Next Article in Journal
Fibroblast Growth Factor Receptors (FGFRs): Structures and Small Molecule Inhibitors
Next Article in Special Issue
The IL4I1 Enzyme: A New Player in the Immunosuppressive Tumor Microenvironment
Previous Article in Journal
Cortisol Excess-Mediated Mitochondrial Damage Induced Hippocampal Neuronal Apoptosis in Mice Following Cold Exposure
Previous Article in Special Issue
Nodal Facilitates Differentiation of Fibroblasts to Cancer-Associated Fibroblasts that Support Tumor Growth in Melanoma and Colorectal Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Distinct Roles of CXCR3 Variants and Their Ligands in the Tumor Microenvironment

1
Immuno-Pharmacology and Interactomics, Department of Infection and Immunity, Luxembourg Institute of Health (LIH), L-1526 Luxembourg, Luxembourg
2
Faculty of Science, Technology and Communication, University of Luxembourg, L-1526 Luxembourg, Luxembourg
3
Laboratory of Molecular Pharmacology, GIGA-Molecular Biology of Diseases, University of Liège, CHU, B-4000 Liège, Belgium
4
Laboratory of Experimental Cancer Research, Department of Oncology, Luxembourg Institute of Health (LIH), L-1526 Luxembourg, Luxembourg
5
Laboratory of Nervous System Diseases and Therapy, GIGA-Neuroscience, University of Liège, B-4000 Liège, Belgium
6
Neurology Department, CHU, Academic Hospital, University of Liège, B-4000 Liège, Belgium
7
NorLux Neuro-Oncology Laboratory, Department of Oncology, Luxembourg Institute of Health (LIH), L-1526 Luxembourg, Luxembourg
8
Research Centre, Saint-Justine Hospital, University of Montreal, Montréal, QC H3T 1C5, Canada
9
Department of Biochemistry, University of Montreal, Montréal, QC H3T 1J4, Canada
10
Laboratory of Medicinal Chemistry, Center for Interdisciplinary Research on Medicine (CIRM), University of Liège, CHU, B-4000 Liège, Belgium
*
Author to whom correspondence should be addressed.
These authors contributed equally.
Cells 2019, 8(6), 613; https://doi.org/10.3390/cells8060613
Submission received: 29 April 2019 / Revised: 14 June 2019 / Accepted: 16 June 2019 / Published: 18 June 2019
(This article belongs to the Special Issue Tumor Microenvironment: Interaction and Metabolism)

Abstract

:
First thought to orchestrate exclusively leukocyte trafficking, chemokines are now acknowledged for their multiple roles in the regulation of cell proliferation, differentiation, and survival. Dysregulation of their normal functions contributes to various pathologies, including inflammatory diseases and cancer. The two chemokine receptor 3 variants CXCR3-A and CXCR3-B, together with their cognate chemokines (CXCL11, CXCL10, CXCL9, CXCL4, and CXCL4L1), are involved in the control but also in the development of many tumors. CXCR3-A drives the infiltration of leukocytes to the tumor bed to modulate tumor progression (paracrine axis). Conversely, tumor-driven changes in the expression of the CXCR3 variants and their ligands promote cancer progression (autocrine axis). This review summarizes the anti- and pro-tumoral activities of the CXCR3 variants and their associated chemokines with a focus on the understanding of their distinct biological roles in the tumor microenvironment.

Graphical Abstract

1. Introduction

Chemokines, or chemotactic cytokines, are small, secretory proteins that form the largest group within the cytokine family [1]. They bind to their cognate chemokine receptors, which belong to the G protein-coupled receptor (GPCR) superfamily, and mediate cellular responses as a result of downstream signaling pathways initiated by the activation of G proteins and regulated by the recruitment of arrestins [1,2,3]. To date, 48 chemokines and 23 receptors have been identified in humans. The chemokine–receptor network is highly intricate; a chemokine can bind one or many receptors, while a receptor usually recognizes several chemokines [1,4]. Moreover, a recent paradigm shift brought to light the concept known as biased signaling according to which GPCRs, including chemokine receptors, depending on the ligand or cellular context, trigger distinct signaling pathways [3,5,6]. Chemokines primarily arbitrate various immune responses through the migration of chemokine-receptor-expressing leukocytes to the target tissues, while they also orchestrate the development and maintenance of lymphoid or non-lymphoid tissues by inducing the migration, proliferation, and differentiation of stromal and immune cells [7,8,9,10,11,12]. In addition, chemokines play key roles in the tumor microenvironment (TME) as they are secreted by and act on the different cellular components comprising tumor cells, T- and B-lymphocytes, natural killer (NK) cells, macrophages, dendritic cells (DCs), fibroblasts, and vascular and lymphatic endothelial cells and their associated pericytes. Together with other soluble factors such as growth factors, inflammatory cytokines, and extracellular matrix enzymes, they contribute to cancer development and the formation of a cancer-specific immunity, influencing the maturation, activation, and trafficking of tumor-infiltrating immune cells [11,13,14,15,16,17,18,19].
The interferon (IFN)-gamma inducible chemokines CXCL11, CXCL10, and CXCL9, and the platelet-derived chemokines CXCL4 and its non-allelic variant CXCL4L1, exert their biological function by binding to the IFN-gamma inducible chemokine receptor CXCR3 [6,20,21,22]. Interestingly, due to alternative splicing of the CXCR3 mRNA, three variants have been reported, CXCR3-A, CXCR3-B, and CXCR3-Alt, which exhibit differences in their expression profiles, their N or C terminus, and the number of transmembrane domains (Figure 1). These structural differences alter the physiological characteristics of the receptors, including ligand-binding properties, signaling pathways, and cellular responses [23]. CXCR3-A and CXCR3-B are the two variants that have been best studied. They differ by a 52-amino acid (AA) extension at the N terminus of CXCR3-B and exert opposing cellular effects [24]. On the one hand, binding of CXCL11, CXCL10, and CXCL9 to the CXCR3-A variant induces chemotaxis and proliferation of cells by activating Gi and Gq proteins and triggering their downstream signaling pathways, including the intracellular Ca2+ release, ERK1/2, and Akt pathways [5,23,25,26,27,28,29,30,31,32,33,34]. The activation of CXCR3-A by CXCL4 and its variant CXCL4L1 remains to be elucidated [30,31,35]. On the other hand, stimulation of CXCR3-B by these chemokines inhibits cell migration and proliferation and induces apoptosis. The mechanism behind this opposite cellular response driven by CXCR3-B is still unclear; however, one study suggests a coupling to the Gs protein [24].
While CXCR3 and its endogenous ligands are mainly involved in inflammation and wound healing [8,36], they have also been described to have a dual role in tumor progression and immunity. This review aims to outline the impact of the CXCR3 ligand–receptor axis and its expression changes on the TME with a focus on the CXCR3-A and CXCR3-B variants.

2. The Crosstalk of the CXCR3 Variants and Their Chemokine Ligands Within the Tumor Microenvironment

Chemokines are abundantly present in the TME and play key roles in inducing proliferation of benign and malignant cells, leukocyte migration, and angiogenesis [11,12,37,38]. These processes can be initiated and maintained in paracrine and autocrine fashions.

2.1. CXCR3-A on Leukocytes Mediates their Migration to the TME to Modulate Tumor Progression

The presence of tumor-infiltrating leukocytes (TIL) in the TME is known to influence tumor development [39]. Following the discovery of CXCR3-A on activated T-lymphocytes [25] and the anti-tumoral activity of CXCL10 [40], the possibility of an anti-tumoral response through the migration of CXCR3-A+ leukocytes to the TME was proposed [41]. The importance of CXCR3-dependent anti-tumoral activity was confirmed by Hensbergen et al. where CXCL11-producing EL4 lymphoma cells, injected in mice, were rejected due to the infiltration of CXCR3+ CD8+ T lymphocytes and macrophages [42]. Similar observations were reported in murine models of renal cell carcinoma (RCC) (RENCA) and spontaneous melanoma (B16F10) where the reduced tumor growth resulted from an increased presence of CXCR3-A-expressing CD4+ and CD8+ lymphocytes and NK cells in the tumor bed [43,44,45]. In another study, melanoma (B16F10) or breast cancer (E0771) cells injected in CXCR3−/− mice showed a significant increase in tumor growth compared to wild type (WT) mice, which was associated with a lower prevalence of CD8+ and CD4+ T cells as well as NK cells. This TIL-dependent anti-tumoral activity was furthermore validated in B16F10 tumor-bearing Rag2−/− mice, which showed a significantly increased tumor growth when transferred with CXCR3−/− cytotoxic T-lymphocytes (CTLs) compared to WT CTLs [46]. Such anti-tumoral activity of CXCR3-A+ leukocytes was also detected in human breast and gastric cancers as well as in RMA lymphoma [47,48,49]. Interestingly, the regressive characteristic of melanoma or certain melanocytic lesions was proposed to depend on the increased attraction of CXCR3-A+ cytotoxic lymphocytes to the TME [50].
In contrast, the higher prevalence of CXCR3+ regulatory T cells (Tregs) in human ovarian carcinomas was suggested to dampen the effector cell response, thus favoring the progression of the tumor [51]. This pro-tumoral effect of CXCR3+ Tregs was also observed in hepatocellular carcinoma (HCC), where a correlation could be made between CXCR3–CXCL10-dependent Treg infiltration and increased tumor growth and HCC recurrence after liver transplantation [52]. Moreover, in a chemically inducible murine model of skin carcinoma, CXCR3 knockout (KO) mice developed fewer tumors compared to WT mice. This observation was linked to a reduced presence of CXCR3-expressing CD3+ T cells, suggesting a cell proliferative effect on epidermal cells and an pro-tumoral activity of these TIL [53].
Taken together, these data suggest that the expression of CXCR3-A on leukocytes is needed to attract them to the TME and allows an anti-tumoral activity that diminishes tumor growth. However, the attraction of Tregs or other T-lymphocytes to the TME could also have a pro-tumoral effect by inducing cell proliferation and inhibiting the antitumor activity of effector leukocytes.
Of note, an adequate recruitment of CTLs in the tumor bed is not always observed. Correlated to a poor survival, it was for instance demonstrated that only 16% of patients with esophageal squamous cell carcinomas (ESCCs) or adenocarcinomas showed detectable CD8+ T cell infiltration within the tumor [54]. In fact, insufficient recruitment of activated T lymphocytes to the tumor bed is also recognized as one of the major hurdles in the current immunotherapeutic approaches [55]. Notably, the relevance of CXCR3-driven responses in immunotherapy has been demonstrated in several studies. In a B16-OVA mouse model, Mikucki et al. identified CXCR3-mediated trafficking at the tumor vasculature as important for effective T-cell-based cancer therapy by showing that CXCR3 is required for the extravasation of CTL to the tumor bed [45]. Similarly, in a B16 melanoma mouse model, the deletion of CXCR3 in CTLs lead to an impaired anti-tumoral response due to the failure of CTL infiltration in the TME upon anti-PD-1 treatment, suggesting CXCR3-dependent anti-PD-1 based anti-tumoral response [46]. Moreover, the blockade of PD-1 on naïve T cells enhanced CXCR3 expression and their absolute number in the spleen [56]. These studies show that CXCR3 is a potential target to increase the effectiveness of current immunotherapies.

