Skip to main content

The roles of metabolic profiles and intracellular signaling pathways of tumor microenvironment cells in angiogenesis of solid tumors

Abstract

Innate and adaptive immune cells patrol and survey throughout the human body and sometimes reside in the tumor microenvironment (TME) with a variety of cell types and nutrients that may differ from those in which they developed. The metabolic pathways and metabolites of immune cells are rooted in cell physiology, and not only provide nutrients and energy for cell growth and survival but also influencing cell differentiation and effector functions. Nowadays, there is a growing awareness that metabolic processes occurring in cancer cells can affect immune cell function and lead to tumor immune evasion and angiogenesis. In order to safely treat cancer patients and prevent immune checkpoint blockade-induced toxicities and autoimmunity, we suggest using anti-angiogenic drugs solely or combined with Immune checkpoint blockers (ICBs) to boost the safety and effectiveness of cancer therapy. As a consequence, there is significant and escalating attention to discovering techniques that target metabolism as a new method of cancer therapy. In this review, a summary of immune-metabolic processes and their potential role in the stimulation of intracellular signaling in TME cells that lead to tumor angiogenesis, and therapeutic applications is provided.

Video abstract

Introduction

Both developed and developing nations continue to be heavily burdened economically and socially by cancer. There are expected to be 19.3 million new cancer cases worldwide in 2020 (excluding nonmelanoma skin cancer) as well as almost 10 million deaths due to cancer (9.9 million excluding nonmelanoma skin cancer) [1]. Therefore, understanding how cancer develops and progresses is imperative for developing interventions aimed at promoting the well-being of cancer patients [2, 3]. So, in order to better understand cancer, we should review the literature on the tumor microenvironment (TME) [46]. In this line, metabolic modifications and physiological processes play a crucial role in the development of resistance to immune checkpoint inhibitors (ICIs) in cancerous cells [7]. Fermentation/anaerobic glycolysis or "Warburg effect" is one of these metabolism modifications producing anabolic precursors, such as lactate, required by rapidly dividing embryonic tissues and tumors despite its low ATP yield/glucose molecule. Also, increased ketone bodies, branched-chain amino acids, and other toxic metabolites produced due to dysfunction of key enzymes of TCA cycle called “oncometabolites” affect the TME cells to promote angiogenesis and tumor growth as well as hypoxia [8]. In response to hypoxia, hypoxia-inducible factor-1 (HIF-1) alpha stabilizes the vascular endothelial growth factor (VEGF) A promoter and activates its gene expression [9]. Upregulated VEGF and numerous signaling pathways activated by HIF-1 promote angiogenesis and tumor growth through several strategies. Moreover, increased lactate levels by upregulating lactate dehydrogenase (LDH) and monocarboxylate transporters (MCTs) via anaerobic glycolysis, acidifying the TME facilitating angiogenesis, tumor growth, and drug resistance through several mechanisms and signaling pathways [10]. So, understanding these TME-modified metabolic pathways and other modifications, such as hypoxia, in TME and solid tumor cells could be beneficial for finding new and potential therapeutic strategies that will be discussed in this review.

Tumor angiogenesis process

Some diseases including cancer, diabetic retinopathy, and rheumatoid arthritis are associated with angiogenesis, regardless of physiological conditions [11]. Angiogenesis is related to tumor growth and metastasis [12]. There are several steps in Angiogenesis, such as the separation of endothelial cells from pericytes and the basement membrane, invasion and migration across the basement membrane, and, finally, an extension of the angiogenesis into the tumor [13]. In this regard, several factors induce angiogenesis, including VEGF, angiopoietins, transforming growth factors (TGF), platelet-derived growth factor (PDGF), tumor necrosis factor-α (TNF-α), interleukins, and the members of the fibroblast growth factor (FGF) family [14, 15]. Of these, the tumor-secreted cytokine VEGF family has a crucial role in both normal and tumor-induced angiogenesis [16]. VEGF A, B, C, and E can interact with VEGF receptor-1 (VEGFR-1), VEGFR-2 on vascular endothelial cells (ECs) and neurons. Hematopoietic stem cells (HSCs), monocytes, and osteoblasts can also be stimulated by VEGF-A. It can also induce the production of nitric oxide (NO) causing vasodilation [17, 18]. Additionally, Angiopoietin-2 (Ang-2) is a downstream target of VEGF signaling [19]. It should be mentioned that NO and Ang-2 act as angiogenic switches [19, 20].

Moreover, in response to the activation of VEGFR1, different signaling pathways are activated, including phosphoinositide-3-kinase (PI3K)/protein kinase B (PKB/Akt), mitogen-activated protein kinase (MAPK)/extracellular signal-regulated kinase (ERK) pathway (p38-MAPK/ERK1/2) [21, 22] facilitating the migration of inflammatory cells, the release of inflammatory cytokines, and the release of proteolytic enzymes into the extracellular matrix (ECM) [23, 24]. As a result, endothelial cells are proliferated through activation of phospholipase-Cγ (PLCγ)/protein kinase C (PKC) and Ras/ Raf/ERK/MAPK due to the binding of VEGF to the extracellular domain of VEGFR-2 [21, 25,26,27].

Ang-2 attaches to Tie-2 receptors on endothelial and leukemia cells [28,29,30]. The receptor activates Src homology 2 containing tyrosine phosphatase protein (SHP2), growth factor receptor-bound protein 7 (GRB7), and focal adhesion kinase (FAK) which in turn promotes cell survival and migration [31]. On another hand, competition between Ang-1 and Ang-2 for the Tie-2 receptor causes Ang-1-mediated stabilization to be blocked by Ang-2 [32].

VEGF-A, Ang-2, and MMPs also, begin to destabilize pre-existing capillaries, which prepares the capillaries for sprouting endothelial cells. The tips of the vessels are endothelial cells that sprout filopodia-like extensions from the primary vessel. Two cell surface proteins called Delta-like 4 (Dll4) and Notch with its ligands, regulate the number and activity of the cells in the tip [33]. So, the Notch/Dll4 pathway is responsible for regulating vessel sprouting and maturation. In response to increasing the Dll4, the basic fibroblast growth factor (FGF), VEGF-A, hepatocyte growth factor (HGF) and VEGFR-1 levels increase [34].

At last, in order for a vessel to function properly, it must be stabilized, and this requires endothelial cell-to-cell contact or interactions between endothelial cells and pericytes. This interaction is promoted by Ang-1 and PDGF-A, -B, -C, -D [9]. In turn, tumors overexpressing Ang-1 and PDGF fail to promote blood vessel maturation as well as resist the effects of antiangiogenic treatment and chemotherapy. Although, the data suggest that tumor microenvironments contain more destabilizing vascular factors than Ang-1 and PDGF [35]. Notably, Endothelial cells secrete a large amount of Ang-2 in response to hypoxia or VEGF-A, which prevents the normalization of blood vessels. In a study, inhibiting Ang-2 and VEGF synergically increased pericyte coverage, VE-cadherin tight junction, and decreased permeability [36]. MMPs are also one of the factors that exert their destabilizing activity through pericyte detachment, cell–cell adhesion cleavage, and degradation of ECM [37].

Innate and adaptive immune cells in TME that lead to tumor angiogenesis

There are many mechanisms that contribute to tumor angiogenesis and immunosuppression in TME [38]. Several laboratory studies have demonstrated that stromal cells, such as fibroblasts and myeloid cells, can promote tumor angiogenesis via expressing various pro-angiogenic factors, such as Bv8/PROK2, members of the VEGF, FGF, PDGF, and angiopoietin families [39,40,41,42,43,44]. TME myeloid cells produce increased fatty acid synthase in response to CSF1, which leads to the expression of PPARβ/δ-dependent genes, such as VEGF, arginase1 (Arg1), and IL-10 contributing to angiogenesis and immunosuppression [45]. Tumor-associated macrophage (TAM) and tumor-associated neutrophil (TAN) have pro-tumor activities through extracellular matrix (ECM) remodeling, enhanced invasion and metastasis of cancer cells, angiogenesis, cancer cell proliferation, lymphangiogenesis, and inhibition of anti-tumor immune surveillance [46]. Therefore, the investigation of TME cells is crucial to our understanding of tumor angiogenesis, and to enhancing the effectiveness of cancer therapy (Fig. 1.).

Fig. 1
figure 1

Contribution of innate and adaptive immune cells in tumor microenvironment to tumor angiogenesis. Within the tumor microenvironment, soluble mediators (cytokines, chemokines, and enzymes) exert their role directly as proangiogenic factors expressed by M2-like tumor-associated macrophage (TAM), myeloid-derived-suppressor cell (MDSC), N2-like tumor-associated neutrophil (TAN), natural killer (NK) cells, mast cells, and dendritic cell (DC), cancer-associated fibroblast (CAF), Tie2-expressing monocytes (TEM), eosinophil, Innate lymphoid cells (ILCs), and T cells (regulatory T cell, γδT17)

Natural killer (NK) cells

According to the pro-tumorigenic phenotype of cancer infiltrating NK cells, they can secrete pro-angiogenic factors, such as VEGF, transforming growth factor-beta (TGF-β), IL-8, IL-10, placental growth factor (PlGF), Ang-1, and Ang-2 [47]. The most relevant seems to be VEGF. Accordingly, a loss of HIF-1α in NK cells increased the bioavailability of VEGF by reducing the infiltration of NK cells through the VEGF receptor-1 (VEGFR-1) [48]. Additionally, zoledronic acid enhances NK cell activity on VEGF by synergizing with IL-2 [49]. With regard to the reports, in non-small cell lung cancers (NSCLC), the levels of VEGF, IL-8, and PlGF synthesized by NK cells are higher than in controls [50]. In a study of colorectal cancer patients, Bruno et al. [50] demonstrated that NK cells express matrix metalloproteinase-2 (MMP-2), and tissue inhibitor of MMP (TIMP), and angiogenin. Additionally, STAT3/STAT5 activation has been found in tumor-associated NK cells (TANKs) [51], and treatment with a STAT5 inhibitor called pimozide reduced the ability of endothelial cells to produce VEGF [51].

Dendritic cells (DCs)

Dendritic cells (DCs) are generally divided into two subgroups: myeloid DC (MDCs) and plasmacytoid DC (PDCs). MDCs within the bone marrow are immature DCs with high phagocytic capability. MDCs, also called conventional DCs (cDCs), are composed of a number of distinct subsets of cells with potent antigen-presenting capacities [52], playing a crucial role in the activation of T cell responses in response to pathogens and tumor cells [53]. Despite this, DCs in the TME containing tumor-associated cDCs or regulatory DCs (regDCs) exhibit altered functions with impaired cross-presentation capacity, express low levels of co-stimulatory molecules, as well as have high-proangiogenic activity. During tumor progression, these changes depend on various conditions, such as hypoxia, production of PGE2, IL-10, adenosine, and lactate level increase [54]. There is evidence that soluble factors derived from tumor cells may interfere with this maturation process and hinder the development of mature DCs [55]. The accumulation of immature DC in tumors is consistent with this, as only very few mature MDC are observed in tumors [55]. Immature DC can be recruited to the TME by tumor-derived factors such as VEGF, HGF, b-defensin, CXCL8, and CXCL12 [55, 56]. Angiogenic cytokines released by tumor-associated DCs, including VEGF, TGF-β, TNF-α, IL-8, and osteopontin, directly contribute to tumor angiogenesis [54, 56, 57]. According to the studies, tumors could attract PDCs to enhance angiogenesis while excluding MDCs to inhibit angiogenesis, demonstrating a novel mechanism for modulating tumor neovascularization [56]. Thus, blocking tumor-associated DCs that stimulate angiogenesis in TME may be a potential strategy in cancer therapy.

Mast cells (MCs)

According to the mast cells (MCs) pro-tumorigenic phenotype, they produce several pro-angiogenic factors, such as VEGF, FGF-2, TGF-β, IL-8, TNF-α, and nerve growth factor (NGF). A correlation between MCs and VEGF with angiogenesis has been shown in laryngeal squamous cell carcinoma and in lung cancer [58,59,60,61]. Moreover, as part of angiogenic factors stored in mast cell granules, two proteases, namely tryptase and chymase, play an important role in the angiogenic responses of MCs [62]. Tryptase stimulates endothelial cell proliferation, degrades connective tissue matrix, promotes vascular tube formation in vitro, and activates matrix metalloproteinases (MMPs) and plasminogen activator (PA), degrading the extracellular matrix, thereby releasing FGF-2 or VEGF from matrix-bound cells [62]. Several hematological and solid tumors, such as breast cancer [63], melanoma, colon-rectal cancer, uterine cervix cancer, and pulmonary adenocarcinoma [64], as well as vascular tumors, such as haemangioma and haemangioblastoma, have been shown to have an increased number of MCs associated with angiogenesis. Hence, the accumulation of MCs leads to an increase in neovascularization, mast cell VEGF and FGF-2 expression, tumor aggressiveness, and poor prognosis [62]. Also, mast cells produce TIMPs, which have a role in the regulation of extracellular matrix degradation, allowing the secretion of angiogenic factors [62].

Myeloid-derived suppressor cells (MDSC)

The immune-suppressive TME causes the appearance of the phenotype and functional alterations of various players, such as Myeloid-derived suppressor cells (MDSCs). MDSCs are capable of invading directly into the tumor endothelium. They also secrete many pro-angiogenic factors. In TME, MDSCs are able to infiltrate tumor tissues and also differentiate into TAM [65]. In addition, MDSCs have a key role in the production of MMPs, chemoattractants, and the creation of pre-metastatic environments that contribute to cancer invasion, metastasis, and angiogenesis [65]. MDSCs can lead to the induction of an immune-suppressive environment [66] and angiogenesis directly as well as indirectly (by interacting with several components of innate and adaptive immunity) [67]. MDSCs are capable of boosting angiogenesis and promoting tumor neovasculature, through the production of high levels of MMPs, such as MMP2, MMP8, MMP9, MMP13, and MMP14. Additionally, MDSCs can also stimulate tumor growth and blood vessel growth [67, 68]. As a result of a recent study, it has demonstrated that MDSCs with the production of high levels of MMP9 stimulate VEGF function through boosting its bioavailability [69]. As part of the regulation of angiogenesis, Bombina variegate peptide 8 (Bv8) and vascular endothelial growth factor (EG-VEGF), and TGF-β are the key molecules. Of these, Bv8 is also capable of recruiting MDSCs to tumor tissues [54, 70]. Recruitment of MDSCs is mediated by chemokines and chemoattractants. Most solid tumors show necrotic regions induced by hypoxia. HIF-1α (hypoxia-inducible factor 1α), as a transcription factor, is expressed under hypoxic conditions. HIF1α overexpression can trigger tumor cells to secrete chemoattractants, including stromal-derived factor 1α (SDF-1α or CXCL12), CXCL5, and CCL2. To recruit MDSCs, these ligands could bind to the receptors on MDSCs [65]. Also, Sorrentino et al. and Iannone et al. observed that MDSCs triggered adenosine receptor A2B-induced VEGF production, vessel density, and angiogenic activity [71, 72]. Moreover, MDSCs have crosstalk with NK cells. MDSCs can also inhibit the anti-tumor responses of NK cells, increase angiogenesis [67], establish pre-metastatic niches [73], and recruit other immunosuppressive cells [74]. It has been found that MDSCs significantly reduce NK cell cytotoxicity in breast cancer, resulting in an increase in metastatic potential [75]. High ROS levels (radical oxygen species), created in cancerous conditions, exert a crucial role in stimulating the MDSCs and VEGF receptor expression on the MDSCs and their expansion in the TME [67].

Tumor associated macrophages (TAM)

The most frequent immune cells of the tumor microenvironment (TME) are tumor-associated macrophages (TAMs). The macrophages displayed different phenotypes based on the microenvironment they populated. The two main types of activated macrophages are classical activated macrophages (M1) and alternatively activated macrophages (M2) [76]. Macrophages of type M1 promote inflammatory responses against pathogens and tumor cells, while macrophages of type M2 have anti-inflammatory properties that promote wound healing and tumor progression [77]. Increasing evidence points to the fact that TAMs contribute to tumor progression via multiple mechanisms. TAMs have been found to release pro-angiogenic factors, such as VEGF-A, EGF, PlGF, PDGF, FGF, HGF, TGF-β, TNF-α, IL-1β, IL-8, CCL2, CXCL8, and CXCL12 [54, 76,77,78]. TAMs contribute to tumor progression within the TME through cross-talking with other leukocytes, inflammatory and stromal cells. Also, TAMs can directly recruit T regulatory cells (Treg) by secreting CCL20 and CCL22 chemokines and they can activate them by secreting IL-10 and TGF-β [54]. On the other hand, several transcription factors are involved in M1/M2 polarization: NF-κB, STAT1, and IRF5 involved in M1 polarization, whereas MYC, STAT6, KLF4, IRF4, and PPARγ are associated with M2 polarization [79]. Studies have shown that hypoxic TME polarizes macrophages into the M2 phenotype and TAMs have a significant role in inducing tumor angiogenesis [80]. In this line, expression of HIF-1α in TAMs induces VEGF-A production [77]. Werno et al. revealed that HIF-1α expression in macrophages plays a key role in tumor angiogenesis when breast cancer cells are co-cultured with wild-type or HIF-1α knockout macrophages [81]. Through the production of growth factors, chemokines, and cytokines, TAMs produce an immunosuppressive TME that inhibits anti-tumor responses [77]. Moreover, they function as angiogenesis promoters by the production of pro-angiogenic factors and MMPs such as MMP-1, MMP-7, MMP-9, and MMP-12, as well as vascular construction which supplies nutrients and oxygen to solid tumor cells [54]. Additionally, an ovarian cancer mouse model found that TAMs were major sources of MMP-9, and MMP9-producing TAMs, were positively related to tumor angiogenesis and tumor growth [82]. Through the production of pro-inflammatory mediators, such as EGFR family ligands, TNF-α, IL-6, IFN-γ, proteases, reactive oxygen species (ROS), and nitrogen species, TAMs create a mutagenic microenvironment. It has been observed that microvascular density correlates positively with TAMs and VEGF levels in mammary tumors [83]. Vascular neuropilin-1 (NRP-1) has been shown to be a variable receptor for two secreted glycoproteins, VEGF-A and Semaphorin 3A (Sema3A), but also its role as an adhesion receptor is poorly understood [84, 85]. As a response to Sema3A, NRP-1 plays a key role in the entry of TAMs into hypoxic niches, but the loss of NRP-1 promotes anti-tumor immunity and hinders angiogenesis [77]. In cervical cancer, NRP-1 plays a crucial role in hypoxic TME-induced activation and TAM-induced pro-tumoral effects [80]. According to biochemical studies, NRP-1 might have a role in VEGF-mediated induction of ERK, Akt, P38 MAPK, SRC, and p130 CAS pathways [84, 86]. So, NRP-1 may play roles in angiogenesis, which likely synergize with its known status as a co-receptor for VEGFR2 [84].

