Next Article in Journal
GLP-1 Receptor Agonists in Neurodegeneration: Neurovascular Unit in the Spotlight
Next Article in Special Issue
β2 Integrins on Dendritic Cells Modulate Cytokine Signaling and Inflammation-Associated Gene Expression, and Are Required for Induction of Autoimmune Encephalomyelitis
Previous Article in Journal
Maternal Vitamin D Deficiency in Mice Increases White Adipose Tissue Inflammation in Offspring
Previous Article in Special Issue
LFA1 Activation: Insights from a Single-Molecule Approach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Integrin Regulators in Neutrophils

1
Department of Immunology, School of Medicine, UConn Health, 263 Farmington Ave., Farmington, CT 06030, USA
2
Academy of Integrative Medicine, Shanghai University of Traditional Chinese Medicine, 1200 Cai Lun Road, Shanghai 201203, China
3
Department of Biochemistry and Molecular Biology, University of Texas Medical Branch, 301 University Blvd., Galveston, TX 77555, USA
4
Department of Pathology, University of Texas Medical Branch, 301 University Blvd., Galveston, TX 77555, USA
5
Department of Medicine, University of California San Diego, 9500 Gilman Drive, La Jolla, CA 92093, USA
*
Authors to whom correspondence should be addressed.
Cells 2022, 11(13), 2025; https://doi.org/10.3390/cells11132025
Submission received: 29 April 2022 / Revised: 17 June 2022 / Accepted: 22 June 2022 / Published: 25 June 2022
(This article belongs to the Special Issue Integrin Activation and Signal Transduction)

Abstract

:
Neutrophils are the most abundant leukocytes in humans and are critical for innate immunity and inflammation. Integrins are critical for neutrophil functions, especially for their recruitment to sites of inflammation or infections. Integrin conformational changes during activation have been heavily investigated but are still not fully understood. Many regulators, such as talin, Rap1-interacting adaptor molecule (RIAM), Rap1, and kindlin, are critical for integrin activation and might be potential targets for integrin-regulating drugs in treating inflammatory diseases. In this review, we outline integrin activation regulators in neutrophils with a focus on the above critical regulators, as well as newly discovered modulators that are involved in integrin activation.

1. Introduction

Neutrophils account for ~60% of human circulating leukocytes, and they are central components of the host defense system for fighting against infections such as bacteria [1] and fungi [2]. Apart from their immune function against bacterial and fungal pathogens, their roles in antiviral host defense, such as human immunodeficiency virus (HIV) infection [3] and poxvirus [4], are also important.
Neutrophils are often considered effector cells of immune and inflammatory reactions due to their short-lived terminally differentiated nature [5] and effector functions, such as phagocytosis, respiratory burst, degranulation, or neutrophil extracellular trap (NET) release [6]. Mature neutrophils are technically challenging to study owing to their low mRNA content and sensitivity to environmental stimulations. Recent advancements in neutrophil biology, such as in vivo imaging, high-dimensional transcriptomic and epigenomic approaches, and studies performed at single-cell resolution, suggest that neutrophils serve functions that extend beyond their microbe-killing machinery. They respond to multiple biological signals to produce several cytokines and other inflammatory factors to regulate various functions in homeostasis and autoimmune disorders [7,8]. As heterogeneous neutrophils gained importance for their responses to different environmental challenges, determining their polarization and activation programs became critical.
Highly motile neutrophils leave the blood circulation for sites of infection through a recruitment cascade (Figure 1), which was described in previous reviews [5,9,10,11]. In the cascade, neutrophils first roll on the vascular endothelium by interacting with endothelial selectins using P-selectin glycoprotein ligand-1 (PSGL-1) [12,13]. When neutrophils encounter endothelial chemokines during rolling, integrins on neutrophils are activated, and neutrophils firmly adhere, which we call arrest [14]. After arrest, neutrophils can spread on the endothelium, perform trans-endothelium migration (TEM) [9], and migrate to the site of infection or inflammation [15,16]. Integrins play a central role in this cascade [17,18,19], as their activation and related signaling pathways mediate neutrophil arrest [20,21], TEM [9], and in-tissue migration to the site of infection or inflammation [15,16,22].
Integrins are well-studied molecules that belong to a family of transmembrane cell-surface adhesion receptors consisting of 24 transmembrane heterodimeric pairs that are generated from 18 α subunits and 8 β subunits by noncovalent association [23,24,25]. According to ligand specificity, they can be broadly separated into four groups: arginine-glycine-aspartate (RGD)-binding receptors, leukocyte-specific receptors, laminin-binding receptors, and collagen-binding receptors [26,27]. Both α and β subunits of the 700–1000 amino acid integrin extracellular domain form an elongated structure composed of a ligand-binding headpiece and a tailpiece [28,29,30,31]. Both integrin subunits have a single-pass transmembrane helical domain (∼20 amino acids), which interacts with two interfaces termed the outer and inner membrane clasps via hydrophobic interaction on the N-terminus and a salt bridge association on the C terminus. From the published data on integrin αIIbβ3, the outer membrane clasp is formed by G972/G976 in the αIIb subunit and G703 in the β3 subunit, and the inner membrane clasp is formed by F992/F993 and R995 in the αIIb subunit and W715/I719 and D723 in the β3 subunit [24,32]. The transmembrane domains are essential for integrin functions, as they allow bidirectional conformational changes. Transduction of biochemical signals and mechanical force across the plasma membrane requires the engagement of intracellular signaling and cytoskeletal proteins by integrin cytoplasmic tails and the engagement of extracellular ligands by integrin extracellular domains. This eventually disrupts the transmembrane domain interactions, separation, and angle changes of the transmembrane and intracellular domains [28,29,33,34]. Most integrin cytoplasmic tails are less than 75 amino acids in length, with the exception of β4, which is 1000 amino acids and contains four fibronectin type III repeats [35]. These repeats contain two NPXY (Asn-Pro-x-Tyr) motifs that can interact with phospho-tyrosine binding domains of intracellular proteins, such as talins [36], kindlins [37,38], and various other signaling and scaffolding molecules [39].

2. Types of Integrins Expressed on Neutrophils

Neutrophils express all four β2 integrins, including lymphocyte function-associated antigen 1 (LFA-1, also known as αLβ2 and CD11a/CD18), macrophage-1 antigen (Mac-1, also known as αMβ2, CD11b/CD18, and complement receptor 3), αXβ2 (also known as CD11c/CD18 and complement receptor 4) [40,41], and αDβ2 (also known as CD11d/CD18) [18,42], and five kinds of β1 integrins, such as α2β1 [43], α4β1 [44], α5β1 [45], α6β1 [46], and α9β1 [47].
β2 Integrins are leukocyte-specific integrins and are involved in most steps of the leukocyte recruitment cascade [48]. As mentioned above, the recruitment cascade is initiated with PSGL–selectin-mediated neutrophil rolling [12,13]. PSGL–selectin interaction induces the extension of β2 integrins, especially LFA-1 [49,50]. The extended low-affinity LFA-1 interacts with its ligand, intercellular adhesion molecule 1 (ICAM-1), and slows down the rolling velocity, which we call slow rolling [49,51,52]. Neutrophils encounter endothelial chemokines, such as interleukin-8 (IL-8) [53], or C-X-C Motif Chemokine Ligand 1 (CXCL1) [54], to initiate integrin inside-out signaling and fully activate β2 integrins so that they have high-affinity binding to ligands in trans, such as LFA-1 (both humans and mice) [49,50] and Mac-1 (humans but not mice) [12,13]. Fully activated β2 integrins bind endothelial ICAM-1 with high affinity and lead to neutrophil arrest [55,56]. Crawling through the inflamed vessel requires Mac-1 but not LFA-1. A Mac-1 knockout model resulted in crawling neutrophils with decreased distance and velocity [57]. It has also been shown that Mac-1 prevents neutrophils from migrating upstream, as their blockade allows neutrophil upstream migration in an LFA-1-dependent manner [58]. During transmigration, neutrophils move through the space in the disrupted endothelial membrane in a Mac-1 and LFA-1-dependent manner [59]. Neutrophils prefer the paracellular route (between endothelial cells) rather than the transcellular route (through endothelial cells) [22] through LFA1 binding to ICAM-1 [58,60,61,62], ICAM-2 [62,63,64], and JAM-A [64,65]. The release of myeloperoxidases by neutrophils during migration and their interactive products with the extracellular matrix (ECM) can, in turn, activate αMβ2 and αDβ2 integrins on macrophages at the site of inflammation [66]. Neutrophil β2 integrins play critical roles in other neutrophil functions, such as phagocytosis [40,41,67], cell differentiation [68,69], degranulation [70,71,72,73,74,75], and formation of NETs [3,76].
β1 Integrins are widely expressed on various cells, and some of them are expressed on neutrophils. They bind vascular adhesion molecules, such as vascular cell adhesion molecule 1 (VCAM-1), and components of ECM, such as fibronectin, collagen, and laminin [77]. Integrin α4β1 is expressed on rat and mouse neutrophils and mediates neutrophil adhesion to endothelial cells by interacting with VCAM-1 [47,78,79,80]. Human neutrophils may not express integrin α4β1 [44]. However, it has been shown that integrin α4β1 may be expressed on neutrophils from sickle cell disease [81] or sepsis patients [82] and contributes to neutrophil adhesion. It has been shown that human neutrophils adhere to laminin via an α6β1 integrin-dependent mechanism [46], to fibronectin via an α5β1-dependent mechanism [45], and to collagen via an α2β1-dependent mechanism [83]. Integrin α9β1 expression is upregulated two- to three-fold on human neutrophils after activation, and it binds VCAM-1 to stabilize interactions between rolling neutrophils and the endothelium and enhance β2 integrin-dependent neutrophil slow rolling and arrest. Neutrophil integrin α9β1 also binds to matrix proteins tenascin C and osteopontin [47]. Rodent studies reported the upregulation of α2β1 and α6β1 integrin membrane expression during neutrophil recruitment to inflammation sites [83,84,85,86]. The major role of β1 integrins is to mediate cell–matrix adhesion and promote leukocyte motility in the perivascular and ECM areas [87].

3. Integrin Conformational Changes during Activation

As mentioned above, integrins have long ectodomains, transmembrane domains, and short cytoplasmic tails [25,28,88]. They undergo large conformational changes in their extracellular domains during activation (Figure 2) to change their ability to bind ligands [25,89,90,91,92,93]. The ectodomain consists of a headpiece and a tailpiece. The conformational changes in integrins have been heavily discussed in previous reviews [20,25,34,42,66,90,94,95,96,97].
The canonical conformational change model in integrin activation is the switchblade model [25,49,96,98]. Resting integrins have bent extracellular domains, where their membrane-distal headpiece bends toward their membrane-proximal tailpiece [31]. In the switchblade model, bent low-affinity (EH−) integrins first extend to an extended low-affinity (E+H) conformation. Then, the hybrid domain swing-out leads to high ligand-binding affinity in the headpiece and becomes an extended high-affinity (E+H+) conformation. However, the switchblade model does not cover the existing bent high-affinity (EH+) conformation reported in αVβ3 [99], αXβ2 [100], and three Mac-1 mutants [101]. The identification of a ligand-bound bent-open integrin αVβ3 [31] suggested a deadbolt model [28]. This model proposes that a hairpin loop in the β tail domain acts as a deadbolt to restrain the displacement of the β-A/I domain β6-α7 loop and maintain the integrin in the H- state. The displacement of this loop releases to attain EH+ conformation, allowing ligand binding with an auto-inhibitory effect on integrin αVβ3 in a binding complex with fibronectin activation. Then, the binding of ligands provides a pulling force that extends the ectodomain. This model is controversial in that the “deadbolt” works on Mac-1 [101] but not in β3 integrin [102]. Using live-cell quantitative dynamic footprinting microscopy, we observed both transitions of E+H to E+H+ and EH+ to E+H+ in neutrophil β2 integrins, suggesting that both models may exist at the same time [102].

4. Integrin Activation Modulators

The integrin α and β chains contain conserved regions to bind different modulator proteins, such as NPXY motifs in the β cytoplasmic tail [39,103,104] and a GFFKR sequence in the α cytoplasmic tail [105]. Talins [106] and kindlins [37,107,108] are the most studied and well-accepted modulators that bind to the integrin β cytoplasmic tail and regulate integrin activation. Filamin A [109,110], DOK1 [111], 14-3-3ζ [112], α-actinin [113], Rap1-interacting adaptor molecule (RIAM) [114,115], and Rap-1 GTPases [116,117,118,119] were also reported to bind the integrin β cytoplasmic tail directly or indirectly. Calreticulin [120,121], RapL [122], paxillin [123], and SHARPIN [124] were reported to bind the integrin α cytoplasmic tail. SHARPIN also binds to the β cytoplasmic tail of β1 integrin [124]. There are some other molecules that reportedly modulate integrin activation. Whether they bind integrins directly is unknown. We discuss these integrin activation modulators in another section.

4.1. Talins

Talins are important integrin-binding proteins that exist in two isoforms, talin-1 and talin-2. Talin is composed of a 50 kDa N-terminal FERM (protein 4.1, ezrin, radixin, moesin) talin head domain (THD) and a 220 kDa C-terminal rod domain (Figure 3A) [125]. Talin-1 is expressed in all cell types, while talin-2 is highly expressed in skeletal and cardiac muscles and brain tissue [126,127]. A knockout study on mice showed that talin-1 is required for neutrophil LFA-1 extension and neutrophil slow rolling and arrest [128]. Talin-1-deficient mice showed defective neutrophil adhesion and spreading, along with impaired extravasation [119]. A knockdown study showed that talin-1 is required for Mac-1-mediated phagocytosis [129].
Both the THD and the rod domain have been reported to have integrin binding sites. The THD contains four subunits: F0, F1, F2, and F3 [130]. The F3 subunit of the THD interacts with the highly conserved membrane-proximal NPXY motifs of the cytoplasmic tails of integrin β subunits and induces integrin activation [131,132,133,134,135,136]. Mutations in the THD identified two sites that are involved in regulating integrin activation [136] and neutrophil adhesion [137]. The W359A mutation blocked talin interaction with integrins and showed similar neutrophil defects compared to the talin knockout. The L325R mutation did not block talin interaction with integrins or neutrophil slow rolling but showed a defect in neutrophil arrest and spreading. Mutations in the β2 integrin cytoplasmic tail showed that W747 and F754 are required for talin binding [129]. Talin recruitment to the integrin cytosolic tail is mediated by the GTPase Rap1 and RIAM (gene name: amyloid-beta precursor protein-binding family b member 1 interacting protein, APBB1IP) complex [118,138]. Recent works indicate that direct Rap1–talin1 binding plays a critical role in integrin activation in platelets [118,139,140] and also regulates integrin activation in blood cells [114,140,141]. The interaction of the THD F3 with the membrane-proximal α helix of the β integrin cytoplasmic tail disrupts the connection between cytoplasmic tails [142]. Consecutive electrostatic interactions between the THD and phospholipids on the plasma membrane contribute to the separation and re-orientation of integrin cytoplasmic tails [143]. The THD F3 forms a well-defined complex with the helix-forming membrane-proximal (MP) region of the β-integrin tail, and this interaction holds the key to the molecular recognition required for activation [136]. The talin rod contains a second integrin-binding site (IBS2) that may bind the membrane-proximal region of the integrin cytoplasmic domain, which may only happen after integrin activation [144,145,146].
A resting talin is autoinhibited. The integrin-binding sites of talin are masked in its autoinhibitory conformation, where the rod segment folds back onto its F3 subdomain [36]. Autoinhibition of talin has been shown to regulate its recruitment to adhesome sites and the maturation of focal adhesions in mouse embryonic fibroblast cells [147]. THD binding to the rod inhibits its binding to the integrin cytoplasmic tail. There are interactions between talin molecules to form dimers, showing an inhibitory effect on talin [148]. It has been shown that in β1 integrins, a mutation that blocks talin autoinhibition leads to increased integrin activation in mouse embryonic fibroblast cell lines, along with increased focal adhesion maturation and stability [147]. Phosphatidylinositol 4,5-bisphosphate (PIP2) binding to talin F2 and F3 domains [149], calpain cleavage [150], and phosphorylation events [147,151] can activate talin. However, these have not been demonstrated in neutrophils.
Although the F3 domain forms a complex with integrin, F0 and F1 domains are also required for integrin activation. Although talin F3 can bind β1A and β3 tails with similar affinity on CHO or HT1080 cells (FN9–11 binding), the expression of the whole talin head (residues 1–405) is required for β1A-integrin activation. The activation especially requires the F0 F1 domain, in which the N-terminal residues 1–85 precede the FERM domain (F0 domain), the F1 FERM domain, and the integrin-binding F3 domain [152]. A large loop within the F1 domain shows that it has a propensity to adopt a helical structure in which basic residues are clustered on one surface and that it interacts with vesicles containing acidic phospholipids [153]. These findings have not been demonstrated in neutrophils as well.