2.2. CXCR3-A on Malignant Cells in the TME Contributes to Tumor Growth and Dissemination

As described above, the expression of CXCR3-A on leukocytes mediates their migration to the TME to exert their anti- and pro-tumoral activity; however, the CXCR3-A variant is also known to induce proliferation of various cell types, including human mesangial cells [24]. Therefore, the presence of CXCR3-A on tumor cells can induce an adverse effect by promoting and sustaining tumor development. Indeed, an increased proliferation of malignant glioma cells compared with normal astrocytes [57] and an anti-apoptotic effect in CXCR3-A-overexpressing human myeloma cell lines (HMCLs) was observed after CXCL10 treatment [58]. The increased expression of CXCR3-A on stage II colorectal cancer (CRC) and papillary thyroid cancer (PTC) cells was also associated with increased tumor development and negative prognosis [59,60]. In line with this, immunohistochemical analysis of localized clear cell RCC patient samples showed strong staining for CXCR3, which was correlated with a negative disease prognosis [61].
In addition to tumor development, expression of CXCR3-A on tumor cells promotes their dissemination. B16F10 melanoma injected into the footpads of mice showed macroscopic metastatic tumors in the lymph nodes already after one week, which could be attenuated with antisense CXCR3-A RNA [62]. An immunohistological analysis of primary invasive cutaneous melanoma samples from patients also supported the implication of CXCR3-A in tissue invasion and metastasis of malignant cells [63]. Furthermore, a metastatic potential conferred by CXCR3-A was observed in other human diseases, including epidermotropic and other B-cell disorders [64,65], glioblastoma (GBM) [66,67], CRC [68,69], high-grade serous ovarian cancer (HGSC) [70], lung adenocarcinoma [71] and breast cancer [72,73].
In conclusion, the expression of CXCR3-A on malignant cells leads to an increased proliferation and dissemination.

2.3. CXCR3-B Has an Anti-Proliferative Effect in the Tumor Microenvironment

In contrast to CXCR3-A, the CXCR3-B variant was shown to inhibit cell migration and proliferation [24]. CXCR3-B was first detected on human microvascular endothelial cells (HuMVECs) [24], but has also been found on CD4+ T lymphocytes [74], airway epithelial cells [75], and pericytes [76]. Although CXCR3-B is often expressed concomitantly with CXCR3-A, distinct CXCR3 variant expression levels can be observed on diverse cell types. Endothelial cells and pericytes abundantly express CXCR3-B [24,76], whereas CXCR3-A is not detectable [76,77], and activated T lymphocytes show higher CXCR3-A expression levels compared with CXCR3-B [31]. Various cancers evolve mechanisms to shift the CXCR3 variant expression levels towards ratios that are more beneficial for tumor maintenance or progression. For instance, human RCC downregulate CXCR3-B, which results in the upregulation of the anti-apoptotic heme oxygenase-1 (HO-1) and an increased tumor cell proliferation and dissemination [78,79]. Similarly, lower CXCR3-B mRNA levels were observed in ovarian and breast cancer cells compared with non-cancerous epithelial cells [80,81]. Moreover, decreasing the expression of CXCR3-B in breast cancer cells in vitro, using CXCR3-B-specific siRNA, resulted in increased cell proliferation and expression of anti-apoptotic proteins [81,82]. In addition to the inhibitory effect of CXCR3-B on cell proliferation, it has been observed that a rise in CXCR3-B expression could lead to increased tumor necrosis and lower the metastatic ability of tumor cells [83,84]. Of note, cancer stem cell-like properties upon CXCR3-B overexpression were reported in vitro in a breast cancer cell line, which could be prevented by silencing its mRNA [84].
Overall, it seems that cancer cells lower their expression of CXCR3-B to promote tumor proliferation, survival and dissemination and that this cellular effect can be reversed by enhancing CXCR3-B expression.

2.4. Is the CXCR3-B Variant Involved in Angiogenesis Regulation?

Tumors require adequate blood supply to obtain the sufficient oxygen, nutrients and other metabolic components to ensure their growth and survival. In contrast to normal angiogenesis, in which new blood vessels mature rapidly and show slow cell turn-over, tumor angiogenesis is an uncontrolled process where new blood vessels are formed in an irregular fashion [85,86]. Due to the high expression of CXCR3-B on endothelial cells and the angiostatic effect of CXCL10, CXCL9, and CXCL4 [24,87,88,89,90], it was proposed that the CXCR3-B variant could be responsible for this negative impact on tumor angiogenesis. Arenberg et al. indeed reported a decrease in angiogenesis and neovascularization in squamous cell carcinoma (SCC) after addition of CXCL10 in the TME that resulted in an attenuation in tumor growth and dissemination to the lung [91]. A similar observation was made following injection of CXCL9 to the tumor site in a mouse RENCA model [43]. In contrast, a rise in CXCL11 had no impact on the angiogenesis of mouse EL4 lymphoma cells nor human CRC [42,92]. Considering the expression of CXCR3 variants was not analyzed in these studies, one can only speculate that the inhibitory effect of these chemokines on tumor angiogenesis is driven through CXCR3-B present on endothelial cells, a phenomenon which may be cancer-type-dependent.
Although the above-mentioned data point to CXCR3-B-dependent angiostatic and inhibitory effects on cell proliferation, it has also been suggested that these are instead GAG-mediated. Indeed, both WT and CXCR3 KO endothelial cells showed an equivalent inhibitory effect on cell proliferation in the presence of CXCL10, which could not be reversed by neutralizing anti-CXCR3 antibodies. Moreover, the generation of a CXCL10 mutant devoid of CXCR3 signaling properties could also inhibit human umbilical cord endothelial cell (HUVEC) proliferation [93].

2.5. The CXCR3 Ligands are Upregulated in the TME and Act in an Autocrine and Paracrine Manner

Tumor cells as well as endothelial and immune cells secrete chemokines in the TME, which induce the recruitment of immune cells, modulate tumor immunity and promote tumor cell proliferation as well as metastasis [12]. Cancer cells also control the production and the activity of chemokines through the transcriptional silencing [94,95,96] or the post-translational modifications, via dipeptidyl peptidase 4 (DPP4 or CD26) [97,98]. These modifications undermine the attraction of tumor-invading lymphocytes, which impairs their anti-tumor effect [99,100] and the success of immunotherapy. In addition, CXCR3-endogenous ligands have been described to activate distinct signaling pathways, leading to specific cellular responses. For example, CXCL9 and CXCL10 induce effector Th1/Th17 cells and CXCL11 drives a Treg and Th2-biased effector cell polarization upon activation of CXCR3 on CD4+ T cells, while they upregulate tumor cell-related proteins that are beneficial for their survival, such as PD-L1 [21,101,102]. Here, we give an outline on the impact of the TME on the expression and secretion of CXCL11, CXCL10, CXCL9, CXCL4 and its variant CXCL4L1, and the effect of these chemokines on the different cellular components present within the TME.

2.5.1. CXCL11

Although CXCL11 is the most potent CXCR3 ligand [103], its role in CXCR3-driven processes is sometimes overlooked due to its impaired production in certain mouse strains; such as in C57BL/6 mice where a 2-bp insertion after the start codon results in a premature stop codon and ultimately CXCL11 null mutant mice [101,104,105]. CXCL11 is also less available compared with CXCL9 and CXCL10, whether in health or disease state [106,107,108]. For CXCL11, being the full agonist of CXCR3 [109], its production may also be submitted to stricter regulation. In addition, ACKR3, formerly known as CXCR7, is a scavenger receptor for CXCL11 that may reduce the availability of CXCL11 in vivo and in the tumor bed. Indeed, ACKR3 has been shown to be upregulated on various tumor cells as well as on tumor-associated vasculature, therefore potentially interfering with the CXCR3–CXCL11 interactions [103,110,111,112].
Nevertheless, several studies pinpointed the importance of CXCL11–CXCR3 axis within the TME. It has been reported that mice challenged with EL4 T-cell lymphoma cells genetically modified to produce murine CXCL11 attracted more CD8+ lymphocytes and macrophages with anti-tumoral activity to the TME compared to WT EL4 T-cell lymphoma cells [42]. On the other hand, a study on human colon adenocarcinoma showed enhanced tumor growth and invasiveness after injection of CXCL11 to the TME [92]. It has also been described that the repression of CXCL11 in CRC tissues diminished the tumor cell growth and metastasis [113]. In addition, the observed increased expression of CXCL11 in HCC cells was linked with the upregulation of stem cell-related genes and the acquisition and maintenance of self-renewal and tumorigenic properties of tumor-initiating cells through an autocrine axis [114].

2.5.2. CXCL10

Although CXCL10 is only a partial agonist of CXCR3, its impact on many human pathologies including cancer has been extensively studied. In glioma cells, for example, the expression of CXCL10 is significantly upregulated when compared with healthy astrocytes and has been shown to increase cell proliferation by acting on CXCR3 in an autocrine manner [57]. This pro-tumoral CXCL10 feed-forward loop was also observed in breast cancer; albeit, CXCL10 also attracted more CD4+ and CD8+ lymphocytes to the TME, in a paracrine manner, which attenuated tumor progression [115].
This pro- and anti-tumoral activity of CXCL10 could also be observed in CRC. Stage IV CRC patients showed an increased presence of CXCL10 in the TME, which was associated with a poor prognosis and promoted metastasis to the liver [116,117]. In contrast, higher secretion of CXCL10 in CRC increased the infiltration of CD8+ T lymphocytes and inhibited the ability of cancer cells to metastasize [118]. In addition, patients with ESCC showing higher CXCL10 expression had a better overall survival rate when compared with CXCL10-negative patients [119], while enhanced proliferation and metastatic behavior of melanoma tumor cells could be linked to higher CXCL10 secretion [120].
CXCL10 having opposing effects on TME that may occur simultaneously, the overall outcome of the disease seems to be tumor-type dependent and thus hard to predict.

