ISSN: 1885-5857 Impact factor 2023 7.2
Vol. 62. Num. 5.
Pages 552-562 (May 2009)

Molecular Basis of Leukocyte-Endothelium Interactions During the Inflammatory Response

Bases moleculares de las interacciones leucocito-endotelio durante la respuesta inflamatoria

Olga BarreiroaFrancisco Sánchez-Madrida

Options

The process of leukocyte extravasation, a critical step in the inflammatory response, involves the migration of leukocytes from the bloodstream towards target tissues, where they exert their effector function. Leukocyte extravasation is orchestrated by the combined action of cellular adhesion receptors and chemotactic factors, and involves radical morphological changes in both leukocytes and endothelial cells. Thus, it constitutes an active process for both cell types and promotes the rapid and efficient influx of leukocytes to inflammatory foci without compromising the integrity of the endothelial barrier. This article provides a review of leukocyte extravasation from both molecular and mechanical points of view, with a particular emphasis on the most recent findings on the topic. It includes a description of newly revealed steps in the adhesion cascade, such as slow rolling motion, intraluminal crawling and alternative pathways for transcellular migration, and discusses the functional role of novel adhesion receptors, the spatiotemporal organization of receptors at the plasma membrane, and the signaling pathways that control different phases of the extravasation process. for transcellular migration, and discusses the functional role of novel adhesion receptors, the spatiotemporal organization of receptors at the plasma membrane and the signaling pathways that control different phases of the extravasation process.

Keywords

Inflammation
Extravasation
Leukocyte- endothelial interaction
Adhesion
Transmigration

Adhesion receptors play an essential physiological role in maintaining tissue integrity by regulating numerous processes such as cell activation, migration, growth, differentiation, and death.1,2 This regulation is achieved through direct signal transduction and modulation of intracellular signaling triggered by different growth factors.3 Cell-cell interactions are essential for regulating hematopoiesis4,5 and the inflammatory response.6,7 Adhesion molecules are therefore particularly implicated in a wide variety of cardiovascular disorders that involve inflammation, such as atherogenic processes and progression of atherosclerotic plaque, myocardial infarction, ischemia-reperfusion injury or transplant rejection, and, to a lesser extent, valve stenosis, and myocardiopathy.

The coordinated functioning of adhesion receptors, the cytoskeleton, and signaling molecules is crucial for leukocyte extravasation, a key process in immune response. Thus, correct integration of signals from "outside in" and "inside out" in leukocytes and the endothelium during each extravasation step—the so-called multi-step paradigm—is essential for this phenomenon to occur6,8 (Figure 1). Leukocyte extravasation takes place not only during inflammatory response, but also during recirculation of the lymphocytes to the secondary lymphoid organs, although this latter process will not be considered in this review.

Figure. 1. The adhesion cascade. The scanning electron microscope image shows a human endothelial monolayer treated with proinflammatory stimuli and perfused with human peripheral blood lymphocytes and monocytes at physiological flow (1.8 dyn/cm2). Several unpolarized leukocytes have come into contact with the endothelium and have been captured during the rolling process. Also shown is a lymphocyte that has managed to firmly adhere to the endothelium and drastically changed its morphology from rounded to polarized.

INITIAL INTERACTIONS BETWEEN CIRCULATING LEUKOCYTES AND THE ENDOTHELIUM: TETHERING AND ROLLING MEDIATED BY SELECTINS AND THEIR LIGANDS

To initiate the inflammatory response, circulating leukocytes in the bloodstream have to establish contact (tethering) with the vascular wall and adhere to it, while withstanding the shear forces. Tethering and rolling of the leukocytes over the activated endothelium are the first steps in the sequential process of extravasation. They are followed by firm adhesion and transendothelial migration. The initial contact or tethering is largely mediated by selectins and their ligands, and blood flow must be present for it to be efficient.9 Although selectins and their ligands tend to interact with a variable affinity, the high frequency of association-dissociation of interactions allows them to mediate labile and transient tethers between leukocytes and the endothelium.10,11 Tethering slows the speed of travel of the leukocytes and allows them to roll over the endothelial surface, favoring subsequent interactions mediated by integrins and their ligands and increasing leukocyte adherence. As a result, the leukocytes finally come to a halt on the vascular wall.12

The selectins (P, E, and L) are type 1 transmembrane glycoproteins that bind to fucosylated and sialylated hydrocarbons present in their ligands in a Ca2+-dependent fashion. L-selectin is expressed by most leukocytes, whereas the E and P forms are expressed on endothelial cells activated by proinflammatory stimuli. P-selectin is also expressed by activated platelets (reviewed by Barreiro et al13). In addition to the interaction of leukocyte selectin (L-selectin) with endothelial selectin (P- and E-selectin), the P-Selectin glycoprotein ligand-1 (PSGL1) protein is a major ligand of these 3 selectins. In fact, the binding of PSGL1 to P- and E-selectin promotes the interaction of leukocytes with the endothelium, whereas the binding of PSGL1 to L-selectin enables leukocyte-leukocyte interactions, whereby the adhered leukocytes facilitate the capture of other circulating leukocytes at sites where the endothelium is inflamed, regardless of whether these cells express ligands for endothelial selectins in a process denoted secondary recruitment.14 In addition to PSGL1, selectins can also bind to other glycoproteins such as CD44 or E-selectinligand-1 (ESL1) in the case of E-selectin. Each particular ligand seems to have a distinct function during the process of neutrophil capture. Thus, PSGL1 is primarily implicated in initial leukocyte tethering, whereas ESL1 is necessary to convert the transient initial tethers into a slower and more stable rolling. Finally, CD44 controls the rolling velocity and intervenes in the polarization of PSGL1 and L-selectin, probably to allow secondary recruitment.15 Platelets can also act as secondary recruiters of leukocytes thanks to their capacity to interact with both the circulating leukocytes and the endothelium at the same time. In addition, they can release chemokines that are immobilized on the luminal endothelial surface, thereby favoring the adhesion process.16

Besides the selectins and their ligands, the α4β1-andα4β7-integrins—through their interaction with vascular cell adhesion molecule (VCAM) 1 and mucosal addressin cell adhesion molecule (MAdCAM) 1, respectively—can independently mediate this initial tethering.17-19 On the other hand, the interaction between lymphocyte function-associated antigen (LFA) 1 and cell-cell adhesion molecule (ICAM) 1 collaborates with the function of L-selectin, thereby stabilizing the transient contact phase and reducing the rolling velocity.20,21

The adhesion receptors must be correctly localized on the cells for them to function correctly during leukocyte trafficking.22 Selectins and their ligands, and the α4 integrins are clustered at the ends of the microvilli of the leukocytes. The tethering of the selectins to the cytoskeleton of actin with proteins such as α-actinin or ezrin/radixin/ moesin (ERM) is necessary for them to function properly.23-26

It has been shown that selectins activate multiple signaling pathways that are linked to processes such as actin cytoskeletal reorganization, like the MAPK, p56lck, Ras, or Rac2 cascade (reviewed by Barreiro et al13). On the other hand, PSGL1 also activates different intracellular signaling pathways with an inductive effect on the activation of leukocytes, thereby increasing the expression of different molecules that are implicated in the following steps of the extravasation process and in effector function. They also play an unexpected role in the induction of tolerogenic function in dendritic cells.27-30

