Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 + 2033 word(s) 2033 2021-04-08 05:40:39 |
2 format correct Meta information modification 2033 2021-04-14 09:12:35 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Bajdak-Rusinek, K. CXCL16. Encyclopedia. Available online: https://encyclopedia.pub/entry/8683 (accessed on 20 April 2024).
Bajdak-Rusinek K. CXCL16. Encyclopedia. Available at: https://encyclopedia.pub/entry/8683. Accessed April 20, 2024.
Bajdak-Rusinek, Karolina. "CXCL16" Encyclopedia, https://encyclopedia.pub/entry/8683 (accessed April 20, 2024).
Bajdak-Rusinek, K. (2021, April 14). CXCL16. In Encyclopedia. https://encyclopedia.pub/entry/8683
Bajdak-Rusinek, Karolina. "CXCL16." Encyclopedia. Web. 14 April, 2021.
CXCL16
Edit

CXCL16 is a chemotactic cytokine belonging to the α-chemokine subfamily. It plays a significant role in the progression of cancer, as well as the course of atherosclerosis, renal fibrosis, and non-alcoholic fatty liver disease (NAFLD). 

CXCL16 CXCR6 cancer tumor NKT cells SR-PSOX ADAM10 tumor microenvironment inflammation

1. Introduction

A tumor mass comprises not just cancer cells but also tumor growth supporting cells and tumor-suppressing immune cells [1]. All these cells forming the tumor microenvironment secrete various factors into the intercellular space. To date, many of these factors have been characterized. The most significant of these are chemokines, a group of approximately 50 chemotactic cytokines [2]. Earlier studies have characterized chemokines as a significant factor in immune system cell function [3][4]. However, over time, more studies have emerged showing the significance of a variety of chemokines in cancer development and growth [5]. One such chemokine is CXC motif chemokine ligand 16 (CXCL16), belonging to the sub-family of α-chemokines. To date, there has not been a review summarizing current knowledge on CXCL16 and its receptor CXC motif chemokine receptor 6 (CXCR6) in a tumor. 

2. CXC Motif Chemokine Ligand 16 (CXCL16): Background Information

CXCL16 is a chemokine distinct from other CXC chemokines. The CXCL16 gene is located on chromosome 17p13 [6], separate from other chemokine genes, and it has poor homology with the other chemokines [7]. There are two CXCL16 transcripts, 1.8 kb and 2.5 kb in length, formed by alternative splicing [8]. Both transcripts differ in the 3′-noncoding regions and the location of their expression. The 1.8 kb transcript is found mostly in the spleen, thymus, and testis. In contrast, the 2.5 kb transcript is found in the heart, kidney, liver, lung, peripheral blood leukocytes, pancreas, and prostate [8]. Alternative RNA splicing produces two other transcripts having 70 (CXCL16v1) and 121 (CXCL16v2) extra nucleotides, as demonstrated in dendritic cells [9]. In these additional sequences, a STOP codon is present. This results in a shorter protein having only the chemokine domain. The other domains typically found in CXCL16 are not present.

After translation, a hydrophobic signal peptide is cleaved at the N-terminus of the CXCL16 polypeptide [7]. The human CXCL16 protein is 254 aa long [6][7], while the murine CXCL16 is 246 aa long and is 44% similar to the human CXCL16 [8]. The newly synthesized CXCL16 for both species is the so-called membrane-bound form of CXCL16 (mCXCL16) [6]. The structure of this protein is very similar to transmembrane CX3CL1 [10] because it consists of a small (24–27 aa long) cytoplasmic tail with a YXPV motif [6]. This motif can be phosphorylated on tyrosine to provide an SH2-binding site. mCXCL16 also consists of a CXC chemokine domain and a transmembrane domain (Figure 1) [6]. Both of these domains are separated from each other by a spacer region, approximately 100 aa long, rich in serine, threonine, and proline. It is a site of heavy O-glycosylation which results in the formation of a mucin-like stalk [6][7], essential for chemokine domain presentation and, therefore, significant for mCXCL16 properties [11][12]. mCXCL16 has a molecular weight of 60 kDa.

Figure 1. mCXCL16 structure. CXC motif chemokine ligand 16 (CXCL16) is a membrane protein consisting of a chemokine domain, highly glycosylated mucin-like stalk, transmembrane domain and cytoplasmic tail.

After CXCL16 cleavage, the 35 kDa chemokine domain is released and becomes a soluble form of CXCL16 (sCXCL16). This process is regulated by disintegrin and metalloproteinase 10 (ADAM10) [13][14][15]. However, disintegrin and metalloproteinase 17 (ADAM17) may also be responsible for sCXCL16 shedding in the absence of ADAM10 [16] or by using phorbol 12-myristate 13-acetate (PMA) as an inflammatory inducer [13]. Studies on mesangial cells have shown that pro-inflammatory cytokines increase sCXCL16 shedding via ADAM10 and ADAM17 [17]. Although sCXCL16 is released from mCXCL16, the remainder of the mCXCL16 (mucin-like stalk, transmembrane domain, and cytoplasmic tail) is still present in the cell membrane. These domains are firstly proteolytically cleaved by γ-secretases and then proteolytically degraded [15].