4.2. RIAM

RIAM is a member of the Mig-10/RIAM/Lamellipodin (MRL) protein family (Figure 3B), which is directly bound to talin via a short N-terminal sequence that was predicted to form amphipathic helices. RIAM binds Rap1, which is discussed in a later section, and functions as a scaffold that connects the membrane-targeting sequences in Ras GTPases to talin [138]. The inhibitory (IN) segment of RIAM and the RA domain at its Rap1-binding site form an autoinhibitory conformation to lock RIAM in an inactive state. Phosphorylations of the IN segment and PH domain by FAK and Src, respectively, release the inhibitory state and activate RIAM, facilitating RIAM interaction with Rap1 and PIP2 to induce talin-1-dependent integrin activation [154,155]. RIAM-induced integrin activation requires its capacity to bind to both Rap1 and talin [115]. Interestingly, RIAM and vinculin have mutually exclusive binding sites on talin [156]. In adhesive cells, RIAM-containing adhesions are primarily in the lamellipodium. RIAM is subsequently reduced in mature focal adhesions due to direct competition with vinculin for talin-binding sites [157]. Whether it has a similar mechanism during neutrophil migration remains to be investigated.
RIAM is abundant in hematopoietic cells, and its absence blocks agonist-mediated αIIbβ3 activation in primary mouse megakaryocytes [158] without affecting development, homeostasis, or platelet integrin functions, as RIAM levels are low in normal platelets, whereas Rap1, talin1, and integrins are highly expressed in platelets, indicating the existence of RIAM-independent Rap1 regulation [159]. RIAM has an indispensable role in the activation of β2 integrins in neutrophils, macrophages, and T cells [119,160]. Studies conducted on mice that both carry Rap1-binding mutant talin1 and lack RIAM expression showed increased neutrophil rolling velocities and decreased adhesion to inflamed cremaster muscle venules compared to wild type or single-mutant mice (Rap1-binding mutant talin1 knock-in or RIAM knockout) by affecting conformational changes in the β2 integrin ectodomain [114]. This specifically regulates leukocyte β2 integrins [119]. Extravasation, αMβ2-mediated adhesion, and spreading to immobilized immune complexes were found to be impaired in RIAM knockout neutrophils [119]. RIAM mainly regulates the activation of β2 integrins; however, it only partially affects integrin β1 [119]. Moreover, RIAM is dispensable for integrins in other cell types, such as fibroblasts or platelets. This may be because (1) the expression of RIAM is low in these cells; (2) the RIAM-dependent β2 activation complex is only formed in certain hematopoietic cells [114]. The hypothesis is supported by recent work that showed that RIAM is dispensable for β2 integrin activation in regulatory T cells [97]. In regulatory T cells, β2 integrins can be activated by Lamellipodin, which is the RIAM paralog protein and is highly expressed in regulatory T cells in the absence of RIAM. These results revealed that RIAM function differs depending on the cell type and integrin class, suggesting a potential method to specifically manipulate the trafficking and function of selective cell types.
Adaptor molecules such as 55 kDa src kinase-associated phosphoprotein (SKAP-55) and the adhesion and degranulation-promoting adapter protein (ADAP) consistently interact with RIAM to promote the membrane targeting of the RIAM-Rap1 module for antigen stimulation of T cells to facilitate LFA-1 integrin activation [161].

4.3. Rap1

Rap1 proteins are small GTPases of the Ras family and are encoded by two Rap1 genes, Rap1A and Rap1B [162]. GDP-bound inactive Rap1 is activated when GTP is exchanged for GDP, which is regulated by several guanine nucleotide-exchange factors (GEFs), and they participate in various signaling pathways, including integrin activation, ERK activation, and other effector pathways [163]. Rap1 signaling is terminated by the hydrolysis of bound GTP to GDP by specific GTPase-activating proteins (GAPs) such as Rap1GAP1 [164] and signal-induced proliferation-associated gene-1 (SPA-1) [165]. Activation of Rap1 GTPase effectors was found to have different consequences on various cells, which has been reviewed before [166,167]. They have a wide variety of functions, such as control of the establishment of cell polarity [91,168], activation of integrin-mediated cell adhesion [169,170], and the regulation of cell–cell contacts [171,172], migration [172,173], cell proliferation [174], and secretion [163,166]. Rap1 deficiency can markedly suppress neutrophil functions by inhibiting the activation of β2 integrin [118].
Despite the Rap1/RIAM/talin axis for integrin activation [114,138], the membrane-anchored Rap1 binds to the F0 domain of talin [118,139]. Although the binding of Rap1 and the talin F0 domain is critical in Drosophila [139,175], it displayed mild defects in talin1-induced αIIbβ3 activation and platelet function, including aggregation and hemostasis [139]. A novel Rap1-binding site in the talin F1 domain was recently identified [176]. Blocking the ability of Rap1 to bind to the talin F1 domain profoundly disrupted the Rap1 function that mediates talin1-induced integrin activation in platelets [140,176]. The association of Rap1 and the talin F1 domain plays a central role in Rap1-mediated integrin activation in platelets [140,177]. Moreover, recent work reported that the Rap1 and talin association plays a critical role in integrin activation in leukocytes as well [114,178]. The Tln13mut mouse with K15A, R30A, and R35A mutations in the F0 domain showed a mild neutrophil adhesion defect and reduced extravasation [118]. However, the Tln13mut mutation combined with RIAM deficiency dramatically blocked neutrophil integrin functions [114]. However, the talin F1 mutation R118E, which blocks the binding of Rap1 to the talin F1 domain, significantly affected both neutrophils and lymphocytes [140,178], indicating that Rap1 binding to talin F1 plays a more important role than the Rap1–F0 interaction. The role of direct Rap1–talin binding in neutrophil function warrants further investigation.
The RASGRP2 gene encodes the Ca2+ and diacylglycerol-regulated guanine nucleotide exchange factor I (CalDAG-GEFI) protein, a guanine nucleotide-exchange factor for the small GTPase Rap1; this is indispensable for platelet Rap1 activity, as mutations in this gene abrogate Rap1 activation and cause platelet dysfunction [179]. This gene is highly expressed in neutrophils and also showed defective integrin activation in mutational studies [180]. Neutrophils from CalDAG-GEFI knockout mice exhibited strong defects in Rap1 and β1 and β2 integrin activation while maintaining normal calcium flux, degranulation, and ROS generation. Hence, neutrophils in these mice failed to adhere firmly to stimulated venules and migrate into sites of inflammation [180].
The PH domain of SWAP-70-like adaptor of T cells (SLAT), also known as DEF6, has a regulatory function on the active form of the small GTPase Rap1 in LFA-1 activation on T cells [181]. Similarly, B cell adaptor molecule of 32 kDa (Bam32) has a suppressive role in chemokine-induced neutrophil chemotaxis and transmigration by regulating Rap1 activation in neutrophils [182].
A recent study reported that phostensin (PTSN), a regulatory subunit of protein phosphatase 1, mediates dephosphorylation of Rap1 and regulates integrin activation [183]. PTSN is mainly expressed in leukocytes [184], and its tissue distribution is similar to that of RIAM. PTSN-deficient mice exhibited a significant increase in peripheral blood neutrophils, suggesting that PTSN may play an important role in modulating neutrophil function [183].

4.4. Kindlin

Kindlin is another integrin-regulating protein that contains a 4.1-ezrin-radixin-moesin (FERM) domain and a pleckstrin homology (PH) domain (Figure 3C). There are three kindlin paralogs, kindlin-1, -2, and -3. Kindlin-1 is mainly expressed in epithelial cells and keratinocytes. Kindlin-2 is expressed in many cell types, such as muscle cells and epidermal cells, but not in hematopoietic cells [185]. Kindlin-3 is expressed in all hematopoietic cells [186]. Kindlin-1 binds integrins β1, β3, and β6, and kindlin-2 and -3 bind integrins β1, β2, and β3 [187] to regulate integrin activation [188].
Mutations in the kindlin-3 gene FERMT3 cause LAD-III syndrome [189,190]. In a mouse knockout study, kindlin-3 was found necessary for the LFA-1 high-affinity conformation and neutrophil arrest [128,191,192]. The FERM-like kindlin molecule contains the F0, F1, F2, and F3 subdomains linearly with a unique PH domain inserted into the F2 subdomain [193,194]. The kindlin-3 F3 subdomain binds to the distal NxxY/F motif [188] and the TTT759−761 motif [195] of the integrin β2 cytoplasmic tail. Kindlin-3 binds β2 integrin through its QW614/615 residues in its F3 subdomain, and a mouse strain carrying the QW/AA mutation was generated [196]. Neutrophils from this mouse strain showed defects in both adhesion and NET release [197]. However, a follow-up study from the same group showed that overexpression of kindlin-3, regardless of WT or QW/AA mutation, in neutrophil-like HL60 cells inhibited NET generation [198]. Kindlin-3 knockdown HL60 cells and neutrophils from kindlin-3flox/floxMx1-Cre mice showed enhanced NET generation [198].
Kindlin-3 is also important for podosome assembly by regulating integrin activation and clustering [199]. Furthermore, the kindlin-3 F0 domain binds leupaxin, recruits leupaxin into podosomes, dephosphorylates paxillin, and increases podosome stability [199]. Kindlin-3 was recruited to the plasma membrane in response to interleukin-8 (IL-8) before the induction of high-affinity β2-integrin conformations [107,200,201]. The PH domain of kindlin is necessary for the plasma membrane recruitment of kindlin-3, which occurs before rolling. The PH domain of kindlin-3 interacts with the scaffold protein receptor of activated protein C kinase 1 (RACK1) [202]. A highly conserved poly-lysine motif in the loop of the domain of kindlin-1 [203] and kindlin-3 [204] supports binding to the negatively charged phospholipid head group. A recent study showed that the F2 PH and F3 subdomains are important for kindlin self-association, which negatively regulates integrin binding, and they identified kindlin-3 point mutations that decrease self-association and enhance integrin binding while maintaining the ability to localize to focal adhesions [24,205].
In addition to integrins, kindlin proteins also interact with binding partners such as integrin-linked kinase (ILK), migfilin, and RACK1 [202]. ILK is reported to have a PKCα-mediated phosphorylation function on kindlin-3, which is required for the chemokine-induced full activation of LFA-1 [206].

4.5. Linking of Integrins to the Actin Cytoskeleton

Linking integrins to the actin cytoskeleton is important for force transmission from extracellular ligands to neutrophil cytoskeletons to support neutrophil adhesion and migration. Moreover, treatment with latrunculin B or blebbistatin to block actin polymerization in human neutrophils impaired β2 integrin activation to high-affinity conformations and inhibited human neutrophil arrest on ICAM-1 [207]. Thus, knowing how the actin cytoskeleton tethers to leukocyte integrins is important in understanding β2 integrin function.
Several integrin activation modulators mentioned above, such as talin, RIAM, and kindlin, have actin-binding sites. Talin (Figure 3A), the THD, and rod domains contain direct actin-binding sites and indirect sites that bind actin-binding vinculin, etc. [125,142,207,208,209]. The F0 domain of kindlin-2 (Figure 3C) was reported to bind actin directly [210]. RIAM (Figure 3B) binds actin indirectly through profilin [211].
α-Actinin is a cytoskeletal actin-binding protein that binds to β2 integrins directly [212]. The cytoplasmic tails of the integrin β2 subunit bind to α-actinin in focal adhesions, adherent junctions, and hemidesmosomes [213]. In fibroblasts, talin and α-actinin compete for binding to β3 integrins [214]. Deletion of α-actinin enhances force generation in initial adhesions on fibronectin but impairs mechanotransduction in a subsequent step, preventing adhesion maturation [214]. It has been shown that α-actinin may be expressed in neutrophils [215,216]; however, the role of α-actinin in regulating neutrophil integrins remains unknown.
Coronin 1A is a family member of evolutionarily conserved actin-binding proteins that regulate actin cytoskeleton-dependent processes. It is predominantly expressed in leukocytes [217] and has been identified as a regulator of β2 integrins that interacts with the cytoplasmic tail of β2 integrins and is crucial for the induction of neutrophil adhesion, spreading, and migration [218]. Coronin 1A knockout neutrophils showed less β2 integrin-dependent soluble ICAM-1 binding and LFA-1 clustering. Another study showed that loss of Coronin 1A results in less β2 integrin translocation to the platelet surface [219].
Transient receptor potential channel 6 (TRPC6) knockdown showed attenuated chronic hypoxia-induced actin assembly and actin reorganization in human mesangial cells [220]. TRPC6 has a significant role in the CXCL1-dependent recruitment of murine neutrophil granulocytes. The recruitment of neutrophils after renal reperfusion injury was attenuated in TRPC6 knockout mice. TRPC6 knockout neutrophils showed diminished Rap1 and β2 integrin activation, resulting in decreased CXCL1-induced neutrophil adhesion and transmigration [221].

