Skip to main content

Structure, signal transduction, activation, and inhibition of integrin αIIbβ3

Abstract

Integrins are heterodimeric receptors comprising α and β subunits. They are expressed on the cell surface and play key roles in cell adhesion, migration, and growth. Several types of integrins are expressed on the platelets, including αvβ3, αIIbβ3, α2β1, α5β1, and α6β1. Among these, physically αIIbβ3 is exclusively expressed on the platelet surface and their precursor cells, megakaryocytes. αIIbβ3 adopts at least three conformations: i) bent-closed, ii) extended-closed, and iii) extended–open. The transition from conformation i) to iii) occurs when αIIbβ3 is activated by stimulants. Conformation iii) possesses a high ligand affinity, which triggers integrin clustering and platelet aggregation. Platelets are indispensable for maintaining vascular system integrity and preventing bleeding. However, excessive platelet activation can result in myocardial infarction (MI) and stroke. Therefore, finding a novel strategy to stop bleeding without accelerating the risk of thrombosis is important. Regulation of αIIbβ3 activation is vital for this strategy. There are a large number of molecules that facilitate or inhibit αIIbβ3 activation. The interference of these molecules can accurately control the balance between hemostasis and thrombosis. This review describes the structure and signal transduction of αIIbβ3, summarizes the molecules that directly or indirectly affect integrin αIIbβ3 activation, and discusses some novel antiαIIbβ3 drugs. This will advance our understanding of the activation of αIIbβ3 and its essential role in platelet function and tumor development.

Introduction

Integrins are key mediators of cell-matrix and cell-cell adhesion in physiology and disease. Recently, the 2022 Albert Lasker Basic Medical Research Award was presented to three scientists for their discoveries concerning the integrins [1]. Integrins are a large family of cell surface adhesion receptors. They comprise heterodimeric transmembrane glycoprotein complexes assembled from noncovalently bound α and β subunits. Every integrin subunit contains a large extracellular domain, transmembrane domain, and short cytoplasmic domain [2]. The α subunit comprises a β-propeller, followed by thigh, genu, calf-1, and calf-2 domains. Some α-subunits have an inserted domain (αI domain), serving as a major ligand-binding site [3]. The β subunit includes a βI domain, a hybrid domain, a plexin-sempahorin-integrin (PSI) domain, four cysteine-rich epidermal growth factor (EGF) repeats, and a membrane-proximal β-tail domain (βTD). Integrins adopt at least three conformations: bent-closed (resting state), extended-closed (intermediate state), and extended-open (activated state) [3, 4] (Fig. 1). The α and β cytoplasmic domains are in close proximity to each other in the resting state. The separation of the two tails is triggered by signal transmission from the intracellular domain to the extracellular domain. This is called ‘inside-out signaling’, which is stimulated by agonists, converts integrin into an extended conformation and increases the affinity of αIIbβ3 for ligands containing a common Arg-Gly-Asp (RGD) motif [5]. Subsequently, ligand binding to integrin initiates the outside-in signaling pathway. This triggers and amplifies various cellular events, including cell spreading, stable thrombus formation, and clot retraction. Transmembrane domain of integrin is involved in bidirectional signaling. Once integrin is activated, the transmembrane domain separates and serves as a signal transmitter. Transmembrane domain of αIIbβ3 is most stably associated, when compared to that of αvβ3 and α5β 1[6]. Truncated integrins lacking the transmembrane domain and cytoplastic domain are constitutively active [7], while αvβ3 ectodomain without transmembrane domain is in the resting bent conformation.

Fig. 1
figure 1

The structure of integrin. The conformational changes from bent-closed, to extended-closed, and finally to extended-open state [3, 4]

Integrin αIIbβ3 is the most abundant receptor expressed on the platelet surface [8]. It binds fibrinogen, which is central to platelet aggregation and spreading. The hemostasis process occurs as follows: vessel wall injury results in collagen exposure, which is bound by von Willebrand factor (vWF) and results in platelet recruitment to the site of the damaged vessel wall. vWF mediates the linkage between platelets and collagen via receptors on the platelet membrane [9]. Platelets are activated by modification from a disc shape into a dendritic form. Activated platelets release several prothrombotic molecules such as adenosine diphosphate (ADP). ADP triggers platelet aggregation, thus recruits more platelets to the injury site. Moreover, activated platelets can trigger the coagulation cascade, resulting in the full activation of platelets, which contribute to the formation of solid clots. αIIbβ3 has become a validated antithrombotic target since it is the main integrin in platelets. This review focuses on the role of integrin αIIbβ3 in platelet activation, which lays the foundation for our understanding of antiplatelet and antitumor therapy.

Structure of αIIbβ3

Integrin αIIbβ3 contains extracellular-, transmembrane-, and cytoplasmic domains. The transmembrane segments of the α- and β-subunits are responsible for signaling events [1]. The αIIb transmembrane segment is a linear α spiral perpendicular to the cell membrane, followed by a corner region of the main chain, whereas β3 is a linear α-helix at an angle of 25° to αIIb [10]. Moreover, there is a conserved membrane-proximal GFFKR region in the αIIb and HDRxE region in the β3 tail, which form a “clasp” and contribute to maintaining an inactive state of αIIbβ3 [11]. Interactions between the transmembrane domains of αIIb and β3 are vital for maintaining integrin in the resting state. The first clasp is called the “outer membrane clasp” (OMC) and is formed by the interaction of the αIIb subunit (Gly972 and Gly976) with the β3 subunit (Gly708) near the outer membrane of the transmembrane segment of integrin in the lipid bilayer. The second clasp is termed the “inner membrane clasp” (IMC) and is assembled from a hydrophobic region of αIIb (F992-F993) and electrostatic αIIb salt bridge (R995)/β3 (D723) located near the inner membrane of the transmembrane segment of the integrin. Mutations in the IMC and the OMC can disturb the interaction of the transmembrane domain. The electrostatic interaction between R995 in αIIb and D723 in β3 is important for maintaining IMC stability, thus stabilizing the low-affinity state and regulating integrin activation [12]. Artificial dimerization or disulfide binding can inhibit integrin activity, whereas the R995 mutation in αIIb or D723 mutation in β3 can stretch and activate integrins [13]. The β3 cytoplasmic segment is longer than αIIb and is more conserved than αIIb. Furthermore, β3 plays an important role in integrin signal transduction. Each extracellular domain of the integrin is composed of a headpiece and a leg piece. In the resting state, the headpiece folds onto the leg piece such that the conformation has a low affinity for the ligands. Once stimulated, the αIIbβ3 headpieces are exposed to ligands such as vWF, fibrinogen, fibronectin, and vitronectin [3]. The swing-out motion of the hybrid domain and switchblade-like movement of the headpiece from the leg domains are the two major conformational rearrangements of αIIbβ3 [14]. A mutation of L33 to P33 in β3 makes the PSI-, I-EGF-1-, and I-EGF-2 domains more flexible, thus improving the ability of β3 to bind fibrinogen and increasing the risk of thrombus formation [15]. However, there are still controversial results. Zhou et al. designed a chimeric β3 with both L33 and P33 forms and discovered that L33 to P33 does not cause conformational changes or elevated ligand binding in β3 integrin. A full-length β3 fragment that adopts intermediate and fully extended conformations of β3 was also designed to provide high-resolution information on how β3 integrin extends at its PSI-, I-EGF1-, and I-EGF2 domains (β-knee region) [16]. Moreover, in platelets with β3 P33 mutation, the Src and FAK are highly activated, especially at high shear rates and β3 P33 mutation knockin mice show increased platelet adhesion, aggregation, and clot formation [17, 18]. They also demonstrated that the I-EGF2 C1-C2 loop plays a critical role in integrin extension, which may be an excellent target for regulating integrin function [19].

Interaction of β3 with signaling factors

Interaction of β3 NPxY747 motif with Talin

Talin is a cytoskeletal protein that consists of an N-terminal head domain (talin-H) and C-terminal flexible rod domain (talin-R). Talin-H includes an F0 domain and an FERM domain (band 4.1, ezrin, radixin, moesin), which contains three subdomains: F1, F2, and F3 [20, 21] (Fig. 2). F3 resembles a phosphotyrosine-binding domain and is responsible for binding the membrane-proximal NPxY motif of β3, phosphatidyl inositol phosphate kinase type 1γ (PIPK1γ), and layilin [22]. Talin adopts a dual auto-inhibited conformation in the resting state. Talin-R masks the β3 tail binding sites in the F3 domain. Meanwhile, talin-R is negatively charged and repels the talin-F2F3/talin-R complex from the membrane. Furthermore, phosphatidylinositol 4,5-bis-phosphate (PIP2)-enriched membranes contribute to talin activation. The strong affinity of PIP2 for positively charged talin-F2F3 pulls talin-F2F3 domains to the membrane. Meanwhile, talin-R is pushed away; this is called the pull-push model [21]. There are two integrin binding sites (IBS) on talin: one is located in talin-H (IBS1), while the other is in talin-R (IBS2). Binding of talin to integrin tails is the final and most important step leading to integrin activation; therefore, its activity should be well regulated [23]. Interactions between the talin-F3 domain and β3 tail can disrupt the salt bridge formed by R995 in αIIb and D723 in β3, resulting in conformational changes that increase fibrinogen affinity. During this process, the talin-F3 domain interacts with PIPKIγ to catalyze PIP2 production, with talin relieving its auto-inhibition between the FERM domain and the C-terminal rod domain to become activated [24]. Gα13 switch region 2 competes with talin-R and directly binds to talin-F3, thus relieving its autoinhibition and activating talin [25]. Furthermore, talin-F3 binds to the highly conserved NPxY747 motif of β3. The loop structure in talin- (309–405) interacts with the α-helix in the membrane-proximal cytoplasmic domain of β3 which causes rearrangement of the transmembrane segment and interferes with D723; this disrupts the salt bridge resulting in αIIbβ3 activation [26]. Mutation of the NPxY747 motif in β3 and mutation of talin-F3 can impair talin binding and reduce integrin affinity. Talin-R includes several vinculin-binding sites (VBSs) and several actin-binding sites that connect to the extracellular matrix (ECM). Furthermore, full-length vinculin is autoinhibited unless activated, which is similar to mode of operation of talin. Activated talin binds to vinculin via the two vinculin-binding sites (VBSs) within the talin R3 helical bundle. This activates vinculin by disrupting the vinculin head-tail interaction and directly links it to the ECM.