Importantly, many papers do not distinguish between two forms of CXCL16. Researchers just alter the expression of the CXCL16 gene or change the expression of two CXCL16 forms at the same time; then they test selected parameters of the experiment. In those cases, we report the findings as referring to the unspecified “CXCL16” which may denote both forms at the same time. When writing about mCXCL16 and sCXCL16, we mean the specific form of CXCL16.

CXCL16 is classified as an α sub-family chemokine because it has a CXC motif at the N-terminus [7]. Nevertheless, this chemokine lacks the ELR motif, which is common in other pro-angiogenic CXC chemokines [6][7]. CXCL16 is expressed in lymphoid organs such as the thymus, spleen, lymph nodes, and Peyer’s patches, but not in bone marrow. It is also expressed in the liver, lung, small intestine, and kidney [6][7]. Moreover, CXCL16 is highly expressed in the epidermis, where it is produced by keratinocytes [18]. mCXCL16 is expressed in macrophages [8][19] and is also present in splenic and lymph node dendritic cells (DC), blood myeloid DC, and monocyte-derived DC. Interestingly, its expression increases after DC maturation and after exposure to pro-inflammatory factors [6][19]. It has been also shown that CXCL16 is expressed on CD19+ B cells [7].

Both forms of CXCL16 have different functions. sCXCL16 is a chemokine that is responsible for the chemotaxis of cells bearing the CXCR6 receptor [6][7] while mCXCL16 is a transmembrane protein. mCXCL16 may have adhesion protein properties that bind to CXCR6 (Table 1) [12][20]. However, signal transduction from CXCR6 via mCXCL16→CXCR6 is not necessary for cell adhesion. This property of mCXCL16 is important in immune cell accumulation at an inflammation site due to increased mCXCL16 expression on vascular walls by pro-inflammatory cytokines [20]. mCXCL16 also mediates the adhesion of Gram-negative and Gram-positive bacteria [11]. This leads to bacterial phagocytosis by cells expressing mCXCL16, such as macrophages and DC [8][11]. mCXCL16 may also act as a receptor that causes signal transduction into the cell. This signal can be induced by sCXCL16 [21] as well as by CXCR6 [22]. In the first case, this effect is called inverse signaling; in the second case–reverse signaling. Additionally, signal transduction from mCXCL16 is insensitive to pertussis toxin [23], and so it has a different mechanism than signaling mediated by the G protein-coupled receptor (CXCR6). The sCXCL16-induced activation of mCXCL16 causes the activation of extracellular signal-regulated kinase (ERK) mitogen-activated protein kinase (MAPK) [21][23] and Akt/protein kinase B (PKB) pathways [21]. This leads to increased proliferation and formation of apoptosis-resistant tumor cells [21][23]. In glioblastoma or melanoma cells, CXCR6 activates the mCXCL16→ERK MAPK pathway, which causes migration but not proliferation of tumor cells [22].

Table 1. Activities of the two different forms of CXCL16.

Ligand Receptor Activated Signaling Pathways Physiological Significance References
sCXCL16 CXCR6 G protein-coupled receptor, PKB, ERK MAPK, calcium mobilization Proliferation, migration, fibrosis, VEGF and CXCL8/IL-8 expression [7][12][24][25][26][27][28][29][30][31][32][33]
sCXCL16 mCXCL16 insensitive to pertussis toxin, ERK MAPK, PKB Proliferation, apoptosis resistance [21][23]
CXCR6 mCXCL16 insensitive to pertussis toxin, ERK MAPK Migration but not proliferation [22][23]
mCXCL16 CXCR6   Cell adhesion [12][20]

CXCL16 expression is increased by pro-inflammatory cytokines. This is important for the accumulation of immune cells at the inflammatory reaction sites, but the influence of the various cytokines is cell-specific. In ectopic endometrial stromal cells, tumor necrosis factor α (TNF-α) increases CXCL16 expression [24]. In vascular smooth muscle cells, interferon-γ (IFN-γ) increases the expression of this chemokine, but TNF-α does not have such an effect [34]. In human umbilical vein endothelial cells (HUVEC), TNF-α, IFN-γ, interleukin (IL)-1β, and IL-6 increase the expression of CXCL16 [20]. Research on keratinocytes has shown that TNF-α, IFN-γ and peptidoglycan increase the expression of CXCL16 in these cells [18]. Also, the expression of CXCL16 is increased by ionizing radiation, which is important during radiotherapy [35][36]. Another factor that increases the expression of CXCL16 is hypoxia and hypoxia-inducible factor-1 (HIF-1) [37][38].