4.6. Other Molecules That Regulate Neutrophil Integrins

Mitofusins (MFNs) are GTPases embedded in the outer membrane of the mitochondria and are essential for mitochondrial fusion. It was found that interfering with MFN2 expression using shRNA can significantly suppress the chemotaxis of neutrophil-like DMSO-differentiated HL-60 (dHL60) cells [222]. This was further demonstrated by neutrophil recruitment defects in MFN2 knockout zebrafish and MFN2flox/floxMRP8-cre mice [223]. MFN2 knockdown dHL60 cells showed a defect in adhesion on inflamed human umbilical vein endothelial cells (HUVEC) [223]. Using microfluidic devices, our group has shown that MFN2 knockdown dHL60 cells have defects in slow rolling and arrest on P-selectin/ICAM-1 and P-selectin/ICAM-1/IL8 substrates, respectively, suggesting a role of MFN2 in regulating β2 integrin activation on neutrophils [224]. MFN2 knockdown dHL60 cells have reduced formyl peptide receptor (FPR) expression and FPR-dependent (fMLP stimulation) and independent (PMA stimulation) β2 integrin activation defects. MFN2 knockdown dHL60 cells also show defects in actin polymerization after fMLP or PMA stimulation and Mn2+-induced cell spreading on ICAM-1. We demonstrated that MFN2 knockdown HL60 cells are unable to differentiate into neutrophil-like dHL60 cells by assessing the nuclear morphology and maturation markers CD35 and CD87. Using the CD87 maturation marker, we found that in the mature CD87high population, MFN2 knockdown HL60 cells still show defects in cell slow rolling, adhesion, and β2 integrin activation, indicating that, besides its effects on differentiation, MFN2 is directly involved in regulating β2 integrin activation. Please note that in these mature populations, MFN2 only affects extension (which is reported by the KIM127 antibody) but not headpiece opening (which is reported by the mAb24 antibody) under PMA stimulation. This suggests that MFN2 might be important for the conformational changes of bent-open to extended-open β2 integrins, which is an alternative allosteric pathway of β2 integrins that we observed before [21,102,224,225,226], in addition to the conical switchblade model [25]. However, since MFN2 is an important mitochondrial regulator, whether MFN2 directly affects the integrin activation pathway or indirectly affects integrin activation by altering mitochondrial function is unclear. It has been shown that MFN2 is critical for mitochondrial respiration in fibroblasts [227] and macrophages (Tur et al., 2020). Whether it is the same in neutrophils is unknown. MFN2 is also known to regulate the tethering of mitochondria and endoplasmic reticulum, which is important for intracellular calcium regulation [228,229]. It has been shown that MFN2 is involved in mitochondria–endoplasmic reticulum interaction in neutrophil-like HL60 cells and may regulate intracellular calcium [223]. Thus, it is also possible that MFN2 regulates integrin activation through intracellular calcium. In macrophages, MFN2 silencing leads to reduced ER–mitochondria contacts and mitochondrial activity. The inflammatory and pro-inflammatory responses upon administration of inflammatory agents also showed increased responses in MFN2-deficient macrophages [230]. Deficient ROS production in the absence of MFN2 impairs the induction of cytokines and nitric oxide and is associated with dysfunctional autophagy, apoptosis, phagocytosis, and antigen processing. The lack of MFN2 in macrophages causes an impaired response in a model of non-septic inflammation in mice, as well as a failure in protection from Listeria, Mycobacterium tuberculosis, or LPS endotoxemia [231,232].
Several cytoplasmic components regulate the interactions of integrin through the process of phosphorylation and dephosphorylation. Leukocyte integrin α chain phosphorylation at Ser-1140 (αL), Ser-1126 (αM), or Ser-1158 (αX) is essential for leukocyte adhesion and intracellular signaling [233,234,235]. The β2 chain becomes phosphorylated after activation through chemokines, the TCR, or phorbol esters. The protein kinase C (PKC) enzyme phosphorylates the β2 chain at Thr-758, leads to the release of bound filamin, and promotes the binding of 14-3-3 proteins, whereas talin can bind to both the Thr-758 phosphorylated and unphosphorylated chains [112]. Following T cell receptor stimulation, phosphorylation of the LFA-1 β2 chain on Thr-758 leads to 14-3-3 recruitment to the integrin, actin cytoskeleton reorganization, and increased adhesion [236]. Another study conducted using a phosphorylation model on the LFA-1 α chain at Ser-1140 showed that it affects β chain phosphorylation at Thr-758 with significantly reduced binding of α-actinin and 14-3-3 on SDF-1α-activated mutant cells [113].
Dok1 and Dok2 negatively regulate immune cell signaling [237], and Dok1 specifically negatively regulates the Ras-ERK pathway by binding to p120RasGAP [238]. Dok1 is an integrin inhibitor that competes with talin for binding to the tyrosine-phosphorylated proximal NPXY sequence in β3 [106]. Dok1 and Dok2 bind weakly to β2 cytoplasmic peptides but bind strongly to peptides phosphorylated at S756. In the neutrophil αMβ2 integrin, the small G protein Rap1 binds to the phosphorylated S756 and blocks Dok1 binding-mediated integrin inhibition [239].
Shank-associated RH domain-interacting protein (SHARPIN) is a 45 kDa cytosolic protein that is widely expressed [240]. It is one of the three subunits of the linear ubiquitin chain assembly complex (LUBAC), an E3 ubiquitin ligase enzyme complex [241,242]. This complex comprises two subunits other than SHARPIN: a large isoform of heme-oxidized iron regulatory protein 2 (IRP2) ubiquitin ligase 1 (HOIL-1L) and HOIL-1L interacting protein (HOIP), on which the RING-in-between-RING (RBR) domain of HOIP is the catalytic center for linear ubiquitination [243,244]. A high-throughput RNAi screen in PC3 prostate cancer cells identified SHARPIN as an endogenous inhibitor of β1-integrin activity. SHARPIN silencing induced a significant increase in active β1 integrins on the cell surface without altering total surface β1 integrin expression [245]. SHARPIN could inactivate integrins by modulating the expression and/or function of the β1-integrin activators talin or kindlin or via the LUBAC-stimulated formation of linear ubiquitin chains involved in signaling [246]. SHARPIN regulates β1 integrin-dependent cell adhesion and migration through its unique binding to the β1-associated α chain and by inhibiting the recruitment and binding of kindlin and talin to the integrin in several normal and malignant cell types [124]. The loss of SHARPIN and increased integrin activity in mice in vivo depicts its negative regulation in integrin activation [245]. Mice deficient in SHARPIN had increased neutrophils in the spleen and peripheral blood [247]. However, whether it is due to global inflammation in these mice or a cell-intrinsic abnormality of integrin activation in neutrophils requires further investigation.

Author Contributions

S.P. drafted the manuscript and prepared the figures; L.H., Y.C., H.S. and Z.F. edited and revised the manuscript; H.S. and Z.F. approved the final version of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Institutes of Health, USA, grant number [R01 HL145454 (Z.F.) and R00 HL153678 (Y.C.)] and UConn Health (a startup fund) and The APC was funded by [R01 HL145454].

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We acknowledge Christopher “Kit” Bonin and Geneva Hargis from UConn Health School of Medicine for their help with scientific writing and editing of this manuscript.

Conflicts of Interest

No conflict of interest, financial or otherwise, are declared by the authors.

