1
|
Coffman RE, Bidone TC. Application of Funnel Metadynamics to the Platelet Integrin αIIbβ3 in Complex with an RGD Peptide. Int J Mol Sci 2024; 25:6580. [PMID: 38928286 PMCID: PMC11203998 DOI: 10.3390/ijms25126580] [Citation(s) in RCA: 0] [Impact Index Per Article: 0] [Reference Citation Analysis] [Abstract] [Key Words] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 03/25/2024] [Revised: 06/11/2024] [Accepted: 06/11/2024] [Indexed: 06/28/2024] Open
Abstract
Integrin αIIbβ3 mediates platelet aggregation by binding the Arginyl-Glycyl-Aspartic acid (RGD) sequence of fibrinogen. RGD binding occurs at a site topographically proximal to the αIIb and β3 subunits, promoting the conformational activation of the receptor from bent to extended states. While several experimental approaches have characterized RGD binding to αIIbβ3 integrin, applying computational methods has been significantly more challenging due to limited sampling and the need for a priori information regarding the interactions between the RGD peptide and integrin. In this study, we employed all-atom simulations using funnel metadynamics (FM) to evaluate the interactions of an RGD peptide with the αIIb and β3 subunits of integrin. FM incorporates an external history-dependent potential on selected degrees of freedom while applying a funnel-shaped restraint potential to limit RGD exploration of the unbound state. Furthermore, it does not require a priori information about the interactions, enhancing the sampling at a low computational cost. Our FM simulations reveal significant molecular changes in the β3 subunit of integrin upon RGD binding and provide a free-energy landscape with a low-energy binding mode surrounded by higher-energy prebinding states. The strong agreement between previous experimental and computational data and our results highlights the reliability of FM as a method for studying dynamic interactions of complex systems such as integrin.
Collapse
Affiliation(s)
- Robert E. Coffman
- Scientific Computing and Imaging Institute, University of Utah, Salt Lake City, UT 84112, USA;
| | - Tamara C. Bidone
- Scientific Computing and Imaging Institute, University of Utah, Salt Lake City, UT 84112, USA;
- Department of Biomedical Engineering, University of Utah, Salt Lake City, UT 84112, USA
- Department of Biochemistry, University of Utah, Salt Lake City, UT 84112, USA
- Department of Molecular Pharmaceutics, University of Utah, Salt Lake City, UT 84112, USA
| |
Collapse
|
2
|
Takada YK, Shimoda M, Takada Y. CD40L Activates Platelet Integrin αIIbβ3 by Binding to the Allosteric Site (Site 2) in a KGD-Independent Manner and HIGM1 Mutations Are Clustered in the Integrin-Binding Sites of CD40L. Cells 2023; 12:1977. [PMID: 37566056 PMCID: PMC10416995 DOI: 10.3390/cells12151977] [Citation(s) in RCA: 2] [Impact Index Per Article: 1.0] [Reference Citation Analysis] [Abstract] [Key Words] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 04/25/2023] [Revised: 07/20/2023] [Accepted: 07/25/2023] [Indexed: 08/12/2023] Open
Abstract
CD40L is expressed in activated T cells, and it plays a major role in immune response and is a major therapeutic target for inflammation. High IgM syndrome type 1 (HIGM1) is a congenital functional defect in CD40L/CD40 signaling due to defective CD40L. CD40L is also stored in platelet granules and transported to the surface upon platelet activation. Platelet integrin αIIbβ3 is known to bind to fibrinogen and activation of αIIbβ3 is a key event that triggers platelet aggregation. Also, the KGD motif is critical for αIIbβ3 binding and the interaction stabilizes thrombus. Previous studies showed that CD40L binds to and activates integrins αvβ3 and α5β1 and that HIGM1 mutations are clustered in the integrin-binding sites. However, the specifics of CD40L binding to αIIbβ3 were unclear. Here, we show that CD40L binds to αIIbβ3 in a KGD-independent manner using CD40L that lacks the KGD motif. Two HIGM1 mutants, S128E/E129G and L155P, reduced the binding of CD40L to the classical ligand-binding site (site 1) of αIIbβ3, indicating that αIIbβ3 binds to the outer surface of CD40L trimer. Also, CD40L bound to the allosteric site (site 2) of αIIbβ3 and allosterically activated αIIbβ3 without inside-out signaling. Two HIMG1 mutants, K143T and G144E, on the surface of trimeric CD40L suppressed CD40L-induced αIIbβ3 activation. These findings suggest that CD40L binds to αIIbβ3 in a manner different from that of αvβ3 and α5β1 and induces αIIbβ3 activation. HIGM1 mutations are clustered in αIIbβ3 binding sites in CD40L and are predicted to suppress thrombus formation and immune responses through αIIbβ3.
Collapse
Affiliation(s)
- Yoko K. Takada
- Department of Dermatology, School of Medicine, UC Davis, Sacramento, CA 95817, USA (M.S.)
| | - Michiko Shimoda
- Department of Dermatology, School of Medicine, UC Davis, Sacramento, CA 95817, USA (M.S.)
| | - Yoshikazu Takada
- Department of Dermatology, School of Medicine, UC Davis, Sacramento, CA 95817, USA (M.S.)
- Department of Biochemistry and Molecular Medicine, School of Medicine, UC Davis, Sacramento, CA 95817, USA
| |
Collapse
|
3
|
Takada YK, Simon SI, Takada Y. The C-type lectin domain of CD62P (P-selectin) functions as an integrin ligand. Life Sci Alliance 2023; 6:e202201747. [PMID: 37184585 PMCID: PMC10130748 DOI: 10.26508/lsa.202201747] [Citation(s) in RCA: 0] [Impact Index Per Article: 0] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 09/29/2022] [Revised: 04/14/2023] [Accepted: 04/17/2023] [Indexed: 05/16/2023] Open
Abstract
Recognition of integrins by CD62P has not been reported and this motivated a docking simulation using integrin αvβ3 as a target. We predicted that the C-type lectin domain of CD62P functions as a potential integrin ligand and observed that it specifically bound to soluble β3 and β1 integrins. Known inhibitors of the interaction between CD62P-PSGL-1 did not suppress the binding, whereas the disintegrin domain of ADAM-15, a known integrin ligand, suppressed recognition by the lectin domain. Furthermore, an R16E/K17E mutation in the predicted integrin-binding interface located outside of the glycan-binding site within the lectin domain, strongly inhibited CD62P binding to integrins. In contrast, the E88D mutation that strongly disrupts glycan binding only slightly affected CD62P-integrin recognition, indicating that the glycan and integrin-binding sites are distinct. Notably, the lectin domain allosterically activated integrins by binding to the allosteric site 2. We conclude that CD62P-integrin binding may function to promote a diverse set of cell-cell adhesive interactions given that β3 and β1 integrins are more widely expressed than PSGL-1 that is limited to leukocytes.
Collapse
Affiliation(s)
- Yoko K Takada
- Department of Dermatology, UC Davis School of Medicine, Sacramento, CA, USA
| | - Scott I Simon
- Department of Dermatology, UC Davis School of Medicine, Sacramento, CA, USA
- Department of Biomedical Engineering, UC Davis, Davis, CA, USA
| | - Yoshikazu Takada
- Department of Dermatology, UC Davis School of Medicine, Sacramento, CA, USA
- Department of Biochemistry and Molecular Medicine, UC Davis School of Medicine, Sacramento, CA, USA
| |
Collapse
|
4
|
Takada YK, Fujita M, Takada Y. Pro-Inflammatory Chemokines CCL5, CXCL12, and CX3CL1 Bind to and Activate Platelet Integrin αIIbβ3 in an Allosteric Manner. Cells 2022; 11:cells11193059. [PMID: 36231020 PMCID: PMC9563052 DOI: 10.3390/cells11193059] [Citation(s) in RCA: 0] [Impact Index Per Article: 0] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Download PDF] [Figures] [Journal Information] [Subscribe] [Scholar Register] [Received: 09/08/2022] [Revised: 09/23/2022] [Accepted: 09/24/2022] [Indexed: 11/16/2022] Open
Abstract
Activation of platelet integrin αIIbβ3, a key event for hemostasis and thrombus formation, is known to be mediated exclusively by inside-out signaling. We showed that inflammatory chemokines CX3CL1 and CXCL12 in previous studies, and CCL5 in this study, bound to the allosteric binding site (site 2) of vascular integrin αvβ3, in addition to the classical ligand binding site (site 1), and allosterically activated integrins independent of inside-out signaling. Since αIIbβ3 is exposed to inflammatory chemokines at increased concentrations during inflammation (e.g., cytokine/chemokine storm) and platelet activation, we hypothesized that these chemokines bind to and activate αIIbβ3 in an allosteric activation mechanism. We found that these chemokines bound to αIIbβ3. Notably, they activated soluble αIIbβ3 in 1 mM Ca2+ by binding to site 2. They activated cell-surface αIIbβ3 on CHO cells, which lack machinery for inside-out signaling or chemokine receptors, quickly (<1 min) and at low concentrations (1–10 ng/mL) compared to activation of soluble αIIbβ3, probably because chemokines bind to cell surface proteoglycans. Furthermore, activation of αIIbβ3 by the chemokines was several times more potent than 1 mM Mn2+. We propose that CCL5 and CXCL12 (stored in platelet granules) may allosterically activate αIIbβ3 upon platelet activation and trigger platelet aggregation. Transmembrane CX3CL1 on activated endothelial cells may mediate platelet–endothelial interaction by binding to and activating αIIbβ3. Additionally, these chemokines in circulation over-produced during inflammation may trigger αIIbβ3 activation, which is a possible missing link between inflammation and thrombosis.
Collapse
Affiliation(s)
- Yoko K. Takada
- Department of Dermatology, School of Medicine, University of California–Davis, 4645 Second Ave., Research III Suite 3300, Sacramento, CA 95817, USA
- Department of Biochemistry and Molecular Medicine, University of California Davis School of Medicine, Sacramento, CA 95817, USA
| | - Masaaki Fujita
- Department of Dermatology, School of Medicine, University of California–Davis, 4645 Second Ave., Research III Suite 3300, Sacramento, CA 95817, USA
| | - Yoshikazu Takada
- Department of Dermatology, School of Medicine, University of California–Davis, 4645 Second Ave., Research III Suite 3300, Sacramento, CA 95817, USA
- Department of Biochemistry and Molecular Medicine, University of California Davis School of Medicine, Sacramento, CA 95817, USA
- Correspondence:
| |
Collapse
|
5
|
Specifications of the variant curation guidelines for ITGA2B/ITGB3: ClinGen Platelet Disorder Variant Curation Panel. Blood Adv 2021; 5:414-431. [PMID: 33496739 PMCID: PMC7839359 DOI: 10.1182/bloodadvances.2020003712] [Citation(s) in RCA: 19] [Impact Index Per Article: 4.8] [Reference Citation Analysis] [Abstract] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 10/29/2020] [Accepted: 12/08/2020] [Indexed: 12/17/2022] Open
Abstract
Accurate and consistent sequence variant interpretation is critical to the correct diagnosis and appropriate clinical management and counseling of patients with inherited genetic disorders. To minimize discrepancies in variant curation and classification among different clinical laboratories, the American College of Medical Genetics and Genomics (ACMG), along with the Association for Molecular Pathology (AMP), published standards and guidelines for the interpretation of sequence variants in 2015. Because the rules are not universally applicable to different genes or disorders, the Clinical Genome Resource (ClinGen) Platelet Disorder Expert Panel (PD-EP) has been tasked to make ACMG/AMP rule specifications for inherited platelet disorders. ITGA2B and ITGB3, the genes underlying autosomal recessive Glanzmann thrombasthenia (GT), were selected as the pilot genes for specification. Eight types of evidence covering clinical phenotype, functional data, and computational/population data were evaluated in the context of GT by the ClinGen PD-EP. The preliminary specifications were validated with 70 pilot ITGA2B/ITGB3 variants and further refined. In the final adapted criteria, gene- or disease-based specifications were made to 16 rules, including 7 with adjustable strength; no modification was made to 5 rules; and 7 rules were deemed not applicable to GT. Employing the GT-specific ACMG/AMP criteria to the pilot variants resulted in a reduction of variants classified with unknown significance from 29% to 20%. The overall concordance with the initial expert assertions was 71%. These adapted criteria will serve as guidelines for GT-related variant interpretation to increase specificity and consistency across laboratories and allow for better clinical integration of genetic knowledge into patient care.
Collapse
|
6
|
Cai C, Sun H, Hu L, Fan Z. Visualization of integrin molecules by fluorescence imaging and techniques. ACTA ACUST UNITED AC 2021; 45:229-257. [PMID: 34219865 PMCID: PMC8249084 DOI: 10.32604/biocell.2021.014338] [Citation(s) in RCA: 4] [Impact Index Per Article: 1.0] [Reference Citation Analysis] [Abstract] [Key Words] [Track Full Text] [Download PDF] [Figures] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/15/2022]
Abstract
Integrin molecules are transmembrane αβ heterodimers involved in cell adhesion, trafficking, and signaling. Upon activation, integrins undergo dynamic conformational changes that regulate their affinity to ligands. The physiological functions and activation mechanisms of integrins have been heavily discussed in previous studies and reviews, but the fluorescence imaging techniques -which are powerful tools for biological studies- have not. Here we review the fluorescence labeling methods, imaging techniques, as well as Förster resonance energy transfer assays used to study integrin expression, localization, activation, and functions.