3. The Importance of CXCL16 in Non-Neoplastic Diseases

CXCL16 is involved in the pathogenesis of atherosclerosis. Inflammatory cytokines such as TNF-α, IFN-γ, IL-1β and IL-6 increase the expression of CXCL16 on endothelial cells, as shown in a study on HUVEC [20]. This leads to the adhesion of THP-1 monocytes, which are human monocytic leukaemia cells. A study on peripheral blood mononuclear cells (PBMC) shows that monocytes do not express CXCR6 [39], which means that in the pathogenesis of atherosclerosis, monocytes may not be directly recruited through the CXCL16→CXCR6 axis. In principle, this effect may be indirect by recruiting CXCR6+ T cells [40]. CXCL16 plays an important role in the later stages of atherosclerosis. A high expression of CXCL16 is present in macrophages in the intima of atherosclerotic lesions but not in normal aortas [41]. mCXCL16 has scavenger receptor activity and can be a receptor for oxidized lipoproteins. For this reason, it was first named the scavenger receptor for phosphatidylserine and oxidized lipoprotein (SR-PSOX) [41]. As such, it can participate in the uptake of oxidized low-density lipoprotein (oxLDL) by macrophages [41] and vascular smooth muscle cells [34]. CXCL16 can also cause proliferation of aortic smooth muscle cells and increase the expression of TNF-α in these cells [26], which is associated with atherosclerotic vascular disease. A study in low-density lipoprotein (LDL) receptor-deficient mice showed that CXCL16 activity is athero-protective [42]. For this reason, more research is needed on the role of CXCL16 in the pathogenesis of atherosclerosis.

Another situation where CXCL16 plays a significant role is in liver diseases, in particular, nonalcoholic fatty liver disease (NAFLD). During inflammatory reactions, the expressions of CXCL16, CXCR6, and ADAM10 increase in the liver [43]. In NAFLD, hepatocytes produce CXCL16 [44] which activates hepatic stellate cells that produce collagen and transform into myofibroblasts [44]. Interestingly, activation of CXCR6 on hepatocytes reduces inflammation, fibrosis, and the death of these cells, as demonstrated in mice with an activated NF-κB pathway [45]. Despite this, CXCL16 increases lipid accumulation, extracellular matrix (ECM) excretion and reactive oxygen species (ROS) production in hepatocytes [43]. CXCL16 is also responsible for the accumulation of NKT cells in the liver [45][46][47][48]. These cells participate in inflammatory reactions, leading to liver fibrosis progression. Also, during septic shock, the expression of CXCL16 is increased in the liver [49]. This leads to the recruitment of activated T cells and consequently to endotoxin-induced lethal liver injury. Importantly, the NK and NKT cells are not responsible for septic shock symptoms in the liver.

Transplantology is another issue where CXCL16 plays an important role. During allogeneic heart transplantation in mice, the CXCL16→CXCR6 axis is involved in the recruitment of NKT cells into the transplanted organ [50], which is related to allograft tolerance. Nevertheless, CXCL16 can also cause graft-vs-host disease (GvHD) due to its participation in the recruitment of activated CD8+ T cells into the liver, leading to GvHD-induced hepatitis [51].

The CXCL16→CXCR6 axis is also involved in fibrosis in various organs. For example, during inflammatory reactions, the expression of CXCL16 increases on tubular epithelial cells [52]. This leads to the recruitment of bone marrow-derived fibroblast precursors, cells expressing CXCR6 [52][53], which participate in the pathogenesis of renal fibrosis. CXCL16 is also engaged in pulmonary fibrosis [30], which was demonstrated in human pulmonary fibroblasts (MRC-5 cell line), where this chemokine caused an increase in proliferation and collagen production. This was associated with the activation of the forkhead box O3a (FOXO3a) via the CXCR6→Akt/PKB pathway [30].

CXCL16 is also essential in the homeostasis of intestinal defense. It is responsible for the distribution of intestinal group 3 innate lymphoid cell (ILC3) subsets and IL-22 production [54]. This leads to an increase in antimicrobial peptide expression, which is involved in the defense against bacteria. CXCL16 can also be taken as a marker of inflammatory bowel disease (IBD) [55]. It seems that CXCR6+CD4+ T cells do not show colitogenic properties, in contrast to CXCR6CD4+ T cells [56]. However, in inflamed colonic tissues, CXCL16 expression occurs on macrophages where it participates in the Th1 immune response [55].

The CXCL16→CXCR6 axis also participates in the development of endometriosis [24]. In ectopic endometrial stromal cells, TNF-α increases the expression of CXCL16. This causes migration and invasion of these cells, which is associated with the development of endometriosis.