2.5.3. CXCL9

Similar to CXCL10, CXCL9 is one of the most extensively studied CXCR3 chemokines in cancer and also plays a dual role in the TME. CXCL9 gene upregulation in CRC and non-small-cell-lung-cancer (NSCLC) patient samples correlated with an increased overall survival, suggesting a beneficial impact of higher secretion of this chemokine in the TME [121,122]. In line with this observation, higher secretion of CXCL9 in HGSC and fibrosarcoma tumors also resulted in an increased overall survival presumable due to an enhanced attraction of leukocytes, in particular T and NK cells, to the tumor bed [95,123]. The anti-tumoral activity of CXCL9 could also be observed in lymphocyte-rich gastric carcinoma (GC) and regressing Burkitt’s lymphoma tumors. These regressing tumors showed a higher prevalence of CXCL9 in the TME compared with conventional tumors and exhibited an increased presence of TIL and tumor necrosis rate [124,125]. In contrast, it has been reported that the higher CXCL9 levels in melanoma—secreted by the melanoma endothelial cells—disrupt the endothelial barrier, which in turn leads to a higher dissemination of melanoma cells. This transendothelial migration of the tumor cells was diminished upon addition of anti-CXCL9 monoclonal antibody (mAb) [126]. The CXCL9-mediated metastasis was also observed in tongue squamous cell carcinoma (TSCC). TSCC cell show an increased CXCL9 and CXCR3 expression, which results in cytoskeleton alterations and decreased cell adhesion molecules that promotes tumor cell invasion and migration to the lymph nodes [127]. Interestingly, in follicular lymphoma (FL), elevated levels of CXCL9 result in lower event-free survival after treatment with immunotherapy implying a potential negative role of this chemokine after cancer treatment [128]. Thus, depending on the tumor cell type, CXCL9 can have either an adverse or a beneficial effect. This beneficial effect appears to rely on leukocyte attraction to the TME in a paracrine fashion; however, the adverse effect of CXCL9 on the malignant cell might not be direct as for CXCL10.

2.5.4. CXCL4 and its CXCL4L1 Variant

While CXCL4 is mainly released after activation of platelets, it can also be secreted by stromal cells and tumor cells. In the two most frequent types of human lung adenocarcinomas (KRAS and TP53) and in CRC, an increase in CXCL4 expression was detected and correlated to lower patient survival probably due to its proliferative effect on tumor cells and anti-proliferative effect on CTLs [129,130,131]. The CXCL4L1 non-allelic variant has also been described to bind to CXCR3-A and -B [132]. The production of CXCL4L1 in melanoma, lung carcinoma and pancreatic adenocarcinoma (PDAC) inhibits tumor growth as well as neovascularization. Furthermore, CXCL4L1, as well as CXCL4, are able to recruit T cells, NK cells and DC to the TME, which may contribute to their anti-tumoral activity [132,133,134,135].

3. Technical Hurdles to Studying the CXCR3 Variants

The identification of the CXCR3-B variant, with its opposite cellular responses, raised new issues pertaining to the study and validation of the exact expression profile, signaling and the differential contribution of the two isoforms in CXCR3-mediated processes. Variant-specific tools were thus needed to be able to attribute the various cellular responses on the different cell types to CXCR3-A or CXCR3-B.
Nucleotide probes, or primers, were the first easily accessible tools to detect and distinguish endogenous expression of CXCR3-A- and CXCR3-B-specific transcripts in tissue and cells. The most frequently used primers, described by Lasagni et al., detect the spliced mRNA encoding both CXCR3-A and CXCR3-B as they bind to sequences that are common to the two variants. Nevertheless, the PCR products can be distinguished based on their different DNA band size. Primers designed for specific detection of CXCR3-A or CXCR3-B were also reported [24] (Figure 1). The CXCR3-A-specific primer aligns with the nucleotide sequence coding for the first seven amino acids present only in this variant [136] while the primer specific to CXCR3-B aligns within its extended N terminus. Besides PCR-based approaches, another application of nucleotide probes is in situ hybridization. This technique was used to screen for mRNA expression of the CXCR3 variants directly in slices of fixed tissues allowing the direct visualization of the receptor distribution and identification of receptor-positive cell types [137].
Variant-specific nucleotide probes proved to be easily manufactured and useful tools to evaluate the gene expression levels. However, confirming the presence of CXCR3 at protein level and, in particular, on the cell surface turned out to be a more difficult task. For the time being, the anti-CXCR3-A antibodies also recognize the CXCR3-B variant due to the usage of CXCR3-A N-terminus-derived peptides for immunization. The production of the first CXCR3-B-specific mAb raised against the extended 52-AA N terminus and its use in immunofluorescence and immunohistochemistry was first described by Lasagni et al. [24]. Meanwhile, other CXCR3-B-specific mAbs, using different CXCR3-B N-terminus-derived peptides as immunogen, are available but usually show an application spectrum limited to ELISA or Western Blot. To circumvent the staining of both CXCR3 isoforms, an alternative may be the use of conformational antibodies against each variant as previously generated for other GPCRs or chemokine receptors, including CXCR4 [138] CCR5 [139] and ACKR3 [109]. Indeed, the opposite cellular response of the two CXCR3 variants suggests that the presence of the N-terminal extension of CXCR3-B modifies the general tertiary structure of the receptor and might therefore expose different discontinued epitopes compared with CXCR3-A. Although their use would be limited to the detection of CXCR3 in its native conformation, it would help to confirm and quantify the presence of the two variants on the cell surface and understand the biology behind the selective receptor expression on the various cell types of the TME.
The diverse cellular effects of the CXCR3 variants were proposed to rely on the activation of different downstream signaling pathways. To investigate and identify CXCR3 variant-specific pathways, novel variant-selective modulators (antagonists, agonists or allosteric modulators) are needed as the currently used endogenous chemokines bind and activate both variants almost indifferently. Although CXCR3-A small-molecule modulators have been generated, their activity on CXCR3-B remains to be elucidated [106,140,141,142]. In addition, no CXCR3-B-specific small molecules have been reported so far. Therefore, the generation of small modulators is timely and would enable to unravel the CXCR3 variants signaling and could hold potential therapeutic value.

4. Conclusions, Challenges, and Future Directions

Over the last two decades, much effort has been made to understand the biology of the chemokine receptor CXCR3. Although the mechanism controlling the alternative splicing of both CXCR3 variants remains to be elucidated, numerous data on their signaling pathways and their respective contribution in tumor development and immunity have been generated.
On the one hand, it was shown that tumor cells can manipulate the expression of the CXCR3 variants in a way that is overall more beneficial for their maintenance. Indeed, by upregulating the expression of CXCR3-A and lowering the expression of CXCR3-B at their surface, tumor-initiating cells are able to enhance their proliferative potential and survival through autocrine stimulation loops that are further reinforced by the simultaneous upregulation of the expression of the CXCR3 ligands in the TME.
On the other hand, the increased secretion of CXCR3 ligands within the TME was shown to enhance the recruitment and the infiltration of various leukocytes into the tumor bed in a paracrine fashion, thereby modulating tumor progression. In a similar manner, chemokines participate in the control of metastasis of CXCR3-expressing tumor cells (summarized in Figure 2).
The demonstration of the impact of the CXCR3 axis in cancer points to its potential interest as a therapeutic target. However, the dual activity of CXCR3-A in the proliferative loops and the attraction of immune cells to the tumor bed, and the opposing roles of the CXCR3 variants’ ligand–receptor axis, as well as ligand and cell context bias, are important aspects that need to be considered for the development and use of anti-CXCR3 drugs or neutralizing antibodies. In addition, other players in the TME, including growth factors, inflammatory cytokines, and extracellular matrix enzymes, and the impact of the tumor on the microenvironment, such as induction of hypoxia, influence the CXCR3 axis and add yet another level of complexity to CXCR3-regulated processes, ultimately making the outcome of potential blockade hard to predict.
Therefore, a better understanding of all aspects of the biology of CXCR3 variants, including their exact signaling pathways and cellular responses in the tumor-microenvironment context, is urgently needed before envisaging the CXCR3 variants as potential anti-tumoral targets and the development of CXCR3 variant-specific drugs to be used for combined therapy, for example with anti-PD-1 immunotherapy, to turn cold to hot tumors.