Quaranta et al. found that TAMs have a significant role in tumor-associated fibroblasts (TAFs) activation for producing an excessive amount of extracellular matrix (ECM), by secreting granulin [87]. Moreover, releasing urokinase-type plasminogen activator (UPA) and thymidine phosphorylase (TP) by TAMs stimulates tumor angiogenesis by increasing endothelial cells (ECs) migration, degradation of ECM, as well as a vascular invasion [88]. The M2 macrophages, also known as alternatively activated macrophages, can be further classified into M2a, M2b, M2c, and M2d. Through adenosine, the M2d phenotype can be induced in pro-inflammatory M1 macrophages via adenosine 2A receptor (A2AR) activation [89]. Indirectly IL-17 triggers differentiation of M2 macrophages through stimulation of the COX-2/PGE2 pathway in cancer cells [90]. Additionally, the studies suggest that FGF signaling may contribute to M2a-induced angiogenesis and PlGF signaling to M2c-induced angiogenesis, but that more research should clarify the exact mechanisms involved [91]. Therefore, M2d macrophage can produce VEGF and VEGF receptors [92] indicating its pro-angiogenic phenotype.

Tumor-associated neutrophils (TANs)

In the TME, it has been reported that macrophages and fibroblasts promote the growth of colorectal cancer. While neutrophils were originally thought to have defensive functions, it has been shown that some populations of neutrophils, called tumor-associated neutrophils (TANs), are cancer-supportive via TGF-β and interferon-β signaling controlling the plasticity between tumor-supportive and tumor-suppressive neutrophils [93].

TANs play an important role in tumor metastasis and angiogenesis due to their ability to release a variety of proangiogenic and immunosuppressive factors such as VEGF, IL-1β, TGF-α, FGF2, HGF, and Ang-1; as well as chemokines including CXCL1, CXCL10, CXCL9, CXCL8, CCL4, and CCL3; and enzymes involved in ECM remodeling (MMP-9) [54].

G-CSF (CSF3) and its receptor CSF3R are required for neutrophil proliferation and growth. STAT3, which is downstream of an activated CSF3R, is important for cancer inflammation. Neutrophils increase the expression of BV8 (prokineticin-2) via CSF3, which causes myeloid cell mobility and myeloid-dependent tumor angiogenesis. The activation of STAT3 is required for the synthesis of BV8 [54]. In addition, G-CSF, IL-6, VEGF, and IL1-β are among the cytokines produced by tumor and stromal cells that cause neutrophilia and make these neutrophils more suppressive. TANs are thought to circulate longer than other circulating neutrophils. Interactions between TANs and neoplastic cells result in the release of GM-CSF by the tumor cells. GM-CSF stimulates the release of Oncostatin M (OSM) and neutrophil synthesis. OSM is a member of the IL-6 family of cytokines that can enhance VEGF production via the Jak/STAT pathway [94].

It is believed that reactive oxygen species (ROS), reactive nitrogen species (RNS), and proteases released by neutrophils are involved in tumor initiation [95]. In tumor cells, neutrophil elastase (NE) can inhibit insulin receptor substrate-1 (IRS-1). A decrease in IRS-1 levels can enhance the interaction between PI3K and PDGF-R, a factor that promotes tumor cell proliferation [96]. In addition, through cyclooxygenase-2 (COX-2)-mediated prostaglandin E2 (PGE2) synthesis, it also increases tumor cell growth [97].

TANs also play a role in tumor invasion and angiogenesis in primary and metastatic sites by generating MMP9, VEGF, HGF, PAF, IL-10 [54, 93]. Modulating MMP-9 enhances angiogenesis by activating VEGF. It is reported that neutrophil-derived tissue inhibitors of metalloproteinase (TIMP)-free MMP-9 induce angiogenesis strongly [98]. These MMP-9 s generated by tumor-infiltrating neutrophils regulate tumor cell invasion and tumor angiogenesis at the same time [98]. Neutrophils also aid tumor cell spreading by trapping circulating tumor cells with neutrophil extracellular traps and enhancing their migration to distant locations [93].

Another factor is IL-8, which is a multifunctional cytokine secreted by neutrophils after cell activation [99]. Studies have shown IL-8 signaling promotes angiogenic responses in endothelial cells, increases the proliferation and survival of both cancer and endothelial cells, and stimulates the migration of the cancer cells, endothelial cells, and neutrophils [100]. A major mechanism of IL-8's biological effects is its binding to two cell-surface G protein-coupled receptors called CXCR1 and CXCR2 [101, 102]. Several studies have shown tumor cells overexpress IL-8 in response to chemotherapy or environmental stress, such as hypoxia. Given the presence of CXCR1 and CXCR2 receptors on endothelial cells, cancer cells, and TANs, increased IL-8 secretion from tumor cells has a broader significance for the TME [103].

Another factor that plays a vital role in TME modification and angiogenesis is HGF. HGF acts as a cell adhesion complex and indirectly increases the production of IL-8 and VEGF. The HGF role in cancer growth can be regulated by several signaling pathways, including the PI3K and MAPK pathways. Besides that, HGF also regulates metastasis and invasion [94].

So, according to the TANs' role in tumor metastasis and angiogenesis, some therapeutic methods targeting TANs were suggested, with two basic approaches: (a) targeting the CXCL-8/CXCR-1/CXCR-2 axis to block TANs, or (b) targeting substances released by polymorpho-nuclear cells that stimulate cancer growth [104].

Carcinoma-associated fibroblasts (CAFs)

Fibroblasts are primarily responsible for the synthesis of the extracellular matrix (ECM) and are not epithelial, vascular, or hematopoietic cells [105]. Although angiogenesis, ECM remodeling, and epithelial proliferation are adaptive for healing wounds, within the tumor microenvironment, they promote tumor growth and development [106].

In research by Orimo et al., fibroblasts were isolated from human breast carcinomas and normal breast fibroblasts were isolated from the same person. The fibroblasts were co-injected into nude mice with breast carcinoma cells, and carcinoma-associated fibroblasts (CAFs) significantly grew tumors more than normal fibroblasts did. It has been shown that this is due to the high levels of CXCL12 secreted by CAFs, which recruit endothelial progenitors to tumors and increase vascularization [107]. In addition, Yang et al. showed that CAFs from human prostate cancers also incited xenograft growth via connective tissue growth factor (CTGF). In the xenograft model, CTGF expression proved to be induced by TGF-β, and overexpression of CTGF in 3T3 fibroblasts led to enhanced microvessel density and tumor growth [108]. CAFs can also contribute to angiogenesis indirectly by releasing active growth factors from the ECM when they express MMPs. In this line, CAFs produce MMP-9 and MMP-13, both involved in angiogenesis. There is evidence that both MMP-9 and MMP-13 increase angiogenesis in tumors by releasing VEGF from the ECM [106]. Tumor vascularization was decreased in integrin α1 knock-out mice that lack integrin α1β1, a blocker of MMP synthesis. This was due to the increased production of angiostatin [109]. MMP-7 and MMP-9 also, act on circulating plasminogen to produce angiostatin. So, these factors indicate that MMPs play contradictory roles in angiogenesis [110].

In addition, CAFs produce more IL-6 than fibroblasts in normal tissues preventing tumor cell apoptosis through a STAT3-dependent mechanism [111] and enhancing angiogenesis [112].

Furthermore, HGFs are expressed by CAFs and they play a key role in angiogenesis. The CAFs produce angiogenic factors, such as VEGF, TGF-β1, EGF, Ang-1, Ang-2, PDGF, MMPs, and FDF, which are essential for hepatocellular carcinoma (HCC) initiation, progression, and metastatic development, as well as the growth of new vessels. When CAFs are activated, the VEGF receptor, the PDGF receptor, and the Tie-2 receptor are upregulated, which results in increased mitogenesis through VEGF [113]. Hormonal stimulation such as leptin, or physical stress like hypoxia, is known to induce VEGF secretion by hepatic stellate cells (HSCs) and has been found to be upregulated in HCC [113]. HCC cells' conditioned medium can activate CAFs and promote VEGF production by activating the Akt-VEGF pathway and subsequently, increasing the oxidative stress in hepatocellular carcinomas (HCC) promoting their potential malignancy [114].

Most believe that there is a strong link between tumor angiogenesis and notch signaling cascades, FGF, VEGF, and angiopoietin (ANGPT). Pro-angiogenic FGF2 stimulates proliferation and migration of endothelial cells directly through activation of FGFR1 (or FGFR2) as well as indirectly by inducing VEGF and ANGPT2 from endothelial cells. ANGPT1 is made by pericytes, and it induces Tie2 signaling, which controls endothelial quiescence or stability. ANGPT2 is secreted by endothelial cells and inhibits Tie2 signaling to promote endothelial activation and growth [115]. Dll4 expression is induced by VEGF signaling in endothelial tip cells, which then activates Notch signaling in endothelial stalk cells for vascular inactivation via downregulation of VEGFR [116]. In endothelial activation, VEGF, FGF2, and ANGPT2 participate, while ANGPT1 and Notch contribute to inactivity. To stimulate tumor angiogenesis in endothelial cells, the VEGFR2 and FGFR1/2 are two crucial receptors as tyrosine kinases (RTKs) [116]. So, a monoclonal antibody (mAb) or small-molecule VEGFR inhibitor is commonly used to target VEGF signaling in cancer patients. But some tumors fail to respond to the VEGF blockade treatment and others recur after it is stopped. One of the major mechanisms responsible for VEGF blockade therapy resistance is the activation of FGF signaling in endothelial cells [115]. Therefore, FGFR inhibitors may be effective in overcoming resistance to this therapy. Two options are available to block the dual signal cascade of FGF and VEGF. It is better to use monotherapy using small-molecule FGFR/VEGFR2 dual inhibitors such as AZD4547 and dovitinib in order to reduce medical costs, but combination therapy using anti-VEGF mAb and FGFR inhibitors may be better to prevent adverse effects. In order to optimize FGF/VEGF dual blockade therapy, safety issues, and medical costs must be considered [116]. While in the context of anti-angiogenic therapy targeting the VEGF pathway, there are adverse effects such as hypertension, bleeding, and thrombosis [117]. Multi-kinase inhibitors such as AZD4547, dovitinib, and ponatinib also, target FGFRs and other tyrosine kinases. Selective FGFR targeting is intended to lessen side effects, whereas the dual targeting of VEGFR/CSF1R and FGFR is predicted to increase anti-tumor effects indirectly by normalizing the TME [116].

Innate lymphoid cells (ILC)

The innate lymphoid cell (ILC), found mostly in solid tissues, is a family of mononuclear hematopoietic cells [54]. ILC family members include NK cells, Group-1 ILCs (ILC1), Group-2 ILCs (ILC2), Group-3 ILCs (ILC3), and lymphoid tissue inducer cells (LTis) [118, 119]. ILCs exhibit high cell plasticity and can easily be converted into different subtypes after exposure to TME stimuli [54]. Moreover, ILC1 can promote tumorigenesis when it is converted into NCR (NKp46, NKp44)+ ILC3 [120].

It is still debated whether ILCs play a role in cancer progression or prevention [54]. ILC2 releases type 2-cytokines such as IL-5 and IL-13; both stimulate angiogenesis [54]. By secreting IL-22, ILC3s support epithelial stability and maintain tissue homeostasis. ILC3s are known to produce IL-17 and CXCL12, which play a role in tumorigenesis, angiogenesis, and tumor growth. There is increasing evidence to suggest that ILC3s play a role in recruiting Treg cells and MDSCs to TME and promoting M2-like macrophages there [54]. CCL21 was used in a syngeneic 4T1.2 mouse breast model to attract ILC3s to primary tumors, which then induces tumor stromal cells to secrete CXCL13, and then results in lymphotoxin and activates receptor of NF-κB ligands, which stimulate tumor cell migration and lymphangiogenesis [121]. Breast cancer patients with invasive behavior also show an association with genes expressed by ILC3, such as CXCL13, CCL21, CCL19, CCR7, and CXCR5. Researchers have shown that ILC3 creates the tertiary lymphoid structures (TLS), which are responsible for tumor growth and lymph node metastasis. On the other hand, the tumor-preventing and tumor-promoting effects of TLS remain contested [54]. In inflamed tissue from patients who suffer from chronic obstructive pulmonary disease or smokers, NRP-1+ LTi-like ILC3s have been detected, which were associated with VEGF production [122]. Immunohistochemical studies of inflamed tissues indicated that RORγτ+NRP-1+ cells were associated with blood vessels as well as in the alveolar parenchyma, showing their role in angiogenesis and triggering of lung TLS. Aside from IL-22 and IL-17, the pro-inflammatory LTi-like NRP-1+ ILC3 subgroup was discovered to produce CSF2, TNF-α, B-cell activating factor, and CXCL8, all of which might lead to angiogenesis [54].

Thus, while the abilities of ILC3 may encourage tumor growth, neoangiogenesis, epithelial-mesenchymal transition, and metastasis; others may instead promote anti-tumor responses [120]. Researchers have also studied the effects of ILCs on tumor vessels in cancer suppression mediated by cytokines, such as IL-12 [123, 124]. It has been shown that IL-12 inhibits angiogenesis by interacting with NK cells in lymphomas [125]. IL-12-activated NK cells have the ability to cause endothelial cell cytotoxicity in vitro, which reduces tumor angiogenesis [125]. There is evidence that other populations of ILCs mediate IL-12's antitumor activity in melanoma. Eisenring et al. discovered that a group of IL-12-driven NKp46+ ILC3s causes overexpression of the adhesion molecules ICAM and VCAM, resulting in enhanced leukocyte infiltration and tumor control [123]. Tumor-infiltrating natural cytotoxicity receptor (NCR)+ ILC3 cells in NSCLC tissues also induced upregulation of these adhesion molecules [126]. Additionally, IL-17 produced by ILC3s may influence tumor vasculature. As a matter of fact, IL-17 stimulates angiogenic factors in stromal cells, including VEGF, TGF-β, and IL-8 [127]. Additionally, IL-17 increased blood vessel permeability and E-Selectin and VCAM-1 expression in lung endothelial cells, resulting in pulmonary metastasis [128]. So, a targeted strategy has yet to be developed for non-NK ILCs because of their recent discovery and incomplete understanding of their role in tumor growth and angiogenesis [54].

Eosinophils

Eosinophils characterization is the expression of CCR3 and CD125. Eosinophils have been found to increase in several human tumors, including gastrointestinal tumors, oral squamous cell carcinoma, nasopharyngeal carcinoma, and Hodgkin lymphoma [129]. In specimens of non-small cell lung carcinoma (NSCLC), the association between eosinophil density and angiogenesis was demonstrated, as well as their relationship to tumor stage [130]. It has been demonstrated that CCL11 (eotaxin), a highly potent and specific eosinophil chemoattractant, via binding to CCR3 is responsible for attracting eosinophils to the TME [131]. Additionally, eosinophil recruitment at tumor sites may lead to angiogenesis, because the secretory granules of eosinophils contain VEGF, which is rapidly secreted when activated by IL-15 [129]. However, the exact role exerted by eosinophils in the TME remains controversial.

Tie2-expressing monocytes (TEM)

As opposed to other monocyte populations, the newly discovered Tie2-expressing monocyte (TEM) expresses the Tie2 receptor for angiopoietin, an attribute that is unique to this population [132, 133] and various human tumor entities have been reported to contain TEM [134].

In this regard, angiopoietin-2 (ANGPT2), as the Tie2 ligand, plays an important role in regulating TEM recruitment [133, 134]. Tumor-infiltrating TEM has been demonstrated to accumulate in close proximity to blood vessels and to hypoxic areas of tumors [134, 135]. By localizing TEM near tumor blood vessels, sanctions would potentially affect the process of tumor angiogenesis. Moreover, studies show that selective removal of TEM from the TME significantly reduced angiogenesis and impaired glioma growth [135]. Interestingly, in spite of the fact that TEM numbers are lower than TAM and granulocytes within the tumor, TEM has a significant role in contributing to vessel neoformation, which suggests that TEM is a potent driver of tumor angiogenesis [135]. In recent studies, it has been found that TEM frequency correlated with angiogenesis in tumor tissues and may serve as a diagnostic marker for NSCLC [136], glioblastoma [137], HCC [138], and Renal Cell Carcinoma (RCC) [139]. In addition, Tie2 and VEGFR pathways play varying roles in TEM angiogenic and lymphangiogenic activities across breast cancer (BC) patients; nevertheless, a combination of Tie2 and VEGFR kinase inhibitors inhibited these activities and overcame inter-patient variability [140]. The expression of MMP-9, b-FGF and angiogenesis-modulating cytokines are the critical factors in transmitting angiogenic signals by TEM [55, 135]. Although the mechanisms of how TEM stimulates angiogenesis are still under debate, current studies are studying the effects of TEM on tumor angiogenesis.

Regulatory T cells (Tregs)

It is well known that regulatory T cells (Tregs) play an important role in tumor progression and tumor angiogenesis, as they are highly enriched in the TME [38]. A number of factors may contribute to the increased number of Tregs at tumor sites [38]. The environment of tumors such as ovarian cancer and Hodgkin lymphoma contains high concentrations of CC-chemokine ligand 22 (CCL22), which is secreted from both tumor macrophages and tumor cells. Through CCR4, CCL22 recruits Tregs, and Treg migration is inhibited through CCR4 blockade in vitro [38]. Tregs in hypoxic areas are capable of stimulating angiogenesis via VEGF production. As well as VEGF, other angiogenic factors produced by Tregs are Leptin and NRP-1 [141]. The expression of NRP-1 on Treg cells correlates with Foxp3 expression and suppressor function in vitro. NRP-1 was found to promote angiogenesis via interaction with VEGF-A165 (and other VEGFs), and VEGF-R2 enhances signaling through this pathway [141]. Moreover, Tregs indirectly promote angiogenesis by blocking the angiostatic cytokines IFN-γ and CXCL-10 released by effector cells [38]. So, VEGFR2 plays a key role in tumor angiogenesis, and it revealed that this receptor expresses on the surface of Tregs [142].