Collapse
Affiliation(s)
- Chen Cai
- Department of Immunology, School of Medicine, UConn Health, Farmington, 06030, USA
| | - Hao Sun
- Department of Medicine, University of California, San Diego, La Jolla, 92093, USA
| | - Liang Hu
- Cardiovascular Institute of Zhengzhou University, Department of Cardiology, The First Affiliated Hospital of Zhengzhou University, Zhengzhou, 450051, China
| | - Zhichao Fan
- Department of Immunology, School of Medicine, UConn Health, Farmington, 06030, USA
| |
Collapse
|
7
|
Diabetes and Hyperglycemia Affect Platelet GPIIIa Expression. Effects on Adhesion Potential of Blood Platelets from Diabetic Patients under In Vitro Flow Conditions. Int J Mol Sci 2020; 21:ijms21093222. [PMID: 32370146 PMCID: PMC7247361 DOI: 10.3390/ijms21093222] [Citation(s) in RCA: 10] [Impact Index Per Article: 2.0] [Reference Citation Analysis] [Abstract] [Key Words] [Track Full Text] [Download PDF] [Figures] [Journal Information] [Subscribe] [Scholar Register] [Received: 03/25/2020] [Revised: 04/28/2020] [Accepted: 04/30/2020] [Indexed: 02/07/2023] Open
Abstract
Blood platelets play a crucial role in the early stages of atherosclerosis development. The process is believed to require firm adhesion of platelets to atherosclerosis-prone sites of the artery. However, little evidence exists regarding whether the blood platelets of individuals with pathological conditions associated with atherosclerosis have higher potential for adhesion. This process is to a large extent dependent on receptors present on the platelet membrane. Therefore, the aim of the presented study was to determine whether blood platelets from diabetic patients have higher capacity of adhesion under flow conditions and how diabetes affects one of the crucial platelet receptors involved in the process of adhesion-GPIIIa. The study compares the ability of platelets from non-diabetic and diabetic humans to interact with fibrinogen and von Willebrand factor, two proteins found in abundance on an inflamed endothelium, under flow conditions. The activation and reactivity of the blood platelets were also characterized by flow cytometry. Platelets from diabetic patients did not demonstrate enhanced adhesion to either studied protein, although they presented increased basal activation and responsiveness towards low concentrations of agonists. Platelets from diabetic patients were characterized by lower expression of GPIIIa, most likely due to an enhanced formation of platelet-derived microparticles PMPs, as supported by the observation of elevated concentration of this integrin and of GPIIIa-positive PMPs in plasma. We conclude that altered functionality of blood platelets in diabetes does not increase their adhesive potential. Increased glycation and decrease in the amount of GPIIIa on platelets may be partially responsible for this effect. Therefore, higher frequency of interactions of platelets with the endothelium, which is observed in animal models of diabetes, is caused by other factors. A primary cause may be a dysfunctional vascular wall.
Collapse
|
8
|
Zhu K, Takada Y, Nakajima K, Sun Y, Jiang J, Zhang Y, Zeng Q, Takada Y, Zhao M. Expression of integrins to control migration direction of electrotaxis. FASEB J 2019; 33:9131-9141. [PMID: 31116572 PMCID: PMC6662972 DOI: 10.1096/fj.201802657r] [Citation(s) in RCA: 14] [Impact Index Per Article: 2.3] [Reference Citation Analysis] [Abstract] [Key Words] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 12/08/2018] [Accepted: 04/15/2019] [Indexed: 02/05/2023]
Abstract
Proper control of cell migration is critically important in many biologic processes, such as wound healing, immune surveillance, and development. Much progress has been made in the initiation of cell migration; however, little is known about termination and sometimes directional reversal. During active cell migration, as in wound healing, development, and immune surveillance, the integrin expression profile undergoes drastic changes. Here, we uncovered the extensive regulatory and even opposing roles of integrins in directional cell migration in electric fields (EFs), a potentially important endogenous guidance mechanism. We established cell lines that stably express specific integrins and determined their responses to applied EFs with a high throughput screen. Expression of specific integrins drove cells to migrate to the cathode or to the anode or to lose migration direction. Cells expressing αMβ2, β1, α2, αIIbβ3, and α5 migrated to the cathode, whereas cells expressing β3, α6, and α9 migrated to the anode. Cells expressing α4, αV, and α6β4 lost directional electrotaxis. Manipulation of α9 molecules, one of the molecular directional switches, suggested that the intracellular domain is critical for the directional reversal. These data revealed an unreported role for integrins in controlling stop, go, and reversal activity of directional migration of mammalian cells in EFs, which might ensure that cells reach their final destination with well-controlled speed and direction.-Zhu, K., Takada, Y., Nakajima, K., Sun, Y., Jiang, J., Zhang, Y., Zeng, Q., Takada, Y., Zhao, M. Expression of integrins to control migration direction of electrotaxis.
Collapse
Affiliation(s)
- Kan Zhu
- Department of Dermatology, School of Medicine, University of California–Davis, Sacramento, California, USA
- Institute for Regenerative Cures, University of California–Davis, Sacramento, California, USA
- State Key Laboratory of Trauma, Burns, and Combined Injury, Institute of Surgery Research, Third Military Medical University, Chongqing, China
| | - Yoko Takada
- Department of Dermatology, School of Medicine, University of California–Davis, Sacramento, California, USA
| | - Kenichi Nakajima
- Department of Dermatology, School of Medicine, University of California–Davis, Sacramento, California, USA
- Institute for Regenerative Cures, University of California–Davis, Sacramento, California, USA
| | - Yaohui Sun
- Department of Dermatology, School of Medicine, University of California–Davis, Sacramento, California, USA
- Institute for Regenerative Cures, University of California–Davis, Sacramento, California, USA
| | - Jianxin Jiang
- State Key Laboratory of Trauma, Burns, and Combined Injury, Institute of Surgery Research, Third Military Medical University, Chongqing, China
| | - Yan Zhang
- Department of Dermatology, School of Medicine, University of California–Davis, Sacramento, California, USA
- Institute for Regenerative Cures, University of California–Davis, Sacramento, California, USA
- Bioelectromagnetics Laboratory, Zhejiang University School of Medicine, Hangzhou, China
| | - Qunli Zeng
- Bioelectromagnetics Laboratory, Zhejiang University School of Medicine, Hangzhou, China
| | - Yoshikazu Takada
- Department of Dermatology, School of Medicine, University of California–Davis, Sacramento, California, USA
| | - Min Zhao
- Department of Dermatology, School of Medicine, University of California–Davis, Sacramento, California, USA
- Institute for Regenerative Cures, University of California–Davis, Sacramento, California, USA
- Department of Ophthalmology and Vision Science, School of Medicine, University of California–Davis, Sacramento, California, USA
| |
Collapse
|
9
|
Waters L, Padula MP, Marks DC, Johnson L. Cryopreserved platelets demonstrate reduced activation responses and impaired signaling after agonist stimulation. Transfusion 2017; 57:2845-2857. [PMID: 28905392 DOI: 10.1111/trf.14310] [Citation(s) in RCA: 22] [Impact Index Per Article: 2.8] [Reference Citation Analysis] [Abstract] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 04/09/2017] [Revised: 07/11/2017] [Accepted: 07/11/2017] [Indexed: 12/25/2022]
Abstract
BACKGROUND Room temperature-stored (20-24°C) platelets (PLTs) have a shelf life of 5 days, making it logistically challenging to supply remote medical centers with PLT products. Cryopreservation of PLTs in dimethyl sulfoxide (DMSO) and storage at -80°C enables an extended shelf life up to 2 years. Although cryopreserved PLTs have been widely characterized under resting conditions, their ability to undergo agonist-induced activation is yet to be fully explored. STUDY DESIGN AND METHODS Buffy coat PLTs were cryopreserved at -80°C with 5% to 6% DMSO and sampled before freezing and after thawing. PLTs were analyzed under resting conditions and after agonist stimulation with adenosine diphosphate, collagen, or thrombin receptor-activating peptide-6. The expression of activation markers, microparticle formation, and calcium mobilization were analyzed by flow cytometry. Soluble PLT proteins present in the PLT supernatant were examined by enzyme-linked immunosorbent assay. Protein phosphorylation was investigated with Western blotting. RESULTS After cryopreservation, PLTs displayed increased surface activation markers and higher basal calcium levels. Cryopreserved PLTs demonstrated diminished aggregation responses. Additionally, cryopreserved PLTs showed a limited ability to become activated (as measured by CD62P and phosphatidylserine exposure and cytokine release) after agonist stimulation. A reduction in the abundance and phosphorylation of key signaling proteins (Akt, Src, Lyn, ERK, and p38) was seen in cryopreserved PLTs. CONCLUSIONS Cryopreservation of PLTs induces dramatic changes to the basal PLT phenotype and renders them largely nonresponsive to agonist stimulation, likely due to the alterations in signal transduction. Therefore, further efforts are required to understand how cryopreserved PLTs achieve their hemostatic effect once transfused.
Collapse
Affiliation(s)
- Lauren Waters
- Research & Development, Australian Red Cross Blood Service, Alexandria, NSW, Australia
| | - Matthew P Padula
- Proteomics Core Facility, University of Technology Sydney, Sydney, NSW, Australia
| | - Denese C Marks
- Research & Development, Australian Red Cross Blood Service, Alexandria, NSW, Australia
| | - Lacey Johnson
- Research & Development, Australian Red Cross Blood Service, Alexandria, NSW, Australia
| |
Collapse
|
10
|
αIIbβ3 binding to a fibrinogen fragment lacking the γ-chain dodecapeptide is activation dependent and EDTA inducible. Blood Adv 2017; 1:417-428. [PMID: 29296957 DOI: 10.1182/bloodadvances.2017004689] [Citation(s) in RCA: 11] [Impact Index Per Article: 1.4] [Reference Citation Analysis] [Abstract] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 01/11/2017] [Accepted: 01/24/2017] [Indexed: 12/24/2022] Open
Abstract
Platelet integrin receptor αIIbβ3 supports platelet aggregation by binding fibrinogen. The interaction between the fibrinogen C-terminal γ-chain peptide composed of residues γ-404-411 (GAKQAGDV) and the Arg-Gly-Asp (RGD) binding pocket on αIIbβ3 is required for fibrinogen-mediated platelet aggregation, but data suggest that other ancillary binding sites on both fibrinogen and αIIbβ3 may lead to higher-affinity fibrinogen binding and clot retraction. To identify additional sites, we analyzed the ability of platelets and cells expressing normal and mutant αIIbβ3 to adhere to an immobilized fibrinogen plasmin fragment that lacks intact γ-404-411 ('D98'). We found the following: (1) Activated, but not unactivated, platelets adhere well to immobilized 'D98.' (2) Cells expressing constitutively active αIIbβ3 mutants, but not cells expressing normal αIIbβ3 or αVβ3, adhere well to 'D98.' (3) Monoclonal antibodies 10E5 and 7E3 inhibit the adhesion to 'D98' of activated platelets and cells expressing constitutively active αIIbβ3, as do small-molecule inhibitors that bind to the RGD pocket. (4) EDTA paradoxically induces normal αIIbβ3 to interact with 'D98.' Because molecular modeling and molecular dynamics simulations suggested that the αIIb L151-D159 helix may contribute to the interaction with 'D98,' we studied an αIIbβ3 mutant in which the αIIb 148-166 loop was swapped with the corresponding αV loop; it failed to bind to fibrinogen or 'D98.' Our data support a model in which conformational changes in αIIbβ3 and/or fibrinogen after platelet activation and the interaction between γ-404-411 and the RGD binding pocket make new ancillary sites available that support higher-affinity fibrinogen binding and clot retraction.
Collapse
|
11
|
Wood B, Padula MP, Marks DC, Johnson L. Refrigerated storage of platelets initiates changes in platelet surface marker expression and localization of intracellular proteins. Transfusion 2016; 56:2548-2559. [PMID: 27460096 DOI: 10.1111/trf.13723] [Citation(s) in RCA: 40] [Impact Index Per Article: 4.4] [Reference Citation Analysis] [Abstract] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 03/31/2016] [Revised: 05/18/2016] [Accepted: 06/01/2016] [Indexed: 12/30/2022]
Abstract
BACKGROUND Platelets (PLTs) are currently stored at room temperature (22°C), which limits their shelf life, primarily due to the risk of bacterial growth. Alternatives to room temperature storage include PLT refrigeration (2-6°C), which inhibits bacterial growth, thus potentially allowing an extension of shelf life. Additionally, refrigerated PLTs appear more hemostatically active than conventional PLTs, which may be beneficial in certain clinical situations. However, the mechanisms responsible for this hemostatic function are not well characterized. The aim of this study was to assess the protein profile of refrigerated PLTs in an effort to understand these functional consequences. STUDY DESIGN AND METHODS Buffy coat PLTs were pooled, split, and stored either at room temperature (20-24°C) or under refrigerated (2-6°C) conditions (n = 8 in each group). PLTs were assessed for changes in external receptor expression and actin filamentation using flow cytometry. Intracellular proteomic changes were assessed using two-dimensional gel electrophoresis and Western blotting. RESULTS PLT refrigeration significantly reduced the abundance of glycoproteins (GPIb, GPIX, GPIIb, and GPIV) on the external membrane. However, refrigeration resulted in the increased expression of high-affinity integrins (αIIbβ3 and β1) and activation and apoptosis markers (CD62P, CD63, and phosphatidylserine). PLT refrigeration substantially altered the abundance and localization of several cytoskeletal proteins and resulted in an increase in actin filamentation, as measured by phalloidin staining. CONCLUSION Refrigerated storage of PLTs induces significant changes in the expression and localization of both surface-expressed and intracellular proteins. Understanding these proteomic changes may help to identify the mechanisms resulting in the refrigeration-associated alterations in PLT function and clearance.