References

  1. Witter, A.R.; Okunnu, B.M.; Berg, R.E. The Essential Role of Neutrophils During Infection with the Intracellular Bacterial Pathogen Listeria Monocytogenes. J. Immunol. 2016, 197, 1557–1565. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Drummond, R.A.; Collar, A.L.; Swamydas, M.; Rodriguez, C.A.; Lim, J.K.; Mendez, L.M.; Fink, D.L.; Hsu, A.P.; Zhai, B.; Karauzum, H.; et al. CARD9-Dependent Neutrophil Recruitment Protects against Fungal Invasion of the Central Nervous System. PLoS Pathog. 2015, 11, e1005293. [Google Scholar] [CrossRef] [PubMed]
  3. Saitoh, T.; Komano, J.; Saitoh, Y.; Misawa, T.; Takahama, M.; Kozaki, T.; Uehata, T.; Iwasaki, H.; Omori, H.; Yamaoka, S.; et al. Neutrophil Extracellular Traps Mediate a Host Defense Response to Human Immunodeficiency Virus-1. Cell Host Microbe 2012, 12, 109–116. [Google Scholar] [CrossRef] [Green Version]
  4. Jenne, C.N.; Wong, C.H.Y.; Zemp, F.J.; McDonald, B.; Rahman, M.M.; Forsyth, P.A.; McFadden, G.; Kubes, P. Neutrophils Recruited to Sites of Infection Protect from Virus Challenge by Releasing Neutrophil Extracellular Traps. Cell Host Microbe 2013, 13, 169–180. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Ley, K.; Hoffman, H.M.; Kubes, P.; Cassatella, M.A.; Zychlinsky, A.; Hedrick, C.C.; Catz, S.D. Neutrophils: New Insights and Open Questions. Sci. Immunol. 2018, 3, eaat4579. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Mócsai, A. Diverse Novel Functions of Neutrophils in Immunity, Inflammation, and Beyond. J. Exp. Med. 2013, 210, 1283–1299. [Google Scholar] [CrossRef] [Green Version]
  7. Bonaventura, A.; Montecucco, F.; Dallegri, F.; Carbone, F.; Lüscher, T.F.; Camici, G.G.; Liberale, L. Novel Findings in Neutrophil Biology and Their Impact on Cardiovascular Disease. Cardiovasc. Res. 2019, 115, 1266–1285. [Google Scholar] [CrossRef]
  8. Scapini, P.; Cassatella, M.A. Social Networking of Human Neutrophils within the Immune System. Blood 2014, 124, 710–719. [Google Scholar] [CrossRef]
  9. Filippi, M.-D. Neutrophil Transendothelial Migration: Updates and New Perspectives. Blood 2019, 133, 2149–2158. [Google Scholar] [CrossRef]
  10. Liew, P.X.; Kubes, P. The Neutrophil’s Role During Health and Disease. Physiol. Rev. 2019, 99, 1223–1248. [Google Scholar] [CrossRef]
  11. Margraf, A.; Ley, K.; Zarbock, A. Neutrophil Recruitment: From Model Systems to Tissue-Specific Patterns. Trends Immunol. 2019, 40, 613–634. [Google Scholar] [CrossRef] [PubMed]
  12. Marki, A.; Esko, J.D.; Pries, A.R.; Ley, K. Role of the Endothelial Surface Layer in Neutrophil Recruitment. J. Leukoc. Biol. 2015, 98, 503–515. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Morikis, V.A.; Simon, S.I. Neutrophil Mechanosignaling Promotes Integrin Engagement with Endothelial Cells and Motility Within Inflamed Vessels. Front. Immunol. 2018, 9, 2774. [Google Scholar] [CrossRef] [PubMed]
  14. Morikis, V.A.; Hernandez, A.A.; Magnani, J.L.; Sperandio, M.; Simon, S.I. Targeting Neutrophil Adhesive Events to Address Vaso-Occlusive Crisis in Sickle Cell Patients. Front. Immunol. 2021, 12, 1256. [Google Scholar] [CrossRef]
  15. Maas, S.L.; Soehnlein, O.; Viola, J.R. Organ-Specific Mechanisms of Transendothelial Neutrophil Migration in the Lung, Liver, Kidney, and Aorta. Front. Immunol. 2018, 9, 2739. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Sadik, C.D.; Kim, N.D.; Luster, A.D. Neutrophils Cascading Their Way to Inflammation. Trends Immunol. 2011, 32, 452–460. [Google Scholar] [CrossRef] [Green Version]
  17. Herter, J.; Zarbock, A. Integrin Regulation during Leukocyte Recruitment. J. Immunol. 2013. [Google Scholar] [CrossRef] [Green Version]
  18. Kourtzelis, I.; Mitroulis, I.; von Renesse, J.; Hajishengallis, G.; Chavakis, T. From Leukocyte Recruitment to Resolution of Inflammation: The Cardinal Role of Integrins. J. Leukoc. Biol. 2017, 102, 677–683. [Google Scholar] [CrossRef] [Green Version]
  19. Sökeland, G.; Schumacher, U. The Functional Role of Integrins during Intra- and Extravasation within the Metastatic Cascade. Mol. Cancer 2019, 18, 12. [Google Scholar] [CrossRef]
  20. Fan, Z.; Ley, K. Leukocyte Arrest: Biomechanics and Molecular Mechanisms of Β2 Integrin Activation. Biorheology 2015, 52, 353–377. [Google Scholar] [CrossRef] [Green Version]
  21. Sun, H.; Hu, L.; Fan, Z. Β2 Integrin Activation and Signal Transduction in Leukocyte Recruitment. Am. J. Physiol. Cell Physiol. 2021, 321, C308–C316. [Google Scholar] [CrossRef] [PubMed]
  22. Ley, K.; Laudanna, C.; Cybulsky, M.I.; Nourshargh, S. Getting to the Site of Inflammation: The Leukocyte Adhesion Cascade Updated. Nat. Rev. Immunol. 2007, 7, 678–689. [Google Scholar] [CrossRef] [PubMed]
  23. Arias-Mejias, S.M.; Warda, K.Y.; Quattrocchi, E.; Alonso-Quinones, H.; Sominidi-Damodaran, S.; Meves, A. The Role of Integrins in Melanoma: A Review. Int. J. Derm. 2020, 59, 525–534. [Google Scholar] [CrossRef]
  24. Kadry, Y.A.; Calderwood, D.A. Chapter 22: Structural and Signaling Functions of Integrins. Biochim. Biophys. Acta Biomembr. 2020, 1862, 183206. [Google Scholar] [CrossRef] [PubMed]
  25. Luo, B.-H.; Carman, C.V.; Springer, T.A. Structural Basis of Integrin Regulation and Signaling. Annu. Rev. Immunol. 2007, 25, 619–647. [Google Scholar] [CrossRef] [Green Version]
  26. Moreno-Layseca, P.; Icha, J.; Hamidi, H.; Ivaska, J. Integrin Trafficking in Cells and Tissues. Nat. Cell Biol. 2019, 21, 122–132. [Google Scholar] [CrossRef]
  27. Ou, Z.; Dolmatova, E.; Lassègue, B.; Griendling, K.K. Β1- and Β2-Integrins: Central Players in Regulating Vascular Permeability and Leukocyte Recruitment during Acute Inflammation. Am. J. Physiol. Heart Circ. Physiol. 2021, 320, H734–H739. [Google Scholar] [CrossRef]
  28. Arnaout, M.A.; Mahalingam, B.; Xiong, J.-P. Integrin Structure, Allostery, and Bidirectional Signaling. Annu. Rev. Cell Dev. Biol. 2005, 21, 381–410. [Google Scholar] [CrossRef] [Green Version]
  29. Hu, P.; Luo, B.-H. Integrin Bi-Directional Signaling across the Plasma Membrane. J. Cell. Physiol. 2013, 228, 306–312. [Google Scholar] [CrossRef]
  30. Stadtmann, A.; Zarbock, A. The Role of Kindlin in Neutrophil Recruitment to Inflammatory Sites. Curr. Opin. Hematol. 2017, 24, 38–45. [Google Scholar] [CrossRef]
  31. Xiong, J.P.; Stehle, T.; Diefenbach, B.; Zhang, R.; Dunker, R.; Scott, D.L.; Joachimiak, A.; Goodman, S.L.; Arnaout, M.A. Crystal Structure of the Extracellular Segment of Integrin Alpha Vbeta3. Science 2001, 294, 339–345. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Lau, T.-L.; Kim, C.; Ginsberg, M.H.; Ulmer, T.S. The Structure of the Integrin AIIbβ3 Transmembrane Complex Explains Integrin Transmembrane Signalling. EMBO J. 2009, 28, 1351–1361. [Google Scholar] [CrossRef] [PubMed]
  33. Morse, E.M.; Brahme, N.N.; Calderwood, D.A. Integrin Cytoplasmic Tail Interactions. Biochemistry 2014, 53, 810–820. [Google Scholar] [CrossRef] [PubMed]
  34. Rocco, M.; Rosano, C.; Weisel, J.W.; Horita, D.A.; Hantgan, R.R. Integrin Conformational Regulation: Uncoupling Extension/Tail Separation from Changes in the Head Region by a Multiresolution Approach. Structure 2008, 16, 954–964. [Google Scholar] [CrossRef] [Green Version]
  35. Yang, H.; Xu, Z.; Peng, Y.; Wang, J.; Xiang, Y. Integrin Β4 as a Potential Diagnostic and Therapeutic Tumor Marker. Biomolecules 2021, 11, 1197. [Google Scholar] [CrossRef]
  36. Goksoy, E.; Ma, Y.-Q.; Wang, X.; Kong, X.; Perera, D.; Plow, E.F.; Qin, J. Structural Basis for the Autoinhibition of Talin in Regulating Integrin Activation. Mol. Cell 2008, 31, 124–133. [Google Scholar] [CrossRef] [Green Version]
  37. Harburger, D.S.; Bouaouina, M.; Calderwood, D.A. Kindlin-1 and -2 Directly Bind the C-Terminal Region of β Integrin Cytoplasmic Tails and Exert Integrin-Specific Activation Effects. J. Biol. Chem. 2009, 284, 11485–11497. [Google Scholar] [CrossRef] [Green Version]
  38. Moser, M.; Nieswandt, B.; Ussar, S.; Pozgajova, M.; Fässler, R. Kindlin-3 Is Essential for Integrin Activation and Platelet Aggregation. Nat. Med. 2008, 14, 325–330. [Google Scholar] [CrossRef]
  39. .Calderwood, D.A.; Fujioka, Y.; de Pereda, J.M.; García-Alvarez, B.; Nakamoto, T.; Margolis, B.; McGlade, C.J.; Liddington, R.C.; Ginsberg, M.H. Integrin Beta Cytoplasmic Domain Interactions with Phosphotyrosine-Binding Domains: A Structural Prototype for Diversity in Integrin Signaling. Proc. Natl. Acad. Sci. USA 2003, 100, 2272–2277. [Google Scholar] [CrossRef] [Green Version]
  40. Greenberg, S.; Grinstein, S. Phagocytosis and Innate Immunity. Curr. Opin. Immunol. 2002, 14, 136–145. [Google Scholar] [CrossRef]
  41. Torres-Gomez, A.; Cabañas, C.; Lafuente, E.M. Phagocytic Integrins: Activation and Signaling. Front. Immunol. 2020, 11, 738. [Google Scholar] [CrossRef] [PubMed]
  42. Blythe, E.N.; Weaver, L.C.; Brown, A.; Dekaban, G.A. Β2 Integrin CD11d/CD18: From Expression to an Emerging Role in Staged Leukocyte Migration. Front. Immunol. 2021, 12, 775447. [Google Scholar] [CrossRef] [PubMed]
  43. Madamanchi, A.; Santoro, S.A.; Zutter, M.M. A2β1 Integrin. Adv. Exp. Med. Biol. 2014, 819, 41–60. [Google Scholar] [CrossRef] [PubMed]
  44. Kirveskari, J.; Bono, P.; Granfors, K.; Leirisalo-Repo, M.; Jalkanen, S.; Salmi, M. Expression of Alpha4-Integrins on Human Neutrophils. J. Leukoc. Biol. 2000, 68, 243–250. [Google Scholar]
  45. Pierini, L.M.; Lawson, M.A.; Eddy, R.J.; Hendey, B.; Maxfield, F.R. Oriented Endocytic Recycling of Alpha5beta1 in Motile Neutrophils. Blood 2000, 95, 2471–2480. [Google Scholar] [CrossRef]
  46. Bohnsack, J.F. CD11/CD18-Independent Neutrophil Adherence to Laminin Is Mediated by the Integrin VLA-6. Blood 1992, 79, 1545–1552. [Google Scholar] [CrossRef] [Green Version]
  47. Mambole, A.; Bigot, S.; Baruch, D.; Lesavre, P.; Halbwachs-Mecarelli, L. Human Neutrophil Integrin A9β1: Up-Regulation by Cell Activation and Synergy with Β2 Integrins during Adhesion to Endothelium under Flow. J. Leukoc. Biol. 2010, 88, 321–327. [Google Scholar] [CrossRef]
  48. Langereis, J.D. Neutrophil Integrin Affinity Regulation in Adhesion, Migration, and Bacterial Clearance. Cell Adhes. Migr. 2013, 7, 476–481. [Google Scholar] [CrossRef] [Green Version]
  49. Lefort, C.; Ley, K. Neutrophil Arrest by LFA-1 Activation. Front. Immunol. 2012, 3, 157. [Google Scholar] [CrossRef] [Green Version]
  50. Salas, A.; Shimaoka, M.; Kogan, A.N.; Harwood, C.; von Andrian, U.H.; Springer, T.A. Rolling Adhesion through an Extended Conformation of Integrin AlphaLbeta2 and Relation to Alpha I and Beta I-like Domain Interaction. Immunity 2004, 20, 393–406. [Google Scholar] [CrossRef] [Green Version]
  51. Chesnutt, B.C.; Smith, D.F.; Raffler, N.A.; Smith, M.L.; White, E.J.; Ley, K. Induction of LFA-1-Dependent Neutrophil Rolling on ICAM-1 by Engagement of E-Selectin. Microcirculation 2006, 13, 99–109. [Google Scholar] [CrossRef] [PubMed]
  52. Kuwano, Y.; Spelten, O.; Zhang, H.; Ley, K.; Zarbock, A. Rolling on E- or P-Selectin Induces the Extended but Not High-Affinity Conformation of LFA-1 in Neutrophils. Blood 2010, 116, 617–624. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. DiVietro, J.A.; Smith, M.J.; Smith, B.R.; Petruzzelli, L.; Larson, R.S.; Lawrence, M.B. Immobilized IL-8 Triggers Progressive Activation of Neutrophils Rolling in Vitro on P-Selectin and Intercellular Adhesion Molecule-1. J. Immunol. 2001, 167, 4017–4025. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Bai, W.; Wang, Q.; Deng, Z.; Li, T.; Xiao, H.; Wu, Z. TRAF1 Suppresses Antifungal Immunity through CXCL1-Mediated Neutrophil Recruitment during Candida Albicans Intradermal Infection. Cell Commun. Signal. 2020, 18, 30. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Gopalan, P.K.; Smith, C.W.; Lu, H.; Berg, E.L.; McIntire, L.V.; Simon, S.I. Neutrophil CD18-Dependent Arrest on Intercellular Adhesion Molecule 1 (ICAM-1) in Shear Flow Can Be Activated through L-Selectin. J. Immunol. 1997, 158, 367–375. [Google Scholar]
  56. Lyck, R.; Enzmann, G. The Physiological Roles of ICAM-1 and ICAM-2 in Neutrophil Migration into Tissues. Curr. Opin. Hematol 2015, 22, 53–59. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Phillipson, M.; Heit, B.; Colarusso, P.; Liu, L.; Ballantyne, C.M.; Kubes, P. Intraluminal Crawling of Neutrophils to Emigration Sites: A Molecularly Distinct Process from Adhesion in the Recruitment Cascade. J. Exp. Med. 2006, 203, 2569–2575. [Google Scholar] [CrossRef]
  58. Buffone, A.; Anderson, N.R.; Hammer, D.A. Human Neutrophils Will Crawl Upstream on ICAM-1 If Mac-1 Is Blocked. Biophys. J. 2019, 117, 1393–1404. [Google Scholar] [CrossRef]
  59. Proebstl, D.; Voisin, M.-B.; Woodfin, A.; Whiteford, J.; D’Acquisto, F.; Jones, G.E.; Rowe, D.; Nourshargh, S. Pericytes Support Neutrophil Subendothelial Cell Crawling and Breaching of Venular Walls in Vivo. J. Exp. Med. 2012, 209, 1219–1234. [Google Scholar] [CrossRef] [Green Version]
  60. Chong, D.L.W.; Rebeyrol, C.; José, R.J.; Williams, A.E.; Brown, J.S.; Scotton, C.J.; Porter, J.C. ICAM-1 and ICAM-2 Are Differentially Expressed and Up-Regulated on Inflamed Pulmonary Epithelium, but Neither ICAM-2 nor LFA-1: ICAM-1 Are Required for Neutrophil Migration Into the Airways In Vivo. Front. Immunol. 2021, 12, 691957. [Google Scholar] [CrossRef]
