–4789. 2007. View Article : Google Scholar : PubMed/NCBI

323 

Nagai T, Tanaka M, Tsuneyoshi Y, Xu B, Michie SA, Hasui K, Hirano H, Arita K and Matsuyama T: Targeting tumor-associated macrophages in an experimental glioma model with a recombinant immunotoxin to folate receptor beta. Cancer Immunol Immunother. 58:1577–1586. 2009. View Article : Google Scholar : PubMed/NCBI

324 

Naser R, Dilabazian H, Bahr H, Barakat A and El-Sibai M: A guide through conventional and modern cancer treatment modalities: A specific focus on glioblastoma cancer therapy (review). Oncol Rep. 48:1902022. View Article : Google Scholar : PubMed/NCBI

325 

Liu Y and Zheng P: Preserving the CTLA-4 checkpoint for safer and more effective cancer immunotherapy. Trends Pharmacol Sci. 41:4–12. 2020. View Article : Google Scholar : PubMed/NCBI

326 

Tauriello DVF, Palomo-Ponce S, Stork D, Berenguer-Llergo A, Badia-Ramentol J, Iglesias M, Sevillano M, Ibiza S, Cañellas A, Hernando-Momblona X, et al: TGFβ drives immune evasion in genetically reconstituted colon cancer metastasis. Nature. 554:538–543. 2018. View Article : Google Scholar : PubMed/NCBI

327 

Greten FR, Arkan MC, Bollrath J, Hsu LC, Goode J, Miething C, Göktuna SI, Neuenhahn M, Fierer J, Paxian S, et al: NF-kappaB is a negative regulator of IL-1beta secretion as revealed by genetic and pharmacological inhibition of IKKbeta. Cell. 130:918–931. 2007. View Article : Google Scholar : PubMed/NCBI

328 

Marsh JL, Jackman CP, Tang SN, Shankar S and Srivastava RK: Embelin suppresses pancreatic cancer growth by modulating tumor immune microenvironment. Front Biosci (Landmark Ed). 19:113–125. 2014. View Article : Google Scholar : PubMed/NCBI

329 

Xu CL, Zheng B, Pei JH, Shen SJ and Wang JZ: Embelin induces apoptosis of human gastric carcinoma through inhibition of p38 MAPK and NF-κB signaling pathways. Mol Med Rep. 14:307–312. 2016. View Article : Google Scholar : PubMed/NCBI

330 

Waldmann TA: Cytokines in cancer immunotherapy. Cold Spring Harb Perspect Biol. 10:a0284722018. View Article : Google Scholar : PubMed/NCBI

331 

Rosenberg SA, Restifo NP, Yang JC, Morgan RA and Dudley ME: Adoptive cell transfer: A clinical path to effective cancer immunotherapy. Nat Rev Cancer. 8:299–308. 2008. View Article : Google Scholar : PubMed/NCBI

332 

Redeker A and Arens R: Improving adoptive T cell therapy: The particular role of T cell costimulation, cytokines, and post-transfer vaccination. Front Immunol. 7:3452016. View Article : Google Scholar : PubMed/NCBI

333 

Chruściel E, Urban-Wójciuk Z, Arcimowicz Ł, Kurkowiak M, Kowalski J, Gliwiński M, Marjański T, Rzyman W, Biernat W, Dziadziuszko R, et al: Adoptive cell therapy-harnessing antigen-specific T cells to target solid tumours. Cancers (Basel). 12:6832020. View Article : Google Scholar : PubMed/NCBI

334 

Benmebarek MR, Karches CH, Cadilha BL, Lesch S, Endres S and Kobold S: Killing mechanisms of chimeric antigen receptor (CAR) T cells. Int J Mol Sci. 20:12832019. View Article : Google Scholar : PubMed/NCBI

335 

Brown CE, Badie B, Barish ME, Weng L, Ostberg JR, Chang WC, Naranjo A, Starr R, Wagner J, Wright C, et al: Bioactivity and safety of IL13Rα2-redirected chimeric antigen receptor CD8+ T cells in patients with recurrent glioblastoma. Clin Cancer Res. 21:4062–4072. 2015. View Article : Google Scholar : PubMed/NCBI