CXCR6 is also an entry cofactor for human immunodeficiency virus (HIV)-1 [6][57][58][59] and HIV-2 [60] and so may be significant in the course of HIV infection. This is also supported by the association of polymorphisms in the CXCR6 gene with the control of viraemic disease [61][62]—its expression is downregulated in this disease [63]. In particular, the presence of specific polymorphisms in CXCR6 leads to rapid progression of acquired immunodeficiency syndrome (AIDS) [64] and influences the effectiveness of highly active antiretroviral therapy (HAART) [65]. Also, CXCL16 levels increase in HIV infection as the virus increases the release of this chemokine by macrophages [66]. This is associated with disease progression as CXCL16 increases HIV replication.

References

  1. Hinshaw, D.C.; Shevde, L.A. The Tumor Microenvironment Innately Modulates Cancer Progression. Cancer Res. 2019, 79, 4557–4566.
  2. Hughes, C.E.; Nibbs, R.J.B. A guide to chemokines and their receptors. FEBS J. 2018, 285, 2944–2971.
  3. Schall, T.J.; Jongstra, J.; Dyer, B.J.; Jorgensen, J.; Clayberger, C.; Davis, M.M.; Krensky, A.M. A human T cell-specific molecule is a member of a new gene family. J. Immunol. 1988, 141, 1018–1025.
  4. Kodelja, V.; Müller, C.; Politz, O.; Hakij, N.; Orfanos, C.E.; Goerdt, S. Alternative macrophage activation-associated CC-chemokine-1, a novel structural homologue of macrophage inflammatory protein-1 alpha with a Th2-associated expression pattern. J. Immunol. 1998, 160, 1411–1418.
  5. Do, H.T.T.; Lee, C.H.; Cho, J. Chemokines and their Receptors: Multifaceted Roles in Cancer Progression and Potential Value as Cancer Prognostic Markers. Cancers 2020, 12, 287.
  6. Matloubian, M.; David, A.; Engel, S.; Ryan, J.E.; Cyster, J.G. A transmembrane CXC chemokine is a ligand for HIV-coreceptor Bonzo. Nat. Immunol. 2000, 1, 298–304.
  7. Wilbanks, A.; Zondlo, S.C.; Murphy, K.; Mak, S.; Soler, D.; Langdon, P.; Andrew, D.P.; Wu, L.; Briskin, M. Expression cloning of the STRL33/BONZO/TYMSTRligand reveals elements of CC, CXC, and CX3C chemokines. J. Immunol. 2001, 166, 5145–5154.
  8. Shimaoka, T.; Kume, N.; Minami, M.; Hayashida, K.; Kataoka, H.; Kita, T.; Yonehara, S. Molecular cloning of a novel scavenger receptor for oxidized low density lipoprotein, SR-PSOX, on macrophages. J. Biol. Chem. 2000, 275, 40663–40666.
  9. Van der Voort, R.; Verweij, V.; de Witte, T.M.; Lasonder, E.; Adema, G.J.; Dolstra, H. An alternatively spliced CXCL16 isoform expressed by dendritic cells is a secreted chemoattractant for CXCR6+ cells. J. Leukoc. Biol. 2010, 87, 1029–1039.
  10. Fong, A.M.; Erickson, H.P.; Zachariah, J.P.; Poon, S.; Schamberg, N.J.; Imai, T.; Patel, D.D. Ultrastructure and function of the fractalkine mucin domain in CX(3)C chemokine domain presentation. J. Biol. Chem. 2000, 275, 3781–3786.
  11. Shimaoka, T.; Nakayama, T.; Kume, N.; Takahashi, S.; Yamaguchi, J.; Minami, M.; Hayashida, K.; Kita, T.; Ohsumi, J.; Yoshie, O.; et al. Cutting edge: SR-PSOX/CXC chemokine ligand 16 mediates bacterial phagocytosis by APCs through its chemokine domain. J. Immunol. 2003, 171, 1647–1651.
  12. Shimaoka, T.; Nakayama, T.; Fukumoto, N.; Kume, N.; Takahashi, S.; Yamaguchi, J.; Minami, M.; Hayashida, K.; Kita, T.; Ohsumi, J.; et al. Cell surface-anchored SR-PSOX/CXC chemokine ligand 16 mediates firm adhesion of CXC chemokine receptor 6-expressing cells. J. Leukoc. Biol. 2004, 75, 267–274.