  61. Kinoshita, K.; Leung, A.; Simon, S.; Evans, E. Long-Lived, High-Strength States of ICAM-1 Bonds to Beta2 Integrin, II: Lifetimes of LFA-1 Bonds under Force in Leukocyte Signaling. Biophys. J. 2010, 98, 1467–1475. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Sarantos, M.R.; Lum, A.F.H.; Staunton, D.E.; Simon, S.I. Kinetics of LFA-1 Binding to ICAM-1 Studied in a Cell-Free System. Conf. Proc. IEEE Eng. Med. Biol. Soc. 2004, 2004, 4974–4977. [Google Scholar] [CrossRef] [PubMed]
  63. Gorina, R.; Lyck, R.; Vestweber, D.; Engelhardt, B. Β2 Integrin-Mediated Crawling on Endothelial ICAM-1 and ICAM-2 Is a Prerequisite for Transcellular Neutrophil Diapedesis across the Inflamed Blood-Brain Barrier. J. Immunol. 2014, 192, 324–337. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Li, N.; Yang, H.; Wang, M.; Lü, S.; Zhang, Y.; Long, M. Ligand-Specific Binding Forces of LFA-1 and Mac-1 in Neutrophil Adhesion and Crawling. Mol. Biol. Cell 2018, 29, 408–418. [Google Scholar] [CrossRef]
  65. Ostermann, G.; Weber, K.S.C.; Zernecke, A.; Schröder, A.; Weber, C. JAM-1 Is a Ligand of the Beta(2) Integrin LFA-1 Involved in Transendothelial Migration of Leukocytes. Nat. Immunol. 2002, 3, 151–158. [Google Scholar] [CrossRef]
  66. Yakubenko, V.P.; Cui, K.; Ardell, C.L.; Brown, K.E.; West, X.Z.; Gao, D.; Stefl, S.; Salomon, R.G.; Podrez, E.A.; Byzova, T.V. Oxidative Modifications of Extracellular Matrix Promote the Second Wave of Inflammation via Β2 Integrins. Blood 2018, 132, 78–88. [Google Scholar] [CrossRef]
  67. Chen, J.; Zhong, M.-C.; Guo, H.; Davidson, D.; Mishel, S.; Lu, Y.; Rhee, I.; Pérez-Quintero, L.-A.; Zhang, S.; Cruz-Munoz, M.-E.; et al. SLAMF7 Is Critical for Phagocytosis of Haematopoietic Tumour Cells via Mac-1 Integrin. Nature 2017, 544, 493–497. [Google Scholar] [CrossRef] [Green Version]
  68. Bose, T.O.; Colpitts, S.L.; Pham, Q.-M.; Puddington, L.; Lefrançois, L. CD11a Is Essential for Normal Development of Hematopoietic Intermediates. J. Immunol. 2014, 193, 2863–2872. [Google Scholar] [CrossRef] [Green Version]
  69. Lampiasi, N.; Russo, R.; Zito, F. The Alternative Faces of Macrophage Generate Osteoclasts. Biomed Res. Int. 2016, 2016, 9089610. [Google Scholar] [CrossRef] [Green Version]
  70. Herrero-Turrión, M.J.; Calafat, J.; Janssen, H.; Fukuda, M.; Mollinedo, F. Rab27a Regulates Exocytosis of Tertiary and Specific Granules in Human Neutrophils. J. Immunol. 2008, 181, 3793–3803. [Google Scholar] [CrossRef] [Green Version]
  71. Martín-Martín, B.; Nabokina, S.M.; Blasi, J.; Lazo, P.A.; Mollinedo, F. Involvement of SNAP-23 and Syntaxin 6 in Human Neutrophil Exocytosis. Blood 2000, 96, 2574–2583. [Google Scholar] [CrossRef] [PubMed]
  72. Mollinedo, F. Neutrophil Degranulation, Plasticity, and Cancer Metastasis. Trends Immunol. 2019, 40, 228–242. [Google Scholar] [CrossRef] [PubMed]
  73. Mollinedo, F.; Calafat, J.; Janssen, H.; Martín-Martín, B.; Canchado, J.; Nabokina, S.M.; Gajate, C. Combinatorial SNARE Complexes Modulate the Secretion of Cytoplasmic Granules in Human Neutrophils. J. Immunol. 2006, 177, 2831–2841. [Google Scholar] [CrossRef] [Green Version]
  74. Mollinedo, F.; Martín-Martín, B.; Calafat, J.; Nabokina, S.M.; Lazo, P.A. Role of Vesicle-Associated Membrane Protein-2, Through Q-Soluble N-Ethylmaleimide-Sensitive Factor Attachment Protein Receptor/R-Soluble N-Ethylmaleimide-Sensitive Factor Attachment Protein Receptor Interaction, in the Exocytosis of Specific and Tertiary Granules of Human Neutrophils. J. Immunol. 2003, 170, 1034–1042. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Ramadass, M.; Johnson, J.L.; Catz, S.D. Rab27a Regulates GM-CSF-Dependent Priming of Neutrophil Exocytosis. J. Leukoc. Biol. 2017, 101, 693–702. [Google Scholar] [CrossRef]
  76. Masuda, S.; Nakazawa, D.; Shida, H.; Miyoshi, A.; Kusunoki, Y.; Tomaru, U.; Ishizu, A. NETosis Markers: Quest for Specific, Objective, and Quantitative Markers. Clin. Chim. Acta 2016, 459, 89–93. [Google Scholar] [CrossRef]
  77. Werr, J.; Xie, X.; Hedqvist, P.; Ruoslahti, E.; Lindbom, L. Beta1 Integrins Are Critically Involved in Neutrophil Locomotion in Extravascular Tissue In Vivo. J. Exp. Med. 1998, 187, 2091–2096. [Google Scholar] [CrossRef] [Green Version]
  78. Davenpeck, K.L.; Sterbinsky, S.A.; Bochner, B.S. Rat Neutrophils Express Alpha4 and Beta1 Integrins and Bind to Vascular Cell Adhesion Molecule-1 (VCAM-1) and Mucosal Addressin Cell Adhesion Molecule-1 (MAdCAM-1). Blood 1998, 91, 2341–2346. [Google Scholar] [CrossRef]
  79. Henderson, R.B.; Lim, L.H.; Tessier, P.A.; Gavins, F.N.; Mathies, M.; Perretti, M.; Hogg, N. The Use of Lymphocyte Function-Associated Antigen (LFA)-1-Deficient Mice to Determine the Role of LFA-1, Mac-1, and Alpha4 Integrin in the Inflammatory Response of Neutrophils. J. Exp. Med. 2001, 194, 219–226. [Google Scholar] [CrossRef] [Green Version]
  80. Issekutz, T.B.; Miyasaka, M.; Issekutz, A.C. Rat Blood Neutrophils Express Very Late Antigen 4 and It Mediates Migration to Arthritic Joint and Dermal Inflammation. J. Exp. Med. 1996, 183, 2175–2184. [Google Scholar] [CrossRef] [Green Version]
  81. Canalli, A.A.; Proença, R.F.; Franco-Penteado, C.F.; Traina, F.; Sakamoto, T.M.; Saad, S.T.O.; Conran, N.; Costa, F.F. Participation of Mac-1, LFA-1 and VLA-4 Integrins in the in Vitro Adhesion of Sickle Cell Disease Neutrophils to Endothelial Layers, and Reversal of Adhesion by Simvastatin. Haematologica 2011, 96, 526–533. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Ibbotson, G.C.; Doig, C.; Kaur, J.; Gill, V.; Ostrovsky, L.; Fairhead, T.; Kubes, P. Functional Alpha4-Integrin: A Newly Identified Pathway of Neutrophil Recruitment in Critically Ill Septic Patients. Nat. Med. 2001, 7, 465–470. [Google Scholar] [CrossRef] [PubMed]
  83. Werr, J.; Johansson, J.; Eriksson, E.E.; Hedqvist, P.; Ruoslahti, E.; Lindbom, L. Integrin Alpha(2)Beta(1) (VLA-2) Is a Principal Receptor Used by Neutrophils for Locomotion in Extravascular Tissue. Blood 2000, 95, 1804–1809. [Google Scholar] [CrossRef]
  84. Dangerfield, J.; Larbi, K.Y.; Huang, M.-T.; Dewar, A.; Nourshargh, S. PECAM-1 (CD31) Homophilic Interaction up-Regulates Alpha6beta1 on Transmigrated Neutrophils in Vivo and Plays a Functional Role in the Ability of Alpha6 Integrins to Mediate Leukocyte Migration through the Perivascular Basement Membrane. J. Exp. Med. 2002, 196, 1201–1211. [Google Scholar] [CrossRef]
  85. Dangerfield, J.P.; Wang, S.; Nourshargh, S. Blockade of Alpha6 Integrin Inhibits IL-1beta- but Not TNF-Alpha-Induced Neutrophil Transmigration in Vivo. J. Leukoc. Biol. 2005, 77, 159–165. [Google Scholar] [CrossRef] [PubMed]
  86. Roussel, E.; Gingras, M.C. Transendothelial Migration Induces Rapid Expression on Neutrophils of Granule-Release VLA6 Used for Tissue Infiltration. J. Leukoc. Biol. 1997, 62, 356–362. [Google Scholar] [CrossRef] [PubMed]
  87. Ridger, V.C.; Wagner, B.E.; Wallace, W.A.; Hellewell, P.G. Differential Effects of CD18, CD29, and CD49 Integrin Subunit Inhibition on Neutrophil Migration in Pulmonary Inflammation. J. Immunol. 2001, 166, 3484–3490. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Campbell, I.D.; Humphries, M.J. Integrin Structure, Activation, and Interactions. Cold Spring Harb. Perspect. Biol. 2011, 3, a004994. [Google Scholar] [CrossRef] [Green Version]
  89. Arnaout, M.A. Biology and Structure of Leukocyte β 2 Integrins and Their Role in Inflammation. F1000Research 2016, 5, 2433. [Google Scholar] [CrossRef] [Green Version]
  90. Nishida, N.; Xie, C.; Shimaoka, M.; Cheng, Y.; Walz, T.; Springer, T.A. Activation of Leukocyte Β2 Integrins by Conversion from Bent to Extended Conformations. Immunity 2006, 25, 583–594. [Google Scholar] [CrossRef] [Green Version]
  91. Shimonaka, M.; Katagiri, K.; Nakayama, T.; Fujita, N.; Tsuruo, T.; Yoshie, O.; Kinashi, T. Rap1 Translates Chemokine Signals to Integrin Activation, Cell Polarization, and Motility across Vascular Endothelium under Flow. J. Cell Biol. 2003, 161, 417–427. [Google Scholar] [CrossRef] [PubMed]
  92. Tadokoro, S.; Shattil, S.J.; Eto, K.; Tai, V.; Liddington, R.C.; de Pereda, J.M.; Ginsberg, M.H.; Calderwood, D.A. Talin Binding to Integrin Beta Tails: A Final Common Step in Integrin Activation. Science 2003, 302, 103–106. [Google Scholar] [CrossRef] [PubMed]
  93. Vinogradova, O.; Velyvis, A.; Velyviene, A.; Hu, B.; Haas, T.; Plow, E.; Qin, J. A Structural Mechanism of Integrin Alpha(IIb)Beta(3) “inside-out” Activation as Regulated by Its Cytoplasmic Face. Cell 2002, 110, 587–597. [Google Scholar] [CrossRef] [Green Version]
  94. Liddington, R.C.; Ginsberg, M.H. Integrin Activation Takes Shape. J. Cell Biol. 2002, 158, 833–839. [Google Scholar] [CrossRef]
  95. Shimaoka, M.; Takagi, J.; Springer, T.A. Conformational Regulation of Integrin Structure and Function. Annu. Rev. Biophys. Biomol. Struct. 2002, 31, 485–516. [Google Scholar] [CrossRef] [Green Version]
  96. Xiong, J.-P.; Stehle, T.; Goodman, S.L.; Arnaout, M.A. New Insights into the Structural Basis of Integrin Activation. Blood 2003, 102, 1155–1159. [Google Scholar] [CrossRef]
  97. Sun, H.; Lagarrigue, F.; Wang, H.; Fan, Z.; Lopez-Ramirez, M.A.; Chang, J.T.; Ginsberg, M.H. Distinct Integrin Activation Pathways for Effector and Regulatory T Cell Trafficking and Function. J. Exp. Med. 2021, 218, e20201524. [Google Scholar] [CrossRef]
  98. Beglova, N.; Blacklow, S.C.; Takagi, J.; Springer, T.A. Cysteine-Rich Module Structure Reveals a Fulcrum for Integrin Rearrangement upon Activation. Nat. Struct. Biol. 2002, 9, 282–287. [Google Scholar] [CrossRef]
  99. Adair, B.D.; Xiong, J.-P.; Maddock, C.; Goodman, S.L.; Arnaout, M.A.; Yeager, M. Three-Dimensional EM Structure of the Ectodomain of Integrin {alpha}V{beta}3 in a Complex with Fibronectin. J. Cell Biol. 2005, 168, 1109–1118. [Google Scholar] [CrossRef]
  100. Sen, M.; Yuki, K.; Springer, T.A. An Internal Ligand-Bound, Metastable State of a Leukocyte Integrin, AXβ2. J. Cell Biol. 2013, 203, 629–642. [Google Scholar] [CrossRef]
  101. Gupta, V.; Gylling, A.; Alonso, J.L.; Sugimori, T.; Ianakiev, P.; Xiong, J.-P.; Amin Arnaout, M. The β-Tail Domain (ΒTD) Regulates Physiologic Ligand Binding to Integrin CD11b/CD18. Blood 2007, 109, 3513–3520. [Google Scholar] [CrossRef] [PubMed]
  102. Fan, Z.; McArdle, S.; Marki, A.; Mikulski, Z.; Gutierrez, E.; Engelhardt, B.; Deutsch, U.; Ginsberg, M.; Groisman, A.; Ley, K. Neutrophil Recruitment Limited by High-Affinity Bent Β2 Integrin Binding Ligand in Cis. Nat. Commun. 2016, 7, 12658. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Calderwood, D.A. Talin Controls Integrin Activation. BioChem. Soc. Trans. 2004, 32, 434–437. [Google Scholar] [CrossRef] [PubMed]
  104. Calderwood, D.A.; Yan, B.; de Pereda, J.M.; Alvarez, B.G.; Fujioka, Y.; Liddington, R.C.; Ginsberg, M.H. The Phosphotyrosine Binding-like Domain of Talin Activates Integrins. J. Biol. Chem. 2002, 277, 21749–21758. [Google Scholar] [CrossRef] [Green Version]
  105. Li, A.; Guo, Q.; Kim, C.; Hu, W.; Ye, F. Integrin AII b Tail Distal of GFFKR Participates in Inside-out AII b Β3 Activation. J. Thromb. Haemost 2014, 12, 1145–1155. [Google Scholar] [CrossRef] [Green Version]
  106. Oxley, C.L.; Anthis, N.J.; Lowe, E.D.; Vakonakis, I.; Campbell, I.D.; Wegener, K.L. An Integrin Phosphorylation Switch: The Effect of Beta3 Integrin Tail Phosphorylation on Dok1 and Talin Binding. J. Biol. Chem. 2008, 283, 5420–5426. [Google Scholar] [CrossRef] [Green Version]
  107. Kammerer, P.; Aretz, J.; Fässler, R. Lucky Kindlin: A Cloverleaf at the Integrin Tail. Proc. Natl. Acad. Sci. USA 2017, 114, 9234–9236. [Google Scholar] [CrossRef] [Green Version]
  108. Larjava, H.; Plow, E.F.; Wu, C. Kindlins: Essential Regulators of Integrin Signalling and Cell-Matrix Adhesion. EMBO Rep. 2008, 9, 1203–1208. [Google Scholar] [CrossRef]
  109. Kiema, T.; Lad, Y.; Jiang, P.; Oxley, C.L.; Baldassarre, M.; Wegener, K.L.; Campbell, I.D.; Ylänne, J.; Calderwood, D.A. The Molecular Basis of Filamin Binding to Integrins and Competition with Talin. Mol. Cell 2006, 21, 337–347. [Google Scholar] [CrossRef]
  110. Razinia, Z.; Mäkelä, T.; Ylänne, J.; Calderwood, D.A. Filamins in Mechanosensing and Signaling. Annu. Rev. Biophys. 2012, 41, 227–246. [Google Scholar] [CrossRef] [Green Version]
  111. Gupta, S.; Chit, J.C.-Y.; Feng, C.; Bhunia, A.; Tan, S.-M.; Bhattacharjya, S. An Alternative Phosphorylation Switch in Integrin Β2 (CD18) Tail for Dok1 Binding. Sci. Rep. 2015, 5, 11630. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Takala, H.; Nurminen, E.; Nurmi, S.M.; Aatonen, M.; Strandin, T.; Takatalo, M.; Kiema, T.; Gahmberg, C.G.; Ylänne, J.; Fagerholm, S.C. Beta2 Integrin Phosphorylation on Thr758 Acts as a Molecular Switch to Regulate 14-3-3 and Filamin Binding. Blood 2008, 112, 1853–1862. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Jahan, F.; Madhavan, S.; Rolova, T.; Viazmina, L.; Grönholm, M.; Gahmberg, C.G. Phosphorylation of the α-Chain in the Integrin LFA-1 Enables Β2-Chain Phosphorylation and α-Actinin Binding Required for Cell Adhesion. J. Biol. Chem. 2018, 293, 12318–12330. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Bromberger, T.; Klapproth, S.; Rohwedder, I.; Weber, J.; Pick, R.; Mittmann, L.; Min-Weißenhorn, S.J.; Reichel, C.A.; Scheiermann, C.; Sperandio, M.; et al. Binding of Rap1 and Riam to Talin1 Fine-Tune Β2 Integrin Activity During Leukocyte Trafficking. Front. Immunol. 2021, 12, 702345. [Google Scholar] [CrossRef] [PubMed]