336 

Xu S, Tang L, Li X, Fan F and Liu Z: Immunotherapy for glioma: Current management and future application. Cancer Lett. 476:1–12. 2020. View Article : Google Scholar : PubMed/NCBI

337 

Liu J, Fu M, Wang M, Wan D, Wei Y and Wei X: Cancer vaccines as promising immuno-therapeutics: Platforms and current progress. J Hematol Oncol. 15:282022. View Article : Google Scholar : PubMed/NCBI

338 

Keskin DB, Anandappa AJ, Sun J, Tirosh I, Mathewson ND, Li S, Oliveira G, Giobbie-Hurder A, Felt K, Gjini E, et al: Neoantigen vaccine generates intratumoral T cell responses in phase Ib glioblastoma trial. Nature. 565:234–239. 2019. View Article : Google Scholar : PubMed/NCBI

339 

Hilf N, Kuttruff-Coqui S, Frenzel K, Bukur V, Stevanović S, Gouttefangeas C, Platten M, Tabatabai G, Dutoit V, van der Burg SH, et al: Actively personalized vaccination trial for newly diagnosed glioblastoma. Nature. 565:240–245. 2019. View Article : Google Scholar : PubMed/NCBI

340 

Hollingsworth RE and Jansen K: Turning the corner on therapeutic cancer vaccines. NPJ Vaccines. 4:72019. View Article : Google Scholar : PubMed/NCBI

341 

Rüttinger D, Winter H, van den Engel NK, Hatz R, Jauch KW, Fox BA and Weber JS: Immunotherapy of cancer: Key findings and commentary on the third tegernsee conference. Oncologist. 15:112–118. 2010. View Article : Google Scholar : PubMed/NCBI

342 

AIVITA, Biomedical Inc., . AIVITA biomedical's phase 2 glioblastoma trial shows improved progression free survival. 2021.

343 

Busby J, McMenamin Ú, Spence A, Johnston BT, Hughes C and Cardwell CR: Angiotensin receptor blocker use and gastro-oesophageal cancer survival: A population-based cohort study. Aliment Pharmacol Ther. 47:279–288. 2018. View Article : Google Scholar : PubMed/NCBI

344 

Murphy JE, Wo JY, Ryan DP, Clark JW, Jiang W, Yeap BY, Drapek LC, Ly L, Baglini CV, Blaszkowsky LS, et al: Total neoadjuvant therapy with FOLFIRINOX in combination with losartan followed by chemoradiotherapy for locally advanced pancreatic cancer: A phase 2 clinical trial. JAMA Oncol. 5:1020–1027. 2019. View Article : Google Scholar : PubMed/NCBI

345 

Fennell DA, Baas P, Taylor P, Nowak AK, Gilligan D, Nakano T, Pachter JA, Weaver DT, Scherpereel A, Pavlakis N, et al: Maintenance defactinib versus placebo after first-line chemotherapy in patients with merlin-stratified pleural mesothelioma: COMMAND-A double-blind, randomized, phase II study. J Clin Oncol. 37:790–798. 2019. View Article : Google Scholar : PubMed/NCBI

346 

Wang-Gillam A: Targeting stroma: A tale of caution. J Clin Oncol. 37:1041–1043. 2019. View Article : Google Scholar : PubMed/NCBI

347 

Ulisse S, Baldini E, Sorrenti S and D'Armiento M: The urokinase plasminogen activator system: A target for anti-cancer therapy. Curr Cancer Drug Targets. 9:32–71. 2009. View Article : Google Scholar : PubMed/NCBI

348 

Feng S, Agoulnik IU, Bogatcheva NV, Kamat AA, Kwabi-Addo B, Li R, Ayala G, Ittmann MM and Agoulnik AI: Relaxin promotes prostate cancer progression. Clin Cancer Res. 13:1695–1702. 2007. View Article : Google Scholar : PubMed/NCBI