  13. Abel, S.; Hundhausen, C.; Mentlein, R.; Schulte, A.; Berkhout, T.A.; Broadway, N.; Hartmann, D.; Sedlacek, R.; Dietrich, S.; Muetze, B.; et al. The transmembrane CXC-chemokine ligand 16 is induced by IFN-gamma and TNF-alpha and shed by the activity of the disintegrin-like metalloproteinase ADAM10. J. Immunol. 2004, 172, 6362–6372.
  14. Gutwein, P.; Schramme, A.; Sinke, N.; Abdel-Bakky, M.S.; Voss, B.; Obermüller, N.; Doberstein, K.; Koziolek, M.; Fritzsche, F.; Johannsen, M.; et al. Tumoural CXCL16 expression is a novel prognostic marker of longer survival times in renal cell cancer patients. Eur. J. Cancer 2009, 45, 478–489.
  15. Schulte, A.; Schulz, B.; Andrzejewski, M.G.; Hundhausen, C.; Mletzko, S.; Achilles, J.; Reiss, K.; Paliga, K.; Weber, C.; John, S.R.; et al. Sequential processing of the transmembrane chemokines CX3CL1 and CXCL16 by alpha- and gamma-secretases. Biochem. Biophys. Res. Commun. 2007, 358, 233–240.
  16. Gutwein, P.; Abdel-Bakky, M.S.; Schramme, A.; Doberstein, K.; Kämpfer-Kolb, N.; Amann, K.; Hauser, I.A.; Obermüller, N.; Bartel, C.; Abdel-Aziz, A.A.; et al. CXCL16 is expressed in podocytes and acts as a scavenger receptor for oxidized low-density lipoprotein. Am. J. Pathol. 2009, 174, 2061–2072.
  17. Schramme, A.; Abdel-Bakky, M.S.; Kämpfer-Kolb, N.; Pfeilschifter, J.; Gutwein, P. The role of CXCL16 and its processing metalloproteinases ADAM10 and ADAM17 in the proliferation and migration of human mesangial cells. Biochem. Biophys. Res. Commun. 2008, 370, 311–316.
  18. Tohyama, M.; Sayama, K.; Komatsuzawa, H.; Hanakawa, Y.; Shirakata, Y.; Dai, X.; Yang, L.; Tokumaru, S.; Nagai, H.; Hirakawa, S.; et al. CXCL16 is a novel mediator of the innate immunity of epidermal keratinocytes. Int. Immunol. 2007, 19, 1095–1102.
  19. Tabata, S.; Kadowaki, N.; Kitawaki, T.; Shimaoka, T.; Yonehara, S.; Yoshie, O.; Uchiyama, T. Distribution and kinetics of SR-PSOX/CXCL16 and CXCR6 expression on human dendritic cell subsets and CD4+ T cells. J. Leukoc. Biol. 2005, 77, 777–786.
  20. Hofnagel, O.; Engel, T.; Severs, N.J.; Robenek, H.; Buers, I. SR-PSOX at sites predisposed to atherosclerotic lesion formation mediates monocyte-endothelial cell adhesion. Atherosclerosis 2011, 217, 371–378.
  21. Hattermann, K.; Bartsch, K.; Gebhardt, H.H.; Mehdorn, H.M.; Synowitz, M.; Schmitt, A.D.; Mentlein, R.; Held-Feindt, J. “Inverse signaling” of the transmembrane chemokine CXCL16 contributes to proliferative and anti-apoptotic effects in cultured human meningioma cells. Cell Commun. Signal. 2016, 14, 26.
  22. Adamski, V.; Mentlein, R.; Lucius, R.; Synowitz, M.; Held-Feindt, J.; Hattermann, K. The Chemokine Receptor CXCR6 Evokes Reverse Signaling via the Transmembrane Chemokine CXCL16. Int. J. Mol. Sci. 2017, 18, 1468.
  23. Hattermann, K.; Gebhardt, H.; Krossa, S.; Ludwig, A.; Lucius, R.; Held-Feindt, J.; Mentlein, R. Transmembrane chemokines act as receptors in a novel mechanism termed inverse signaling. Elife 2016, 5, e10820.
  24. Peng, Y.; Ma, J.; Lin, J. Activation of the CXCL16/CXCR6 Axis by TNF-α Contributes to Ectopic Endometrial Stromal Cells Migration and Invasion. Reprod. Sci. 2019, 26, 420–427.
  25. Koenen, A.; Babendreyer, A.; Schumacher, J.; Pasqualon, T.; Schwarz, N.; Seifert, A.; Deupi, X.; Ludwig, A.; Dreymueller, D. The DRF motif of CXCR6 as chemokine receptor adaptation to adhesion. PLoS ONE 2017, 12, e0173486.