IL-17-producing T cells: γδT and Th17

The evidence suggests that IL-17 is a key cytokine that plays a significant role in various inflammatory diseases, as well as tumorigenesis [143]. A variety of T-cell subsets produce IL-17, including CD8+ T cells, CD4+ T cells, NKT cells and γδT cells [143,144,145]. Th17 exerts its role by several transcription factors, including HIF1α, RORγt, RORα, IRF4, AHR, c-Rel, IκBζ, BATF, and RUNX1 [146]. Also, γδT cells have various transcription factors, such as RORγt, RelB, RUNX1, AHR, and Hes1 [146]. IL-17-producing γσT cells are tumor-promoting cells that induce angiogenesis in response to the TME [145]. Also according to studies, Th17 cells and IL-17 have been reported to promote anti-VEGF therapy resistance by recruiting immunosuppressive and proangiogenic myeloid cells to the TME [43, 147]. The expression of IL-17 in tumor microenvironments is well established [148, 149]. Studies revealed that as a result of the IL-17 produced by IL-17-producing γδT cells, VEGF, Ang-2, GM-CSF, IL-8, and other angiogenesis factors productions were induced [54]. However, it also stimulates the production of anti-angiogenic factors such as thrombospondin-1 (TSP1), TIMP1, serpine-1, and platelet factor 4 [54]. Furthermore, angiogenesis may be accelerated by the accumulation of other intratumoral IL-17-producing cells, such as Th17 [150]. According to a recent study, IL-17-producing γδT cells homing to inflamed skin depend on CCR6. However, it is not clear precisely how IL-17 influences tumor development [151]. By doing so, a potential cancer immunotherapy approach would be to manipulate the production of IL-17, and other proangiogenic factors by human γδT and Th17 cells, as well as CCR6, or factors involved in γδT17 cell proangiogenic polarization.

Metabolic alterations of cancer cells

Cancer cells are assumed to need to reprogram their catabolic and anabolic metabolism for energy intake and biomass synthesis for cell survival and development in order to initiate and progress, especially in unfavorable microenvironmental conditions [152,153,154]. Otto Warburg discovered nearly a century ago that cancer cells used glucose a lot through aerobic glycolysis [155]. Based on investigations in the field over the last two decades have not only proved that oncogenic defects are mostly responsible for the Warburg Effect in cancer cells, but they have also indicated that metabolic reprogramming has developed considerably beyond what was initially expected [154]. Given the increasing importance of this dysregulated metabolism in cancer biology, it seems appropriate to review what we know about cancer metabolic reprogramming other than the Warburg effect by addressing the following main points. How do cancer cells manage their anabolic metabolism to support rapid proliferation? What are the other energy sources besides the main ones like glucose and glutamine used by cancer cells? How can cancer cells use metabolic reprogramming to communicate with and guide their microenvironment? Because glycolysis and the tricarboxylic acid cycle occur in cells, glucose and glutamine are two major nutritional sources for cancer cell survival and growth (TCA). Conventional waste products from cells, such as lactate, ketone bodies, acetate, ammonia, and other foreign proteins, have long been considered useless metabolites in association with these processes. Remarkably, advances in recent years have described a variety of new features in cancer cells for those conventional waste products, which have gradually evolved into unconventional nutrient sources for ATP production and biomass synthesis for essential components during cancer cells' reaction to stressed stated. We mentioned these unusual nutrient functions in tumor progression bellow.

In recent studies, several typical waste products have been recognized as unconventional nutrient sources, including lactate, ketone bodies, acetate, ammonia, and exogenous proteins.

  1. 1.

    Lactate produced by lactate dehydrogenase A (LDHA) can be exported to the extracellular environment by monocarboxylate transporter 4 (MCT4) in particular; exogenous lactate imported by monocarboxylate transporter 1 (MCT1) can be converted to pyruvate by LDHA in the cytosol or lactate dehydrogenase B (LDHB) in the mitochondria while simultaneously reducing NAD+ to NADH to enter the TCA cycle. Lactate dehydrogenase (LDH), which is commonly activated by oncoproteins like cMyc, HIF-1α, and mTOR in cancer cells, converts pyruvate to lactate while simultaneously oxidizing NADH to NAD+ [156,157,158,159].

For many years, lactate was supposed to be a normal byproduct of cellular metabolism. But recent research indicates that lactate behaves as a complex immunomodulatory component that regulates the activity or functions of the innate and adaptive immune systems. The innate and adaptive immune responses in the intestine and other systemic regions are thus shaped by lactate, a crucial new signaling molecule. In addition, lactate's pleiotropic effects modulate several immune cell functions in the microenvironment and pathological cases [160,161,162].

Excessive lactate released by reprogramming cancer cells' metabolism affects immunological responses through extracellular acidification, serving as an energy source by migrating among various cell populations and blocking the mTOR pathway in immune cells [163, 164].

Exogenous lactate increases the migration and invasion of cancer cells in a concentration-dependent approach using the Boyden chamber assay [165], stimulates various oncogenic signaling pathways [166], and positively correlates with radioresistance [167]. Lactate has also been found to acidify the tumor microenvironment and modify numerous immune cells, allowing them to evade immune surveillance and radiotherapy resistance in cancerous patients [168].

  1. 2.

    Acetate is absorbed and converted to cytosolic acetyl-CoA by cytoplasmic acetyl-CoA synthetase 2 (ACSS2), which recent research suggests that acetate potentially plays an essential role in sterol, cholesterol synthesis, and fatty acid production and histone acetylation, particularly in hypoxic and low-lipid [169].

ACSS2 has also been identified as a critical enzyme in the production of acetyl-CoA from acetate, and its expression has been linked to tumor aggressiveness in various organs. These results point to the possibility that acetate use is a common characteristic of many tumors [170].

  1. 3.

    In mitochondria, enzymes such as D-OHB dehydrogenase (BDH) and succinyl-CoA: 3-oxoacid-CoA transferase (OXCT) induce ketolysis (ketone body catabolism), To restore the acetyl-CoA pool. Studies have identified a surprising relationship between ketolysis and liver cancer progression, suggesting novel targets for liver cancer therapy based on its pathological mechanism for liver cancer [171].

  2. 4.

    Ammonia generated by glutaminase in the primary organ can be used by glutamate dehydrogenase in the secondary organ to produce glutamate or α-ketoglutarate under certain conditions. The production of proline, aspartate, and BCAAs (branched-chain amino acids) is facilitated by increased glutamate levels. Because of deficient vascularization, ammonia frequently concentrates in the tumor microenvironment, leading to its repletion into cancer cells. The destination and functions of the ammonia produced by cancer cells remain unknown. It would be better sense if the ammonia could be utilized in a metabolic pathway by the tumor cells, as it may be a nitrogen donor [172].

  3. 5.

    The RAS and PI3K pathways enable cancer cells to grow and survive. In this regard, under the control of RAS and PI3K, extracellular proteins can be absorbed, digested, and degraded into amino acids in tumor cell lysosomes, a process known as macropinocytosis. Macropinocytosis is a conserved endosomal process in lysosomes that utilizes extracellular protein degradation to free amino acids [154] (Fig. 2).

Fig. 2
figure 2

Interactions between metabolism and the tumor microenvironment. The chemical properties of the extracellular space are altered by cancer cells, which has complex effects on the characteristics of normal cells in the tumor microenvironments, as well as the extracellular matrix. In addition, the cancer cells; metabolic and signaling responses are influenced by the microenvironment

The effects of metabolic alterations in cancer cells and TME cells on each other

One of the important risk factors in cancer angiogenesis is the metabolic alteration of cancer cells. So, targeting key metabolic enzymes and/or mitochondrial metabolic pathways in the hypoxic conditions of the TME, can be a valuable and new anti-cancer therapy [173,174,175].

Information about a cell's metabolism has the ability to influence not just the cell's own programming, but also the destiny of other cells in its proximity. However, a range of genetically stable cell types, including endothelial cells, TAFs, and innate and adaptive immune system components, have been shown to develop phenotypic alterations as a result of living close to developing tumors [4]. Although it is unclear how cancer cells reprogram their microenvironment to promote tumor growth and dissemination, it is apparent that certain reprogramming involves a variety of mechanisms, such as secreted growth factors, changes in cell–cell interactions, and the extracellular matrix play a crucial role in this line. So, proliferating cancer cells affect the metabolic content of the extracellular environment around them. For instance, extracellular lactate accumulates as a result of cancer cell's excessive use of extracellular glucose and glutamine, which has been reported to influence a variety of cell types in tumor microenvironments. High lactate levels, which inhibit monocyte migration and dendritic and T cell activation, promote the formation of immune-permissive microenvironments [165, 176, 177]. Lactate also causes local macrophages to polarize into the so-called M2 state, which is essential in immunosuppression and wound healing [178, 179]. Moreover, lactate accumulation is important in the stimulation of angiogenesis. Lactate increases HIF-1α stability and VEGF release from tumor-associated stromal cells, and also activation of NF-κB and PI3K signaling in endothelial cells [180,181,182,183]. Lactate stimulates the formation of hyaluronic acid by fibroblasts, which may promote tumor invasion [184]. In hypoxia conditions, lactate secretion into the extracellular environment via the monocarboxylate transporter MCT1 is associated with H+ co-transport, causing the cell microenvironment to become acidic. Also, the excess CO2 produced during mitochondrial decarboxylation pathways increases extracellular acidity. CO2 diffuses into the extracellular environment, where it is converted to H+ and HCO3 by an extracellular type of carbonic anhydrases [185]. So, during this condition, the expression of carbonic anhydrases, particularly the CAIX isoform, is upregulated through hyper-activated HIF-1. In the following, the proteolytic activity of MMPs and cathepsins are promoted by greater extracellular acidity, enhancing the breakdown of extracellular matrix proteins and increasing tumor invasion. Although lactate accumulation and extracellular acidification may be considered a side effect of cancer-specific metabolic reprogramming. On the other hand, ROS produced by the tumor induces oxidative stress in the cancer adjacent fibroblasts leading to a reduction in their mitochondrial function and increase of glucose uptake, which facilitate metabolic reprogramming and differentiation into CAFs. CAFs, as neighborhood cells of tumor cells representing a significant portion of the tumor mass, have quite similarly metabolic reprogramming relative to the tumor cells. CAFs produce high-energy metabolites, such as lactate, pyruvate, and ketone bodies that fuel the neighboring tumor cells. So, these acidic products, particularly lactic acid, acidify TME, as an important stimulator for tumor progression angiogenesis [186,187,188,189,190].

Various tumors use a different process to stimulate the development of an immune-permissive microenvironment around them. Especially, the tryptophan-degrading dioxygenases indoleamine-2, 3- ioxygenase (IDO1), and tryptophan-2, 3-dioxygenase (TDO2), which catalyze the converting tryptophan, into its derivative, kynurenine, are overexpressed in a variety of solid tumor types [191]. As a result, tryptophan deficiency induces effector T cells death as a result of amino acid starvation [192]. Additionally, kynurenine accumulates can act as a ligand for aryl hydrocarbon receptors (AhR) [193]. Kynurenine enhances the regulatory T-cell phenotype in a mechanism that is dependent on AhR, contributing to the inhibition of anti-tumor immune responses [194]. Finally, kynurenine stimulates extracellular matrix breakdown and invasion by influencing autocrine signaling through AhR on cancer cells [193]. Clinical trials for small molecule inhibitors of IDO1 are presently underway [195]. In addition, the conditions in the TME have a significant impact on a cancer cell's metabolism. Tumors are constantly faced with nutrient and oxygen-depleted environments and they adopt a variety of nutrient-scavenging mechanisms to overcome these constraint conditions. Hypoxia affects cell's ability to carry out oxidative phosphorylation and other oxygen-dependent activities, affecting the redox balance and altering cellular signaling and transcriptional pathways. On the other hand, a disorder in the TCA cycle of cancerous cells through loss-of-function mutations in its enzymes such as succinate dehydrogenase, fumarate hydratase and isocitrate dehydrogenase cause to toxic accumulation of succinate, fumarate and L-2-hydroxyglutarate and D-2-hydroxyglutarat called Oncometabolites. These Oncometabolites contribute to angiogenesis and cancer cell growth by altering the expression of the related genes required for malignant features such as HIF-1 [187, 196,197,198]. Other metabolites such as lysophosphatidic acid (LPA) and adenosine, accumulated in the extracellular milieu of the TME, also promote tumor growth by affecting the immune functions. In cancer cells, an excessive amount of synthesized LPAs are released into the TME. On the other hand, some TME cells, like Cancer-associated adipocytes (CAAs) and TAMs, also contribute to increase the LPA into the TME. Increased LPA of TME induces aerobic glycolysis in the cancerous cells, stimulates pro-tumorigenic features of TAMs, suppresses immune activity of T lymphocytes and subsequently, promotes cancer cell proliferation and migration [199,200,201,202]. Adenosine is also, abundantly produced by the tumor as well as by cells in the TME such as CAFs and immune cells infiltrating the area by the action of the ecto-enzymes of CD39 and CD73 overexpressed to catabolize the ATP. Adenosine riched TME directly leads to tumor growth by binding the adenosines to four distinct GPCRs. Moreover, an adenosine rich TME has immunosuppressive effects through reducing cytotoxic activities of T lymphocytes and natural killer (NK) cells and reduced capacity of neutrophils to phagocytose, degranulate, adhere to endothelial cells and produce ROS and inhibiting functions of Tregs [187, 203,204,205]. Interestingly, some metabolic enzymes, such as fructose-bisphosphatase 1 (FBP1), pyruvate kinase M2 (PKM2) and malate dehydrogenase 1 (MDH1), act as a tumor suppressor or oncogenic factor alongside their canonical role. For instance, FBP1 enzyme inhibits Notch signaling in breast cancer by HIF-1 un-stabilizing and regulating the Wnt/β-catenin pathway. So, FBP1 enzyme plays a tumor suppressive role downregulated in tumor cells. On the other hand, PKM2 and MDH1 enzymes act as oncogenic factors. In this line, PKM2 enzyme interacts with anti-apoptotic protein Bcl-2 increasing its stability and tumor cells' survival. PKM2 enzyme also increases the activity of STAT3 and HIF-1α and their downstream factors/genes promoting angiogenesis and tumor growth. To support the cancerous cells survival, MDH1 enzyme directly binds to p53 increasing its stability and transcriptional activity [206,207,208,209,210]. Hence, taken together, reciprocal interactions among cancer cells and their microenvironments generate a selective influence on cancer cell metabolism, promoting the development of a more aggressive state (Fig. 3).

Fig. 3
figure 3

Links between metabolism and the microenvironment. Cancer cells change the extracellular milieu's chemical composition, which has pleiotropic effects on the phenotypes of cells around the tumor and the extracellular matrix. The microenvironment influences the metabolic and signaling responses of cancer cells reciprocally. MMPs: matrix metalloproteinases, AhR: aryl -hydrocarbon receptor, HA, hyaluronic acid, MCT1: monocarboxylate transporter 1, Kyn, kynurenine, TDO2: tryptophan-2, 3-dioxygenase 2, IDO1: indoleamine-2, 3-dioxygenase 1, CAIX: carbonic anhydrase IX, Treg: regulatory T cells, ECM: extracellular matrix

Induced tumor angiogenesis by stimulated signaling pathways

Obviously, signaling pathways play a key role in angiogenesis. A similar orchestra leads to tumor expansion. Hypoxia results from insufficient blood supply in the TME area and lead to activation of HIF-1 and upregulation of VEGF [211, 212]. HIF-1 consists of HIF-1α that is sensitive to oxygen pressure unlike HIF-1β (structural subunit). In normoxia, HIF-1α is hydroxylated with Poly Hydroxylase Domain protein (PHD), the Von Hippel-Lindau protein (pVHL) recognizes hydroxylated HIF-1α and links to Elogin C and musters ubiquitin ligase complex, eventually, proteasome degrades hydroxylated HIF-1α after polyubiquitination [213]. Inevitably, a significant amount of redox changes are always seen in the TME that lead to stability of HIF-1α. For instance, increased amounts of ROS inhibit PHD activity through iron oxidation, or NO targets the pVHL oxygen-dependent region commonly known as S-Nitrosylation, and prevents HIF-1α detection by pVHL [214]. According to studies PHD needs α-KG to hydroxylate HIF-1α, but there is a condition called pseudohypoxia that is caused by substances such as 2-hydroxyglutarate (2-HG). 2-HG due to very strong competition with α-KG prevents α-KG connection to PHD and inhibits it [215].

In absence of oxygen, PI3K/Akt/mTOR pathway causes activation of HIF-1α, Importantly Akt as a serine/tyrosine contributes to the activation of mTOR [208, 216, 217]. According to studies by Carbonneau, M. et al. in addition to preventing PHD activity, 2-HG also activates mTOR as a transcriptional factor (3) [218]. Likewise, activated HIF-1α gets dimerized in the nucleus with HIF-1β and forms HIF-1. Binding this heterodimer to the hypoxia response element (HRE) and adjoining to CBP/P300 in the promoter area leads to rising above transcription of VGEF genes [219, 220].

On surface of Tip cells as a leader with numerous filopodia have been expressed VEGFR2, VEGFR3 in a vast number and interaction with VEGF and these receptors lead to activation of the promoter of VEGFR2 and delta-like ligand 4 (Dll4) [221,222,223]. Dll4 is a type of notch ligand that has a high expression in Tip cells, on the other hand, its notch receptor is expressed on another type of EC that is called stalk cells. Interestingly, engagement of Dll4/notch cleavages notch intracellular domain (NICD), likewise, NICD translocates into the nucleus and links with Rbpj/Cbf1 transcription factors results in gene expressions that are involved in differentiation and proliferation of cells [224, 225]. Moreover, this signaling pathway leads to a decrease of VEGFR2, VEGFR3 expression and an increase of VEGFR1 expression which finally prevents stalk cells from turning into Tip cells [221].

In order for new vessels to form a vascular network, Tip cells are able to link each other with filopodia [226], but the role of myeloid cells such as Microglial and macrophage shouldn't be overlooked. Macrophages under the influence of Ang-2 secreted from ECs migrate forward Tip cells [226]. Importantly it is essential that both Tip cells in junction site turn into stalk cells. Macrophages are the main leaders in this event, and secrete VEGF that affects VEGFR3 on Tip cells, the downstream signaling pathway activates FoxC2 transcription factor that advances gene transcription in favor of converting Tip to stalk [227]. According to studies by Zonneville et al. TGF-β causes Smad3/4 signaling pathway in fibroblast cells via interaction with TGFRB1. This signaling pathway increases fibronectin production and deposition. Interestingly fibronectin, not only strengthens the pericytes and ECs bond but also holds PDGF in TME in favor of pericytes signaling pathways [228]. PDGF-β that is secreted from ECs effects on PDGFBR-β and causes proliferation and migration of pericytes [229]. Migration of pericytes is up to Ras/Raf/MAPK signaling pathway that recruits cellular cytoskeleton in favor of migration. PI3K/PKC/TGF-β and PI3K/Akt/NF-κB signaling pathways both are involved in the proliferation of pericytes [230].