Fig. 2
figure 2

The structural change of talin from the inactive form to the active form [21]

Recent findings suggested that the Rap1/talin-1 (Talin includes talin1 and talin2) interaction is vital for platelet integrin activity. Rap1a and Rap1b are Ras subfamily members which share a high degree of similarity in vertebrates [27]. Rap1a and Rap1b may have some overlapping functions [4]. Rap1b deficient mice show impaired platelet function. Rap1-GTP-interacting adaptor molecule (RIAM) was previously considered a Rap1 effector [28]. Talin binding sites to RIAM are identified in R3, and R8 domains and F3 of talin [29]. However, increasing studies indicate that the RIAM gene in mice is dispensable for platelet integrin functions [30, 31]. RIAM depleted mice exhibit severe leukocytosis and impaired leukocyte extravasation, suggesting that RIAM may be a new target for human inflammatory diseases [32]. Furthermore, Rap1 directly interacts with talin1 and enhances talin-induced integrin activation [29]. Platelet functions such as aggregation, spreading, hemostasis, and thrombosis are impaired in Rap1 binding-deficient talin1 knock-in mice [33]. There are two identified talin1 binding sites to Rap1: one is located in the F0 domain and the other is in the F1 domain. Rap1 binding to the F1 domain makes a greater contribution to αIIbβ3 activation than binding to the F0 domain [4]. αIIbβ3 activation requires the talin1 F1/Rap1 interaction whereas the talin 1 F0/Rap1 interaction is not essential [34].

Interaction of the β3 NxxY759 motif with kindlin

Kindlins, similarly as talin, have important implications for integrin signal transduction. Kindlin is named after the Kindler syndrome, a rare skin blister disease characterized by blistering, skin atrophy, photosensitivity, pigmentation defects and so on [35]. All these symptoms are related to cell adhesion defects [36]. Members of the kindlin family include kindlin-1, kindlin-2, and kindlin-3, and they exhibit a high degree of similarity [37]. Kindlin-1 is mainly expressed in epithelial cells [38]. Kindlin-2 is ubiquitously expressed in tissues and is mainly found in skeletal- and smooth muscle cells. Knockout of kindlin-2 results in mouse embryonic lethality [39, 40]. A recent study reported that kindlin-2 depletion induces cardiac myocyte hypertrophy and increases the heart weight [41]. Surprisingly, kindlin-2 may play a central role in the occurrence and development of fatty liver and diabetes mellitus [42, 43], suggesting that it may be involved in the metabolism of blood sugars and lipids. Additionally, kindlin-2 binds to integrin-linked kinase (ILK) and promotes focal adhesions [44, 45]. Kindlin-3 was previously thought to be restricted to hematopoietic cells; however, more recent work shows that is also present in endothelial cells [39]. Kindlin-3-deficient leukocytes fail to arrest and extravasate into inflammatory tissues which causes leukocyte adhesion deficiency syndrome type III; this manifests in massive bleeding and recurrent infection [46]. Patients with kindlin-3 gene variants present with impaired platelet function, lymphocytosis and granulocytosis [47].

Kindlins and talins are FERM-containing proteins. Moreover, kindlin comprises a loop region in the F1 domain and a pleckstrin homology (PH) domain inserted in the F2 domain [48]. Unlike the extended structure of talin, kindlin adopts a cloverleaf structure [49] and multiple salt bridges in the F1 and F3 subdomains stabilize the cloverleaf-like FERM domain which supports integrin αIIbβ3 activation [49]. Mutations of interface residues in the F3 subdomain impair the ability of kindlin to support integrin αIIbβ3 activation [49]. The binding of kindlin to the cytoplasmic tail of β3 integrin cooperates with talin to activate β3. The β3 membrane-proximal NxxP747 motif is the binding site for talin, while the membrane-distal NxxY759 motif is the binding site for kindlin. Both talin and kindlin are necessary for regulating integrin affinity; however, the mechanism of cooperative activation remains unclear [50]. One possible mechanism involves kindlin binding to the β3 tail to expose and stabilize the proximal talin-binding site in β3, resulting in talin destroying the αIIbβ3 transmembrane domain and activating αIIbβ3. Alternatively, talin binds to β3, followed by the recruitment of kindlin to trigger integrin αIIbβ3 clustering, and enhances its affinity for extracellular ligands [51].

Many proteins have been identified to interact with kindlin, including paxillin, adhesion and degranulation promoting adaptor protein (ADAP), migfilin, α-actinin, Src, and F-actin. In recent years, paxillin was thought to interact with kindlin-3 via its F0 subdomain [52], while other studies suggest that binding occurs through its F1- and PH domain [51, 53]. Therefore, the precise mechanism underlying the interaction between paxillin and kindlin-3 remains unclear.

Interaction of β3 R760GT762 with Src family kinase (SFK)

Src binds to the last three amino acid residues in the β3 cytoplasmic tail (R760, G761, and T762) via its SH3 domain in platelets mainly mediating outside-in signaling transduction [54]. The platelets with R760GT762-deleted β3 show impaired spreading on immobilized fibrinogen, but intact soluble fibrinogen binding [55, 56]. Src contains the following domains: an N-terminal 14-carbon myristoyl group, two highly conserved Src homology domains (SH2 and SH3), a catalytic domain, and a C-terminal tail [57]. Recently our group found that the residues, especially E97, in the RT loop of Src SH3 are critical for interacting with β3 [58]. DCDBS84, a small molecule disrupting β3/Src interaction, inhibits the outside-in signaling-regulated platelet functions [58]. Src myristoylation promotes cell membrane binding and is imperative for function in vivo. The SH2 region of Src interacts with its C-terminal phosphorylated Y529, which is maintained by the c-Src kinase, Csk. Csk also binds to the β3 cytoplasmic tail. Once αIIbβ3 binds to fibrinogen, Src Y529 is dephosphorylated and Y418 is phosphorylated [59]. This is followed by Csk dissociation from β3, and the activation of Src triggers a tyrosine phosphorylation cascade leading to phosphorylation of the β3 tail at Y747 and Y759 [60, 61]. Src can interact with FAK (focal adhesion kinase) to form a complex and mediates cell growth and migration. In addition, tumor growth and metastasis are associated with Src-FAK complex too [62]. Spleen tyrosine kinase (Syk) is downstream of Src and implicated in outside-in signaling of αIIbβ3. Platelets in Syk−/− mice show impaired spreading [59]. Moreover, the cleavage of the β3 Src-binding site by calpain is regarded as a molecular switch between cell spreading and retraction [63]. Loss of Src binding to the β3 tail results in defective spreading and enhanced clot retraction mediated by RhoA, indicating that Src may impose restrictions on RhoA [63]. There are four types of SFKs in mouse platelets: Src, Fgr, Fyn, and Lyn. Each plays a different role in platelet activation. Mice deficient in SFK are characterized by impaired tyrosine phosphorylation [61]. However, individual depletion of each type does not result in severe bleeding, indicating that overlapping functions may exist in the four SFKs [64]. Fyn binds to the β3721–725 domain (I721HDRK725). Fyn−/− mice show prolonged bleeding and delayed platelet spreading [65]. The β3721–725 domain is also required for binding of other proteins, including focal adhesion kinase, skelemin, and paxillin. The β3721–725 domain is close to the salt bridge; therefore, mutations that affect the stability of the transmembrane domain induce conformational changes. Junctional adhesion molecule A (JAM-A) is associated with the regulation of platelet function by inhibiting the recruitment of c-Src kinase, which inhibits c-Src activation and outside-in signaling [66]. JAM-A-knockout mice have shorter bleeding times and enhanced thrombosis capability [67].

Interaction E731XE733 of β3 with Gα13

Gα13 is a heterotrimeric guanine nucleotide-binding protein and can interact with E731XE733 motif in β3 tail. αIIbβ3-expressing CHO cells or platelets interfered with Gα13 siRNA neither spread well on fibrinogen-coated surface nor activate Src [68]. While interference of Gα13 stimulates the small guanosine triphosphatase RhoA with accelerated cell retraction [68]. Platelets treated with the myristoylated EXE-motif-containing peptide, which disrupts Gα13/β3 E731XE733 interaction, demonstrate poor spreading and aggregation, while strong clot retraction [69].

The αIIbβ3 bidirectional signaling and the interactions are shown in Fig. 3.

Fig. 3
figure 3

αIIbβ3 bidirectional signaling. The agonists, such as thrombin, ADP, epinephrine, 5-HT, TXA2, or collagen, bind to their receptors causing intracellular signaling events, which then lead talin to bind to β3 tail with the help of kindlin (A and B). Then, αIIbβ3 conformation changes from bent-closed to extended-open state. Thus, the affinity of αIIbβ3 to fibrinogen increases (C). This process is called inside-out signaling transduction (C). Once fibrinogen binds to αIIbβ3, signal transfers from outside to inside, named as outside-in signaling transduction, causing Gα13 to bind to E731XE733 motif, and Src to bind to R760GT762 motif, accompanied by phosphorylation of Y747 and Y759 in β3 tail, as well as phosphorylation of Y416 in Src. At last, the platelet spreading, aggregation, and clot retraction develop (D). For representative purposes, β tail is outsized

Other factors involved in αIIbβ3 bidirectional signaling

Proteins which promote αIIbβ3 activation

Phosphoinositide 3-kinases (PI3Ks) are a group of lipid kinases, including different isoforms [70]. PI3Kα is involved in the outside-in signaling of vWF-engaged αIIbβ3 integrin. An absence or inhibition of PI3Kα results in a decreased thrombus size after superficial injuries of mouse mesenteric arteries and an increased time to arterial occlusion after carotid lesion, without prolonged tail bleeding time [71]. PI3Kβ is responsible for PIP2 production and is linked to the activation of Rap1b and serine/threonine kinase [72]. Inhibition of PI3Kβ or PI3Kγ leads to a profound defect in platelet aggregation, hemostatic plug formation, and arterial thrombosis [72].