CENTRAL ROLE OF LEUKOCYTE INTEGRINS AND THEIR ENDOTHELIAL LIGANDS IN PROCESSES OF ACTIVATION, ARREST, FIRM ADHESION, AND CRAWLING

Leukocyte trafficking through the different tissues and organs and the subsequent leukocyte interaction with other immune cells are essential for developing innate and acquired immunity.31 Integrins are fundamental molecules in cell migration. They control the cell-cell and cell-extracellular matrix interactions during recirculation and inflammation. One of their most important characteristics lies in their ability to alter their adherent activity, regardless of how extensively expressed they are on the membrane.32 Thus, circulating leukocytes in blood maintain their integrins in an inactive conformation to avoid nonspecific contact with uninflamed vascular walls, but when they arrive at the inflammatory focus, a rapid in situ activation of the integrins occurs.33 As in the case of the selectins, the spatial distribution of the integrins and their ligands on specialized membrane structures is essential for proper function. This spatial organization requires a precise regulation of the cytoskeleton to allow recruitment of signaling intermediates and second messengers that trigger cell activation.34,35

The integrins comprise a family of 24 heterodimeric receptors, each of which is composed of an α subunit and another β subunit. These molecules dynamically alter their adhesive properties through conformational changes (affinity) as well as through spatial redistribution on the cell surface (avidity).36 Recent observations predict the existence of 3 conformational states (bent conformation with low affinity, extended conformation with intermediate affinity, and extended conformation with high affinity).37,38 The most relevant integrins for leukocyte adhesion to the endothelium are members of the β2 subfamily, particularly LFA-1 (CD11a/CD18 or αLβ2) and myeloid-specific integrin Mac-1 (CD11b/CD18 or αMβ2), as well as integrins α4 VLA-4 (a4β1) and α4β7. Most of the ligands are transmembrane proteins that belong to the immunoglobulin superfamily. LFA-1 can bind to 5 intercellular adhesion molecules (ICAM-1 to ICAM-5), although the most important of these are ICAM-1 and ICAM-3.39 ICAM-1 is expressed on leukocytes, dendritic cells, and epithelial cells. In addition, expression is low on quiescent endothelial cells and increases with proinflammatory stimuli.40 ICAM-3 is expressed constitutively on all leukocytes.41 In addition to LFA-1, another ligand is the junctional adhesion molecule JAM-A, which is selectively concentrated on the apical region of the tight junctions of the endothelium and is partially redistributed to the apical surface of the endothelium with certain proinflammatory stimuli.42 On the other hand, Mac-1 interacts with ICAM-1, JAM-C, and the receptor for advanced glycation endproducts (RAGE).43,44 The integrin VLA-4 interacts with VCAM-1,45 which is an adhesion molecule that is expressed de novo after endothelial activation46 and also binds to JAM-B.47 VLA-4 also interacts with ADAM-28, fibronectin, osteopontin, thrombospondin, the von Willebrand coagulation factor, and the invasin bacterial protein.48

Finally, αβ7 integrin, apart from interacting with VCAM-1 and fibronectin, specifically recognizes MAdCAM-1, a receptor expressed in the lymphoid tissues of the mucosa.19

Modulation of Chemokine-Mediated Integrin Activity

As tethering to the vascular endothelium occurs, the rolling velocity of the leukocytes slows and they are activated on encountering immobilized chemokines and integrin ligands exposed on the apical endothelial surface. This activation step enables the arrest of leukocytes and their subsequent firm adhesion to the endothelium under physiological flow conditions.49,50 Leukocyte activation implies a marked morphological change: the rounded circulating cell is transformed into a promigratory cell with polarized morphology, in which at least 2 regions can be identified, the cell front and the cellular uropod.51 The polarization of the leukocytes allows the cell to coordinate the intracellular forces to produce the necessary cell crawling during the extravasation process.52

The chemokines bound to the glycosaminoglycans of the apical endothelial membrane act by signaling via the G protein coupled receptors (GPCR) located on the microvilli of the leukocyte, thereby inducing a wide range of "outside in" signals in a fraction of a second. These lead to multiple conformational changes in the integrins.53-55 The complexity and the short time period of the signaling mechanisms induced by the chemokines that control the activation of the integrins are consistent with the existence of compartmentalized and pre-formed protein networks ("signalosomes") in the leukocytes.56 The presence of specific chemokines in different vascular beds helps orchestrate the selective recruitment of different leukocyte subpopulations to the inflammatory foci or to the secondary lymphoid organs.57 In addition, chemokines can exert a differential effect on specific integrins within the same microenvironment.58

Modulation of Ligand-Mediated Integrin Affinity

After chemokine-induced activation, the conformation of the integrins changes reversibly from the inactive (bent) form to the extended form with intermediate affinity. This event prepares the integrin for binding to its endothelial ligand. The integrins that contain an I domain inserted into their α subunits undergo a subsequent conformational change after binding to the ligand, culminating in the complete activation of the integrin and leukocyte arrest.59-61 Therefore, the high-affinity conformational state for immediate arrest of the leukocyte on the endothelium requires immobilized chemokines and the integrin ligands.55,62 However, the α4 integrins, which contain an I-like domain on the β chains, can spontaneously interact with their endothelial ligands without a prior chemotactic trigger.17

The signaling induced by the binding to the ligand leads to the separation of the cytoplasmic regions of the subunits of the integrin, thereby favoring its association with the cortical actin cytoskeleton. The α4 integrins are basically linked through paxillin while β2 integrins are linked through talin, filamin, and other structural molecules. In addition, the binding to the ligand increases the recruitment of additional integrins to increase the firm adhesion of the leukocyte in conditions of flow stress.63 This clustering of integrins depends on the release from their tether to the cytoskeleton—a process mediated by protein kinase C (PKC) and calpain—to increase their lateral mobility on the membrane.64 In addition, the role of Rap-1 and its activator CalDAG-GEFI have recently been described, as well as the coordinated action of kindlin-3 with talin in the activation of integrins to mediate the leukocyte firm adhesion in different types of hematopoietic cells.65,66 With regard to the spatial organization of the integrins, it has been shown that nanoclusters of LFA-1 not bound to its ligand are present and these enable the efficient formation of microclusters induced by binding to a ligand.67,68

On the other hand, several studies indicate that flow stress also regulates the integrins, reinforcing their bonds and even increasing their affinity.69,70 The integration of signaling derived from chemokines and the external forces to favor transmigration has been defined as the phenomenon of chemorheotaxis.71