In vacuolization VEGF is the radical factor and as mentioned hypoxia has an effect on VEGF production directly. According to the studies, mTOR has a crucial role in angiogenesis [231]. For instance, VEGF and VEGFR2 connection results Y951 site autophosphorylation in VEGFR2 that results in TSAd/Src/PI3K/Akt/mTOR signaling pathway and significantly increases cell survival [25, 232,233,234]. Y1214 tyrosine residue actives NCK/FYN/PAK2/CDC42/P38MAPK signaling pathway and at the end P38MAPK causes migration by the use of cellular fibers [235]. RAS/RAF/ERK signaling pathway which probably is activated under PLCγ activity nearby Y1175 site gives rise to proliferation [234, 236]. After engagement of VEGF and VEGFR2, JAK2 phosphorylates STAT3, thus STAT3 plays its angiogenic role by upregulation of BCL2 and ANG2 gene expression [233].

As mentioned above, VEGF is the most important angiogenic factor in TME. Moreover, there are some factors that strengthen the VEGF/VEGFR axis. For example, bone morphogenetic protein (BMP) 2/6, increase the expression of VEGFR2, Dll4 and enhance the notch signaling pathway in the Tip cells. In contrast, it can be said that VEGF also increases the expression of BMP2/4 [237]. BMP9/10 also phosphorylate smad1/5 factors through BMP9/10-ALK1-endoglin signaling pathway, which upregulation of expression of id1 gene is the angiogenic activity of this pathway. In addition, it is important to note that increasing BMPs in cancers significantly increases resistance to conventional anti-VEGF therapies [238]. Erkasap et al. studies have proven that in most solid cancers, elevated levels of leptin as an adipokine and IL-1 have been associated with angiogenesis [239]. The binding of IL-1 to its receptor causes activation of IL-1 receptor kinase 1 (IRAK1) and IRAK4. IRAK1/4 causes recruitment of TNF receptor (TNFR)-associated factor-6 (TRAF6), ubiquitin conjugating enzyme 13 (Ubc13), and ubiquitin E2 variant 1a (Uev1a). Eventually, these factors activate TGF-β-activated kinase 1 (TAK1). TAK1, activity causes phosphorylation of MAP kinase kinase4/7 (MKK4/7), inhibitory kappa B kinase (IKK), and MKK3/6. These phosphorylated factors activate NF-κB, JNK, and p38 [240]. Yasmine F. Elesawy et al. have been reported that elevated leptin levels are directly related to MMP2/9 levels in cancers, and in addition, the induction of JNK,p38, and MAPK/ERK is another effect of leptin on angiogenesis in tumors [241].

According to studies, EphB4/ephrinB2 signaling pathways are involved in the migration of ECs and the formation of veins and arteries [242]. EphB4 and ephrinB2 are transmembrane receptor and ligand respectively which their connection brings on forward signaling in the cell that contains EphB4 and on the opposite side, reverse signaling occurs [243]. Forward signaling includes PI3K/Akt/NO pathway that stimulates PKG/Raf/Ras/MAPK and FAK pathways. These two pathways lead to proliferation and migration respectively. As well reverse signaling leads to the migration of ECs results from PAK/FAK signaling pathway [244]. Moreover, EphB4 directly phosphorylates STAT3 [243, 244] and STAT3 transmits the message of extracellular matrix (ECM) and pericytes into the nucleus that causes cell assembly [243, 245, 246].

Interestingly, in Wnt canonical signaling, β-catenin as a transcription factor is phosphorylated by glycogen synthase kinase 3β (GSK 3β), adenomatous polyposis coli (APC) and Axin. This phosphorylation brings degradation of β-catenin, but in contrast, when Wnt connects to FZD, APC/GSK-3β/Axin complex is inhibited [247, 248] and β-catenin translocates into the nucleus and expresses target genes along with LEF/TCF as co-transcriptional factors [249].

Aalso, Ang-1 and Ang-2 have the same receptor with the same extracellular site on their receptor and almost equal affinity. But the point is that connection of Ang-1 with Tie-2 as its receptor results PI3K/PDK-1/Akt signaling pathway while connection of Ang-2 with Tie-2 blocks this pathway and in the other word Ang-2 is an antagonist of Tie-2 [220, 250] [10, 50]. Survival and migration of epithelial cells are caused by PI3K/PDK-1/Akt/mTOR and PI3K/PDK-1/Akt/eNOS/NO signaling pathways [251, 252]. Obviously, increased expression of immune checkpoint ligands such as PD-L1 in TME is one of the immune escape mechanisms of tumor cells, but according to recent studies, increased PD-L1 in TME is directly related to angiogenesis. To that end, PD-L1 binds to VEGFR2 and actives C-JUN as a transcription factor in favor of the proliferation [253] (Fig. 4).

Fig. 4
figure 4

Illustration of EC signaling due to TME angiogenic activity and hypoxia-induced VEGF production in tumor cells: Hypoxia prevents HIF-1α from degradation in the proteasome and HIF-1α in the nucleus links with HIF-1β and CBP/P300 and upregulates VEGF gene transcription. Briefly, the most important pathways in EC lead to cell proliferation, survival, and migration that results in angiogenesis under the effect of soluble mediators secreted by TME cells. Proliferation: Ras/Raf/ERK related to VEGF/VEGFR2 signaling, and PI3K/AKT/MAPK related to EphB4/ephrinB2 forward signaling. Survival: TSAd/Akt/mTOR in VEGF/VEGFR2 signaling events, and AKT/PDK-1/Akt/mTOR in ANG-2/Tie-2 signaling events. Migration: NCK/ FYN/p38MAPK results from VEGF/VEGFR2 interaction, PI3K/AKT/FAK results from EphB4/ephrinB2 forward signaling, and PAK/FAK results from EphB4/ephrinB2 reverse signaling pathway

Therapeutic perspectives and future direction

In the TME, a variety of metabolic processes and signaling molecules/pathways influence tumor angiogenesis. For the development of new therapeutic strategies, it is necessary to understand how these components function as angiogenic stimuli or as repressors. According to studies, anti-angiogenic drugs can reduce immunosuppressive cell numbers in TME and alleviate tumor-associated immunosuppression [254]. There are various challenges in cancer therapy. One of them is providing effective therapies while minimizing side effects [255]. So, patients may benefit from a variety of combinations of immune checkpoint blockers (ICBs). Recent studies revealed that there are multiple adverse events (AEs) and immune-related AEs (irAEs) due to the utilization of ICIs in cancer patients [256]. In addition, since cancer and coronavirus disease 2019 (COVID-19) have very different immune systems, ICI may cause some unpredictable irAEs in cancer patients with COVID-19 infection [257]. As a way to safely treat cancer patients and prevent immune checkpoint blockade-induced toxicities and autoimmunity, we can use anti-angiogenic drugs solely or combined with ICBs to enhance the safety and effectiveness of therapy for these patients. Transcriptional factors hypoxia-inducible factors (HIFs) regulate gene expression, phenotypic and metabolic changes including tumor angiogenesis, metastasis, and invasion, during hypoxic tumor response [258, 259]. Thus, directly targeting HIF-1 or its indirectly targeting by agents inhibiting its up and down-stream signaling/factors, such as translational inhibitor and microtubule-targeting metabolite 2-methoxyestradiol (2ME2), heat shock protein inhibitors, histone deacetylase inhibitors, and even topotecan (a topoisomerase I inhibitor), are potential strategies for cancer therapy. Moreover, hypoxia conditions in cancerous cells and TME represent an opportunity to develop hypoxia pro-drugs such as evofosfamide and gemcitabine activated via redox under extremely hypoxic conditions. As mentioned previously, carbonic anhydrase enzymes are upregulated via hypoxia and hyper-activation of HIF-1. To prevent the high activity of carbonic anhydrase enzymes, Girentuximab, an anti-carbonic anhydrase enzyme (CAIX) antibody, was developed to stimulate both innate and adaptive immune-mediated killing of tumor cells. Other drugs, directly and indirectly, targeting HIF-1 are 2ME2, 17-AAG, Vorinostat, PT2977, EZN-2208, and CRLX101 whose detailed properties are described in Table 1 [260,261,262,263,264].

Table 1 Drugs for cancer treatment with targeting signaling pathways and factors associated with tumor angiogenesis

Tumor development is closely associated with angiogenesis, and VEGFR2 plays a crucial role in tumor angiogenesis. There is extensive VEGFR2 expression in the blood vessels, especially in tumor microvessels. As well, VEGFR2 is found on the surface of a variety of immune cells, including macrophages, DCs, and Tregs [142]. Therefore, drugs directly/indirectly inhibiting VEGFR2 activity may be a potential anti-angiogenic therapy for different solid tumors. For instance, cetuximab, tivozanib, cediranib and Tyrosine kinase inhibitors (TKIs) like axitinib, sorafenib, sunitinib and pazopanib. Besides VEGFR-1, 2 and 3 inhibitions, Pazopanib also, suppresses c-KIT and platelet-derived growth factor receptor (PDGFR) tyrosine kinases (TKs). Moreover, Bevacizumab (also called avastin) is an inhibitor of VEGFA in patients with lung cancer and colorectal cancer (CRC) [265,266,267,268]. Programmed death-1 receptor (PD-1) is a checkpoint mediator acting primarily in the priming phase of immune responses, that interaction with its ligand, programmed death-ligand 1 (PD-L1), promotes the immunosuppressive state in the TME. PD-1 or PD-L1 blockade can thus offer promising therapeutic options for patients with advanced cancers. In this line, durvalumab, as a PD-L1 inhibitor, showed a promising response rate in patients with ovarian, endometrial and triple-negative breast cancer (TNBC) when it was combined with cediranib. Also, it was shown that combination therapy of bevacizumab with atezolizumab, another PD-L1 inhibitor, created an immunogenic microenvironment for tumor regression [269,270,271]. Also, Ang/Tie2 receptor axis stimulates angiogenic signaling, and studies suggest Ang-1, 2 and Tie2 receptor as an anti-angiogenic target for therapy [272]. Due to the researches, it has been described an angiogenic function for Wnt/FZD signaling pathway [273] via inducing VEGF upregulation, which leads to unstable and leaky tumor angiogenesis [273, 274]. Thus anti-FZD drugs such as Vantictumab and OTSA101 or anti-Wnt drugs such as CGX1321 and Ipafricept may have anti-angiogenic effects [274, 275].

Rebastinib is one of the Tie2 receptor inhibitors with picomolar potency increasing tumor growth, angiogenesis, and metastasis in metastatic mammary carcinoma and pancreatic neuroendocrine tumors [276]. A combinatorial approach targeting Dll4/Notch and EphB2/EphB4 may also lead to disrupting tumor angiogenesis [277]. Trebananib (AMG386) is a selective antagonist peptide-Fc fusion protein inhibiting the interaction between Ang-1, Ang-2 and Tie2 suppressing the endothelial cell proliferation and subsequently, tumor growth [278]. Angiogenesis and immunosuppressive cell recruitment can also be increased by IL-8. Researchers found that targeting IL-8 or IL-8R, such as HuMax-IL8 (BMS-986253) as an anti-IL-8 monoclonal antibody [279], would be able to provide anti-tumor and anti-angiogenic responses (Table 1) [280]. Despite the fact that metabolic processes and signaling pathways associated with tumor angiogenesis are still not fully known, this review opens new windows of therapeutic insight into intervention for the treatment of cancer patients. Therefore, we encourage researchers to target TME cells and their mediators (Fig. 1), metabolic profiles (Figs. 2, 3), and intracellular signaling pathways (Fig. 4) in order to inhibit angiogenesis in solid tumors.

Conclusion

Nowadays, cancer therapy has been revolutionized by Immune checkpoint blockers (ICBs). While, a large number of patients fail to respond to the ICBs, or suffer a relapse with long-term toxicity (i.e., autoimmunity). The polarized TME is crucial in the outcome of the patient response to an ICB, therefore theoretically treating a vascularized TME could improve the effectiveness of these therapies. We suggest using drugs that inhibit angiogenic signaling pathways stimulated due to metabolic processes or employing drugs that eliminate the tumor or tumor-associated cells that contribute to tumor angiogenesis and invasion. Also, it can combine with other cancer therapy approaches such as ICBs.

Availability of data and materials

Not applicable.

Abbreviations

TME:

Tumor microenvironment

ECM:

Extracellular matrix

CAF:

Cancer-associated fibroblast

TAM:

Tumor-associated macrophage

TAN:

Tumor-associated neutrophil

MDSC:

Myeloid-derived suppressor cells

NK cells:

Natural killer cells

DC:

Dendritic cell

TEM:

Tie2-expressing monocytes

ILCs:

Innate lymphoid cells

Treg:

Regulatory T cell

ICBs:

Immune checkpoint blockers

PI3K:

Phosphatidylinositol 3-kinase

HIF:

Hypoxia-inducible factor

mTOR:

Mammalian target of rapamycin

LDHA:

Lactate dehydrogenase A

LDHB:

Lactate dehydrogenase B

MCT:

Monocarboxylate transporter

ACSS2:

Acetyl-CoA synthetase 2

ACLY:

ATP citrate lyase

OXCT:

Succinyl-CoA:3-oxoacid-CoA transferase

BDH:

D-βOHB dehydrogenase

GDH:

Glutamate dehydrogenase

GLS:

Glutaminase

GS:

Glutamine synthetase

OAA:

Oxaloacetic acid

ECM:

Extracellular matrix

MCT1:

Monocarboxylate transporter 1

MMPs:

Matrix metalloproteinases

Treg:

Regulatory T cells

HA:

Hyaluronic acid

CAIX:

Carbonic anhydrase IX

IDO1:

Indoleamine-2, 3-dioxygenase 1

TDO2:

Tryptophan-2, 3-dioxygenase 2

Kyn:

Kynurenine

AhR:

Aryl—hydrocarbon receptor

PDGFR:

Platelet-derived growth factor receptor

PD-L1:

Programmed death-ligand 1

TKIs:

Tyrosine kinase inhibitors

CRC:

Colorectal cancer

TNBC:

Triple-negative breast cancer

RCC:

Renal cell carcinoma

HCC:

Hepatocellular carcinoma

NSCLC:

Non-small cell lung cancer

References

  1. Sung H, Ferlay J, Siegel RL, Laversanne M, Soerjomataram I, Jemal A, et al. Global cancer statistics 2020: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J Clin. 2021;71(3):209–49.

    Article  PubMed  Google Scholar 

  2. Anderson BO, Cazap E, El Saghir NS, Yip C-H, Khaled HM, Otero IV, et al. Optimisation of breast cancer management in low-resource and middle-resource countries: executive summary of the Breast Health Global Initiative consensus, 2010. Lancet Oncol. 2011;12(4):387–98.

    Article  PubMed  Google Scholar 

  3. Cronin KA, Ravdin PM, Edwards BK. Sustained lower rates of breast cancer in the United States. Breast Cancer Res Treat. 2009;117(1):223.

    Article  PubMed  Google Scholar 

  4. Hanahan D, Coussens LM. Accessories to the crime: functions of cells recruited to the tumor microenvironment. Cancer Cell. 2012;21(3):309–22.

    Article  CAS  PubMed  Google Scholar 

  5. Hanahan D, Weinberg RA. Hallmarks of cancer: the next generation. Cell. 2011;144(5):646–74.

    Article  CAS  PubMed  Google Scholar 

  6. Arneth B. Tumor microenvironment. Medicina. 2020;56(1):15.

    Article  Google Scholar 

  7. Giannone G, Ghisoni E, Genta S, Scotto G, Tuninetti V, Turinetto M, et al. Immuno-metabolism and microenvironment in cancer: key players for immunotherapy. Int J Mol Sci. 2020;21(12):4414.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Cassim S, Pouyssegur J. Tumor microenvironment: a metabolic player that shapes the immune response. Int J Mol Sci. 2020;21(1):157.

    Article  CAS  Google Scholar 

  9. Hall K, Ran S. Regulation of tumor angiogenesis by the local environment. Front Biosci. 2010;15(15):195–212.

    Article  CAS  Google Scholar 

  10. de la Cruz-López KG, Castro-Muñoz LJ, Reyes-Hernández DO, García-Carrancá A, Manzo-Merino J. Lactate in the regulation of tumor microenvironment and therapeutic approaches. Front Oncol. 2019;9:1143.

    Article  PubMed  PubMed Central  Google Scholar 

  11. Risau W. Mechanisms of angiogenesis. Nature. 1997;386(6626):671–4.

    Article  CAS  PubMed  Google Scholar 

  12. Folkman J. What is the evidence that tumors are angiogenesis dependent? JNCI J Natl Cancer Inst. 1990;82(1):4–7.

    Article  CAS  PubMed  Google Scholar 

  13. Carmeliet P. Mechanisms of angiogenesis and arteriogenesis. Nat Med. 2000;6(4):389–95.

    Article  CAS  PubMed  Google Scholar 

  14. Papetti M, Herman IM. Mechanisms of normal and tumor-derived angiogenesis. Am J Physiol Cell Physiol. 2002;282(5):C947–70.

    Article  CAS  PubMed  Google Scholar 

  15. Presta M, Dell’Era P, Mitola S, Moroni E, Ronca R, Rusnati M. Fibroblast growth factor/fibroblast growth factor receptor system in angiogenesis. Cytokine Growth Factor Rev. 2005;16(2):159–78.

    Article  CAS  PubMed  Google Scholar 

  16. Rini BI, Small EJ. Biology and clinical development of vascular endothelial growth factor–targeted therapy in renal cell carcinoma. J Clin Oncol. 2005;23(5):1028–43.

    Article  CAS  PubMed  Google Scholar 

  17. Otrock ZK, Mahfouz RA, Makarem JA, Shamseddine AI. Understanding the biology of angiogenesis: review of the most important molecular mechanisms. Blood Cells Mol Dis. 2007;39(2):212–20.

    Article  CAS  PubMed  Google Scholar 

  18. Clauss M, Weich H, Breier G, Knies U, Röckl W, Waltenberger J, et al. The vascular endothelial growth factor receptor Flt-1 mediates biological activities: implications for a functional role of placenta growth factor in monocyte activation and chemotaxis. J Biol Chem. 1996;271(30):17629–34.

    Article  CAS  PubMed  Google Scholar 

  19. Oh H, Takagi H, Suzuma K, Otani A, Matsumura M, Honda Y. Hypoxia and vascular endothelial growth factor selectively up-regulate angiopoietin-2 in bovine microvascular endothelial cells. J Biol Chem. 1999;274(22):15732–9.