Collapse
Affiliation(s)
- Ben Wood
- Research & Development, Australian Red Cross Blood Service, Alexandria, NSW, Australia.,Proteomics Core Facility, University of Technology Sydney, Sydney, NSW, Australia
| | - Matthew P Padula
- Proteomics Core Facility, University of Technology Sydney, Sydney, NSW, Australia
| | - Denese C Marks
- Research & Development, Australian Red Cross Blood Service, Alexandria, NSW, Australia
| | - Lacey Johnson
- Research & Development, Australian Red Cross Blood Service, Alexandria, NSW, Australia.
| |
Collapse
|
12
|
Müller MA, Brunie L, Bächer AS, Kessler H, Gottschalk KE, Reuning U. Cytoplasmic salt bridge formation in integrin αvß3 stabilizes its inactive state affecting integrin-mediated cell biological effects. Cell Signal 2014; 26:2493-503. [PMID: 25041847 DOI: 10.1016/j.cellsig.2014.07.013] [Citation(s) in RCA: 12] [Impact Index Per Article: 1.1] [Reference Citation Analysis] [Abstract] [Key Words] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 06/16/2014] [Accepted: 07/09/2014] [Indexed: 02/01/2023]
Abstract
Heterodimeric integrin receptors are mediators of cell adhesion, motility, invasion, proliferation, and survival. By this, they are crucially involved in (tumor) cell biological behavior. Integrins trigger signals bidirectionally across cell membranes: by outside-in, following binding of protein ligands of the extracellular matrix, and by inside-out, where proteins are recruited to ß-integrin cytoplasmic tails resulting in conformational changes leading to increased integrin binding affinity and integrin activation. Computational modeling and experimental/mutational approaches imply that associations of integrin transmembrane domains stabilize the low-affinity integrin state. Moreover, a cytoplasmic interchain salt bridge is discussed to contribute to a tight clasp of the α/ß-membrane-proximal regions; however, its existence and physiological relevance for integrin activation are still a controversial issue. In order to further elucidate the functional role of salt bridge formation, we designed mutants of the tumor biologically relevant integrin αvß3 by mutually exchanging the salt bridge forming amino acid residues on each chain (αvR995D and ß3D723R). Following transfection of human ovarian cancer cells with different combinations of wild type and mutated integrin chains, we showed that loss of salt bridge formation strengthened αvß3-mediated adhesion to vitronectin, provoked recruitment of cytoskeletal proteins, such as talin, and induced integrin signaling, ultimately resulting in enhanced cell migration, proliferation, and activation of integrin-related signaling molecules. These data support the notion of a functional relevance of integrin cytoplasmic salt bridge disruption during integrin activation.
Collapse
Affiliation(s)
- Martina A Müller
- Clinical Research Unit, Dept. for Obstetrics & Gynecology, Technische Universitaet München, Munich, Germany
| | - Leonora Brunie
- Clinical Research Unit, Dept. for Obstetrics & Gynecology, Technische Universitaet München, Munich, Germany
| | - Anne-Sophie Bächer
- Clinical Research Unit, Dept. for Obstetrics & Gynecology, Technische Universitaet München, Munich, Germany
| | - Horst Kessler
- Institute for Advanced Study and Centre of Integrated Protein Science, Department Chemie, Technische Universitaet München, Garching, Germany; Chemistry Department, Faculty of Science, King Abdulaziz University, Jeddah, Saudi Arabia
| | | | - Ute Reuning
- Clinical Research Unit, Dept. for Obstetrics & Gynecology, Technische Universitaet München, Munich, Germany.
| |
Collapse
|
13
|
Rui X, Mehrbod M, Van Agthoven JF, Anand S, Xiong JP, Mofrad MRK, Arnaout MA. The α-subunit regulates stability of the metal ion at the ligand-associated metal ion-binding site in β3 integrins. J Biol Chem 2014; 289:23256-23263. [PMID: 24975416 DOI: 10.1074/jbc.m114.581470] [Citation(s) in RCA: 7] [Impact Index Per Article: 0.6] [Reference Citation Analysis] [Abstract] [Key Words] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/06/2022] Open
Abstract
The aspartate in the prototypical integrin-binding motif Arg-Gly-Asp binds the integrin βA domain of the β-subunit through a divalent cation at the metal ion-dependent adhesion site (MIDAS). An auxiliary metal ion at a ligand-associated metal ion-binding site (LIMBS) stabilizes the metal ion at MIDAS. LIMBS contacts distinct residues in the α-subunits of the two β3 integrins αIIbβ3 and αVβ3, but a potential role of this interaction on stability of the metal ion at LIMBS in β3 integrins has not been explored. Equilibrium molecular dynamics simulations of fully hydrated β3 integrin ectodomains revealed strikingly different conformations of LIMBS in unliganded αIIbβ3 versus αVβ3, the result of stronger interactions of LIMBS with αV, which reduce stability of the LIMBS metal ion in αVβ3. Replacing the αIIb-LIMBS interface residue Phe(191) in αIIb (equivalent to Trp(179) in αV) with Trp strengthened this interface and destabilized the metal ion at LIMBS in αIIbβ3; a Trp(179) to Phe mutation in αV produced the opposite but weaker effect. Consistently, an F191/W substitution in cellular αIIbβ3 and a W179/F substitution in αVβ3 reduced and increased, respectively, the apparent affinity of Mn(2+) to the integrin. These findings offer an explanation for the variable occupancy of the metal ion at LIMBS in αVβ3 structures in the absence of ligand and provide new insights into the mechanisms of integrin regulation.
Collapse
Affiliation(s)
- Xianliang Rui
- Leukocyte Biology and Inflammation Program and Massachusetts General Hospital and Harvard Medical School, Charlestown, Massachusetts 02129
| | - Mehrdad Mehrbod
- Departments of Bioengineering and Mechanical Engineering, University of California, Berkeley, California 94720
| | - Johannes F Van Agthoven
- Structural Biology Program, Department of Medicine, Massachusetts General Hospital and Harvard Medical School, Charlestown, Massachusetts 02129 and
| | - Saurabh Anand
- Leukocyte Biology and Inflammation Program and Massachusetts General Hospital and Harvard Medical School, Charlestown, Massachusetts 02129
| | - Jian-Ping Xiong
- Structural Biology Program, Department of Medicine, Massachusetts General Hospital and Harvard Medical School, Charlestown, Massachusetts 02129 and
| | - Mohammad R K Mofrad
- Departments of Bioengineering and Mechanical Engineering, University of California, Berkeley, California 94720.
| | - M Amin Arnaout
- Leukocyte Biology and Inflammation Program and Massachusetts General Hospital and Harvard Medical School, Charlestown, Massachusetts 02129; Structural Biology Program, Department of Medicine, Massachusetts General Hospital and Harvard Medical School, Charlestown, Massachusetts 02129 and.
| |
Collapse
|
14
|
Podolnikova NP, Yakovlev S, Yakubenko VP, Wang X, Gorkun OV, Ugarova TP. The interaction of integrin αIIbβ3 with fibrin occurs through multiple binding sites in the αIIb β-propeller domain. J Biol Chem 2013; 289:2371-83. [PMID: 24338009 DOI: 10.1074/jbc.m113.518126] [Citation(s) in RCA: 40] [Impact Index Per Article: 3.3] [Reference Citation Analysis] [Abstract] [Key Words] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 12/15/2022] Open
Abstract
The currently available antithrombotic agents target the interaction of platelet integrin αIIbβ3 (GPIIb-IIIa) with fibrinogen during platelet aggregation. Platelets also bind fibrin formed early during thrombus growth. It was proposed that inhibition of platelet-fibrin interactions may be a necessary and important property of αIIbβ3 antagonists; however, the mechanisms by which αIIbβ3 binds fibrin are uncertain. We have previously identified the γ370-381 sequence (P3) in the γC domain of fibrinogen as the fibrin-specific binding site for αIIbβ3 involved in platelet adhesion and platelet-mediated fibrin clot retraction. In the present study, we have demonstrated that P3 can bind to several discontinuous segments within the αIIb β-propeller domain of αIIbβ3 enriched with negatively charged and aromatic residues. By screening peptide libraries spanning the sequence of the αIIb β-propeller, several sequences were identified as candidate contact sites for P3. Synthetic peptides duplicating these segments inhibited platelet adhesion and clot retraction but not platelet aggregation, supporting the role of these regions in fibrin recognition. Mutant αIIbβ3 receptors in which residues identified as critical for P3 binding were substituted for homologous residues in the I-less integrin αMβ2 exhibited reduced cell adhesion and clot retraction. These residues are different from those that are involved in the coordination of the fibrinogen γ404-411 sequence and from auxiliary sites implicated in binding of soluble fibrinogen. These results map the binding of fibrin to multiple sites in the αIIb β-propeller and further indicate that recognition specificity of αIIbβ3 for fibrin differs from that for soluble fibrinogen.
Collapse
|
15
|
Epitope mapping for monoclonal antibody reveals the activation mechanism for αVβ3 integrin. PLoS One 2013; 8:e66096. [PMID: 23840404 PMCID: PMC3688720 DOI: 10.1371/journal.pone.0066096] [Citation(s) in RCA: 5] [Impact Index Per Article: 0.4] [Reference Citation Analysis] [Abstract] [Track Full Text] [Download PDF] [Figures] [Journal Information] [Subscribe] [Scholar Register] [Received: 01/30/2013] [Accepted: 05/02/2013] [Indexed: 12/02/2022] Open
Abstract
Epitopes for a panel of anti-αVβ3 monoclonal antibodies (mAbs) were investigated to explore the activation mechanism of αVβ3 integrin. Experiments utilizing αV/αIIb domain-swapping chimeras revealed that among the nine mAbs tested, five recognized the ligand-binding β-propeller domain and four recognized the thigh domain, which is the upper leg of the αV chain. Interestingly, the four mAbs included function-blocking as well as non-functional mAbs, although they bound at a distance from the ligand-binding site. The epitopes for these four mAbs were further determined using human-to-mouse αV chimeras. Among the four, P3G8 recognized an amino acid residue, Ser-528, located on the side of the thigh domain, while AMF-7, M9, and P2W7 all recognized a common epitope, Ser-462, that was located close to the α-genu, where integrin makes a sharp bend in the crystal structure. Fibrinogen binding studies for cells expressing wild-type αVβ3 confirmed that AMF-7, M9, and P2W7 were inhibitory, while P3G8 was non-functional. However, these mAbs were all unable to block binding when αVβ3 was constrained in its extended conformation. These results suggest that AMF-7, M9, and P2W7 block ligand binding allosterically by stabilizing the angle of the bend in the bent conformation. Thus, a switchblade-like movement of the integrin leg is indispensable for the affinity regulation of αVβ3 integrin.
Collapse
|
16
|
Recognition of highly restricted regions in the β-propeller domain of αIIb by platelet-associated anti-αIIbβ3 autoantibodies in primary immune thrombocytopenia. Blood 2012; 120:1499-509. [DOI: 10.1182/blood-2012-02-409995] [Citation(s) in RCA: 23] [Impact Index Per Article: 1.8] [Reference Citation Analysis] [Abstract] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/20/2022] Open
Abstract
AbstractPlatelet-associated (PA) IgG autoantibodies play an essential role in primary immune thrombocytopenia (ITP). However, little is known about the epitopes of these Abs. This study aimed to identify critical binding regions for PA anti-αIIbβ3 Abs. Because PA anti-αIIbβ3 Abs bound poorly to mouse αIIbβ3, we created human-mouse chimera constructs. We first examined 76 platelet eluates obtained from patients with primary ITP. Of these, 26 harbored PA anti-αIIbβ3 Abs (34%). Further analysis of 15 patients who provided sufficient materials showed that the epitopes of these Abs were mainly localized in the N-terminal half of the β-propeller domain in αIIb (L1-W235). We could identify 3 main recognition sites in the region; 2 eluates recognized a conformation formed by the W1:1-2 and W2:3-4 loops, 5 recognized W1:2-3, and 4 recognized W3:4-1. The remaining 4 eluates could not be defined by the binding sites. Within these regions, we identified residues critical for binding, including S29 and R32 in W1:1-2; G44 and P45 in W1:2-3; and P135, E136, and R139 in W2:3-4. Of 11 eluates whose recognition sites were identified, 5 clearly showed restricted κ/λ-chain usage. These results suggested that PA anti-αIIbβ3 Abs in primary ITP tended to recognize highly restricted regions of αIIb with clonality.
