Next Article in Journal
Micro-RNA and Proteomic Profiles of Plasma-Derived Exosomes from Irradiated Mice Reveal Molecular Changes Preventing Apoptosis in Neonatal Cerebellum
Next Article in Special Issue
Skin Inflammation Modulation via TNF-α, IL-17, and IL-12 Family Inhibitors Therapy and Cancer Control in Patients with Psoriasis
Previous Article in Journal
Utilizing Genomically Targeted Molecular Data to Improve Patient-Specific Outcomes in Autism Spectrum Disorder
Previous Article in Special Issue
CXCL1: Gene, Promoter, Regulation of Expression, mRNA Stability, Regulation of Activity in the Intercellular Space
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

CXCR2 Receptor: Regulation of Expression, Signal Transduction, and Involvement in Cancer

by
Jan Korbecki
1,
Patrycja Kupnicka
1,
Mikołaj Chlubek
1,
Jarosław Gorący
2,
Izabela Gutowska
3 and
Irena Baranowska-Bosiacka
1,*
1
Department of Biochemistry and Medical Chemistry, Pomeranian Medical University in Szczecin, Powstańców Wlkp. 72 Av., 70-111 Szczecin, Poland
2
Clinic of Cardiology, Pomeranian Medical University in Szczecin, Powstańców Wlkp. 72 Av., 70-111 Szczecin, Poland
3
Department of Medical Chemistry, Pomeranian Medical University in Szczecin, Powstańców Wlkp. 72 Av., 70-111 Szczecin, Poland
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(4), 2168; https://doi.org/10.3390/ijms23042168
Submission received: 4 February 2022 / Revised: 12 February 2022 / Accepted: 14 February 2022 / Published: 16 February 2022
(This article belongs to the Special Issue Inflammation and Cancer 2021)

Abstract

:
Chemokines are a group of about 50 chemotactic cytokines crucial for the migration of immune system cells and tumor cells, as well as for metastasis. One of the 20 chemokine receptors identified to date is CXCR2, a G-protein-coupled receptor (GPCR) whose most known ligands are CXCL8 (IL-8) and CXCL1 (GRO-α). In this article we present a comprehensive review of literature concerning the role of CXCR2 in cancer. We start with regulation of its expression at the transcriptional level and how this regulation involves microRNAs. We show the mechanism of CXCR2 signal transduction, in particular the action of heterotrimeric G proteins, phosphorylation, internalization, intracellular trafficking, sequestration, recycling, and degradation of CXCR2. We discuss in detail the mechanism of the effects of activated CXCR2 on the actin cytoskeleton. Finally, we describe the involvement of CXCR2 in cancer. We focused on the importance of CXCR2 in tumor processes such as proliferation, migration, and invasion of tumor cells as well as the effects of CXCR2 activation on angiogenesis, lymphangiogenesis, and cellular senescence. We also discuss the importance of CXCR2 in cell recruitment to the tumor niche including tumor-associated neutrophils (TAN), tumor-associated macrophages (TAM), myeloid-derived suppressor cells (MDSC), and regulatory T (Treg) cells.

Graphical Abstract

1. Introduction

Forty-six human chemokines, grouped into four sub-families according to the number and spacing of two to four highly conserved N-terminal cysteines, are known to activate 20 chemokine receptors [1] and an additional 4 atypical chemokine receptors [2]. The names of the receptors correspond to the sub-family of chemokines they bind, with 10 receptors for CC chemokines, one receptor each for CX3C and XC chemokines, and 7 receptors for CXC chemokines [1]. One of the CXC receptors is CXC motif chemokine receptor 2 (CXCR2)—the receptor for the well-known studied chemokine CXC motif chemokine ligand 8 (CXCL8)/interleukin-8 (IL-8) [3].
These receptors are responsible for the properties of the chemokines. This is associated with the fact that a chemokine must activate its receptor in order to act. Signal transduction from such a receptor causes changes in the target cell, which are described as the action of a particular chemokine. Despite this, on PubMed (https://pubmed.ncbi.nlm.nih.gov; accessed on 30 December 2021), much more research is devoted to the chemokines than to their receptors, with more than 15 times as many papers on CXCL8/IL-8 than those about its receptors, CXC motif chemokine receptor 1 (CXCR1) and CXCR2. This is likely due to the fact that a given interaction between two cells is associated with the secretion of a ligand (chemokine) then an increase in the sensitivity of the target cell to the chemokines. For this reason, the importance of the chemokine receptors in biological processes is very often overlooked.
In an attempt to fill this gap, the aim of this review is to provide an insight into the activity of CXCR2—its expression, activation, signal transduction, and involvement in cancer.

2. CXCR2 Gene and Regulation of Expression

2.1. CXCR2 Gene

CXCR2 was described for the first time in 1991 [4]. Its link to chemokines was discovered a year later when, together with CXCR1, it turned out to be a high affinity receptor for CXCL8/IL-8 [5]. CXCR2 was first named interleukin-8 receptor B (IL-8R-B); other variants of this name include IL-8R2 [6] and class II IL-8R [7].
The CXCR2 and CXCR1 genes are located at 2q34-q35 [8,9], forming a gene cluster with CXC motif chemokine receptor 2 pseudogene 1 (CXCR2P1). Proteins CXCR1 and CXCR2 show a 77% similarity in amino acid sequence [5,8], while the nucleotide sequence of CXCR2P1 is 87% similar to CXCR2 [9], although there are many frameshifts and stop codons in the sequence of CXCR2P1 compared to CXCR2. All this information indicates that CXCR1, CXCR2, and CXCR2P1 are derived from a common ancestral gene that has been duplicated.

2.2. Regulation of CXCR2 Expression

The promoter of the CXCR2 gene contains a non-classical TATA-box which means that additional transcription initiation factors are required for the transcription of this gene [10]. Near the TATA-box are specificity protein 1 (Sp1) and activating protein-(AP-)2 binding sites [11]. Further away are CAAT-boxes, NF-IL6, AP-2, AP-3, and another Sp1 binding site [10]. The CXCR2 gene promoter also contains two nuclear factor κB (NF-κB) binding sites at positions -331bp and -288bp [12].
CXCR2 gene expression is elevated in hypoxia, being associated with the action of hypoxia inducible factor-1 (HIF-1) and NF-κB, as shown on prostate cancer cells [13]. However, the effect of hypoxia on CXCR2 expression may depend on the selected model, such as in gastric cancer cells, where hypoxia reduces CXCR2 expression [14].
CXCR2 expression can be increased by signal transducer and activator of transcription 3 (STAT3), when attached to the CXCR2 promoter [15,16,17]. Moreover, the CXCR2 promoter can undergo histone acetylation on histone 3 at Lys9 (H3K9), which increases the expression of CXCR2 in the spinal cord—a process crucial for the development of chronic pain after traumatic brain injury [18]. The promoter can also be attached to p53, increasing CXCR2 expression for cellular senescence.
The CXCR2 gene consists of 11 exons [12]. The size of the open reading frame of this gene is 1065bp, and it is entirely located on the last exon [11]. The remaining exons form 11 different 5′-untranslated regions via alternative splicing [12]. So far, the significance of the formation of such a large number of different CXCR2 mRNAs has not been understood—it may be related to the regulation of CXCR2 expression. The 3′-UTR of CXCR2 mRNA also contains clustered adenine-uridine rich sequence elements (AREs), which indicate that the regulation of CXCR2 expression may take place via the regulation of mRNA stability, although this has never been experimentally confirmed [11].

2.3. Involvement of MicroRNAs in the Regulation of CXCR2 Expression

Most studies on CXCR2 promoter function were carried out in the 1990s and were not followed up with detailed research in the following decades. For this reason, information on this issue is somewhat limited.
Much better research was undertaken on the regulation of CXCR2 expression by miRNA, a very significant factor in cancer (Figure 1). However, we know only one miRNA involved in this process in the context of cancer, namely, miR-940 [19,20]. For example, a decrease in its expression in hepatocellular carcinoma was found to increase CXCR2 expression, which then leads to the increased migration of tumor cells [19]. A similar mechanism was demonstrated for tongue squamous cell carcinoma, where the downregulation of miR-940 expression increased CXCR2 expression and thus the response of these cancer cells to CXCR2 ligands.
The regulation of CXCR2 expression is important for maintaining endothelial integrity. Toxic agents such as oxidized low-density lipoprotein (oxLDL) increase the expression of metastasis associated lung adenocarcinoma transcript 1 (MALAT1), which sponges miR-22-3p, a miRNA that directly decreases CXCR2 expression [21].
Another miRNA, miR-155-5p, regulates CXCR2 expression during osteoclast differentiation [22]. Decreased expression of this miRNA results in an increase in CXCR2 expression, which leads to osteoclast differentiation, as shown in root resorption associated with orthodontic tooth movement.
In granulocytes and NK cells, CXCR2 expression is regulated by miR-4437 [23]. The expression of this miRNA is decreased in these cells in patients with ischemic stroke, and leads to increased expression of CXCR2 on granulocytes and NK cells, triggering the migration of these cells to the brain where they cause ischemic brain injury.
In silico analysis has shown some association of CXCR2 with atrial fibrillation—potential miRNAs that decrease CXCR2 expression are miR-5001-3p and miR-4283 [24]. Another in silico analysis of genes associated with neuropathic pain has shown that CXCR2 expression may be increased by reducing the expression of miR-7688-3p [25]. The study was associated with the earlier observation that CXCR2 is a receptor important in the development of neuropathic pain [26].
Some miRNAs that decrease CXCR2 expression may originate outside the human body. For example, preparations containing Cordyceps militaris, a fungus parasitizing on butterfly and moth caterpillars and that is used in traditional Chinese medicine, may decrease CXCR2 expression. The fungus contains miR-1321 and miR-3188—microRNAs that directly decrease CXCR2 expression [27].
It is also possible that miRNAs can influence CXCR2 activity. An example of this is miR-K12-3 [28], a miRNA from Kaposi’s sarcoma-associated herpesvirus (KSHV) that directly reduces the expression of G protein-coupled receptor kinase 2 (GRK2). The activity of this kinase leads to the desensitization and internalization of CXCR2 by phosphorylating this receptor [29]. The reduction in GRK2 expression increases the length of CXCR2 activation by ligands, a process crucial in Kaposi sarcoma development as it leads to angiogenesis and to KSHV latency.

3. CXCR2 Ligands

CXCR2 is a receptor for seven CXC chemokines: CXCL1, CXCL2, CXCL3, CXCL5, CXCL6, CXCL7, and CXCL8 [30,31]. Its strongest agonist is CXCL3 whose half maximal effective concentration (EC50) is 1 nM [30]; the weakest agonist is CXCL5 whose EC50 is 10.6 nM [30]. In contrast, chemokines CXCL1, CXCL2, CXCL6, CXCL7, and CXCL8 activate CXCR2 in the EC50 range 3 nM–7 nM [30,31]. CXCR2 activation by CXCL6 and CXCL8 occurs at the lowest concentrations required to activate CXCR1 [30,31]. On the other hand, some CXCR2 agonists are weak CXCR1 agonists, activating this receptor at concentrations above 40 nM. CXCR2 can also be activated by macrophage migration inhibitory factor (MIF) [32,33,34], although its affinity to CXCR2 is half that of CXCL8. The dissociation constant (Kd) for MIF is 1.4 nM, while for CXCL8 it is 0.7 nM [32].
CXC chemokines that are ligands for CXCR2 can be divided into two groups: those that activate it at similar concentrations to CXCR1 activation (CXCL6 and CXCL8) and those at much higher concentrations (CXCL1, CXCL2, CXCL3, CXCL5, and CXCL7). Ligand binding by both these receptors occurs via the extracellularly located N-terminus, which determines ligand binding specificity [35]. The receptors show 77% similarity in terms of amino acid sequence [5,8] but the differences at the ends of the sequences result in differences in ligand binding specificity. Differences also occur in the binding of various CXC chemokines to CXCR2. The same or different amino acid residues may be responsible for the binding of individual CXC chemokines by CXCR1 and CXCR2. Asp9, Glu12, Lys108, and Lys120 are crucial for CXCL8 binding by CXCR2 [36], compared to Glu7, Asp9, and Glu12 for binding CXCL1. Differences in the binding of a particular chemokine are also related to differences between the chemokines. For example, significant differences between CXCL8 and CXCL1 in the loop preceding the first β-strand lead to a significantly different affinity of these two chemokines for the CXCR1 receptor [37]. At the same time, the residues responsible for binding a given chemokine may be different from the residues important in CXCR2 activation [36]. This indicates that the binding of the chemokine to CXCR2 is not associated with the activation of this receptor although both processes always occur at the same time.
Another poorly researched aspect of the CXCR2 function is the dimerization of the chemokine receptors. CXCR2 forms homodimers even in the absence of a ligand [38,39]. Homodimerization of CXCR2 occurs via the Ala106-Lys163 region [38]. It is also possible to heterodimerize CXCR1 with CXCR2 if these receptors are expressed simultaneously in the cell [39]. Ligands cause the breakdown of CXCR1:CXCR2 heterodimers but stabilize CXCR1:CXCR1 and CXCR2:CXCR2 homodimers, which then can undergo internalization [40]. Nevertheless, CXCR2 dimerization has been poorly researched and little is known about the role of this process in signal transduction from this receptor.