    Article  CAS  PubMed  Google Scholar 

  20. Kroll J, Waltenberger J. VEGF-A induces expression of eNOS and iNOS in endothelial cells via VEGF receptor-2 (KDR). Biochem Biophys Res Commun. 1998;252(3):743–6.

    Article  CAS  PubMed  Google Scholar 

  21. Koch S, Claesson-Welsh L. Signal transduction by vascular endothelial growth factor receptors. Cold Spring Harb Perspect Med. 2012;2(7): a006502.

    Article  PubMed  PubMed Central  Google Scholar 

  22. Tchaikovski V, Fellbrich G, Waltenberger J. The molecular basis of VEGFR-1 signal transduction pathways in primary human monocytes. Arterioscler Thromb Vasc Biol. 2008;28(2):322–8.

    Article  CAS  PubMed  Google Scholar 

  23. Barleon B, Sozzani S, Zhou D, Weich HA, Mantovani A, Marme D. Migration of human monocytes in response to vascular endothelial growth factor (VEGF) is mediated via the VEGF receptor flt-1. 1996.

  24. Li T, Zhu Y, Han L, Ren W, Liu H, Qin C. VEGFR-1 activation-induced MMP-9-dependent invasion in hepatocellular carcinoma. Future Oncol. 2015;11(23):3143–57.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Ferrara N. Vascular endothelial growth factor: basic science and clinical progress. Endocr Rev. 2004;25(4):581–611.

    Article  CAS  PubMed  Google Scholar 

  26. Hoeben A, Landuyt B, Highley MS, Wildiers H, Van Oosterom AT, De Bruijn EA. Vascular endothelial growth factor and angiogenesis. Pharmacol Rev. 2004;56(4):549–80.

    Article  CAS  PubMed  Google Scholar 

  27. Takahashi H, Shibuya M. The vascular endothelial growth factor (VEGF)/VEGF receptor system and its role under physiological and pathological conditions. Clin Sci. 2005;109(3):227–41.

    Article  CAS  Google Scholar 

  28. Bach F, Uddin F, Burke D. Angiopoietins in malignancy. Eur J Surg Oncol (EJSO). 2007;33(1):7–15.

    Article  CAS  PubMed  Google Scholar 

  29. Schnurch H, Risau W. Expression of tie-2, a member of a novel family of receptor tyrosine kinases, in the endothelial cell lineage. Development. 1993;119(3):957–68.

    Article  CAS  PubMed  Google Scholar 

  30. Schliemann C, Bieker R, Padro T, Kessler T, Hintelmann H, Buchner T, et al. Expression of angiopoietins and their receptor Tie2 in the bone marrow of patients with acute myeloid leukemia. Haematologica. 2006;91(9):1203–11.

    CAS  PubMed  Google Scholar 

  31. Jones N, Iljin K, Dumont DJ, Alitalo K. Tie receptors: new modulators of angiogenic and lymphangiogenic responses. Nat Rev Mol Cell Biol. 2001;2(4):257–67.

    Article  CAS  PubMed  Google Scholar 

  32. Hegen A, Koidl S, Weindel K, Marmé D, Augustin HG, Fiedler U. Expression of angiopoietin-2 in endothelial cells is controlled by positive and negative regulatory promoter elements. Arterioscler Thromb Vasc Biol. 2004;24(10):1803–9.

    Article  CAS  PubMed  Google Scholar 

  33. Suchting S, Freitas C, le Noble F, Benedito R, Bréant C, Duarte A, et al. The Notch ligand Delta-like 4 negatively regulates endothelial tip cell formation and vessel branching. Proc Natl Acad Sci. 2007;104(9):3225–30.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Harrington LS, Sainson RC, Williams CK, Taylor JM, Shi W, Li J-L, et al. Regulation of multiple angiogenic pathways by Dll4 and Notch in human umbilical vein endothelial cells. Microvasc Res. 2008;75(2):144–54.

    Article  CAS  PubMed  Google Scholar 

  35. Bender JG, Cooney EM, Kandel JJ, Yamashiro DJ. Vascular remodeling and clinical resistance to antiangiogenic cancer therapy. Drug Resist Updates. 2004;7(4–5):289–300.

    Google Scholar 

  36. Coutelle O, Schiffmann L, Liwschitz M, Brunold M, Goede V, Hallek M, et al. Dual targeting of Angiopoetin-2 and VEGF potentiates effective vascular normalisation without inducing empty basement membrane sleeves in xenograft tumours. Br J Cancer. 2015;112(3):495–503.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Qin L, Bromberg-White JL, Qian C-N. Opportunities and challenges in tumor angiogenesis research: back and forth between bench and bed. Adv Cancer Res. 2012;113:191–239.

    Article  CAS  PubMed  Google Scholar 

  38. Facciabene A, Motz GT, Coukos G. T-regulatory cells: key players in tumor immune escape and angiogenesis. Can Res. 2012;72(9):2162–71.

    Article  CAS  Google Scholar 

  39. Negri L, Ferrara N. The prokineticins: neuromodulators and mediators of inflammation and myeloid cell-dependent angiogenesis. Physiol Rev. 2018;98(2):1055–82.

    Article  CAS  PubMed  Google Scholar 

  40. Pietras K, Pahler J, Bergers G, Hanahan D. Functions of paracrine PDGF signaling in the proangiogenic tumor stroma revealed by pharmacological targeting. PLoS Med. 2008;5(1): e19.

    Article  PubMed  PubMed Central  Google Scholar 

  41. Beenken A, Mohammadi M. The FGF family: biology, pathophysiology and therapy. Nat Rev Drug Discovery. 2009;8(3):235–53.

    Article  CAS  PubMed  Google Scholar 

  42. Ellis LM, Hicklin DJ. VEGF-targeted therapy: mechanisms of anti-tumour activity. Nat Rev Cancer. 2008;8(8):579–91.

    Article  CAS  PubMed  Google Scholar 

  43. Protopsaltis NJ, Liang W, Nudleman E, Ferrara N. Interleukin-22 promotes tumor angiogenesis. Angiogenesis. 2019;22(2):311–23.

    Article  CAS  PubMed  Google Scholar 

  44. Mostafavi S, Zalpoor H, Hassan ZM. The promising therapeutic effects of metformin on metabolic reprogramming of cancer-associated fibroblasts in solid tumors. Cell Mol Biol Lett. 2022;27(1):58.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Park J, Lee SE, Hur J, Hong EB, Choi J-I, Yang J-M, et al. M-CSF from cancer cells induces fatty acid synthase and PPARβ/δ activation in tumor myeloid cells, leading to tumor progression. Cell Rep. 2015;10(9):1614–25.

    Article  CAS  PubMed  Google Scholar 

  46. Galdiero MR, Bonavita E, Barajon I, Garlanda C, Mantovani A, Jaillon S. Tumor associated macrophages and neutrophils in cancer. Immunobiology. 2013;218(11):1402–10.

    Article  CAS  PubMed  Google Scholar 

  47. Ribatti D, Tamma R, Crivellato E. Cross talk between natural killer cells and mast cells in tumor angiogenesis. Inflamm Res. 2019;68(1):19–23.

    Article  CAS  PubMed  Google Scholar 

  48. Krzywinska E, Kantari-Mimoun C, Kerdiles Y, Sobecki M, Isagawa T, Gotthardt D, et al. Loss of HIF-1α in natural killer cells inhibits tumour growth by stimulating non-productive angiogenesis. Nat Commun. 2017;8(1):1–13.

    Article  CAS  Google Scholar 

  49. Hoeres T, Wilhelm M, Smetak M, Holzmann E, Schulze-Tanzil G, Birkmann J. Immune cells regulate VEGF signalling via release of VEGF and antagonistic soluble VEGF receptor-1. Clin Exp Immunol. 2018;192(1):54–67.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Bruno A, Focaccetti C, Pagani A, Imperatori AS, Spagnoletti M, Rotolo N, et al. The proangiogenic phenotype of natural killer cells in patients with non-small cell lung cancer. Neoplasia. 2013;15(2):133.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Bruno A, Bassani B, D’Urso DG, Pitaku I, Cassinotti E, Pelosi G, et al. Angiogenin and the MMP9-TIMP2 axis are up-regulated in proangiogenic, decidual NK-like cells from patients with colorectal cancer. FASEB J. 2018;32(10):5365–77.

    Article  CAS  PubMed  Google Scholar 

  52. Shapoorian H, Zalpoor H, Ganjalikhani-Hakemi M. The correlation between Flt3-ITD mutation in dendritic cells with TIM-3 expression in acute myeloid leukemia. Blood Sci. 2021;3(4):132–5.

    Article  PubMed  PubMed Central  Google Scholar 

  53. Khesht AMS, Karpisheh V, Saeed BQ, Zekiy AO, Yapanto LM, Afjadi MN, et al. Different T cell related immunological profiles in COVID-19 patients compared to healthy controls. Int Immunopharmacol. 2021;97: 107828.

    Article  Google Scholar 

  54. Albini A, Bruno A, Noonan DM, Mortara L. Contribution to tumor angiogenesis from innate immune cells within the tumor microenvironment: implications for immunotherapy. Front Immunol. 2018;9:527.

    Article  PubMed  PubMed Central  Google Scholar 

  55. Stockmann C, Schadendorf D, Klose R, Helfrich I. The impact of the immune system on tumor: angiogenesis and vascular remodeling. Front Oncol. 2014;4:69.

    Article  PubMed  PubMed Central  Google Scholar 

  56. Curiel TJ, Cheng P, Mottram P, Alvarez X, Moons L, Evdemon-Hogan M, et al. Dendritic cell subsets differentially regulate angiogenesis in human ovarian cancer. Can Res. 2004;64(16):5535–8.

    Article  CAS  Google Scholar 

  57. Feijoó E, Alfaro C, Mazzolini G, Serra P, Penuelas I, Arina A, et al. Dendritic cells delivered inside human carcinomas are sequestered by interleukin-8. Int J Cancer. 2005;116(2):275–81.

    Article  PubMed  Google Scholar 

  58. Sawatsubashi M, Yamada T, Fukushima N, Mizokami H, Tokunaga O, Shin T. Association of vascular endothelial growth factor and mast cells with angiogenesis in laryngeal squamous cell carcinoma. Virchows Arch. 2000;436(3):243–8.

    Article  CAS  PubMed  Google Scholar 

  59. Imada A, Shijubo N, Kojima H, Abe S. Mast cells correlate with angiogenesis and poor outcome in stage I lung adenocarcinoma. Eur Respir J. 2000;15(6):1087–93.

    Article  CAS  PubMed  Google Scholar 

  60. Takanami I, Takeuchi K, Naruke M. Mast cell density is associated with angiogenesis and poor prognosis in pulmonary adenocarcinoma. Cancer. 2000;88(12):2686–92.

    Article  CAS  PubMed  Google Scholar 

  61. Tomita M, Matsuzaki Y, Onitsuka T. Effect of mast cells on tumor angiogenesis in lung cancer. Ann Thorac Surg. 2000;69(6):1686–90.

    Article  CAS  PubMed  Google Scholar 

  62. Ribatti D. Mast cells and macrophages exert beneficial and detrimental effects on tumor progression and angiogenesis. Immunol Lett. 2013;152(2):83–8.

    Article  CAS  PubMed  Google Scholar 

  63. Bowrey P, King J, Magarey C, Schwartz P, Marr P, Bolton E, et al. Histamine, mast cells and tumour cell proliferation in breast cancer: does preoperative cimetidine administration have an effect? Br J Cancer. 2000;82(1):167–70.

    Article  CAS  PubMed  Google Scholar 

  64. Ullah E, Nagi AH, Lail RA. Angiogenesis and mast cell density in invasive pulmonary adenocarcinoma. J Cancer Res Ther. 2012;8(4):537.

    Article  CAS  PubMed  Google Scholar 

  65. Ye X-Z, Yu S-C, Bian X-W. Contribution of myeloid-derived suppressor cells to tumor-induced immune suppression, angiogenesis, invasion and metastasis. J Genet Genom. 2010;37(7):423–30.

    Article  CAS  Google Scholar 

  66. Mirghorbani M, Van Gool S, Rezaei N. Myeloid-derived suppressor cells in glioma. Exp Rev Neurother. 2013;13(12):1395–406.

    Article  CAS  Google Scholar 

  67. Bruno A, Mortara L, Baci D, Noonan DM, Albini A. Myeloid derived suppressor cells interactions with natural killer cells and pro-angiogenic activities: roles in tumor progression. Front Immunol. 2019;10:771.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Safarzadeh E, Orangi M, Mohammadi H, Babaie F, Baradaran B. Myeloid-derived suppressor cells: important contributors to tumor progression and metastasis. J Cell Physiol. 2018;233(4):3024–36.

    Article  CAS  PubMed  Google Scholar 

  69. Jacob A, Prekeris R. The regulation of MMP targeting to invadopodia during cancer metastasis. Front Cell Dev Biol. 2015;3(4):2015.

    Google Scholar 

  70. Shojaei F, Wu X, Zhong C, Yu L, Liang X-H, Yao J, et al. Bv8 regulates myeloid-cell-dependent tumour angiogenesis. Nature. 2007;450(7171):825–31.

    Article  CAS  PubMed  Google Scholar 

  71. Sorrentino C, Miele L, Porta A, Pinto A, Morello S. Myeloid-derived suppressor cells contribute to A2B adenosine receptor-induced VEGF production and angiogenesis in a mouse melanoma model. Oncotarget. 2015;6(29):27478.

    Article  PubMed  PubMed Central  Google Scholar 

  72. Iannone R, Miele L, Maiolino P, Pinto A, Morello S. Blockade of A2b adenosine receptor reduces tumor growth and immune suppression mediated by myeloid-derived suppressor cells in a mouse model of melanoma. Neoplasia. 2013;15(12):1400.

    Article  PubMed  PubMed Central  Google Scholar 

  73. Condamine T, Ramachandran I, Youn J-I, Gabrilovich DI. Regulation of tumor metastasis by myeloid-derived suppressor cells. Annu Rev Med. 2015;66:97–110.

    Article  CAS  PubMed  Google Scholar 

  74. Marvel D, Gabrilovich DI. Myeloid-derived suppressor cells in the tumor microenvironment: expect the unexpected. J Clin Investig. 2015;125(9):3356–64.

    Article  PubMed  PubMed Central  Google Scholar 

  75. Sceneay J, Chow MT, Chen A, Halse HM, Wong CS, Andrews DM, et al. Primary tumor hypoxia recruits CD11b+/Ly6Cmed/Ly6G+ immune suppressor cells and compromises NK cell cytotoxicity in the premetastatic niche. Can Res. 2012;72(16):3906–11.

    Article  CAS  Google Scholar 

  76. Nasrollahzadeh E, Razi S, Keshavarz-Fathi M, Mazzone M, Rezaei N. Pro-tumorigenic functions of macrophages at the primary, invasive and metastatic tumor site. Cancer Immunol Immunother. 2020;69:1673–97.

    Article  CAS  PubMed  Google Scholar 

  77. Fu L-Q, Du W-L, Cai M-H, Yao J-Y, Zhao Y-Y, Mou X-Z. The roles of tumor-associated macrophages in tumor angiogenesis and metastasis. Cell Immunol. 2020;353: 104119.

    Article  CAS  PubMed  Google Scholar 

  78. Williams CB, Yeh ES, Soloff AC. Tumor-associated macrophages: unwitting accomplices in breast cancer malignancy. NPJ Breast Cancer. 2016;2(1):1–12.

    Article  Google Scholar 

  79. Mantovani A, Allavena P. The interaction of anticancer therapies with tumor-associated macrophages. J Exp Med. 2015;212(4):435–45.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Chen XJ, Wu S, Yan RM, Fan LS, Yu L, Zhang YM, et al. The role of the hypoxia-Nrp-1 axis in the activation of M2-like tumor-associated macrophages in the tumor microenvironment of cervical cancer. Mol Carcinog. 2019;58(3):388–97.

    Article  CAS  PubMed  Google Scholar 

  81. Werno C, Menrad H, Weigert A, Dehne N, Goerdt S, Schledzewski K, et al. Knockout of HIF-1α in tumor-associated macrophages enhances M2 polarization and attenuates their pro-angiogenic responses. Carcinogenesis. 2010;31(10):1863–72.

    Article  CAS  PubMed  Google Scholar 

  82. Huang S, Van Arsdall M, Tedjarati S, McCarty M, Wu W, Langley R, et al. Contributions of stromal metalloproteinase-9 to angiogenesis and growth of human ovarian carcinoma in mice. J Natl Cancer Inst. 2002;94(15):1134–42.

    Article  CAS  PubMed  Google Scholar 

  83. Valković T, Dobrila F, Melato M, Sasso F, Rizzardi C, Jonjić N. Correlation between vascular endothelial growth factor, angiogenesis, and tumor-associated macrophages in invasive ductal breast carcinoma. Virchows Arch. 2002;440(6):583–8.

    Article  PubMed  Google Scholar 

  84. Plein A, Fantin A, Ruhrberg C. Neuropilin regulation of angiogenesis, arteriogenesis, and vascular permeability. Microcirculation. 2014;21(4):315–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Zalpoor H, Akbari A, Samei A, Forghaniesfidvajani R, Kamali M, Afzalnia A, et al. The roles of Eph receptors, neuropilin-1, P2X7, and CD147 in COVID-19-associated neurodegenerative diseases: inflammasome and JaK inhibitors as potential promising therapies. Cell Mol Biol Lett. 2022;27(1):1–21.

    Article  Google Scholar 

  86. Zalpoor H, Shapourian H, Akbari A, Shahveh S, Haghshenas L. Increased neuropilin-1 expression by COVID-19: a possible cause of long-term neurological complications and progression of primary brain tumors. Human Cell. 2022:1–3.

  87. Quaranta V, Rainer C, Nielsen SR, Raymant ML, Ahmed MS, Engle DD, et al. Macrophage-derived granulin drives resistance to immune checkpoint inhibition in metastatic pancreatic cancer. Can Res. 2018;78(15):4253–69.

    Article  CAS  Google Scholar 

  88. Kzhyshkowska J, Riabov V, Gudima A, Wang N, Orekhov A, Mickley A. Role of tumor associated macrophages in tumor angiogenesis and lymphangiogenesis. Front Physiol. 2014;5:75.

    PubMed  PubMed Central  Google Scholar 

  89. Wang Y, Smith W, Hao D, He B, Kong L. M1 and M2 macrophage polarization and potentially therapeutic naturally occurring compounds. Int Immunopharmacol. 2019;70:459–66.