Collapse
|
17
|
Yan Y, Li Y, Zhang S, Ai C. Studies of tricyclic piperazine/piperidine furnished molecules as novel integrin αvβ3/αIIbβ3 dual antagonists using 3D-QSAR and molecular docking. J Mol Graph Model 2011; 29:747-62. [PMID: 21273104 DOI: 10.1016/j.jmgm.2010.12.008] [Citation(s) in RCA: 3] [Impact Index Per Article: 0.2] [Reference Citation Analysis] [Abstract] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 07/30/2010] [Revised: 12/10/2010] [Accepted: 12/20/2010] [Indexed: 10/18/2022]
Abstract
The development of injectable integrin α(v)β(3)/α(IIb)β(3) dual antagonists attracts much attention of research for treating of acute ischemic diseases in recent years. In this work, based on a dataset composed of 102 tricyclic piperazine/piperidine furnished dual α(v)β(3) and α(IIb)β(3) antagonists, a variety of in silico modeling approaches including the comparative molecular field analysis (CoMFA), comparative similarity indices analysis (CoMSIA), and molecular docking were applied to reveal the requisite 3D structural features impacting the biological activities. Our statistical results show that the ligand-based 3D-QSAR models for both the α(v)β(3) and α(IIb)β(3) studies exhibited satisfactory internal and external predictability, i.e., for the CoMFA models, results of Q(2)=0.48, R(ncv)(2)=0.87, R(pred)(2)=0.71 for α(v)β(3) and Q(2)=0.50, R(ncv)(2)=0.85, R(pred)(2)=0.72 for α(IIb)β(3) analysis were obtained, and for the CoMSIA ones, the outcomes of Q(2)=0.55, R(ncv)(2)=0.90, R(pred)(2)=0.72 for α(v)β(3) and Q(2)=0.52, R(ncv)(2)=0.88, R(pred)(2)=0.74 for α(IIb)β(3) were achieved respectively. In addition, through a comparison between 3D-QSAR contour maps and docking results, it is revealed that that the most crucial interactions occurring between the tricyclic piperazine/piperidine derivatives and α(v)β(3)/α(IIb)β(3) receptor ligand binding pocket are H-bonding, and the key amino acids impacting the interactions are Arg214, Asn215, Ser123, and Lys253 for α(v)β(3), but Arg214, Asn215, Ser123 and Tyr190 for α(IIb)β(3) receptors, respectively. Halogen-containing groups at position 15 and 16, benzene sulfonamide substituent at position 23, and the replacement of piperazine with 4-aminopiperidine of ring B may increase the α(v)β(3)/α(IIb)β(3) antagonistic activity. The potencies for antagonists to inhibit isolated α(v)β(3) and α(IIb)β(3) are linear correlated, indicating that similar interaction mechanisms may exist for the series of molecules. To our best knowledge this is the first report on 3D-QSAR modeling of these dual α(v)β(3)/α(IIb)β(3) antagonists. The results obtained should provide information for better understanding of the mechanism of antagonism and thus be helpful in design of novel potent dual α(v)β(3)/α(IIb)β(3) antagonists.
Collapse
Affiliation(s)
- Yulian Yan
- School of Chemical Engineering, Dalian University of Technology, Dalian, 116012, Liaoning, China
| | | | | | | |
Collapse
|
18
|
Cilia La Corte AL, Philippou H, Ariëns RAS. Role of fibrin structure in thrombosis and vascular disease. ADVANCES IN PROTEIN CHEMISTRY AND STRUCTURAL BIOLOGY 2011; 83:75-127. [PMID: 21570666 DOI: 10.1016/b978-0-12-381262-9.00003-3] [Citation(s) in RCA: 55] [Impact Index Per Article: 3.9] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Subscribe] [Scholar Register] [Indexed: 12/11/2022]
Abstract
Fibrin clot formation is a key event in the development of thrombotic disease and is the final step in a multifactor coagulation cascade. Fibrinogen is a large glycoprotein that forms the basis of a fibrin clot. Each fibrinogen molecule is comprised of two sets of Aα, Bβ, and γ polypeptide chains that form a protein containing two distal D regions connected to a central E region by a coiled-coil segment. Fibrin is produced upon cleavage of the fibrinopeptides by thrombin, which can then form double-stranded half staggered oligomers that lengthen into protofibrils. The protofibrils then aggregate and branch, yielding a three-dimensional clot network. Factor XIII, a transglutaminase, cross-links the fibrin stabilizing the clot protecting it from mechanical stress and proteolytic attack. The mechanical properties of the fibrin clot are essential for its function as it must prevent bleeding but still allow the penetration of cells. This viscoelastic property is generated at the level of each individual fiber up to the complete clot. Fibrinolysis is the mechanism of clot removal, and involves a cascade of interacting zymogens and enzymes that act in concert with clot formation to maintain blood flow. Clots vary significantly in structure between individuals due to both genetic and environmental factors and this has an effect on clot stability and susceptibility to lysis. There is increasing evidence that clot structure is a determinant for the development of disease and this review will discuss the determinants for clot structure and the association with thrombosis and vascular disease.
Collapse
Affiliation(s)
- Amy L Cilia La Corte
- Division of Cardiovascular and Diabetes Research, Section on Mechanisms of Thrombosis, Leeds Institute for Genetics Health and Therapeutics, Faculty of Medicine and Health, University of Leeds, Leeds, United Kingdom
| | | | | |
Collapse
|
19
|
Kamata T, Handa M, Ito S, Sato Y, Ohtani T, Kawai Y, Ikeda Y, Aiso S. Structural requirements for activation in alphaIIb beta3 integrin. J Biol Chem 2010; 285:38428-37. [PMID: 20884611 DOI: 10.1074/jbc.m110.139667] [Citation(s) in RCA: 21] [Impact Index Per Article: 1.4] [Reference Citation Analysis] [Abstract] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 12/14/2022] Open
Abstract
Integrins are postulated to undergo structural rearrangement from a low affinity bent conformer to a high affinity extended conformer upon activation. However, some reports have shown that a bent conformer is capable of binding a ligand, whereas another report has shown that integrin extension does not absolutely lead to activation. To clarify whether integrin affinity is indeed regulated by the so-called switchblade-like movement, we have engineered a series of mutant αIIbβ3 integrins that are constrained specifically in either a bent or an extended conformation. These mutant αIIbβ3 integrins were expressed in mammalian cells, and fibrinogen binding to these cells was examined. The bent integrins were created through the introduction of artificial disulfide bridges in the β-head/β-tail interface. Cells expressing bent integrins all failed to bind fibrinogen unless pretreated with DTT to disrupt the disulfide bridges. The extended integrins were created by introducing N-glycosylation sites in amino acid residues located close to the α-genu, where the integrin legs fold backward. Among these mutants, activation was maximized in one integrin with an N-glycosylation site located behind the α-genu. This extension-induced activation was completely blocked when the swing-out of the hybrid domain was prevented. These results suggest that the bent and extended conformers represent low affinity and high affinity conformers, respectively, and that extension-induced activation depends on the swing-out of the hybrid domain. Taken together, these results are consistent with the current hypothesis that integrin affinity is regulated by the switchblade-like movement of the integrin legs.
Collapse
Affiliation(s)
- Tetsuji Kamata
- Department of Anatomy, Keio University School of Medicine, 35 Shinanomachi, Shinjuku-ku, Tokyo 160-8582, Japan.
| | | | | | | | | | | | | | | |
Collapse
|
20
|
Pillitteri D, Pilgrimm AK, Kirchmaier CM. Novel Mutations in the GPIIb and GPIIIa Genes in Glanzmann Thrombasthenia. Transfus Med Hemother 2010; 37:268-277. [PMID: 21113249 PMCID: PMC2980511 DOI: 10.1159/000320258] [Citation(s) in RCA: 13] [Impact Index Per Article: 0.9] [Reference Citation Analysis] [Abstract] [Key Words] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 05/06/2010] [Accepted: 08/17/2010] [Indexed: 02/03/2023] Open
Abstract
BACKGROUND: Glanzmann thrombasthenia (GT) is an inherited autosomal recessive platelet disorder characterized by a complete or partial lack, or mutation, of the GPIIb/IIIa complex (integrin α(IIb)β(3)) on the thrombocytes' surface, leading to a severe bleeding syndrome. MATERIAL AND METHODS: Molecular genetic analysis was performed in patients with suspected GT. The aim of the present study was the identification of new natural variants, their impact on platelet function, and their relation to the risk of bleeding. RESULTS: Expression of the platelet integrin α(IIb)β(3) was determined by flow cytometry. Mutations were identified through sequencing of cDNA and genomic DNA. In addition, platelet function studies (PAC-binding, aggregations) were implemented. The study included 25 patients revealing 13 mutations (GPIIb: n = 9; GPIIIa: n = 4). Two of the 13 mutations were previously described (T207I; L214P). The remaining mutations have not been published yet, whereas 1 mutation in 2 unrelated families was identical (3062 T→C). CONCLUSION: All patients with less than 25% of present α(IIb)β(3) have a medical history of bleeding.
Collapse
Affiliation(s)
- Daniele Pillitteri
- Deutsche Klinik für Diagnostik, Sektion Innere Medizin I, Arbeitsgruppe: «Thrombose, Hämostase und vaskuläre Medizin», Wiesbaden, Germany
| | | | | |
Collapse
|
21
|
Basani RB, Zhu H, Thornton MA, Soto CS, Degrado WF, Kowalska MA, Bennett JS, Poncz M. Species differences in small molecule binding to alpha IIb beta 3 are the result of sequence differences in 2 loops of the alpha IIb beta propeller. Blood 2009; 113:902-10. [PMID: 18987357 PMCID: PMC2630275 DOI: 10.1182/blood-2008-09-177337] [Citation(s) in RCA: 11] [Impact Index Per Article: 0.7] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 09/03/2008] [Accepted: 10/07/2008] [Indexed: 11/20/2022] Open
Abstract
Compared with human platelets, rodent platelets are less responsive to peptides and peptidomimetics containing an arginine-glycine-aspartic acid (RGD) motif. Using chimeric human-rat alphaIIbbeta3 molecules, we found that this difference in Arg-Gly-Asp-Ser (RGDS) sensitivity was the result of amino acid substitutions at residues 157, 159, and 162 in the W3:4-1 loop and an Asp-His replacement at residue 232 in the W4:4-1 loop of the alphaIIb beta propeller. Introducing the entire rat W3:4-1 and W4:4-1 loops into human alphaIIbbeta3 also decreased the inhibitory effect of the disintegrins, echistatin and eristostatin, and the alphaIIbbeta3 antagonists, tirofiban and eptifibatide, on fibrinogen binding, whereas the specific point mutations did not. This suggests that RGDS interacts with alphaIIb in a different manner than with these small molecules. None of these species-based substitutions affected the ability of alphaIIbbeta3 to interact with RGD-containing macromolecules. Thus, human von Willebrand factor contains an RGD motif and binds equally well to adenosine diphosphate-stimulated human and rodent platelets, implying that other motifs are responsible for maintaining ligand binding affinity. Many venoms contain RGD-based toxins. Our data suggest that these species amino acids differences in the alphaIIb beta-propeller represent an evolutionary response by rodents to maintain hemostasis while concurrently protecting against RGD-containing toxins.
Collapse
Affiliation(s)
- Ramesh B Basani
- Division of Hematology, Children's Hospital of Philadelphia, PA19104, USA
| | | | | | | | | | | | | | | |
Collapse
|
22
|
Matsumoto A, Kamata T, Takagi J, Iwasaki K, Yura K. Key interactions in integrin ectodomain responsible for global conformational change detected by elastic network normal-mode analysis. Biophys J 2008; 95:2895-908. [PMID: 18515366 PMCID: PMC2527288 DOI: 10.1529/biophysj.108.131045] [Citation(s) in RCA: 19] [Impact Index Per Article: 1.1] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Download PDF] [Figures] [Journal Information] [Subscribe] [Scholar Register] [Received: 02/05/2008] [Accepted: 05/06/2008] [Indexed: 01/03/2023] Open
Abstract
Integrin, a membrane protein with a huge extracellular domain, participates in cell-cell and cell-extracellular-matrix interactions for metazoan. A group of integrins is known to perform a large-scale structural change when the protein is activated, but the activation mechanism and generality of the conformational change remain to be elucidated. We performed normal-mode analysis of the elastic network model on integrin alpha(V)beta(3) ectodomain in the bent form and identified key residues that influenced molecular motions. Iterative normal-mode calculations demonstrated that the specific nonbonded interactions involving the key residues work as a snap to keep integrin in the bent form. The importance of the key residues for the conformational change was further verified by mutation experiments, in which integrin alpha(IIb)beta(3) was used. The conservation pattern of amino acid residues among the integrin family showed that the characteristic pattern of residues seen around these key residues is found in the limited groups of integrin beta-chains. This conservation pattern suggests that the molecular mechanism of the conformational change relying on the interactions found in integrin alpha(V)beta(3) is unique to the limited types of integrins.