Paxillin is a key focal adhesion adaptor/scaffold protein which binds with the F0 subdomain of kindlin-1 or the PH domain of kindlin-2 and kindlin-3 to support αIIbβ3 activation [73]. Interaction between paxillin and the kindlin-2 F0 domain contributes to the recruitment of paxillin to focal adhesions which leads to cell migration and transmission of mechanical force along the cytoskeleton [74]. Recently, it is found that the interaction between paxillin and PH domain of kindlin-3 plays an important role in supporting integrin αIIbβ3 outside-in signaling in platelets [53]. Furthermore, co-existence of paxillin and kindlin-1 greatly improves binding of the talin-H to β3, suggesting that paxillin may facilitate the interaction between talin and β3. Suppression of paxillin expression significantly inhibits αIIbβ3 activation [52]. Moreover, paxillin binds vinculin and talin, leading to talin and vinculin recruitment into nascent adhesions and inducing focal adhesion maturation [75]. Besides the indirect interaction, paxillin possibly directly interacts with talin and β3 integrin. It functions as a signaling molecule, recruiting other focal adhesion proteins, modulating F-actin polymerization, and thus generating feedback signaling [76].

Integrin-linked kinase (ILK) is a component of focal adhesion, which participates in various biological processes including cell migration, proliferation, and survival. ILK binds with pinch and parvin to form the ILK/pinch/parvin complex (IPP), thereby helping kindlin-2 activate αIIbβ3 [77]. Additionally, pinch participates in the modulation of chondrogenesis and bone mass, thus maintaining bone homeostasis [78]. ILK-deficient mice or cells are defective in ligand-binding, granule secretion, and platelet aggregation [79, 80].

Calcium and integrin-binding protein 1 (CIB1) is an αIIb-binding protein participating in c-Src activation. CIB1 is indispensable for platelet spreading [81] and CIB1−/− mice are characterized by defective thrombosis [82]. Additionally, CIB1 is a regulator that enhances the activity of focal adhesion kinases and contributes to the formation of focal adhesions [83].

Vinculin is a focal adhesion protein with an autoinhibited structure and regulates mechanical coupling between the ECM and cytoskeleton [84]. The release of talin from autoinhibition causes a change in its conformation which allows vinculin to bind; this disrupts auto-inhibition and causes activation [85]. There is a PIP2 binding site in the vinculin tail and the PIP2/vinculin interaction results in a conformational change in vinculin, followed by vinculin activation. Vinculin-deficient mice are embryonic lethal [86].

Filamin (FLN) is a large actin-binding protein that contains three isoforms: FLNα, FLNb, and FLNc. FLNa mutations are associated with many diseases including skeletal dysplasia, neuronal migration abnormalities, and intestinal malrotation. FLNa mutations alter platelet production from the megakaryocytes, causing macrothrombocytopenia. In platelets per se, FLNa mutations may lead to impaired αIIbβ3 activation [87]. A recent study demonstrated that FLNα/αIIbβ3 interactions downregulate RhoA activity, which may have a significant positive effect on platelet count [88].

14–3-3ς is an integrin scaffolding adaptor which cooperates with c-Src and β3 to form the 14–3-3ς/c-Src/β3 complex. The binding site of 14–3-3ς to β3 is on the highly conserved EL17 motif [89]. Destruction of 14–3-3ς/c-Src/β3 complex inhibits thrombosis without influencing hemostasis [89]. The inhibitor of 14–3-3ς, 3′,4′,7′-trihydroxyisoflavone, is a potential antithrombotic drug, without the side effects of severe bleeding [89]. Meanwhile, 14–3-3ς interacts with GPIb-IX, contributing to the binding of vWF to GPIb-IX, thus supporting αIIbβ3 activation [90].

Profilin 1 is a small actin-binding protein that is essential for actin rearrangements. Platelets lacking profilin 1 are characterized by integrin inactivation, impaired platelet function, and microthrombocytopenia [91].

ADAP also attends to αIIbβ3 activation, stable adhesion, and cytoskeletal reorganization [92]. Under shear flow, ADAP−/− mice fail to form stable thrombus in response to injury. ADAP depleted mice display an increased rebleeding time after tail cutting [92]. In addition, ADAP participates in platelet GPIb-IX-V signaling, which contributes to hemostasis and enhances the affinity of platelets for fibrinogen [93, 94]. ADAP helps talin and kindlin-3 transfer to the β3 cytoplasmic tail to support αIIbβ3 activation [95]. However, it is unclear whether ADAP directly binds to Talin and kindlin or not.

Smad4 is a transcription factor associated with transforming growth factor-beta induced integrin transcription and αIIbβ3 signal transduction. Smad4-deficient platelets display damaged functions, including platelet aggregation and spreading, fibrinogen binding and α-granule secretion, as well as clot retraction. Smad4−/− mice suffer from mild thrombocytopenia, impaired platelet aggregation, and prolonged bleeding time [96].

Many members of the protein disulfide isomerase (PDI) family (including ERp5, ERp57, ERp72 and ERp46) are required for integrin activation and platelet function. Mice depleted in these proteins have impaired platelet aggregation and defective thrombosis [97, 98].

Calcium- and DAG-regulated guanine nucleotide exchange factor I (CalDAG-GEFI) is associated with Rap1 activity, integrin activation, and leukocyte adhesion. CalDAG-GEFI−/− mice have increased bleeding time, impaired platelet aggregation, decreased thrombus formation, and neutrophil dysfunction. A 16-year-old girl with CalDAG-GEFI deficiency displays delayed αIIbβ3 activation, prolonged bleeding time, and impaired platelet granule release [99, 100]. Moreover, patients with a mutation in RASGRP2, encoding CalDAG-GEFI, demonstrate severe bleeding tendencies [101].

C-type lectin-like receptor 2 (CLEC-2) is considered a regulator of αIIbβ3 activation triggered by vWF binding to GPIbα in patients with thrombotic thrombocytopenic purpura (TTP). The hallmark of TTP is microvascular thrombosis. However, platelets lacking CLEC-2 reduce the formation of thrombosis by directly interacting with the extracellular domain of GPIbα [102].

Apoptosis signal-regulating kinase 1 (ASK1) is a member of the mitogen-activated protein kinase family. Ask1−/− platelets have defects in integrin αIIbβ3 activation and aggregation and show a severe reduction in granule secretion [103].

Proteins which inhibit αIIbβ3 activation

SHANK-associated RH domain interactor (SHARPIN) acts as a suppressor during integrin activation by directly interacting with the αIIb subunit and blocking the recruitment of talin to the cytoplasmic tail [104]. SHARPIN-null mouse platelets exhibit increased spreading on fibrinogen, increased soluble fibrinogen binding in response to sub-maximal ADP concentrations with an increased colocalization of αIIbβ3 with talin [104]. However, SHARPIN-null mice show impaired thrombus growth and slightly prolonged tail bleeding time [104]. The V267 and L276 residues of SHARPIN play a vital role in SHARPIN-mediated inhibition of αIIbβ3 [105]. SHARPIN is also associated with tumor proliferation, migration, and invasion. A recent study showed that SHARPIN negatively impacts melanoma prognosis by upregulating Rap1 [105].

Twinfilin 2a (Twf2a) is a small actin-binding protein and is indispensable for suppression of cytoskeletal dynamics by inhibiting the activity of cofilin and profilin 1. As shown above, talin is the final step of integrin activation. Talin and β3 colocalization increases in Twf2a−/− mice. Twf2a−/− mice display accelerated arterial thrombus formation, shorter tail bleeding time, and mild macrothrombocytopenia [106]. In contrast, twinfilin1 deficient mice have no apparent change in platelet function. However, deficiencies in twinfilin1 and cofilin1 result in severe macrothrombocytopenia and impaired platelet function [107].

Thioredoxin-related transmembrane protein 1 (TMX1) is a member of the polydispersity index family associated with platelet function. TMX1 exerts a negative influence on platelet function which contrasts with the positive regulation of platelet function by the other polydispersity index members, such as ERp5, ERp57, and ERp72 [97, 108, 109]. TMX1 inhibits platelet aggregation, ATP release, and αIIbβ3 activation [106, 108].

Downstream of the kinase (DOK) negatively affects αIIbβ3 activation by competing with talin [110]. Clot retraction is impaired in Dok1−/− mice [111]. DOK2 directly interacts with the NPxY747 motif of β3, while DOK1 does not [112]. Additionally, DOK2 is responsible for the adhesion function of integrin αIIbβ3. Dok1−/− mice are characterized by enhanced platelet aggregation and thrombus formation [113]. DOK1 forms a complex with 14–3-3ς via the phosphotyrosine-binding domain which interacts with the β3 cytoplasmic tail to regulate integrin αIIbβ3 bidirectional signaling [114].

Adenosine 5-diphosphate-ribosylation factors (Arfs) are small guanosine triphosphate-binding proteins that regulate endocytic trafficking, actin cytoskeleton remodeling, and lipid metabolism. Arf-6 knockout mice have enhanced platelet spreading on fibrinogen and faster clot retraction. This may be not the result of alterations in αIIbβ3 signaling because myosin light-chain phosphorylation and Rac1/RhoA activation are unaffected [115].

Other proteins also inhibit platelet function, such as CEACAM1/2, PECAM-1, CLP36, c-Cbl, PKC delta, PLD2, and so on [116]. Although deeper understanding of the proteins interacting with αIIbβ3 was explored, the exact mechanism of its activation and the switch point of bidirectional signal transduction have not been clarified fully. The purpose of exploring the mechanism is to develop safer clinical treatment. At present, αIIbβ3 has been confirmed to have prospects in thrombosis, hemostasis and tumor treatment. It is known to all that the understanding of integrin bidirectional signal is far from complete. High-resolution imaging, proteomics and structural biology should be highly valued in further research.

Above interacting factors are shown in Table 1.

Table 1 Proteins associated with the connection of ECM to the cytoskeleton

Integrin αIIbβ3 and shear stress

Under shear conditions, GPIb-IX-V/vWF interaction is regarded as the main contributor for capture and arrest of flowing platelets at injured sites of vessels, and for initiation of platelet adhesion [125]. While, αIIbβ3 is responsible for platelet stable adhesion and required to resist the washing or shedding by shear stress [56, 126,127,128], contributing to maintain the attachment-detachment balance of thrombus formation under flow [129]. β3-knockout or β3-blocked platelets hardly adhere and aggregate to the wound or wound-mimicking surface at shear stress in vivo [58] or in vitro [129, 130]. Under flow in vivo, the thrombus formed in the mice with β3 double mutations in Y747 and Y759, which have a defect in αIIbβ3 outside-in signaling, have a loosely-packed and less-activated core, compared that in wild-type mice [131]. High supraphysiologic shear stress can induce αIIbβ3 activation, but also shed αIIbβ3 from platelet surface [132]. Recently, it is found that shear stress can downregulate platelet αIIbβ3 via redistribution to αIIbβ3-enriched microparticles, rather than via enzymatic shedding of their ectodomains, and sheared platelets show reduced levels of activated αIIbβ3 and lower aggregation under the stimulation by biochemical agonists [133].