Regulation of Leukocyte Crawling by Integrins

The signals implicated in the integrin-mediated firm adhesion of leukocytes to the endothelium have to be attenuated and the original contact has to be weakened enough to allow the migration of the leukocyte towards the appropriate site to start the process of endothelial transmigration. The β2 integrins seem to play an important role in crawling, as blockade of the integrin or its ligands leads to random migration, failure to position the cell at the interendothelial junctions, and defective diapedesis.72 In vivo studies using genetically modified mice lacking LFA-1 or Mac-1 clearly delineate the different underlying mechanisms for each of these β2 integrins. Whereas firm adhesion is mediated by LFA-1, crawling depends on Mac-1; both processes contribute to an efficient migration.73 After activation by binding to a ligand, the integrins regulate different effectors of myosin contractility, actin-remodeling GTPases, and molecules implicated in the regulation of the microtubule network both at the cell front and in the uropod. Thus, the integration of signals generated at both cell poles leads to coordinated movement of the leukocyte.34

Functional Role of the Endothelial Adhesion Molecules VCAM-1 and ICAM-1 in Leukocyte Capture

VCAM-1 and ICAM-1 molecules, both members of the immunoglobulin superfamily, are the main endothelial adhesion molecules implicated in the binding to integrins VLA-4 and LFA-1, respectively.45,74 ICAM-1 is scarcely expressed on the quiescent endothelium, whereas the expression of both molecules is induced after cell activation by proinflammatory cytokines such as interleukin (IL) 1 and tumor necrosis factor (TNF) α.40,46 In addition, binding of VCAM-1 and ICAM-1 to the actin cytoskeleton has been reported through 2 members of the ERM family, namely, ezrin and meosin.75,76 These molecules act as links between the membrane and the actin cytoskeleton, regulating cortical morphogenesis and cell adhesion.

The dynamics of VCAM-1 and ICAM-1 has been studied in human umbilical vein endothelial cells (HUVECs) activated with TNF during the process of leukocyte-endothelium interaction. It has been observed that, after leukocyte arrest on the endothelium, the binding of VCAM-1 and ICAM-1 with their ligands triggers the reorganization of the endothelial cortical actin cytoskeleton and generates a 3-dimensional docking structure that surrounds the leukocyte and prevents leukocytes adhered in conditions of physiological flow from becoming unbound. This structure has a large accumulation of adhesion receptors, as well as the activated ezrin and moesin proteins. The endothelial docking structure is supported by the actin cytoskeleton, docking actin-bundling proteins such as α actinin, proteins typical of focal adhesions such as talin, paxillin, and vinculin, and actin nucleating proteins. In addition, second messengers such as PI(4,5)P2 and the Rho/160ROCK signaling pathway are also important for generating and maintaining the endothelial docking structure (Figure 2).75 Furthermore, both ICAM-1 and VCAM-1 cluster together in the endothelial docking structure, although one of them is not bound to its corresponding ligand. This joint clustering also proceeds independently of the anchoring to the actin cytoskeleton and of the formation of ICAM-1/VCAM-1 heterodimers, as this is due to the inclusion of VCAM-1 and ICAM-1 in microdomains rich in tetraspanins, which act as endothelial platforms for specialized adhesion77 (Figure 2). The tetraspanins are small proteins that cross the membrane 4 times with lateral interaction of their second extracellular domain with other integral proteins of the membrane, regulate membrane function, and form multiple protein domains on the plasma membrane. They have been implicated in several cell functions, including migration, homotypic and heterotypic cell-cell adhesion, and antigen presentation, viral infection, and gamete fusion.78-82

Figure 2. Active role of the endothelium during extravasation. The scanning electron microscope image shows the organization of endothelial adhesion receptors in nanoclusters on the apical membrane (endothelial adhesive platforms; staining corresponds to ICAM-1 using antibodies coupled to colloidal gold). When a leukocyte establishes contact with the endothelium, the endothelial adhesion receptors are concentrated in the endothelial docking structure, which keeps the leukocyte firmly adhered and prevents it from becoming separated due to the force of the flow that it has to withstand.

The use of innovative microscopy analytical techniques has enabled the characterization of the diffusive properties, organization at the nano scale, and specific molecular interactions within the microdomains in live primary human endothelial cells. Such studies have provided convincing evidence of the existence of endothelial adhesion platforms as physical entities distinct from the lipid rafts in the plasma membrane.77 Scanning electron microscopy in samples treated with a specific peptide blocker of tetraspanins reveals the nanoclustering or avidity of VCAM-1 and ICAM-1 induced by the endothelial adhesion platforms as a new supramolecular organization mechanism that regulates the efficient adhesive capacity of both endothelial adhesion receptors to their counter-receptors, the leukocyte integrins.77 The functional relevance of the inclusion of ICAM-1 and VCAM-1 in tetraspanin microdomains in endothelial cells has been demonstrated through the use of an experimental strategy with interfering RNA targeting the tetraspanins CD9 and CD151 in primary human endothelial cells and through competitive blockade with glutathione-S-transferase (GST) fusion proteins that contain the second extracellular region of CD9.83 Therefore, the inclusion of ICAM-1 and VCAM-1 in tetraspanin domains is necessary for these domains to function properly in stringent dynamic conditions such as flow stress. VCAM-1 and ICAM-1 are not the only adhesion receptors that interact with tetraspanin microdomains, others such as JAM-A, PECAM-1, ICAM-2, or CD44 also do so. It could therefore be postulated that the tetraspanin microdomains act as specialized platforms that constitutively organize the appropriate adhesion receptors in the membrane for fast kinetics and efficient leukocyte extravasation.77

The endothelial adhesion receptors VCAM-1 and ICAM-1 are able to transmit signals after binding to the ligand. The VCAM-1 molecule is implicated in opening the interendothelial junctions to facilitate leukocyte extravasation. In fact, VCAM-1 induces the activation of NADPH oxidase (possibly NOX2) and production of reactive oxygen species (ROS) dependent on GTPase rac activity, with the subsequent activation of matrix metalloproteinases and loss of adhesion mediated by VE-cadherin due to phosphorylation of β-catenin by Pyk-2,84-89 thereby favoring the extravasation process. On the other hand, VCAM-1 and ICAM-1 are able to induce a rapid increase in Ca2+ concentration, leading to the activation of Src kinase and the subsequent phosphorylation of cortactin.90-93 ICAM-1 can also activate RhoA, inducing the formation of stress fibers and the phosphorylation of focal adhesion kinase (FAK), paxillin, and p130Cas, which in turn are implicated in signaling routes involving c-Jun N-terminal kinase (JNK) and p38.94-97 This increases endothelial permeability and is associated with increased transendothelial leukocyte migration. The induction of c-fos and rhoA transcription has also been reported via ICAM-1.96 Finally, ICAM-1 can induce its own expression and that of VCAM-1, acting as a regulating mechanism to facilitate leukocyte transmigration.98

INTEGRINS AND THEIR LIGANDS DURING ENDOTHELIAL TRANSMIGRATION

During endothelial transmigration, the endothelial junctions are partially dismantled to avoid damage to the monolayer or substantial changes in permeability. Thus, the leukocyte membranes and the endothelium remain in close contact during diapedesis and, afterwards, the endothelial membranes reseal their links.