Collapse
Affiliation(s)
- Atsushi Matsumoto
- Quantum Bioinformatics Team, Center for Computational Science and Engineering, Japan Atomic Energy Agency, 8-1 Umemidai, Kizugawa, Kyoto 619-0215, Japan.
| | | | | | | | | |
Collapse
|
23
|
Bunch TA, Kendall TL, Shakalya K, Mahadevan D, Brower DL. Modulation of ligand binding by alternative splicing of the alphaPS2 integrin subunit. J Cell Biochem 2007; 102:211-23. [PMID: 17372926 DOI: 10.1002/jcb.21288] [Citation(s) in RCA: 2] [Impact Index Per Article: 0.1] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 12/27/2022]
Abstract
The Drosophila alphaPS2 integrin subunit is found in two isoforms. alphaPS2C contains 25 residues not found in alphaPS2m8, encoded by the alternative eighth exon. Previously, it was shown that cells expressing alphaPS2C spread more effectively than alphaPS2m8 cells on fragments of the ECM protein Tiggrin, and that alphaPS2C-containing integrins are relatively insensitive to depletion of Ca(2+). Using a ligand mimetic probe for Tiggrin affinity (TWOW-1), we show that the affinity of alphaPS2CbetaPS for this ligand is much higher than that of alphaPS2m8betaPS. However, the two isoforms become more similar in the presence of activating levels of Mn(2+). Modeling indicates that the exon 8-encoded residues replace the third beta strand of the third blade of the alpha subunit beta-propeller structure, and generate an exaggerated loop between this and the fourth strand. alphaPS2 subunits with the extra loop structure but with an m8-like third strand, or subunits with a C-like strand but an m8-like short loop, both fail to show alphaPS2C-like affinity for TWOW-1. Surprisingly, a single C > m8-like change at the third strand-loop transition point is sufficient to make alphaPS2C require Ca(2+) for function, despite the absence of any known cation binding site in this region. These data indicate that alternative splicing in integrin alpha subunit extracellular domains may affect ligand affinity via relatively subtle alterations in integrin conformation. These results may have relevance for vertebrate alpha6 and alpha7, which are alternatively spliced at the same site.
Collapse
Affiliation(s)
- Thomas A Bunch
- Department of Molecular and Cellular Biology, Arizona Cancer Center, 1515 N. Campbell Ave., Tucson, Arizona 85724, USA.
| | | | | | | | | |
Collapse
|
24
|
Martínez M, Labiós M, Gabriel F. [Platelet activation and hypercholesterolemia in the pathogenesis of deep vein thrombosis]. Med Clin (Barc) 2006; 127:669-72. [PMID: 17169286 DOI: 10.1157/13094823] [Citation(s) in RCA: 0] [Impact Index Per Article: 0] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/21/2022]
Abstract
Currently it is accepted that deep vein thrombosis is a multifactorial event in which the presence of activated platelets and also plasmatic lipids seems to play a pivotal role that it is not well established in the scientific bibliography. Due to the non consensus state about these topics between the different groups working in these aspects, the topic involving deep vein thrombosis-platelets-lipids, and also their interactions, still is an interesting area of investigation, in which it is necessary to carry out studies with the aim of establishing risk factors, initial diagnostic methods and clinical assays to probe the efficacy of new therapies.
Collapse
Affiliation(s)
- Marcial Martínez
- Departamento de Biopatología Clínica, Hospital Universitario La Fe, Valencia, España.
| | | | | |
Collapse
|
25
|
Abstract
The platelet integrin alpha(IIb)beta(3) is required for platelet aggregation. Like other integrins, alpha(IIb)beta(3) resides on cell surfaces in an equilibrium between inactive and active conformations. Recent experiments suggest that the shift between these conformations involves a global reorganization of the alpha(IIb)beta(3) molecule and disruption of constraints imposed by the heteromeric association of the alpha(IIb) and beta(3) transmembrane and cytoplasmic domains. The biochemical, biophysical, and ultrastructural results that support this conclusion are discussed in this Review.
Collapse
Affiliation(s)
- Joel S Bennett
- Hematology-Oncology Division, Department of Medicine, University of Pennsylvania School of Medicine, Philadelphia, Pennsylvania 19104-6058, USA.
| |
Collapse
|
26
|
Labiós M, Martínez M, Gabriel F, Guiral V, Martínez E, Aznar J. Effect of atorvastatin upon platelet activation in hypercholesterolemia, evaluated by flow cytometry. Thromb Res 2005; 115:263-70. [PMID: 15668185 DOI: 10.1016/j.thromres.2004.08.016] [Citation(s) in RCA: 28] [Impact Index Per Article: 1.4] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 05/27/2004] [Revised: 07/30/2004] [Accepted: 08/05/2004] [Indexed: 11/16/2022]
Abstract
Hyperlipidemia is a well established risk factor for cardiovascular disease and atherothrombotic events, in which platelet activation also plays a significant role. However, very few studies have addressed platelet activation in hypercholesterolemia, the potential effect of lipid lowering drugs upon platelet hyperfunction, and the question of whether changes in the latter are correlated to normalization of plasma lipids. This study used whole blood flow cytometry to assess in vivo and in vitro platelet activation in a group of 33 patients with hypercholesterolemia, and also the ex vivo effect of atorvastatin (20 mg/day) upon such activation. A control group of 40 normolipidemic volunteers matched in terms of age, sex and added risk factors to the patient group was used. The results showed that hypercholesterolemic patients had in vivo a significantly greater percentage of GPIIb/IIIa- and phosphatidylserine-positive platelets compared with the control group (4.62+/-3.51% and 2.58+/-1.19% versus 2.73+/-1.08% and 1.54+/-0.68%, respectively). In vitro response of CD62 expression to thrombin was also greater in the patients than in the controls (92.51+/-6.00% versus 89.63+/-10.72%, p<0.05). Atorvastatin therapy normalized platelet hyperfunction in the patients studied and reduced GPIIb/IIIa response to ADP (from 82.65+/-6.43% to 75.84+/-4.89%, p<0.01). A significant correlation can be seen between such normalization and the decrease in plasma levels of total and LDL cholesterol.
Collapse
Affiliation(s)
- Manuel Labiós
- Department of Internal Medicine, Clinic University Hospital, Valencia, Spain
| | | | | | | | | | | |
Collapse
|
27
|
Kamata T, Handa M, Sato Y, Ikeda Y, Aiso S. Membrane-proximal α/β Stalk Interactions Differentially Regulate Integrin Activation. J Biol Chem 2005; 280:24775-83. [PMID: 15863495 DOI: 10.1074/jbc.m409548200] [Citation(s) in RCA: 41] [Impact Index Per Article: 2.1] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/06/2022] Open
Abstract
The affinity of integrin-ligand interaction is regulated extracellularly by divalent cations and intracellularly by inside-out signaling. We report here that the extracellular, membrane-proximal alpha/beta stalk interactions not only regulate cation-induced integrin activation but also play critical roles in propagating inside-out signaling. Two closely related integrins, alphaIIbbeta3 and alphaVbeta3, share high structural homology and bind to similar ligands in an RGD-dependent manner. Despite these structural and functional similarities, they exhibit distinct responses to Mn(2+). Although alphaVbeta3 showed robust ligand binding in the presence of Mn(2+), alphaIIbbeta3 showed a limited increase but failed to achieve full activation. Swapping alpha stalk regions between alphaIIb and alphaV revealed that the alpha stalk, but not the ligand-binding head region, was responsible for the difference. A series of alphaIIb/alphaV domain-swapping chimeras were constructed to identify the responsible domain. Surprisingly, the minimum component required to render alphaIIbbeta3 susceptible to Mn(2+) activation was the alphaV calf-2 domain, which does not contain any divalent cation-binding sites. The calf-2 domain makes interface with beta epidermal growth factor 4 and beta tail domain in three-dimensional structure. The effect of calf-2 domain swapping was partially reproduced by mutating the specific amino acid residues in the calf-2/epidermal growth factor 4-beta tail domain interface. When this interface was constrained by an artificially introduced disulfide bridge, the Mn(2+)-induced alphaVbeta3-fibrinogen interaction was significantly impaired. Notably, a similar disulfide bridge completely abrogated fibrinogen binding to alphaIIbbeta3 when alphaIIbbeta3 was activated by cytoplasmic tail truncation to mimic inside-out signaling. Thus, disruption/formation of the membrane-proximal alpha/beta stalk interface may act as an on/off switch that triggers integrin-mediated bidirectional signaling.
Collapse
Affiliation(s)
- Tetsuji Kamata
- Departments of Anatomy, Transfusion Medicine and Cell Therapy, and Internal Medicine, Keio University School of Medicine, Tokyo 160-8582, Japan.
| | | | | | | | | |
Collapse
|
28
|
Kamata T, Ambo H, Puzon-McLaughlin W, Tieu KK, Handa M, Ikeda Y, Takada Y. Critical cysteine residues for regulation of integrin alphaIIbbeta3 are clustered in the epidermal growth factor domains of the beta3 subunit. Biochem J 2004; 378:1079-82. [PMID: 14690453 PMCID: PMC1224037 DOI: 10.1042/bj20031701] [Citation(s) in RCA: 56] [Impact Index Per Article: 2.7] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Received: 11/06/2003] [Accepted: 12/22/2003] [Indexed: 11/17/2022]
Abstract
Chemical or enzymic reduction/oxidation of integrin cysteine residues (e.g. by reducing agents and protein disulphide isomerase) may be a mechanism for regulating integrin function. It has also been proposed that unique cysteine residues in the integrin beta3 subunit are involved in the regulation of alphaIIbbeta3. In the present study, we studied systematically the role of disulphide bonds in beta3 on the ligand-binding function of alphaIIbbeta3 by mutating individual cysteine residues of beta3 to serine. We found that the disulphide bonds that are critical for alphaIIbbeta3 regulation are clustered within the EGF (epidermal growth factor) domains. Interestingly, disrupting only a single disulphide bond in the EGF domains was enough to activate alphaIIbbeta3 fully. In contrast, only two (of 13) disulphide bonds tested outside the EGF domains activated alphaIIbbeta3. These results suggest that the disulphide bonds in the EGF domains should be intact to keep alphaIIbbeta3 in an inactive state, and that there is no unique cysteine residue in the EGF domain critical for regulating the receptor. The cysteine residues in the EGF domains are potential targets for chemical or enzymic reduction.
Collapse
Affiliation(s)
- Tetsuji Kamata
- Department of Cell Biology, The Scripps Research Institute, 10550 N. Torrey Pines Road, La Jolla, CA 92037, USA
| | | | | | | | | | | | | |
Collapse
|
29
|
Chung J, Rader C, Popkov M, Hur YM, Kim HK, Lee YJ, Barbas CF. Integrin αIIbβ3 specific synthetic human monoclonal antibodies and HCDR3 peptides that potently inhibit platelet aggregation. FASEB J 2003; 18:361-3. [PMID: 14688205 DOI: 10.1096/fj.03-0586fje] [Citation(s) in RCA: 14] [Impact Index Per Article: 0.6] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/11/2022]
Abstract
The interaction of fibrinogen with integrin alphaIIbbeta3 (GPIIb/IIIa), in part mediated by an RGD tripeptide motif, is an essential step in platelet aggregation. Based on their inhibition of platelet aggregation, three integrin alphaIIbbeta3 inhibitors are clinically approved. The clinically most widely used integrin alphaIIbbeta3 inhibitor abciximab is a chimeric mouse/human antibody that induces thrombocytopenia, often severe, in 1-2% of patients due to a human anti-mouse antibody (HAMA) response. In addition, unlike other ligands mimicking small molecular drugs, abciximab cross-reacts with integrin alphavbeta3 and alphaMbeta2. Here we used phage display to select monoclonal antibodies specific to integrin alphaIIbbeta3 from a synthetic human antibody library based on the randomized HCDR3 sequence VGXXXRADXXXYAMDV. The selected antibodies revealed a strong consensus in HCDR3 (V(V/W)CRAD(K/R)RC) and high specificity toward integrin alphaIIbbeta3 but not to other RGD binding integrins such as alphavbeta3, alphavbeta5, and alpha5beta1. The selected antibodies as well as three synthetic peptides (VWCRADRRC, VWCRADKRC, and VVCRADRRC) whose sequences were derived from the HCDR3 sequences of the selected antibodies strongly inhibited the interaction between integrin alphaIIbbeta3 and fibrinogen and platelet aggregation ex vivo. To our knowledge, these are the first fully human monoclonal antibodies that are specific to integrin alphaIIbbeta3 and can potently inhibit platelet aggregation.