  26. Chandrasekar, B.; Bysani, S.; Mummidi, S. CXCL16 signals via Gi, phosphatidylinositol 3-kinase, Akt, I kappa B kinase, and nuclear factor-kappa B and induces cell-cell adhesion and aortic smooth muscle cell proliferation. J. Biol. Chem. 2004, 279, 3188–3196.
  27. Hong, L.; Wang, S.; Li, W.; Wu, D.; Chen, W. Tumor-associated macrophages promote the metastasis of ovarian carcinoma cells by enhancing CXCL16/CXCR6 expression. Pathol. Res. Pract. 2018, 214, 1345–1351.
  28. Liang, K.; Liu, Y.; Eer, D.; Liu, J.; Yang, F.; Hu, K. High CXC Chemokine Ligand 16 (CXCL16) Expression Promotes Proliferation and Metastasis of Lung Cancer via Regulating the NF-κB Pathway. Med. Sci. Monit. 2018, 24, 405–411.
  29. Wang, J.; Lu, Y.; Wang, J.; Koch, A.E.; Zhang, J.; Taichman, R.S. CXCR6 induces prostate cancer progression by the AKT/mammalian target of rapamycin signaling pathway. Cancer Res. 2008, 68, 10367–10376.
  30. Ma, Z.; Yu, R.; Zhu, Q.; Sun, L.; Jian, L.; Wang, X.; Zhao, J.; Li, C.; Liu, X. CXCL16/CXCR6 axis promotes bleomycin-induced fibrotic process in MRC-5 cells via the PI3K/AKT/FOXO3a pathway. Int. Immunopharmacol. 2020, 81, 106035.
  31. Xiao, G.; Wang, X.; Wang, J.; Zu, L.; Cheng, G.; Hao, M.; Sun, X.; Xue, Y.; Lu, J.; Wang, J. CXCL16/CXCR6 chemokine signaling mediates breast cancer progression by pERK1/2-dependent mechanisms. Oncotarget 2015, 6, 14165–14178.
  32. Gao, Q.; Zhao, Y.J.; Wang, X.Y.; Qiu, S.J.; Shi, Y.H.; Sun, J.; Yi, Y.; Shi, J.Y.; Shi, G.M.; Ding, Z.B.; et al. CXCR6 upregulation contributes to a proinflammatory tumor microenvironment that drives metastasis and poor patient outcomes in hepatocellular carcinoma. Cancer Res. 2012, 72, 3546–3556.
  33. Xu, J.M.; Weng, M.Z.; Song, F.B.; Chen, J.Y.; Zhang, J.Y.; Wu, J.Y.; Qin, J.; Jin, T.; Wang, X.L. Blockade of the CXCR6 signaling inhibits growth and invasion of hepatocellular carcinoma cells through inhibition of the VEGF expression. Int. J. Immunopathol. Pharmacol. 2014, 27, 553–561.
  34. Wågsäter, D.; Olofsson, P.S.; Norgren, L.; Stenberg, B.; Sirsjö, A. The chemokine and scavenger receptor CXCL16/SR-PSOX is expressed in human vascular smooth muscle cells and is induced by interferon gamma. Biochem. Biophys. Res. Commun. 2004, 325, 1187–1193.
  35. Matsumura, S.; Wang, B.; Kawashima, N.; Braunstein, S.; Badura, M.; Cameron, T.O.; Babb, J.S.; Schneider, R.J.; Formenti, S.C.; Dustin, M.L.; et al. Radiation-induced CXCL16 release by breast cancer cells attracts effector T cells. J. Immunol. 2008, 181, 3099–3107.
  36. Xiao, Z.; Yang, S.; Su, Y.; Wang, W.; Zhang, H.; Zhang, M.; Zhang, K.; Tian, Y.; Cao, Y.; Yin, L.; et al. Alteration of the inflammatory molecule network after irradiation of soft tissue. Adv. Exp. Med. Biol. 2013, 765, 335–341.
  37. Ye, L.Y.; Chen, W.; Bai, X.L.; Xu, X.Y.; Zhang, Q.; Xia, X.F.; Sun, X.; Li, G.G.; Hu, Q.D.; Fu, Q.H.; et al. Hypoxia-Induced Epithelial-to-Mesenchymal Transition in Hepatocellular Carcinoma Induces an Immunosuppressive Tumor Microenvironment to Promote Metastasis. Cancer Res. 2016, 76, 818–830.
  38. Yu, X.; Zhao, R.; Lin, S.; Bai, X.; Zhang, L.; Yuan, S.; Sun, L. CXCL16 induces angiogenesis in autocrine signaling pathway involving hypoxia-inducible factor 1α in human umbilical vein endothelial cells. Oncol. Rep. 2016, 35, 1557–1565.
  39. Unutmaz, D.; Xiang, W.; Sunshine, M.J.; Campbell, J.; Butcher, E.; Littman, D.R. The primate lentiviral receptor Bonzo/STRL33 is coordinately regulated with CCR5 and its expression pattern is conserved between human and mouse. J. Immunol. 2000, 165, 3284–3292.