Integrin αIIbβ3 and tumor

Platelet-cancer interaction has been recognized for a long time. The tumor cells contribute to platelet activation, adhesion and aggregation. Conversely, activated platelets can interact with tumor cells facilitating tumor progression, migration, and growth [134, 135]. Under surgery, a main treatment for solid tumor, platelets adopt a TLR4-dependent manner to aggregate circulating tumor cells and lead to distant metastasis through ERK5-αIIbβ3 signaling [136]. Antiplatelet drugs contribute to tumor prevention and treatment [137, 138]. After treatment with αIIbβ3 antagonist, abciximab, the release of platelet-derived vascular endothelial growth factor and the adhesion of platelets to tumor cells are significantly impaired, restricting the extravasation of tumor cells, thus inhibiting metastasis [139]. Another study of eptifibatide, another αIIbβ3 antagonist, also confirms the role of αIIbβ3 in adhesion and hematogenous metastasis of breast cancer [140]. The oral αIIbβ3 inhibitor, XV454, inhibits lung metastasis of tumor cells [141]. Cadherin 6 (CDH6), highly expressed in renal cancer and ovarian cancer, interplays with αIIbβ3 via its RGD sequence, regulating tumor adhesion, migration and invasion [142]. These functions can be inhibited by RGD-based monoclonal antibodies or eptifibatide [142]. 1-(4-isopropyl-phenyl)-β-carboline-3-carboxylic acid (ICCA) is designed to inhibit both αIIbβ3 and P-selectin and exerts a negative role in tumor growth and thrombosis [143, 144]. Additionally, some tumor cells, such as breast cancer, melanoma, or lung carcinoma, also ectopically express αIIbβ3, besides αvβ3 [145]. The αIIbβ3-expressing tumor cell lines and the role of αIIbβ3 in them are listed in Table 2.

Table 2 Integrin αIIbβ3 in tumor cell lines

Novel inhibitors target αIIbβ3

αIIbβ3 has become a validated target for inhibiting aberrant platelet aggregation in light of its important role in platelet activation and function. Many oral αIIbβ3 antagonists, such as orbofiban, sibrafiban, xemilofiban, lotrafiban and roxifiban, have been tested, but clinical trials were stopped due to prolonged bleeding time, increased incidence of thrombocytopenia, and growing mortality [150, 151]. To date, only three drugs are approved by FDA to use in humans, mainly in percutaneous coronary intervention (PCI) to improve prognosis and reduce mortality in patients with acute myocardial infarction [152], including abciximab, eptifibatide, and tirofiban. Recently, some novel agents based on integrin have been developed. All of the antagonists list above are characterized by affecting hemostasis and lead to a risk of bleeding or prolonged bleeding time. In addition, intravenous injection limits their clinical application when they display their antiplatelet function [150]. This enlightens us to determine the switch that exhibits anti-platelet activity without affecting hemostasis. Some new targets have been found based on αIIbβ3, compared with antagonists above,they have made improvements in reducing bleeding side effects and inhibiting integrin conformational changes. RUC-4 is a novel αIIbβ3 antagonist that plays a key role in blocking platelet aggregation, thrombosis, and αIIbβ3-fibrinogen binding [153]. RUC-4 is more effective and soluble, and shows greater specificity for integrin αIIbβ3 [154]. RUC-4 was first used in 2020 on patients with coronary artery disease; it demonstrated good inhibition of platelet aggregation with no serious side effects [153]. What’s more, RUC-4 does not induce integrin extension [154].

Hr10 is a peptide designed based on the RGD sequence. Unlike tirofiban, Hr10 has no effect on the conformational changes of αIIbβ3. Mice treated with Hr10 show inhibited thrombosis, while without serious bleeding. A modified tirofiban serving as a “pure” antagonist disrupts the interaction between fibrinogen and αIIbβ3 to negatively affect platelet aggregation. Notably, it has no influence on clot contraction; thus, the hemostatic function is preserved [155]. However, a recent study confirms that clot retraction and platelet aggregation can be inhibited by modified tirofiban at an appropriate concentration. Contrary to previous conclusions, M-tirofiban can enhance ligand-induced binding sites exposure and induce conformational changes [156].

TMV-7 is another type of disintegrin isolated from snake venom. TMV-7 selectively inhibits the outside-in signal transduction mediated by Gα13, but does not affect the inside-out signal. The binding site of TMV-7 is located between the αIIb and β3 subunits around the β-propeller, and its binding does not lead to conformational changes in αIIbβ3 [157].

ANTP266 plays a pivotal role in αIIbβ3 inhibition. It has a short plasma half-life and is used to treat emergent heart diseases, such as acute myocardial infarction without increased bleeding time [158].

Dieckol, a phlorotannin, can reduce thrombosis and induce fibrinolysis [159, 160]. Dieckol suppresses αIIbβ3 activation through cAMP-PKA-VASP pathway, thereby impairing platelet aggregation and granule release, without severe bleeding risk [161].

For a more in-depth learning of integrin αIIbβ3 antagonists, several comprehensive reviews have been selected for further reading [5, 150, 162].

Conclusion

An increasing number of people die of cardiovascular thrombotic diseases each year, mainly caused by abnormal platelet activation. Integrin αIIbβ3 is the most abundant receptor in platelets and plays a key role in platelet activation. Recently, a deeper understanding of the internal mechanism of αIIbβ3 activation and more interactive proteins are discovered. However, the interactions between αIIbβ3 and related proteins are too complex to elucidate. The goal of understanding αIIbβ3 activation is to identify new targets for the development of safe antithrombotic and antitumor drugs. Further research is needed to develop an efficient antiplatelet drug that does not alter hemostasis.

Availability of data and materials

Not applicable.

Abbreviations

ADP:

Adenosine diphosphate

ADAP:

Adhesion and degranulation promoting adapter protein

Arf:

Adenosine 5-diphosphate-ribosylation factors

Ask1:

Apoptosis signal-regulating kinase 1

βTD:

β tail domain

CIB1:

Calcium- and integrin-binding protein 1

CalDAG-GEFI:

Ca22+ and DAG-regulated guanine nucleotide exchange factor I

CLEC-2:

C-type lectin-like receptor 2

Csk:

c-Src kinase

DOK:

Down-stream of tyrosine kinase

EGF:

Epidermal growth factor

ECM:

Extracellular matrix

FLN:

Filamin

FDA:

Food and Drug Administration

FERM:

band 4.1, ezrin, radixin, moesin

HPA:

Human platelet antigen

IMC:

Inner membrane clasp

IBS:

Integrin binding site

ILK:

Integrin-linked kinase

ICAP-1:

Integrin cytoplasmic domain associated protein-1

JAM-A:

Junctional adhesion molecule A

OMC:

Outer membrane clasp

PDI:

Protein disulfide isomerase

PH:

Pleckstrin homology

PSI:

Plexin-sempahorin-integrin

PIPK1γ:

Phosphatidyl inositol phosphate kinase type 1γ

PIP2:

Phosphotidylinositol 4,5-bis- phosphate

PCI:

Percutaneous coronary intervention

PI3Ks:

Phosphoinositide 3-kinases

RIAM:

Rap1-GTP-interacting adaptor molecule

SHARPIN:

Shank-associated RH domain interacting protein

SFK:

Src family kinase

Syk:

Spleen tyrosine kinase

TMX1:

Thioredoxin-related transmembrane protein 1

Talin-H:

N-terminal head domain of talin

Talin-R:

C-terminal flexible rod domain of talin

Twf2a:

Twinfilin 2a

VBS:

Vinculin binding site

References

  1. Hynes RO, Ruoslahti E, Springer TA. Reflections on Integrins-past, present, and future: the Albert Lasker basic medical research award. JAMA. 2022;328(13):1291–2.

    Article  PubMed  Google Scholar 

  2. Hynes RO. Integrins: bidirectional, allosteric signaling machines. Cell J. 2002;110(6):673–87.

    Article  CAS  Google Scholar 

  3. Luo BH, Carman CV, Springer TA. Structural basis of integrin regulation and signaling. Annu Rev Immunol. 2007;25:619–47.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Stefanini L, Lee RH, Paul DS, O'Shaughnessy EC, Ghalloussi D, Jones CI, et al. Functional redundancy between RAP1 isoforms in murine platelet production and function. Blood. 2018;132(18):1951–62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Huang J, Li X, Shi X, Zhu M, Wang J, Huang S, et al. Platelet integrin alphaIIbbeta3: signal transduction, regulation, and its therapeutic targeting. J Hematol Oncol. 2019;12(1):26.

    Article  PubMed  PubMed Central  Google Scholar 

  6. Pagani G, Gohlke H. On the contributing role of the transmembrane domain for subunit-specific sensitivity of integrin activation. Sci Rep. 2018;8(1):5733.

    Article  PubMed  PubMed Central  Google Scholar 

  7. Bennett JS, Berger BW, Billings PC. The structure and function of platelet integrins. J Thromb Haemost. 2009;7(Suppl):1200–5.

    Google Scholar 

  8. Shattil SJ, Newman PJ. Integrins: dynamic scaffolds for adhesion and signaling in platelets. Blood. 2004;104(6):1606–15.

    Article  CAS  PubMed  Google Scholar 

  9. Andre P, Denis CV, Ware J, Saffaripour S, Hynes RO, Ruggeri ZM, et al. Platelets adhere to and translocate on von Willebrand factor presented by endothelium in simulated veins. Blood. 2000;96(10):3322–8.

    Article  CAS  PubMed  Google Scholar 

  10. Lau TL, Dua V, Ulmer TS. Structure of the integrin alphaIIb transmembrane segment. J Biol Chem. 2008;283(23):16162–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Lau TL, Kim C, Ginsberg MH, Ulmer TS. The structure of the integrin alpha IIb beta 3 transmembrane complex explains integrin transmembrane signalling. Embo Journal. 2009;28(9):1351–61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Kim C, Ye F, Ginsberg MH. Regulation of integrin activation. Annu Rev Cell Dev Biol. 2011;27:321–45.