4. CXCR2 Signal Transduction

4.1. Heterotrimeric G Proteins and CXCR2

The CXCR2 receptor is a seven transmembrane G protein-coupled receptor (GPCR) due to the role heterotrimeric G proteins play in signal transduction from this receptor (Figure 2) [7,41]. CXCR2 couples to inhibitory guanine nucleotide regulatory protein (Gαi) [42,43,44,45,46]. As Gαi is inhibited by pertussis toxin [47], signal transduction from CXCR2 is sensitive to pertussis toxin, yet it does not completely inhibit signal transduction. In particular, it does not completely inhibit Ca2+ flux which indicates the involvement of other G proteins or signaling pathways [47]. After CXCR2 activation and signal transduction on Gαi, heterotrimeric G protein dissociates into a heterodimer of Gβγ complex and Gαi subunit [48]. Gβγ can cause PI3K activation and signal transduction via the phosphatidylinositol-4,5-bisphosphate 3-kinase (PI3K)-protein kinase B (PKB)/Akt pathway [49]. Gβγ can also activate phospholipase C-β (PLC-β) [50,51]. This enzyme generates diacylglycerol (DAG) and inositol triphosphate (IP3); the latter of these second messengers is responsible for Ca2+ mobilization. In turn, the Gαi subunit is responsible for the inhibition of adenylyl cyclase [47,52,53,54]. With that said, the effect of CXCR2 on adenylyl cyclase appears to be a minor pathway in CXCR2 action compared to PI3K and PLC-β activation.
Gαi activity—and thus signal transduction from CXCR2—can be regulated. Immediately after CXCR2 activation, a G protein-coupled receptor kinase 6 (GRK6) complex is formed with the activator of G protein signaling 3 (AGS3) and Gαi2 [45]. The process of formation of this complex is dependent on the activation of Gαi2. This is followed by inhibition of the activity of Gαi. With this process of complex formation, phosphorylation of AGS3 by protein kinase C (PKC) occurs. However, the effect of this phosphorylation is not known. There are also other mechanisms that decrease CXCR2 activity. Moreover, the regulator of G protein signaling 5 (RGS5) inhibits CXCR2 activity in neutrophils and inactivates Gαi [55].
Physical coupling of CXCR2 to Gαi isoforms and the exact signal transduction is dependent on the cell type. For example, in neutrophils CXCR2 couples to Gαi2 and Gαi3 [44,45,46]. However, in human embryonic kidney (HEK)-293 cells transfected with CXCR2, this receptor couples to Gαi2 but not Gαi1 or Gαi3 [42,43]. Coupling with a given G protein is dependent on the expression level of the given Gαi in the cell [46]. Importantly, various Gαi have different roles in signal transduction from CXCR2. In neutrophils, Gαi2 is responsible for Ca2+ flux and neutrophil arrest in blood vessels [44,46]. In contrast, Gαi3 in neutrophils is responsible for neutrophil transmigration, migration, and activation of the PI3K-PKB/Akt pathway [46].

4.2. Phosphorylation of CXCR2

During signal transduction, CXCR2 becomes phosphorylated at the C-terminus residues Ser342, Ser346, Ser347, and Ser348 (Figure 3) which leads to desensitization, internalization, and sequestration of CXCR2 [7,56,57]. However, phosphorylation of CXCR2 at Ser346 and Ser348 does not induce internalization of CXCR2 [57]. CXCR2 phosphorylation is triggered by serine-threonine protein kinase: G protein-coupled receptor kinases (GRKs), in particular GRK2 [28,29] and GRK6 [58]. After internalization, CXCR2 undergoes dephosphorylation by protein phosphatase 2A (PP2A) [59,60]. This is followed by resensitization, i.e., the receptor is sensitive to its ligands again. Then it is transported back to the cell membrane.
CXCR2 can also undergo PKC-dependent desensitization [61]. In particular, activation of PKCε by CXCR1 and CCR5 causes cross-desensitization of CXCR2 [62]. In this process, PKCε phosphorylates the cytoplasmic tail of CXCR2. CXCR1 is the receptor for CXCL8, while CC motif chemokine receptor 5 (CCR5) is the receptor for CC motif chemokine ligand (CCL)3, CCL4, and CCL5 [1]. There are other mechanisms that decrease the activity of CXCR2. Therefore, these chemokines activate the CXCR1 and CCR5 receptors. This causes activation of PKCε, which phosphorylates and desensitizes CXCR2. In this way these chemokines disrupt CXCR2.

4.3. CXCR2 Chemosynapse

Many proteins interact with CXCR2 on the cytoplasmic side (Figure 4). Significantly, various proteins can attach to activated and non-activated CXCR2. Following activation of the receptor, some of these proteins stop directly binding to CXCR2, some start interacting with CXCR2, and some are still bound to CXCR2 regardless of its activation [63]. All proteins bound to CXCR2 have been called CXCR2 “chemosynapse” [63,64].
One of the proteins that interacts with CXCR2 is Na+/H+ exchanger regulatory factor-1 (NHERF1), an adaptor/scaffold protein [65]. Adaptor/scaffold proteins help in signal transduction by binding signaling pathway elements. NHERF1 has PDZ motifs that bind CXCR2 and PLC-β [65,66,67]. This interaction is essential for the activation of PLC-β by CXCR2 and thus signal transduction by this bound protein.
Another protein associated with the C-terminus of CXCR2 is IQ motif containing GTPase activating protein 1 (IQGAP1) [63]. This protein is associated with non-activated as well as with activated CXCR2. CXCR2 activation results in binding of activated cell division control protein 42 homolog (Cdc42) to IQGAP1. This interaction is important in Cdc42 activity because IQGAP1 inhibits the GTPase activity of Cdc42 and thus increases the activity of Cdc42 [68]. Moreover, CXCR2 activation results in F-actin binding to IQGAP1. Binding of actin filaments (F-actin) to IQGAP1 and Cdc42 activity leads to the formation of cross-link actin filaments which results in chemotaxis [63].
CXCR2 activation is followed by the attachment of vasodilator-stimulated phosphoprotein (VASP) to this receptor [63,69]. This is caused by the phosphorylation of Ser239 VASP by PKCδ and Ser157 by protein kinase A (PKA). The activation of PKCδ by CXCR2 most likely occurs through the activation of PLC-β by Gβγ [46,50,51]. PLC-β induces the generation of DAG and IP3, the former being the activator of PKCδ [70]. At the same time, PKCδ does not have a Ca2+-binding domain and thus Ca2+ flux does not affect the activity of this enzyme [71]. In contrast, the mechanism of PKA activation by CXCR2 is unclear. CXCR2 activation induces a signal transduction on Gαi [42,43,44,45,46]. This G protein causes inhibition of adenylyl cyclase, thus reducing the generation of PKA activator cyclic adenosine monophosphate (cAMP) [47,53,54]. Although this means that CXCR2 should not increase but rather decrease PKA activity, it has been postulated that CXCR2 does increase cAMP generation through Ca2+ flux and activation of calcium-dependent adenylyl cyclase [46]. Upon phosphorylation, VASP begins to bind F-actin [69], resulting in VASP binding to CXCR2 and F-actin binding to the CXCR2 chemosynapse. This process is important in changes in the actin cytoskeleton under the influence of CXCR2 activity, resulting in chemotaxis.
The C-terminus of CXCR2 contains a leucine-rich domain—an LLKIL motif spanning 325–329aa of CXCR2 [72]. This domain is essential for the chemotactic response and is important but not essential in CXCR2 internalization and PKB/Akt and extracellular signal-regulated kinase (ERK) mitogen-activated protein kinase (MAPK) activation [73]. Various proteins bind to the domain, in particular, LIM and SH3 protein 1 (LASP-1) which bind independently of CXCR2 activation [74]. LASP-1 reduces basal c-Src activation and also participates in the activation of this protein. In turn, Src activates paxillin [74] and the engulfment and cell motility (ELMO)-dedicator of cytokinesis 2 (Dock2)- Rac family small GTPase 2 (Rac2) pathway [75] which is important in chemotaxis. Moreover, LASP-1 is important in the activation of the p130CAS-Rac1 pathway [74,76]. Rac1 [77] and Rac2 [75] can also be activated by CXCR2 through PI3K, which leads to further signal transduction as Rac family small GTPase 1 (Rac1) can activate p38 MAPK [78] and PAK [77]. LASP-1 is crucial for the activation of Cdc42-p21-activated kinase 1 (PAK1) cascade and ERK MAPK by CXCR2 [74,79]. Significantly, PAK1 activation is independent of ERK MAPK activation. All pathways in which LASP-1 participates affect cytoskeletal organization and thus cell migration.
In addition, the activation of CXCR2 is associated with the binding of Hsc70-interacting protein (Hip) to the leucine-rich domain at the C-terminus [72,73,80], resulting in the formation of a CXCR2-Hip-heat shock protein family A (Hsc70) complex essential for the internalization of CXCR2 and chemotaxis. CXCR2 activation also results in the binding of adaptor protein 2 (AP2) to the LLKIL motif [72,73,81]. By binding to membrane phosphatidylinositol phospholipid, AP2 docks in the plasma membrane. With the involvement of α-tubulin acetyl transferase, AP2 causes microtubule acetylation which is essential for directional cell locomotion [81,82].
Other proteins also start to interact with activated CXCR2, such as secretory carrier membrane protein 2 (SCAMP2), Nipsnap homolog 1, and dynein heavy chain 5, all important in intracellular transport [63]. Moreover, proteasome 20S subunit α 2 (PSMA2) and 26S proteosome subunit 9 begin to bind to activated CXCR2 [63].
Proteins that only interact with non-activated CXCR2 include P21-Arc, gelsolin, and plastin, important for the function of the actin cytoskeleton [63]. It is likely that these three proteins are uncoupled from CXCR2 upon activation of this receptor, which leads to changes in the actin cytoskeleton. Other proteins binding only to non-activated CXCR2 are Rab7, annexin 1, kinesin light chain-2, valosin-containing protein (VCP), important in intracellular transport. Furthermore, binding to inactivated CXCR2 are 14-3-3γ, an adaptor/scaffold protein, and Hsp90, Hsp75, and chaperonin TCP1 [63].
In contrast, some proteins bind to both activated and non-activated CXCR2 [63]. These include actin related protein 2/3 complex subunit 2, important in actin cytoskeleton function. Another protein from this group is 14-3-3ζ which is an adaptor/scaffold protein. Nevertheless, the role of the proteins mentioned in this subsection in signal transduction from CXCR2 is yet to be fully understood.

4.4. Other Pathways Activated by CXCR2

CXCR2 also activates pathways whose mechanisms are much less understood than the mentioned signaling pathways. For example, CXCR2 activates ERK MAPK via a variety of pathways. Activation of ERK MAPK by CXCR2 is sensitive to pertussis toxin, showing that it depends on G proteins which activate mitogen-activated protein kinase kinase 1 (MEK1) [52,83]. ERK MAPK activation by CXCR2 may also depend on Raf-MEK-ERK MAPK activation by activated STAT3 [15]. ERK MAPK activation may also be induced by reactive oxidant species (ROS) [84]. The generation of ROS in signal transduction from CXCR2 is dependent on NADPH oxidase activation—a process whose initiation is associated with Rac, activated by CXCR2 via PI3K [77]. ROS participates in the activation of c-Raf-MEK and subsequently ERK MAPK, as well as other MAPK cascades [84]. This entire process is inhibited by β-arrestin, preventing oxidative bursts and thus protecting the cell from death. β-arrestin also causes the inhibition of activation of other MAPK kinase cascades.
Activated CXCR2 can also indirectly activate ERK MAPK via the transactivation of epidermal growth factor receptor (EGFR) [85,86], in a mechanism identified in ovarian cancer cells.
CXCR2 also increases integrin β1 expression, which enhances focal adhesion kinase (FAK) activation [87]. At the same time, integrin activation may be induced by Ras-related protein 1 (Rap1) activation by CXCR2 [88]. FAK activation may also be associated with the PI3K-PKB/Akt or Gαi subunit pathway, although this mechanism is yet to be confirmed [89]. FAK activation is followed by an increase in the expression of MMP2 and MMP9 which leads to the migration of tumor cells [89]. In particular, this process is important in lymphatic metastasis due to the secretion of CXCR2 ligands (CXCL1) by lymphatic endothelial cells.
CXCR2 activation also causes the phosphorylation of janus tyrosine kinase 2 (JAK2) and STAT3 [17,90]. The interaction between STAT3 and CXCR2 is a feed-forward loop where the activation of STAT3 increases the expression of CXCR2, which then activates STAT3 [16,17]. STAT3 activated by CXCR2 is important in the activation of the Raf-MEK-ERK MAPK pathway [15]. However, the very mechanism of STAT3 activation by CXCR2 is still unclear as no proteins relevant for direct activation of JAK2 bind directly to CXCR2 [63]. It is possible that CXCR2 may increase JAK2 expression which then may induce STAT3 activation [91]. In contrast, in granulocyte and macrophage progenitor cells (GMPs), CXCR2 activation results in decreased expression of Sin3-associated 18 kDa polypeptide (SAP18), which leads to an increase in PI3K expression and thus increased activation of ERK MAPK and STAT3 [92]. However, this mechanism of STAT3 activation appears to occur only in these cells.
CXCR2 has also been found to activate ras homolog family member A (RhoA), a representative of the Rho family of monomeric GTPases [93]. The activation of RhoA is important in endothelial cell chemotaxis and thus in angiogenesis induced by CXCR2 ligands. The activation of this protein by CXCR2 remains poorly understood.