  40. Galkina, E.; Harry, B.L.; Ludwig, A.; Liehn, E.A.; Sanders, J.M.; Bruce, A.; Weber, C.; Ley, K. CXCR6 promotes atherosclerosis by supporting T-cell homing, interferon-gamma production, and macrophage accumulation in the aortic wall. Circulation 2007, 116, 1801–1811.
  41. Minami, M.; Kume, N.; Shimaoka, T.; Kataoka, H.; Hayashida, K.; Yonehara, S.; Kita, T. Expression of scavenger receptor for phosphatidylserine and oxidized lipoprotein (SR-PSOX) in human atheroma. Ann. N. Y. Acad. Sci. 2001, 947, 373–376.
  42. Aslanian, A.M.; Charo, I.F. Targeted disruption of the scavenger receptor and chemokine CXCL16 accelerates atherosclerosis. Circulation 2006, 114, 583–590.
  43. Ma, K.L.; Wu, Y.; Zhang, Y.; Wang, G.H.; Hu, Z.B.; Ruan, X.Z. Activation of the CXCL16/CXCR6 pathway promotes lipid deposition in fatty livers of apolipoprotein E knockout mice and HepG2 cells. Am. J. Transl. Res. 2018, 10, 1802–1816.
  44. Jiang, L.; Yang, M.; Li, X.; Wang, Y.; Zhou, G.; Zhao, J. CXC Motif Ligand 16 Promotes Nonalcoholic Fatty Liver Disease Progression via Hepatocyte-Stellate Cell Crosstalk. J. Clin. Endocrinol. Metab. 2018, 103, 3974–3985.
  45. Liepelt, A.; Wehr, A.; Kohlhepp, M.; Mossanen, J.C.; Kreggenwinkel, K.; Denecke, B.; Costa, I.G.; Luedde, T.; Trautwein, C.; Tacke, F. CXCR6 protects from inflammation and fibrosis in NEMOLPC-KO mice. Biochim. Biophys. Acta Mol. Basis Dis. 2019, 1865, 391–402.
  46. Xu, H.B.; Gong, Y.P.; Cheng, J.; Chu, Y.W.; Xiong, S.D. CXCL16 participates in pathogenesis of immunological liver injury by regulating T lymphocyte infiltration in liver tissue. World J. Gastroenterol. 2005, 11, 4979–4985.
  47. Wehr, A.; Baeck, C.; Heymann, F.; Niemietz, P.M.; Hammerich, L.; Martin, C.; Zimmermann, H.W.; Pack, O.; Gassler, N.; Hittatiya, K.; et al. Chemokine receptor CXCR6-dependent hepatic NK T Cell accumulation promotes inflammation and liver fibrosis. J. Immunol. 2013, 190, 5226–5236.
  48. Wehr, A.; Baeck, C.; Ulmer, F.; Gassler, N.; Hittatiya, K.; Luedde, T.; Neumann, U.P.; Trautwein, C.; Tacke, F. Pharmacological inhibition of the chemokine CXCL16 diminishes liver macrophage infiltration and steatohepatitis in chronic hepatic injury. PLoS ONE 2014, 9, e112327.
  49. Xu, H.; Xu, W.; Chu, Y.; Gong, Y.; Jiang, Z.; Xiong, S. Involvement of up-regulated CXC chemokine ligand 16/scavenger receptor that binds phosphatidylserine and oxidized lipoprotein in endotoxin-induced lethal liver injury via regulation of T-cell recruitment and adhesion. Infect. Immun. 2005, 73, 4007–4016.
  50. Jiang, X.; Shimaoka, T.; Kojo, S.; Harada, M.; Watarai, H.; Wakao, H.; Ohkohchi, N.; Yonehara, S.; Taniguchi, M.; Seino, K. Cutting edge: Critical role of CXCL16/CXCR6 in NKT cell trafficking in allograft tolerance. J. Immunol. 2005, 175, 2051–2055.
  51. Sato, T.; Thorlacius, H.; Johnston, B.; Staton, T.L.; Xiang, W.; Littman, D.R.; Butcher, E.C. Role for CXCR6 in recruitment of activated CD8+ lymphocytes to inflamed liver. J. Immunol. 2005, 174, 277–283.
  52. Chen, G.; Lin, S.C.; Chen, J.; He, L.; Dong, F.; Xu, J.; Han, S.; Du, J.; Entman, M.L.; Wang, Y. CXCL16 recruits bone marrow-derived fibroblast precursors in renal fibrosis. J. Am. Soc. Nephrol. 2011, 22, 1876–1886.
  53. Xia, Y.; Yan, J.; Jin, X.; Entman, M.L.; Wang, Y. The chemokine receptor CXCR6 contributes to recruitment of bone marrow-derived fibroblast precursors in renal fibrosis. Kidney Int. 2014, 86, 327–337.