Collapse
Affiliation(s)
- Junho Chung
- The Skaggs Institute for Chemical Biology and the Department of Molecular Biology, The Scripps Research Institute, La Jolla, California 92037, USA
| | | | | | | | | | | | | |
Collapse
|
30
|
Tarui T, Andronicos N, Czekay RP, Mazar AP, Bdeir K, Parry GC, Kuo A, Loskutoff DJ, Cines DB, Takada Y. Critical role of integrin alpha 5 beta 1 in urokinase (uPA)/urokinase receptor (uPAR, CD87) signaling. J Biol Chem 2003; 278:29863-72. [PMID: 12754207 DOI: 10.1074/jbc.m304694200] [Citation(s) in RCA: 57] [Impact Index Per Article: 2.6] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/06/2022] Open
Abstract
Urokinase-type plasminogen activator (uPA) induces cell adhesion and chemotactic movement. uPA signaling requires its binding to uPA receptor (uPAR/CD87), but how glycosylphosphatidylinositol-anchored uPAR mediates signaling is unclear. uPAR is a ligand for several integrins (e.g. alpha 5 beta 1) and supports cell-cell interaction by binding to integrins on apposing cells (in trans). We studied whether binding of uPAR to alpha 5 beta 1 in cis is involved in adhesion and migration of Chinese hamster ovary cells in response to immobilized uPA. This process was temperature-sensitive and required mitogen-activated protein kinase activation. Anti-uPAR antibody or depletion of uPAR blocked, whereas overexpression of uPAR enhanced, cell adhesion to uPA. Adhesion to uPA was also blocked by deletion of the growth factor domain (GFD) of uPA and by anti-GFD antibody, whereas neither the isolated uPA kringle nor serine protease domain supported adhesion directly. Interestingly, anti-alpha 5 antibody, RGD peptide, and function-blocking mutations in alpha 5 beta 1 blocked adhesion to uPA. uPA-induced cell migration also required GFD, uPAR, and alpha 5 beta 1, but alpha 5 beta 1 alone did not support uPA-induced adhesion and migration. Thus, binding of uPA causes uPAR to act as a ligand for alpha 5 beta 1 to induce cell adhesion, intracellular signaling, and cell migration. We demonstrated that uPA induced RGD-dependent binding of uPAR to alpha 5 beta 1 in solution. These results suggest that uPA-induced adhesion and migration of Chinese hamster ovary cells occurs as a consequence of (a) uPA binding to uPAR through GFD, (b) the subsequent binding of a uPA.uPAR complex to alpha 5 beta 1 via uPAR, and (c) signal transduction through alpha 5 beta 1.
Collapse
Affiliation(s)
- Takehiko Tarui
- Department of Cell Biology, The Scripps Research Institute, La Jolla, California 92037, USA
| | | | | | | | | | | | | | | | | | | |
Collapse
|
31
|
Labiós M, Martínez M, Gabriel F, Gómez-Biedma S, Guiral V, Vivó M, Aznar J. Flow cytometric analysis of platelet activation in hypertensive patients. Effect of doxazosin. Thromb Res 2003; 110:203-8. [PMID: 14512082 DOI: 10.1016/s0049-3848(03)00377-3] [Citation(s) in RCA: 9] [Impact Index Per Article: 0.4] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 10/27/2022]
Abstract
The percentage of spontaneously activated platelets and the platelet response to several agonists were studied in 26 hypertensive patients. The percentage of platelets expressing glycoprotein (GP) IIb/IIIa in its active conformation (GPIIb/IIIa*), P-selectin and phosphatidylserine (PS) was measured by flow cytometry at baseline and 1 and 2 months after treatment with doxazosin (4 mg/day). The response to ADP and Ca2+ ionophore was also evaluated. The results were compared with those of a control group of 71 normotensive volunteers. Spontaneous platelet activation was higher in patients than in controls (P-selectin-positive results in 4.4+/-2.0% patients vs. 2.7+/-1.7 controls, p<0.05; phosphatidylserine-positive results in 0.7+/-0.4% vs. 0.5+/-0.3%, respectively, p<0.05), and higher in response to ionophore action (phosphatidylserine-positive results 51.8+/-11.1% vs. 43.4+/-11.7%, p<0.01). Platelet activation in patients decreased after 2 months of doxazosin administration compared to baseline (P-selectin-positive results 2.7+/-1.4% vs. 4.4+/-2.0%, p<0.05; phosphatidylserine-positive results 0.3+/-0.2% vs. 0.7+/-0.4%, p<0.05). No significant differences were noted in GPIIb/IIIa*. The clinical significance of normalization of platelet activity by doxazosin remains to be established.
Collapse
Affiliation(s)
- Manuel Labiós
- Hypertension Unit, Internal Medicine, Hospital Clínico Universitario, Valencia, Spain
| | | | | | | | | | | | | |
Collapse
|
32
|
Abstract
AlphaIIbbeta3, the major membrane protein on the surface of platelets, is a member of the integrin family of heterodimeric adhesion receptors. The alphaIIb and beta3 subunits are each composed of a short cytoplasmic tail, a single transmembrane domain, and a large, extracellular region that consists of a series of linked domains. Recent structural analyses have provided insights into the organization of this and other integrins and how a signal is initiated at its cytoplasmic tail to transform the extracellular domain of alphaIIbbeta3 into a functional receptor for fibrinogen or von Willebrand factor to support platelet aggregation and thrombus formation. These functions of alphaIIbbeta3 have been targeted for antithrombotic therapy, and intravenous alphaIIbbeta3 antagonists have been remarkably effective in the setting of percutaneous coronary interventions, showing both short-term and long-term mortality benefits. However, the development of oral antagonists has been abandoned on the basis of excess of mortality in clinical trials, and the extension of therapy with existing alphaIIbbeta3 antagonists to broadly treat acute coronary syndromes has not fully met expectations. An in-depth understanding of how antagonists engage and influence the function of alphaIIbbeta3 and platelets in the context of the new structural insights may explain its salutary and potential deleterious effects.
Collapse
Affiliation(s)
- Martin J Quinn
- Joseph J. Jacobs Center for Thrombosis and Vascular Biology, Department of Molecular Cardiology/NB50, Cleveland Clinic Foundation, 9500 Euclid Ave, Cleveland, Ohio 44195, USA
| | | | | | | | | |
Collapse
|
33
|
Vella F, Thielens NM, Bersch B, Arlaud GJ, Frachet P. A recombinant chimeric epidermal growth factor-like module with high binding affinity for integrins. J Biol Chem 2003; 278:19834-43. [PMID: 12654911 DOI: 10.1074/jbc.m301470200] [Citation(s) in RCA: 6] [Impact Index Per Article: 0.3] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/06/2022] Open
Abstract
Integrins are cell surface receptors involved in numerous pathological processes such as metastasis invasion and abnormal angiogenesis. To target these receptors, the epidermal growth factor (EGF)-like domain of human complement protease C1r was used as a natural scaffold to design chimeric modules containing the RGD motif. Here we report a high yield bacterial expression system and its application to the production of two such modules, EGF-RGD and V2, the latter variant mimicking the RGD-containing domain of disintegrins. These modules were characterized chemically, and their biological activity was investigated by cellular assays using various Chinese hamster ovary cell lines expressing beta1 and beta3 integrins and by surface plasmon resonance spectroscopy. Remarkably, the modifications leading to the V2 variant had differential effects on the interaction with beta3 and beta1 integrins. The disintegrin-like V2 module exhibited enhanced binding affinities compared with EGF-RGD, with KD values of 7.2 nm for alpha5beta1 (a 4-fold decrease) and 3.5 nm for alphavbeta3 (a 1.5-fold decrease), comparable with the values determined for natural integrin ligands. Analysis by NMR spectroscopy also revealed a differential dynamic behavior of the RGD motif in the EGF-RGD and V2 variants, providing insights into the structural basis of their relative binding efficiency. These novel RGD-containing EGF modules open the way to the design of improved variants with selective affinity for particular integrins and their use as carriers for other biologically active modules.
Collapse
Affiliation(s)
- Fanny Vella
- Laboratoire d'Enzymologie Moléculaire, Institut de Biologie Structurale Jean-Pierre Ebel, (Commissariat à l'Energie Atomique CNRS, Université Joseph Fourier), Grenoble 38027 Cedex 1, France
| | | | | | | | | |
Collapse
|
34
|
Abstract
Integrins are alphabeta heterodimeric cell-surface receptors that are vital to the survival and function of nucleated cells. They recognize aspartic-acid- or a glutamic-acid-based sequence motifs in structurally diverse ligands. Integrin recognition of most ligands is divalent cation dependent and conformationally sensitive. In addition to this common property, there is an underlying binding specificity between integrins and ligands for which there has been no structural basis. The recently reported crystal structures of the extracellular segment of an integrin in its unliganded state and in complex with a prototypical Arg-Gly-Asp (RGD) ligand have provided an atomic basis for cation-mediated binding of aspartic-acid-based ligands to integrins. They also serve as a basis for modelling other integrins in complex with larger physiologic ligands. These models provide new insights into the molecular basis for ligand binding specificity in integrins and its regulation by activation-driven tertiary and quaternary changes.
Collapse
Affiliation(s)
- M Amin Arnaout
- Renal Unit, Leukocyte Biology and Inflammation Program, Massachusetts General Hospital, and Harvard Medical School, Charlestown, MA 02129, USA.
| | | | | |
Collapse
|
35
|
Abstract
The divalent-cation-dependent binding of alphabeta heterodimeric integrins to their ligands regulates most cellular processes. Integrin-ligand interactions are tightly controlled by inside-out activation signals. Ligand-bound integrins in turn transduce outside-in signals typical of other receptors. Precise information of how ligands bind to integrins is restricted to that of a small vWF A-type domain present in some alpha-subunits (alphaA). Both inside-out and outside-in signals elicit tertiary and quaternary changes in integrins, but the precise nature and scope and of these changes are unknown. The recently solved structures of the extracellular segment of integrin alphaVbeta3 in its unliganded and liganded states are generating exciting new insights into the design, wiring, function and regulation of this protein family. The structures reveal a surprising degree of flexibility at defined regions in the structure that is potentially controlled by cations. The quaternary structure of the ligand-binding region bears a striking resemblance to the nucleotide-binding pocket of G-proteins, implying analogous activation and signaling mechanisms. Structural links exist through which ligand-induced tertiary changes may be translated into quaternary changes and vice versa. The structures also raise the tantalizing hypothesis that alphaA is a regulated endogenous integrin ligand, so that no special regulatory features are needed in this integrin. These findings provide the framework for new investigations of structure-activity relationships in integrins, with important implications for targeting these receptors therapeutically [corrected].
Collapse
Affiliation(s)
- M Amin Arnaout
- Renal Unit, Leukocyte Biology & Inflammation Program, Structural Biology Program, Massachusetts General Hospital and Harvard Medical School, Charlestown, MA 02129, USA.
| |
Collapse
|
36
|
Kamata T, Takada Y. Platelet integrin alphaIIbbeta3-ligand interactions: what can we learn from the structure? Int J Hematol 2001; 74:382-9. [PMID: 11794692 DOI: 10.1007/bf02982080] [Citation(s) in RCA: 12] [Impact Index Per Article: 0.5] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 12/24/2022]
Abstract
Upon vascular injury, platelets initiate interaction with exposed subendothelial matrices through various receptors such as glycoprotein (GP) Ib/IX/V complex, alpha2beta1 integrin, and GPVI/FcRgamma. Although these interactions cannot sustain stable platelet thrombus formation by themselves, they ultimately lead to the activation of alphaIIbbeta3 integrin (GPIIb-IIIa complex [GPIIb-IIIa]), the most abundant receptor in platelets. The alphaIIbbeta3 integrin plays a central role in primary hemostasis by serving as a receptor for fibrinogen and von Willebrand factor (vWf). It establishes a stable interaction with vWf bound to the extracellular matrices and uses fibrinogen as a bridging molecule in platelet aggregate formation. The alphaIIbbeta3 integrin also plays an important role in the pathogenesis of thrombosis. Over the past decades, a tremendous amount of effort has been made to elucidate the ligand-binding mechanisms of alphaIIbbeta3, in part because of its clinical significance. Most of the studies have relied on biochemical analyses of purified alphaIIbbeta3 or recombinant proteins generated in vitro. With the lack of actual 3-dimensional structure, molecular modeling has provided a useful framework for interpreting such experimental data on structure-function correlation of integrin molecules. However, it has also generated disagreement between different models. The aim of this minireview is to summarize the past efforts as well as the recent accomplishments in elucidating the structure/function of alphaIIbbeta3. Finally, we will try to explain all those experimental data using the recently published crystal structure of the extracellular domains of the alphaVbeta3 heterodimeric complex.