349 

Raue R, Frank AC, Syed SN and Brüne B: Therapeutic targeting of MicroRNAs in the tumor microenvironment. Int J Mol Sci. 22:22102021. View Article : Google Scholar : PubMed/NCBI

350 

Huang S, Zhang J, Lai X, Zhuang L and Wu J: Identification of novel tumor microenvironment-related long noncoding RNAs to determine the prognosis and response to immunotherapy of hepatocellular carcinoma patients. Front Mol Biosci. 8:7813072021. View Article : Google Scholar : PubMed/NCBI

351 

Ting NLN, Lau HCH and Yu J: Cancer pharmacomicrobiomics: Targeting microbiota to optimise cancer therapy outcomes. Gut. 71:1412–1425. 2022. View Article : Google Scholar : PubMed/NCBI

352 

Spanogiannopoulos P, Bess EN, Carmody RN and Turnbaugh PJ: The microbial pharmacists within us: A metagenomic view of xenobiotic metabolism. Nat Rev Microbiol. 14:273–287. 2016. View Article : Google Scholar : PubMed/NCBI

353 

Imai H, Saijo K, Komine K, Otsuki Y, Ohuchi K, Sato Y, Okita A, Takahashi M, Takahashi S, Shirota H, et al: Antibiotic therapy augments the efficacy of gemcitabine-containing regimens for advanced cancer: A retrospective study. Cancer Manag Res. 11:7953–7965. 2019. View Article : Google Scholar : PubMed/NCBI

354 

Nakano S, Komatsu Y, Kawamoto Y, Saito R, Ito K, Nakatsumi H, Yuki S and Sakamoto N: Association between the use of antibiotics and efficacy of gemcitabine plus nab-paclitaxel in advanced pancreatic cancer. Medicine (Baltimore). 99:e222502020. View Article : Google Scholar : PubMed/NCBI

355 

Sunakawa Y, Arai H, Izawa N, Mizukami T, Horie Y, Doi A, Hirakawa M, Ogura T, Tsuda T and Nakajima TE: Antibiotics may enhance the efficacy of gemcitabine treatment for advanced pancreatic cancer. Ann Oncol. 29 (Suppl 8):viii251–viii252. 2018. View Article : Google Scholar

356 

Frankel AE, Coughlin LA, Kim J, Froehlich TW, Xie Y, Frenkel EP and Koh AY: Metagenomic shotgun sequencing and unbiased metabolomic profiling identify specific human gut microbiota and metabolites associated with immune checkpoint therapy efficacy in melanoma patients. Neoplasia. 19:848–855. 2017. View Article : Google Scholar : PubMed/NCBI

357 

Routy B, Le Chatelier E, Derosa L, Duong CPM, Alou MT, Daillère R, Fluckiger A, Messaoudene M, Rauber C, Roberti MP, et al: Gut microbiome influences efficacy of PD-1-based immunotherapy against epithelial tumors. Science. 359:91–97. 2018. View Article : Google Scholar : PubMed/NCBI

358 

Matson V, Fessler J, Bao R, Chongsuwat T, Zha Y, Alegre ML, Luke JJ and Gajewski TF: The commensal microbiome is associated with anti-PD-1 efficacy in metastatic melanoma patients. Science. 359:104–108. 2018. View Article : Google Scholar : PubMed/NCBI

359 

Gopalakrishnan V, Spencer CN, Nezi L, Reuben A, Andrews MC, Karpinets TV, Prieto PA, Vicente D, Hoffman K, Wei SC, et al: Gut microbiome modulates response to anti-PD-1 immunotherapy in melanoma patients. Science. 359:97–103. 2018. View Article : Google Scholar : PubMed/NCBI

360 

Vétizou M, Pitt JM, Daillère R, Lepage P, Waldschmitt N, Flament C, Rusakiewicz S, Routy B, Roberti MP, Duong CP, et al: Anticancer immunotherapy by CTLA-4 blockade relies on the gut microbiota. Science. 350:1079–1084. 2015. View Article : Google Scholar : PubMed/NCBI