  54. Satoh-Takayama, N.; Serafini, N.; Verrier, T.; Rekiki, A.; Renauld, J.C.; Frankel, G.; Di Santo, J.P. The chemokine receptor CXCR6 controls the functional topography of interleukin-22 producing intestinal innate lymphoid cells. Immunity 2014, 41, 776–788.
  55. Uza, N.; Nakase, H.; Yamamoto, S.; Yoshino, T.; Takeda, Y.; Ueno, S.; Inoue, S.; Mikami, S.; Matsuura, M.; Shimaoka, T.; et al. SR-PSOX/CXCL16 plays a critical role in the progression of colonic inflammation. Gut 2011, 60, 1494–1505.
  56. Mandai, Y.; Takahashi, D.; Hase, K.; Obata, Y.; Furusawa, Y.; Ebisawa, M.; Nakagawa, T.; Sato, T.; Katsuno, T.; Saito, Y.; et al. Distinct Roles for CXCR6+ and CXCR6− CD4+ T Cells in the Pathogenesis of Chronic Colitis. PLoS ONE 2013, 8, e65488.
  57. Liao, F.; Alkhatib, G.; Peden, K.W.; Sharma, G.; Berger, E.A.; Farber, J.M. STRL33, A novel chemokine receptor-like protein, functions as a fusion cofactor for both macrophage-tropic and T cell line-tropic HIV-1. J. Exp. Med. 1997, 185, 2015–2023.
  58. Deng, H.K.; Unutmaz, D.; KewalRamani, V.N.; Littman, D.R. Expression cloning of new receptors used by simian and human immunodeficiency viruses. Nature 1997, 388, 296–300.
  59. Zhang, Y.J.; Zhang, L.; Ketas, T.; Korber, B.T.; Moore, J.P. HIV type 1 molecular clones able to use the Bonzo/STRL-33 coreceptor for virus entry. AIDS Res. Hum. Retrovir. 2001, 17, 217–227.
  60. Blaak, H.; Boers, P.H.; Gruters, R.A.; Schuitemaker, H.; van der Ende, M.E.; Osterhaus, A.D. CCR5, GPR15, and CXCR6 are major coreceptors of human immunodeficiency virus type 2 variants isolated from individuals with and without plasma viremia. J. Virol. 2005, 79, 1686–1700.
  61. Limou, S.; Coulonges, C.; Herbeck, J.T.; van Manen, D.; An, P.; Le Clerc, S.; Delaneau, O.; Diop, G.; Taing, L.; Montes, M.; et al. Multiple-cohort genetic association study reveals CXCR6 as a new chemokine receptor involved in long-term nonprogression to AIDS. J. Infect. Dis. 2010, 202, 908–915.
  62. Picton, A.C.P.; Paximadis, M.; Chaisson, R.E.; Martinson, N.A.; Tiemessen, C.T. CXCR6 gene characterization in two ethnically distinct South African populations and association with viraemic disease control in HIV-1-infected black South African individuals. Clin. Immunol. 2017, 180, 69–79.
  63. Zhang, W.; Ambikan, A.T.; Sperk, M.; van Domselaar, R.; Nowak, P.; Noyan, K.; Russom, A.; Sönnerborg, A.; Neogi, U. Transcriptomics and Targeted Proteomics Analysis to Gain Insights Into the Immune-control Mechanisms of HIV-1 Infected Elite Controllers. EBioMedicine 2018, 27, 40–50.
  64. Duggal, P.; An, P.; Beaty, T.H.; Strathdee, S.A.; Farzadegan, H.; Markham, R.B.; Johnson, L.; O’Brien, S.J.; Vlahov, D.; Winkler, C.A. Genetic influence of CXCR6 chemokine receptor alleles on PCP-mediated AIDS progression among African Americans. Genes Immun. 2003, 4, 245–250.
  65. Passam, A.M.; Sourvinos, G.; Krambovitis, E.; Miyakis, S.; Stavrianeas, N.; Zagoreos, I.; Spandidos, D.A. Polymorphisms of Cx(3)CR1 and CXCR6 receptors in relation to HAART therapy of HIV type 1 patients. AIDS Res. Hum. Retrovir. 2007, 23, 1026–1032.
  66. Landrø, L.; Damås, J.K.; Halvorsen, B.; Fevang, B.; Ueland, T.; Otterdal, K.; Heggelund, L.; Frøland, S.S.; Aukrust, P. CXCL16 in HIV infection—A link between inflammation and viral replication. Eur. J. Clin. Investig. 2009, 39, 1017–1024.
More
Information
Contributor MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register :
View Times: 554
Revisions: 2 times (View History)
Update Date: 14 Apr 2021
1000/1000