    Article  CAS  PubMed  Google Scholar 

  13. Luo BH, Springer TA, Takagi J. A specific interface between integrin transmembrane helices and affinity for ligand. PLoS Biol. 2004;2(6):e153.

    Article  PubMed  PubMed Central  Google Scholar 

  14. Cheng M, Li J, Negri A, Coller BS. Swing-out of the beta3 hybrid domain is required for alphaIIbbeta3 priming and normal cytoskeletal reorganization, but not adhesion to immobilized fibrinogen. PLoS One. 2013;8(12):e81609.

    Article  PubMed  PubMed Central  Google Scholar 

  15. Jallu V, Poulain P, Fuchs PF, Kaplan C, de Brevern AG. Modeling and molecular dynamics of HPA-1a and -1b polymorphisms: effects on the structure of the beta3 subunit of the alphaIIbbeta3 integrin. PLoS One. 2012;7(11):e47304.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Zhou DW, Thinn AMM, Zhao Y, Wang ZL, Zhu JQ. Structure of an extended beta (3) integrin. Blood. 2018;132(9):962–72.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Huynh K, Nguyen TH, Nguyen PT, Tran NQ, Vo VT, Gyenes M, et al. Leu33Pro (PlA) polymorphism of integrin beta3 modulates platelet Src pY418 and focal adhesion kinase pY397 phosphorylation in response to abnormally high shear stress. Blood Coagul Fibrinolysis. 2018;29(6):488–95.

    Article  CAS  PubMed  Google Scholar 

  18. Oliver KH, Jessen T, Crawford EL, Chung CY, Sutcliffe JS, Carneiro AM. Pro32Pro33 mutations in the integrin beta (3) PSI domain result in alpha IIb beta (3) priming and enhanced adhesion: reversal of the hypercoagulability phenotype by the Src inhibitor SKI-606. Mol Pharmacol. 2014;85(6):921–31.

    Article  PubMed  PubMed Central  Google Scholar 

  19. Zhu J, Luo BH, Xiao T, Zhang C, Nishida N, Springer TA. Structure of a complete integrin ectodomain in a physiologic resting state and activation and deactivation by applied forces. Mol Cell. 2008;32(6):849–61.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Dedden D, Schumacher S, Kelley CF, Zacharias M, Biertumpfel C, Fassler R, et al. The architecture of Talin1 reveals an autoinhibition mechanism. Cell J. 2019;179(1):120–31 e13.

    Article  CAS  Google Scholar 

  21. Wang JH. Pull and push: Talin activation for integrin signaling. Cell Res. 2012;22(11):1512–4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Wegener KL, Basran J, Bagshaw CR, Campbell ID, Roberts GC, Critchley DR, et al. Structural basis for the interaction between the cytoplasmic domain of the hyaluronate receptor layilin and the Talin F3 subdomain. J Mol Biol. 2008;382(1):112–26.

    Article  CAS  PubMed  Google Scholar 

  23. Tadokoro S, Shattil SJ, Eto K, Tai V, Liddington RC, de Pereda JM, et al. Talin binding to integrin beta tails: a final common step in integrin activation. Science. 2003;302(5642):103–6.

    Article  CAS  PubMed  Google Scholar 

  24. Goksoy E, Ma YQ, Wang X, Kong X, Perera D, Plow EF, et al. Structural basis for the autoinhibition of Talin in regulating integrin activation. Mol Cell. 2008;31(1):124–33.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Schiemer J, Bohm A, Lin L, Merrill-Skoloff G, Flaumenhaft R, Huang JS, et al. Galpha13 switch region 2 relieves Talin autoinhibition to activate alphaIIbbeta3 integrin. J Biol Chem. 2016;291(52):26598–612.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Calderwood DA, Yan B, de Pereda JM, Alvarez BG, Fujioka Y, Liddington RC, et al. The phosphotyrosine binding-like domain of Talin activates integrins. J Biol Chem. 2002;277(24):21749–58.

    Article  CAS  PubMed  Google Scholar 

  27. Caron E. Cellular functions of the Rap1 GTP-binding protein: a pattern emerges. J Cell Sci. 2003;116(3):435–40.

    Article  CAS  PubMed  Google Scholar 

  28. Han J, Lim CJ, Watanabe N, Soriani A, Ratnikov B, Calderwood DA, et al. Reconstructing and deconstructing agonist-induced activation of integrin alphaIIbbeta3. Curr Biol. 2006;16(18):1796–806.

    Article  CAS  PubMed  Google Scholar 

  29. Zhu L, Yang J, Bromberger T, Holly A, Lu F, Liu H, et al. Structure of Rap1b bound to Talin reveals a pathway for triggering integrin activation. Nat Commun. 2017;8(1):1744.

    Article  PubMed  PubMed Central  Google Scholar 

  30. Chrzanowska-Wodnicka M, Smyth SS, Schoenwaelder SM, Fischer TH, White GC 2nd. Rap1b is required for normal platelet function and hemostasis in mice. J Clin Invest. 2005;115(3):680–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Stritt S, Wolf K, Lorenz V, Vogtle T, Gupta S, Bosl MR, et al. Rap1-GTP-interacting adaptor molecule (RIAM) is dispensable for platelet integrin activation and function in mice. Blood. 2015;125(2):219–22.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Lagarrigue F, Kim C, Ginsberg MH. The Rap1-RIAM-Talin axis of integrin activation and blood cell function. Blood. 2016;128(4):479–87.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Bromberger T, Klapproth S, Rohwedder I, Zhu L, Mittmann L, Reichel CA, et al. Direct Rap1/Talin1 interaction regulates platelet and neutrophil integrin activity in mice. Blood. 2018;132(26):2754–62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Lagarrigue F, Gingras AR, Paul DS, Valadez AJ, Cuevas MN, Sun H, et al. Rap1 binding to the Talin 1 F0 domain makes a minimal contribution to murine platelet GPIIb-IIIa activation. Blood Adv. 2018;2(18):2358–68.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Has C, Chmel N, Levati L, Neri I, Sonnenwald T, Pigors M, et al. FERMT1 promoter mutations in patients with kindler syndrome. Clin Genet. 2015;88(3):248–54.

    Article  CAS  PubMed  Google Scholar 

  36. Ussar S, Moser M, Widmaier M, Rognoni E, Harrer C, Genzel-Boroviczeny O, et al. Loss of Kindlin-1 causes skin atrophy and lethal neonatal intestinal epithelial dysfunction. PLoS Genet. 2008;4(12):e1000289.

    Article  PubMed  PubMed Central  Google Scholar 

  37. Khan AA, Janke A, Shimokawa T, Zhang H. Phylogenetic analysis of kindlins suggests subfunctionalization of an ancestral unduplicated kindlin into three paralogs in vertebrates. Evol Bioinform Online. 2011;7:7–19.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Zhan J, Zhang H. Kindlins: roles in development and cancer progression. Int J Biochem Cell Biol. 2018;98:93–103.

    Article  CAS  PubMed  Google Scholar 

  39. Bialkowska K, Ma YQ, Bledzka K, Sossey-Alaoui K, Izem L, Zhang X, et al. The integrin co-activator Kindlin-3 is expressed and functional in a non-hematopoietic cell, the endothelial cell. J Biol Chem. 2010;285(24):18640–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Malinin NL, Plow EF, Byzova TV. Kindlins in FERM adhesion. Blood. 2010;115(20):4011–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Qi L, Yu Y, Chi X, Lu D, Song Y, Zhang Y, et al. Depletion of Kindlin-2 induces cardiac dysfunction in mice. Sci China-Life Sci. 2016;59(11):1123–30.

    Article  CAS  PubMed  Google Scholar 

  42. Gao H, Zhou L, Zhong Y, Ding Z, Lin S, Hou X, et al. Kindlin-2 haploinsufficiency protects against fatty liver by targeting Foxo1 in mice. Nat Commun. 2022;13(1):1025.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Zhu K, Lai Y, Cao H, Bai X, Liu C, Yan Q, et al. Kindlin-2 modulates MafA and beta-catenin expression to regulate beta-cell function and mass in mice. Nat Commun. 2020;11(1):484.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Huet-Calderwood C, Brahme NN, Kumar N, Stiegler AL, Raghavan S, Boggon TJ, et al. Differences in binding to the ILK complex determines kindlin isoform adhesion localization and integrin activation. J Cell Sci. 2014;127(Pt 19):4308–21.

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Guan SY, Chng CP, Ong LT, Tan HF, Alex Law SK, Tan SM. The binding interface of kindlin-2 and ILK involves Asp344/Asp352/Thr356 in kindlin-2 and Arg243/Arg334 in ILK. FEBS Lett. 2018;592(1):112–21.

    Article  CAS  PubMed  Google Scholar 

  46. Moser M, Nieswandt B, Ussar S, Pozgajova M, Fassler R. Kindlin-3 is essential for integrin activation and platelet aggregation. Nat Med. 2008;14(3):325–30.

    Article  CAS  PubMed  Google Scholar 

  47. Yahya AM, AlMulla AA, AlRufaye HJ, Al Dhaheri A, Elomami AS, Al-Hammadi S, et al. Case report: a case of leukocyte adhesion deficiency, type III presenting with impaired platelet function, lymphocytosis and Granulocytosis. Front Pediatr. 2021;9:713921.

  48. Li HD, Deng Y, Sun K, Yang HB, Liu J, Wang ML, et al. Structural basis of kindlin-mediated integrin recognition and activation. Proc Natl Acad Sci USA. 2017;114(35):9349–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Sun J, Xiao D, Ni Y, Zhang T, Cao Z, Xu Z, et al. Structure basis of the FERM domain of kindlin-3 in supporting integrin alphaIIbbeta3 activation in platelets. Blood Adv. 2020;4(13):3128–35.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Calderwood DA, Campbell ID, Critchley DR. Talins and kindlins: partners in integrin-mediated adhesion. Nat Rev Mol Cell Biol. 2013;14(8):503–17.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Malinin NL, Zhang L, Choi J, Ciocea A, Razorenova O, Ma YQ, et al. A point mutation in KINDLIN3 ablates activation of three integrin subfamilies in humans. Nat Med. 2009;15(3):313–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Gao J, Huang M, Lai J, Mao K, Sun P, Cao Z, et al. Kindlin supports platelet integrin alphaIIbbeta3 activation by interacting with paxillin. J Cell Sci. 2017;130(21):3764–75.