  115. Lee, H.-S.; Lim, C.J.; Puzon-McLaughlin, W.; Shattil, S.J.; Ginsberg, M.H. RIAM Activates Integrins by Linking Talin to Ras GTPase Membrane-Targeting Sequences. J. Biol. Chem. 2009, 284, 5119–5127. [Google Scholar] [CrossRef] [Green Version]
  116. Bos, J.L.; de Rooij, J.; Reedquist, K.A. Rap1 Signalling: Adhering to New Models. Nat. Rev. Mol. Cell Biol. 2001, 2, 369–377. [Google Scholar] [CrossRef]
  117. Bos, J.L.; de Bruyn, K.; Enserink, J.; Kuiperij, B.; Rangarajan, S.; Rehmann, H.; Riedl, J.; de Rooij, J.; van Mansfeld, F.; Zwartkruis, F. The Role of Rap1 in Integrin-Mediated Cell Adhesion. Biochem. Soc. Trans. 2003, 31, 83–86. [Google Scholar] [CrossRef]
  118. Bromberger, T.; Klapproth, S.; Rohwedder, I.; Zhu, L.; Mittmann, L.; Reichel, C.A.; Sperandio, M.; Qin, J.; Moser, M. Direct Rap1/Talin1 Interaction Regulates Platelet and Neutrophil Integrin Activity in Mice. Blood 2018, 132, 2754–2762. [Google Scholar] [CrossRef] [Green Version]
  119. Klapproth, S.; Moretti, F.A.; Zeiler, M.; Ruppert, R.; Breithaupt, U.; Mueller, S.; Haas, R.; Mann, M.; Sperandio, M.; Fässler, R.; et al. Minimal Amounts of Kindlin-3 Suffice for Basal Platelet and Leukocyte Functions in Mice. Blood 2015, 126, 2592–2600. [Google Scholar] [CrossRef] [Green Version]
  120. Ghiran, I.; Klickstein, L.B.; Nicholson-Weller, A. Calreticulin Is at the Surface of Circulating Neutrophils and Uses CD59 as an Adaptor Molecule. J. Biol. Chem. 2003, 278, 21024–21031. [Google Scholar] [CrossRef] [Green Version]
  121. Ohkuro, M.; Kim, J.-D.; Kuboi, Y.; Hayashi, Y.; Mizukami, H.; Kobayashi-Kuramochi, H.; Muramoto, K.; Shirato, M.; Michikawa-Tanaka, F.; Moriya, J.; et al. Calreticulin and Integrin Alpha Dissociation Induces Anti-Inflammatory Programming in Animal Models of Inflammatory Bowel Disease. Nat. Commun. 2018, 9, 1982. [Google Scholar] [CrossRef] [PubMed]
  122. Kinashi, T.; Katagiri, K. Regulation of Immune Cell Adhesion and Migration by Regulator of Adhesion and Cell Polarization Enriched in Lymphoid Tissues. Immunology 2005, 116, 164–171. [Google Scholar] [CrossRef] [PubMed]
  123. Hyduk, S.J.; Oh, J.; Xiao, H.; Chen, M.; Cybulsky, M.I. Paxillin Selectively Associates with Constitutive and Chemoattractant-Induced High-Affinity Alpha4beta1 Integrins: Implications for Integrin Signaling. Blood 2004, 104, 2818–2824. [Google Scholar] [CrossRef] [PubMed]
  124. Gao, J.; Bao, Y.; Ge, S.; Sun, P.; Sun, J.; Liu, J.; Chen, F.; Han, L.; Cao, Z.; Qin, J.; et al. Sharpin Suppresses Β1-Integrin Activation by Complexing with the Β1 Tail and Kindlin-1. Cell Commun. Signal. 2019, 17, 101. [Google Scholar] [CrossRef] [Green Version]
  125. Roberts, G.C.K.; Critchley, D.R. Structural and Biophysical Properties of the Integrin-Associated Cytoskeletal Protein Talin. Biophys. Rev. 2009, 1, 61–69. [Google Scholar] [CrossRef] [Green Version]
  126. Monkley, S.J.; Pritchard, C.A.; Critchley, D.R. Analysis of the Mammalian Talin2 Gene TLN2. Biochem. Biophys. Res. Commun. 2001, 286, 880–885. [Google Scholar] [CrossRef]
  127. Senetar, M.A.; Moncman, C.L.; McCann, R.O. Talin2 Is Induced during Striated Muscle Differentiation and Is Targeted to Stable Adhesion Complexes in Mature Muscle. Cell Motil. Cytoskelet. 2007, 64, 157–173. [Google Scholar] [CrossRef]
  128. Lefort, C.T.; Rossaint, J.; Moser, M.; Petrich, B.G.; Zarbock, A.; Monkley, S.J.; Critchley, D.R.; Ginsberg, M.H.; Fässler, R.; Ley, K. Distinct Roles for Talin-1 and Kindlin-3 in LFA-1 Extension and Affinity Regulation. Blood 2012, 119, 4275–4282. [Google Scholar] [CrossRef]
  129. Lim, J.; Wiedemann, A.; Tzircotis, G.; Monkley, S.J.; Critchley, D.R.; Caron, E. An Essential Role for Talin during AMβ2-Mediated Phagocytosis. Mol. Biol. Cell 2007, 18, 976–985. [Google Scholar] [CrossRef] [Green Version]
  130. Calderwood, D.A.; Campbell, I.D.; Critchley, D.R. Talins and Kindlins: Partners in Integrin-Mediated Adhesion. Nat. Rev. Mol. Cell Biol. 2013, 14, 503–517. [Google Scholar] [CrossRef] [Green Version]
  131. García-Alvarez, B.; de Pereda, J.M.; Calderwood, D.A.; Ulmer, T.S.; Critchley, D.; Campbell, I.D.; Ginsberg, M.H.; Liddington, R.C. Structural Determinants of Integrin Recognition by Talin. Mol. Cell 2003, 11, 49–58. [Google Scholar] [CrossRef]
  132. Hemmings, L.; Rees, D.J.; Ohanian, V.; Bolton, S.J.; Gilmore, A.P.; Patel, B.; Priddle, H.; Trevithick, J.E.; Hynes, R.O.; Critchley, D.R. Talin Contains Three Actin-Binding Sites Each of Which Is Adjacent to a Vinculin-Binding Site. J. Cell Sci. 1996, 109, 2715–2726. [Google Scholar] [CrossRef] [PubMed]
  133. Hynes, R.O. Integrins: Bidirectional, Allosteric Signaling Machines. Cell 2002, 110, 673–687. [Google Scholar] [CrossRef] [Green Version]
  134. Kim, C.; Ye, F.; Ginsberg, M.H. Regulation of Integrin Activation. Annu. Rev. Cell Dev. Biol. 2011, 27, 321–345. [Google Scholar] [CrossRef]
  135. Shattil, S.J.; Kim, C.; Ginsberg, M.H. The Final Steps of Integrin Activation: The End Game. Nat. Rev. Mol. Cell Biol. 2010, 11, 288–300. [Google Scholar] [CrossRef] [Green Version]
  136. Wegener, K.L.; Partridge, A.W.; Han, J.; Pickford, A.R.; Liddington, R.C.; Ginsberg, M.H.; Campbell, I.D. Structural Basis of Integrin Activation by Talin. Cell 2007, 128, 171–182. [Google Scholar] [CrossRef]
  137. Yago, T.; Petrich, B.G.; Zhang, N.; Liu, Z.; Shao, B.; Ginsberg, M.H.; McEver, R.P. Blocking Neutrophil Integrin Activation Prevents Ischemia–Reperfusion Injury. J. Exp. Med. 2015, 212, 1267–1281. [Google Scholar] [CrossRef]
  138. Lagarrigue, F.; Kim, C.; Ginsberg, M.H. The Rap1-RIAM-Talin Axis of Integrin Activation and Blood Cell Function. Blood 2016, 128, 479–487. [Google Scholar] [CrossRef] [Green Version]
  139. Lagarrigue, F.; Gingras, A.R.; Paul, D.S.; Valadez, A.J.; Cuevas, M.N.; Sun, H.; Lopez-Ramirez, M.A.; Goult, B.T.; Shattil, S.J.; Bergmeier, W.; et al. Rap1 Binding to the Talin 1 F0 Domain Makes a Minimal Contribution to Murine Platelet GPIIb-IIIa Activation. Blood Adv. 2018, 2, 2358–2368. [Google Scholar] [CrossRef] [Green Version]
  140. Lagarrigue, F.; Paul, D.S.; Gingras, A.R.; Valadez, A.J.; Sun, H.; Lin, J.; Cuevas, M.N.; Ablack, J.N.; Lopez-Ramirez, M.A.; Bergmeier, W.; et al. Talin-1 Is the Principal Platelet Rap1 Effector of Integrin Activation. Blood 2020, 136, 1180–1190. [Google Scholar] [CrossRef] [PubMed]
  141. Sun, H.; Lagarrigue, F.; Ginsberg, M.H. The Connection Between Rap1 and Talin1 in the Activation of Integrins in Blood Cells. Front. Cell Dev. Biol. 2022, 10. [Google Scholar] [CrossRef] [PubMed]
  142. Moser, M.; Legate, K.R.; Zent, R.; Fässler, R. The Tail of Integrins, Talin, and Kindlins. Science 2009, 324, 895–899. [Google Scholar] [CrossRef] [PubMed]
  143. Moore, D.T.; Nygren, P.; Jo, H.; Boesze-Battaglia, K.; Bennett, J.S.; DeGrado, W.F. Affinity of Talin-1 for the Β3-Integrin Cytosolic Domain Is Modulated by Its Phospholipid Bilayer Environment. Proc. Natl. Acad. Sci. USA 2012, 109, 793–798. [Google Scholar] [CrossRef] [Green Version]
  144. Gingras, A.R.; Ziegler, W.H.; Bobkov, A.A.; Joyce, M.G.; Fasci, D.; Himmel, M.; Rothemund, S.; Ritter, A.; Grossmann, J.G.; Patel, B.; et al. Structural Determinants of Integrin Binding to the Talin Rod*. J. Biol. Chem. 2009, 284, 8866–8876. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Moes, M.; Rodius, S.; Coleman, S.J.; Monkley, S.J.; Goormaghtigh, E.; Tremuth, L.; Kox, C.; van der Holst, P.P.G.; Critchley, D.R.; Kieffer, N. The Integrin Binding Site 2 (IBS2) in the Talin Rod Domain Is Essential for Linking Integrin β Subunits to the Cytoskeleton*. J. Biol. Chem. 2007, 282, 17280–17288. [Google Scholar] [CrossRef] [Green Version]
  146. Rodius, S.; Chaloin, O.; Moes, M.; Schaffner-Reckinger, E.; Landrieu, I.; Lippens, G.; Lin, M.; Zhang, J.; Kieffer, N. The Talin Rod IBS2 α-Helix Interacts with the Β3 Integrin Cytoplasmic Tail Membrane-Proximal Helix by Establishing Charge Complementary Salt Bridges*. J. Biol. Chem. 2008, 283, 24212–24223. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Haage, A.; Goodwin, K.; Whitewood, A.; Camp, D.; Bogutz, A.; Turner, C.T.; Granville, D.J.; Lefebvre, L.; Plotnikov, S.; Goult, B.T.; et al. Talin Autoinhibition Regulates Cell-ECM Adhesion Dynamics and Wound Healing In Vivo. Cell Rep. 2018, 25, 2401–2416.e5. [Google Scholar] [CrossRef] [Green Version]
  148. Goult, B.T.; Xu, X.-P.; Gingras, A.R.; Swift, M.; Patel, B.; Bate, N.; Kopp, P.M.; Barsukov, I.L.; Critchley, D.R.; Volkmann, N.; et al. Structural Studies on Full-Length Talin1 Reveal a Compact Auto-Inhibited Dimer: Implications for Talin Activation. J. Struct. Biol. 2013, 184, 21–32. [Google Scholar] [CrossRef] [Green Version]
  149. Saltel, F.; Mortier, E.; Hytönen, V.P.; Jacquier, M.-C.; Zimmermann, P.; Vogel, V.; Liu, W.; Wehrle-Haller, B. New PI(4,5)P2- and Membrane Proximal Integrin-Binding Motifs in the Talin Head Control Beta3-Integrin Clustering. J. Cell Biol. 2009, 187, 715–731. [Google Scholar] [CrossRef] [Green Version]
  150. Fong, K.P.; Molnar, K.S.; Agard, N.; Litvinov, R.I.; Kim, O.V.; Wells, J.A.; Weisel, J.W.; DeGrado, W.F.; Bennett, J.S. Cleavage of Talin by Calpain Promotes Platelet-Mediated Fibrin Clot Contraction. Blood Adv. 2021, 5, 4901–4909. [Google Scholar] [CrossRef]
  151. Huang, C.; Rajfur, Z.; Yousefi, N.; Chen, Z.; Jacobson, K.; Ginsberg, M.H. Talin Phosphorylation by Cdk5 Regulates Smurf1-Mediated Talin Head Ubiquitylation and Cell Migration. Nat. Cell Biol. 2009, 11, 624–630. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Bouaouina, M.; Lad, Y.; Calderwood, D.A. The N-Terminal Domains of Talin Cooperate with the Phosphotyrosine Binding-like Domain to Activate Beta1 and Beta3 Integrins. J. Biol. Chem. 2008, 283, 6118–6125. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Goult, B.T.; Bouaouina, M.; Elliott, P.R.; Bate, N.; Patel, B.; Gingras, A.R.; Grossmann, J.G.; Roberts, G.C.K.; Calderwood, D.A.; Critchley, D.R.; et al. Structure of a Double Ubiquitin-like Domain in the Talin Head: A Role in Integrin Activation. EMBO J. 2010, 29, 1069–1080. [Google Scholar] [CrossRef] [PubMed]
  154. Chang, Y.-C.; Su, W.; Cho, E.-A.; Zhang, H.; Huang, Q.; Philips, M.R.; Wu, J. Molecular Basis for Autoinhibition of RIAM Regulated by FAK in Integrin Activation. Proc. Natl. Acad. Sci. USA 2019, 116, 3524–3529. [Google Scholar] [CrossRef] [Green Version]
  155. Cho, E.-A.; Zhang, P.; Kumar, V.; Kavalchuk, M.; Zhang, H.; Huang, Q.; Duncan, J.S.; Wu, J. Phosphorylation of RIAM by Src Promotes Integrin Activation by Unmasking the PH Domain of RIAM. Structure 2021, 29, 320–329.e4. [Google Scholar] [CrossRef]
  156. Goult, B.T.; Zacharchenko, T.; Bate, N.; Tsang, R.; Hey, F.; Gingras, A.R.; Elliott, P.R.; Roberts, G.C.K.; Ballestrem, C.; Critchley, D.R.; et al. RIAM and Vinculin Binding to Talin Are Mutually Exclusive and Regulate Adhesion Assembly and Turnover. J. Biol. Chem. 2013, 288, 8238–8249. [Google Scholar] [CrossRef] [Green Version]
  157. Lee, H.-S.; Anekal, P.; Lim, C.J.; Liu, C.-C.; Ginsberg, M.H. Two Modes of Integrin Activation Form a Binary Molecular Switch in Adhesion Maturation. Mol. Biol. Cell 2013, 24, 1354–1362. [Google Scholar] [CrossRef]
  158. Watanabe, N.; Bodin, L.; Pandey, M.; Krause, M.; Coughlin, S.; Boussiotis, V.A.; Ginsberg, M.H.; Shattil, S.J. Mechanisms and Consequences of Agonist-Induced Talin Recruitment to Platelet Integrin AlphaIIbbeta3. J. Cell Biol. 2008, 181, 1211–1222. [Google Scholar] [CrossRef] [Green Version]
  159. Stritt, S.; Wolf, K.; Lorenz, V.; Vögtle, T.; Gupta, S.; Bösl, M.R.; Nieswandt, B. Rap1-GTP–Interacting Adaptor Molecule (RIAM) Is Dispensable for Platelet Integrin Activation and Function in Mice. Blood 2015, 125, 219–222. [Google Scholar] [CrossRef] [Green Version]
  160. Su, W.; Wynne, J.; Pinheiro, E.M.; Strazza, M.; Mor, A.; Montenont, E.; Berger, J.; Paul, D.S.; Bergmeier, W.; Gertler, F.B.; et al. Rap1 and Its Effector RIAM Are Required for Lymphocyte Trafficking. Blood 2015, 126, 2695–2703. [Google Scholar] [CrossRef] [Green Version]
  161. Ménasché, G.; Kliche, S.; Chen, E.J.H.; Stradal, T.E.B.; Schraven, B.; Koretzky, G. RIAM Links the ADAP/SKAP-55 Signaling Module to Rap1, Facilitating T-Cell-Receptor-Mediated Integrin Activation. Mol. Cell Biol. 2007, 27, 4070–4081. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Bos, J.L. Ras-like GTPases. Biochim. Biophys. Acta 1997, 1333, M19–M31. [Google Scholar] [CrossRef]
  163. Stork, P.J.S.; Dillon, T.J. Multiple Roles of Rap1 in Hematopoietic Cells: Complementary versus Antagonistic Functions. Blood 2005, 106, 2952–2961. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Polakis, P.G.; Rubinfeld, B.; Evans, T.; McCormick, F. Purification of a Plasma Membrane-Associated GTPase-Activating Protein Specific for Rap1/Krev-1 from HL60 Cells. Proc. Natl. Acad. Sci. USA 1991, 88, 239–243. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Kurachi, H.; Wada, Y.; Tsukamoto, N.; Maeda, M.; Kubota, H.; Hattori, M.; Iwai, K.; Minato, N. Human SPA-1 Gene Product Selectively Expressed in Lymphoid Tissues Is a Specific GTPase-Activating Protein for Rap1 and Rap2. Segregate Expression Profiles from a Rap1GAP Gene Product. J. Biol. Chem. 1997, 272, 28081–28088. [Google Scholar] [CrossRef] [Green Version]