361 

Sivan A, Corrales L, Hubert N, Williams JB, Aquino-Michaels K, Earley ZM, Benyamin FW, Lei YM, Jabri B, Alegre ML, et al: Commensal bifidobacterium promotes antitumor immunity and facilitates anti-PD-L1 efficacy. Science. 350:1084–1089. 2015. View Article : Google Scholar : PubMed/NCBI

362 

Khalesi S, Bellissimo N, Vandelanotte C, Williams S, Stanley D and Irwin C: A review of probiotic supplementation in healthy adults: Helpful or hype? Eur J Clin Nutr. 73:24–37. 2019. View Article : Google Scholar : PubMed/NCBI

363 

Pushalkar S, Hundeyin M, Daley D, Zambirinis CP, Kurz E, Mishra A, Mohan N, Aykut B, Usyk M, Torres LE, et al: The pancreatic cancer microbiome promotes oncogenesis by induction of innate and adaptive immune suppression. Cancer Discov. 8:403–416. 2018. View Article : Google Scholar : PubMed/NCBI

364 

Yu T, Guo F, Yu Y, Sun T, Ma D, Han J, Qian Y, Kryczek I, Sun D, Nagarsheth N, et al: Fusobacterium nucleatum promotes chemoresistance to colorectal cancer by modulating autophagy. Cell. 170:548–563.e16. 2017. View Article : Google Scholar : PubMed/NCBI

365 

Zheng JH, Nguyen VH, Jiang SN, Park SH, Tan W, Hong SH, Shin MG, Chung IJ, Hong Y, Bom HS, et al: Two-step enhanced cancer immunotherapy with engineered Salmonella typhimurium secreting heterologous flagellin. Sci Transl Med. 9:eaak95372017. View Article : Google Scholar : PubMed/NCBI

366 

Juul FE, Garborg K, Bretthauer M, Skudal H, Øines MN, Wiig H, Rose Ø, Seip B, Lamont JT, Midtvedt T, et al: Fecal microbiota transplantation for primary clostridium difficile infection. N Engl J Med. 378:2535–2536. 2018. View Article : Google Scholar : PubMed/NCBI

367 

van Nood E, Vrieze A, Nieuwdorp M, Fuentes S, Zoetendal EG, de Vos WM, Visser CE, Kuijper EJ, Bartelsman JF, Tijssen JG, et al: Duodenal infusion of donor feces for recurrent clostridium difficile. N Engl J Med. 368:407–415. 2013. View Article : Google Scholar : PubMed/NCBI

368 

Arbel LT, Hsu E and McNally K: Cost-effectiveness of fecal microbiota transplantation in the treatment of recurrent clostridium difficile infection: A literature review. Cureus. 9:e15992017.PubMed/NCBI

369 

Bayat Mokhtari R, Homayouni TS, Baluch N, Morgatskaya E, Kumar S, Das B and Yeger H: Combination therapy in combating cancer. Oncotarget. 8:38022–38043. 2017. View Article : Google Scholar : PubMed/NCBI

370 

Lane D: Designer combination therapy for cancer. Nat Biotechnol. 24:163–164. 2006. View Article : Google Scholar : PubMed/NCBI

371 

Allen E, Jabouille A, Rivera LB, Lodewijckx I, Missiaen R, Steri V, Feyen K, Tawney J, Hanahan D, Michael IP and Bergers G: Combined antiangiogenic and anti-PD-L1 therapy stimulates tumor immunity through HEV formation. Sci Transl Med. 9:eaak96792017. View Article : Google Scholar : PubMed/NCBI

372 

Aparicio LMA, Fernandez IP and Cassinello J: Tyrosine kinase inhibitors reprogramming immunity in renal cell carcinoma: Rethinking cancer immunotherapy. Clin Transl Oncol. 19:1175–1182. 2017. View Article : Google Scholar : PubMed/NCBI