4.5. Internalization of CXCR2

The aforementioned LLKIL motif is also important in the internalization of CXCR2 because it binds Hip [72,80] and AP2 [81], proteins involved not only in migration but also in the internalization of CXCR2. The latter protein, AP2, binds clathrin and β-arrestin and thus participates in the internalization of CXCR2 (Figure 5) [81]. CXCR2 also binds β-arrestin independently of AP2 [81,94] but is dependent on the phosphorylation of CXCR2. The internalization of CXCR2 requires the ubiquitination of Lys327 CXCR2 [95], a residue located in the aforementioned leucine-rich domain, which suggests that the ubiquitination is important in the activity of AP2 and Hip, although this needs to be verified.
The first step of CXCR2 internalization involves the formation of a clathrin-coated pit, while the subsequent steps involve Rab5, adaptin 2, and dynamin [59,94,96]. Nevertheless, CXCR2 internalization may occur independently of clathrin [81]. After internalization, CXCR2 is no longer in the cell membrane and is found in endosomes in the cytoplasm. This state is called sequestration. Importantly, internalization of CXCR2 is faster than that of CXCR1 [97]. However, recycling of CXCR2 takes much longer than CXCR1.

4.6. CXCR2 Recycling or Degradation

Upon activation, CXCR2 is internalized and then undergoes intracellular trafficking in a process dependent on Rab proteins [98]. The internalization of CXCR2 and formation of early endosomes with this receptor are dependent on clathrin, Rab5, and dynamin [96,99]. In this location, i.e., in the cytoplasm on endosomes, CXCR2 is not activated because it is no longer in the cell membrane—this state is called receptor sequestration. Subsequently, CXCR2 undergoes sorting decisions. It can be transported to recycling endosomes in a Rab11a-dependent process [99], and then recycled to the cell membrane, where it may become activated again and be involved in signal transduction. However, prolonged action by CXCR2 agonists is associated with a translocation of this receptor to early endosomes, then to late endosomes, and finally to lysosomes, which results in the degradation of CXCR2. Rab7 is responsible for directing CXCR2 into the degradation pathway [99]. Ras homolog B (RhoB), a GTPase activated by CXCR2, is responsible for sorting decisions [100]. GDP-bound RhoB induces the recycling of CXCR2, while GTP-bound RhoB is important in directing CXCR2 to lysosomes. In the absence of RhoB expression, the translocation of CXCR2 from endosomes to the cell membrane occurs via alternative recycling pathways, a part of the trans-Golgi network dependent on Rab4 and mannose-6-phosphate receptor.

5. The Role of CXCR2 in Cancer

CXCR2 has significant pro-tumor functions, where increased CXCR2 expression is associated with poorer patient prognosis, as shown by studies on acute myeloid leukemia [101], invasive ductal breast cancer [102], colorectal cancer [103], esophageal cancer [104], gastric cancer [16,17,105,106], intrahepatic cholangiocellular carcinoma [107], laryngeal squamous cell carcinoma [108], lung adenocarcinoma [109], non-small cell lung cancer [91], ovarian cancer [110,111], pancreatic ductal adenocarcinoma [112]. In contrast, in studies on triple-negative breast cancer [113], colorectal cancer [114], clear cell renal cell carcinoma [115], gastric cancer [116] and pancreatic ductal adenocarcinoma [117] (Table 1), no such association is shown, which may be related to the initiation of an antitumor immune response resulting in tumor infiltration by tumor-infiltrating lymphocytes (TILs), in particular CD8+ cytotoxic lymphocytes [113]. Additionally, CXCR2 activation may be significant in the anticancer functions of neutrophils [118].
In tumors, particularly in melanoma, CXCR2 is activated in an autocrine manner [119]. CXCR2-induced NF-κB activation increases the expression of the CXCR2 ligands which activate CXCR2. Activation of NF-κB via CXCR2 happens through a process dependent on Ras, mitogen-activated protein kinase kinase kinase 1 (MEKK1), and p38 [120]. Moreover, CXCR2-induced NF-κB activation occurs via the PI3K-PKB/Akt pathway [121] and via transforming growth factor (TGF)-β-activating kinase 1 (TAK1) [122]. At the same time, activation of this pathway may not so much directly depend on CXCR2 but rather on the transactivation of EGFR [121]. This activation of NF-κB leads to the expression of pro-inflammatory genes, including ligands for CXCR2 [119]. Moreover, the activation of NF-κB by CXCR2 causes cancer cell migration [121,122].
CXCR2 increases tumor cell proliferation in cancer. First of all, this happens via EGFR as a result of metalloprotease activation that cleaves heparin-binding epidermal growth factor-like growth factor (HB-EGF) [86], an EGFR ligand. HB-EGF, probably released via the actions of enzymes a disintegrin and metalloproteinase 17 (ADAM17) [123,124] or cathepsin B [125], induces the transactivation of EGFR, which leads to increased proliferation of epithelial ovarian cancer cells [86]. Second, the activation of CXCR2 results in increased expression of early growth response-1 (EGR-1) in esophageal cancer [126], which increases proliferation by elevating the expression of cyclin-dependent kinase 4 (CDK4). Third, CXCR2 activates PKB/Akt, which affects the action of p53 [127] via the activation of murine double minute 2 (Mdm2), an enzyme causing the ubiquitination and subsequent degradation of p53 [128]. Decreased expression of p53 results in decreased expression of p21, a cell cycle inhibitor—in this way CXCR2 activation results in decreased p21 expression and thus increased cell proliferation.
CXCR2 appears to inhibit proliferation in normal cells, which is associated with senescence dependent on p53 [129] and p38 MAPK [130]. Senescence itself causes an increase in the expression of CXCR2 and ligands for this receptor and thus autocrine enhancement of senescence [129,130]. If there is a mutation in the TP53 gene or a decrease in p53 expression, then the described mechanism will not cause growth arrest but will stimulate proliferation [129], which is significant in cancer cell proliferation where p53 dysfunction is common [131]. The association of CXCR2 with senescence is important in the tumor microenvironment, as CXCR2 ligands secreted by cancer cells induce senescence of cancer-associated fibroblast (CAF) [132]. Then, these cells exhibit a senescence-associated secretory phenotype (SASP) that enhances cancer tumor growth [133]. In the absence of CXCR2 activation on CAFs, these cells become myofibroblasts [134].
CXCR2 activation also causes tumor cell migration. For example, it causes epithelial-mesenchymal transition (EMT). As a consequence of signal transduction on PI3K, the activation and accumulation of snail family transcriptional repressor 1 (Snail1) occur in the nucleus, which leads to the expression of EMT-related genes [135]. EMT may also involve the activation of STAT3 by CXCR2 [17]. CXCR2 is also significant in metastasis—if circulating tumor cells express CXCR2, then such cells will metastasize in organs with the high expression of ligands for this receptor. This activation of CXCR2 on a cancer cell in the blood is associated with an increase in vascular cell adhesion molecule-1 (VCAM-1) expression on cells with activated CXCR2, something that has been observed in the metastasis to the lung by osteosarcoma [136]. Moreover, CXCR2 may participate in bone metastasis, in particular in bone destruction during bone metastasis [137], something which is related to the recruitment and increased activity of osteoclasts. This is associated with the expression of CXCR2 on osteoclast precursors which migrate to a cancer cell releasing CXCR2 ligands [138,139]. Finally, one of the significant symptoms of bone cancer and bone metastases is bone pain, which is related to increased expression of CXCR2 ligands in the spinal cord [140,141,142]—CXCR2 activation in the spinal cord leads to bone cancer pain.
CXCR2 is involved in angiogenesis. For example, it is expressed on epithelial cells [143,144,145], which causes proliferation, tube formation, and migration of epithelial cells, leading to the formation of new blood vessels [143,145,146]. CXCR2 expression also occurs on lymphatic endothelial cells [147], which causes migration of these cells and tube formation, leading to the formation of new lymphatic vessels or lymphangiogenesis [147]. CXCR2 and its ligands are also part of the cascade of angiogenic factors. Vascular endothelial growth factor (VEGF), the most significant angiogenic factor, increases the expression of CXCL1 and CXCL8 in epithelial cells [148] and cancer cells [149]. VEGF and the aforementioned ELR+ CXC chemokines jointly participate in angiogenesis. CXCR2 activation also results in the secretion of epidermal growth factor (EGF) by epithelial cells [145], VEGF via STAT3 activation [90], and lymphangiogenic factors such as VEGF-C and VEGF-D by cancer cells [150]. The activation of EGFR on cancer cells also results in the activation of NF-κB and thus expression of pro-angiogenic CXCR2 ligands [151]. Additionally, CXCR2 is important in the mobilization and function of bone marrow-derived endothelial progenitor cells (EPCs) recruited to growing blood vessels [152].
CXCR2 is also significant in the recruitment of different cells to the tumor niche. CXCR2 ligands, upon binding to the receptor, induce recruitment and enhance the pro-tumor properties of tumor-associated macrophages (TAM) [153,154,155]. Recruitment of these cells is associated with CXCR2 expression on CD14 monocytes [156]. CXCR2 is also important in tumor-associated neutrophils (TAN) recruitment [153,157], with pro-tumorigenic N2 neutrophils showing increased CXCR2 expression relative to N1 polarized neutrophils [158]. Other cells that are recruited to the tumor niche via CXCR2 include granulocytic myeloid-derived suppressor cells (MDSCs) [153,159,160,161], regulatory T cells (Treg) [162], and bone marrow-derived mesenchymal cells (BM-MCs) [163]. CXCR2 ligands also indirectly induce the recruitment of monocytic MDSCs to the tumor niche. Increased levels of CXCR2 ligands result in the expansion of monocytic MDSCs in the bone marrow which increases the number of these cells [164].
CXCR2 plays an important role in cancer processes, a fact which can potentially be used in designing effective anti-cancer therapies. One possible course is the use of specific CXCR2 inhibitors to reduce the activity of this receptor [165,166,167,168]. This not only decreases cancer tumor growth but also increases the efficacy of the chemotherapy. In particular, due to the pro-angiogenic properties of CXCR2, inhibitors of this receptor increase the effectiveness of antiangiogenic therapy [169]. Another therapeutic approach is immunotherapy; as antiangiogenic lymphocytes lack CXCR2 receptor expression [170], they are not recruited to malignant tumors with a high concentration of CXCR2 ligands. CXCR2 expression may be increased via transduction in lymphocytes isolated from the patient [171,172,173] and when such transduced lymphocytes are reintroduced into the patient’s body, the cells will migrate to the tumor and destroy the tumor cells.

6. Perspective for Further Research

Current knowledge of CXCR2 is sufficient to understand the principles of the function and the importance of this receptor. However, some aspects require more comprehensive research. In particular, we know very little about the mechanisms regulating CXCR2 expression. Only at the outset of research on CXCR2 was the mechanism of regulation of the expression of this receptor described at the transcriptional level. Furthermore, very little is known about the mechanism of regulation of CXCR2 expression changes in mRNA stability—the CXCR2 mRNA sequence indicates that theoretically such regulation is possible although it is yet to be confirmed experimentally. There is also a considerable gap in knowledge on the significance of some proteins that directly interact with CXCR2 (14-3-3γ, 14-3-3ζ, Hsp90, Hsp75, and chaperonin TCP1). The biggest deficiency in current research on CXCR2 is the lack of research on the receptor itself and the focus on only one or two ligands despite the fact it is activated by seven different chemokines. Many research papers are content with only showing an increase in expression of one of the CXCR2 ligands, not paying any attention to the significance of CXCR2 itself and changes in its expression.