  166. Frische, E.W.; Zwartkruis, F.J.T. Rap1, a Mercenary among the Ras-like GTPases. Dev. Biol. 2010, 340, 1–9. [Google Scholar] [CrossRef] [Green Version]
  167. Jaśkiewicz, A.; Pająk, B.; Orzechowski, A. The Many Faces of Rap1 GTPase. Int. J. Mol. Sci. 2018, 19, 2848. [Google Scholar] [CrossRef] [Green Version]
  168. Schwamborn, J.C.; Püschel, A.W. The Sequential Activity of the GTPases Rap1B and Cdc42 Determines Neuronal Polarity. Nat. Neurosci. 2004, 7, 923–929. [Google Scholar] [CrossRef]
  169. Bivona, T.G.; Wiener, H.H.; Ahearn, I.M.; Silletti, J.; Chiu, V.K.; Philips, M.R. Rap1 Up-Regulation and Activation on Plasma Membrane Regulates T Cell Adhesion. J. Cell Biol. 2004, 164, 461–470. [Google Scholar] [CrossRef] [Green Version]
  170. Bromberger, T.; Zhu, L.; Klapproth, S.; Qin, J.; Moser, M. Rap1 and Membrane Lipids Cooperatively Recruit Talin to Trigger Integrin Activation. J. Cell Sci. 2019, 132, jcs235531. [Google Scholar] [CrossRef]
  171. Hogan, C.; Serpente, N.; Cogram, P.; Hosking, C.R.; Bialucha, C.U.; Feller, S.M.; Braga, V.M.M.; Birchmeier, W.; Fujita, Y. Rap1 Regulates the Formation of E-Cadherin-Based Cell-Cell Contacts. Mol. Cell. Biol. 2004. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Sawant, K.; Chen, Y.; Kotian, N.; Preuss, K.M.; McDonald, J.A. Rap1 GTPase Promotes Coordinated Collective Cell Migration in Vivo. Mol. Biol. Cell 2018, 29, 2656–2673. [Google Scholar] [CrossRef] [PubMed]
  173. Chang, Y.-C.; Wu, J.-W.; Hsieh, Y.-C.; Huang, T.-H.; Liao, Z.-M.; Huang, Y.-S.; Mondo, J.A.; Montell, D.; Jang, A.C.-C. Rap1 Negatively Regulates the Hippo Pathway to Polarize Directional Protrusions in Collective Cell Migration. Cell Rep. 2018, 22, 2160–2175. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Zhang, X.; Liu, Z.; Liu, X.; Wang, S.; Zhang, Y.; He, X.; Sun, S.; Ma, S.; Shyh-Chang, N.; Liu, F.; et al. Telomere-Dependent and Telomere-Independent Roles of RAP1 in Regulating Human Stem Cell Homeostasis. Protein Cell 2019, 10, 649–667. [Google Scholar] [CrossRef] [Green Version]
  175. Camp, D.; Haage, A.; Solianova, V.; Castle, W.M.; Xu, Q.A.; Lostchuck, E.; Goult, B.T.; Tanentzapf, G. Direct Binding of Talin to Rap1 Is Required for Cell-ECM Adhesion in Drosophila. J. Cell Sci. 2018, 131, jcs225144. [Google Scholar] [CrossRef] [Green Version]
  176. Gingras, A.R.; Lagarrigue, F.; Cuevas, M.N.; Valadez, A.J.; Zorovich, M.; McLaughlin, W.; Lopez-Ramirez, M.A.; Seban, N.; Ley, K.; Kiosses, W.B.; et al. Rap1 Binding and a Lipid-Dependent Helix in Talin F1 Domain Promote Integrin Activation in Tandem. J. Cell Biol. 2019, 218, 1799–1809. [Google Scholar] [CrossRef] [Green Version]
  177. Stefanini, L.; Lee, R.H.; Paul, D.S.; O’Shaughnessy, E.C.; Ghalloussi, D.; Jones, C.I.; Boulaftali, Y.; Poe, K.O.; Piatt, R.; Kechele, D.O.; et al. Functional Redundancy between RAP1 Isoforms in Murine Platelet Production and Function. Blood 2018, 132, 1951–1962. [Google Scholar] [CrossRef] [Green Version]
  178. Lagarrigue, F.; Tan, B.; Du, Q.; Fan, Z.; Lopez-Ramirez, M.A.; Gingras, A.R.; Wang, H.; Qi, W.; Sun, H. Direct Binding of Rap1 to Talin1 and to MRL Proteins Promotes Integrin Activation in CD4+ T Cells. J. Immunol. 2022, 208, 1378–1388. [Google Scholar] [CrossRef]
  179. Lozano, M.L.; Cook, A.; Bastida, J.M.; Paul, D.S.; Iruin, G.; Cid, A.R.; Adan-Pedroso, R.; Ramón González-Porras, J.; Hernández-Rivas, J.M.; Fletcher, S.J.; et al. Novel Mutations in RASGRP2, Which Encodes CalDAG-GEFI, Abrogate Rap1 Activation, Causing Platelet Dysfunction. Blood 2016, 128, 1282–1289. [Google Scholar] [CrossRef] [Green Version]
  180. Bergmeier, W.; Goerge, T.; Wang, H.-W.; Crittenden, J.R.; Baldwin, A.C.W.; Cifuni, S.M.; Housman, D.E.; Graybiel, A.M.; Wagner, D.D. Mice Lacking the Signaling Molecule CalDAG-GEFI Represent a Model for Leukocyte Adhesion Deficiency Type III. J. Clin. Investig. 2007, 117, 1699–1707. [Google Scholar] [CrossRef] [Green Version]
  181. Côte, M.; Fos, C.; Canonigo-Balancio, A.J.; Ley, K.; Bécart, S.; Altman, A. SLAT Promotes TCR-Mediated, Rap1-Dependent LFA-1 Activation and Adhesion through Interaction of Its PH Domain with Rap1. J. Cell Sci. 2015, 128, 4341–4352. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Hao, L.; Marshall, A.J.; Liu, L. Suppressive Role of Bam32/DAPP1 in Chemokine-Induced Neutrophil Recruitment. Int. J. Mol. Sci. 2021, 22, 1825. [Google Scholar] [CrossRef] [PubMed]
  183. Lee, H.-S.; Sun, H.; Lagarrigue, F.; Fox, J.W.; Sherman, N.E.; Gingras, A.R.; Ginsberg, M.H. Phostensin Enables Lymphocyte Integrin Activation and Population of Peripheral Lymphoid Organs. bioRxiv 2021. [Google Scholar] [CrossRef]
  184. Lin, Y.-S.; Huang, K.-Y.; Wang, T.-F.; Huang, H.; Yu, H.-C.; Yen, J.; Hung, S.; Liu, S.-Q.; Lai, N.-S.; Huang, H. Immunolocalization of Phostensin in Lymphatic Cells and Tissues. J. Histochem. Cytochem. 2011, 59, 741–749. [Google Scholar] [CrossRef] [Green Version]
  185. Ussar, S.; Wang, H.-V.; Linder, S.; Fässler, R.; Moser, M. The Kindlins: Subcellular Localization and Expression during Murine Development. Exp. Cell Res. 2006, 312, 3142–3151. [Google Scholar] [CrossRef]
  186. Bialkowska, K.; Ma, Y.-Q.; Bledzka, K.; Sossey-Alaoui, K.; Izem, L.; Zhang, X.; Malinin, N.; Qin, J.; Byzova, T.; Plow, E.F. The Integrin Co-Activator Kindlin-3 Is Expressed and Functional in a Non-Hematopoietic Cell, the Endothelial Cell. J. Biol. Chem. 2010, 285, 18640–18649. [Google Scholar] [CrossRef] [Green Version]
  187. Rognoni, E.; Ruppert, R.; Fässler, R. The Kindlin Family: Functions, Signaling Properties and Implications for Human Disease. J. Cell Sci. 2016, 129, 17–27. [Google Scholar] [CrossRef] [Green Version]
  188. Moser, M.; Bauer, M.; Schmid, S.; Ruppert, R.; Schmidt, S.; Sixt, M.; Wang, H.-V.; Sperandio, M.; Fässler, R. Kindlin-3 Is Required for Beta2 Integrin-Mediated Leukocyte Adhesion to Endothelial Cells. Nat. Med. 2009, 15, 300–305. [Google Scholar] [CrossRef]
  189. Svensson, L.; Howarth, K.; McDowall, A.; Patzak, I.; Evans, R.; Ussar, S.; Moser, M.; Metin, A.; Fried, M.; Tomlinson, I.; et al. Leukocyte Adhesion Deficiency-III Is Caused by Mutations in KINDLIN3 Affecting Integrin Activation. Nat. Med. 2009, 15, 306–312. [Google Scholar] [CrossRef] [Green Version]
  190. Malinin, N.L.; Zhang, L.; Choi, J.; Ciocea, A.; Razorenova, O.; Ma, Y.-Q.; Podrez, E.A.; Tosi, M.; Lennon, D.P.; Caplan, A.I.; et al. A Point Mutation in KINDLIN3 Ablates Activation of Three Integrin Subfamilies in Humans. Nat. Med. 2009, 15, 313–318. [Google Scholar] [CrossRef] [Green Version]
  191. Dixit, N.; Kim, M.-H.; Rossaint, J.; Yamayoshi, I.; Zarbock, A.; Simon, S.I. Leukocyte Function Antigen-1, Kindlin-3, and Calcium Flux Orchestrate Neutrophil Recruitment during Inflammation. J. Immunol. 2012, 189, 5954–5964. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Fagerholm, S.C.; Lek, H.S.; Morrison, V.L. Kindlin-3 in the Immune System. Am. J. Clin. Exp. Immunol. 2014, 3, 37–42. [Google Scholar] [PubMed]
  193. Karaköse, E.; Schiller, H.B.; Fässler, R. The Kindlins at a Glance. J. Cell Sci. 2010, 123, 2353–2356. [Google Scholar] [CrossRef] [Green Version]
  194. Yates, L.A.; Füzéry, A.K.; Bonet, R.; Campbell, I.D.; Gilbert, R.J.C. Biophysical Analysis of Kindlin-3 Reveals an Elongated Conformation and Maps Integrin Binding to the Membrane-Distal β-Subunit NPXY Motif. J. Biol. Chem. 2012, 287, 37715–37731. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Morrison, V.L.; MacPherson, M.; Savinko, T.; San Lek, H.; Prescott, A.; Fagerholm, S.C. The Β2 Integrin–Kindlin-3 Interaction Is Essential for T-Cell Homing but Dispensable. for T-Cell Activation in Vivo. Blood 2013, 122, 1428–1436. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Xu, Z.; Chen, X.; Zhi, H.; Gao, J.; Bialkowska, K.; Byzova, T.V.; Pluskota, E.; White, G.C.; Liu, J.; Plow, E.F.; et al. Direct Interaction of Kindlin-3 With Integrin AIIbβ3 in Platelets Is Required for Supporting Arterial Thrombosis in Mice. Arterioscler. Thromb. Vasc. Biol. 2014, 34, 1961–1967. [Google Scholar] [CrossRef] [Green Version]
  197. Xu, Z.; Cai, J.; Gao, J.; White, G.C.; Chen, F.; Ma, Y.-Q. Interaction of Kindlin-3 and Β2-Integrins Differentially Regulates Neutrophil Recruitment and NET Release in Mice. Blood 2015, 126, 373–377. [Google Scholar] [CrossRef] [Green Version]
  198. Xu, Z.; Ni, B.; Cao, Z.; Zielonka, J.; Gao, J.; Chen, F.; Kalyanaraman, B.; White, G.C.; Ma, Y.-Q. Kindlin-3 Negatively Regulates the Release of Neutrophil Extracellular Traps. J. Leukoc. Biol. 2018, 104, 597–602. [Google Scholar] [CrossRef]
  199. Klapproth, S.; Bromberger, T.; Türk, C.; Krüger, M.; Moser, M. A Kindlin-3–Leupaxin–Paxillin Signaling Pathway Regulates Podosome Stability. J. Cell Biol. 2019, 218, 3436–3454. [Google Scholar] [CrossRef] [Green Version]
  200. Plow, E.F.; Qin, J. The Kindlin Family of Adapter Proteins: A Past, Present and Future Prospectus. Circ. Res. 2019, 124, 202–204. [Google Scholar] [CrossRef]
  201. Wen, L.; Marki, A.; Roy, P.; McArdle, S.; Sun, H.; Fan, Z.; Gingras, A.R.; Ginsberg, M.H.; Ley, K. Kindlin-3 Recruitment to the Plasma Membrane Precedes High-Affinity Β2-Integrin and Neutrophil Arrest from Rolling. Blood 2021, 137, 29–38. [Google Scholar] [CrossRef] [PubMed]
  202. Feng, C.; Li, Y.-F.; Yau, Y.-H.; Lee, H.-S.; Tang, X.-Y.; Xue, Z.-H.; Zhou, Y.-C.; Lim, W.-M.; Cornvik, T.C.; Ruedl, C.; et al. Kindlin-3 Mediates Integrin ALβ2 Outside-in Signaling, and It Interacts with Scaffold Protein Receptor for Activated-C Kinase 1 (RACK1). J. Biol. Chem. 2012, 287, 10714–10726. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Bouaouina, M.; Goult, B.T.; Huet-Calderwood, C.; Bate, N.; Brahme, N.N.; Barsukov, I.L.; Critchley, D.R.; Calderwood, D.A. A Conserved Lipid-Binding Loop in the Kindlin FERM F1 Domain Is Required for Kindlin-Mediated AIIbβ3 Integrin Coactivation*. J. Biol. Chem. 2012, 287, 6979–6990. [Google Scholar] [CrossRef] [Green Version]
  204. Chua, G.-L.; Tan, S.-M.; Bhattacharjya, S. NMR Characterization and Membrane Interactions of the Loop Region of Kindlin-3 F1 Subdomain. PLoS ONE 2016, 11, e0153501. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Orré, T.; Joly, A.; Karatas, Z.; Kastberger, B.; Cabriel, C.; Böttcher, R.T.; Lévêque-Fort, S.; Sibarita, J.-B.; Fässler, R.; Wehrle-Haller, B.; et al. Molecular Motion and Tridimensional Nanoscale Localization of Kindlin Control Integrin Activation in Focal Adhesions. Nat. Commun. 2021, 12, 3104. [Google Scholar] [CrossRef]
  206. Margraf, A.; Germena, G.; Drexler, H.C.A.; Rossaint, J.; Ludwig, N.; Prystaj, B.; Mersmann, S.; Thomas, K.; Block, H.; Gottschlich, W.; et al. The Integrin-Linked Kinase Is Required for Chemokine-Triggered High-Affinity Conformation of the Neutrophil Β2-Integrin LFA-1. Blood 2020, 136, 2200–2205. [Google Scholar] [CrossRef]
  207. Shao, B.; Yago, T.; Coghill, P.A.; Klopocki, A.G.; Mehta-D’souza, P.; Schmidtke, D.W.; Rodgers, W.; McEver, R.P. Signal-Dependent Slow Leukocyte Rolling Does Not Require Cytoskeletal Anchorage of P-Selectin Glycoprotein Ligand-1 (PSGL-1) or Integrin ALβ2. J. Biol. Chem. 2012, 287, 19585–19598. [Google Scholar] [CrossRef] [Green Version]
  208. Critchley, D.R. Biochemical and Structural Properties of the Integrin-Associated Cytoskeletal Protein Talin. Annu. Rev. Biophys. 2009, 38, 235–254. [Google Scholar] [CrossRef]
  209. Zhu, L.; Yang, J.; Bromberger, T.; Holly, A.; Lu, F.; Liu, H.; Sun, K.; Klapproth, S.; Hirbawi, J.; Byzova, T.V.; et al. Structure of Rap1b Bound to Talin Reveals a Pathway for Triggering Integrin Activation. Nat. Commun. 2017, 8, 1744. [Google Scholar] [CrossRef] [Green Version]
  210. Bledzka, K.; Bialkowska, K.; Sossey-Alaoui, K.; Vaynberg, J.; Pluskota, E.; Qin, J.; Plow, E.F. Kindlin-2 Directly Binds Actin and Regulates Integrin Outside-in Signaling. J. Cell Biol. 2016, 213, 97–108. [Google Scholar] [CrossRef] [Green Version]
  211. Patsoukis, N.; Bardhan, K.; Weaver, J.D.; Sari, D.; Torres-Gomez, A.; Li, L.; Strauss, L.; Lafuente, E.M.; Boussiotis, V.A. The Adaptor Molecule RIAM Integrates Signaling Events Critical for Integrin-Mediated Control of Immune Function and Cancer Progression. Sci. Signal. 2017, 10, eaam8298. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Otey, C.A.; Pavalko, F.M.; Burridge, K. An Interaction between Alpha-Actinin and the Beta 1 Integrin Subunit in Vitro. J. Cell Biol. 1990, 111, 721–729. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Carpén, O.; Pallai, P.; Staunton, D.E.; Springer, T.A. Association of Intercellular Adhesion Molecule-1 (ICAM-1) with Actin-Containing Cytoskeleton and Alpha-Actinin. J. Cell Biol. 1992, 118, 1223–1234. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. Roca-Cusachs, P.; del Rio, A.; Puklin-Faucher, E.; Gauthier, N.C.; Biais, N.; Sheetz, M.P. Integrin-Dependent Force Transmission to the Extracellular Matrix by α-Actinin Triggers Adhesion Maturation. Proc. Natl. Acad. Sci. USA 2013, 110, E1361–E1370. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Arandjelovic, S.; Perry, J.S.A.; Lucas, C.D.; Penberthy, K.K.; Kim, T.-H.; Zhou, M.; Rosen, D.A.; Chuang, T.-Y.; Bettina, A.M.; Shankman, L.S.; et al. A Noncanonical Role for the Engulfment Gene ELMO1 in Neutrophils That Promotes Inflammatory Arthritis. Nat. Immunol. 2019, 20, 141–151. [Google Scholar] [CrossRef] [PubMed]