    Article  CAS  PubMed  Google Scholar 

  90. Li Q, Liu L, Zhang Q, Liu S, Ge D, You Z. Interleukin-17 indirectly promotes M2 macrophage differentiation through stimulation of COX-2/PGE2 pathway in the cancer cells. Cancer Res Treat. 2014;46(3):297.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Jetten N, Verbruggen S, Gijbels MJ, Post MJ, De Winther MP, Donners MM. Anti-inflammatory M2, but not pro-inflammatory M1 macrophages promote angiogenesis in vivo. Angiogenesis. 2014;17(1):109–18.

    Article  CAS  PubMed  Google Scholar 

  92. Lai YS, Wahyuningtyas R, Aui SP, Chang KT. Autocrine VEGF signalling on M2 macrophages regulates PD-L1 expression for immunomodulation of T cells. J Cell Mol Med. 2019;23(2):1257–67.

    CAS  PubMed  Google Scholar 

  93. Mizuno R, Kawada K, Itatani Y, Ogawa R, Kiyasu Y, Sakai Y. The role of tumor-associated neutrophils in colorectal cancer. Int J Mol Sci. 2019;20(3):529.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Hajizadeh F, Maleki LA, Alexander M, Mikhailova MV, Masjedi A, Ahmadpour M, et al. Tumor-associated neutrophils as new players in immunosuppressive process of the tumor microenvironment in breast cancer. Life Sci. 2020:118699.

  95. Antonio N, Bønnelykke-Behrndtz ML, Ward LC, Collin J, Christensen IJ, Steiniche T, et al. The wound inflammatory response exacerbates growth of pre-neoplastic cells and progression to cancer. EMBO J. 2015;34(17):2219–36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Houghton AM, Rzymkiewicz DM, Ji H, Gregory AD, Egea EE, Metz HE, et al. Neutrophil elastase–mediated degradation of IRS-1 accelerates lung tumor growth. Nat Med. 2010;16(2):219–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Ma X, Aoki T, Tsuruyama T, Narumiya S. Definition of prostaglandin E2–EP2 signals in the colon tumor microenvironment that amplify inflammation and tumor growth. Can Res. 2015;75(14):2822–32.

    Article  CAS  Google Scholar 

  98. Hajizadeh F, Maleki LA, Alexander M, Mikhailova MV, Masjedi A, Ahmadpour M, et al. Tumor-associated neutrophils as new players in immunosuppressive process of the tumor microenvironment in breast cancer. Life Sci. 2021;264: 118699.

    Article  CAS  PubMed  Google Scholar 

  99. Strieter RM, Kasahara K, Allen RM, Standiford TJ, Rolfe MW, Becker FS, et al. Cytokine-induced neutrophil-derived interleukin-8. Am J Pathol. 1992;141(2):397.

    CAS  PubMed  PubMed Central  Google Scholar 

  100. Waugh DJ, Wilson C. The interleukin-8 pathway in cancer. Clin Cancer Res. 2008;14(21):6735–41.

    Article  CAS  PubMed  Google Scholar 

  101. Holmes WE, Lee J, Kuang W-J, Rice GC, Wood WI. Structure and functional expression of a human interleukin-8 receptor. Science. 1991;253(5025):1278–80.

    Article  CAS  PubMed  Google Scholar 

  102. Murphy PM, Tiffany HL. Cloning of complementary DNA encoding a functional human interleukin-8 receptor. Science. 1991;253(5025):1280–3.

    Article  CAS  PubMed  Google Scholar 

  103. Knall C, Worthen GS, Johnson GL. Interleukin 8-stimulated phosphatidylinositol-3-kinase activity regulates the migration of human neutrophils independent of extracellular signal-regulated kinase and p38 mitogen-activated protein kinases. Proc Natl Acad Sci. 1997;94(7):3052–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. Masucci MT, Minopoli M, Carriero MV. Tumor associated neutrophils Their role in tumorigenesis, metastasis, prognosis and therapy. Front Oncol. 2019;9:1146.

    Article  PubMed  PubMed Central  Google Scholar 

  105. Tarin D, Croft C. Ultrastructural features of wound healing in mouse skin. J Anat. 1969;105(Pt 1):189–90.

    CAS  PubMed  Google Scholar 

  106. Denton AE, Roberts EW, Fearon DT. Stromal cells in the tumor microenvironment. Stromal Immunol. 2018:99–114.

  107. Orimo A, Gupta PB, Sgroi DC, Arenzana-Seisdedos F, Delaunay T, Naeem R, et al. Stromal fibroblasts present in invasive human breast carcinomas promote tumor growth and angiogenesis through elevated SDF-1/CXCL12 secretion. Cell. 2005;121(3):335–48.

    Article  CAS  PubMed  Google Scholar 

  108. Yang F, Tuxhorn JA, Ressler SJ, McAlhany SJ, Dang TD, Rowley DR. Stromal expression of connective tissue growth factor promotes angiogenesis and prostate cancer tumorigenesis. Can Res. 2005;65(19):8887–95.

    Article  CAS  Google Scholar 

  109. Pozzi A, Moberg PE, Miles LA, Wagner S, Soloway P, Gardner HA. Elevated matrix metalloprotease and angiostatin levels in integrin α1 knockout mice cause reduced tumor vascularization. Proc Natl Acad Sci. 2000;97(5):2202–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Mantovani A, Schioppa T, Porta C, Allavena P, Sica A. Role of tumor-associated macrophages in tumor progression and invasion. Cancer Metastasis Rev. 2006;25(3):315–22.

    Article  PubMed  Google Scholar 

  111. Naugler WE, Karin M. The wolf in sheep’s clothing: the role of interleukin-6 in immunity, inflammation and cancer. Trends Mol Med. 2008;14(3):109–19.

    Article  CAS  PubMed  Google Scholar 

  112. Nilsson MB, Langley RR, Fidler IJ. Interleukin-6, secreted by human ovarian carcinoma cells, is a potent proangiogenic cytokine. Can Res. 2005;65(23):10794–800.

    Article  CAS  Google Scholar 

  113. Kubo N, Araki K, Kuwano H, Shirabe K. Cancer-associated fibroblasts in hepatocellular carcinoma. World J Gastroenterol. 2016;22(30):6841.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Jo M, Nishikawa T, Nakajima T, Okada Y, Yamaguchi K, Mitsuyoshi H, et al. Oxidative stress is closely associated with tumor angiogenesis of hepatocellular carcinoma. J Gastroenterol. 2011;46(6):809–21.

    Article  CAS  PubMed  Google Scholar 

  115. Yan W, Wang X, Dai Y, Zhao B, Yang X, Fan J, et al. Discovery of 3-(5′-Substituted)-Benzimidazole-5-(1-(3, 5-dichloropyridin-4-yl) ethoxy)-1 H-indazoles as potent fibroblast growth factor receptor inhibitors: design, synthesis, and biological evaluation. J Med Chem. 2016;59(14):6690–708.

    Article  CAS  PubMed  Google Scholar 

  116. Katoh M. FGFR inhibitors: effects on cancer cells, tumor microenvironment and whole-body homeostasis. Int J Mol Med. 2016;38(1):3–15.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Gacche RN, Meshram RJ. Angiogenic factors as potential drug target: efficacy and limitations of anti-angiogenic therapy. Biochim Biophys Acta BBA Rev Cancer. 2014;1846(1):161–79.

    Article  CAS  Google Scholar 

  118. Klose CS, Artis D. Innate lymphoid cells as regulators of immunity, inflammation and tissue homeostasis. Nat Immunol. 2016;17(7):765–74.

    Article  CAS  PubMed  Google Scholar 

  119. Vivier E, Artis D, Colonna M, Diefenbach A, Di Santo JP, Eberl G, et al. Innate lymphoid cells: 10 years on. Cell. 2018;174(5):1054–66.

    Article  CAS  PubMed  Google Scholar 

  120. Vacca P, Munari E, Tumino N, Moretta F, Pietra G, Vitale M, et al. Human natural killer cells and other innate lymphoid cells in cancer: friends or foes? Immunol Lett. 2018;201:14–9.

    Article  CAS  PubMed  Google Scholar 

  121. Irshad S, Flores-Borja F, Lawler K, Monypenny J, Evans R, Male V, et al. RORγt+ innate lymphoid cells promote lymph node metastasis of breast cancers. Can Res. 2017;77(5):1083–96.

    Article  CAS  Google Scholar 

  122. Shikhagaie MM, Björklund ÅK, Mjösberg J, Erjefält JS, Cornelissen AS, Ros XR, et al. Neuropilin-1 is expressed on lymphoid tissue residing LTi-like group 3 innate lymphoid cells and associated with ectopic lymphoid aggregates. Cell Rep. 2017;18(7):1761–73.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Eisenring M, Vom Berg J, Kristiansen G, Saller E, Becher B. IL-12 initiates tumor rejection via lymphoid tissue–inducer cells bearing the natural cytotoxicity receptor NKp46. Nat Immunol. 2010;11(11):1030–8.

    Article  CAS  PubMed  Google Scholar 

  124. Sgadari C, Angiolillo AL, Tosato G. Inhibition of angiogenesis by interleukin-12 is mediated by the interferon-inducible protein 10. 1996.

  125. Tugues S, Ducimetiere L, Friebel E, Becher B, editors. Innate lymphoid cells as regulators of the tumor microenvironment. Seminars in immunology; 2019: Elsevier.

  126. Carrega P, Loiacono F, Di Carlo E, Scaramuccia A, Mora M, Conte R, et al. NCR+ ILC3 concentrate in human lung cancer and associate with intratumoral lymphoid structures. Nat Commun. 2015;6(1):1–13.

    Article  Google Scholar 

  127. Murugaiyan G, Saha B. Protumor vs antitumor functions of IL-17. J Immunol. 2009;183(7):4169–75.

    Article  CAS  PubMed  Google Scholar 

  128. Kulig P, Burkhard S, Mikita-Geoffroy J, Croxford AL, Hövelmeyer N, Gyülvészi G, et al. IL17A-mediated endothelial breach promotes metastasis formation. Cancer Immunol Res. 2016;4(1):26–32.

    Article  CAS  PubMed  Google Scholar 

  129. Schmid MC, Varner JA. Myeloid cells in the tumor microenvironment: modulation of tumor angiogenesis and tumor inflammation. J Oncol. 2010;2010.

  130. Tataroğlu C, Kargı A, Özkal S, Eşrefoğlu N, Akkoçlu A. Association of macrophages, mast cells and eosinophil leukocytes with angiogenesis and tumor stage in non-small cell lung carcinomas (NSCLC). Lung Cancer. 2004;43(1):47–54.

    Article  PubMed  Google Scholar 

  131. Lorena S, Oliveira DT, Dorta R, Landman G, Kowalski LP. Eotaxin expression in oral squamous cell carcinomas with and without tumour associated tissue eosinophilia. Oral Dis. 2003;9(6):279–83.

    Article  CAS  PubMed  Google Scholar 

  132. Nowak G, Karrar A, Holmén C, Nava S, Uzunel M, Hultenby K, et al. Expression of vascular endothelial growth factor receptor-2 or Tie-2 on peripheral blood cells defines functionally competent cell populations capable of reendothelialization. Circulation. 2004;110(24):3699–707.

    Article  CAS  PubMed  Google Scholar 

  133. Murdoch C, Tazzyman S, Webster S, Lewis CE. Expression of Tie-2 by human monocytes and their responses to angiopoietin-2. J Immunol. 2007;178(11):7405–11.

    Article  CAS  PubMed  Google Scholar 

  134. Venneri MA, Palma MD, Ponzoni M, Pucci F, Scielzo C, Zonari E, et al. Identification of proangiogenic TIE2-expressing monocytes (TEMs) in human peripheral blood and cancer. Blood J Am Soc Hematol. 2007;109(12):5276–85.

    CAS  Google Scholar 

  135. De Palma M, Venneri MA, Galli R, Sergi LS, Politi LS, Sampaolesi M, et al. Tie2 identifies a hematopoietic lineage of proangiogenic monocytes required for tumor vessel formation and a mesenchymal population of pericyte progenitors. Cancer Cell. 2005;8(3):211–26.

    Article  PubMed  Google Scholar 

  136. Xue R, Sheng Y, Duan X, Yang Y, Ma S, Xu J, et al. Tie2-expressing monocytes as a novel angiogenesis-related cellular biomarker for non-small cell lung cancer. Int J Cancer. 2021;148(6):1519–28.

    Article  CAS  PubMed  Google Scholar 

  137. Gabrusiewicz K, Liu D, Cortes-Santiago N, Hossain MB, Conrad CA, Aldape KD, et al. Anti-vascular endothelial growth factor therapy-induced glioma invasion is associated with accumulation of Tie2-expressing monocytes. Oncotarget. 2014;5(8):2208.

    Article  PubMed  PubMed Central  Google Scholar 

  138. De Palma M, Coukos G, Semela D. TIE2‐expressing monocytes: A novel cellular biomarker for hepatocellular carcinoma? Wiley Online Library; 2013.

  139. Ji J, Zhang G, Sun B, Yuan H, Huang Y, Zhang J, et al. The frequency of tumor-infiltrating Tie–2-expressing monocytes in renal cell carcinoma: its relationship to angiogenesis and progression. Urology. 2013;82(4):974.

    Article  Google Scholar 

  140. Bron S, Henry L, Faes-Van’t Hull E, Turrini R, Vanhecke D, Guex N, et al. TIE-2-expressing monocytes are lymphangiogenic and associate specifically with lymphatics of human breast cancer. Oncoimmunology. 2016;5(2): e1073882.

    Article  PubMed  Google Scholar 

  141. Pucino V, De Rosa V, Procaccini C, Matarese G. Regulatory T cells, leptin and angiogenesis. Angiogene Lymphangiogene Clin Implic. 2014;99:155–69.

    Article  CAS  Google Scholar 

  142. Zhu P, Hu C, Hui K, Jiang X. The role and significance of VEGFR2+ regulatory T cells in tumor immunity. Onco Targets Ther. 2017;10:4315.

    Article  PubMed  PubMed Central  Google Scholar 

  143. Dong C. Diversification of T-helper-cell lineages: finding the family root of IL-17-producing cells. Nat Rev Immunol. 2006;6(4):329–34.

    Article  CAS  PubMed  Google Scholar 

  144. Michel M-L, Keller AC, Paget C, Fujio M, Trottein F, Savage PB, et al. Identification of an IL-17–producing NK1 1neg iNKT cell population involved in airway neutrophilia. J Exp Med. 2007;204(5):995–1001.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  145. Wakita D, Sumida K, Iwakura Y, Nishikawa H, Ohkuri T, Chamoto K, et al. Tumor-infiltrating IL-17-producing γδ T cells support the progression of tumor by promoting angiogenesis. Eur J Immunol. 2010;40(7):1927–37.

    Article  CAS  PubMed  Google Scholar 

  146. Kim JS, Jordan MS. Diversity of IL-17-producing T lymphocytes. Cell Mol Life Sci. 2013;70(13):2271–90.

    Article  CAS  PubMed  Google Scholar 

  147. Chung AS, Wu X, Zhuang G, Ngu H, Kasman I, Zhang J, et al. An interleukin-17–mediated paracrine network promotes tumor resistance to anti-angiogenic therapy. Nat Med. 2013;19(9):1114–23.

    Article  CAS  PubMed  Google Scholar 

  148. Numasaki M, Watanabe M, Suzuki T, Takahashi H, Nakamura A, McAllister F, et al. IL-17 enhances the net angiogenic activity and in vivo growth of human non-small cell lung cancer in SCID mice through promoting CXCR-2-dependent angiogenesis. J Immunol. 2005;175(9):6177–89.

    Article  CAS  PubMed  Google Scholar 

  149. Kryczek I, Wei S, Zou L, Altuwaijri S, Szeliga W, Kolls J, et al. Cutting edge: Th17 and regulatory T cell dynamics and the regulation by IL-2 in the tumor microenvironment. J Immunol. 2007;178(11):6730–3.

    Article  CAS  PubMed  Google Scholar 

  150. Zhang J-P, Yan J, Xu J, Pang X-H, Chen M-S, Li L, et al. Increased intratumoral IL-17-producing cells correlate with poor survival in hepatocellular carcinoma patients. J Hepatol. 2009;50(5):980–9.

    Article  CAS  PubMed  Google Scholar 

  151. Campbell JJ, Ebsworth K, Ertl LS, McMahon JP, Newland D, Wang Y, et al. IL-17–Secreting γδ T cells are completely dependent upon CCR6 for homing to inflamed skin. J Immunol. 2017;199(9):3129–36.

    Article  CAS  PubMed  Google Scholar 

  152. Vander Heiden MG, Cantley LC, Thompson CB. Understanding the Warburg effect: the metabolic requirements of cell proliferation. Science. 2009;324(5930):1029–33.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. Ward PS, Thompson CB. Metabolic reprogramming: a cancer hallmark even warburg did not anticipate. Cancer Cell. 2012;21(3):297–308.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  154. Pavlova NN, Thompson CB. The emerging hallmarks of cancer metabolism. Cell Metab. 2016;23(1):27–47.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  155. Warburg O, Wind F, Negelein E. The metabolism of tumors in the body. J Gen Physiol. 1927;8(6):519–30.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  156. Dang CV. c-Myc target genes involved in cell growth, apoptosis, and metabolism. Mol Cell Biol. 1999;19(1):1–11.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  157. Shim H, Dolde C, Lewis BC, Wu C-S, Dang G, Jungmann RA, et al. c-Myc transactivation of LDH-A: implications for tumor metabolism and growth. Proc Natl Acad Sci. 1997;94(13):6658–63.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  158. Semenza GL, Jiang B-H, Leung SW, Passantino R, Concordet J-P, Maire P, et al. Hypoxia response elements in the aldolase A, enolase 1, and lactate dehydrogenase A gene promoters contain essential binding sites for hypoxia-inducible factor 1. J Biol Chem. 1996;271(51):32529–37.

    Article  CAS  PubMed  Google Scholar 

  159. Babaei M, Alizadeh-Fanalou S, Nourian A, Yarahmadi S, Farahmandian N, Nabi-Afjadi M, et al. Evaluation of testicular glycogen storage, FGF21 and LDH expression and physiological parameters of sperm in hyperglycemic rats treated with hydroalcoholic extract of Securigera Securidaca seeds, and Glibenclamide. Reprod Biol Endocrinol. 2021;19(1):1–15.

    Article  Google Scholar 

  160. Mishra D, Banerjee D. Metabolic interactions between tumor and stromal cells in the tumor microenvironment. Tumor microenvironment: cellular, metabolic and immunologic interactions: Springer; 2021. p. 101–21.

  161. Certo M, Tsai C-H, Pucino V, Ho P-C, Mauro C. Lactate modulation of immune responses in inflammatory versus tumour microenvironments. Nat Rev Immunol. 2021;21(3):151–61.