Author Contributions

J.K., writing—original draft preparation; P.K., investigation; M.C., investigation; J.G., investigation; I.G., investigation; I.B.-B., writing—original draft preparation, writing—review and editing, supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the statutory budget of the Department of Biochemistry and Medical Chemistry Pomeranian Medical University in Szczecin, Poland.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hughes, C.E.; Nibbs, R.J.B. A guide to chemokines and their receptors. FEBS J. 2018, 285, 2944–2971. [Google Scholar] [CrossRef] [PubMed]
  2. Graham, G.J.; Locati, M.; Mantovani, A.; Rot, A.; Thelen, M. The biochemistry and biology of the atypical chemokine receptors. Immunol. Lett. 2012, 145, 30–38. [Google Scholar] [CrossRef] [PubMed]
  3. Han, Z.J.; Li, Y.B.; Yang, L.X.; Cheng, H.J.; Liu, X.; Chen, H. Roles of the CXCL8-CXCR1/2 axis in the tumor microenvironment and immunotherapy. Molecules 2021, 27, 137. [Google Scholar] [CrossRef] [PubMed]
  4. Murphy, P.M.; McDermott, D. Functional expression of the human formyl peptide receptor in Xenopus oocytes requires a complementary human factor. J. Biol. Chem. 1991, 266, 12560–12567. [Google Scholar] [CrossRef]
  5. Lee, J.; Horuk, R.; Rice, G.C.; Bennett, G.L.; Camerato, T.; Wood, W.I. Characterization of two high affinity human interleukin-8 receptors. J. Biol. Chem. 1992, 267, 16283–16287. [Google Scholar] [CrossRef]
  6. Loetscher, P.; Seitz, M.; Clark-Lewis, I.; Baggiolini, M.; Moser, B. Both interleukin-8 receptors independently mediate chemotaxis. Jurkat cells transfected with IL-8R1 or IL-8R2 migrate in response to IL-8, GRO alpha and NAP-2. FEBS Lett. 1994, 341, 187–192. [Google Scholar] [CrossRef] [Green Version]
  7. Mueller, S.G.; Schraw, W.P.; Richmond, A. Melanoma growth stimulatory activity enhances the phosphorylation of the class II interleukin-8 receptor in non-hematopoietic cells. J. Biol. Chem. 1994, 269, 1973–1980. [Google Scholar] [CrossRef]
  8. Ahuja, S.K.; Ozçelik, T.; Milatovitch, A.; Francke, U.; Murphy, P.M. Molecular evolution of the human interleukin-8 receptor gene cluster. Nat. Genet. 1992, 2, 31–36. [Google Scholar] [CrossRef]
  9. Mollereau, C.; Muscatelli, F.; Mattei, M.G.; Vassart, G.; Parmentier, M. The high-affinity interleukin 8 receptor gene (IL8RA) maps to the 2q33-q36 region of the human genome: Cloning of a pseudogene (IL8RBP) for the low-affinity receptor. Genomics 1993, 16, 248–251. [Google Scholar] [CrossRef]
  10. Sprenger, H.; Lloyd, A.R.; Kelvin, D.J. Promoter analysis of the human interleukin-8 receptor genes, IL-8RA and IL-8RB. Immunobiology 1995, 193, 334–340. [Google Scholar] [CrossRef]
  11. Sprenger, H.; Lloyd, A.R.; Lautens, L.L.; Bonner, T.I.; Kelvin, D.J. Structure, genomic organization, and expression of the human interleukin-8 receptor B gene. J. Biol. Chem. 1994, 269, 11065–11072. [Google Scholar] [CrossRef]
  12. Ahuja, S.K.; Shetty, A.; Tiffany, H.L.; Murphy, P.M. Comparison of the genomic organization and promoter function for human interleukin-8 receptors A and B. J. Biol. Chem. 1994, 269, 26381–26389. [Google Scholar] [CrossRef]
  13. Maxwell, P.J.; Gallagher, R.; Seaton, A.; Wilson, C.; Scullin, P.; Pettigrew, J.; Stratford, I.J.; Williams, K.J.; Johnston, P.G.; Waugh, D.J. HIF-1 and NF-κB-mediated upregulation of CXCR1 and CXCR2 expression promotes cell survival in hypoxic prostate cancer cells. Oncogene 2007, 26, 7333–7345. [Google Scholar] [CrossRef] [Green Version]
  14. Kinoshita, H.; Yashiro, M.; Fukuoka, T.; Hasegawa, T.; Morisaki, T.; Kasashima, H.; Masuda, G.; Noda, S.; Hirakawa, K. Diffuse-type gastric cancer cells switch their driver pathways from FGFR2 signaling to SDF1/CXCR4 axis in hypoxic tumor microenvironments. Carcinogenesis 2015, 36, 1511–1520. [Google Scholar] [CrossRef] [Green Version]
  15. Nguyen-Jackson, H.; Panopoulos, A.D.; Zhang, H.; Li, H.S.; Watowich, S.S. STAT3 controls the neutrophil migratory response to CXCR2 ligands by direct activation of G-CSF-induced CXCR2 expression and via modulation of CXCR2 signal transduction. Blood 2010, 115, 3354–3363. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Xiang, Z.; Zhou, Z.J.; Xia, G.K.; Zhang, X.H.; Wei, Z.W.; Zhu, J.T.; Yu, J.; Chen, W.; He, Y.; Schwarz, R.E.; et al. A positive crosstalk between CXCR4 and CXCR2 promotes gastric cancer metastasis. Oncogene 2017, 36, 5122–5133. [Google Scholar] [CrossRef] [PubMed]
  17. Zhou, Z.; Xia, G.; Xiang, Z.; Liu, M.; Wei, Z.; Yan, J.; Chen, W.; Zhu, J.; Awasthi, N.; Sun, X.; et al. A C-X-C chemokine receptor type 2-dominated cross-talk between tumor cells and macrophages drives gastric cancer metastasis. Clin. Cancer Res. 2019, 25, 3317–3328. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Liang, D.Y.; Shi, X.; Liu, P.; Sun, Y.; Sahbaie, P.; Li, W.W.; Yeomans, D.C.; Clark, J.D. The chemokine receptor CXCR2 supports nociceptive sensitization after traumatic brain injury. Mol. Pain 2017, 13, 1744806917730212. [Google Scholar] [CrossRef] [Green Version]
  19. Ding, D.; Zhang, Y.; Yang, R.; Wang, X.; Ji, G.; Huo, L.; Shao, Z.; Li, X. MiR-940 suppresses tumor cell invasion and migration via regulation of CXCR2 in hepatocellular carcinoma. Biomed. Res. Int. 2016, 2016, 7618342. [Google Scholar] [CrossRef] [Green Version]
  20. Ma, T.; Zhao, Z.; Wang, Z.; Wang, C.; Zhang, L. MiR-940 inhibits migration and invasion of tongue squamous cell carcinoma via regulatingCXCR2/NF-κB system-mediated epithelial-mesenchymal transition. Naunyn Schmiedebergs Arch. Pharm. 2019, 392, 1359–1369. [Google Scholar] [CrossRef]
  21. Tang, Y.; Jin, X.; Xiang, Y.; Chen, Y.; Shen, C.X.; Zhang, Y.C.; Li, Y.G. The lncRNA MALAT1 protects the endothelium against ox-LDL-induced dysfunction via upregulating the expression of the miR-22-3p target genes CXCR2 and AKT. FEBS Lett. 2015, 589, 3189–3196. [Google Scholar] [CrossRef] [Green Version]
  22. Jiang, H.; Kitaura, H.; Liu, L.; Mizoguchi, I.; Liu, S. The miR-155-5p inhibits osteoclast differentiation through targeting CXCR2 in orthodontic root resorption. J. Periodontal Res. 2021, 56, 761–773. [Google Scholar] [CrossRef] [PubMed]
  23. He, Q.; Shi, X.; Zhou, B.; Teng, J.; Zhang, C.; Liu, S.; Lian, J.; Luo, B.; Zhao, G.; Lu, H.; et al. Interleukin 8 (CXCL8)-CXC chemokine receptor 2 (CXCR2) axis contributes to MiR-4437-associated recruitment of granulocytes and natural killer cells in ischemic stroke. Mol. Immunol. 2018, 101, 440–449. [Google Scholar] [CrossRef] [PubMed]
  24. Fan, G.; Wei, J. Identification of potential novel biomarkers and therapeutic targets involved in human atrial fibrillation based on bioinformatics analysis. Kardiol. Pol. 2020, 78, 694–702. [Google Scholar] [CrossRef] [PubMed]
  25. Li, H.; Wan, H.Q.; Zhao, H.J.; Luan, S.X.; Zhang, C.G. Identification of candidate genes and miRNAs associated with neuropathic pain induced by spared nerve injury. Int. J. Mol. Med. 2019, 44, 1205–1218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Moraes, T.R.; Elisei, L.S.; Malta, I.H.; Galdino, G. Participation of CXCL1 in the glial cells during neuropathic pain. Eur. J. Pharmacol. 2020, 875, 173039. [Google Scholar] [CrossRef] [PubMed]
  27. Liu, S.; Tang, J.; Huang, L.; Xu, Q.; Ling, X.; Liu, J. Cordyceps militaris alleviates severity of murine acute lung injury through miRNAs-mediated CXCR2 inhibition. Cell. Physiol. Biochem. 2015, 36, 2003–2011. [Google Scholar] [CrossRef]
  28. Li, W.; Jia, X.; Shen, C.; Zhang, M.; Xu, J.; Shang, Y.; Zhu, K.; Hu, M.; Yan, Q.; Qin, D.; et al. A KSHV microRNA enhances viral latency and induces angiogenesis by targeting GRK2 to activate the CXCR2/AKT pathway. Oncotarget 2016, 7, 32286–32305. [Google Scholar] [CrossRef]
  29. Alves-Filho, J.C.; Sônego, F.; Souto, F.O.; Freitas, A.; Verri, W.A., Jr.; Auxiliadora-Martins, M.; Basile-Filho, A.; McKenzie, A.N.; Xu, D.; Cunha, F.Q.; et al. Interleukin-33 attenuates sepsis by enhancing neutrophil influx to the site of infection. Nat. Med. 2010, 16, 708–712. [Google Scholar] [CrossRef]
  30. Ahuja, S.K.; Murphy, P.M. The CXC chemokines growth-regulated oncogene (GRO) alpha, GRObeta, GROgamma, neutrophil-activating peptide-2, and epithelial cell-derived neutrophil-activating peptide-78 are potent agonists for the type B, but not the type A, human interleukin-8 receptor. J. Biol. Chem. 1996, 271, 20545–20550. [Google Scholar] [CrossRef] [Green Version]
  31. Wuyts, A.; Proost, P.; Lenaerts, J.P.; Ben-Baruch, A.; van Damme, J.; Wang, J.M. Differential usage of the CXC chemokine receptors 1 and 2 by interleukin-8, granulocyte chemotactic protein-2 and epithelial-cell-derived neutrophil attractant-78. Eur. J. Biochem. 1998, 255, 67–73. [Google Scholar] [CrossRef] [PubMed]
  32. Bernhagen, J.; Krohn, R.; Lue, H.; Gregory, J.L.; Zernecke, A.; Koenen, R.R.; Dewor, M.; Georgiev, I.; Schober, A.; Leng, L.; et al. MIF is a noncognate ligand of CXC chemokine receptors in inflammatory and atherogenic cell recruitment. Nat. Med. 2007, 13, 587–596. [Google Scholar] [CrossRef] [PubMed]
  33. Xu, L.; Li, Y.; Li, D.; Xu, P.; Tian, S.; Sun, H.; Liu, H.; Hou, T. Exploring the binding mechanisms of MIF to CXCR2 using theoretical approaches. Phys. Chem. Chem. Phys. 2015, 17, 3370–3382. [Google Scholar] [CrossRef] [PubMed]
  34. Zhang, H.; Ye, Y.L.; Li, M.X.; Ye, S.B.; Huang, W.R.; Cai, T.T.; He, J.; Peng, J.Y.; Duan, T.H.; Cui, J.; et al. CXCL2/MIF-CXCR2 signaling promotes the recruitment of myeloid-derived suppressor cells and is correlated with prognosis in bladder cancer. Oncogene 2017, 36, 2095–2104. [Google Scholar] [CrossRef]
  35. LaRosa, G.J.; Thomas, K.M.; Kaufmann, M.E.; Mark, R.; White, M.; Taylor, L.; Gray, G.; Witt, D.; Navarro, J. Amino terminus of the interleukin-8 receptor is a major determinant of receptor subtype specificity. J. Biol. Chem. 1992, 267, 25402–25406. [Google Scholar] [CrossRef]
  36. Katancik, J.A.; Sharma, A.; de Nardin, E. Interleukin 8, neutrophil-activating peptide-2 and GRO-alpha bind to and elicit cell activation via specific and different amino acid residues of CXCR2. Cytokine 2000, 12, 1480–1488. [Google Scholar] [CrossRef]
  37. Lowman, H.B.; Slagle, P.H.; DeForge, L.E.; Wirth, C.M.; Gillece-Castro, B.L.; Bourell, J.H.; Fairbrother, W.J. Exchanging interleukin-8 and melanoma growth-stimulating activity receptor binding specificities. J. Biol. Chem. 1996, 271, 14344–14352. [Google Scholar] [CrossRef] [Green Version]
  38. Trettel, F.; di Bartolomeo, S.; Lauro, C.; Catalano, M.; Ciotti, M.T.; Limatola, C. Ligand-independent CXCR2 dimerization. J. Biol. Chem. 2003, 278, 40980–40988. [Google Scholar] [CrossRef] [Green Version]
  39. Wilson, S.; Wilkinson, G.; Milligan, G. The CXCR1 and CXCR2 receptors form constitutive homo- and heterodimers selectively and with equal apparent affinities. J. Biol. Chem. 2005, 280, 28663–28674. [Google Scholar] [CrossRef] [Green Version]
  40. Martínez Muñoz, L.; Lucas, P.; Navarro, G.; Checa, A.I.; Franco, R.; Martínez-A, C.; Rodríguez-Frade, J.M.; Mellado, M. Dynamic regulation of CXCR1 and CXCR2 homo- and heterodimers. J. Immunol. 2009, 183, 7337–7346. [Google Scholar] [CrossRef] [Green Version]
  41. Weis, W.I.; Kobilka, B.K. The molecular basis of G protein-coupled receptor activation. Annu. Rev. Biochem. 2018, 87, 897–919. [Google Scholar] [CrossRef] [PubMed]
  42. Damaj, B.B.; McColl, S.R.; Neote, K.; Songqing, N.; Ogborn, K.T.; Hébert, C.A.; Naccache, P.H. Identification of G-protein binding sites of the human interleukin-8 receptors by functional mapping of the intracellular loops. FASEB J. 1996, 10, 1426–1434. [Google Scholar] [CrossRef] [PubMed]