373 

Duchnowska R, Loibl S and Jassem J: Tyrosine kinase inhibitors for brain metastases in HER2-positive breast cancer. Cancer Treat Rev. 67:71–77. 2018. View Article : Google Scholar : PubMed/NCBI

374 

Ntanasis-Stathopoulos I, Fotopoulos G, Tzanninis IG and Kotteas EA: The emerging role of tyrosine kinase inhibitors in ovarian cancer treatment: A systematic review. Cancer Invest. 34:313–339. 2016. View Article : Google Scholar : PubMed/NCBI

375 

Bożyk A, Wojas-Krawczyk K, Krawczyk P and Milanowski J: Tumor microenvironment-a short review of cellular and interaction diversity. Biology (Basel). 11:9292022.PubMed/NCBI

376 

Russo M and Nastasi C: Targeting the tumor microenvironment: A close up of tumor-associated macrophages and neutrophils. Front Oncol. 12:8715132022. View Article : Google Scholar : PubMed/NCBI

377 

Mao D, Xu R, Chen H, Chen X, Li D, Song S, He Y, Wei Z and Zhang C: Cross-talk of focal adhesion-related gene defines prognosis and the immune microenvironment in gastric cancer. Front Cell Dev Biol. 9:7164612021. View Article : Google Scholar : PubMed/NCBI

378 

Qian H, Li H, Xie J, Lu X, Li F, Wang W, Tang X, Shi M, Jiang L, Li H, et al: Immunity-related gene signature identifies subtypes benefitting from adjuvant chemotherapy or potentially responding to PD1/PD-L1 blockage in pancreatic cancer. Front Cell Dev Biol. 9:6822612021. View Article : Google Scholar : PubMed/NCBI

379 

Zhou S, Sun Y, Chen T, Wang J, He J, Lyu J, Shen Y, Chen X and Yang R: The landscape of the tumor microenvironment in skin cutaneous melanoma reveals a prognostic and immunotherapeutically relevant gene signature. Front Cell Dev Biol. 9:7395942021. View Article : Google Scholar : PubMed/NCBI

380 

Wu J, Zhou J, Xu Q, Foley R..Guo J, Zhang X, Tian C, Mu M, Xing Y, Liu Y, et al: Identification of key genes driving tumor associated macrophage migration and polarization based on immune fingerprints of lung adenocarcinoma. Front Cell Dev Biol. 9:7518002021. View Article : Google Scholar : PubMed/NCBI

381 

Haibe Y, Kreidieh M, El Hajj H, Khalifeh I, Mukherji D, Temraz S and Shamseddine A: Resistance mechanisms to anti-angiogenic therapies in cancer. Front Oncol. 10:2212020. View Article : Google Scholar : PubMed/NCBI

382 

Navab R, Strumpf D, To C, Pasko E, Kim KS, Park CJ, Hai J, Liu J, Jonkman J, Barczyk M, et al: Integrin α11β1 regulates cancer stromal stiffness and promotes tumorigenicity and metastasis in non-small cell lung cancer. Oncogene. 35:1899–1908. 2016. View Article : Google Scholar : PubMed/NCBI

383 

Junttila MR and de Sauvage FJ: Influence of tumour micro-environment heterogeneity on therapeutic response. Nature. 501:346–354. 2013. View Article : Google Scholar : PubMed/NCBI

384 

Giuliano S and Pagès G: Mechanisms of resistance to anti-angiogenesis therapies. Biochimie. 95:1110–1119. 2013. View Article : Google Scholar : PubMed/NCBI

385 

Flaherty KT, Manola JB, Pins M, McDermott DF, Atkins MB, Dutcher JJ, George DJ, Margolin KA and DiPaola RS: BEST: A randomized phase II study of vascular endothelial growth factor, RAF kinase, and mammalian target of rapamycin combination targeted therapy with bevacizumab, sorafenib, and temsirolimus in advanced renal cell carcinoma-a trial of the ECOG-ACRIN cancer research group (E2804). J Clin Oncol. 33:2384–2391. 2015. View Article : Google Scholar : PubMed/NCBI