Collapse
Affiliation(s)
- T Kamata
- Department of Anatomy, Keio University School of Medicine, Tokyo, Japan.
| | | |
Collapse
|
37
|
Kamata T, Tieu KK, Irie A, Springer TA, Takada Y. Amino acid residues in the alpha IIb subunit that are critical for ligand binding to integrin alpha IIbbeta 3 are clustered in the beta-propeller model. J Biol Chem 2001; 276:44275-83. [PMID: 11557768 DOI: 10.1074/jbc.m107021200] [Citation(s) in RCA: 62] [Impact Index Per Article: 2.6] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/06/2022] Open
Abstract
Several distinct regions of the integrin alpha(IIb) subunit have been implicated in ligand binding. To localize the ligand binding sites in alpha(IIb), we swapped all 27 predicted loops with the corresponding sequences of alpha(4) or alpha(5). 19 of the 27 swapping mutations had no effect on binding to both fibrinogen and ligand-mimetic antibodies (e.g. LJ-CP3), suggesting that these regions do not contain major ligand binding sites. In contrast, swapping the remaining 8 predicted loops completely blocked ligand binding. Ala scanning mutagenesis of these critical predicted loops identified more than 30 discontinuous residues in repeats 2-4 and at the boundary between repeats 4 and 5 as critical for ligand binding. Interestingly, these residues are clustered in the predicted beta-propeller model, consistent with this model. Most of the critical residues are located at the edge of the upper face of the propeller, and several critical residues are located on the side of the propeller domain. None of the predicted loops in repeats 1, 6, and 7, and none of the four putative Ca(2+)-binding predicted loops on the lower surface of the beta-propeller were important for ligand binding. The results map an important ligand binding interface at the edge of the top and on the side of the beta-propeller toroid, centering on repeat 3.
Collapse
Affiliation(s)
- T Kamata
- Department of Cell Biology, the Scripps Research Institute, La Jolla, California 92037, USA.
| | | | | | | | | |
Collapse
|
38
|
Kosugi S, Tomiyama Y, Honda S, Kato H, Kiyoi T, Kashiwagi H, Kurata Y, Matsuzawa Y. Platelet-associated anti-GPIIb-IIIa autoantibodies in chronic immune thrombocytopenic purpura recognizing epitopes close to the ligand-binding site of glycoprotein (GP) IIb. Blood 2001; 98:1819-27. [PMID: 11535516 DOI: 10.1182/blood.v98.6.1819] [Citation(s) in RCA: 28] [Impact Index Per Article: 1.2] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/20/2022] Open
Abstract
Localization of epitopes for platelet-associated (PA) anti-GPIIb-IIIa (alpha(IIb)beta(3)) autoantibodies in chronic immune thrombocytopenic purpura remains elusive. Previous studies suggest that PA antibodies recognize the tertiary structure of intact glycoprotein (GP) IIb-IIIa. To localize their epitopes using antigen-capture enzyme-linked immunosorbent assay (ELISA), the reactivity of 34 PA anti-GPIIb-IIIa antibodies was examined with recombinant GPIIb-IIIa having a defect in ligand-binding sites in either GPIIb or GPIIIa, and no major conformational change was induced: KO variant GPIIb-IIIa was attributed to a 2-amino acid insertion between residues 160 and 161 in the W3 4-1 loop in GPIIb, and CAM variant GPIIb-IIIa was attributed to D119Y in GPIIIa. In one third (11 of 34) of the patients, PA antibodies showed a marked decrease (less than 50%) in reactivity with KO compared with wild-type GPIIb-IIIa. Their reactivity was also impaired against GPIIbD163A-IIIa. In sharp contrast, they reacted normally with CAM GPIIb-IIIa. OP-G2, a ligand-mimetic monoclonal antibody, markedly inhibited their binding to GPIIb-IIIa in patients with impaired binding to KO GPIIb-IIIa, but small GPIIb-IIIa antagonists did not. In addition, a newly developed sensitive ELISA indicated that autoantibodies showing impaired binding to KO are more potent inhibitors for fibrinogen binding. The present data suggest that certain PA anti-GPIIb-IIIa autoantibodies recognize epitopes close to the ligand-binding site in GPIIb, but not in GPIIIa.
Collapse
Affiliation(s)
- S Kosugi
- Department of Internal Medicine and Molecular Science, Graduate School of Medicine, Osaka University, Japan
| | | | | | | | | | | | | | | |
Collapse
|
39
|
Abstract
Binding of fibrinogen to GPIIb-IIIa on agonist-stimulated platelets results in platelet aggregation, presumably by crosslinking adjacent activated platelets. Although unactivated platelets express numerous copies of GPIIb-IIIa on their surface, spontaneous, and potentially deleterious, platelet aggregation is prevented by tightly regulating the fibrinogen binding activity of GPIIb-IIIa. Preliminary evidence suggests that it is the submembranous actin or actin-associated proteins that constrains GPIIb-IIIa in a low affinity state and that relief of this constraint by initiating actin filament turnover enables GPIIb-IIIa to bind fibrinogen. Two regions of the fibrinogen alpha chain that contain an RGD motif, as well as the carboxyl-terminus of the fibrinogen gamma chain, represent potential binding sites for GPIIb-IIIa in the fibrinogen molecule. However, ultrastructural studies using purified fibrinogen and GPIIb-IIIa, and studies using recombinant fibrinogen in which the RGD and relevant gamma chain motifs were mutated indicate that sequences located at the carboxyl-terminal end of the gamma chain mediates fibrinogen binding to GPIIb-IIIa. There is evidence that fibrinogen itself binds to regions in the amino terminal portions of both GPIIb and GPIIIa and that the sites interacting with the fibrinogen gamma chain and with RGD-containing peptides are spatially distinct. Nonetheless, there appears to be allosteric linkage between these sites, accounting for the ability of RGD-containing peptides to inhibit platelet aggregation and arterial thrombosis.
Collapse
Affiliation(s)
- J S Bennett
- Hematology-Oncology Division, Department of Medicine, University of Pennsylvania School of Medicine, Philadelphia, Pennsylvania, USA.
| |
Collapse
|
40
|
Tarui T, Mazar AP, Cines DB, Takada Y. Urokinase-type plasminogen activator receptor (CD87) is a ligand for integrins and mediates cell-cell interaction. J Biol Chem 2001; 276:3983-90. [PMID: 11053440 DOI: 10.1074/jbc.m008220200] [Citation(s) in RCA: 127] [Impact Index Per Article: 5.3] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/06/2022] Open
Abstract
Binding of urokinase-type plasminogen activator (uPA) to its receptor (uPAR/CD87) regulates cellular adhesion, migration, and tumor cell invasion. However, it is unclear how glycosyl phosphatidylinositol-anchored uPAR, which lacks a transmembrane structure, mediates signal transduction. It has been proposed that uPAR forms cis-interactions with integrins as an associated protein and thereby transduces proliferative or migratory signals to cells upon binding of uPA. We provide evidence that soluble uPAR (suPAR) specifically binds to integrins alpha4beta1, alpha6beta1, alpha9beta1, and alphavbeta3 on Chinese hamster ovary cells in a cation-dependent manner. Anti-integrin and anti-uPAR antibodies effectively block binding of suPAR to these integrins. Binding of suPAR to alpha4beta1 and alphavbeta3 is blocked by known soluble ligands and by the integrin mutations that inhibit ligand binding. These results suggest that uPAR is an integrin ligand rather than, or in addition to, an integrin-associated protein. In addition, we demonstrate that glycosyl phosphatidylinositol-anchored uPAR on the cell surface specifically binds to integrins on the apposing cells, suggesting that uPAR-integrin interaction may mediate cell-cell interaction (trans-interaction). These previously unrecognized uPAR-integrin interactions may allow uPAR to transduce signals through the engaged integrin without a hypothetical transmembrane adapter and may provide a potential therapeutic target for control of inflammation and cancer.
Collapse
Affiliation(s)
- T Tarui
- Department of Vascular Biology, The Scripps Research Institute, La Jolla, California 92037, USA
| | | | | | | |
Collapse
|
41
|
Honda S, Tomiyama Y, Pampori N, Kashiwagi H, Kiyoi T, Kosugi S, Tadokoro S, Kurata Y, Shattil SJ, Matsuzawa Y. Ligand binding to integrin alpha(v)beta(3) requires tyrosine 178 in the alpha(v) subunit. Blood 2001; 97:175-82. [PMID: 11133758 DOI: 10.1182/blood.v97.1.175] [Citation(s) in RCA: 15] [Impact Index Per Article: 0.6] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/20/2022] Open
Abstract
Integrin alpha(v)beta(3) has been implicated in angiogenesis and other biological processes. However, the ligand-binding sites in alpha(v), a non-I-domain alpha subunit, remain to be identified. Recently in alpha(IIb), the other partner of the beta(3) subunit, several discontinuous residues important for ligand binding were identified in the predicted loops between repeats 2 and 3 (W3 4-1 loop) and within repeat 3 (W3 2-3 loop). Based on these findings, alanine-scanning mutagenesis in 293 cells was used to investigate the role of these loops (cysteine [C]142-C155 and glycine [G]172-G181) of alpha(v) in ligand binding. Wild-type alpha(v)beta(3) was able to bind soluble fibrinogen following integrin activation either by 0.5 mM manganese dichloride (MnCl(2)) or a mutation of beta(3) threonine (T)562 to asparagine. However, mutation of tyrosine (Y)178 to alanine in the predicted G172-G181 loop of alpha(v) abolished fibrinogen binding, and alanine (A) substitutions at adjacent residues phenylalanine (F)177 and tryptophan (W)179 had a similar effect. Cells expressing Y178Aalpha(v) also failed to bind to immobilized fibrinogen. Moreover, the Y178A mutation abolished the binding of WOW-1 Fab, a monovalent ligand-mimetic anti-alpha(v)beta(3) antibody, and the expression of beta(3) ligand-induced binding sites (LIBS) induced by arginine-glycine-aspartic acid-tryptophan (RGDW). In sharp contrast to the data obtained with alpha(IIb), none of the mutations in the predicted W3 4-1 loop in alpha(v) impaired ligand binding. These results implicate alpha(v) Y178 in ligand binding to alpha(v)beta(3), and they suggest that there are key structural differences in the adhesive ligand-binding sites of alpha(v)beta(3) and alpha(IIb)beta(3).
Collapse
Affiliation(s)
- S Honda
- Department of Internal Medicine and Molecular Science, Graduate School of Medicine, Osaka University, Osaka, Japan
| | | | | | | | | | | | | | | | | | | |
Collapse
|
42
|
Affiliation(s)
- E F Plow
- Joseph J. Jacobs Center for Thrombosis and Vascular Biology, Cleveland Clinic Foundation, Cleveland, Ohio 44195, USA.
| | | | | | | | | |
Collapse
|
43
|
Zang Q, Lu C, Huang C, Takagi J, Springer TA. The top of the inserted-like domain of the integrin lymphocyte function-associated antigen-1 beta subunit contacts the alpha subunit beta -propeller domain near beta-sheet 3. J Biol Chem 2000; 275:22202-12. [PMID: 10781608 DOI: 10.1074/jbc.m002883200] [Citation(s) in RCA: 31] [Impact Index Per Article: 1.2] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/06/2022] Open
Abstract
We find that monoclonal antibody YTA-1 recognizes an epitope formed by a combination of the integrin alpha(L) and beta(2) subunits of LFA-1. Using human/mouse chimeras of the alpha(L) and beta(2) subunits, we determined that YTA-1 binds to the predicted inserted (I)-like domain of the beta(2) subunit and the predicted beta-propeller domain of the alpha(L) subunit. Substitution into mouse LFA-1 of human residues Ser(302) and Arg(303) of the beta(2) subunit and Pro(78), Thr(79), Asp(80), Ile(365), and Asn(367) of the alpha(L) subunit is sufficient to completely reconstitute YTA-1 reactivity. Antibodies that bind to epitopes that are nearby in models of the I-like and beta-propeller domains compete with YTA-1 monoclonal antibody for binding. The predicted beta-propeller domain of integrin alpha subunits contains seven beta-sheets arranged like blades of a propeller around a pseudosymmetry axis. The antigenic residues cluster on the bottom of this domain in the 1-2 loop of blade 2, and on the side of the domain in beta-strand 4 of blade 3. The I domain is inserted between these blades on the top of the beta-propeller domain. The antigenic residues in the beta subunit localize to the top of the I-like domain near the putative Mg(2+) ion binding site. Thus, the I-like domain contacts the bottom or side of the beta-propeller domain near beta-sheets 2 and 3. YTA-1 preferentially reacts with activated LFA-1 and is a function-blocking antibody, suggesting that conformational movements occur near the interface it defines between the LFA-1 alpha and beta subunits.