  43. Damaj, B.B.; McColl, S.R.; Mahana, W.; Crouch, M.F.; Naccache, P.H. Physical association of Gi2α with interleukin-8 receptors. J. Biol. Chem. 1996, 271, 12783–12789. [Google Scholar] [CrossRef] [Green Version]
  44. Zarbock, A.; Deem, T.L.; Burcin, T.L.; Ley, K. Gαi2 is required for chemokine-induced neutrophil arrest. Blood 2007, 110, 3773–3779. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Singh, V.; Raghuwanshi, S.K.; Smith, N.; Rivers, E.J.; Richardson, R.M. G protein-coupled receptor kinase-6 interacts with activator of G protein signaling-3 to regulate CXCR2-mediated cellular functions. J. Immunol. 2014, 192, 2186–2194. [Google Scholar] [CrossRef] [Green Version]
  46. Kuwano, Y.; Adler, M.; Zhang, H.; Groisman, A.; Ley, K. Gαi2 and Gαi3 differentially regulate arrest from flow and chemotaxis in mouse neutrophils. J. Immunol. 2016, 196, 3828–3833. [Google Scholar] [CrossRef] [Green Version]
  47. Hall, D.A.; Beresford, I.J.; Browning, C.; Giles, H. Signalling by CXC-chemokine receptors 1 and 2 expressed in CHO cells: A comparison of calcium mobilization, inhibition of adenylyl cyclase and stimulation of GTPγS binding induced by IL-8 and GROα. Br. J. Pharmacol. 1999, 126, 810–818. [Google Scholar] [CrossRef] [Green Version]
  48. Wettschureck, N.; Offermanns, S. Mammalian G proteins and their cell type specific functions. Physiol. Rev. 2005, 85, 1159–1204. [Google Scholar] [CrossRef] [Green Version]
  49. Stephens, L.R.; Eguinoa, A.; Erdjument-Bromage, H.; Lui, M.; Cooke, F.; Coadwell, J.; Smrcka, A.S.; Thelen, M.; Cadwallader, K.; Tempst, P.; et al. The Gβγ sensitivity of a PI3K is dependent upon a tightly associated adaptor, p101. Cell 1997, 89, 105–114. [Google Scholar] [CrossRef] [Green Version]
  50. Boyer, J.L.; Graber, S.G.; Waldo, G.L.; Harden, T.K.; Garrison, J.C. Selective activation of phospholipase C by recombinant G-protein alpha- and beta gamma-subunits. J. Biol. Chem. 1994, 269, 2814–2819. [Google Scholar] [CrossRef]
  51. Metzner, B.; Elsner, J.; Dobos, G.; Kownatzki, E.; Parlow, F.; Schraufstätter, I.; Norgauer, J. [Ca2+]i-transients and actin polymerization in human neutrophils under stimulation with GROα and complement fragment C5a. Agents Actions 1994, 42, 101–106. [Google Scholar] [CrossRef] [PubMed]
  52. Shyamala, V.; Khoja, H. Interleukin-8 receptors R1 and R2 activate mitogen-activated protein kinases and induce c-fos, independent of Ras and Raf-1 in Chinese hamster ovary cells. Biochemistry 1998, 37, 15918–15924. [Google Scholar] [CrossRef] [PubMed]
  53. Ha, H.; Neamati, N. Pyrimidine-based compounds modulate CXCR2-mediated signaling and receptor turnover. Mol. Pharm. 2014, 11, 2431–2441. [Google Scholar] [CrossRef] [PubMed]
  54. Jin, L.; Li, Z.F.; Wang, D.K.; Sun, M.; Qi, W.; Ma, Q.; Zhang, L.; Chu, C.; Chan, E.Y.M.; Lee, S.S.T.; et al. Molecular and functional characterization of tumor-induced factor (TIF): Hamster homolog of CXCL3 (GROγ) displays tumor suppressive activity. Cytokine 2018, 102, 62–75. [Google Scholar] [CrossRef]
  55. Chan, E.C.; Ren, C.; Xie, Z.; Jude, J.; Barker, T.; Koziol-White, C.A.; Ma, M.; Panettieri, R.A., Jr.; Wu, D.; Rosenberg, H.F.; et al. Regulator of G protein signaling 5 restricts neutrophil chemotaxis and trafficking. J. Biol. Chem. 2018, 293, 12690–12702. [Google Scholar] [CrossRef] [Green Version]
  56. Mueller, S.G.; White, J.R.; Schraw, W.P.; Lam, V.; Richmond, A. Ligand-induced desensitization of the human CXC chemokine receptor-2 is modulated by multiple serine residues in the carboxyl-terminal domain of the receptor. J. Biol. Chem. 1997, 272, 8207–8214. [Google Scholar] [CrossRef] [Green Version]
  57. Schraufstätter, I.U.; Burger, M.; Hoch, R.C.; Oades, Z.G.; Takamori, H. Importance of the carboxy-terminus of the CXCR2 for signal transduction. Biochem. Biophys. Res. Commun. 1998, 244, 243–248. [Google Scholar] [CrossRef]
  58. Raghuwanshi, S.K.; Su, Y.; Singh, V.; Haynes, K.; Richmond, A.; Richardson, R.M. The chemokine receptors CXCR1 and CXCR2 couple to distinct G protein-coupled receptor kinases to mediate and regulate leukocyte functions. J. Immunol. 2012, 189, 2824–2832. [Google Scholar] [CrossRef] [Green Version]
  59. Yang, W.; Wang, D.; Richmond, A. Role of clathrin-mediated endocytosis in CXCR2 sequestration, resensitization, and signal transduction. J. Biol. Chem. 1999, 274, 11328–11333. [Google Scholar] [CrossRef] [Green Version]
  60. Fan, G.H.; Yang, W.; Sai, J.; Richmond, A. Phosphorylation-independent association of CXCR2 with the protein phosphatase 2A core enzyme. J. Biol. Chem. 2001, 276, 16960–16968. [Google Scholar] [CrossRef] [Green Version]
  61. Mueller, S.G.; Schraw, W.P.; Richmond, A. Activation of protein kinase C enhances the phosphorylation of the type B interleukin-8 receptor and stimulates its degradation in non-hematopoietic cells. J. Biol. Chem. 1995, 270, 10439–10448. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Nasser, M.W.; Marjoram, R.J.; Brown, S.L.; Richardson, R.M. Cross-desensitization among CXCR1, CXCR2, and CCR5: Role of protein kinase C-epsilon. J. Immunol. 2005, 174, 6927–6933. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Neel, N.F.; Sai, J.; Ham, A.J.; Sobolik-Delmaire, T.; Mernaugh, R.L.; Richmond, A. IQGAP1 is a novel CXCR2-interacting protein and essential component of the “chemosynapse”. PLoS ONE 2011, 6, e23813. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Raman, D.; Neel, N.F.; Sai, J.; Mernaugh, R.L.; Ham, A.J.; Richmond, A.J. Characterization of chemokine receptor CXCR2 interacting proteins using a proteomics approach to define the CXCR2 “chemosynapse”. Methods Enzymol. 2009, 460, 315–330. [Google Scholar] [PubMed] [Green Version]
  65. Holcomb, J.; Jiang, Y.; Guan, X.; Trescott, L.; Lu, G.; Hou, Y.; Wang, S.; Brunzelle, J.; Sirinupong, N.; Li, C.; et al. Crystal structure of the NHERF1 PDZ2 domain in complex with the chemokine receptor CXCR2 reveals probable modes of PDZ2 dimerization. Biochem. Biophys. Res. Commun. 2014, 448, 169–174. [Google Scholar] [CrossRef]
  66. Wu, Y.; Wang, S.; Farooq, S.M.; Castelvetere, M.P.; Hou, Y.; Gao, J.L.; Navarro, J.V.; Oupicky, D.; Sun, F.; Li, C. A chemokine receptor CXCR2 macromolecular complex regulates neutrophil functions in inflammatory diseases. J. Biol. Chem. 2012, 287, 5744–5755. [Google Scholar] [CrossRef] [Green Version]
  67. Jiang, Y.; Wang, S.; Holcomb, J.; Trescott, L.; Guan, X.; Hou, Y.; Brunzelle, J.; Sirinupong, N.; Li, C.; Yang, Z. Crystallographic analysis of NHERF1-PLCβ3 interaction provides structural basis for CXCR2 signaling in pancreatic cancer. Biochem. Biophys. Res. Commun. 2014, 446, 638–643. [Google Scholar] [CrossRef]
  68. Swart-Mataraza, J.M.; Li, Z.; Sacks, D.B. IQGAP1 is a component of Cdc42 signaling to the cytoskeleton. J. Biol. Chem. 2002, 277, 24753–24763. [Google Scholar] [CrossRef] [Green Version]
  69. Neel, N.F.; Barzik, M.; Raman, D.; Sobolik-Delmaire, T.; Sai, J.; Ham, A.J.; Mernaugh, R.L.; Gertler, F.B.; Richmond, A. VASP is a CXCR2-interacting protein that regulates CXCR2-mediated polarization and chemotaxis. J. Cell Sci. 2009, 122, 1882–1894. [Google Scholar] [CrossRef] [Green Version]
  70. Kamiya, Y.; Mizuno, S.; Komenoi, S.; Sakai, H.; Sakane, F. Activation of conventional and novel protein kinase C isozymes by different diacylglycerol molecular species. Biochem. Biophys. Rep. 2016, 7, 361–366. [Google Scholar] [CrossRef] [Green Version]
  71. Steinberg, S.F. Distinctive activation mechanisms and functions for protein kinase Cδ. Biochem. J. 2004, 384, 449–459. [Google Scholar] [CrossRef] [PubMed]
  72. Sai, J.; Walker, G.; Wikswo, J.; Richmond, A. The IL sequence in the LLKIL motif in CXCR2 is required for full ligand-induced activation of Erk, Akt, and chemotaxis in HL60 cells. J. Biol. Chem. 2006, 281, 35931–35941. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Sai, J.; Fan, G.H.; Wang, D.; Richmond, A. The C-terminal domain LLKIL motif of CXCR2 is required for ligand-mediated polarization of early signals during chemotaxis. J. Cell Sci. 2004, 117, 5489–5496. [Google Scholar] [CrossRef] [Green Version]
  74. Raman, D.; Sai, J.; Neel, N.F.; Chew, C.S.; Richmond, A. LIM and SH3 protein-1 modulates CXCR2-mediated cell migration. PLoS ONE 2010, 5, e10050. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Sai, J.; Raman, D.; Liu, Y.; Wikswo, J.; Richmond, A. Parallel phosphatidylinositol 3-kinase (PI3K)-dependent and Src-dependent pathways lead to CXCL8-mediated Rac2 activation and chemotaxis. J. Biol. Chem. 2008, 283, 26538–26547. [Google Scholar] [CrossRef] [Green Version]
  76. Schraw, W.; Richmond, A. Melanoma growth stimulatory activity signaling through the class II interleukin-8 receptor enhances the tyrosine phosphorylation of Crk-associated substrate, p130, and a 70-kilodalton protein. Biochemistry 1995, 34, 13760–13767. [Google Scholar] [CrossRef]
  77. Gavard, J.; Hou, X.; Qu, Y.; Masedunskas, A.; Martin, D.; Weigert, R.; Li, X.; Gutkind, J.S. A role for a CXCR2/phosphatidylinositol 3-kinase gamma signaling axis in acute and chronic vascular permeability. Mol. Cell. Biol. 2009, 29, 2469–2480. [Google Scholar] [CrossRef] [Green Version]
  78. Yu, B.; Wong, M.M.; Potter, C.M.; Simpson, R.M.; Karamariti, E.; Zhang, Z.; Zeng, L.; Warren, D.; Hu, Y.; Wang, W.; et al. Vascular stem/progenitor cell migration induced by smooth muscle cell-derived chemokine (C-C motif) ligand 2 and chemokine (C-X-C motif) ligand 1 contributes to neointima formation. Stem Cells 2016, 34, 2368–2380. [Google Scholar] [CrossRef] [Green Version]
  79. Wang, D.; Sai, J.; Carter, G.; Sachpatzidis, A.; Lolis, E.; Richmond, A. PAK1 kinase is required for CXCL1-induced chemotaxis. Biochemistry 2002, 41, 7100–7107. [Google Scholar] [CrossRef] [Green Version]
  80. Fan, G.H.; Yang, W.; Sai, J.; Richmond, A. Hsc/Hsp70 interacting protein (hip) associates with CXCR2 and regulates the receptor signaling and trafficking. J. Biol. Chem. 2002, 277, 6590–6597. [Google Scholar] [CrossRef] [Green Version]
  81. Raman, D.; Sai, J.; Hawkins, O.; Richmond, A. Adaptor protein2 (AP2) orchestrates CXCR2-mediated cell migration. Traffic 2014, 15, 451–469. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Montagnac, G.; Meas-Yedid, V.; Irondelle, M.; Castro-Castro, A.; Franco, M.; Shida, T.; Nachury, M.V.; Benmerah, A.; Olivo-Marin, J.C.; Chavrier, P. αTAT1 catalyses microtubule acetylation at clathrin-coated pits. Nature 2013, 502, 567–570. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Jones, S.A.; Moser, B.; Thelen, M. A comparison of post-receptor signal transduction events in Jurkat cells transfected with either IL-8R1 or IL-8R2. Chemokine mediated activation of p42/p44 MAP-kinase (ERK-2). FEBS Lett. 1995, 364, 211–214. [Google Scholar] [PubMed] [Green Version]
  84. Zhao, M.; Wimmer, A.; Trieu, K.; Discipio, R.G.; Schraufstatter, I.U. Arrestin regulates MAPK activation and prevents NADPH oxidase-dependent death of cells expressing CXCR2. J. Biol. Chem. 2004, 279, 49259–49267. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Venkatakrishnan, G.; Salgia, R.; Groopman, J.E. Chemokine receptors CXCR-1/2 activate mitogen-activated protein kinase via the epidermal growth factor receptor in ovarian cancer cells. J. Biol. Chem. 2000, 275, 6868–6875. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Bolitho, C.; Hahn, M.A.; Baxter, R.C.; Marsh, D.J. The chemokine CXCL1 induces proliferation in epithelial ovarian cancer cells by transactivation of the epidermal growth factor receptor. Endocr.-Relat. Cancer 2010, 17, 929–940. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Wang, Z.; Wang, Z.; Li, G.; Wu, H.; Sun, K.; Chen, J.; Feng, Y.; Chen, C.; Cai, S.; Xu, J.; et al. CXCL1 from tumor-associated lymphatic endothelial cells drives gastric cancer cell into lymphatic system via activating integrin β1/FAK/AKT signaling. Cancer Lett. 2017, 385, 28–38. [Google Scholar] [CrossRef]