    Article  CAS  PubMed  Google Scholar 

  162. Manoharan I, Prasad PD, Thangaraju M, Manicassamy S. Lactate-dependent regulation of immune responses by dendritic cells and macrophages. Front Immunol. 2021;12: 691134.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  163. Wang Z-H, Peng W-B, Zhang P, Yang X-P, Zhou Q. Lactate in the tumour microenvironment: from immune modulation to therapy. EBioMedicine. 2021;73: 103627.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  164. Caslin HL, Abebayehu D, Pinette JA, Ryan JJ. Lactate is a metabolic mediator that shapes immune cell fate and function. Front Physiol. 2021:1785.

  165. Goetze K, Walenta S, Ksiazkiewicz M, Kunz-Schughart LA, Mueller-Klieser W. Lactate enhances motility of tumor cells and inhibits monocyte migration and cytokine release. Int J Oncol. 2011;39(2):453–63.

    CAS  PubMed  Google Scholar 

  166. Hirschhaeuser F, Sattler UG, Mueller-Klieser W. Lactate: a metabolic key player in cancer. Can Res. 2011;71(22):6921–5.

    Article  CAS  Google Scholar 

  167. Sattler UG, Meyer SS, Quennet V, Hoerner C, Knoerzer H, Fabian C, et al. Glycolytic metabolism and tumour response to fractionated irradiation. Radiother Oncol. 2010;94(1):102–9.

    Article  CAS  PubMed  Google Scholar 

  168. Groussard C, Morel I, Chevanne M, Monnier M, Cillard J, Delamarche A. Free radical scavenging and antioxidant effects of lactate ion: an in vitro study. J Appl Physiol. 2000;89(1):169–75.

    Article  CAS  PubMed  Google Scholar 

  169. Schug ZT, Vande Voorde J, Gottlieb E. The metabolic fate of acetate in cancer. Nat Rev Cancer. 2016;16(11):708–17.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  170. Schug ZT, Voorde JV, Gottlieb E. The metabolic fate of acetate in cancer. Nat Rev Cancer. 2016;16(11):708–17.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  171. Huang D, Li T, Wang L, Zhang L, Yan R, Li K, et al. Hepatocellular carcinoma redirects to ketolysis for progression under nutrition deprivation stress. Cell Res. 2016;26(10):1112–30.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  172. Yang L, Achreja A, Yeung T-L, Mangala LS, Jiang D, Han C, et al. Targeting stromal glutamine synthetase in tumors disrupts tumor microenvironment-regulated cancer cell growth. Cell Metab. 2016;24(5):685–700.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  173. Stine ZE, Schug ZT, Salvino JM, Dang CV. Targeting cancer metabolism in the era of precision oncology. Nat Rev Drug Discovery. 2022;21(2):141–62.

    Article  CAS  PubMed  Google Scholar 

  174. Luengo A, Gui DY, Vander Heiden MG. Targeting metabolism for cancer therapy. Cell Chem Biol. 2017;24(9):1161–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  175. Mishra D, Banerjee D. Lactate dehydrogenases as metabolic links between tumor and stroma in the tumor microenvironment. Cancers. 2019;11(6):750.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  176. Fischer K, Hoffmann P, Voelkl S, Meidenbauer N, Ammer J, Edinger M, et al. Inhibitory effect of tumor cell–derived lactic acid on human T cells. Blood. 2007;109(9):3812–9.

    Article  CAS  PubMed  Google Scholar 

  177. Gottfried E, Kunz-Schughart LA, Ebner S, Mueller-Klieser W, Hoves S, Andreesen R, et al. Tumor-derived lactic acid modulates dendritic cell activation and antigen expression. Blood. 2006;107(5):2013–21.

    Article  CAS  PubMed  Google Scholar 

  178. Carmona-Fontaine C, Bucci V, Akkari L, Deforet M, Joyce JA, Xavier JB. Emergence of spatial structure in the tumor microenvironment due to the Warburg effect. Proc Natl Acad Sci. 2013;110(48):19402–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  179. Colegio OR, Chu N-Q, Szabo AL, Chu T, Rhebergen AM, Jairam V, et al. Functional polarization of tumour-associated macrophages by tumour-derived lactic acid. Nature. 2014;513(7519):559–63.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  180. Constant JS, Feng JJ, Zabel DD, Yuan H, Suh DY, Scheuenstuhl H, et al. Lactate elicits vascular endothelial growth factor from macrophages: a possible alternative to hypoxia. Wound Repair Regenerat. 2000;8(5):353–60.

    Article  CAS  Google Scholar 

  181. Fernandez-del Castillo C, Schmidt J, Warshaw AL, Rattner DW. Interstitial protease activation is the central event in progression to necrotizing pancreatitis. Surgery. 1994;116(3):497–504.

    CAS  PubMed  Google Scholar 

  182. Végran F, Boidot R, Michiels C, Sonveaux P, Feron O. Lactate influx through the endothelial cell monocarboxylate transporter MCT1 supports an NF-κB/IL-8 pathway that drives tumor angiogenesis. Can Res. 2011;71(7):2550–60.

    Article  Google Scholar 

  183. Sonveaux P, Copetti T, De Saedeleer C, Végran F, Verrax J. Targeting the lactate transporter MCT1 in endothelial cells inhibits lactate. 2012.

  184. Stern R, Shuster S, Neudecker BA, Formby B. Lactate stimulates fibroblast expression of hyaluronan and CD44: the Warburg effect revisited. Exp Cell Res. 2002;276(1):24–31.

    Article  CAS  PubMed  Google Scholar 

  185. Swietach P, Vaughan-Jones RD, Harris AL. Regulation of tumor pH and the role of carbonic anhydrase 9. Cancer Metastasis Rev. 2007;26(2):299–310.

    Article  CAS  PubMed  Google Scholar 

  186. Arcucci A, Ruocco MR, Granato G, Sacco AM, Montagnani S. Cancer: an oxidative crosstalk between solid tumor cells and cancer associated fibroblasts. BioMed Res Int. 2016;2016.

  187. Eisenberg L, Eisenberg-Bord M, Eisenberg-Lerner A, Sagi-Eisenberg R. Metabolic alterations in the tumor microenvironment and their role in oncogenesis. Cancer Lett. 2020;484:65–71.

    Article  CAS  PubMed  Google Scholar 

  188. Peppicelli S, Andreucci E, Ruzzolini J, Laurenzana A, Margheri F, Fibbi G, et al. The acidic microenvironment as a possible niche of dormant tumor cells. Cell Mol Life Sci. 2017;74(15):2761–71.

    Article  CAS  PubMed  Google Scholar 

  189. Peppicelli S, Bianchini F, Calorini L. Extracellular acidity, a “reappreciated” trait of tumor environment driving malignancy: perspectives in diagnosis and therapy. Cancer Metastasis Rev. 2014;33(2):823–32.

    Article  CAS  PubMed  Google Scholar 

  190. Martinez-Outschoorn UE, Lisanti MP, Sotgia F, editors. Catabolic cancer-associated fibroblasts transfer energy and biomass to anabolic cancer cells, fueling tumor growth. Seminars in cancer biology; 2014: Elsevier.

  191. Munn DH, Mellor AL. Indoleamine 2, 3-dioxygenase and tumor-induced tolerance. J Clin Investig. 2007;117(5):1147–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  192. Fallarino F, Grohmann U, Vacca C, Bianchi R, Orabona C, Spreca A, et al. T cell apoptosis by tryptophan catabolism. Cell Death Differ. 2002;9(10):1069–77.

    Article  CAS  PubMed  Google Scholar 

  193. Opitz CA, Litzenburger UM, Sahm F, Ott M, Tritschler I, Trump S, et al. An endogenous tumour-promoting ligand of the human aryl hydrocarbon receptor. Nature. 2011;478(7368):197–203.

    Article  CAS  PubMed  Google Scholar 

  194. Fallarino F, Grohmann U, You S, McGrath BC, Cavener DR, Vacca C, et al. The combined effects of tryptophan starvation and tryptophan catabolites down-regulate T cell receptor ζ-chain and induce a regulatory phenotype in naive T cells. J Immunol. 2006;176(11):6752–61.

    Article  CAS  PubMed  Google Scholar 

  195. Vacchelli E, Aranda F, Eggermont A, Sautes-Fridman C, Tartour E, Kennedy EP, et al. Trial watch: IDO inhibitors in cancer therapy. Oncoimmunology. 2014;3(10): e957994.

    Article  PubMed  PubMed Central  Google Scholar 

  196. Morin A, Letouzé E, Gimenez-Roqueplo AP, Favier J. Oncometabolites-driven tumorigenesis: from genetics to targeted therapy. Int J Cancer. 2014;135(10):2237–48.

    Article  CAS  PubMed  Google Scholar 

  197. Yong C, Stewart GD, Frezza C. Oncometabolites in renal cancer. Nat Rev Nephrol. 2020;16(3):156–72.

    Article  CAS  PubMed  Google Scholar 

  198. Dang L, White DW, Gross S, Bennett BD, Bittinger MA, Driggers EM, et al. Cancer-associated IDH1 mutations produce 2-hydroxyglutarate. Nature. 2009;462(7274):739–44.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  199. Leblanc R, Peyruchaud O. New insights into the autotaxin/LPA axis in cancer development and metastasis. Exp Cell Res. 2015;333(2):183–9.

    Article  CAS  PubMed  Google Scholar 

  200. Schmid R, Wolf K, Robering JW, Strauß S, Strissel PL, Strick R, et al. ADSCs and adipocytes are the main producers in the autotaxin–lysophosphatidic acid axis of breast cancer and healthy mammary tissue in vitro. BMC Cancer. 2018;18(1):1–11.

    Article  Google Scholar 

  201. Reinartz S, Lieber S, Pesek J, Brandt DT, Asafova A, Finkernagel F, et al. Cell type-selective pathways and clinical associations of lysophosphatidic acid biosynthesis and signaling in the ovarian cancer microenvironment. Mol Oncol. 2019;13(2):185–201.

    Article  CAS  PubMed  Google Scholar 

  202. Ha JH, Radhakrishnan R, Jayaraman M, Yan M, Ward JD, Fung K-M, et al. LPA induces metabolic reprogramming in ovarian cancer via a pseudohypoxic response. Can Res. 2018;78(8):1923–34.

    Article  CAS  Google Scholar 

  203. Ballarin M, Fredholm B, Ambrosio S, Mahy N. Extracellular levels of adenosine and its metabolites in the striatum of awake rats: inhibition of uptake and metabolism. Acta Physiol Scand. 1991;142(1):97–103.

    Article  CAS  PubMed  Google Scholar 

  204. Vigano S, Alatzoglou D, Irving M, Ménétrier-Caux C, Caux C, Romero P, et al. Targeting adenosine in cancer immunotherapy to enhance T-cell function. Front Immunol. 2019;10:925.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  205. Giuliani AL, Sarti AC, Di Virgilio F. Extracellular nucleotides and nucleosides as signalling molecules. Immunol Lett. 2019;205:16–24.

    Article  CAS  PubMed  Google Scholar 

  206. Li B, Qiu B, Lee DS, Walton ZE, Ochocki JD, Mathew LK, et al. Fructose-1, 6-bisphosphatase opposes renal carcinoma progression. Nature. 2014;513(7517):251–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  207. Harris RA, Fenton AW. A critical review of the role of M2PYK in the Warburg effect. Biochim Biophys Acta (BBA) Rev Cancer. 2019;1871(2):225–39.

    Article  CAS  Google Scholar 

  208. Masson N, Ratcliffe PJ. Hypoxia signaling pathways in cancer metabolism: the importance of co-selecting interconnected physiological pathways. Cancer & metabolism. 2014;2(1):1–17.

    Article  Google Scholar 

  209. Dayton TL, Jacks T, Vander Heiden MG. PKM 2, cancer metabolism, and the road ahead. EMBO Rep. 2016;17(12):1721–30.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  210. Liang J, Cao R, Wang X, Zhang Y, Wang P, Gao H, et al. Mitochondrial PKM2 regulates oxidative stress-induced apoptosis by stabilizing Bcl2. Cell Res. 2017;27(3):329–51.

    Article  CAS  PubMed  Google Scholar 

  211. Lee WS, Yang H, Chon HJ, Kim C. Combination of anti-angiogenic therapy and immune checkpoint blockade normalizes vascular-immune crosstalk to potentiate cancer immunity. Exp Mol Med. 2020;52(9):1475–85.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  212. De Palma M, Biziato D, Petrova TV. Microenvironmental regulation of tumour angiogenesis. Nat Rev Cancer. 2017;17(8):457–74.

    Article  PubMed  Google Scholar 

  213. Ma S, Pradeep S, Hu W, Zhang D, Coleman R, Sood A. The role of tumor microenvironment in resistance to anti-angiogenic therapy. F1000Research. 2018;7.

  214. Manuelli V, Pecorari C, Filomeni G, Zito E. Regulation of redox signaling in HIF‐1‐dependent tumor angiogenesis. FEBS J. 2021.

  215. van der Merwe M, van Niekerk G, Fourie C, du Plessis M, Engelbrecht A-M. The impact of mitochondria on cancer treatment resistance. Cell Oncol. 2021;44(5):983–95.

    Article  Google Scholar 

  216. Deep G, Kumar R, Nambiar DK, Jain AK, Ramteke AM, Serkova NJ, et al. Silibinin inhibits hypoxia-induced HIF-1α-mediated signaling, angiogenesis and lipogenesis in prostate cancer cells: In vitro evidence and in vivo functional imaging and metabolomics. Mol Carcinog. 2017;56(3):833–48.

    Article  CAS  PubMed  Google Scholar 

  217. Zalpoor H, Bakhtiyari M, Liaghat M, Nabi‐Afjadi M, Ganjalikhani‐Hakemi M. Quercetin potential effects against SARS‐CoV‐2 infection and COVID‐19‐associated cancer progression by inhibiting mTOR and hypoxia‐inducible factor‐1α (HIF‐1α). Phytotherapy Research. 2022.

  218. Zhang Z, Yao L, Yang J, Wang Z, Du G. PI3K/Akt and HIF-1 signaling pathway in hypoxia-ischemia. Mol Med Rep. 2018;18(4):3547–54.

    CAS  PubMed  PubMed Central  Google Scholar 

  219. Gray MJ, Zhang J, Ellis LM, Semenza GL, Evans DB, Watowich SS, et al. HIF-1 α, STAT3, CBP/p300 and Ref-1/APE are components of a transcriptional complex that regulates Src-dependent hypoxia-induced expression of VEGF in pancreatic and prostate carcinomas. Oncogene. 2005;24(19):3110–20.

    Article  CAS  PubMed  Google Scholar 

  220. Jiang BH, Liu LZ. PI3K/PTEN signaling in angiogenesis and tumorigenesis. Adv Cancer Res. 2009;102:19–65.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  221. Lugano R, Ramachandran M, Dimberg A. Tumor angiogenesis: causes, consequences, challenges and opportunities. Cell Mol Life Sci. 2020;77(9):1745–70.

    Article  CAS  PubMed  Google Scholar 

  222. Beets K, Huylebroeck D, Moya IM, Umans L, Zwijsen A. Robustness in angiogenesis: notch and BMP shaping waves. Trends Genet. 2013;29(3):140–9.

    Article  CAS  PubMed  Google Scholar 

  223. Hellström M, Phng L-K, Gerhardt H. VEGF and Notch signaling: the yin and yang of angiogenic sprouting. Cell Adh Migr. 2007;1(3):133–6.

    Article  PubMed  PubMed Central  Google Scholar 

  224. Serra H, Chivite I, Angulo-Urarte A, Soler A, Sutherland JD, Arruabarrena-Aristorena A, et al. PTEN mediates Notch-dependent stalk cell arrest in angiogenesis. Nat Commun. 2015;6(1):1–13.

    Article  Google Scholar 

  225. Akil A, Gutiérrez-García AK, Guenter R, Rose JB, Beck AW, Chen H, et al. Notch signaling in vascular endothelial cells, angiogenesis, and tumor progression: an update and prospective. Front Cell Dev Biol. 2021;9:177.

    Article  Google Scholar 

  226. Moreira-Soares M, Coimbra R, Rebelo L, Carvalho J, Travasso RD. Angiogenic factors produced by hypoxic cells are a leading driver of anastomoses in sprouting angiogenesis–a computational study. Sci Rep. 2018;8(1):1–12.

    Article  CAS  Google Scholar 

  227. Zhu C, Kros JM, Cheng C, Mustafa D. The contribution of tumor-associated macrophages in glioma neo-angiogenesis and implications for anti-angiogenic strategies. Neuro Oncol. 2017;19(11):1435–46.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  228. Zonneville J, Safina A, Truskinovsky AM, Arteaga CL, Bakin AV. TGF-β signaling promotes tumor vasculature by enhancing the pericyte-endothelium association. BMC Cancer. 2018;18(1):1–13.

    Article  Google Scholar 

  229. Ribatti D, Nico B, Crivellato E. The role of pericytes in angiogenesis. Int J Dev Biol. 2011;55(3):261–8.

    Article  CAS  PubMed  Google Scholar 

  230. Sweeney MD, Ayyadurai S, Zlokovic BV. Pericytes of the neurovascular unit: key functions and signaling pathways. Nat Neurosci. 2016;19(6):771–83.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  231. Roudsari NM, Lashgari N-A, Momtaz S, Abaft S, Jamali F, Safaiepour P, et al. Inhibitors of the PI3K/Akt/mTOR pathway in prostate cancer chemoprevention and intervention. Pharmaceutics. 2021;13(8):1195.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  232. Wang X, Bove AM, Simone G, Ma B. Molecular bases of VEGFR-2-mediated physiological function and pathological role. Front Cell Dev Biol. 2020;8:1314.

    Google Scholar 

  233. Wan L, Zhao Y, Zhang Q, Gao G, Zhang S, Gao Y, et al. Alkaloid extract of Corydalis yanhusuo inhibits angiogenesis via targeting vascular endothelial growth factor receptor signaling. BMC Complement Altern Med. 2019;19(1):1–13.

    Article  Google Scholar 

  234. Melincovici CS, Boşca AB, Şuşman S, Mărginean M, Mihu C, Istrate M, et al. Vascular endothelial growth factor (VEGF)-key factor in normal and pathological angiogenesis. Rom J Morphol Embryol. 2018;59(2):455–67.

    PubMed  Google Scholar 

  235. Zhu X, Zhou W. The emerging regulation of VEGFR-2 in triple-negative breast cancer. Front Endocrinol. 2015;6:159.

    Article  Google Scholar 

  236. Ziyad S, Iruela-Arispe ML. Molecular mechanisms of tumor angiogenesis. Genes Cancer. 2011;2(12):1085–96.

    Article  PubMed  PubMed Central  Google Scholar 

  237. Pulkkinen H, Kiema M, Lappalainen J, Toropainen A, Beter M, Tirronen A, et al. BMP6/TAZ-Hippo signaling modulates angiogenesis and endothelial cell response to VEGF. Angiogenesis. 2021;24(1):129–44.