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Nguyen HTT, Xu Z, Shi X, Liu S, Schulte ML, White GC, et al. Paxillin binding to the PH domain of kindlin-3 in platelets is required to support integrin alphaIIbbeta3 outside-in signaling. J Thromb Haemost. 2021;19(12):3126–38.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Su XY, Mi JQ, Yan JS, Flevaris P, Lu YJ, Liu HC, et al. RGT, a synthetic peptide corresponding to the integrin beta 3 cytoplasmic C-terminal sequence, selectively inhibits outside-in signaling in human platelets by disrupting the interaction of integrin alpha IIb beta 3 with Src kinase. Blood. 2008;112(3):592–602.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Ablooglu AJ, Kang J, Petrich BG, Ginsberg MH, Shattil SJ. Antithrombotic effects of targeting alphaIIbbeta3 signaling in platelets. Blood. 2009;113(15):3585–92.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Shi X, Yang J, Cui X, Huang J, Long Z, Zhou Y, et al. Functional effect of the mutations similar to the cleavage during platelet activation at integrin beta3 cytoplasmic tail when expressed in mouse platelets. PLoS One. 2016;11(11):e0166136.

    Article  PubMed  PubMed Central  Google Scholar 

  57. Cole PA, Shen K, Qiao Y, Wang D. Protein tyrosine kinases Src and Csk: a tail's tale. Curr Opin Chem Biol. 2003;7(5):580–5.

    Article  CAS  PubMed  Google Scholar 

  58. Mao J, Zhu K, Long Z, Zhang H, Xiao B, Xi W, et al. Targeting the RT loop of Src SH3 in Platelets Prevents Thrombosis without Compromising Hemostasis. Adv Sci (Weinh). 2022;9(7):e2103228.

  59. Obergfell A, Eto K, Mocsai A, Buensuceso C, Moores SL, Brugge JS, et al. Coordinate interactions of Csk, Src, and Syk kinases with [alpha] IIb [beta]3 initiate integrin signaling to the cytoskeleton. J Cell Biol. 2002;157(2):265–75.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Arias-Salgado EG, Lizano S, Shattil SJ, Ginsberg MH. Specification of the direction of adhesive signaling by the integrin beta cytoplasmic domain. J Biol Chem. 2005;280(33):29699–707.

    Article  CAS  PubMed  Google Scholar 

  61. Law DA, DeGuzman FR, Heiser P, Ministri-Madrid K, Killeen N, Phillips DR. Integrin cytoplasmic tyrosine motif is required for outside-in alphaIIbbeta3 signalling and platelet function. Nature. 1999;401(6755):808–11.

    Article  CAS  PubMed  Google Scholar 

  62. Mitra SK, Schlaepfer DD. Integrin-regulated FAK-Src signaling in normal and cancer cells. Curr Opin Cell Biol. 2006;18(5):516–23.

    Article  CAS  PubMed  Google Scholar 

  63. Flevaris P, Stojanovic A, Gong H, Chishti A, Welch E, Du X. A molecular switch that controls cell spreading and retraction. J Cell Biol. 2007;179(3):553–65.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Severin S, Nash CA, Mori J, Zhao Y, Abram C, Lowell CA, et al. Distinct and overlapping functional roles of Src family kinases in mouse platelets. J Thromb Haemost. 2012;10(8):1631–45.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Senis YA, Mazharian A, Mori J. Src family kinases: at the forefront of platelet activation. Blood. 2014;124(13):2013–24.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Naik MU, Caplan JL, Naik UP. Junctional adhesion molecule-a suppresses platelet integrin alpha (IIb)beta (3) signaling by recruiting Csk to the integrin-c-Src complex. Blood. 2014;123(9):1393–402.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Naik MU, Stalker TJ, Brass LF, Naik UP. JAM-A protects from thrombosis by suppressing integrin alpha (IIb)beta (3)-dependent outside-in signaling in platelets. Blood. 2012;119(14):3352–60.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Gong H, Shen B, Flevaris P, Chow C, Lam SC, Voyno-Yasenetskaya TA, et al. G protein subunit Galpha13 binds to integrin alphaIIbbeta3 and mediates integrin “outside-in” signaling. Science. 2010;327(5963):340–3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Shen B, Zhao X, O'Brien KA, Stojanovic-Terpo A, Delaney MK, Kim K, et al. A directional switch of integrin signalling and a new anti-thrombotic strategy. Nature. 2013;503(7474):131–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Setiabakti NM, Larsson P, Hamilton JR. Phosphoinositide 3-kinases as potential targets for thrombosis prevention. Int J Mol Sci. 2022;23(9):4840.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Laurent PA, Hechler B, Solinhac R, Ragab A, Cabou C, Anquetil T, et al. Impact of PI3Kalpha (phosphoinositide 3-kinase alpha) inhibition on hemostasis and thrombosis. Arterioscler Thromb Vasc Biol. 2018;38(9):2041–53.

    Article  CAS  PubMed  Google Scholar 

  72. Schoenwaelder SM, Ono A, Sturgeon S, Chan SM, Mangin P, Maxwell MJ, et al. Identification of a unique co-operative phosphoinositide 3-kinase signaling mechanism regulating integrin alpha IIb beta 3 adhesive function in platelets. J Biol Chem. 2007;282(39):28648–58.

    Article  CAS  PubMed  Google Scholar 

  73. Theodosiou M, Widmaier M, Bottcher RT, Rognoni E, Veelders M, Bharadwaj M, et al. Kindlin-2 cooperates with Talin to activate integrins and induces cell spreading by directly binding paxillin. Elife. 2016;5:e10130.

  74. Zhu L, Liu H, Lu F, Yang J, Byzova TV, Qin J. Structural basis of Paxillin recruitment by Kindlin-2 in regulating cell adhesion. Structure [J]. 2019;27(11):1686–97 e5.

    Article  CAS  PubMed  Google Scholar 

  75. Atherton P, Lausecker F, Carisey A, Gilmore A, Critchley D, Barsukov I, et al. Relief of Talin autoinhibition triggers a force-independent association with vinculin. J Cell Biol. 2020;219(1):e201903134.

    Article  PubMed  Google Scholar 

  76. Ripamonti M, Wehrle-Haller B, de Curtis I. Paxillin: a hub for Mechano-transduction from the beta3 integrin-Talin-Kindlin Axis. Front Cell Dev Biol. 2022;10:852016.

  77. Ghatak S, Morgner J, Wickstrom SA. ILK: a pseudokinase with a unique function in the integrin-actin linkage. Biochem Soc Trans. 2013;41(4):995–1001.

    Article  CAS  PubMed  Google Scholar 

  78. Wang Y, Yan Q, Zhao Y, Liu X, Lin S, Zhang P, et al. Focal adhesion proteins Pinch1 and Pinch2 regulate bone homeostasis in mice. JCI Insight. 2019;4(22):e131692.

    Article  PubMed  PubMed Central  Google Scholar 

  79. Honda S, Shirotani-Ikejima H, Tadokoro S, Maeda Y, Kinoshita T, Tomiyama Y, et al. Integrin-linked kinase associated with integrin activation. Blood. 2009;113(21):5304–13.

    Article  CAS  PubMed  Google Scholar 

  80. Tucker KL, Sage T, Stevens JM, Jordan PA, Jones S, Barrett NE, et al. A dual role for integrin-linked kinase in platelets: regulating integrin function and alpha-granule secretion. Blood. 2008;112(12):4523–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Naik MU, Naik TU, Summer R, Naik UP. Binding of CIB1 to the alphaIIb tail of alphaIIbbeta3 is required for FAK recruitment and activation in platelets. PLoS One. 2017;12(5):e0176602.

    Article  PubMed  PubMed Central  Google Scholar 

  82. Naik MU, Nigam A, Manrai P, Millili P, Czymmek K, Sullivan M, et al. CIB1 deficiency results in impaired thrombosis: the potential role of CIB1 in outside-in signaling through integrin alpha IIb beta 3. J Thromb Haemost. 2009;7(11):1906–14.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Naik MU, Naik UP. Calcium-and integrin-binding protein regulates focal adhesion kinase activity during platelet spreading on immobilized fibrinogen. Blood. 2003;102(10):3629–36.

    Article  CAS  PubMed  Google Scholar 

  84. Gauthier NC, Roca-Cusachs P. Mechanosensing at integrin-mediated cell-matrix adhesions: from molecular to integrated mechanisms. Curr Opin Cell Biol. 2018;50:20–6.

    Article  CAS  PubMed  Google Scholar 

  85. Atherton P, Stutchbury B, Wang DY, Jethwa D, Tsang R, Meiler-Rodriguez E, et al. Vinculin controls Talin engagement with the actomyosin machinery. Nat Commun. 2015;6:10038.

  86. Atherton P, Stutchbury B, Jethwa D, Ballestrem C. Mechanosensitive components of integrin adhesions: role of vinculin. Exp Cell Res. 2016;343(1):21–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Rosa JP, Raslova H, Bryckaert M. Filamin a: key actor in platelet biology. Blood. 2019;134(16):1279–88.

    Article  PubMed  Google Scholar 

  88. Donada A, Balayn N, Shwa D, Lordier L, Ceglia V, Baschieri F, et al. Disrupted filamin a/alpha (IIb)beta (3) interaction induces macrothrombocytopenia by increasing RhoA activity. Blood. 2019;133(16):1778–88.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Shen CB, Liu M, Xu RJ, Wang G, Li J, Chen PG, et al. The 14-3-3 zeta-c-Src-integrin-beta 3 complex is vital for platelet activation. Blood. 2020;136(8):974–88.

    Article  PubMed  Google Scholar 

  90. Chen YF, Ruggeri ZM, Du XP. 14-3-3 proteins in platelet biology and glycoprotein Ib-IX signaling. Blood. 2018;131(22):2436–48.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Stritt S, Birkholz I, Beck S, Sorrentino S, Sapra KT, Viaud J, et al. Profilin 1-mediated cytoskeletal rearrangements regulate integrin function in mouse platelets. Blood Adv. 2018;2(9):1040–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Kasirer-Friede A, Ruggeri ZM, Shattil SJ. Role for ADAP in shear flow-induced platelet mechanotransduction. Blood. 2010;115(11):2274–82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Spindler M, van Eeuwijk JMM, Schurr Y, Nurden P, Nieswandt B, Stegner D, et al. ADAP deficiency impairs megakaryocyte polarization with ectopic proplatelet release and causes microthrombocytopenia. Blood. 2018;132(6):635–46.