  88. Artz, A.; Butz, S.; Vestweber, D. GDF-15 inhibits integrin activation and mouse neutrophil recruitment through the ALK-5/TGF-βRII heterodimer. Blood 2016, 128, 529–541. [Google Scholar] [CrossRef] [Green Version]
  89. Antonosante, A.; Brandolini, L.; D’Angelo, M.; Benedetti, E.; Castelli, V.; Maestro, M.D.; Luzzi, S.; Giordano, A.; Cimini, A.; Allegretti, M. Autocrine CXCL8-dependent invasiveness triggers modulation of actin cytoskeletal network and cell dynamics. Aging (Albany NY) 2020, 12, 1928–1951. [Google Scholar] [CrossRef]
  90. Wei, Z.W.; Xia, G.K.; Wu, Y.; Chen, W.; Xiang, Z.; Schwarz, R.E.; Brekken, R.A.; Awasthi, N.; He, Y.L.; Zhang, C.H. CXCL1 promotes tumor growth through VEGF pathway activation and is associated with inferior survival in gastric cancer. Cancer Lett. 2015, 359, 335–343. [Google Scholar] [CrossRef]
  91. Wei, L.; Liu, Y.; Ma, Y.; Ding, C.; Zhang, H.; Lu, Z.; Gu, Z.; Zhu, C. C-X-C chemokine receptor 2 correlates with unfavorable prognosis and facilitates malignant cell activities via activating JAK2/STAT3 pathway in non-small cell lung cancer. Cell Cycle 2019, 18, 3456–3471. [Google Scholar] [CrossRef]
  92. Han, X.; Shi, H.; Sun, Y.; Shang, C.; Luan, T.; Wang, D.; Ba, X.; Zeng, X. CXCR2 expression on granulocyte and macrophage progenitors under tumor conditions contributes to mo-MDSC generation via SAP18/ERK/STAT3. Cell Death Dis. 2019, 10, 598. [Google Scholar] [CrossRef] [Green Version]
  93. Moldobaeva, A.; Baek, A.; Wagner, E.M. MIP-2 causes differential activation of RhoA in mouse aortic versus pulmonary artery endothelial cells. Microvasc. Res. 2008, 75, 53–58. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Fan, G.H.; Yang, W.; Wang, X.J.; Qian, Q.; Richmond, A. Identification of a motif in the carboxyl terminus of CXCR2 that is involved in adaptin 2 binding and receptor internalization. Biochemistry 2001, 40, 791–800. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Leclair, H.M.; Dubois, S.M.; Azzi, S.; Dwyer, J.; Bidère, N.; Gavard, J. Control of CXCR2 activity through its ubiquitination on K327 residue. BMC Cell Biol. 2014, 15, 38. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Rose, J.J.; Foley, J.F.; Murphy, P.M.; Venkatesan, S. On the mechanism and significance of ligand-induced internalization of human neutrophil chemokine receptors CXCR1 and CXCR2. J. Biol. Chem. 2004, 279, 24372–24386. [Google Scholar] [CrossRef] [Green Version]
  97. Richardson, R.M.; Pridgen, B.C.; Haribabu, B.; Ali, H.; Snyderman, R. Differential cross-regulation of the human chemokine receptors CXCR1 and CXCR2. Evidence for time-dependent signal generation. J. Biol. Chem. 1998, 273, 23830–23836. [Google Scholar] [CrossRef] [Green Version]
  98. Wandinger-Ness, A.; Zerial, M. Rab proteins and the compartmentalization of the endosomal system. Cold Spring Harb. Perspect. Biol. 2014, 6, a022616. [Google Scholar] [CrossRef]
  99. Fan, G.H.; Lapierre, L.A.; Goldenring, J.R.; Richmond, A. Differential regulation of CXCR2 trafficking by Rab GTPases. Blood 2003, 101, 2115–2124. [Google Scholar] [CrossRef] [Green Version]
  100. Neel, N.F.; Lapierre, L.A.; Goldenring, J.R.; Richmond, A. RhoB plays an essential role in CXCR2 sorting decisions. J. Cell Sci. 2007, 120, 1559–1571. [Google Scholar] [CrossRef] [Green Version]
  101. Tang, W.; Li, Z.; Li, X.; Huo, Z. High CXCR2 expression predicts poor prognosis in adult patients with acute myeloid leukemia. Ther. Adv. Hematol. 2020, 11, 2040620720958586. [Google Scholar] [CrossRef] [PubMed]
  102. Xu, H.; Lin, F.; Wang, Z.; Yang, L.; Meng, J.; Ou, Z.; Shao, Z.; Di, G.; Yang, G. CXCR2 promotes breast cancer metastasis and chemoresistance via suppression of AKT1 and activation of COX2. Cancer Lett. 2018, 412, 69–80. [Google Scholar] [CrossRef]
  103. Zhao, J.; Ou, B.; Feng, H.; Wang, P.; Yin, S.; Zhu, C.; Wang, S.; Chen, C.; Zheng, M.; Zong, Y.; et al. Overexpression of CXCR2 predicts poor prognosis in patients with colorectal cancer. Oncotarget 2017, 8, 28442–28454. [Google Scholar] [CrossRef] [Green Version]
  104. Sui, P.; Hu, P.; Zhang, T.; Zhang, X.; Liu, Q.; Du, J. High expression of CXCR-2 correlates with lymph node metastasis and predicts unfavorable prognosis in resected esophageal carcinoma. Med. Oncol. 2014, 31, 809. [Google Scholar] [CrossRef]
  105. Wang, Z.; Liu, H.; Shen, Z.; Wang, X.; Zhang, H.; Qin, J.; Xu, J.; Sun, Y.; Qin, X. The prognostic value of CXC-chemokine receptor 2 (CXCR2) in gastric cancer patients. BMC Cancer 2015, 15, 766. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Kasashima, H.; Yashiro, M.; Nakamae, H.; Masuda, G.; Kinoshita, H.; Morisaki, T.; Fukuoka, T.; Hasegawa, T.; Nakane, T.; Hino, M.; et al. Clinicopathologic significance of the CXCL1-CXCR2 axis in the tumor microenvironment of gastric carcinoma. PLoS ONE 2017, 12, e0178635. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Sueoka, H.; Hirano, T.; Uda, Y.; Iimuro, Y.; Yamanaka, J.; Fujimoto, J. Blockage of CXCR2 suppresses tumor growth of intrahepatic cholangiocellular carcinoma. Surgery 2014, 155, 640–649. [Google Scholar] [CrossRef]
  108. Han, L.; Jiang, B.; Wu, H.; Wang, X.; Tang, X.; Huang, J.; Zhu, J. High expression of CXCR2 is associated with tumorigenesis, progression, and prognosis of laryngeal squamous cell carcinoma. Med. Oncol. 2012, 29, 2466–2472. [Google Scholar] [CrossRef]
  109. Saintigny, P.; Massarelli, E.; Lin, S.; Ahn, Y.H.; Chen, Y.; Goswami, S.; Erez, B.; O’Reilly, M.S.; Liu, D.; Lee, J.J.; et al. CXCR2 expression in tumor cells is a poor prognostic factor and promotes invasion and metastasis in lung adenocarcinoma. Cancer Res. 2013, 73, 571–582. [Google Scholar] [CrossRef] [Green Version]
  110. Yang, G.; Rosen, D.G.; Liu, G.; Yang, F.; Guo, X.; Xiao, X.; Xue, F.; Mercado-Uribe, I.; Huang, J.; Lin, S.H.; et al. CXCR2 promotes ovarian cancer growth through dysregulated cell cycle, diminished apoptosis, and enhanced angiogenesis. Clin. Cancer Res. 2010, 16, 3875–3886. [Google Scholar] [CrossRef] [Green Version]
  111. Henriques, T.B.; dos Santos, D.Z.; dos Santos Guimarães, I.; Tessarollo, N.G.; Lyra-Junior, P.C.M.; Mesquita, P.; Pádua, D.; Amaral, A.L.; Cavadas, B.; Pereira, L.; et al. Inhibition of CXCR2 plays a pivotal role in re-sensitizing ovarian cancer to cisplatin treatment. Aging (Albany NY) 2021, 13, 13405–13420. [Google Scholar] [CrossRef] [PubMed]
  112. Steele, C.W.; Karim, S.A.; Leach, J.D.G.; Bailey, P.; Upstill-Goddard, R.; Rishi, L.; Foth, M.; Bryson, S.; McDaid, K.; Wilson, Z.; et al. CXCR2 inhibition profoundly suppresses metastases and augments immunotherapy in pancreatic ductal adenocarcinoma. Cancer Cell 2016, 29, 832–845. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Boissière-Michot, F.; Jacot, W.; Massol, O.; Mollevi, C.; Lazennec, G. CXCR2 levels correlate with immune infiltration and a better prognosis of triple-negative breast cancers. Cancers Basel 2021, 13, 2328. [Google Scholar] [CrossRef] [PubMed]
  114. Oladipo, O.; Conlon, S.; O’Grady, A.; Purcell, C.; Wilson, C.; Maxwell, P.J.; Johnston, P.G.; Stevenson, M.; Kay, E.W.; Wilson, R.H.; et al. The expression and prognostic impact of CXC-chemokines in stage II and III colorectal cancer epithelial and stromal tissue. Br. J. Cancer 2011, 104, 480–487. [Google Scholar] [CrossRef] [PubMed]
  115. Wu, Z.; Zhang, Y.; Chen, X.; Tan, W.; He, L.; Peng, L. Characterization of the prognostic values of the CXCR1-7 in clear cell renal cell carcinoma (ccRCC) microenvironment. Front. Mol. Biosci. 2020, 7, 601206. [Google Scholar] [CrossRef] [PubMed]
  116. Yu, C.; Zhang, Y. Characterization of the prognostic values of CXCR family in gastric cancer. Cytokine 2019, 123, 154785. [Google Scholar] [CrossRef]
  117. Maeda, S.; Kuboki, S.; Nojima, H.; Shimizu, H.; Yoshitomi, H.; Furukawa, K.; Miyazaki, M.; Ohtsuka, M. Duffy antigen receptor for chemokines (DARC) expressing in cancer cells inhibits tumor progression by suppressing CXCR2 signaling in human pancreatic ductal adenocarcinoma. Cytokine 2017, 95, 12–21. [Google Scholar] [CrossRef]
  118. Timaxian, C.; Vogel, C.F.A.; Orcel, C.; Vetter, D.; Durochat, C.; Chinal, C.; NGuyen, P.; Aknin, M.L.; Mercier-Nomé, F.; Davy, M.; et al. Pivotal role for Cxcr2 in regulating tumor-associated neutrophil in breast cancer. Cancers 2021, 13, 2584. [Google Scholar] [CrossRef]
  119. Yang, J.; Richmond, A. Constitutive IκB kinase activity correlates with nuclear factor-κB activation in human melanoma cells. Cancer Res. 2001, 61, 4901–4909. [Google Scholar]
  120. Wang, D.; Richmond, A. Nuclear factor-κB activation by the CXC chemokine melanoma growth-stimulatory activity/growth-regulated protein involves the MEKK1/p38 mitogen-activated protein kinase pathway. J. Biol. Chem. 2001, 276, 3650–3659. [Google Scholar] [CrossRef] [Green Version]
  121. Dong, Y.L.; Kabir, S.M.; Lee, E.S.; Son, D.S. CXCR2-driven ovarian cancer progression involves upregulation of proinflammatory chemokines by potentiating NF-κB activation via EGFR-transactivated Akt signaling. PLoS ONE 2013, 8, e83789. [Google Scholar] [CrossRef] [PubMed]
  122. Yung, M.M.; Tang, H.W.; Cai, P.C.; Leung, T.H.; Ngu, S.F.; Chan, K.K.; Xu, D.; Yang, H.; Ngan, H.Y.; Chan, D.W. GRO-α and IL-8 enhance ovarian cancer metastatic potential via the CXCR2-mediated TAK1/NFκB signaling cascade. Theranostics 2018, 8, 1270–1285. [Google Scholar] [CrossRef] [PubMed]
  123. Kveiborg, M.; Instrell, R.; Rowlands, C.; Howell, M.; Parker, P.J. PKCα and PKCδ regulate ADAM17-mediated ectodomain shedding of heparin binding-EGF through separate pathways. PLoS ONE 2011, 6, e17168. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Lisi, S.; Sisto, M.; Lofrumento, D.D.; D’Amore, M.; de Lucro, R.; Ribatti, D. GRO-α/CXCR2 system and ADAM17 correlated expression in Sjögren’s syndrome. Inflammation 2013, 36, 759–766. [Google Scholar] [CrossRef]
  125. Schraufstatter, I.U.; Trieu, K.; Zhao, M.; Rose, D.M.; Terkeltaub, R.A.; Burger, M. IL-8-mediated cell migration in endothelial cells depends on cathepsin B activity and transactivation of the epidermal growth factor receptor. J. Immunol. 2003, 171, 6714–6722. [Google Scholar] [CrossRef] [Green Version]
  126. Wang, B.; Khachigian, L.M.; Esau, L.; Birrer, M.J.; Zhao, X.; Parker, M.I.; Hendricks, D.T. A key role for early growth response-1 and nuclear factor-κB in mediating and maintaining GRO/CXCR2 proliferative signaling in esophageal cancer. Mol. Cancer Res. 2009, 7, 755–764. [Google Scholar] [CrossRef] [Green Version]
  127. Ignacio, R.M.C.; Dong, Y.L.; Kabir, S.M.; Choi, H.; Lee, E.S.; Wilson, A.J.; Beeghly-Fadiel, A.; Whalen, M.M.; Son, D.S. CXCR2 is a negative regulator of p21 in p53-dependent and independent manner via Akt-mediated Mdm2 in ovarian cancer. Oncotarget 2018, 9, 9751–9765. [Google Scholar] [CrossRef] [Green Version]
  128. Midgley, C.A.; Desterro, J.M.; Saville, M.K.; Howard, S.; Sparks, A.; Hay, R.T.; Lane, D.P. An N-terminal p14ARF peptide blocks Mdm2-dependent ubiquitination in vitro and can activate p53 in vivo. Oncogene 2000, 19, 2312–2323. [Google Scholar] [CrossRef] [Green Version]