Once the leukocytes have reached an appropriate site for transmigration (preferably the intercellular junctions), they deploy exploratory pseudopods between 2 adjacent endothelial cells. The pseudopods then transform into a lamella that moves across the open space on the monolayer. During this process, the LFA-1 molecule is the integrin with the predominant role. This molecule is quickly relocalized to form a ring-shaped cluster at the contact interface between the leukocyte and endothelium, where it interacts with ICAM-199 and, in some other cell models, with JAM-A.100 When the transmigration process is over, LFA-1 is finally concentrated in the uropod.101 Other proteins implicated in the transmigration process are ICAM-2, JAM-B, JAM-C, PECAM-1 (CD31), ESAM, and CD99. Many of these are able to interact both homophilically and heterophilically maintaining the interendothelial junctions or the leukocyteendothelial interactions.102-105

In the leukocyte transmigration process, in addition to the classic diapedesis route, in which the leukocytes cross the interendothelial junctions (paracellular route), there is increasing evidence of an alternative route in which the leukocytes can migrate through individual endothelial cells (transcellular route) without perturbing the interendothelial junctions. This process takes place preferentially in the microvasculature, the blood-brain barrier, or high endothelial venules of the secondary lymphoid organs rather than in the macrovasculature.106-108 Recent observations on the mechanism of this transcellular migration process indicate that, initially, the leukocytes generate invasive podosomes dependent on Src kinase and Wiskott-Aldrich syndrome protein (WASP) activity to palpate the endothelial surface. These podosomes subsequently develop into the transcellular pore. In the endothelium. It is necessary membrane fusion regulated by calcium and SNARE-containing complexes, as well as new membrane supply by vacuole-vesicular organulles.108 It has been also reported the translocation of ICAM-1 to caveolae after leukocyte adhesion and the subsequent formation of a kind of multivesicular channel containing ICAM-1 and caveolin-1 around a leukocyte pseudopod that penetrates through the endothelial cell. Both proteins, ICAM-1 and caveolin, follow the path of the entire leukocyte, moving towards the basal endothelial membrane.109 In addition, the intermediate filament protein vimentin also seems to play an important role in the transcellular route.110 The in vivo presence of dome-shaped endothelial structures that cover the leukocyte during transendothelial migration have also recently been described.111 These observations seem to indicate that the endothelial docking structures might become domes that completely envelope the leukocytes on the luminal face of the endothelium, thereby allowing rupture of the basolateral membrane without compromising the endothelial barrier function.

ANTIADHESION-BASED THERAPIES

The advances in our knowledge of the molecular mechanisms that underlie cell migration and the extravasation cascade have allowed molecular targets to be identified for antiadhesion therapy for inflammation. Monoclonal antibodies against the α4 and αL chains have shown a clear beneficial effect in different animal models of inflammatory and autoimmune states, as well as in human diseases such as multiple sclerosis, inflammatory bowel disease, and psoriasis. Similar results have been obtained in animal models with different VLA-4 synthetic peptides.

The promising results obtained in these animal studies have encouraged formation of to different groups and pharmaceutical companies dedicated to developing new drugs for clinical trials. Thus, a humanized anti-VLA-4 monoclonal antibody has shown a clear therapeutic effect in relapses of multiple sclerosis112 and Crohn disease.113 Other potential uses of this type of therapeutic monoclonal antibody would be in inflammatory and/or autoimmune diseases that are widespread in the general population, such as rheumatoid arthritis, asthma, and type 1 diabetes mellitus.114,115 As in the case of VLA-4, the anti-LFA-1 antibody has been shown to have a significant therapeutic effect in humans, as a humanized monoclonal anti-aL antibody has been approved to treated moderate to severe psoriasis.116

There is no doubt that the therapeutic monoclonal antibodies directed against adhesion molecules or costimulators represent an important step forward in treating inflammation and autoimmune diseases. However, biological agents, such as monoclonal antibodies against a4 and αL leukocyte integrin chains are directed against receptors with a range of different biological functions: the generation of immune response, differentiation of lymphocytes into Th1/Th2,117 the effector phase of immune cells, and the extravasation of leukocytes to the inflammatory foci, among others. In addition, it is evident that some monoclonal antibodies may act as agonist molecules, thereby generating intracellular signals after binding to their antigen. Thus, long-term administration of these types of drugs could have unexpected and even undesirable consequences. It would be very important to take into account all information derived from both basic studies and preclinical research to enable the appropriate design of future clinical trials with this type of biologic agent (reviewed González Amaro et al118).

A large number of anti-inflammatory therapies against different target molecules in the endothelium have also been investigated. The efficacy of blockade of P-selectin to prevent damage caused during ischemia-reperfusion processes (transplantation, thrombosis, stroke, etc) as well as the beneficial effects of anti-ICAM-1 antibodies for preventing restenotic lesions in animals have been investigated and, as a result, they are emerging as possible therapeutic targets in humans.119,120 There are also numerous studies of chronic autoimmune or inflammatory diseases in which therapies based on anti-TNF, anti-VCAM-1, or anti-ICAM-1 have been applied. Recent research points to tetraspanins and, specifically, CD9, as a potential general anti-inflammatory target that could regulate the adhesive function of multiple adhesion receptors and be more effective than individual inhibition of each one separately. However, the tetraspanin CD9 is ubiquitously expressed throughout the organism, and so the release of CD9 blockers should be done locally and restricted to the area of inflammation. Prior studies on knock-out mice lacking expression of CD9 and other tetraspanins are necessary to clarify whether this hypothesis is plausible. Indeed, recently, the role of the endothelial tetraspanin CD81 was described as a possible diagnostic and therapeutic marker of atherogenesis in humans. The expression of CD81 on the luminal surface of the endothelium increases in the initial states of the disease, and so this molecule could play a crucial role in the formation of atherosclerotic plaque by favoring adhesion of monocytes in the stage prior to the triggering of the inflammatory response.121

Section sponsored by Laboratory Dr Esteve

Financed by project RD06/0014-0030 of the Thematic Networks for Collaborative Research into Cardiovascular Diseases (RECAVA). The Centro Nacional de Investigaciones Cardiovasculares is funded by the Ministry for Health and Consumer Affairs and the Fundación Pro-CNIC.


Correspondence:
Prof. F. Sánchez-Madrid.
Servicio de Inmunología. Hospital Universitario de la Princesa. Universidad Autónoma de Madrid. Diego de León, 62. 28006 Madrid. España.
E-mail: fsanchez.hlpr@salud.madrid.org