    Article  CAS  PubMed  Google Scholar 

  94. Kasirer-Friede A, Moran B, Nagrampa-Orje J, Swanson K, Ruggeri ZM, Schraven B, et al. ADAP is required for normal alpha IIb beta 3 activation by VWF/GP Ib-IX-V and other agonists. Blood. 2007;109(3):1018–25.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Bennett JS. Platelet alpha IIb beta 3 activation: filling in the pieces. Blood. 2014;123(20):3065–6.

    Article  CAS  PubMed  Google Scholar 

  96. Wang Y, Jiang L, Mo X, Lan Y, Yang X, Liu X, et al. Megakaryocytic Smad4 regulates platelet function through Syk and ROCK2 expression. Mol Pharmacol. 2017;92(3):285–96.

    Article  CAS  PubMed  Google Scholar 

  97. Wang L, Wu Y, Zhou J, Ahmad SS, Mutus B, Garbi N, et al. Platelet-derived ERp57 mediates platelet incorporation into a growing thrombus by regulation of the alphaIIbbeta3 integrin. Blood. 2013;122(22):3642–50.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Zhou J, Wu Y, Rauova L, Koma G, Wang L, Poncz M, et al. A novel role for endoplasmic reticulum protein 46 (ERp46) in platelet function and arterial thrombosis in mice. Blood. 2022;139(13):2050–65.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Stefanini L, Roden RC, Bergmeier W. CalDAG-GEFI is at the nexus of calcium-dependent platelet activation. Blood. 2009;114(12):2506–14.

    Article  CAS  PubMed  Google Scholar 

  100. Kato H, Nakazawa Y, Kurokawa Y, Kashiwagi H, Morikawa Y, Morita D, et al. Human CalDAG-GEFI deficiency increases bleeding and delays alpha IIb beta 3 activation. Blood. 2016;128(23):2729–33.

    Article  CAS  PubMed  Google Scholar 

  101. Cattaneo M. Inherited CalDAG-GEFI deficiency. Blood. 2016;128(9):1165–7.

    Article  CAS  PubMed  Google Scholar 

  102. Shao B, Hoover C, Shi H, Kondo Y, Lee RH, Chen J, et al. Deletion of platelet CLEC-2 decreases GPIbalpha-mediated integrin alphaIIbbeta3 activation and decreases thrombosis in TTP. Blood. 2022;139(16):2523–33.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Naik MU, Patel P, Derstine R, Turaga R, Chen X, Golla K, et al. Ask1 regulates murine platelet granule secretion, thromboxane A2 generation, and thrombus formation. Blood. 2017;129(9):1197–209.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. Kasirer-Friede A, Peuhu E, Ivaska J, Shattil SJ. Platelet SHARPIN regulates platelet adhesion and inflammatory responses through associations with alphaIIbbeta3 and LUBAC. Blood Adv. 2022;6(8):2595–607.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  105. Zhou S, Liang Y, Zhang X, Liao L, Yang Y, Ouyang W, et al. SHARPIN promotes melanoma progression via Rap1 signaling pathway. J Invest Dermatol. 2020;140(2):395–403 e6.

    Article  CAS  PubMed  Google Scholar 

  106. Stritt S, Beck S, Becker IC, Vogtle T, Hakala M, Heinze KG, et al. Twinfilin 2a regulates platelet reactivity and turnover in mice. Blood. 2017;130(15):1746–56.

    Article  CAS  PubMed  Google Scholar 

  107. Becker IC, Scheller I, Wackerbarth LM, Beck S, Heib T, Aurbach K, et al. Actin/microtubule crosstalk during platelet biogenesis in mice is critically regulated by Twinfilin1 and Cofilin1. Blood Adv. 2020;4(10):2124–34.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Holbrook LM, Sasikumar P, Stanley RG, Simmonds AD, Bicknell AB, Gibbins JM. The platelet-surface thiol isomerase enzyme ERp57 modulates platelet function. J Thromb Haemost. 2012;10(2):278–88.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  109. Zhou J, Wu Y, Chen F, Wang L, Rauova L, Hayes VM, et al. The disulfide isomerase ERp72 supports arterial thrombosis in mice. Blood. 2017;130(6):817–28.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Oxley CL, Anthis NJ, Lowe ED, Vakonakis I, Campbell ID, Wegener KL. An integrin phosphorylation switch: the effect of beta3 integrin tail phosphorylation on Dok1 and Talin binding. J Biol Chem. 2008;283(9):5420–6.

    Article  CAS  PubMed  Google Scholar 

  111. Niki M, Nayak MK, Jin H, Bhasin N, Plow EF, Pandolfi PP, et al. Dok-1 negatively regulates platelet integrin alphaIIbbeta3 outside-in signalling and inhibits thrombosis in mice. Thromb Haemost. 2016;115(5):969–78.

    Article  PubMed  PubMed Central  Google Scholar 

  112. Hughan SC, Watson SP. Differential regulation of adapter proteins Dok2 and Dok1 in platelets, leading to an association of Dok2 with integrin alphaIIbbeta3. J Thromb Haemost. 2007;5(2):387–94.

    Article  CAS  PubMed  Google Scholar 

  113. Hughan SC, Spring CM, Schoenwaelder SM, Sturgeon S, Alwis I, Yuan Y, et al. Dok-2 adaptor protein regulates the shear-dependent adhesive function of platelet integrin alphaIIbbeta3 in mice. J Biol Chem. 2014;289(8):5051–60.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Chatterjee D, D'Souza A, Zhang Y, Bin W, Tan SM, Bhattacharjya S. Interaction analyses of 14-3-3zeta, Dok1, and phosphorylated integrin beta cytoplasmic tails reveal a bi-molecular switch in integrin regulation. J Mol Biol. 2018;430(21):4419–30.

    Article  CAS  PubMed  Google Scholar 

  115. Huang YJ, Joshi S, Xiang BG, Kanaho Y, Li ZY, Bouchard BA, et al. Arf6 controls platelet spreading and clot retraction via integrin alpha (IIb)beta (3) trafficking. Blood. 2016;127(11):1459–67.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  116. Li YJ, Zhu HX, Zhang D, Li HC, Ma P, Huang LY. Novel endogenous negative modulators of platelet function as potential anti-thrombotic targets. Eur Rev Med Pharmacol Sci. 2017;21(13):3146–58.

    PubMed  Google Scholar 

  117. Ginsberg MH. Integrin activation. BMB Rep. 2014;47(12):655–9.

    Article  PubMed  PubMed Central  Google Scholar 

  118. Hagel M, George EL, Kim A, Tamimi R, Opitz SL, Turner CE, et al. The adaptor protein paxillin is essential for normal development in the mouse and is a critical transducer of fibronectin signaling. Mol Cell Biol. 2002;22(3):901–15.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. Mitsios JV, Prevost N, Kasirer-Friede A, Gutierrez E, Groisman A, Abrams CS, et al. What is vinculin needed for in platelets? J Thromb Haemost. 2010;8(10):2294–304.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  120. Kiema T, Lad Y, Jiang PJ, Oxley CL, Baldassarre M, Wegener KL, et al. The molecular basis of filamin binding to integrins and competition with Talin. Molecular Cell. 2006;21(3):337–47.

    Article  CAS  PubMed  Google Scholar 

  121. Roca-Cusachs P, del Rio A, Puklin-Faucher E, Gauthier NC, Biais N, Sheetz MP. Integrin-dependent force transmission to the extracellular matrix by alpha-actinin triggers adhesion maturation. Proc Natl Acad Sci U S A. 2013;110(15):E1361–70.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  122. Pleines I, Hagedorn I, Gupta S, May F, Chakarova L, van Hengel J, et al. Megakaryocyte-specific RhoA deficiency causes macrothrombocytopenia and defective platelet activation in hemostasis and thrombosis. Blood. 2012;119(4):1054–63.

    Article  CAS  PubMed  Google Scholar 

  123. Aslan JE, McCarty OJ. Rac and Cdc42 team up for platelets. Blood [J]. 2013;122(18):3096–7.

    Article  CAS  PubMed  Google Scholar 

  124. Pleines I, Dutting S, Cherpokova D, Eckly A, Meyer I, Morowski M, et al. Defective tubulin organization and proplatelet formation in murine megakaryocytes lacking Rac1 and Cdc42. Blood. 2013;122(18):3178–87.

    Article  CAS  PubMed  Google Scholar 

  125. Feghhi S, Munday AD, Tooley WW, Rajsekar S, Fura AM, Kulman JD, et al. Glycoprotein Ib-IX-V complex transmits cytoskeletal forces that enhance platelet adhesion. Biophys J. 2016;111(3):601–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Nesbitt WS, Kulkarni S, Giuliano S, Goncalves I, Dopheide SM, Yap CL, et al. Distinct glycoprotein Ib/V/IX and integrin alpha IIbbeta 3-dependent calcium signals cooperatively regulate platelet adhesion under flow. J Biol Chem. 2002;277(4):2965–72.

    Article  CAS  PubMed  Google Scholar 

  127. Goncalves I, Hughan SC, Schoenwaelder SM, Yap CL, Yuan Y, Jackson SP. Integrin alpha IIb beta 3-dependent calcium signals regulate platelet-fibrinogen interactions under flow. Involvement of phospholipase C gamma 2. J Biol Chem [J]. 2003;278(37):34812–22.

    Article  CAS  PubMed  Google Scholar 

  128. Giuliano S, Nesbitt WS, Rooney M, Jackson SP. Bidirectional integrin alphaIIbbeta3 signalling regulating platelet adhesion under flow: contribution of protein kinase C. Biochem J. 2003;372(Pt 1):163–72.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Shi X, Yang J, Huang J, Long Z, Ruan Z, Xiao B, et al. Effects of different shear rates on the attachment and detachment of platelet thrombi. Mol Med Rep. 2016;13(3):2447–56.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  130. Huang J, Shi X, Xi W, Liu P, Long Z, Xi X. Evaluation of targeting c-Src by the RGT-containing peptide as a novel antithrombotic strategy. J Hematol Oncol. 2015;8(1):62.