  129. Acosta, J.C.; O’Loghlen, A.; Banito, A.; Guijarro, M.V.; Augert, A.; Raguz, S.; Fumagalli, M.; da Costa, M.; Brown, C.; Popov, N.; et al. Chemokine signaling via the CXCR2 receptor reinforces senescence. Cell 2008, 133, 1006–1018. [Google Scholar] [CrossRef] [Green Version]
  130. Guo, H.; Liu, Z.; Xu, B.; Hu, H.; Wei, Z.; Liu, Q.; Zhang, X.; Ding, X.; Wang, Y.; Zhao, M.; et al. Chemokine receptor CXCR2 is transactivated by p53 and induces p38-mediated cellular senescence in response to DNA damage. Aging Cell 2013, 12, 1110–1121. [Google Scholar] [CrossRef] [Green Version]
  131. O’Connor, P.M.; Jackman, J.; Bae, I.; Myers, T.G.; Fan, S.; Mutoh, M.; Scudiero, D.A.; Monks, A.; Sausville, E.A.; Weinstein, J.N.; et al. Characterization of the p53 tumor suppressor pathway in cell lines of the National Cancer Institute anticancer drug screen and correlations with the growth-inhibitory potency of 123 anticancer agents. Cancer Res. 1997, 57, 4285–4300. [Google Scholar] [PubMed]
  132. Yang, G.; Rosen, D.G.; Zhang, Z.; Bast, R.C., Jr.; Mills, G.B.; Colacino, J.A.; Mercado-Uribe, I.; Liu, J. The chemokine growth-regulated oncogene 1 (Gro-1) links RAS signaling to the senescence of stromal fibroblasts and ovarian tumorigenesis. Proc. Natl. Acad. Sci. USA 2006, 103, 16472–16477. [Google Scholar] [CrossRef] [Green Version]
  133. Sano, M.; Ijichi, H.; Takahashi, R.; Miyabayashi, K.; Fujiwara, H.; Yamada, T.; Kato, H.; Nakatsuka, T.; Tanaka, Y.; Tateishi, K.; et al. Blocking CXCLs-CXCR2 axis in tumor-stromal interactions contributes to survival in a mouse model of pancreatic ductal adenocarcinoma through reduced cell invasion/migration and a shift of immune-inflammatory microenvironment. Oncogenesis 2019, 8, 8. [Google Scholar] [CrossRef] [PubMed]
  134. Awaji, M.; Saxena, S.; Wu, L.; Prajapati, D.R.; Purohit, A.; Varney, M.L.; Kumar, S.; Rachagani, S.; Ly, Q.P.; Jain, M.; et al. CXCR2 signaling promotes secretory cancer-associated fibroblasts in pancreatic ductal adenocarcinoma. FASEB J. 2020, 34, 9405–9418. [Google Scholar] [CrossRef] [PubMed]
  135. Chen, L.; Pan, X.W.; Huang, H.; Gao, Y.; Yang, Q.W.; Wang, L.H.; Cui, X.G.; Xu, D.F. Epithelial-mesenchymal transition induced by GRO-α-CXCR2 promotes bladder cancer recurrence after intravesical chemotherapy. Oncotarget 2017, 8, 45274–45285. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Chao, C.C.; Lee, C.W.; Chang, T.M.; Chen, P.C.; Liu, J.F. CXCL1/CXCR2 paracrine axis contributes to lung metastasis in osteosarcoma. Cancers Basel 2020, 12, 459. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Sharma, B.; Nannuru, K.C.; Saxena, S.; Varney, M.L.; Singh, R.K. CXCR2: A novel mediator of mammary tumor bone metastasis. Int. J. Mol. Sci. 2019, 20, 1237. [Google Scholar] [CrossRef] [Green Version]
  138. Grassi, F.; Piacentini, A.; Cristino, S.; Toneguzzi, S.; Cavallo, C.; Facchini, A.; Lisignoli, G. Human osteoclasts express different CXC chemokines depending on cell culture substrate: Molecular and immunocytochemical evidence of high levels of CXCL10 and CXCL12. Histochem. Cell Biol. 2003, 120, 391–400. [Google Scholar] [CrossRef]
  139. Onan, D.; Allan, E.H.; Quinn, J.M.; Gooi, J.H.; Pompolo, S.; Sims, N.A.; Gillespie, M.T.; Martin, T.J. The chemokine Cxcl1 is a novel target gene of parathyroid hormone (PTH)/PTH-related protein in committed osteoblasts. Endocrinology 2009, 150, 2244–2253. [Google Scholar] [CrossRef] [Green Version]
  140. Xu, J.; Zhu, M.D.; Zhang, X.; Tian, H.; Zhang, J.H.; Wu, X.B.; Gao, Y.J. NFκB-mediated CXCL1 production in spinal cord astrocytes contributes to the maintenance of bone cancer pain in mice. J. Neuroinflamm. 2014, 11, 38. [Google Scholar] [CrossRef] [Green Version]
  141. Ni, H.; Wang, Y.; An, K.; Liu, Q.; Xu, L.; Zhu, C.; Deng, H.; He, Q.; Wang, T.; Xu, M.; et al. Crosstalk between NFκB-dependent astrocytic CXCL1 and neuron CXCR2 plays a role in descending pain facilitation. J. Neuroinflamm. 2019, 16, 1. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Ni, H.; Xu, M.; Xie, K.; Fei, Y.; Deng, H.; He, Q.; Wang, T.; Liu, S.; Zhu, J.; Xu, L.; et al. Liquiritin alleviates pain through inhibiting CXCL1/CXCR2 signaling pathway in bone cancer pain rat. Front. Pharmacol. 2020, 11, 436. [Google Scholar] [CrossRef]
  143. Addison, C.L.; Daniel, T.O.; Burdick, M.D.; Liu, H.; Ehlert, J.E.; Xue, Y.Y.; Buechi, L.; Walz, A.; Richmond, A.; Strieter, R.M. The CXC chemokine receptor 2, CXCR2, is the putative receptor for ELR+ CXC chemokine-induced angiogenic activity. J. Immunol. 2000, 165, 5269–5277. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Mestas, J.; Burdick, M.D.; Reckamp, K.; Pantuck, A.; Figlin, R.A.; Strieter, R.M. The role of CXCR2/CXCR2 ligand biological axis in renal cell carcinoma. J. Immunol. 2005, 175, 5351–5357. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Miyake, M.; Goodison, S.; Urquidi, V.; Gomes Giacoia, E.; Rosser, C.J. Expression of CXCL1 in human endothelial cells induces angiogenesis through the CXCR2 receptor and the ERK1/2 and EGF pathways. Lab. Investig. 2013, 93, 768–778. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Keane, M.P.; Belperio, J.A.; Xue, Y.Y.; Burdick, M.D.; Strieter, R.M. Depletion of CXCR2 inhibits tumor growth and angiogenesis in a murine model of lung cancer. J. Immunol. 2004, 172, 2853–2860. [Google Scholar] [CrossRef] [Green Version]
  147. Xu, J.; Zhang, C.; He, Y.; Wu, H.; Wang, Z.; Song, W.; Li, W.; He, W.; Cai, S.; Zhan, W. Lymphatic endothelial cell-secreted CXCL1 stimulates lymphangiogenesis and metastasis of gastric cancer. Int. J. Cancer 2012, 130, 787–797. [Google Scholar] [CrossRef]
  148. Karl, E.; Warner, K.; Zeitlin, B.; Kaneko, T.; Wurtzel, L.; Jin, T.; Chang, J.; Wang, S.; Wang, C.Y.; Strieter, R.M.; et al. Bcl-2 acts in a proangiogenic signaling pathway through nuclear factor-κB and CXC chemokines. Cancer Res. 2005, 65, 5063–5069. [Google Scholar] [CrossRef] [Green Version]
  149. Kaneko, T.; Zhang, Z.; Mantellini, M.G.; Karl, E.; Zeitlin, B.; Verhaegen, M.; Soengas, M.S.; Lingen, M.; Strieter, R.M.; Nunez, G.; et al. Bcl-2 orchestrates a cross-talk between endothelial and tumor cells that promotes tumor growth. Cancer Res. 2007, 67, 9685–9693. [Google Scholar] [CrossRef] [Green Version]
  150. Yu, M.; Berk, R.; Kosir, M.A. CXCL7-mediated stimulation of lymphangiogenic factors VEGF-C, VEGF-D in human breast cancer cells. J. Oncol. 2010, 2010, 939407. [Google Scholar] [CrossRef] [Green Version]
  151. Huang, Z.; Zhang, M.; Chen, G.; Wang, W.; Zhang, P.; Yue, Y.; Guan, Z.; Wang, X.; Fan, J. Bladder cancer cells interact with vascular endothelial cells triggering EGFR signals to promote tumor progression. Int. J. Oncol. 2019, 54, 1555–1566. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Li, A.; Cheng, X.J.; Moro, A.; Singh, R.K.; Hines, O.J.; Eibl, G. CXCR2-dependent endothelial progenitor cell mobilization in pancreatic cancer growth. Transl. Oncol. 2011, 4, 20–28. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Sharma, B.; Nannuru, K.C.; Varney, M.L.; Singh, R.K. Host Cxcr2-dependent regulation of mammary tumor growth and metastasis. Clin. Exp. Metastasis 2015, 32, 65–72. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Di Mitri, D.; Mirenda, M.; Vasilevska, J.; Calcinotto, A.; Delaleu, N.; Revandkar, A.; Gil, V.; Boysen, G.; Losa, M.; Mosole, S.; et al. Re-education of tumor-associated macrophages by CXCR2 blockade drives senescence and tumor inhibition in advanced prostate cancer. Cell Rep. 2019, 28, 2156–2168.e5. [Google Scholar] [CrossRef] [Green Version]
  155. Zhang, M.; Huang, L.; Ding, G.; Huang, H.; Cao, G.; Sun, X.; Lou, N.; Wei, Q.; Shen, T.; Xu, X.; et al. Interferon gamma inhibits CXCL8-CXCR2 axis mediated tumor-associated macrophages tumor trafficking and enhances anti-PD1 efficacy in pancreatic cancer. J. Immunother. Cancer 2020, 8, e000308. [Google Scholar] [CrossRef] [Green Version]
  156. Geissmann, F.; Jung, S.; Littman, D.R. Blood monocytes consist of two principal subsets with distinct migratory properties. Immunity 2003, 19, 71–82. [Google Scholar] [CrossRef] [Green Version]
  157. Chao, T.; Furth, E.E.; Vonderheide, R.H. CXCR2-dependent accumulation of tumor-associated neutrophils regulates T-cell immunity in pancreatic ductal adenocarcinoma. Cancer Immunol. Res. 2016, 4, 968–982. [Google Scholar] [CrossRef] [Green Version]
  158. Ohms, M.; Möller, S.; Laskay, T. An attempt to polarize human neutrophils toward N1 and N2 phenotypes in vitro. Front. Immunol 2020, 11, 532. [Google Scholar] [CrossRef]
  159. Yang, L.; Huang, J.; Ren, X.; Gorska, A.E.; Chytil, A.; Aakre, M.; Carbone, D.P.; Matrisian, L.M.; Richmond, A.; Lin, P.C.; et al. Abrogation of TGFβ signaling in mammary carcinomas recruits Gr-1+CD11b+ myeloid cells that promote metastasis. Cancer Cell 2008, 13, 23–35. [Google Scholar] [CrossRef] [Green Version]
  160. Acharyya, S.; Oskarsson, T.; Vanharanta, S.; Malladi, S.; Kim, J.; Morris, P.G.; Manova-Todorova, K.; Leversha, M.; Hogg, N.; Seshan, V.E.; et al. A CXCL1 paracrine network links cancer chemoresistance and metastasis. Cell 2012, 150, 165–178. [Google Scholar] [CrossRef] [Green Version]
  161. Katoh, H.; Wang, D.; Daikoku, T.; Sun, H.; Dey, S.K.; Dubois, R.N. CXCR2-expressing myeloid-derived suppressor cells are essential to promote colitis-associated tumorigenesis. Cancer Cell 2013, 24, 631–644. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Lv, M.; Xu, Y.; Tang, R.; Ren, J.; Shen, S.; Chen, Y.; Liu, B.; Hou, Y.; Wang, T. MiR141-CXCL1-CXCR2 signaling-induced Treg recruitment regulates metastases and survival of non-small cell lung cancer. Mol. Cancer Ther. 2014, 13, 3152–3162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Kasashima, H.; Yashiro, M.; Nakamae, H.; Kitayama, K.; Masuda, G.; Kinoshita, H.; Fukuoka, T.; Hasegawa, T.; Nakane, T.; Hino, M.; et al. CXCL1-chemokine (C-X-C motif) receptor 2 signaling stimulates the recruitment of bone marrow-derived mesenchymal cells into diffuse-type gastric cancer stroma. Am. J. Pathol. 2016, 186, 3028–3039. [Google Scholar] [CrossRef] [PubMed]
  164. Shi, H.; Han, X.; Sun, Y.; Shang, C.; Wei, M.; Ba, X.; Zeng, X. Chemokine (C-X-C motif) ligand 1 and CXCL2 produced by tumor promote the generation of monocytic myeloid-derived suppressor cells. Cancer Sci. 2018, 109, 3826–3839. [Google Scholar] [CrossRef] [PubMed]
  165. Ijichi, H.; Chytil, A.; Gorska, A.E.; Aakre, M.E.; Bierie, B.; Tada, M.; Mohri, D.; Miyabayashi, K.; Asaoka, Y.; Maeda, S.; et al. Inhibiting Cxcr2 disrupts tumor-stromal interactions and improves survival in a mouse model of pancreatic ductal adenocarcinoma. J. Clin. Investig. 2011, 121, 4106–4117. [Google Scholar] [CrossRef]
  166. Ning, Y.; Labonte, M.J.; Zhang, W.; Bohanes, P.O.; Gerger, A.; Yang, D.; Benhaim, L.; Paez, D.; Rosenberg, D.O.; Venkata, K.C.N.; et al. The CXCR2 antagonist, SCH-527123, shows antitumor activity and sensitizes cells to oxaliplatin in preclinical colon cancer models. Mol. Cancer Ther. 2012, 11, 1353–1364. [Google Scholar] [CrossRef] [Green Version]
  167. Angara, K.; Borin, T.F.; Rashid, M.H.; Lebedyeva, I.; Ara, R.; Lin, P.C.; Iskander, A.; Bollag, R.J.; Achyut, B.R.; Arbab, A.S. CXCR2-expressing tumor cells drive vascular mimicry in antiangiogenic therapy-resistant glioblastoma. Neoplasia 2018, 20, 1070–1082. [Google Scholar] [CrossRef]
  168. Cheng, Y.; Mo, F.; Li, Q.; Han, X.; Shi, H.; Chen, S.; Wei, Y.; Wei, X. Targeting CXCR2 inhibits the progression of lung cancer and promotes therapeutic effect of cisplatin. Mol. Cancer 2021, 20, 62. [Google Scholar] [CrossRef]