Funding

This study was supported by the Luxembourg Institute of Health (LIH) MESR (grants 20160116 and 20170113), Luxembourg National Research Fund (INTER/FWO “Nanokine” - grant 15/10358798), F.R.S.-FNRS-Télévie (grants 7456814, 7461515, 34985948 and 35013979). N. Reynders is part of the Luxembourg National Research Fund PhD fellows (PRIDE-11012546 “NextImmune”). Julian Hanson is a F.R.S.-FNRS Research Associate.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zlotnik, A.; Yoshie, O. The chemokine superfamily revisited. Immunity 2012, 36, 705–716. [Google Scholar] [CrossRef] [PubMed]
  2. Thelen, M. Dancing to the tune of chemokines. Nat. Immunol. 2001, 2, 129–134. [Google Scholar] [CrossRef] [PubMed]
  3. Kleist, A.B.; Getschman, A.E.; Ziarek, J.J.; Nevins, A.M.; Gauthier, P.-A.; Chevigné, A.; Szpakowska, M.; Volkman, B.F. New paradigms in chemokine receptor signal transduction: Moving beyond the two-site model. Biochem. Pharmacol. 2016, 114, 53–68. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Szpakowska, M.; Dupuis, N.; Baragli, A.; Counson, M.; Hanson, J.; Piette, J.; Chevigné, A. Human herpesvirus 8-encoded chemokine vCCL2/vMIP-II is an agonist of the atypical chemokine receptor ACKR3/CXCR7. Biochem. Pharmacol. 2016, 114, 14–21. [Google Scholar] [CrossRef] [PubMed]
  5. Smith, J.S.; Alagesan, P.; Desai, N.K.; Pack, T.F.; Wu, J.-H.; Inoue, A.; Freedman, N.J.; Rajagopal, S. CXC motif chemokine receptor 3 splice variants differentially activate beta-arrestins to regulate downstream signaling pathways. Mol. Pharmacol. 2017, 92, 136–150. [Google Scholar] [CrossRef]
  6. Metzemaekers, M.; Vanheule, V.; Janssens, R.; Struyf, S.; Proost, P. Overview of the mechanisms that may contribute to the non-redundant activities of interferon-inducible CXC chemokine receptor 3 ligands. Front. Immunol. 2018, 8, 1970. [Google Scholar] [CrossRef] [PubMed]
  7. Moser, B.; Loetscher, P. Lymphocyte traffic control by chemokines. Nat. Immunol. 2001, 2, 123–128. [Google Scholar] [CrossRef]
  8. Griffith, J.W.; Sokol, C.L.; Luster, A.D. Chemokines and chemokine receptors: Positioning cells for host defense and immunity. Annu. Rev. Immunol. 2014, 32, 659–702. [Google Scholar] [CrossRef]
  9. Thelen, M.; Stein, J.V. How chemokines invite leukocytes to dance. Nat. Immunol. 2008, 9, 953–959. [Google Scholar] [CrossRef]
  10. Luster, A.D. Chemokines—Chemotactic cytokines that mediate inflammation. N. Engl. J. Med. 1998, 338, 436–445. [Google Scholar] [CrossRef]
  11. Balkwill, F. Cancer and the chemokine network. Nat. Rev. Cancer 2004, 4, 540–550. [Google Scholar] [CrossRef] [PubMed]
  12. Nagarsheth, N.; Wicha, M.S.; Zou, W. Chemokines in the cancer microenvironment and their relevance in cancer immunotherapy. Nat. Rev. Immunol. 2017, 17, 559–572. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Relation, T.; Dominici, M.; Horwitz, E.M. Concise review: An (Im) Penetrable shield: How the tumor microenvironment protects cancer stem cells. Stem Cells 2017, 35, 1123–1130. [Google Scholar] [CrossRef] [PubMed]
  14. Binnewies, M.; Roberts, E.W.; Kersten, K.; Chan, V.; Fearon, D.F.; Merad, M.; Coussens, L.M.; Gabrilovich, D.I.; Ostrand-Rosenberg, S.; Hedrick, C.C. Understanding the tumor immune microenvironment (TIME) for effective therapy. Nat. Med. 2018, 24, 541–550. [Google Scholar] [CrossRef] [PubMed]
  15. Zou, W. Immunosuppressive networks in the tumour environment and their therapeutic relevance. Nat. Rev. Cancer 2005, 5, 263–274. [Google Scholar] [CrossRef] [PubMed]
  16. Franciszkiewicz, K.; Boissonnas, A.; Boutet, M.; Combadière, C.; Mami-Chouaib, F. Role of chemokines and chemokine receptors in shaping the effector phase of the antitumor immune response. Cancer Res. 2012, 72, 6325–6332. [Google Scholar] [CrossRef]
  17. Witsch, E.; Sela, M.; Yarden, Y. Roles for growth factors in cancer progression. Physiology 2010, 25, 85–101. [Google Scholar] [CrossRef]
  18. Wang, M.; Zhao, J.; Zhang, L.; Wei, F.; Lian, Y.; Wu, Y.; Gong, Z.; Zhang, S.; Zhou, J.; Cao, K. Role of tumor microenvironment in tumorigenesis. J. Cancer 2017, 8, 761–773. [Google Scholar] [CrossRef]
  19. Balkwill, F.R.; Capasso, M.; Hagemann, T. The tumor microenvironment at a glance. J. Cell Sci. 2012, 125, 5591–5596. [Google Scholar] [CrossRef] [Green Version]
  20. Groom, J.R.; Luster, A.D. CXCR3 ligands: Redundant, collaborative and antagonistic functions. Immunol. Cell Biol. 2011, 89, 207–215. [Google Scholar] [CrossRef]
  21. Tokunaga, R.; Zhang, W.; Naseem, M.; Puccini, A.; Berger, M.D.; Soni, S.; McSkane, M.; Baba, H.; Lenz, H.-J. CXCL9, CXCL10, CXCL11/CXCR3 axis for immune activation—A target for novel cancer therapy. Cancer Treat. Rev. 2018, 63, 40–47. [Google Scholar] [CrossRef] [PubMed]
  22. Van Raemdonck, K.; Van den Steen, P.E.; Liekens, S.; Van Damme, J.; Struyf, S. CXCR3 ligands in disease and therapy. Cytokine Growth Factor Rev. 2015, 26, 311–327. [Google Scholar] [CrossRef] [PubMed]
  23. Berchiche, Y.A.; Sakmar, T.P. CXC chemokine receptor 3 alternative splice variants selectively activate different signaling pathways. Mol. Pharmacol. 2016, 90, 483–495. [Google Scholar] [CrossRef] [PubMed]
  24. Lasagni, L.; Francalanci, M.; Annunziato, F.; Lazzeri, E.; Giannini, S.; Cosmi, L.; Sagrinati, C.; Mazzinghi, B.; Orlando, C.; Maggi, E. An alternatively spliced variant of CXCR3 mediates the inhibition of endothelial cell growth induced by IP-10, Mig, and I-TAC, and acts as functional receptor for platelet factor 4. J. Exp. Med. 2003, 197, 1537–1549. [Google Scholar] [CrossRef] [PubMed]
  25. Loetscher, M.; Gerber, B.; Loetscher, P.; Jones, S.A.; Piali, L.; Clark-Lewis, I.; Baggiolini, M.; Moser, B. Chemokine receptor specific for IP10 and mig: Structure, function, and expression in activated T-lymphocytes. J. Exp. Med. 1996, 184, 963–969. [Google Scholar] [CrossRef]
  26. Cole, K.E.; Strick, C.A.; Paradis, T.J.; Ogborne, K.T.; Loetscher, M.; Gladue, R.P.; Lin, W.; Boyd, J.G.; Moser, B.; Wood, D.E. Interferon–inducible T cell alpha chemoattractant (I-TAC): A novel Non-ELR CXC Chemokine with potent activity on activated T cells through selective high affinity binding to CXCR3. J. Exp. Med. 1998, 187, 2009–2021. [Google Scholar] [CrossRef]
  27. Loetscher, M.; Loetscher, P.; Brass, N.; Meese, E.; Moser, B. Lymphocyte-specific chemokine receptor CXCR3: Regulation, chemokine binding and gene localization. Eur. J. Immunol. 1998, 28, 3696–3705. [Google Scholar] [CrossRef]
  28. Thompson, B.D.; Jin, Y.; Wu, K.H.; Colvin, R.A.; Luster, A.D.; Birnbaumer, L.; Wu, M.X. Inhibition of Gai2 activation by Gai3 in CXCR3-mediated signaling. J. Biol. Chem. 2007, 282, 9547–9555. [Google Scholar] [CrossRef]
  29. Smit, M.J.; Verdijk, P.; van der Raaij-Helmer, E.M.; Navis, M.; Hensbergen, P.J.; Leurs, R.; Tensen, C.P. CXCR3-mediated chemotaxis of human T cells is regulated by a Gi-and phospholipase C-dependent pathway and not via activation of MEK/p44/p42 MAPK nor Akt/PI-3 kinase. Blood 2003, 102, 1959–1965. [Google Scholar] [CrossRef]
  30. Mueller, A.; Meiser, A.; McDonagh, E.M.; Fox, J.M.; Petit, S.J.; Xanthou, G.; Williams, T.J.; Pease, J.E. CXCL4-induced migration of activated T lymphocytes is mediated by the chemokine receptor CXCR3. J. Leukoc. Biol. 2008, 83, 875–882. [Google Scholar] [CrossRef]
  31. Korniejewska, A.; McKnight, A.J.; Johnson, Z.; Watson, M.L.; Ward, S.G. Expression and agonist responsiveness of CXCR3 variants in human T lymphocytes. Immunology 2011, 132, 503–515. [Google Scholar] [CrossRef] [PubMed]
  32. Watts, A.O.; Scholten, D.J.; Heitman, L.H.; Vischer, H.F.; Leurs, R. Label-free impedance responses of endogenous and synthetic chemokine receptor CXCR3 agonists correlate with G i-protein pathway activation. Biochem. Biophys. Res. Commun. 2012, 419, 412–418. [Google Scholar] [CrossRef] [PubMed]
  33. Romagnani, P.; Beltrame, C.; Annunziato, F.; Lasagni, L.; Luconi, M.; Galli, G.; Cosmi, L.; Maggi, E.; Salvadori, M.; Pupilli, C. Role for interactions between IP-10/Mig and CXCR3 in proliferative glomerulonephritis. J. Am. Soc. Nephrol. 1999, 10, 2518–2526. [Google Scholar] [PubMed]
  34. Bonacchi, A.; Romagnani, P.; Romanelli, R.G.; Efsen, E.; Annunziato, F.; Lasagni, L.; Francalanci, M.; Serio, M.; Laffi, G.; Pinzani, M. Signal transduction by the chemokine receptor CXCR3: Activation of Ras/ERK, Src, and phosphatidylinositol 3-kinase/Akt controls cell migration and proliferation in human vascular pericytes. J. Biol. Chem. 2001, 276, 9945–9954. [Google Scholar] [CrossRef] [PubMed]
  35. Chen, Y.-P.; Wu, H.-L.; Boyé, K.; Pan, C.-Y.; Chen, Y.-C.; Pujol, N.; Lin, C.-W.; Chiu, L.-Y.; Billottet, C.; Alves, I.D. Oligomerization state of CXCL4 chemokines regulates G protein-coupled receptor activation. Acs Chem. Biol. 2017, 12, 2767–2778. [Google Scholar] [CrossRef] [PubMed]