    Article  PubMed  PubMed Central  Google Scholar 

  131. Stalker TJ, Welsh JD, Tomaiuolo M, Wu J, Colace TV, Diamond SL, et al. A systems approach to hemostasis: 3. Thrombus consolidation regulates intrathrombus solute transport and local thrombin activity. Blood. 2014;124(11):1824–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  132. Chen Z, Mondal NK, Ding J, Koenig SC, Slaughter MS, Griffith BP, et al. Activation and shedding of platelet glycoprotein IIb/IIIa under non-physiological shear stress. Mol Cell Biochem. 2015;409(1–2):93–101.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  133. Roka-Moiia Y, Miller-Gutierrez S, Palomares DE, Italiano JE, Sheriff J, Bluestein D, et al. Platelet dysfunction during mechanical circulatory support: elevated shear stress promotes downregulation of alphaIIbbeta3 and GPIb via microparticle shedding decreasing platelet Aggregability. Arterioscler Thromb Vasc Biol. 2021;41(4):1319–36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  134. Nurden AT. Platelets, inflammation and tissue regeneration. Thromb Haemost. 2011;105(Suppl 1):S13-33.

    CAS  PubMed  Google Scholar 

  135. Bazou D, Santos-Martinez MJ, Medina C, Radomski MW. Elucidation of flow-mediated tumour cell-induced platelet aggregation using an ultrasound standing wave trap. Br J Pharmacol. 2011;162(7):1577–89.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  136. Ren J, He J, Zhang H, Xia Y, Hu Z, Loughran P, et al. Platelet TLR4-ERK5 Axis facilitates NET-mediated capturing of circulating tumor cells and distant metastasis after surgical stress. Cancer Res. 2021;81(9):2373–85.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  137. Goubran HA, Burnouf T, Radosevic M, El-Ekiaby M. The platelet-cancer loop. Eur J Intern Med. 2013;24(5):393–400.

    Article  CAS  PubMed  Google Scholar 

  138. Xu XR, Yousef GM, Ni H. Cancer and platelet crosstalk: opportunities and challenges for aspirin and other antiplatelet agents. Blood. 2018;131(16):1777–89.

    Article  CAS  PubMed  Google Scholar 

  139. Amirkhosravi A, Amaya M, Siddiqui F, Biggerstaff JP, Meyer TV, Francis JL. Blockade of GpIIb/IIIa inhibits the release of vascular endothelial growth factor (VEGF) from tumor cell-activated platelets and experimental metastasis. Platelets. 1999;10(5):285–92.

    Article  CAS  PubMed  Google Scholar 

  140. Zhao F, Li L, Guan L, Yang H, Wu C, Liu Y. Roles for GP IIb/IIIa and alphavbeta3 integrins in MDA-MB-231 cell invasion and shear flow-induced cancer cell mechanotransduction. Cancer Lett. 2014;344(1):62–73.

    Article  CAS  PubMed  Google Scholar 

  141. Amirkhosravi A, Mousa SA, Amaya M, Blaydes S, Desai H, Meyer T, et al. Inhibition of tumor cell-induced platelet aggregation and lung metastasis by the oral GpIIb/IIIa antagonist XV454. Thromb Haemost. 2003;90(3):549–54.

    CAS  PubMed  Google Scholar 

  142. Bartolome RA, Robles J, Martin-Regalado A, Pintado-Berninches L, Burdiel M, Jaen M, et al. CDH6-activated alphaIIbbeta3 crosstalks with alpha2beta1 to trigger cellular adhesion and invasion in metastatic ovarian and renal cancers. Mol Oncol. 2021;15(7):1849–65.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  143. Chen H, Lu A, Zhang X, Gui L, Wang Y, Wu J, et al. Design and development of ICCA as a dual inhibitor of GPIIb/IIIa and P-selectin receptors. Drug Des Devel Ther. 2018;12:2097–110.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  144. Zhang X, Zhang Y, Wang Y, Wu J, Chen H, Zhao M, et al. Modifying ICCA with Trp-Phe-Phe to enhance in vivo activity and form Nano-medicine. Int J Nanomedicine. 2020;15:465–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  145. Chen YQ, Trikha M, Gao X, Bazaz R, Porter AT, Timar J, et al. Ectopic expression of platelet integrin alphaIIb beta3 in tumor cells from various species and histological origin. Int J Cancer. 1997;72(4):642–8.

    Article  CAS  PubMed  Google Scholar 

  146. Pang JH, Coupland LA, Freeman C, Chong BH, Parish CR. Activation of tumour cell ECM degradation by thrombin-activated platelet membranes: potentially a P-selectin and GPIIb/IIIa-dependent process. Clin Exp Metastasis. 2015;32(5):495–505.

    Article  CAS  PubMed  Google Scholar 

  147. Lonsdorf AS, Kramer BF, Fahrleitner M, Schonberger T, Gnerlich S, Ring S, et al. Engagement of alphaIIbbeta3 (GPIIb/IIIa) with alphanubeta3 integrin mediates interaction of melanoma cells with platelets: a connection to hematogenous metastasis. J Biol Chem. 2012;287(3):2168–78.

    Article  CAS  PubMed  Google Scholar 

  148. Boukerche H, Berthier-Vergnes O, Tabone E, Dore JF, Leung LL, McGregor JL. Platelet-melanoma cell interaction is mediated by the glycoprotein IIb-IIIa complex. Blood. 1989;74(2):658–63.

    Article  CAS  PubMed  Google Scholar 

  149. Karpatkin S, Pearlstein E, Ambrogio C, Coller BS. Role of adhesive proteins in platelet tumor interaction in vitro and metastasis formation in vivo. J Clin Invest. 1988;81(4):1012–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  150. van den Kerkhof DL, van der Meijden PEJ, Hackeng TM, Dijkgraaf I. Exogenous integrin alphaIIbbeta3 inhibitors revisited: past, present and future applications. Int J Mol Sci. 2021;22(7):3366.

    Article  PubMed  PubMed Central  Google Scholar 

  151. Yeh ET, Khan BV. The potential role of antiplatelet agents in modulating inflammatory markers in atherothrombosis. J Thromb Haemost. 2006;4(11):2308–16.

    Article  CAS  PubMed  Google Scholar 

  152. Chew DP, Bhatt DL, Sapp S, Topol EJ. Increased mortality with oral platelet glycoprotein IIb/IIIa antagonists - a meta-analysis of phase III multicenter randomized trials. Circulation. 2001;103(2):201–6.

    Article  CAS  PubMed  Google Scholar 

  153. Kereiakes DJ, Henry TD, DeMaria AN, Bentur O, Carlson M, Seng Yue C, et al. First human use of RUC-4: a nonactivating second-generation small-molecule platelet glycoprotein IIb/IIIa (integrin alphaIIbbeta3) inhibitor designed for subcutaneous point-of-care treatment of ST-segment-elevation myocardial infarction. J Am Heart Assoc. 2020;9(17):e016552.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  154. Vootukuri S, Li J, Nedelman M, Thomas C, Jiang JK, Babayeva M, et al. Preclinical studies of RUC-4, a novel platelet alphaIIbbeta3 antagonist, in non-human Primates and with human platelets. J Clin Transl Sci. 2019;3(2–3):65–74.

    Article  PubMed  PubMed Central  Google Scholar 

  155. Adair BD, Alonso JL, van Agthoven J, Hayes V, Ahn HS, Yu IS, et al. Structure-guided design of pure orthosteric inhibitors of alphaIIbbeta3 that prevent thrombosis but preserve hemostasis. Nat Commun. 2020;11(1):398.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  156. Lin FY, Lin J, Xie YH, Zhu JH, Nguyen TTH, Zhang YH, et al. A general chemical principle for creating closure inhibitors. Cell J. 2022;185(19):3533-+.

    Article  CAS  Google Scholar 

  157. Kuo YJ, Chen YR, Hsu CC, Peng HC, Huang TF. An alphaIIb beta3 antagonist prevents thrombosis without causing fc receptor gamma-chain IIa-mediated thrombocytopenia. J Thromb Haemost. 2017;15(11):2230–44.

    Article  PubMed  Google Scholar 

  158. Liu TD, Ren SH, Ding X, Xie ZL, Kong Y. A short half-life alphaIIbbeta (3) antagonist ANTP266 reduces Thrombus formation. Int J Mol Sci. 2018;19(8):2306.

    Article  PubMed  PubMed Central  Google Scholar 

  159. Yayeh T, Im EJ, Kwon TH, Roh SS, Kim S, Kim JH, et al. Hemeoxygenase 1 partly mediates the anti-inflammatory effect of dieckol in lipopolysaccharide stimulated murine macrophages. Int Immunopharmacol. 2014;22(1):51–8.

    Article  CAS  PubMed  Google Scholar 

  160. Kim TH, Ku SK, Bae JS. Antithrombotic and profibrinolytic activities of eckol and dieckol. J Cell Biochem. 2012;113(9):2877–83.

    Article  CAS  PubMed  Google Scholar 

  161. Irfan M, Kwon TH, Kwon HW, Rhee MH. Pharmacological actions of dieckol on modulation of platelet functions and thrombus formation via integrin alphaIIbbeta3 and cAMP signaling. Pharmacol Res. 2022;177:106088.

  162. Lazarovici P, Marcinkiewicz C, Lelkes PI. From Snake Venom's Disintegrins and C-type lectins to anti-platelet drugs. Toxins (Basel). 2019;11(5):303.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgments

We would like to thank Editage (www.editage.cn) for English language editing.

Funding

National Natural Science Foundation of China (81700130).

Author information

Authors and Affiliations

Authors

Contributions

H.X. and X.S. wrote the manuscript. J.H., Z.S., J.M. and X.X. reviewed the manuscript; All authors approved the manuscript.

Corresponding authors

Correspondence to Xiaodong Xi or Xiaofeng Shi.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

No competing interests.

Additional information

Publisher’s Note

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

Rights and permissions

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

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Xin, H., Huang, J., Song, Z. et al. Structure, signal transduction, activation, and inhibition of integrin αIIbβ3. Thrombosis J 21, 18 (2023). https://doi.org/10.1186/s12959-023-00463-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12959-023-00463-w

Keywords