  169. Devapatla, B.; Sharma, A.; Woo, S. CXCR2 inhibition combined with sorafenib improved antitumor and antiangiogenic response in preclinical models of ovarian cancer. PLoS ONE 2015, 10, e0139237. [Google Scholar]
  170. Sapoznik, S.; Ortenberg, R.; Galore-Haskel, G.; Kozlovski, S.; Levy, D.; Avivi, C.; Barshack, I.; Cohen, C.J.; Besser, M.J.; Schachter, J.; et al. CXCR1 as a novel target for directing reactive T cells toward melanoma: Implications for adoptive cell transfer immunotherapy. Cancer Immunol. Immunother. 2012, 61, 1833–1847. [Google Scholar] [CrossRef]
  171. Peng, W.; Ye, Y.; Rabinovich, B.A.; Liu, C.; Lou, Y.; Zhang, M.; Whittington, M.; Yang, Y.; Overwijk, W.W.; Lizée, G.; et al. Transduction of tumor-specific T cells with CXCR2 chemokine receptor improves migration to tumor and antitumor immune responses. Clin. Cancer Res. 2010, 16, 5458–5468. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Kremer, V.; Ligtenberg, M.A.; Zendehdel, R.; Seitz, C.; Duivenvoorden, A.; Wennerberg, E.; Colón, E.; Scherman-Plogell, A.H.; Lundqvist, A. Genetic engineering of human NK cells to express CXCR2 improves migration to renal cell carcinoma. J. Immunother. Cancer 2017, 5, 73. [Google Scholar] [CrossRef] [PubMed]
  173. Liu, G.; Rui, W.; Zheng, H.; Huang, D.; Yu, F.; Zhang, Y.; Dong, J.; Zhao, X.; Lin, X. CXCR2-modified CAR-T cells have enhanced trafficking ability that improves treatment of hepatocellular carcinoma. Eur. J. Immunol. 2020, 50, 712–724. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Regulation of CXCR2 by miRNAs. CXCR2 expression is regulated by miRNAs. Diseases such as cancer, ischemic stroke, atrial fibrillation, and neuropathic pain are associated with a decrease in the level of some miRNAs which results in a pathological increase in CXCR2 expression. Fungus Cordyceps militaris contains miRNAs that decrease the expression of CXCR2. The increased expression of CXCR2 can be physiological—it is often related to a decrease in the levels of certain miRNAs, resulting in an increase in CXCR2 expression and thus normal cell and tissue function. This is exemplified by osteoclast precursors, where increased CXCR2 expression leads to the differentiation of these cells into osteoclasts. This mechanism has been observed in root resorption associated with orthodontic tooth movement. Furthermore, increased CXCR2 expression on endothelial cells increases their proliferation which is associated with the preservation of endothelial integrity. CXCR2 activity is the effect of miRNAs on GRK2 expression. This mechanism is significant in cell infection by KSHV.
Figure 1. Regulation of CXCR2 by miRNAs. CXCR2 expression is regulated by miRNAs. Diseases such as cancer, ischemic stroke, atrial fibrillation, and neuropathic pain are associated with a decrease in the level of some miRNAs which results in a pathological increase in CXCR2 expression. Fungus Cordyceps militaris contains miRNAs that decrease the expression of CXCR2. The increased expression of CXCR2 can be physiological—it is often related to a decrease in the levels of certain miRNAs, resulting in an increase in CXCR2 expression and thus normal cell and tissue function. This is exemplified by osteoclast precursors, where increased CXCR2 expression leads to the differentiation of these cells into osteoclasts. This mechanism has been observed in root resorption associated with orthodontic tooth movement. Furthermore, increased CXCR2 expression on endothelial cells increases their proliferation which is associated with the preservation of endothelial integrity. CXCR2 activity is the effect of miRNAs on GRK2 expression. This mechanism is significant in cell infection by KSHV.
Ijms 23 02168 g001
Figure 2. Consequences of G-protein activation by CXCR2. CXCR2 belongs to the seven transmembrane G-protein-coupled receptor family. CXCR2 is coupled with Gαi—upon CXCR2 activation, this G protein dissociates into heterodimer of Gβγ and Gαi subunits. Gαi subunit inhibits adenylyl cyclase activity, while Gβγ activates PI3K and PLC-β. Activation of PLC-β results in the production of DAG and IP3. IP3 induces the mobilization of Ca2+.
Figure 2. Consequences of G-protein activation by CXCR2. CXCR2 belongs to the seven transmembrane G-protein-coupled receptor family. CXCR2 is coupled with Gαi—upon CXCR2 activation, this G protein dissociates into heterodimer of Gβγ and Gαi subunits. Gαi subunit inhibits adenylyl cyclase activity, while Gβγ activates PI3K and PLC-β. Activation of PLC-β results in the production of DAG and IP3. IP3 induces the mobilization of Ca2+.
Ijms 23 02168 g002
Figure 3. Importance of phosphorylation in the function of CXCR2 receptor. As a consequence of activation by its ligand, CXCR2 is phosphorylated at the C-terminus by GRK2 and GRK6. This leads to desensitization and subsequent internalization of CXCR2. In endosomes CXCR2 is dephosphorylated by PP2A, which leads to its resensitization and being recycled back to the cell membrane.
Figure 3. Importance of phosphorylation in the function of CXCR2 receptor. As a consequence of activation by its ligand, CXCR2 is phosphorylated at the C-terminus by GRK2 and GRK6. This leads to desensitization and subsequent internalization of CXCR2. In endosomes CXCR2 is dephosphorylated by PP2A, which leads to its resensitization and being recycled back to the cell membrane.
Ijms 23 02168 g003
Figure 4. CXCR2 as a chemosynapse. Activated and inactivated CXCR2 bind various proteins which play an important role in signal transduction from this receptor. One such is NHERF1, which localizes PLC-β close to CXCR2—this allows the receptor to efficiently activate PLC-β. The activation of CXCR1 is followed by the dissociation of gelsolin and plastin, proteins which regulate the formation of F-actin close to the activated CXCR2 receptor. Moreover, upon CXCR2 activation, F-actin begins to bind to CXCR2 via VASP and IQGAP1—these processes lead to cell migration as a consequence of CXCR2 activation. Furthermore, CXCR2 activation mediated by LASP-1 is followed by the activation of various proteins including Cdc42, ERK MAPK, p130CAS, and Src.
Figure 4. CXCR2 as a chemosynapse. Activated and inactivated CXCR2 bind various proteins which play an important role in signal transduction from this receptor. One such is NHERF1, which localizes PLC-β close to CXCR2—this allows the receptor to efficiently activate PLC-β. The activation of CXCR1 is followed by the dissociation of gelsolin and plastin, proteins which regulate the formation of F-actin close to the activated CXCR2 receptor. Moreover, upon CXCR2 activation, F-actin begins to bind to CXCR2 via VASP and IQGAP1—these processes lead to cell migration as a consequence of CXCR2 activation. Furthermore, CXCR2 activation mediated by LASP-1 is followed by the activation of various proteins including Cdc42, ERK MAPK, p130CAS, and Src.
Ijms 23 02168 g004
Figure 5. CXCR2 trafficking. Upon activation, CXCR2 is phosphorylated, and Hip and AP2 attach to the LLKIL motif at the C-terminus, which is also ubiquitinated. Clathrin and β-arrestin bind to AP2, resulting in the formation of a clathrin-coated pit. As a result, CXCR2 undergoes Rab5-dependent internalization into early endosomes. Subsequently, CXCR2 undergoes RhoB-dependent sorting decisions. CXCR2 can transition to recycling endosomes depending on Rab11a and then is transported back into the cell membrane. CXCR2 can also move to late endosomes and eventually to lysosomes in a Rab7-dependent process, which leads to the degradation of CXCR2.
Figure 5. CXCR2 trafficking. Upon activation, CXCR2 is phosphorylated, and Hip and AP2 attach to the LLKIL motif at the C-terminus, which is also ubiquitinated. Clathrin and β-arrestin bind to AP2, resulting in the formation of a clathrin-coated pit. As a result, CXCR2 undergoes Rab5-dependent internalization into early endosomes. Subsequently, CXCR2 undergoes RhoB-dependent sorting decisions. CXCR2 can transition to recycling endosomes depending on Rab11a and then is transported back into the cell membrane. CXCR2 can also move to late endosomes and eventually to lysosomes in a Rab7-dependent process, which leads to the degradation of CXCR2.
Ijms 23 02168 g005
Table 1. Effect of high CXCR2 expression in tumor on prognosis for patients with different cancers.
Table 1. Effect of high CXCR2 expression in tumor on prognosis for patients with different cancers.
Type of CancerImpact on PrognosisNumber of Patients Selected for TestingNotesSource
Breast cancer: invasive ductal breast cancerWorse
prognosis
225disease-free survival and overall survival[102]
Breast cancer: triple-negative breast cancerBetter
prognosis
290recurrence-free survival and overall survival[113]
Colorectal cancerNot associated with
patient prognosis
254recurrence-free survival and overall survival[114]
Colorectal cancerWorse
prognosis
134overall survival and disease-free survival[103]
Esophageal cancerWorse
prognosis
95overall survival[104]
Gastric cancerBetter
prognosis
593from TCGA database, overall survival[116]
Gastric cancerWorse
prognosis
115cumulative survival[16]
Gastric cancerWorse
prognosis
269overall survival[106]
Gastric cancerWorse
prognosis
357overall survival[105]
Gastric cancerWorse
prognosis
155overall survival[17]
Intrahepatic cholangiocellular carcinomaWorse
prognosis
34overall survival[107]
Kidney cancer: clear cell renal cell carcinomaBetter
prognosis
530from TCGA database, overall survival[115]
Laryngeal squamous cell carcinomaWorse
prognosis
109overall survival[108]
Leukemia: acute myeloid leukemiaWorse
prognosis
45recurrence-free survival and overall survival[101]
Lung cancer: Lung adenocarcinomaWorse
prognosis
173recurrence-free survival and overall survival[109]
Lung cancer: non-small cell lung cancerWorse
prognosis
340disease-free survival and overall survival[91]
Ovarian cancerWorse
prognosis
240disease-free survival[110]
Ovarian cancerWorse
prognosis
370from TCGA database, overall survival[111]
Pancreatic ductal adenocarcinomaNot associated with
patient prognosis
102recurrence-free survival and overall survival[117]
Pancreatic ductal adenocarcinomaWorse
prognosis
44overall survival[112]
Red background—increased CXCR2 expression in the tumor is associated with a worse prognosis; blue background—increased CXCR2 expression in the tumor is associated with a better prognosis; gray background—CXCR2 expression in the tumor is not associated with patient prognosis.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Korbecki, J.; Kupnicka, P.; Chlubek, M.; Gorący, J.; Gutowska, I.; Baranowska-Bosiacka, I. CXCR2 Receptor: Regulation of Expression, Signal Transduction, and Involvement in Cancer. Int. J. Mol. Sci. 2022, 23, 2168. https://doi.org/10.3390/ijms23042168

AMA Style

Korbecki J, Kupnicka P, Chlubek M, Gorący J, Gutowska I, Baranowska-Bosiacka I. CXCR2 Receptor: Regulation of Expression, Signal Transduction, and Involvement in Cancer. International Journal of Molecular Sciences. 2022; 23(4):2168. https://doi.org/10.3390/ijms23042168

Chicago/Turabian Style

Korbecki, Jan, Patrycja Kupnicka, Mikołaj Chlubek, Jarosław Gorący, Izabela Gutowska, and Irena Baranowska-Bosiacka. 2022. "CXCR2 Receptor: Regulation of Expression, Signal Transduction, and Involvement in Cancer" International Journal of Molecular Sciences 23, no. 4: 2168. https://doi.org/10.3390/ijms23042168

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