Collapse
Affiliation(s)
- Q Zang
- Center For Blood Research, Department of Pathology, Harvard Medical School, Boston, Massachusetts 02115, USA
| | | | | | | | | |
Collapse
|
44
|
Mould AP, Askari JA, Humphries MJ. Molecular basis of ligand recognition by integrin alpha 5beta 1. I. Specificity of ligand binding is determined by amino acid sequences in the second and third NH2-terminal repeats of the alpha subunit. J Biol Chem 2000; 275:20324-36. [PMID: 10764748 DOI: 10.1074/jbc.m000572200] [Citation(s) in RCA: 79] [Impact Index Per Article: 3.2] [Reference Citation Analysis] [Abstract] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/06/2022] Open
Abstract
The NH(2)-terminal portion (putative ligand-binding domain) of alpha subunits contains 7 homologous repeats, the last 3 or 4 of which possess divalent cation binding sequences. These repeats are predicted to form a seven-bladed beta-propeller structure. To map ligand recognition sites on the alpha(5) subunit we have taken the approach of constructing and expressing alpha(V)/alpha(5) chimeras. Although the NH(2)-terminal repeats of alpha(5) and alpha(V) are >50% identical at the amino acid level, alpha(5)beta(1) and alpha(V)beta(1) show marked differences in their ligand binding specificities. Thus: (i) although both integrins recognize the Arg-Gly-Asp (RGD) sequence in fibronectin, the interaction of alpha(5)beta(1) but not of alpha(V)beta(1) with fibronectin is strongly dependent on the "synergy" sequence Pro-His-Ser-Arg-Asn; (ii) alpha(5)beta(1) binds preferentially to RGD peptides in which RGD is followed by Gly-Trp (GW) whereas alpha(V)beta(1) has a broader specificity; (iii) only alpha(5)beta(1) recognizes peptides containing the sequence Arg-Arg-Glu-Thr-Ala-Trp-Ala (RRETAWA). Therefore, amino acid residues involved in ligand recognition by alpha(5)beta(1) can potentially be identified in gain-of-function experiments by their ability to switch the ligand binding properties of alpha(V)beta(1) to those of alpha(5)beta(1). By introducing appropriate restriction enzyme sites, or using site-directed mutagenesis, parts of the NH(2)-terminal repeats of alpha(V) were replaced with the corresponding regions of the alpha(5) subunit. Chimeric subunits were expressed on the surface of Chinese hamster ovary-B2 cells (which lack endogenous alpha(5)) as heterodimers with hamster beta(1). Stable cell lines were generated and tested for their ability to attach to alpha(5)beta(1)-selective ligands. Our results demonstrate that: (a) the first three NH(2)-terminal repeats contain the amino acid sequences that determine ligand binding specificity and the same repeats include the epitopes of function blocking anti-alpha subunit mAbs; (b) the divalent cation-binding sites (in repeats 4-7) do not confer alpha(5)beta(1)- or alpha(V)beta(1)-specific ligand recognition; (c) amino acid residues Ala(107)-Tyr(226) of alpha(5) (corresponding approximately to repeats 2 and 3) are sufficient to change all the ligand binding properties of alpha(V)beta(1) to those of alpha(5)beta(1); (d) swapping a small part of a predicted loop region of alpha(V) with the corresponding region of alpha(5) (Asp(154)-Ala(159)) is sufficient to confer selectivity for RGDGW and the ability to recognize RRETAWA.
Collapse
Affiliation(s)
- A P Mould
- Wellcome Trust Centre for Cell-Matrix Research, School of Biological Sciences, University of Manchester, Manchester M13 9PT, United Kingdom.
| | | | | |
Collapse
|
45
|
A Leu262Pro mutation in the integrin β3 subunit results in an αIIb-β3 complex that binds fibrin but not fibrinogen. Blood 2000. [DOI: 10.1182/blood.v96.1.161] [Citation(s) in RCA: 26] [Impact Index Per Article: 1.0] [Reference Citation Analysis] [Abstract] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/20/2022] Open
Abstract
AbstractPlatelet retraction of a fibrin clot is mediated by the platelet fibrinogen receptor, IIbβ3. In certain forms of the inherited platelet disorder, Glanzmann thrombasthenia (GT), mutant IIbβ3 may interact normally with fibrin yet fail to support fibrinogen-dependent aggregation. We describe a patient (LD) with such a form of GT. Platelets from LD supported normal clot retraction but failed to bind fibrinogen. Platelet analysis using flow cytometry and immunoblotting showed reduced but clearly detectable IIbβ3, findings consistent with type II GT. Genotyping of LD revealed 2 novel β3 mutations: a deletion of nucleotides 867 to 868, resulting in a premature stop codon at amino acid residue 267, and a T883C missense mutation, resulting in a leucine (Leu) 262-to-proline (Pro) substitution. Leu262 is highly conserved among β integrin subunits and lies within an intrachain loop implicated in subunit association. Leu262Proβ3 cotransfected with wild-type IIb into COS-7 cells showed delayed intracellular maturation and reduced surface expression of easily dissociable complexes. In human embryonic kidney 293 cells, Leu262Proβ3 formed a complex with endogenous av and retracted fibrin clots similarly to wild-type β3. The same cells, however, were unable to bind immobilized fibrinogen. The molecular requirements for IIbβ3 to interact with fibrin compared with fibrinogen, therefore, appear to differ. The region surrounding β3 Leu262 may maintain β3 in a fibrinogen-binding, competent form, but it appears not to be required for receptor interactions with fibrin.
Collapse
|
46
|
A Leu262Pro mutation in the integrin β3 subunit results in an αIIb-β3 complex that binds fibrin but not fibrinogen. Blood 2000. [DOI: 10.1182/blood.v96.1.161.013k50_161_169] [Citation(s) in RCA: 1] [Impact Index Per Article: 0.0] [Reference Citation Analysis] [Abstract] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/20/2022] Open
Abstract
Platelet retraction of a fibrin clot is mediated by the platelet fibrinogen receptor, IIbβ3. In certain forms of the inherited platelet disorder, Glanzmann thrombasthenia (GT), mutant IIbβ3 may interact normally with fibrin yet fail to support fibrinogen-dependent aggregation. We describe a patient (LD) with such a form of GT. Platelets from LD supported normal clot retraction but failed to bind fibrinogen. Platelet analysis using flow cytometry and immunoblotting showed reduced but clearly detectable IIbβ3, findings consistent with type II GT. Genotyping of LD revealed 2 novel β3 mutations: a deletion of nucleotides 867 to 868, resulting in a premature stop codon at amino acid residue 267, and a T883C missense mutation, resulting in a leucine (Leu) 262-to-proline (Pro) substitution. Leu262 is highly conserved among β integrin subunits and lies within an intrachain loop implicated in subunit association. Leu262Proβ3 cotransfected with wild-type IIb into COS-7 cells showed delayed intracellular maturation and reduced surface expression of easily dissociable complexes. In human embryonic kidney 293 cells, Leu262Proβ3 formed a complex with endogenous av and retracted fibrin clots similarly to wild-type β3. The same cells, however, were unable to bind immobilized fibrinogen. The molecular requirements for IIbβ3 to interact with fibrin compared with fibrinogen, therefore, appear to differ. The region surrounding β3 Leu262 may maintain β3 in a fibrinogen-binding, competent form, but it appears not to be required for receptor interactions with fibrin.
Collapse
|
47
|
Yokoyama K, Erickson HP, Ikeda Y, Takada Y. Identification of amino acid sequences in fibrinogen gamma -chain and tenascin C C-terminal domains critical for binding to integrin alpha vbeta 3. J Biol Chem 2000; 275:16891-8. [PMID: 10747940 DOI: 10.1074/jbc.m000610200] [Citation(s) in RCA: 72] [Impact Index Per Article: 2.9] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/06/2022] Open
Abstract
Integrin alpha(v)beta(3) recognizes fibrinogen gamma and alpha(E) chain C-terminal domains (gammaC and alpha(E)C) but does not require the gammaC dodecapeptide sequence HHLGGAKQAGDV(400-411) for binding to gammaC. We have localized the alpha(v)beta(3) binding sites in gammaC using gammaC-derived synthetic peptides. We found that two peptides GWTVFQKRLDGSV(190-202) and GVYYQGGTYSKAS(346-358) block the alpha(v)beta(3) binding to gammaC or alpha(E)C, block the alpha(v)beta(3)-mediated clot retraction, and induce the ligand-induced binding site 2 (LIBS2) epitope in alpha(v)beta(3). Neither peptide affects fibrinogen binding to alpha(IIb)beta(3). Scrambled or inverted peptides were not effective. These results suggest that the two gammaC-derived peptides directly interact with alpha(v)beta(3) and specifically block alpha(v)beta(3)-gammaC or alpha(E)C interaction. The two sequences are located next to each other in the gammaC crystal structure, although they are separate in the primary structure. Asp-199, Ser-201, Gln-350, Thr-353, Lys-356, Ala-357, and Ser-358 residues are exposed to the surface. This suggests that the two sequences are part of alpha(v)beta(3) binding sites in fibrinogen gammaC domain. We also found that tenascin C C-terminal fibrinogen-like domain specifically binds to alpha(v)beta(3). Notably, a peptide WYRNCHRVNLMGRYGDNNHSQGVNWFHWKG from this domain that includes the sequence corresponding to gammaC GVYYQGGTYSKAS(346-358) specifically binds to alpha(v)beta(3), suggesting that fibrinogen and tenascin C C-terminal domains interact with alpha(v)beta(3) in a similar manner.
Collapse
Affiliation(s)
- K Yokoyama
- Department of Vascular Biology, The Scripps Research Institute, La Jolla, California 92037, USA
| | | | | | | |
Collapse
|
48
|
Puzon-McLaughlin W, Kamata T, Takada Y. Multiple discontinuous ligand-mimetic antibody binding sites define a ligand binding pocket in integrin alpha(IIb)beta(3). J Biol Chem 2000; 275:7795-802. [PMID: 10713093 DOI: 10.1074/jbc.275.11.7795] [Citation(s) in RCA: 80] [Impact Index Per Article: 3.2] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/06/2022] Open
Abstract
Integrin alpha(IIb)beta(3), a platelet fibrinogen receptor, is critically involved in thrombosis and hemostasis. However, how ligands interact with alpha(IIb)beta(3) has been controversial. Ligand-mimetic anti-alpha(IIb)beta(3) antibodies (PAC-1, LJ-CP3, and OP-G2) contain the RGD-like RYD sequence in their CDR3 in the heavy chain and have structural and functional similarities to native ligands. We have located binding sites for ligand-mimetic antibodies in alpha(IIb) and beta(3) using human-to-mouse chimeras, which we expect to maintain functional integrity of alpha(IIb)beta(3). Here we report that these antibodies recognize several discontinuous binding sites in both the alpha(IIb) and beta(3) subunits; these binding sites are located in residues 156-162 and 229-230 of alpha(IIb) and residues 179-183 of beta(3). In contrast, several nonligand-mimetic antibodies (e.g. 7E3) recognize single epitopes in either subunit. Thus, binding to several discontinuous sites in both subunits is unique to ligand-mimetic antibodies. Interestingly, these binding sites overlap with several (but not all) of the sequences that have been reported to be critical for fibrinogen binding (e.g. N-terminal repeats 2-3 but not repeats 4-7, of alpha(IIb)). These results suggest that ligand-mimetic antibodies and probably native ligands may make direct contact with these discontinuous binding sites in both subunits, which may constitute a ligand-binding pocket.
Collapse
Affiliation(s)
- W Puzon-McLaughlin
- Department of Vascular Biology, The Scripps Research Institute, La Jolla, California 92037, USA
| | | | | |
Collapse
|
49
|
Gidwitz S, Lyman S, White GC. Expression and function of calcium binding domain chimeras of the integrins alpha(IIb) and alpha(5). J Biol Chem 2000; 275:6680-8. [PMID: 10692478 DOI: 10.1074/jbc.275.9.6680] [Citation(s) in RCA: 5] [Impact Index Per Article: 0.2] [Reference Citation Analysis] [Abstract] [MESH Headings] [Grants] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/06/2022] Open
Abstract
To further identify amino acid domains involved in the ligand binding specificity of alpha(IIb)beta(3), chimeras of the conserved calcium binding domains of alpha(IIb) and the alpha subunit of the fibronectin receptor alpha(5)beta(1) were constructed. Chimeras that replaced all four calcium binding domains, replaced all but the second calcium binding domain of alpha(IIb) with those of alpha(5), or deleted all four calcium binding domains were synthesized but not expressed on the cell surface. Additional chimeras exchanged subsets or all of the variant amino acids in the second calcium binding domain, a region implicated in ligand binding. Cell surface expression of each second calcium binding domain mutant complexed with beta(3) was observed. Each second calcium binding domain mutant was able to 1) bind to immobilized fibrinogen, 2) form fibrinogen-dependent aggregates after treatment with dithiothreitol, and 3) bind the activation-dependent antibody PAC1 after LIBS 6 treatment. Soluble fibrinogen binding studies suggested that there were only small changes in either the K(d) or B(max) of any mutant. We conclude that chimeras of alpha(IIb) containing the second calcium binding domain sequences of alpha(5) are capable of complexing with beta(3), that the complexes are expressed on the cell surface, and that mutant complexes are capable of binding both immobilized and soluble fibrinogen, suggesting that the second calcium binding domain does not determine ligand binding specificity.
Collapse
Affiliation(s)
- S Gidwitz
- Center for Thrombosis and Hemostasis, University of North Carolina, Chapel Hill, North Carolina 27599, USA.
| | | | | |
Collapse
|
50
|
French DL, Seligsohn U. Platelet glycoprotein IIb/IIIa receptors and Glanzmann's thrombasthenia. Arterioscler Thromb Vasc Biol 2000; 20:607-10. [PMID: 10712380 DOI: 10.1161/01.atv.20.3.607] [Citation(s) in RCA: 45] [Impact Index Per Article: 1.8] [Reference Citation Analysis] [MESH Headings] [Track Full Text] [Journal Information] [Subscribe] [Scholar Register] [Indexed: 11/16/2022]
Affiliation(s)
- D L French
- Mount Sinai School of Medicine, New York, NY 10029, USA.
| | | |
Collapse
|