386 

Wang J, Li Y, Nie G and Zhao Y: Precise design of nanomedicines: Perspectives for cancer treatment. Natl Sci Rev. 6:1107–1110. 2019. View Article : Google Scholar : PubMed/NCBI

387 

Atat OE, Farzaneh Z, Pourhamzeh M, Taki F, Abi-Habib R, Vosough M and El-Sibai M: 3D modeling in cancer studies. Hum Cell. 35:23–36. 2022. View Article : Google Scholar : PubMed/NCBI

388 

Selek L, Seigneuret E, Nugue G, Wion D, Nissou MF, Salon C, Seurin MJ, Carozzo C, Ponce F, Roger T and Berger F: Imaging and histological characterization of a human brain xenograft in pig: the first induced glioma model in a large animal. J Neurosci Methods. 221:159–165. 2014. View Article : Google Scholar : PubMed/NCBI

389 

Khoshnevis M, Carozzo C, Bonnefont-Rebeix C, Belluco S, Leveneur O, Chuzel T, Pillet-Michelland E, Dreyfus M, Roger T, Berger F and Ponce F: Development of induced glioblastoma by implantation of a human xenograft in Yucatan minipig as a large animal model. J Neurosci Methods. 282:61–68. 2017. View Article : Google Scholar : PubMed/NCBI

390 

Khoshnevis M, Carozzo C, Brown R, Bardiès M, Bonnefont-Rebeix C, Belluco S, Nennig C, Marcon L, Tillement O, Gehan H, et al: Feasibility of intratumoral 165Holmium siloxane delivery to induced U87 glioblastoma in a large animal model, the Yucatan minipig. PLoS One. 15:e02347722020. View Article : Google Scholar : PubMed/NCBI

391 

Mackenzie NJ, Nicholls C, Templeton AR, Perera MP, Jeffery PL, Zimmermann K, Kulasinghe A, Kenna TJ, Vela I, Williams ED and Thomas PB: Modelling the tumor immune microenvironment for precision immunotherapy. Clin Transl Immunology. 11:e14002022. View Article : Google Scholar : PubMed/NCBI

392 

Mendes N, Dias Carvalho P, Martins F, Mendonça S, Malheiro AR, Ribeiro A, Carvalho J and Velho S: Animal models to study cancer and its microenvironment. Adv Exp Med Biol. 1219:389–401. 2020. View Article : Google Scholar : PubMed/NCBI

393 

Hsu JF, Chu SM, Liao CC, Wang CJ, Wang YS, Lai MY, Wang HC, Huang HR and Tsai MH: Nanotechnology and nanocarrier-based drug delivery as the potential therapeutic strategy for glioblastoma multiforme: An update. Cancers (Basel). 13:1952021. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

February-2023
Volume 62 Issue 2

Print ISSN: 1019-6439
Online ISSN:1791-2423

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Naser R, Fakhoury I, El-Fouani A, Abi-Habib R and El-Sibai M: Role of the tumor microenvironment in cancer hallmarks and targeted therapy (Review). Int J Oncol 62: 23, 2023
APA
Naser, R., Fakhoury, I., El-Fouani, A., Abi-Habib, R., & El-Sibai, M. (2023). Role of the tumor microenvironment in cancer hallmarks and targeted therapy (Review). International Journal of Oncology, 62, 23. https://doi.org/10.3892/ijo.2022.5471
MLA
Naser, R., Fakhoury, I., El-Fouani, A., Abi-Habib, R., El-Sibai, M."Role of the tumor microenvironment in cancer hallmarks and targeted therapy (Review)". International Journal of Oncology 62.2 (2023): 23.
Chicago
Naser, R., Fakhoury, I., El-Fouani, A., Abi-Habib, R., El-Sibai, M."Role of the tumor microenvironment in cancer hallmarks and targeted therapy (Review)". International Journal of Oncology 62, no. 2 (2023): 23. https://doi.org/10.3892/ijo.2022.5471