  36. Billottet, C.; Quemener, C.; Bikfalvi, A. CXCR3, a double-edged sword in tumor progression and angiogenesis. Biochim. Et Biophys. Acta (BBA) Rev. Cancer 2013, 1836, 287–295. [Google Scholar] [CrossRef]
  37. Strieter, R.M.; Burdick, M.D.; Gomperts, B.N.; Belperio, J.A.; Keane, M.P. CXC chemokines in angiogenesis. Cytokine Growth Factor Rev. 2005, 16, 593–609. [Google Scholar] [CrossRef] [Green Version]
  38. Mgrditchian, T.; Arakelian, T.; Paggetti, J.; Noman, M.Z.; Viry, E.; Moussay, E.; Van Moer, K.; Kreis, S.; Guerin, C.; Buart, S. Targeting autophagy inhibits melanoma growth by enhancing NK cells infiltration in a CCL5-dependent manner. Proc. Natl. Acad. Sci. USA 2017, 114, E9271–E9279. [Google Scholar] [CrossRef] [Green Version]
  39. Gooden, M.J.; de Bock, G.H.; Leffers, N.; Daemen, T.; Nijman, H.W. The prognostic influence of tumour-infiltrating lymphocytes in cancer: A systematic review with meta-analysis. Br. J. Cancer 2011, 105, 93–103. [Google Scholar] [CrossRef]
  40. Luster, A.D.; Leder, P. IP-10, a-CXC-chemokine, elicits a potent thymus-dependent antitumor response in vivo. J. Exp. Med. 1993, 178, 1057–1065. [Google Scholar] [CrossRef]
  41. Mullins, I.M.; Slingluff, C.L.; Lee, J.K.; Garbee, C.F.; Shu, J.; Anderson, S.G.; Mayer, M.E.; Knaus, W.A.; Mullins, D.W. CXC chemokine receptor 3 expression by activated CD8+ T cells is associated with survival in melanoma patients with stage III disease. Cancer Res. 2004, 64, 7697–7701. [Google Scholar] [CrossRef] [PubMed]
  42. Hensbergen, P.J.; Wijnands, P.G.B.; Schreurs, M.W.; Scheper, R.J.; Willemze, R.; Tensen, C.P. The CXCR3 targeting chemokine CXCL11 has potent antitumor activity in vivo involving attraction of CD8+ T lymphocytes but not inhibition of angiogenesis. J. Immunother. 2005, 28, 343–351. [Google Scholar] [CrossRef] [PubMed]
  43. Pan, J.; Burdick, M.D.; Belperio, J.A.; Xue, Y.Y.; Gerard, C.; Sharma, S.; Dubinett, S.M.; Strieter, R.M. CXCR3/CXCR3 ligand biological axis impairs RENCA tumor growth by a mechanism of immunoangiostasis. J. Immunol. 2006, 176, 1456–1464. [Google Scholar] [CrossRef] [PubMed]
  44. Reckamp, K.L.; Figlin, R.A.; Moldawer, N.; Pantuck, A.J.; Belldegrun, A.S.; Burdick, M.D.; Strieter, R.M. Expression of CXCR3 on mononuclear cells and CXCR3 ligands in patients with metastatic renal cell carcinoma in response to systemic IL-2 therapy. J. Immunother. 2007, 30, 417–424. [Google Scholar] [CrossRef] [PubMed]
  45. Mikucki, M.; Fisher, D.; Matsuzaki, J.; Skitzki, J.; Gaulin, N.; Muhitch, J.; Ku, A.; Frelinger, J.; Odunsi, K.; Gajewski, T. Non-redundant requirement for CXCR3 signalling during tumoricidal T-cell trafficking across tumour vascular checkpoints. Nat. Commun. 2015, 6, 7458. [Google Scholar] [CrossRef] [PubMed]
  46. Chheda, Z.S.; Sharma, R.K.; Jala, V.R.; Luster, A.D.; Haribabu, B. Chemoattractant receptors BLT1 and CXCR3 regulate antitumor immunity by facilitating CD8+ T cell migration into tumors. J. Immunol. 2016, 197, 2016–2026. [Google Scholar] [CrossRef]
  47. Oghumu, S.; Varikuti, S.; Terrazas, C.; Kotov, D.; Nasser, M.W.; Powell, C.A.; Ganju, R.K.; Satoskar, A.R. CXCR 3 deficiency enhances tumor progression by promoting macrophage M2 polarization in a murine breast cancer model. Immunology 2014, 143, 109–119. [Google Scholar] [CrossRef]
  48. Li, K.; Zhu, Z.; Luo, J.; Fang, J.; Zhou, H.; Hu, M.; Maskey, N.; Yang, G. Impact of chemokine receptor CXCR3 on tumor-infiltrating lymphocyte recruitment associated with favorable prognosis in advanced gastric cancer. Int. J. Clin. Exp. Pathol. 2015, 8, 14725. [Google Scholar]
  49. Wendel, M.; Galani, I.E.; Suri-Payer, E.; Cerwenka, A. Natural killer cell accumulation in tumors is dependent on IFN-γ and CXCR3 ligands. Cancer Res. 2008, 68, 8437–8445. [Google Scholar] [CrossRef]
  50. Wenzel, J.; Bekisch, B.; Uerlich, M.; Haller, O.; Bieber, T.; Tüting, T. Type I interferon–associated recruitment of cytotoxic lymphocytes: A common mechanism in regressive melanocytic lesions. Am. J. Clin. Pathol. 2005, 124, 37–48. [Google Scholar] [CrossRef]
  51. Redjimi, N.; Raffin, C.; Raimbaud, I.; Pignon, P.; Matsuzaki, J.; Odunsi, K.; Valmori, D.; Ayyoub, M. CXCR3+ T regulatory cells selectively accumulate in human ovarian carcinomas to limit type I immunity. Cancer Res. 2012, 72, 4351–4360. [Google Scholar] [CrossRef] [PubMed]
  52. Li, C.X.; Ling, C.C.; Shao, Y.; Xu, A.; Li, X.C.; Ng, K.T.-P.; Liu, X.B.; Ma, Y.Y.; Qi, X.; Liu, H. CXCL10/CXCR3 signaling mobilized-regulatory T cells promote liver tumor recurrence after transplantation. J. Hepatol. 2016, 65, 944–952. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Winkler, A.E.; Brotman, J.J.; Pittman, M.E.; Judd, N.P.; Lewis, J.S.; Schreiber, R.D.; Uppaluri, R. CXCR3 enhances a T cell dependent epidermal proliferative response and promotes skin tumorigenesis. Cancer Res. 2011, 71, 5707–5716. [Google Scholar] [CrossRef] [PubMed]
  54. Schumacher, K.; Haensch, W.; Röefzaad, C.; Schlag, P.M. Prognostic significance of activated CD8+ T cell infiltrations within esophageal carcinomas. Cancer Res. 2001, 61, 3932–3936. [Google Scholar] [PubMed]
  55. Sackstein, R.; Schatton, T.; Barthel, S.R. T-lymphocyte homing: An underappreciated yet critical hurdle for successful cancer immunotherapy. Lab. Investig. 2017, 97, 669–697. [Google Scholar] [CrossRef] [PubMed]
  56. Pauken, K.E.; Jenkins, M.K.; Azuma, M.; Fife, B.T. PD-1, but not PD-L1, expressed by islet-reactive CD4+ T cells suppresses infiltration of the pancreas during type 1 diabetes. Diabetes 2013, 62, 2859–2869. [Google Scholar] [CrossRef] [PubMed]
  57. Maru, S.V.; Holloway, K.A.; Flynn, G.; Lancashire, C.L.; Loughlin, A.J.; Male, D.K.; Romero, I.A. Chemokine production and chemokine receptor expression by human glioma cells: Role of CXCL10 in tumour cell proliferation. J. Neuroimmunol. 2008, 199, 35–45. [Google Scholar] [CrossRef] [PubMed]
  58. Giuliani, N.; Bonomini, S.; Romagnani, P.; Lazzaretti, M.; Morandi, F.; Colla, S.; Tagliaferri, S.; Lasagni, L.; Annunziato, F.; Crugnola, M. CXCR3 and its binding chemokines in myeloma cells: Expression of isoforms and potential relationships with myeloma cell proliferation and survival. Haematologica 2006, 91, 1489–1497. [Google Scholar]
  59. Bai, M.; Chen, X.; Ba, Y. CXCL10/CXCR3 overexpression as a biomarker of poor prognosis in patients with stage II colorectal cancer. Mol. Clin. Oncol. 2016, 4, 23–30. [Google Scholar] [CrossRef]
  60. Urra, S.; Fischer, M.C.; Martínez, J.R.; Véliz, L.; Orellana, P.; Solar, A.; Bohmwald, K.; Kalergis, A.; Riedel, C.; Corvalán, A.H. Differential expression profile of CXCR3 splicing variants is associated with thyroid neoplasia. Potential role in papillary thyroid carcinoma oncogenesis? Oncotarget 2018, 9, 2445–2467. [Google Scholar] [CrossRef]
  61. Klatte, T.; Seligson, D.B.; Leppert, J.T.; Riggs, S.B.; Yu, H.; Zomorodian, N.; Kabbinavar, F.F.; Strieter, R.M.; Belldegrun, A.S.; Pantuck, A.J. The chemokine receptor CXCR3 is an independent prognostic factor in patients with localized clear cell renal cell carcinoma. J. Urol. 2008, 179, 61–66. [Google Scholar] [CrossRef] [PubMed]
  62. Kawada, K.; Sonoshita, M.; Sakashita, H.; Takabayashi, A.; Yamaoka, Y.; Manabe, T.; Inaba, K.; Minato, N.; Oshima, M.; Taketo, M.M. Pivotal role of CXCR3 in melanoma cell metastasis to lymph nodes. Cancer Res. 2004, 64, 4010–4017. [Google Scholar] [CrossRef] [PubMed]
  63. Monteagudo, C.; Martin, J.M.; Jorda, E.; Llombart-Bosch, A. CXCR3 chemokine receptor immunoreactivity in primary cutaneous malignant melanoma: Correlation with clinicopathological prognostic factors. J. Clin. Pathol. 2007, 60, 596–599. [Google Scholar] [CrossRef] [PubMed]
  64. Magro, C.M.; Momtahen, S.; Lee, B.A.; Swanson, D.L.; Pavlovic, M.D. Epidermotropic B-cell lymphoma: A unique subset of CXCR3-positive marginal zone lymphoma. Am. J. Dermatopathol. 2016, 38, 105–112. [Google Scholar] [CrossRef] [PubMed]
  65. Trentin, L.; Agostini, C.; Facco, M.; Piazza, F.; Perin, A.; Siviero, M.; Gurrieri, C.; Galvan, S.; Adami, F.; Zambello, R. The chemokine receptor CXCR3 is expressed on malignant B cells and mediates chemotaxis. J. Clin. Investig. 1999, 104, 115–121. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Liu, C.; Luo, D.; Reynolds, B.A.; Meher, G.; Katritzky, A.R.; Lu, B.; Gerard, C.J.; Bhadha, C.P.; Harrison, J.K. Chemokine receptor CXCR3 promotes growth of glioma. Carcinogenesis 2010, 32, 129–137. [Google Scholar] [CrossRef]