  216. Taylor, A.; Tang, W.; Bruscia, E.M.; Zhang, P.-X.; Lin, A.; Gaines, P.; Wu, D.; Halene, S. SRF Is Required for Neutrophil Migration in Response to Inflammation. Blood 2014, 123, 3027–3036. [Google Scholar] [CrossRef] [Green Version]
  217. Ferrari, G.; Langen, H.; Naito, M.; Pieters, J. A Coat Protein on Phagosomes Involved in the Intracellular Survival of Mycobacteria. Cell 1999, 97, 435–447. [Google Scholar] [CrossRef] [Green Version]
  218. Pick, R.; Begandt, D.; Stocker, T.J.; Salvermoser, M.; Thome, S.; Böttcher, R.T.; Montanez, E.; Harrison, U.; Forné, I.; Khandoga, A.G.; et al. Coronin 1A, a Novel Player in Integrin Biology, Controls Neutrophil Trafficking in Innate Immunity. Blood 2017, 130, 847–858. [Google Scholar] [CrossRef]
  219. Riley, D.R.J.; Khalil, J.S.; Pieters, J.; Naseem, K.M.; Rivero, F. Coronin 1 Is Required for Integrin Β2 Translocation in Platelets. Int. J. Mol. Sci. 2020, 21, 356. [Google Scholar] [CrossRef] [Green Version]
  220. Liao, C.; Yang, H.; Zhang, R.; Sun, H.; Zhao, B.; Gao, C.; Zhu, F.; Jiao, J. The Upregulation of TRPC6 Contributes to Ca2+ Signaling and Actin Assembly in Human Mesangial Cells after Chronic Hypoxia. Biochem. Biophys. Res. Commun. 2012, 421, 750–756. [Google Scholar] [CrossRef]
  221. Lindemann, O.; Rossaint, J.; Najder, K.; Schimmelpfennig, S.; Hofschröer, V.; Wälte, M.; Fels, B.; Oberleithner, H.; Zarbock, A.; Schwab, A. Intravascular Adhesion and Recruitment of Neutrophils in Response to CXCL1 Depends on Their TRPC6 Channels. J. Mol. Med. 2020, 98, 349–360. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Mazaki, Y.; Takada, S.; Nio-Kobayashi, J.; Maekawa, S.; Higashi, T.; Onodera, Y.; Sabe, H. Mitofusin 2 Is Involved in Chemotaxis of Neutrophil-like Differentiated HL-60 cells. Biochem. Biophys. Res. Commun. 2019, 513, 708–713. [Google Scholar] [CrossRef] [PubMed]
  223. Zhou, W.; Hsu, A.Y.; Wang, Y.; Syahirah, R.; Wang, T.; Jeffries, J.; Wang, X.; Mohammad, H.; Seleem, M.N.; Umulis, D.; et al. Mitofusin 2 Regulates Neutrophil Adhesive Migration and the Actin Cytoskeleton. J. Cell Sci. 2020, 133, jcs248880. [Google Scholar] [CrossRef] [PubMed]
  224. Liu, W.; Hsu, A.Y.; Wang, Y.; Lin, T.; Sun, H.; Pachter, J.S.; Groisman, A.; Imperioli, M.; Yungher, F.W.; Hu, L.; et al. Mitofusin-2 Regulates Leukocyte Adhesion and Β2 Integrin Activation. J. Leukoc. Biol. 2021. [Google Scholar] [CrossRef]
  225. Fan, Z.; Kiosses, W.B.; Sun, H.; Orecchioni, M.; Ghosheh, Y.; Zajonc, D.M.; Arnaout, M.A.; Gutierrez, E.; Groisman, A.; Ginsberg, M.H.; et al. High-Affinity Bent Β2-Integrin Molecules in Arresting Neutrophils Face Each Other through Binding to ICAMs In Cis. Cell Rep. 2019, 26, 119–130.e5. [Google Scholar] [CrossRef] [Green Version]
  226. Sun, H.; Zhi, K.; Hu, L.; Fan, Z. The Activation and Regulation of Β2 Integrins in Phagocytes and Phagocytosis. Front. Immunol. 2021, 12, 978. [Google Scholar] [CrossRef]
  227. Yao, C.-H.; Wang, R.; Wang, Y.; Kung, C.-P.; Weber, J.D.; Patti, G.J. Mitochondrial Fusion Supports Increased Oxidative Phosphorylation during Cell Proliferation. eLife 2019, 8, e41351. [Google Scholar] [CrossRef]
  228. de Brito, O.M.; Scorrano, L. Mitofusin 2 Tethers Endoplasmic Reticulum to Mitochondria. Nature 2008, 456, 605–610. [Google Scholar] [CrossRef]
  229. Kuo, I.Y.; Brill, A.L.; Lemos, F.O.; Jiang, J.Y.; Falcone, J.L.; Kimmerling, E.P.; Cai, Y.; Dong, K.; Kaplan, D.L.; Wallace, D.P.; et al. Polycystin 2 Regulates Mitochondrial Ca2+ Signaling, Bioenergetics, and Dynamics through Mitofusin 2. Sci. Signal. 2019, 12, eaat7397. [Google Scholar] [CrossRef]
  230. Khodzhaeva, V.; Schreiber, Y.; Geisslinger, G.; Brandes, R.P.; Brüne, B.; Namgaladze, D. Mitofusin 2 Deficiency Causes Pro-Inflammatory Effects in Human Primary Macrophages. Front. Immunol. 2021, 12, 723683. [Google Scholar] [CrossRef]
  231. Lloberas, J.; Muñoz, J.P.; Hernández-Álvarez, M.I.; Cardona, P.-J.; Zorzano, A.; Celada, A. Macrophage Mitochondrial MFN2 (Mitofusin 2) Links Immune Stress and Immune Response through Reactive Oxygen Species (ROS) Production. Autophagy 2020, 16, 2307–2309. [Google Scholar] [CrossRef]
  232. Tur, J.; Pereira-Lopes, S.; Vico, T.; Marín, E.A.; Muñoz, J.P.; Hernández-Alvarez, M.; Cardona, P.-J.; Zorzano, A.; Lloberas, J.; Celada, A. Mitofusin 2 in Macrophages Links Mitochondrial ROS Production, Cytokine Release, Phagocytosis, Autophagy, and Bactericidal Activity. Cell Rep. 2020, 32, 108079. [Google Scholar] [CrossRef] [PubMed]
  233. Fagerholm, S.C.; Hilden, T.J.; Nurmi, S.M.; Gahmberg, C.G. Specific Integrin Alpha and Beta Chain Phosphorylations Regulate LFA-1 Activation through Affinity-Dependent and -Independent Mechanisms. J. Cell Biol. 2005, 171, 705–715. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Fagerholm, S.C.; Varis, M.; Stefanidakis, M.; Hilden, T.J.; Gahmberg, C.G. Alpha-Chain Phosphorylation of the Human Leukocyte CD11b/CD18 (Mac-1) Integrin Is Pivotal for Integrin Activation to Bind ICAMs and Leukocyte Extravasation. Blood 2006, 108, 3379–3386. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Uotila, L.M.; Aatonen, M.; Gahmberg, C.G. Integrin CD11c/CD18 α-Chain Phosphorylation Is Functionally Important. J. Biol. Chem. 2013, 288, 33494–33499. [Google Scholar] [CrossRef] [Green Version]
  236. Nurmi, S.M.; Autero, M.; Raunio, A.K.; Gahmberg, C.G.; Fagerholm, S.C. Phosphorylation of the LFA-1 Integrin Beta2-Chain on Thr-758 Leads to Adhesion, Rac-1/Cdc42 Activation, and Stimulation of CD69 Expression in Human T Cells. J. Biol. Chem. 2007, 282, 968–975. [Google Scholar] [CrossRef] [Green Version]
  237. Yasuda, T.; Bundo, K.; Hino, A.; Honda, K.; Inoue, A.; Shirakata, M.; Osawa, M.; Tamura, T.; Nariuchi, H.; Oda, H.; et al. Dok-1 and Dok-2 Are Negative Regulators of T Cell Receptor Signaling. Int. Immunol. 2007, 19, 487–495. [Google Scholar] [CrossRef]
  238. Mashima, R.; Hishida, Y.; Tezuka, T.; Yamanashi, Y. The Roles of Dok Family Adapters in Immunoreceptor Signaling. Immunol. Rev. 2009, 232, 273–285. [Google Scholar] [CrossRef]
  239. Lim, J.; Hotchin, N.A.; Caron, E. Ser756 of Β2 Integrin Controls Rap1 Activity during Inside-out Activation of AMβ2. Biochem. J. 2011, 437, 461–467. [Google Scholar] [CrossRef] [Green Version]
  240. Lim, S.; Sala, C.; Yoon, J.; Park, S.; Kuroda, S.; Sheng, M.; Kim, E. Sharpin, a Novel Postsynaptic Density Protein That Directly Interacts with the Shank Family of Proteins. Mol. Cell Neurosci. 2001, 17, 385–397. [Google Scholar] [CrossRef]
  241. Ikeda, F.; Deribe, Y.L.; Skånland, S.S.; Stieglitz, B.; Grabbe, C.; Franz-Wachtel, M.; van Wijk, S.J.L.; Goswami, P.; Nagy, V.; Terzic, J.; et al. SHARPIN Forms a Linear Ubiquitin Ligase Complex Regulating NF-ΚB Activity and Apoptosis. Nature 2011, 471, 637–641. [Google Scholar] [CrossRef] [PubMed]
  242. Tokunaga, F.; Nakagawa, T.; Nakahara, M.; Saeki, Y.; Taniguchi, M.; Sakata, S.; Tanaka, K.; Nakano, H.; Iwai, K. SHARPIN Is a Component of the NF-ΚB-Activating Linear Ubiquitin Chain Assembly Complex. Nature 2011, 471, 633–636. [Google Scholar] [CrossRef] [PubMed]
  243. IWAI, K. LUBAC-Mediated Linear Ubiquitination: A Crucial Regulator of Immune Signaling. Proc. Jpn. Acad. Ser. B Phys. Biol. Sci. 2021, 97, 120–133. [Google Scholar] [CrossRef] [PubMed]
  244. Kirisako, T.; Kamei, K.; Murata, S.; Kato, M.; Fukumoto, H.; Kanie, M.; Sano, S.; Tokunaga, F.; Tanaka, K.; Iwai, K. A Ubiquitin Ligase Complex Assembles Linear Polyubiquitin Chains. EMBO J. 2006, 25, 4877–4887. [Google Scholar] [CrossRef] [PubMed]
  245. Rantala, J.K.; Pouwels, J.; Pellinen, T.; Veltel, S.; Laasola, P.; Potter, C.S.; Duffy, T.; Sundberg, J.P.; Kallioniemi, O.; Askari, J.A.; et al. SHARPIN Is an Endogenous Inhibitor of Beta1-Integrin Activation. Nat. Cell Biol. 2011, 13, 1315–1324. [Google Scholar] [CrossRef] [Green Version]
  246. Tokunaga, F.; Iwai, K. LUBAC, a Novel Ubiquitin Ligase for Linear Ubiquitination, Is Crucial for Inflammation and Immune Responses. Microbes Infect. 2012, 14, 563–572. [Google Scholar] [CrossRef]
  247. Gurung, P.; Sharma, B.R.; Kanneganti, T.-D. Distinct Role of IL-1β in Instigating Disease in Sharpincpdm Mice. Sci. Rep. 2016, 6, 36634. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Schematic of neutrophil recruitment cascade: Flowing neutrophils roll on vascular endothelial cells through the interaction between PSGL-1 and selectins. Engagement of PSGL-1 with selectins induces the extension of LFA-1, which binds ICAM-1 or other ligands with an intermediate affinity and slows down rolling velocity. We call this process slow rolling. β1 Integrins are also involved in neutrophil rolling and slow rolling. When rolling neutrophils encounter chemokines, LFA-1 or Mac-1 will be fully activated and bind ICAM-1 or other ligands with a high affinity and stop rolling neutrophils. We call this process arrest. After arrest, neutrophils spread and crawl on endothelial cells, which is predominantly mediated by Mac-1. Neutrophils undergo trans-endothelial migration in an LFA-1/Mac-1-dependent manner to the tissue or site of injury. Created in BioRender.com (accessed on 12 June 2022).
Figure 1. Schematic of neutrophil recruitment cascade: Flowing neutrophils roll on vascular endothelial cells through the interaction between PSGL-1 and selectins. Engagement of PSGL-1 with selectins induces the extension of LFA-1, which binds ICAM-1 or other ligands with an intermediate affinity and slows down rolling velocity. We call this process slow rolling. β1 Integrins are also involved in neutrophil rolling and slow rolling. When rolling neutrophils encounter chemokines, LFA-1 or Mac-1 will be fully activated and bind ICAM-1 or other ligands with a high affinity and stop rolling neutrophils. We call this process arrest. After arrest, neutrophils spread and crawl on endothelial cells, which is predominantly mediated by Mac-1. Neutrophils undergo trans-endothelial migration in an LFA-1/Mac-1-dependent manner to the tissue or site of injury. Created in BioRender.com (accessed on 12 June 2022).
Cells 11 02025 g001
Figure 2. Schematic of the integrin activation conformational changes and key modulators: Resting integrins have a bent low-affinity (EH) conformation. They attain an extended low-affinity (E+H) confirmation through the RIAM- and Rap1-mediated recruitment of talin-1. Further recruitment of kindlin-3 induces full integrin activation into extended high-affinity integrins (E+H+). There is an alternative allosteric pathway in which EH integrins transition to a bent-high affinity conformation (EH+) and then change to an extended high-affinity integrin (E+H+). Which integrin adaptors are involved in this allosteric pathway remains to be further investigated. Created in BioRender.com (accessed on 12 June 2022).
Figure 2. Schematic of the integrin activation conformational changes and key modulators: Resting integrins have a bent low-affinity (EH) conformation. They attain an extended low-affinity (E+H) confirmation through the RIAM- and Rap1-mediated recruitment of talin-1. Further recruitment of kindlin-3 induces full integrin activation into extended high-affinity integrins (E+H+). There is an alternative allosteric pathway in which EH integrins transition to a bent-high affinity conformation (EH+) and then change to an extended high-affinity integrin (E+H+). Which integrin adaptors are involved in this allosteric pathway remains to be further investigated. Created in BioRender.com (accessed on 12 June 2022).
Cells 11 02025 g002
Figure 3. Schematics showing domains and binding proteins of talin (A), RIAM (B), and kindlin (C). Integrin-NPXY: Asn-Pro-x-Tyr motifs of the β-integrin tail; Integrin MP region: membrane-proximal region of the β-integrin tail; Integrin-NxxY/F: Asn-x-x-Tyr/Phe motifs of the β-integrin tail; PIP2: phosphatidylinositol 4,5-bisphosphate; PIP3: phosphatidylinositol-3,4,5-triphosphate; Profilin-Actin: direct binding to profilin and indirect binding to actin. Created in BioRender.com (accessed on 12 June 2022).
Figure 3. Schematics showing domains and binding proteins of talin (A), RIAM (B), and kindlin (C). Integrin-NPXY: Asn-Pro-x-Tyr motifs of the β-integrin tail; Integrin MP region: membrane-proximal region of the β-integrin tail; Integrin-NxxY/F: Asn-x-x-Tyr/Phe motifs of the β-integrin tail; PIP2: phosphatidylinositol 4,5-bisphosphate; PIP3: phosphatidylinositol-3,4,5-triphosphate; Profilin-Actin: direct binding to profilin and indirect binding to actin. Created in BioRender.com (accessed on 12 June 2022).
Cells 11 02025 g003
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Pulikkot, S.; Hu, L.; Chen, Y.; Sun, H.; Fan, Z. Integrin Regulators in Neutrophils. Cells 2022, 11, 2025. https://doi.org/10.3390/cells11132025

AMA Style

Pulikkot S, Hu L, Chen Y, Sun H, Fan Z. Integrin Regulators in Neutrophils. Cells. 2022; 11(13):2025. https://doi.org/10.3390/cells11132025

Chicago/Turabian Style

Pulikkot, Sunitha, Liang Hu, Yunfeng Chen, Hao Sun, and Zhichao Fan. 2022. "Integrin Regulators in Neutrophils" Cells 11, no. 13: 2025. https://doi.org/10.3390/cells11132025

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