Bibliography
[1]
Frenette PS, Wagner DD..
Adhesion molecules—Part I.
N Engl J Med, (1996), 334 pp. 1526-9
[2]
Frenette PS, Wagner DD..
Adhesion molecules—Part II: Blood vessels and blood cells.
N Engl J Med, (1996), 335 pp. 43-5
[3]
Aplin AE, Howe A, Alahari SK, Juliano RL..
Signal transduction and signal modulation by cell adhesion receptors: the role of integrins, cadherins, immunoglobulin-cell adhesion molecules, and selectins.
Pharmacol Rev, (1998), 50 pp. 197-263
[4]
Levesque JP, Zannettino AC, Pudney M, Niutta S, Haylock DN, Snapp KR, et al..
PSGL-1-mediated adhesion of human hematopoietic progenitors to P-selectin results in suppression of hematopoiesis.
Immunity, (1999), 11 pp. 369-78
[5]
Verfaillie CM..
Adhesion receptors as regulators of the hematopoietic process.
Blood, (1998), 92 pp. 2609-12
[6]
Butcher EC..
Leukocyte-endothelial cell recognition: three (or more) steps to specificity and diversity.
Cell, (1991), 67 pp. 1033-6
[7]
Butcher EC, Picker LJ..
Lymphocyte homing and homeostasis.
Science, (1996), 272 pp. 60-6
[8]
Springer TA..
Traffic signals for lymphocyte recirculation and leukocyte emigration: the multiple paradigm.
Cell, (1994), 76 pp. 301-14
[9]
Alon R, Ley K..
Cells on the run: shear-regulated integrin activation in leukocyte rolling and arrest on endothelial cells.
Curr Opin Cell Biol, (2008), 20 pp. 525-32
[10]
Mehta P, Cummings RD, McEver RP..
Affinity and kinetic analysis of P-selectin binding to P-selectin glycoprotein ligand-1.
J Biol Chem, (1998), 273 pp. 32506-13
[11]
Nicholson MW, Barclay AN, Singer MS, Rosen SD, van der Merwe PA..
Affinity and kinetic analysis of L-selectin (CD62L) binding to glycosylation-dependent cell-adhesion molecule-1.
J Biol Chem, (1998), 273 pp. 763-70
[12]
Evans EA, Calderwood DA..
Forces and bond dynamics in cell adhesion.
Science, (2007), 316 pp. 1148-53
[13]
Barreiro O, Vicente-Manzanares M, Urzainqui A, Yáñez-Mó M, Sánchez-Madrid F..
Interactive protrusive structures during leukocyte adhesion and transendothelial migration.
Front Biosci, (2004), 9 pp. 1849-63
[14]
Eriksson EE, Xie X, Werr J, Thoren P, Lindbom L..
Importance of primary capture and L-selectin-dependent secondary capture in leukocyte accumulation in inflammation and atherosclerosis in vivo.
J Exp Med, (2001), 194 pp. 205-18
[15]
Hidalgo A, Peired AJ, Wild MK, Vestweber D, Frenette PS..
Complete identification of E-selectin ligands on neutrophils reveals distinct functions of PSGL-1, ESL-1, and CD44.
[16]
von Hundelshausen P, Koenen RR, Weber C..
Platelet-mediated enhancement of leukocyte adhesion.
Microcirculation, (2009), 16 pp. 84-96
[17]
Alon R, Kassner P, Carr M, Finger E, Hemler M, Springer T..
The integrin VLA-4 supports tethering and rolling in flow on VCAM-1.
J Cell Biol, (1995), 128 pp. 1243-53
[18]
Berlin C, Bargatze R, Campbell J, von Andrian U, Szabo M, Hasslen S, et al..
Alpha 4 integrins mediate lymphocyte attachment and rolling under physiologic flow.
Cell, (1995), 80 pp. 413-22
[19]
Berlin C, Berg EL, Briskin MJ, Andrew DP, Kilshaw PJ, Holzmann B, et al..
Alpha 4 beta 7 integrin mediates lymphocyte binding to the mucosal vascular addressin MAdCAM-1.
Cell, (1993), 74 pp. 185-95
[20]
Henderson RB, Lim LH, Tessier PA, Gavins FN, Mathies M, Perretti M, et al..
The use of lymphocyte function-associated antigen (LFA)-1-deficient mice to determine the role of LFA-1, Mac-1, and alpha4 integrin in the inflammatory response of neutrophils.
J Exp Med, (2001), 194 pp. 219-26
[21]
Kadono T, Venturi GM, Steeber DA, Tedder TF..
Leukocyte rolling velocities and migration are optimized by cooperative L-selectin and intercellular adhesion molecule-1 functions.
J Immunol, (2002), 169 pp. 4542-50
[22]
von Andrian UH, Hasslen SR, Nelson RD, Erlandsen SL, Butcher EC..
A central role for microvillous receptor presentation in leukocyte adhesion under flow.
Cell, (1995), 82 pp. 989-99
[23]
Dwir O, Kansas GS, Alon R..
Cytoplasmic anchorage of L-selectin controls leukocyte capture and rolling by increasing the mechanical stability of the selectin tether.
J Cell Biol, (2001), 155 pp. 145-56
[24]
Ivetic A, Deka J, Ridley AJ, Ager A..
The cytoplasmic tail of L-selectin interacts with members of the Ezrin-Radixin-Moesin (ERM) family of proteins: cell activation-dependent binding of Moesin but not Ezrin.
J Biol Chem, (2002), 277 pp. 2321-9
[25]
Pavalko FM, Walker DM, Graham L, Goheen M, Doerschuk CM, Kansas GS..
The cytoplasmic domain of L-selectin interacts with cytoskeletal proteins via alpha-actinin: receptor positioning in microvilli does not require interaction with alpha-actinin.
J Cell Biol, (1995), 129 pp. 1155-64
[26]
Killock DJ, Parsons M, Zarrouk M, Ameer-Beg SM, Ridley AJ, Haskard DO, et al..
In vitro and in vivo characterization of molecular interactions between calmodulin, ezrin/radixin/ moesin (ERM) and L-selectin.
J Biol Chem, (2009 Jan 7 [Epub ahead of print]),
[27]
Urzainqui A, Martínez del Hoyo G, Lamana A, de la Fuente H, Barreiro O, Olazabal IM, et al..
Functional role of P-selectin glycoprotein ligand 1/P-selectin interaction in the generation of tolerogenic dendritic cells.
J Immunol, (2007), 179 pp. 7457-65
[28]
Urzainqui A, Serrador JM, Viedma F, Yáñez-Mó M, Rodríguez A, Corbí AL, et al..
ITAM-based interaction of ERM proteins with Syk mediates signaling by the leukocyte adhesion receptor PSGL-1.
Immunity, (2002), 17 pp. 401-12
[29]
Zarbock A, Abram CL, Hundt M, Altman A, Lowell CA, Ley K..
PSGL-1 engagement by E-selectin signals through Src kinase Fgr and ITAM adapters DAP12 and FcR gamma to induce slow leukocyte rolling.
J Exp Med, (2008), 205 pp. 2339-47
[30]
Zarbock A, Lowell CA, Ley K..
Spleen tyrosine kinase Syk is necessary for E-selectin-induced alpha(L)beta(2) integrin-mediated rolling on intercellular adhesion molecule-1.
[31]
von Andrian UH, Mackay CR..
T-cell function and migration. Two sides of the same coin.
N Engl J Med, (2000), 343 pp. 1020-34
[32]
Hynes RO..
Integrins: bidirectional, allosteric signaling machines.
Cell, (2002), 110 pp. 673-87
[33]
Campbell JJ, Hedrick J, Zlotnik A, Siani MA, Thompson DA, Butcher EC..
Chemokines and the arrest of lymphocytes rolling under flow conditions.
Science, (1998), 279 pp. 381-4
[34]
Vicente-Manzanares M, Sánchez-Madrid F..
Role of the cytoskeleton during leukocyte responses.
Nat Rev Immunol, (2004), 4 pp. 110-22
[35]
Barreiro O, de la Fuente H, Mittelbrunn M, Sánchez-Madrid F..
Functional insights on the polarized redistribution of leukocyte integrins and their ligands during leukocyte migration and immune interactions.
Immunol Rev, (2007), 218 pp. 147-64
[36]
Carman CV, Springer TA..
Integrin avidity regulation: are changes in affinity and conformation underemphasized? Curr Opin Cell Biol.
, (2003), 15 pp. 547-56
[37]
Beglova N, Blacklow SC, Takagi J, Springer TA..
Cysteine-rich module structure reveals a fulcrum for integrin rearrangement upon activation.
Nat Struct Biol, (2002), 9 pp. 282-7
[38]
Nishida N, Xie C, Shimaoka M, Cheng Y, Walz T, Springer TA..
Activation of leukocyte beta2 integrins by conversion from bent to extended conformations.
[39]
Gahmberg CG, Nortamo P, Kantor C, Autero M, Kotovuori P, Hemio L, et al..
The pivotal role of the Leu-CAM and ICAM molecules in human leukocyte adhesion.
Cell Differ Dev, (1990), 32 pp. 239-45
[40]
Dustin ML, Rothlein R, Bhan AK, Dinarello CA, Springer TA..
Induction by IL 1 and interferon-gamma: tissue distribution, biochemistry, and function of a natural adherence molecule (ICAM-1).
J Immunol, (1986), 137 pp. 245-54
[41]
Acevedo A, del Pozo MA, Arroyo AG, Sánchez-Mateos P, González-Amaro R, Sánchez-Madrid F..
Distribution of ICAM-3-bearing cells in normal human tissues. Expression of a novel counter-receptor for LFA-1 in epidermal Langerhans cells.
Am J Pathol, (1993), 143 pp. 774-83
[42]
Ostermann G, Weber KS, Zernecke A, Schroder A, Weber C..
JAM-1 is a ligand of the beta(2) integrin LFA-1 involved in transendothelial migration of leukocytes.
Nat Immunol, (2002), 3 pp. 151-8
[43]
Chavakis T, Bierhaus A, Al-Fakhri N, Schneider D, Witte S, Linn T, et al..
The pattern recognition receptor (RAGE) is a counterreceptor for leukocyte integrins: a novel pathway for inflammatory cell recruitment.
J Exp Med, (2003), 198 pp. 1507-15
[44]
Lamagna C, Meda P, Mandicourt G, Brown J, Gilbert RJ, Jones EY, et al..
Dual interaction of JAM-C with JAM-B and alpha(M)beta2 integrin: function in junctional complexes and leukocyte adhesion.
Mol Biol Cell, (2005), 16 pp. 4992-5003
[45]
Elices MJ, Osborn L, Takada Y, Crouse C, Luhowskyj S, Hemler ME, et al..
VCAM-1 on activated endothelium interacts with the leukocyte integrin VLA-4 at a site distinct from the VLA-4/fibronectin binding site.
Cell, (1990), 60 pp. 577-84
[46]
Carlos TM, Harlan JM..
Leukocyte-endothelial adhesion molecules.
Blood, (1994), 84 pp. 2068-101
[47]
Cunningham SA, Rodríguez JM, Arrate MP, Tran TM, Brock TA..
JAM2 interacts with alpha4beta1. Facilitation by JAM3.
J Biol Chem, (2002), 277 pp. 27589-92
[48]
Mittelbrunn M, Cabanas C, Sánchez-Madrid F..
Integrin alpha4. AfCS-Nature Molecule Pages. 2006 20 Jul. doi:10.1038/mp.
a001203, (01),
[49]
Alon R, Grabovsky V, Feigelson S..
Chemokine induction of integrin adhesiveness on rolling and arrested leukocytes local signaling events or global stepwise activation? Microcirculation.
, (2003), 10 pp. 297-311
[50]
Rot A, von Andrian UH..
Chemokines in innate and adaptive host defense: basic chemokinese grammar for immune cells.
Annu Rev Immunol, (2004), 22 pp. 891-928
[51]
Del Pozo MA, Sánchez-Mateos P, Nieto M, Sánchez-Madrid F..
Chemokines regulate cellular polarization and adhesion receptor redistribution during lymphocyte interaction with endothelium and extracellular matrix. Involvement of cAMP signaling pathway.
J Cell Biol, (1995), 131 pp. 495-508
[52]
Geiger B, Bershadsky A..
Exploring the neighborhood: adhesion-coupled cell mechanosensors.
Cell, (2002), 110 pp. 139-42
[53]
Constantin G, Majeed M, Giagulli C, Piccio L, Kim JY, Butcher EC, et al..
Chemokines trigger immediate beta2 integrin affinity and mobility changes: differential regulation and roles in lymphocyte arrest under flow.
Immunity, (2000), 13 pp. 759-69
[54]
Sánchez-Madrid F, del Pozo MA..
Leukocyte polarization in cell migration and immune interactions.
EMBO J, (1999), 18 pp. 501-11
[55]
Shamri R, Grabovsky V, Gauguet JM, Feigelson S, Manevich E, Kolanus W, et al..
Lymphocyte arrest requires instantaneous induction of an extended LFA-1 conformation mediated by endothelium-bound chemokines.
Nat Immunol, (2005), 6 pp. 497-506
[56]
Laudanna C, Alon R..
Right on the spot. Chemokine triggering of integrin-mediated arrest of rolling leukocytes.
Thromb Haemost, (2006), 95 pp. 5-11
[57]
Luster AD..
Chemokines—chemotactic cytokines that mediate inflammation.
N Engl J Med, (1998), 338 pp. 436-45
[58]
Laudanna C..
Integrin activation under flow: a local affair.
Nat Immunol, (2005), 6 pp. 429-30
[59]
Cabanas C, Hogg N..
Ligand intercellular adhesion molecule 1 has a necessary role in activation of integrin lymphocyte function-associated molecule 1.
Proc Natl Acad Sci U S A, (1993), 90 pp. 5838-42
[60]
Jun CD, Shimaoka M, Carman CV, Takagi J, Springer TA..
Dimerization and the effectiveness of ICAM-1 in mediating LFA-1-dependent adhesion.
Proc Natl Acad Sci U S A, (2001), 98 pp. 6830-5
[61]
Salas A, Shimaoka M, Kogan AN, Harwood C, von Andrian UH, Springer TA..
Rolling adhesion through an extended conformation of integrin alphaLbeta2 and relation to alpha I and beta I-like domain interaction.
Immunity, (2004), 20 pp. 393-406
[62]
Grabovsky V, Feigelson S, Chen C, Bleijs DA, Peled A, Cinamon G, et al..
Subsecond induction of alpha4integrin clustering by immobilized chemokines stimulates leukocyte tethering and rolling on endothelial vascular cell adhesion molecule 1 under flow conditions.
J Exp Med, (2000), 192 pp. 495-506
[63]
Dobereiner HG, Dubin-Thaler BJ, Hofman JM, Xenias HS, Sims TN, Giannone G, et al..
Lateral membrane waves constitute a universal dynamic pattern of motile cells.
Phys Rev Lett, (2006), 97 pp. 038102
[64]
Stewart MP, McDowall A, Hogg N..
LFA-1-mediated adhesion is regulated by cytoskeletal restraint and by a Ca2+-dependent protease, calpain.
J Cell Biol, (1998), 140 pp. 699-707
[65]
Mory A, Feigelson SW, Yarali N, Kilic SS, Bayhan GI, Gershoni-Baruch R, et al..
Kindlin-3: a new gene involved in the pathogenesis of LAD-III.
[66]
Pasvolsky R, Feigelson SW, Kilie SS, Simon AJ, Tal-Lapidot G, Grabovsky V, et al..
A LAD-III Syndrome is associated with defective expression of the Rap-1 activator CalDAGGEFI in lymphocytes, neutrophils, and platelets.
J Exp Med, (2007), 204 pp. 1571-82
[67]
Cairo CW, Mirchev R, Golan DE..
Cytoskeletal regulation couples LFA-1 conformational changes to receptor lateral mobility and clustering.
Immunity, (2006), 25 pp. 297-308
[68]
Cambi A, Joosten B, Koopman M, de Lange F, Beeren I, Torensma R, et al..
Organization of the integrin LFA-1 in nanoclusters regulates its activity.
Mol Biol Cell, (2006), 17 pp. 4270-81
[69]
Marschel P, Schmid-Schonbein GW..
Control of fluid shear response in circulating leukocytes by integrins.
Ann Biomed Eng, (2002), 30 pp. 333-43
[70]
Zwartz GJ, Chigaev A, Dwyer DC, Foutz TD, Edwards BS, Sklar LA..
Real-time analysis of very late antigen-4 affinity modulation by shear.
J Biol Chem, (2004), 279 pp. 38277-86
[71]
Cinamon G, Shinder V, Alon R..
Shear forces promote lymphocyte migration across vascular endothelium bearing apical chemokines.
Nat Immunol, (2001), 2 pp. 515-22
[72]
Schenkel AR, Mamdouh Z, Muller WA..
Locomotion of monocytes on endothelium is a critical step during extravasation.
Nat Immunol, (2004), 5 pp. 393-400
[73]
Phillipson M, Heit B, Colarusso P, Liu L, Ballantyne CM, Kubes P..
Intraluminal crawling of neutrophils to emigration sites: a molecularly distinct process from adhesion in the recruitment cascade.
J Exp Med, (2006), 203 pp. 2569-75
[74]
Marlin SD, Springer TA..
Purified intercellular adhesion molecule-1 (ICAM-1) is a ligand for lymphocyte function-associated antigen 1 (LFA-1).
Cell, (1987), 51 pp. 813-9
[75]
Barreiro O, Yánez-Mó M, Serrador JM, Montoya MC, Vicente-Manzanares M, Tejedor R, et al..
Dynamic interaction of VCAM-1 and ICAM-1 with moesin and ezrin in a novel endothelial docking structure for adherent leukocytes.
J Cell Biol, (2002), 157 pp. 1233-45
[76]
Heiska L, Alfthan K, Gronholm M, Vilja P, Vaheri A, Carpen O..
Association of ezrin with intercellular adhesion molecule-1 and -2 (ICAM-1 and ICAM-2). Regulation by phosphatidylinositol 4, 5- bisphosphate.
J Biol Chem, (1998), 273 pp. 21893-900
[77]
Barreiro O, Zamai M, Yánez-Mó M, Tejera E, López-Romero P, Monk PN, et al..
Endothelial adhesion receptors are recruited to adherent leukocytes by inclusion in preformed tetraspanin nanoplatforms.
J Cell Biol, (2008), 183 pp. 527-42
[78]
Hemler ME..
Tetraspanin functions and associated microdomains.
Nat Rev Mol Cell Biol, (2005), 6 pp. 801-11
[79]
Gordon-Alonso M, Yáñez-Mó M, Barreiro O, Álvarez S, Muñoz-Fernández MA, Valenzuela-Fernández A, et al..
Tetraspanins CD9 and CD81 modulate HIV-1-induced membrane fusion.
J Immunol, (2006), 177 pp. 5129-37
[80]
Mittelbrunn M, Yáñez-Mó M, Sancho D, Ursa A, Sánchez-Madrid F..
Cutting edge: dynamic redistribution of tetraspanin CD81 at the central zone of the immune synapse in both T lymphocytes and APC.
J Immunol, (2002), 169 pp. 6691-5
[81]
Yáñez-Mó M, Alfranca A, Cabanas C, Marazuela M, Tejedor R, Ursa MA, et al..
Regulation of endothelial cell motility by complexes of tetraspan molecules CD81/TAPA-1 and CD151/ PETA-3 with alpha3 beta1 integrin localized at endothelial lateral junctions.
J Cell Biol, (1998), 141 pp. 791-804
[82]
García-López MA, Barreiro O, García-Díez A, Sánchez-Madrid F, Penas PF..
Role of tetraspanins CD9 and CD151 in primary melanocyte motility.
J Invest Dermatol, (2005), 125 pp. 1001-9
[83]
Barreiro O, Yáñez-Mó M, Sala-Valdes M, Gutiérrez-López MD, Ovalle S, Higginbottom A, et al..
Endothelial tetraspanin microdomains regulate leukocyte firm adhesion during extravasation.
Blood, (2005), 105 pp. 2852-61
[84]
Cook-Mills JM..
VCAM-1 signals during lymphocyte migration: role of reactive oxygen species.
Mol Immunol, (2002), 39 pp. 499-508
[85]
van Wetering S, van Buul JD, Quik S, Mul FP, Anthony EC, ten Klooster JP, et al..
Reactive oxygen species mediate Rac-induced loss of cell-cell adhesion in primary human endothelial cells.
J Cell Sci, (2002), 115 pp. 1837-46
[86]
van Wetering S, van Den Berk N, van Buul JD, Mul FP, Lommerse I, Mous R, et al..
VCAM-1-mediated Rac signaling controls endothelial cell-cell contacts and leukocyte transmigration.
Am J Physiol Cell Physiol, (2003), 285 pp. C343-52
[87]
Deem TL, Cook-Mills JM..
Vascular cell adhesion molecule 1 (VCAM-1) activation of endothelial cell matrix metalloproteinases: role of reactive oxygen species.
Blood, (2004), 104 pp. 2385-93
[88]
van Buul JD, Fernández-Borja M, Anthony EC, Hordijk PL..
Expression and localization of NOX2 and NOX4 in primary human endothelial cells.
Antioxid Redox Signal, (2005), 7 pp. 308-17
[89]
van Buul JD, Anthony EC, Fernández-Borja M, Burridge K, Hordijk PL..
Proline-rich tyrosine kinase 2 (Pyk2) mediates vascular endothelial-cadherin-based cell-cell adhesion by regulating beta-catenin tyrosine phosphorylation.
J Biol Chem, (2005), 280 pp. 21129-36
[90]
Etienne-Manneville S, Manneville JB, Adamson P, Wilbourn B, Greenwood J, Couraud PO..
ICAM-1-coupled cytoskeletal rearrangements and transendothelial lymphocyte migration involve intracellular calcium signaling in brain endothelial cell lines.
J Immunol, (2000), 165 pp. 3375-83
[91]
Lorenzon P, Vecile E, Nardon E, Ferrero E, Harlan JM, Tedesco F, et al..
Endothelial cell E- and P-selectin and vascular cell adhesion molecule-1 function as signaling receptors.
J Cell Biol, (1998), 142 pp. 1381-91
[92]
Yang L, Kowalski JR, Yacono P, Bajmoczi M, Shaw SK, Froio RM, et al..
Endothelial cell cortactin coordinates intercellular adhesion molecule-1 clustering and actin cytoskeleton remodeling during polymorphonuclear leukocyte adhesion and transmigration.
J Immunol, (2006), 177 pp. 6440-9
[93]
Yang L, Kowalski JR, Zhan X, Thomas SM, Luscinskas FW..
Endothelial cell cortactin phosphorylation by Src contributes to polymorphonuclear leukocyte transmigration in vitro.
[94]
Greenwood J, Etienne-Manneville S, Adamson P, Couraud PO..
Lymphocyte migration into the central nervous system: implication of ICAM-1 signalling at the blood-brain barrier.
Vascul Pharmacol, (2002), 38 pp. 315-22
[95]
Hubbard AK, Rothlein R..
Intercellular adhesion molecule-1 (ICAM-1) expression and cell signaling cascades.
Free Radic Biol Med, (2000), 28 pp. 1379-86
[96]
Thompson PW, Randi AM, Ridley AJ..
Intercellular adhesion molecule (ICAM)-1, but not ICAM-2, activates RhoA and stimulates c-fos and rhoA transcription in endothelial cells.
J Immunol, (2002), 169 pp. 1007-13
[97]
Wang Q, Doerschuk CM..
The signaling pathways induced by neutrophil-endothelial cell adhesion.
Antioxid Redox Signal, (2002), 4 pp. 39-47
[98]
Clayton A, Evans RA, Pettit E, Hallett M, Williams JD, Steadman R..
Cellular activation through the ligation of intercellular adhesion molecule-1.
J Cell Sci, (1998), 111 pp. 443-53
[99]
Shaw SK, Ma S, Kim MB, Rao RM, Hartman CU, Froio RM, et al..
Coordinated redistribution of leukocyte LFA-1 and endothelial cell ICAM-1 accompany neutrophil transmigration.
J Exp Med, (2004), 200 pp. 1571-80
[100]
Woodfin A, Reichel CA, Khandoga A, Corada M, Voisin MB, Scheiermann C, et al..
JAM-A mediates neutrophil transmigration in a stimulus-specific manner in vivo: evidence for sequential roles for JAM-A and PECAM-1 in neutrophil transmigration.
[101]
Sandig M, Negrou E, Rogers KA..
Changes in the distribution of LFA-1, catenins, and F-actin during transendothelial migration of monocytes in culture.
J Cell Sci, (1997), 110 pp. 2 807-18
[102]
Vestweber D..
Adhesion and signaling molecules controlling the transmigration of leukocytes through endothelium.
Immunol Rev, (2007), 218 pp. 178-96
[103]
Weber C, Fraemohs L, Dejana E..
The role of junctional adhesion molecules in vascular inflammation.
Nat Rev Immunol, (2007), 7 pp. 467-77
[104]
Woodfin A, Voisin MB, Nourshargh S..
PECAM-1: a multi-functional molecule in inflammation and vascular biology.
Arterioscler Thromb Vasc Biol, (2007), 27 pp. 2514-23
[105]
Muller WA..
Leukocyte-endothelial-cell interactions in leukocyte transmigration and the inflammatory response.
Trends Immunol, (2003), 24 pp. 327-34
[106]
Engelhardt B, Wolburg H..
Mini-review: Transendothelial migration of leukocytes: through the front door or around the side of the house? Eur J Immunol.
, (2004), 34 pp. 2955-63
[107]
Carman CV, Springer TA..
A transmigratory cup in leukocyte diapedesis both through individual vascular endothelial cells and between them.
J Cell Biol, (2004), 167 pp. 377-88
[108]
Carman CV, Sage PT, Sciuto TE, de la Fuente MA, Geha RS, Ochs HD, et al..
Transcellular diapedesis is initiated by invasive podosomes.
[109]
Millan J, Hewlett L, Glyn M, Toomre D, Clark P, Ridley AJ..
Lymphocyte transcellular migration occurs through recruitment of endothelial ICAM-1 to caveola- and F-actin-rich domains.
Nat Cell Biol, (2006), 8 pp. 113-23
[110]
Nieminen M, Henttinen T, Merinen M, Marttila-Ichihara F, Eriksson JE, Jalkanen S..
Vimentin function in lymphocyte adhesion and transcellular migration.
Nat Cell Biol, (2006), 8 pp. 156-62
[111]
Phillipson M, Kaur J, Colarusso P, Ballantyne CM, Kubes P..
Endothelial domes encapsulate adherent neutrophils and minimize increases in vascular permeability in paracellular and transcellular emigration.
[112]
Chaudhuri A, Behan PO..
Natalizumab for relapsing multiple sclerosis.
N Engl J Med, (2003), 348 pp. 1598-9
[113]
Lew EA, Stoffel EM..
Natalizumab for active Crohn''s disease.
N Engl J Med, (2003), 348 pp. 1599
[114]
Noseworthy JH, Kirkpatrick P..
Natalizumab.
Nat Rev Drug Discov, (2005), 4 pp. 101-2
[115]
von Andrian UH, Engelhardt B..
Alpha4 integrins as therapeutic targets in autoimmune disease.
N Engl J Med, (2003), 348 pp. 68-72
[116]
Marecki S, Kirkpatrick P..
Efalizumab.
Nat Rev Drug Discov, (2004), 3 pp. 473-4
[117]
Mittelbrunn M, Molina A, Escribese MM, Yáñez-Mó M, Escudero E, Ursa A, et al..
VLA-4 integrin concentrates at the peripheral supramolecular activation complex of the immune synapse and drives T helper 1 responses.
Proc Natl Acad Sci U S A, (2004), 101 pp. 11058-63
[118]
González-Amaro R, Mittelbrunn M, Sánchez-Madrid F..
Therapeutic anti-integrin (alpha4 and alphaL) monoclonal antibodies: two-edged swords? Immunology.
, (2005), 116 pp. 289-96
[119]
Chamoun F, Burne M, O''Donnell M, Rabb H..
Pathophysiologic role of selectins and their ligands in ischemia reperfusion injury.
Front Biosci, (2000), 5 pp. E103-9
[120]
Kollum M, Hoefer I, Schreiber R, Bode C, Hehrlein C..
Systemic application of anti-ICAM-1 monoclonal antibodies to prevent restenosis in rabbits: an anti-inflammatory strategy.
Coron Artery Dis, (2007), 18 pp. 117-23
[121]
Tohlena J, Volger OL, van Buul JD, Hekking LH, van Gils JM, Bonta PI, et al..
Endothelial CD81 is a marker of early human atherosclerotic plaques and facilitates monocyte adhesion.
Cardiovasc Res, (2009), 81 pp. 187-96
Are you a healthcare professional authorized to prescribe or dispense medications?