  67. Pu, Y.; Li, S.; Zhang, C.; Bao, Z.; Yang, Z.; Sun, L. High expression of CXCR3 is an independent prognostic factor in glioblastoma patients that promotes an invasive phenotype. J. Neuro-Oncol. 2015, 122, 43–51. [Google Scholar] [CrossRef] [PubMed]
  68. Murakami, T.; Kawada, K.; Iwamoto, M.; Akagami, M.; Hida, K.; Nakanishi, Y.; Kanda, K.; Kawada, M.; Seno, H.; Taketo, M.M. The role of CXCR3 and CXCR4 in colorectal cancer metastasis. Int. J. Cancer 2013, 132, 276–287. [Google Scholar] [CrossRef]
  69. Kawada, K.; Hosogi, H.; Sonoshita, M.; Sakashita, H.; Manabe, T.; Shimahara, Y.; Sakai, Y.; Takabayashi, A.; Oshima, M.; Taketo, M. Chemokine receptor CXCR3 promotes colon cancer metastasis to lymph nodes. Oncogene 2007, 26, 4679–4688. [Google Scholar] [CrossRef]
  70. Windmüller, C.; Zech, D.; Avril, S.; Boxberg, M.; Dawidek, T.; Schmalfeldt, B.; Schmitt, M.; Kiechle, M.; Bronger, H. CXCR3 mediates ascites-directed tumor cell migration and predicts poor outcome in ovarian cancer patients. Oncogenesis 2017, 6, e331. [Google Scholar] [CrossRef]
  71. Duruisseaux, M.; Rabbe, N.; Antoine, M.; Vieira, T.; Poulot, V.; Cadranel, J.; Wislez, M. Pro-tumoural CXCL10/CXCR3-A autocrine loop in invasive mucinous lung adenocarcinoma. ERJ Open Res. 2017, 3. [Google Scholar] [CrossRef] [PubMed]
  72. Ma, X.; Norsworthy, K.; Kundu, N.; Rodgers, W.H.; Gimotty, P.A.; Goloubeva, O.; Lipsky, M.; Li, Y.; Holt, D.; Fulton, A. CXCR3 expression is associated with poor survival in breast cancer and promotes metastasis in a murine model. Mol. Cancer Ther. 2009, 8, 490–498. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Zhu, G.; Yan, H.H.; Pang, Y.; Jian, J.; Achyut, B.R.; Liang, X.; Weiss, J.M.; Wiltrout, R.H.; Hollander, M.C.; Yang, L. CXCR3 as a molecular target in breast cancer metastasis: Inhibition of tumor cell migration and promotion of host anti-tumor immunity. Oncotarget 2015, 6, 43408–43419. [Google Scholar] [CrossRef] [PubMed]
  74. Romagnani, P.; Maggi, L.; Mazzinghi, B.; Cosmi, L.; Lasagni, L.; Liotta, F.; Lazzeri, E.; Angeli, R.; Rotondi, M.; Filì, L. CXCR3-mediated opposite effects of CXCL10 and CXCL4 on TH1 or TH2 cytokine production. J. Allergy Clin. Immunol. 2005, 116, 1372–1379. [Google Scholar] [CrossRef] [PubMed]
  75. Kelsen, S.G.; Aksoy, M.O.; Yang, Y.; Shahabuddin, S.; Litvin, J.; Safadi, F.; Rogers, T.J. The chemokine receptor CXCR3 and its splice variant are expressed in human airway epithelial cells. Am. J. Physiol. Lung Cell. Mol. Physiol. 2004, 287, L584–L591. [Google Scholar] [CrossRef] [PubMed]
  76. Bodnar, R.J.; Wells, A. Differential regulation of pericyte function by the CXC receptor 3. Wound Repair Regen. 2015, 23, 785–796. [Google Scholar] [CrossRef] [PubMed]
  77. Murdoch, C.; Monk, P.N.; Finn, A. Cxc chemokine receptor expression on human endothelial cells. Cytokine 1999, 11, 704–712. [Google Scholar] [CrossRef] [PubMed]
  78. Datta, D.; Contreras, A.G.; Grimm, M.; Waaga-Gasser, A.M.; Briscoe, D.M.; Pal, S. Calcineurin inhibitors modulate CXCR3 splice variant expression and mediate renal cancer progression. J. Am. Soc. Nephrol. 2008, 19, 2437–2446. [Google Scholar] [CrossRef]
  79. Datta, D.; Banerjee, P.; Gasser, M.; Waaga-Gasser, A.M.; Pal, S. CXCR3-B can mediate growth-inhibitory signals in human renal cancer cells by down-regulating the expression of heme oxygenase-1. J. Biol. Chem. 2010, 285, 36842–36848. [Google Scholar] [CrossRef]
  80. Furuya, M.; Yoneyama, T.; Miyagi, E.; Tanaka, R.; Nagahama, K.; Miyagi, Y.; Nagashima, Y.; Hirahara, F.; Inayama, Y.; Aoki, I. Differential expression patterns of CXCR3 variants and corresponding CXC chemokines in clear cell ovarian cancers and endometriosis. Gynecol. Oncol. 2011, 122, 648–655. [Google Scholar] [CrossRef]
  81. Balan, M.; Pal, S. A novel CXCR3-B chemokine receptor-induced growth-inhibitory signal in cancer cells is mediated through the regulation of Bach-1 protein and Nrf2 protein nuclear translocation. J. Biol. Chem. 2014, 289, 3126–3137. [Google Scholar] [CrossRef] [PubMed]
  82. Datta, D.; Flaxenburg, J.A.; Laxmanan, S.; Geehan, C.; Grimm, M.; Waaga-Gasser, A.M.; Briscoe, D.M.; Pal, S. Ras-induced modulation of CXCL10 and its receptor splice variant CXCR3-B in MDA-MB-435 and MCF-7 cells: Relevance for the development of human breast cancer. Cancer Res. 2006, 66, 9509–9518. [Google Scholar] [CrossRef] [PubMed]
  83. Gacci, M.; Serni, S.; Lapini, A.; Vittori, G.; Alessandrini, M.; Nesi, G.; Palli, D.; Carini, M. CXCR3-B expression correlates with tumor necrosis extension in renal cell carcinoma. J. Urol. 2009, 181, 843–848. [Google Scholar] [CrossRef] [PubMed]
  84. Li, Y.; Reader, J.C.; Ma, X.; Kundu, N.; Kochel, T.; Fulton, A.M. Divergent roles of CXCR3 isoforms in promoting cancer stem-like cell survival and metastasis. Breast Cancer Res. Treat. 2015, 149, 403–415. [Google Scholar] [CrossRef] [PubMed]
  85. De Palma, M.; Biziato, D.; Petrova, T.V. Microenvironmental regulation of tumour angiogenesis. Nat. Rev. Cancer 2017, 17, 457–474. [Google Scholar] [CrossRef] [PubMed]
  86. Bergers, G.; Benjamin, L.E. Angiogenesis: Tumorigenesis and the angiogenic switch. Nat. Rev. Cancer 2003, 3, 401–410. [Google Scholar] [CrossRef] [PubMed]
  87. Maione, T.E.; Gray, G.S.; Petro, J.; Hunt, A.J.; Donner, A.L.; Bauer, S.I.; Carson, H.F.; Sharpe, R.J. Inhibition of angiogenesis by recombinant human platelet factor-4 and related peptides. Science 1990, 247, 77–79. [Google Scholar] [CrossRef]
  88. Luster, A.D.; Greenberg, S.M.; Leder, P. The IP-10 chemokine binds to a specific cell surface heparan sulfate site shared with platelet factor 4 and inhibits endothelial cell proliferation. J. Exp. Med. 1995, 182, 219–231. [Google Scholar] [CrossRef]
  89. Strieter, R.M.; Polverini, P.J.; Kunkel, S.L.; Arenberg, D.A.; Burdick, M.D.; Kasper, J.; Dzuiba, J.; Van Damme, J.; Walz, A.; Marriott, D. The functional role of the ELR motif in CXC chemokine-mediated angiogenesis. J. Biol. Chem. 1995, 270, 27348–27357. [Google Scholar] [CrossRef]
  90. Angiolillo, A.L.; Sgadari, C.; Taub, D.D.; Liao, F.; Farber, J.M.; Maheshwari, S.; Kleinman, H.K.; Reaman, G.H.; Tosato, G. Human interferon-inducible protein 10 is a potent inhibitor of angiogenesis in vivo. J. Exp. Med. 1995, 182, 155–162. [Google Scholar] [CrossRef]
  91. Arenberg, D.A.; Kunkel, S.L.; Polverini, P.J.; Morris, S.B.; Burdick, M.D.; Glass, M.C.; Taub, D.T.; Iannettoni, M.D.; Whyte, R.I.; Strieter, R.M. Interferon-gamma-inducible protein 10 (IP-10) is an angiostatic factor that inhibits human non-small cell lung cancer (NSCLC) tumorigenesis and spontaneous metastases. J. Exp. Med. 1996, 184, 981–992. [Google Scholar] [CrossRef] [PubMed]
  92. Rupertus, K.; Sinistra, J.; Scheuer, C.; Nickels, R.M.; Schilling, M.K.; Menger, M.D.; Kollmar, O. Interaction of the chemokines I-TAC (CXCL11) and SDF-1 (CXCL12) in the regulation of tumor angiogenesis of colorectal cancer. Clin. Exp. Metastasis 2014, 31, 447–459. [Google Scholar] [CrossRef] [PubMed]
  93. Campanella, G.S.; Colvin, R.A.; Luster, A.D. CXCL10 can inhibit endothelial cell proliferation independently of CXCR3. PLoS ONE 2010, 5, e12700. [Google Scholar] [CrossRef] [PubMed]
  94. Nagarsheth, N.; Peng, D.; Kryczek, I.; Wu, K.; Li, W.; Zhao, E.; Zhao, L.; Wei, S.; Frankel, T.; Vatan, L. PRC2 epigenetically silences Th1-type chemokines to suppress effector T-cell trafficking in colon cancer. Cancer Res. 2016, 76, 275–282. [Google Scholar] [CrossRef] [PubMed]
  95. Bronger, H.; Singer, J.; Windmüller, C.; Reuning, U.; Zech, D.; Delbridge, C.; Dorn, J.; Kiechle, M.; Schmalfeldt, B.; Schmitt, M.; et al. CXCL9 and CXCL10 predict survival and are regulated by cyclooxygenase inhibition in advanced serous ovarian cancer. Br. J. Cancer 2016, 115, 553–563. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Bronger, H.; Kraeft, S.; Schwarz-Boeger, U.; Cerny, C.; Stöckel, A.; Avril, S.; Kiechle, M.; Schmitt, M. Modulation of CXCR3 ligand secretion by prostaglandin E2 and cyclooxygenase inhibitors in human breast cancer. Breast Cancer Res. 2012, 14, R30. [Google Scholar] [CrossRef]
  97. Havre, P.A.; Abe, M.; Urasaki, Y.; Ohnuma, K.; Morimoto, C.; Dang, N.H. The role of CD26/dipeptidyl peptidase IV in cancer. Front. Biosci. 2008, 13, 1634–1645. [Google Scholar] [CrossRef]
  98. Mortier, A.; Van Damme, J.; Proost, P. Overview of the mechanisms regulating chemokine activity and availability. Immunol. Lett. 2012, 145, 2–9. [Google Scholar] [CrossRef]
  99. Proost, P.; Schutyser, E.; Menten, P.; Struyf, S.; Wuyts, A.; Opdenakker, G.; Detheux, M.; Parmentier, M.; Durinx, C.; Lambeir, A.-M. Amino-terminal truncation of CXCR3 agonists impairs receptor signaling and lymphocyte chemotaxis, while preserving antiangiogenic properties. Blood 2001, 98, 3554–3561. [Google Scholar] [CrossRef] [Green Version]