    Article  CAS  PubMed  Google Scholar 

  238. Al Tabosh T, Al Tarrass M, Bailly S. The BMP9/10-ALK1-endoglin pathway as a target of anti-angiogenic therapy in cancer. 2021.

  239. Erkasap N, Ozyurt R, Ozkurt M, Erkasap S, Yasar F, Ihtiyar E, et al. Role of Notch, IL-1 and leptin expression in colorectal cancer. Exp Ther Med. 2021;21(6):1–11.

    Article  Google Scholar 

  240. Zhu L, Lama S, Tu L, Dusting GJ, Wang J-H, Liu G-S. TAK1 signaling is a potential therapeutic target for pathological angiogenesis. Angiogenesis. 2021;24(3):453–70.

    Article  CAS  PubMed  Google Scholar 

  241. Rasha MR, Yasmine FE, IsmailAmer M, Samar I. Immunohistochemical expression of leptin in mammary carcinoma. Med J Cairo Univ. 2021;89:285–96.

    Article  Google Scholar 

  242. Ieguchi K, Maru Y. Eph/Ephrin signaling in the tumor microenvironment. Tumor Microenvironment: Springer; 2021. p. 45–56.

    Book  Google Scholar 

  243. Yang D, Jin C, Ma H, Huang M, Shi G-P, Wang J, et al. EphrinB2/EphB4 pathway in postnatal angiogenesis: a potential therapeutic target for ischemic cardiovascular disease. Angiogenesis. 2016;19(3):297–309.

    Article  CAS  PubMed  Google Scholar 

  244. Du E, Li X, He S, Li X, He S. The critical role of the interplays of EphrinB2/EphB4 and VEGF in the induction of angiogenesis. Mol Biol Rep. 2020;47(6):4681–90.

    Article  CAS  PubMed  Google Scholar 

  245. Zalpoor H, Akbari A, Nabi-Afjadi M. Ephrin (Eph) receptor and downstream signaling pathways: a promising potential targeted therapy for COVID-19 and associated cancers and diseases. Hum Cell. 2022;35(3):952–4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  246. Joshi N, Hajizadeh F, Dezfouli EA, Zekiy AO, Afjadi MN, Mousavi SM, et al. Silencing STAT3 enhances sensitivity of cancer cells to doxorubicin and inhibits tumor progression. Life Sci. 2021;275: 119369.

    Article  CAS  PubMed  Google Scholar 

  247. Patel S, Alam A, Pant R, Chattopadhyay S. Wnt signaling and its significance within the tumor microenvironment: novel therapeutic insights. Front Immunol. 2019;10:2872.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  248. Ding M, Wang X. Antagonism between Hedgehog and Wnt signaling pathways regulates tumorigenicity. Oncol Lett. 2017;14(6):6327–33.

    PubMed  PubMed Central  Google Scholar 

  249. Kasprzak A. Angiogenesis-related functions of wnt signaling in colorectal carcinogenesis. Cancers. 2020;12(12):3601.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  250. Fiedler U, Augustin HG. Angiopoietins: a link between angiogenesis and inflammation. Trends Immunol. 2006;27(12):552–8.

    Article  CAS  PubMed  Google Scholar 

  251. Gál Z, Gézsi A, Molnár V, Nagy A, Kiss A, Sultész M, et al. Investigation of the possible role of Tie2 pathway and TEK Gene in asthma and allergic conjunctivitis. Front Genet. 2020;11:128.

    Article  PubMed  PubMed Central  Google Scholar 

  252. Karar J, Maity A. PI3K/AKT/mTOR pathway in angiogenesis. Front Mol Neurosci. 2011;4:51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  253. Yang Y, Xia L, Wu Y, Zhou H, Chen X, Li H, et al. Programmed death ligand-1 regulates angiogenesis and metastasis by participating in the c-JUN/VEGFR2 signaling axis in ovarian cancer. Cancer Commun. 2021;41(6):511–27.

    Article  Google Scholar 

  254. Hu C, Jiang X. The effect of anti-angiogenic drugs on regulatory T cells in the tumor microenvironment. Biomed Pharmacother. 2017;88:134–7.

    Article  CAS  PubMed  Google Scholar 

  255. Mansouri V, Beheshtizadeh N, Gharibshahian M, Sabouri L, Varzandeh M, Rezaei N. Recent advances in regenerative medicine strategies for cancer treatment. Biomed Pharmacother. 2021;141: 111875.

    Article  CAS  PubMed  Google Scholar 

  256. Balibegloo M, Nejadghaderi SA, Sadeghalvad M, Soleymanitabar A, Nezamabadi SS, Saghazadeh A, et al. Adverse events associated with immune checkpoint inhibitors in patients with breast cancer: A systematic review and meta-analysis. Int Immunopharmacol. 2021;96: 107796.

    Article  CAS  PubMed  Google Scholar 

  257. Pezeshki PS, Rezaei N. Immune checkpoint inhibition in COVID-19: risks and benefits. Expert Opin Biol Therapy. 2021:1–7.

  258. McIntyre A, Harris AL. Metabolic and hypoxic adaptation to anti-angiogenic therapy: a target for induced essentiality. EMBO Mol Med. 2015;7(4):368–79.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  259. Zalpoor H, Akbari A, Nabi-Afjadi M, Forghaniesfidvajani R, Tavakol C, Barzegar Z, et al. Hypoxia‐inducible factor 1 alpha (HIF‐1α) stimulated and P2X7 receptor activated by COVID-19, as a potential therapeutic target and risk factor for epilepsy. Human Cell. 2022:1–8.

  260. Li Y, Patel SP, Roszik J, Qin Y. Hypoxia-driven immunosuppressive metabolites in the tumor microenvironment: new approaches for combinational immunotherapy. Front Immunol. 2018;9:1591.

    Article  PubMed  PubMed Central  Google Scholar 

  261. Wigerup C, Påhlman S, Bexell D. Therapeutic targeting of hypoxia and hypoxia-inducible factors in cancer. Pharmacol Ther. 2016;164:152–69.

    Article  CAS  PubMed  Google Scholar 

  262. Augustin RC, Delgoffe GM, Najjar YG. Characteristics of the tumor microenvironment that influence immune cell functions: hypoxia, oxidative stress, metabolic alterations. Cancers. 2020;12(12):3802.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  263. Van Cutsem E, Lenz H-J, Furuse J, Tabernero J, Heinemann V, Ioka T, et al. MAESTRO: A randomized, double-blind phase III study of evofosfamide (Evo) in combination with gemcitabine (Gem) in previously untreated patients (pts) with metastatic or locally advanced unresectable pancreatic ductal adenocarcinoma (PDAC). American Society of Clinical Oncology; 2016.

  264. Chang D-K, Moniz RJ, Xu Z, Sun J, Signoretti S, Zhu Q, et al. Human anti-CAIX antibodies mediate immune cell inhibition of renal cell carcinoma in vitro and in a humanized mouse model in vivo. Mol Cancer. 2015;14(1):1–13.

    Article  Google Scholar 

  265. Motzer RJ, Penkov K, Haanen J, Rini B, Albiges L, Campbell MT, et al. Avelumab plus axitinib versus sunitinib for advanced renal-cell carcinoma. N Engl J Med. 2019;380(12):1103–15.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  266. Albini A, Tosetti F, Li VW, Noonan DM, Li WW. Cancer prevention by targeting angiogenesis. Nat Rev Clin Oncol. 2012;9(9):498–509.

    Article  CAS  PubMed  Google Scholar 

  267. Folkman J. Angiogenesis: an organizing principle for drug discovery? Nat Rev Drug Discovery. 2007;6(4):273–86.

    Article  CAS  PubMed  Google Scholar 

  268. Porta C, Cosmai L, Leibovich BC, Powles T, Gallieni M, Bex A. The adjuvant treatment of kidney cancer: a multidisciplinary outlook. Nat Rev Nephrol. 2019;15(7):423–33.

    Article  PubMed  Google Scholar 

  269. Mortezaee K. Immune escape: a critical hallmark in solid tumors. Life Sci. 2020;258: 118110.

    Article  CAS  PubMed  Google Scholar 

  270. Lee MS, Ryoo B-Y, Hsu C-H, Numata K, Stein S, Verret W, et al. Atezolizumab with or without bevacizumab in unresectable hepatocellular carcinoma (GO30140): an open-label, multicentre, phase 1b study. Lancet Oncol. 2020;21(6):808–20.

    Article  CAS  PubMed  Google Scholar 

  271. Sefid F, Payandeh Z, Azamirad G, Baradaran B, Nabi Afjadi M, Islami M, et al. Atezolizumab and granzyme B as immunotoxin against PD-L1 antigen; an insilico study. In Silico Pharmacol. 2021;9(1):1–12.

    Article  Google Scholar 

  272. Stratmann A, Acker T, Burger AM, Amann K, Risau W, Plate KH. Differential inhibition of tumor angiogenesis by tie2 and vascular endothelial growth factor receptor-2 dominant-negative receptor mutants. Int J Cancer. 2001;91(3):273–82.

    Article  CAS  PubMed  Google Scholar 

  273. Parmalee NL, Kitajewski J. Wnt signaling in angiogenesis. Curr Drug Targets. 2008;9(7):558–64.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  274. Katoh M, Katoh M. Molecular genetics and targeted therapy of WNT-related human diseases. Int J Mol Med. 2017;40(3):587–606.

    CAS  PubMed  PubMed Central  Google Scholar 

  275. Ghosh N, Hossain U, Mandal A, Sil PC. The Wnt signaling pathway: a potential therapeutic target against cancer. Ann N Y Acad Sci. 2019;1443(1):54–74.

    Article  PubMed  Google Scholar 

  276. Duran CL, Borriello L, Karagiannis GS, Entenberg D, Oktay MH, Condeelis JS. Targeting Tie2 in the tumor microenvironment: from angiogenesis to dissemination. Cancers. 2021;13(22):5730.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  277. Djokovic D, Trindade A, Gigante J, Badenes M, Silva L, Liu R, et al. Combination of Dll4/Notch and Ephrin-B2/EphB4 targeted therapy is highly effective in disrupting tumor angiogenesis. BMC Cancer. 2010;10(1):1–12.

    Article  Google Scholar 

  278. Chiu JW, Hotte SJ, Kollmannsberger CK, Renouf DJ, Cescon DW, Hedley D, et al. A phase I trial of ANG1/2-Tie2 inhibitor trebaninib (AMG386) and temsirolimus in advanced solid tumors (PJC008/NCI♯ 9041). Invest New Drugs. 2016;34(1):104–11.

    Article  CAS  PubMed  Google Scholar 

  279. Bilusic M, Heery CR, Collins JM, Donahue RN, Palena C, Madan RA, et al. Phase I trial of HuMax-IL8 (BMS-986253), an anti-IL-8 monoclonal antibody, in patients with metastatic or unresectable solid tumors. J Immunother Cancer. 2019;7(1):1–8.

    Article  Google Scholar 

  280. Fousek K, Horn LA, Palena C. Interleukin-8: A chemokine at the intersection of cancer plasticity, angiogenesis, and immune suppression. Pharmacol Ther. 2021;219: 107692.

    Article  CAS  PubMed  Google Scholar 

  281. Chamie K, Donin NM, Klöpfer P, Bevan P, Fall B, Wilhelm O, et al. Adjuvant weekly girentuximab following nephrectomy for high-risk renal cell carcinoma: the ARISER randomized clinical trial. JAMA Oncol. 2017;3(7):913–20.

    Article  PubMed  Google Scholar 

  282. Xu R, Wang K, Rizzi JP, Huang H, Grina JA, Schlachter ST, et al. 3-[(1 S, 2 S, 3 R)-2, 3-Difluoro-1-hydroxy-7-methylsulfonylindan-4-yl] oxy-5-fluorobenzonitrile (PT2977), a hypoxia-inducible factor 2α (HIF-2α) inhibitor for the treatment of clear cell Renal cell carcinoma. ACS Publications; 2019.

  283. Papadopoulos KP, Jonasch E, Zojwalla NJ, Wang K, Bauer TM. A first-in-human phase 1 dose-escalation trial of the oral HIF-2a inhibitor PT2977 in patients with advanced solid tumors. American Society of Clinical Oncology; 2018.

  284. Garrett CR, Bekaii-Saab TS, Ryan T, Fisher GA, Clive S, Kavan P, et al. Randomized phase 2 study of pegylated SN-38 (EZN-2208) or irinotecan plus cetuximab in patients with advanced colorectal cancer. Cancer. 2013;119(24):4223–30.

    Article  CAS  PubMed  Google Scholar 

  285. Clark AJ, Wiley DT, Zuckerman JE, Webster P, Chao J, Lin J, et al. CRLX101 nanoparticles localize in human tumors and not in adjacent, nonneoplastic tissue after intravenous dosing. Proc Natl Acad Sci. 2016;113(14):3850–4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  286. Weiss GJ, Chao J, Neidhart JD, Ramanathan RK, Bassett D, Neidhart JA, et al. First-in-human phase 1/2a trial of CRLX101, a cyclodextrin-containing polymer-camptothecin nanopharmaceutical in patients with advanced solid tumor malignancies. Invest New Drugs. 2013;31(4):986–1000.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  287. Schmidt KT, Chau CH, Strope JD, Huitema AD, Sissung TM, Price DK, et al. Antitumor activity of NLG207 (formerly CRLX101) in combination with enzalutamide in preclinical prostate cancer models. Mol Cancer Ther. 2021;20(5):915–24.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  288. Liu S, Jiang C, Yang L, Huang J, Peng R, Wang X, et al. First-line cetuximab improves the efficacy of subsequent bevacizumab for RAS wild-type left-sided metastatic colorectal cancer: an observational retrospective study. Sci Rep. 2020;10(1):1–12.

    Google Scholar 

  289. Batchelor TT, Duda DG, di Tomaso E, Ancukiewicz M, Plotkin SR, Gerstner E, et al. Phase II study of cediranib, an oral pan–vascular endothelial growth factor receptor tyrosine kinase inhibitor, in patients with recurrent glioblastoma. J Clin Oncol. 2010;28(17):2817.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  290. Kalpathy-Cramer J, Chandra V, Da X, Ou Y, Emblem KE, Muzikansky A, et al. Phase II study of tivozanib, an oral VEGFR inhibitor, in patients with recurrent glioblastoma. J Neurooncol. 2017;131(3):603–10.

    Article  CAS  PubMed  Google Scholar 

  291. Sonpavde G, Hutson TE, Rini BI. Axitinib for renal cell carcinoma. Expert Opin Investig Drugs. 2008;17(5):741–8.

    Article  CAS  PubMed  Google Scholar 

  292. Pick AM, Nystrom KK. Pazopanib for the treatment of metastatic renal cell carcinoma. Clin Ther. 2012;34(3):511–20.

    Article  CAS  PubMed  Google Scholar 

  293. Motzer RJ, Rini BI, Bukowski RM, Curti BD, George DJ, Hudes GR, et al. Sunitinib in patients with metastatic renal cell carcinoma. JAMA. 2006;295(21):2516–24.

    Article  CAS  PubMed  Google Scholar 

  294. Llovet JM, Ricci S, Mazzaferro V, Hilgard P, Gane E, Blanc J-F, et al. Sorafenib in advanced hepatocellular carcinoma. N Engl J Med. 2008;359(4):378–90.

    Article  CAS  PubMed  Google Scholar 

  295. Braghiroli MI, Sabbaga J, Hoff PM. Bevacizumab: overview of the literature. Expert Rev Anticancer Ther. 2012;12(5):567–80.

    Article  CAS  PubMed  Google Scholar 

  296. Majidpoor J, Mortezaee K. Angiogenesis as a hallmark of solid tumors-clinical perspectives. Cell Oncol. 2021;44(4):715–37.

    Article  CAS  Google Scholar 

  297. Socinski MA, Jotte RM, Cappuzzo F, Orlandi F, Stroyakovskiy D, Nogami N, et al. Atezolizumab for first-line treatment of metastatic nonsquamous NSCLC. N Engl J Med. 2018;378(24):2288–301.

    Article  CAS  PubMed  Google Scholar 

  298. Schmid P, Adams S, Rugo HS, Schneeweiss A, Barrios CH, Iwata H, et al. Atezolizumab and nab-paclitaxel in advanced triple-negative breast cancer. N Engl J Med. 2018;379(22):2108–21.

    Article  CAS  PubMed  Google Scholar 

  299. Horn L, Mansfield AS, Szczęsna A, Havel L, Krzakowski M, Hochmair MJ, et al. First-line atezolizumab plus chemotherapy in extensive-stage small-cell lung cancer. N Engl J Med. 2018;379(23):2220–9.

    Article  CAS  PubMed  Google Scholar 

  300. Larasati Y, Boudou C, Koval A, Katanaev VL. Unlocking the Wnt pathway: therapeutic potential of selective targeting FZD7 in cancer. Drug discovery today. 2021.

  301. Wigerinck LS. GLPG1790: the first ephrin (EPH) receptor tyrosine kinase inhibitor for the treatment of triple negative breast cancer. 2014.

  302. Vitiello P, Mele L, Prisco C, Cardone C, Ciardiello D, Poliero L, et al. GLPG 1790, a new selective EPHA2 inhibitor, is active in colorectal cancer cell lines belonging to the CMS4/mesenchymal-like subtype. Ann Oncol. 2019;30:v8–9.

    Article  Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

There is no foundation for this study.

Author information

Authors and Affiliations

Authors

Contributions

H.Z conceived the study and designed the headings. H.Z, F.A, M.L, M.B, M.N, A.A, and R.F wrote and revised the manuscript text. H.Z, F.A, and M.B created the figures. M.N created the table. N.R and H.Z supervised the study. All authors read and approved the final manuscript.

Corresponding authors

Correspondence to Hamidreza Zalpoor or Nima Rezaei.

Ethics declarations

Ethical approval and consent to participate

Not applicable.

Consent for publication

All authors have read the manuscript and given their consent for publication.

Competing interests

The authors declare no conflict of interest.

Additional information

Publisher's Note

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

Rights and permissions

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

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zalpoor, H., Aziziyan, F., Liaghat, M. et al. The roles of metabolic profiles and intracellular signaling pathways of tumor microenvironment cells in angiogenesis of solid tumors. Cell Commun Signal 20, 186 (2022). https://doi.org/10.1186/s12964-022-00951-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12964-022-00951-y

Keywords