  100. Da Silva, R.B.; Laird, M.E.; Yatim, N.; Fiette, L.; Ingersoll, M.A.; Albert, M.L. Dipeptidylpeptidase 4 inhibition enhances lymphocyte trafficking, improving both naturally occurring tumor immunity and immunotherapy. Nat. Immunol. 2015, 16, 850–858. [Google Scholar] [CrossRef]
  101. Karin, N.; Wildbaum, G.; Thelen, M. Biased signaling pathways via CXCR3 control the development and function of CD4+ T cell subsets. J. Leukoc. Biol. 2016, 99, 857–862. [Google Scholar] [CrossRef] [PubMed]
  102. Zhang, C.; Li, Z.; Xu, L.; Che, X.; Wen, T.; Fan, Y.; Li, C.; Wang, S.; Cheng, Y.; Wang, X. CXCL9/10/11, a regulator of PD-L1 expression in gastric cancer. BMC Cancer 2018, 18, 462. [Google Scholar] [CrossRef] [PubMed]
  103. Szpakowska, M.; Nevins, A.M.; Meyrath, M.; Rhainds, D.; D’huys, T.; Guité-Vinet, F.; Dupuis, N.; Gauthier, P.A.; Counson, M.; Kleist, A. Different contributions of chemokine N-terminal features attest to a different ligand binding mode and a bias towards activation of ACKR3/CXCR7 compared with CXCR4 and CXCR3. Br. J. Pharmacol. 2018, 175, 1419–1438. [Google Scholar] [CrossRef] [PubMed]
  104. Sierro, F.; Biben, C.; Martínez-Muñoz, L.; Mellado, M.; Ransohoff, R.M.; Li, M.; Woehl, B.; Leung, H.; Groom, J.; Batten, M. Disrupted cardiac development but normal hematopoiesis in mice deficient in the second CXCL12/SDF-1 receptor, CXCR7. Proc. Natl. Acad. Sci. USA 2007, 104, 14759–14764. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Zohar, Y.; Wildbaum, G.; Novak, R.; Salzman, A.L.; Thelen, M.; Alon, R.; Barsheshet, Y.; Karp, C.L.; Karin, N. CXCL11-dependent induction of FOXP3-negative regulatory T cells suppresses autoimmune encephalomyelitis. J. Clin. Investig. 2014, 124, 2009–2022. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Heise, C.E.; Pahuja, A.; Hudson, S.C.; Mistry, M.S.; Putnam, A.L.; Gross, M.M.; Gottlieb, P.A.; Wade, W.S.; Kiankarimi, M.; Schwarz, D. Pharmacological characterization of CXC chemokine receptor 3 ligands and a small molecule antagonist. J. Pharmacol. Exp. Ther. 2005, 313, 1263–1271. [Google Scholar] [CrossRef] [PubMed]
  107. Proost, P.; Verpoest, S.; Van de Borne, K.; Schutyser, E.; Struyf, S.; Put, W.; Ronsse, I.; Grillet, B.; Opdenakker, G.; Van Damme, J. Synergistic induction of CXCL9 and CXCL11 by Toll-like receptor ligands and interferon-γ in fibroblasts correlates with elevated levels of CXCR3 ligands in septic arthritis synovial fluids. J. Leukoc. Biol. 2004, 75, 777–784. [Google Scholar] [CrossRef] [PubMed]
  108. Sauty, A.; Colvin, R.A.; Wagner, L.; Rochat, S.; Spertini, F.; Luster, A.D. CXCR3 internalization following T cell-endothelial cell contact: Preferential role of IFN-inducible T cell α chemoattractant (CXCL11). J. Immunol. 2001, 167, 7084–7093. [Google Scholar] [CrossRef] [PubMed]
  109. Szpakowska, M.; Meyrath, M.; Reynders, N.; Counson, M.; Hanson, J.; Steyaert, J.; Chevigné, A. Mutational analysis of the extracellular disulphide bridges of the atypical chemokine receptor ACKR3/CXCR7 uncovers multiple binding and activation modes for its chemokine and endogenous non-chemokine agonists. Biochem. Pharmacol. 2018, 153, 299–309. [Google Scholar] [CrossRef]
  110. Burns, J.M.; Summers, B.C.; Wang, Y.; Melikian, A.; Berahovich, R.; Miao, Z.; Penfold, M.E.; Sunshine, M.J.; Littman, D.R.; Kuo, C.J. A novel chemokine receptor for SDF-1 and I-TAC involved in cell survival, cell adhesion, and tumor development. J. Exp. Med. 2006, 203, 2201–2213. [Google Scholar] [CrossRef]
  111. Massara, M.; Bonavita, O.; Mantovani, A.; Locati, M.; Bonecchi, R. Atypical chemokine receptors in cancer: Friends or foes? J. Leukoc. Biol. 2016, 99, 927–933. [Google Scholar] [CrossRef] [PubMed]
  112. Freitas, C.; Desnoyer, A.; Meuris, F.; Bachelerie, F.; Balabanian, K.; Machelon, V. The relevance of the chemokine receptor ACKR3/CXCR7 on CXCL12-mediated effects in cancers with a focus on virus-related cancers. Cytokine Growth Factor Rev. 2014, 25, 307–316. [Google Scholar] [CrossRef] [PubMed]
  113. Gao, Y.J.; De Lin Liu, S.L.; Yuan, G.F.; Li, L.; Zhu, H.Y.; Cao, G.Y. Down-regulation of cXcl11 inhibits colorectal cancer cell growth and epithelial-mesenchymal transition. Oncotargets Ther. 2018, 11, 7333–7343. [Google Scholar] [CrossRef] [PubMed]
  114. Zhang, Y.; Zhao, W.; Li, S.; Lv, M.; Yang, X.; Li, M.; Zhang, Z. CXCL11 promotes self-renewal and tumorigenicity of α2δ1+ liver tumor-initiating cells through CXCR3/ERK1/2 signaling. Cancer Lett. 2019, 449, 163–171. [Google Scholar] [CrossRef] [PubMed]
  115. Mulligan, A.M.; Raitman, I.; Feeley, L.; Pinnaduwage, D.; Nguyen, L.T.; O’Malley, F.P.; Ohashi, P.S.; Andrulis, I.L. Tumoral lymphocytic infiltration and expression of the chemokine CXCL10 in breast cancers from the Ontario Familial Breast Cancer Registry. Clin. Cancer Res. 2013, 19, 336–346. [Google Scholar] [CrossRef] [PubMed]
  116. Toiyama, Y.; Fujikawa, H.; Kawamura, M.; Matsushita, K.; Saigusa, S.; Tanaka, K.; Inoue, Y.; Uchida, K.; Mohri, Y.; Kusunoki, M. Evaluation of CXCL10 as a novel serum marker for predicting liver metastasis and prognosis in colorectal cancer. Int. J. Oncol. 2012, 40, 560–566. [Google Scholar] [CrossRef]
  117. Zipin-Roitman, A.; Meshel, T.; Sagi-Assif, O.; Shalmon, B.; Avivi, C.; Pfeffer, R.M.; Witz, I.P.; Ben-Baruch, A. CXCL10 promotes invasion-related properties in human colorectal carcinoma cells. Cancer Res. 2007, 67, 3396–3405. [Google Scholar] [CrossRef]
  118. Zumwalt, T.J.; Arnold, M.; Goel, A.; Boland, C.R. Active secretion of CXCL10 and CCL5 from colorectal cancer microenvironments associates with GranzymeB+ CD8+ T-cell infiltration. Oncotarget 2014, 6, 2981–2991. [Google Scholar] [CrossRef]
  119. Sato, Y.; Motoyama, S.; Nanjo, H.; Wakita, A.; Yoshino, K.; Sasaki, T.; Nagaki, Y.; Liu, J.; Imai, K.; Saito, H. CXCL10 expression status is prognostic in patients with advanced thoracic esophageal squamous cell carcinoma. Ann. Surg. Oncol. 2016, 23, 936–942. [Google Scholar] [CrossRef]
  120. Wightman, S.; Uppal, A.; Pitroda, S.; Ganai, S.; Burnette, B.; Stack, M.; Oshima, G.; Khan, S.; Huang, X.; Posner, M. Oncogenic CXCL10 signalling drives metastasis development and poor clinical outcome. Br. J. Cancer 2015, 113, 327–335. [Google Scholar] [CrossRef]
  121. Wu, Z.; Huang, X.; Han, X.; Li, Z.; Zhu, Q.; Yan, J.; Yu, S.; Jin, Z.; Wang, Z.; Zheng, Q. The chemokine CXCL9 expression is associated with better prognosis for colorectal carcinoma patients. Biomed. Pharmacother. 2016, 78, 8–13. [Google Scholar] [CrossRef] [PubMed]
  122. Cao, Y.; Huang, H.; Wang, Z.; Zhang, G. The Inflammatory CXC Chemokines, GROαhigh, IP-10low, and MIGlow, in Tumor Microenvironment Can Be Used as New Indicators for Non-small Cell Lung Cancer Progression. Immunol. Investig. 2017, 46, 361–374. [Google Scholar] [CrossRef] [PubMed]
  123. Gorbachev, A.V.; Kobayashi, H.; Kudo, D.; Tannenbaum, C.S.; Finke, J.H.; Shu, S.; Farber, J.M.; Fairchild, R.L. CXC chemokine ligand 9/monokine induced by IFN-γ production by tumor cells is critical for T cell-mediated suppression of cutaneous tumors. J. Immunol. 2007, 178, 2278–2286. [Google Scholar] [CrossRef] [PubMed]
  124. Ohtani, H.; Jin, Z.; Takegawa, S.; Nakayama, T.; Yoshie, O. Abundant expression of CXCL9 (MIG) by stromal cells that include dendritic cells and accumulation of CXCR3+ T cells in lymphocyte-rich gastric carcinoma. J. Pathol. A J. Pathol. Soc. Great Br. Irel. 2009, 217, 21–31. [Google Scholar] [CrossRef] [PubMed]
  125. Sgadari, C.; Farber, J.M.; Angiolillo, A.L.; Liao, F.; Teruya-Feldstein, J.; Burd, P.R.; Yao, L.; Gupta, G.; Kanegane, C.; Tosato, G. Mig, the monokine induced by interferon-γ, promotes tumor necrosis in vivo. Blood 1997, 89, 2635–2643. [Google Scholar] [PubMed]
  126. Amatschek, S.; Lucas, R.; Eger, A.; Pflueger, M.; Hundsberger, H.; Knoll, C.; Grosse-Kracht, S.; Schuett, W.; Koszik, F.; Maurer, D. CXCL9 induces chemotaxis, chemorepulsion and endothelial barrier disruption through CXCR3-mediated activation of melanoma cells. Br. J. Cancer 2011, 104, 469–479. [Google Scholar] [CrossRef] [PubMed]
  127. Li, Z.; Liu, J.; Li, L.; Shao, S.; Wu, J.; Bian, L.; He, Y. Epithelial mesenchymal transition induced by the CXCL9/CXCR3 axis through AKT activation promotes invasion and metastasis in tongue squamous cell carcinoma. Oncol. Rep. 2018, 39, 1356–1368. [Google Scholar] [CrossRef]
  128. Mir, M.A.; Maurer, M.J.; Ziesmer, S.C.; Slager, S.L.; Habermann, T.; Macon, W.R.; Link, B.K.; Syrbu, S.; Witzig, T.; Friedberg, J.W. Elevated serum levels of IL-2R, IL-1RA, and CXCL9 are associated with a poor prognosis in follicular lymphoma. Blood 2015, 125, 992–998. [Google Scholar] [CrossRef] [Green Version]
  129. Pucci, F.; Rickelt, S.; Newton, A.P.; Garris, C.; Nunes, E.; Evavold, C.; Pfirschke, C.; Engblom, C.; Mino-Kenudson, M.; Hynes, R.O. PF4 promotes platelet production and lung cancer growth. Cell Rep. 2016, 17, 1764–1772. [Google Scholar] [CrossRef]
  130. Deng, S.; Deng, Q.; Zhang, Y.; Ye, H.; Yu, X.; Zhang, Y.; Han, G.Y.; Luo, P.; Wu, M.; Yu, Y. Non-platelet-derived CXCL4 differentially regulates cytotoxic and regulatory T cells through CXCR3 to suppress the immune response to colon cancer. Cancer Lett. 2019, 443, 1–12. [Google Scholar] [CrossRef]
  131. Zhang, Y.; Gao, J.; Wang, X.; Deng, S.; Ye, H.; Guan, W.; Wu, M.; Zhu, S.; Yu, Y.; Han, W. CXCL4 mediates tumor regrowth after chemotherapy by suppression of antitumor immunity. Cancer Biol. Ther. 2015, 16, 1775–1783. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Struyf, S.; Salogni, L.; Burdick, M.D.; Vandercappellen, J.; Gouwy, M.; Noppen, S.; Proost, P.; Opdenakker, G.; Parmentier, M.; Gerard, C. Angiostatic and chemotactic activities of the CXC chemokine CXCL4L1 (platelet factor-4 variant) are mediated by CXCR3. Blood 2011, 117, 480–488. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Struyf, S.; Burdick, M.D.; Proost, P.; Van Damme, J.; Strieter, R.M. Platelets release CXCL4L1, a nonallelic variant of the chemokine platelet factor-4/CXCL4 and potent inhibitor of angiogenesis. Circ. Res. 2004, 95, 855–857. [Google Scholar] [CrossRef] [PubMed]
  134. Struyf, S.; Burdick, M.D.; Peeters, E.; Van den Broeck, K.; Dillen, C.; Proost, P.; Van Damme, J.; Strieter, R.M. Platelet factor-4 variant chemokine CXCL4L1 inhibits melanoma and lung carcinoma growth and metastasis by preventing angiogenesis. Cancer Res. 2007, 67, 5940–5948. [Google Scholar] [CrossRef] [PubMed]
  135. Quemener, C.; Baud, J.; Boyé, K.; Dubrac, A.; Billottet, C.; Soulet, F.; Darlot, F.; Dumartin, L.; Sire, M.; Grepin, R. Dual roles for CXCL4 chemokines and CXCR3 in angiogenesis and invasion of pancreatic cancer. Cancer Res. 2016, 76, 6507–6519. [Google Scholar] [CrossRef]
  136. Karjalainen, M.K.; Ojaniemi, M.; Haapalainen, A.M.; Mahlman, M.; Salminen, A.; Huusko, J.M.; Määttä, T.A.; Kaukola, T.; Anttonen, J.; Ulvila, J. CXCR3 polymorphism and expression associate with spontaneous preterm birth. J. Immunol. 2015, 195, 2187–2198. [Google Scholar] [CrossRef] [PubMed]
  137. Wu, Q.; Dhir, R.; Wells, A. Altered CXCR3 isoform expression regulates prostate cancer cell migration and invasion. Mol. Cancer 2012, 11, 3. [Google Scholar] [CrossRef] [PubMed]
  138. Baribaud, F.; Edwards, T.G.; Sharron, M.; Brelot, A.; Heveker, N.; Price, K.; Mortari, F.; Alizon, M.; Tsang, M.; Doms, R.W. Antigenically distinct conformations of CXCR4. J. Virol. 2001, 75, 8957–8967. [Google Scholar] [CrossRef]
  139. Blanpain, C.; Vanderwinden, J.-M.; Cihak, J.; Wittamer, V.; Le Poul, E.; Issafras, H.; Stangassinger, M.; Vassart, G.; Marullo, S.; Schlöndorff, D.; et al. Multiple active states and oligomerization of CCR5 revealed by functional properties of monoclonal antibodies. Mol. Biol. Cell 2002, 13, 723–737. [Google Scholar] [CrossRef]
  140. Jopling, L.; Watt, G.; Fisher, S.; Birch, H.; Coggon, S.; Christie, M. Analysis of the pharmacokinetic/pharmacodynamic relationship of a small molecule CXCR3 antagonist, NBI-74330, using a murine CXCR3 internalization assay. Br. J. Pharmacol. 2007, 152, 1260–1271. [Google Scholar] [CrossRef] [Green Version]
  141. Scholten, D.; Canals, M.; Wijtmans, M.; De Munnik, S.; Nguyen, P.; Verzijl, D.; De Esch, I.; Vischer, H.; Smit, M.; Leurs, R. Pharmacological characterization of a small-molecule agonist for the chemokine receptor CXCR3. Br. J. Pharmacol. 2012, 166, 898–911. [Google Scholar] [CrossRef] [PubMed]
  142. Andrews, S.P.; Cox, R.J. Small molecule CXCR3 antagonists. J. Med. Chem. 2015, 59, 2894–2917. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic representation of CXCR3 variants. Due to alternative splicing of the pre-mRNA of the CXCR3 gene, located in chromosome X, three CXCR3 variants can be generated. The CXCR3-A variant is the product of the splicing of the exon 1 and exon 3 within the CXCR3 gene. The assembly of exon 2 and exon 3 results in the CXCR3-B variant, which has an N terminus longer by 52 amino acids (AAs) compared with CXCR3-A. The removal of the intron, exon 2, and a 337-bp region within the third exon during RNA splicing results in the CXCR3-Alt variant that comprises the N terminus and the first four transmembrane domains identical to CXCR3-A, as well as a possible fifth transmembrane region and a C terminus which are different from CXCR3-A and -B. The primers used to detect CXCR3-A, which also recognize CXCR3-B, and CXCR3-B are represented by the black and purple arrows, respectively.
Figure 1. Schematic representation of CXCR3 variants. Due to alternative splicing of the pre-mRNA of the CXCR3 gene, located in chromosome X, three CXCR3 variants can be generated. The CXCR3-A variant is the product of the splicing of the exon 1 and exon 3 within the CXCR3 gene. The assembly of exon 2 and exon 3 results in the CXCR3-B variant, which has an N terminus longer by 52 amino acids (AAs) compared with CXCR3-A. The removal of the intron, exon 2, and a 337-bp region within the third exon during RNA splicing results in the CXCR3-Alt variant that comprises the N terminus and the first four transmembrane domains identical to CXCR3-A, as well as a possible fifth transmembrane region and a C terminus which are different from CXCR3-A and -B. The primers used to detect CXCR3-A, which also recognize CXCR3-B, and CXCR3-B are represented by the black and purple arrows, respectively.
Cells 08 00613 g001
Figure 2. Illustration of the roles of CXCR3 variants and their ligands in the tumor microenvironment. The increased secretion of CXCR3 ligands in the TME, originating from the tumor cells, results in an increased chemokine gradient (1) and the recruitment of CXCR3-A+ leukocytes towards the tumor (paracrine axis) (2). The Th1-related immune cells drive anti-tumoral responses, while the recruited CXCR3-A+ Tregs have a pro-tumoral effect by suppressing other immune cells. The secreted chemokines also bind to CXCR3-A or CXCR3-B, present on secreting (autocrine axis) and neighboring tumor cells, and depending on the cellular context, may induce a proliferative or inhibitory effect (3). Moreover, the chemokines in close proximity to the endothelial cells attenuate the angiogenesis by activating CXCR3-B (4). In addition, the increased expression of CXCR3-A, together with decreased CXCR3-B expression, on the tumor cells enable them to detach from the tumor mass and disseminate to a distant tissue (5).
Figure 2. Illustration of the roles of CXCR3 variants and their ligands in the tumor microenvironment. The increased secretion of CXCR3 ligands in the TME, originating from the tumor cells, results in an increased chemokine gradient (1) and the recruitment of CXCR3-A+ leukocytes towards the tumor (paracrine axis) (2). The Th1-related immune cells drive anti-tumoral responses, while the recruited CXCR3-A+ Tregs have a pro-tumoral effect by suppressing other immune cells. The secreted chemokines also bind to CXCR3-A or CXCR3-B, present on secreting (autocrine axis) and neighboring tumor cells, and depending on the cellular context, may induce a proliferative or inhibitory effect (3). Moreover, the chemokines in close proximity to the endothelial cells attenuate the angiogenesis by activating CXCR3-B (4). In addition, the increased expression of CXCR3-A, together with decreased CXCR3-B expression, on the tumor cells enable them to detach from the tumor mass and disseminate to a distant tissue (5).
Cells 08 00613 g002

Share and Cite

MDPI and ACS Style

Reynders, N.; Abboud, D.; Baragli, A.; Noman, M.Z.; Rogister, B.; Niclou, S.P.; Heveker, N.; Janji, B.; Hanson, J.; Szpakowska, M.; et al. The Distinct Roles of CXCR3 Variants and Their Ligands in the Tumor Microenvironment. Cells 2019, 8, 613. https://doi.org/10.3390/cells8060613

AMA Style

Reynders N, Abboud D, Baragli A, Noman MZ, Rogister B, Niclou SP, Heveker N, Janji B, Hanson J, Szpakowska M, et al. The Distinct Roles of CXCR3 Variants and Their Ligands in the Tumor Microenvironment. Cells. 2019; 8(6):613. https://doi.org/10.3390/cells8060613

Chicago/Turabian Style

Reynders, Nathan, Dayana Abboud, Alessandra Baragli, Muhammad Zaeem Noman, Bernard Rogister, Simone P. Niclou, Nikolaus Heveker, Bassam Janji, Julien Hanson, Martyna Szpakowska, and et al. 2019. "The Distinct Roles of CXCR3 Variants and Their Ligands in the Tumor Microenvironment" Cells 8, no. 6: 613. https://doi.org/10.3390/cells8060613

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop