THE ENDOTHELIUM has long been viewed as an inert cellophane-like membrane that lines the circulatory system with its primary essential function being the maintenance of vessel wall permeability. Shortly after the first description of circulating blood by William Harvey in 1628, the existence of a network of vessels arose from studies of Malphigi, who described the physical separation between blood and tissue.1 In the 1800s, von Reckingausen established that vessels were not merely tunnels bored through tissues but were lined by cells. The strength of Starling's experiments and his law of capillary exchange proposed in 1896 served to solidify the belief that the endothelium was principally a selective but static physical barrier, not withstanding Heidenhahn's description in 1891 of the endothelium as an active secretory cell system. However, electron microscopic studies of the vessel wall by Palade in 1953 and physiological studies by Gowan in 1959 describing the interaction between lymphocytes and endothelium of postcapillary venules stimulated numerous subsequent studies that led to the current view of the endothelium as a dynamic, heterogeneous, disseminated organ that possesses vital secretory, synthetic, metabolic, and immunologic functions.1 

The endothelial cell (EC) surface in an adult human is composed of approximately 1 to 6 × 1013 cells, weighs approximately 1 kg, and covers a surface area of approximately 1 to 7 m2.2 ECs line vessels in every organ system and regulate the flow of nutrient substances, diverse biologically active molecules, and the blood cells themselves. This gate-keeping role of endothelium is effected through the presence of membrane-bound receptors for numerous molecules including proteins (eg, growth factors, coagulant, and anticoagulant proteins), lipid transporting particle (eg, low-density lipoprotein [LDL]), metabolites (eg, nitrous oxide and serotonin), and hormones (eg, endothelin-1), as well as through specific junctional proteins and receptors that govern cell-cell and cell-matrix interactions.

The endothelium also plays a pivotal role in regulating blood flow. In part, this results from the capacity of quiescent ECs to generate an active antithrombotic surface that facilitates transit of plasma and cellular constituents throughout the vasculature. Perturbations, such as those that may occur at sites of inflammation or high hydrodynamic shear stress, disrupt these activities and induce ECs to create a prothrombotic and antifibrinolytic microenvironment. Blood flow is also regulated, in part, through secretion and uptake of vasoactive substances by the endothelium that act in a paracrine manner to constrict and dilate specific vascular beds in response to stimuli such as endotoxin.

Detailed study of endothelial function first became feasible with the development in the 1970s of techniques to culture ECs in vitro.3-5 Limitations of this approach have become apparent recently with the realization that cell culture perturbs ECs from their quiescent in vivo state (0.1% replications per day) to an activated phenotype (1% to 10% replications per day) with loss of specialized functions associated with diverse vessels and organ systems. More complex analytic systems now exist that incorporate changes in EC properties imparted by plasma and cellular blood elements, by rheologic factors, and by cell-cell interactions that occur within the vessel wall. Genetic recombination studies in mice are likely to advance understanding of ECs in both their physiologic and pathologic roles in thrombosis, atherosclerosis, tumor metastasis, and organ rejection.

The purpose of this review is to provide a broad overview of EC participation in several biological processes judged to be relevant to clinical hematologists and investigators of vascular biology. Part I will principally examine the known physiologic roles of the endothelium, whereas Part II will discuss the interactions between ECs and blood cells and emphasize the contribution of the EC to the pathogenesis of specific diseases. Because of the introductory nature of this review, many topics and important contributions have been omitted, including the involvement of ECs in hematopoiesis, neuroendocrinology, cell aging, cellular integrins and matrix interactions, vascular permeability, lipid metabolism, the lymphatic vasculature, and the endothelium as a target for gene therapy. We hope that the abbreviated bibliographies will serve as an introduction to readers who wish to further investigate EC biology.

Overview of early vascular development.

Recently developed techniques that permit alteration of genomic sequences and manipulation of developing embryonic tissues have provided important insights into molecular and genetic elements that regulate vascular development.6 These studies show that the cardiovasculature is the first system to form in the gastrulating embryo. The de novo organization of ECs into vessels in the absence of any pre-existing vascular system is referred to as vasculogenesis and only occurs in the early embryo. Angiogenesis, the continued expansion of the vascular tree as a result of ECs sprouting from existing vessels, occurs in avascular regions of the embryo and is repeated many times in the mature animal, most commonly during wound healing and tumor metastasis7 (Fig 1). It remains uncertain how the pattern of the vascular tree is established or which factors govern the site of sprouting or the route taken by migrating ECs during angiogenic expansion.

Fig. 1.

The formation of new vessels during vasculogenesis and angiogenesis. Vasculogenesis, the de novo organization of ECs into vessels in the absence of preexisting vascular structures, takes place during embryogenesis in the blood islands of the yolk sac (pictured) and in the embryo through expression of growth factors, in particular fibroblast growth factor (FGF) and vascular endothelial growth factor (VEGF). The tyrosine receptor kinases, VEGFR-1 (flk-1) and VEGFR-2 (flt-1), are expressed on mesenchymal cells and newly formed ECs, respectively, and are essential for the generation and proliferation of new ECs and the formation of tubal EC structures. Angiogenesis, the continued expansion of the vascular tree, is mediated through the expression of additional tyrosine kinase receptors, tie-2 (tek), which binds to Ang1 and Ang2 (angiopoietins), resulting in the maintenance of mature vessels, the development of new vessels, and the regression of formed vessels in processes dependent on a combination of factors, most notably the presence or absence of growth factors.

Fig. 1.

The formation of new vessels during vasculogenesis and angiogenesis. Vasculogenesis, the de novo organization of ECs into vessels in the absence of preexisting vascular structures, takes place during embryogenesis in the blood islands of the yolk sac (pictured) and in the embryo through expression of growth factors, in particular fibroblast growth factor (FGF) and vascular endothelial growth factor (VEGF). The tyrosine receptor kinases, VEGFR-1 (flk-1) and VEGFR-2 (flt-1), are expressed on mesenchymal cells and newly formed ECs, respectively, and are essential for the generation and proliferation of new ECs and the formation of tubal EC structures. Angiogenesis, the continued expansion of the vascular tree, is mediated through the expression of additional tyrosine kinase receptors, tie-2 (tek), which binds to Ang1 and Ang2 (angiopoietins), resulting in the maintenance of mature vessels, the development of new vessels, and the regression of formed vessels in processes dependent on a combination of factors, most notably the presence or absence of growth factors.

Close modal
Origin of the vascular endothelium.

Molecular events involved in EC differentiation from the early mesoderm remain uncertain. Vascular and hematopoietic tissues develop together, beginning shortly after implantation with the formation of blood islands within the primitive yolk sac8 (Fig 1) composed of two cell types: (1) angioblasts that form the outer layer of ECs encasing the blood island; and (2) hematopoietic stem cells, in the inner cluster, from which the first embryonic blood cells develop. Angioblasts committed to EC differentiation are found primarily in embryonic mesoderm,9 whereas the early epiblast also contains a subpopulation of hematopoietic stem cells.10Recently, angioblasts have been identified in the adult as well.11 The endoderm is the initial stimulus for angioblast formation.8,9 Within the embryo proper, the first angioblasts arise from the lateral mesodermal plate and cardiac crescent8,9; some cells migrate into the forming brain, whereas others assemble into the endocardium of the early heart tube. Other angioblasts form a plexus of ECs at the base of the primitive heart tube that assemble into the vitelline vessels, allowing blood cells from the yolk sac to circulate within the body of the embryo.8 The vasculature of the viscera is formed from ECs that differentiate directly from the surrounding mesenchyme incorporated into the angiogenic extensions of invading vessels.12 For example, as the airway of the developing lung expands, endodermally derived cells of the branching airway provide directional queues for advancing branchial arteries and induce formation of angioblasts that become part of the pulmonary vasculature.13 Vasculogenic activity of early organ rudiments was shown through engraftment experiments using pieces of early quail lung and chick embryos. These experiments lead to the hypothesis that endoderm, but not ectoderm, induces vasculogenesis but that both endoderm and ectoderm can support angiogenesis14; recent experiments suggest additional sources of endothelial growth factors may exist.15 

Genetic programs regulating EC differentiation and early vascular development.

The best insight into molecular events required to initiate and maintain vascular development has come from detailed analyses of mouse embryos in which the genes for specific polypeptide growth factors or their transmembrane receptor tyrosine kinases (RTKs) have been inactivated. Such experiments show that initiation of vascular development requires both basic fibroblast growth factor (βFGF) and vascular EC growth factors (VEGF; see Beck and D'Amore16for a detailed discussion of growth factors and vascular development). Three alternatively spliced isoforms of VEGFs, members of the platelet-derived growth factor family (VEGF, VEGF-B, and VEGF-C17), interact with specific tyrosine kinase receptors. The growth factor-receptor interactions include VEGFR-1 (also known as flt-1 or fms-like tyrosine kinase-1) with VEGF and a related placenta growth factor (PlGF); VEGFR-2 (known alternatively as flk-1, fetal liver kinase-1; or Kdr, kinase-inserted domain containing receptor) with both VEGF and VEGF-C; and VEGFR-3 (originally designated flk-4) with VEGF-C. All VEGFs stimulate receptor autophosphorylation and EC replication and migration. The crucial role of this ligand in early vasculogenesis is demonstrated by the fact that loss of the VEGF gene results in embryonic death. Subsequent assembly of ECs into vessels requires activation of VEGFR-1 on the surface of the newly differentiated cells.18,19 The decision for a vessel to become a vein or an artery appears to be under the control of yet another growth factor, VEGFR-3,20 that is expressed later in development only on ECs that will become veins or lymphatic vessels.

Expansion of the vascular tree, continued endocardial and ventricular development, and formation of the vascular wall is controlled by two members of a second family of RTKs and their ligands,19tie-1 (tyrosine kinase with Ig and epidermal growth factor homology domains) and tie-2 or tek (tunica interna EC kinase). Two ligands, termed angiopoietin-1 and angiopoietin-2,21,22 are specific for tie-2 and are synthesized by cells surrounding the developing vessels. Ligand binding results in autophosphorylation of tyrosine residues in the intracellular domain of tie-2, but does not lead to EC replication or tube formation, as is the case for other endothelium-associated receptor-ligand interactions. Interestingly, angiopoietin-2 appears to function as an antagonist for angiopoietin-1, blocking its binding to tie-2. Targeted mutations of the genes for either tie-2 or angiopoietin-1 result in embryos with abnormal hearts and vessels with poorly formed walls.23 This has led to the suggestion that angiopoietin-1 acts via its receptor on ECs to stimulate the production of growth factors that, in turn, stimulate the differentiation of surrounding mesenchyme into pericytes or smooth muscle cells required for vessel wall formation.24 This is consistent with the phenotype of a Tie-2 mutation in humans that leads to smooth muscle deficiencies around small vessels and microaneurysms.25 These observations suggest that carefully regulated activity of tie-1 and 2 is required for continued vascular branching and vessel remodeling. Thus, the assembly of the early vascular tree depends on the programmed expression of at least two sets of RTKs and their ligands, one set for EC differentiation and initiation of vessel formation and the other for subsequent branching, establishment of capillary beds, and vessel wall formation.

Although genetic manipulations now possible in the mouse have provided important insights, there is much that we do not know about vascular development. Random genetic mutations introduced into the zebrafish have also generated many surprising and fascinating cardiovascular anomalies.6,26 For example, zebrafish can be induced to develop with hearts that do not contain an endocardium although the remainder of the vascular system appears functional. Thus, it is likely that a combination of genetics and developmental biology, unencumbered by previous assumptions, will continue to show new genes and suggest new paradigms that will advance our understanding of vascular development.

Extracellular matrix and matrix adhesion receptors in vascular development.

The ability of ECs to form capillary-like tubes is regulated not only by specific cytokine/receptor combinations, but also by the extracellular matrix. For example, human umbilical vein ECs (HUVECs) exposed to transforming growth factor-β (TGF-β) grow as a rapidly dividing monolayer if cultured on a flat surface coated with type I collagen.27 However, under similar TGF-β exposure, but within a type I collagen gel, ECs spontaneously organize into capillary-like tubules and continue to divide. Several kinds of molecules on the EC surface act together to mediate cell-extracellular matrix (ECM) interactions, including proteoglycans as well as proteins. The best studied family of receptors that mediate cell-matrix interactions is the integrins,28 which serve both a tethering and an information transfer function. Integrin-ligand binding triggers cytoskeletal organization at specific sites on the surface membrane to facilitate cell movement or maintain tissue stability. Binding also activates intercellular pathways that can result in either cell replication or programmed cell death.29 

Cells express more than one integrin and the combination of integrins expressed during embryonic development is constantly changing, suggesting that specific combinations are required as development proceeds. Experimental results in the developing mouse embryo suggest that functional compensation by integrins can occur during embryogenesis. It is also possible that receptors required for angiogenesis in early development may differ from those required for collateral vessel formation or tumor angiogenesis or that both gene inactivation and the introduction of inhibitory agents have unknown secondary effects. Similarly, mouse embryos continue to develop normally when genes for certain ECM components have been inactivated, whereas inactivation of other genes, such as the abundant fibronectin, results in early embryonic lethality.30 Similarly, knocking out the gene for a fibronectin receptor subunit (α5β1) also results in a poorly developed heart and vascular system and early embryonic death.30Again, other integrins, most likely αvβ3, appear to compensate for the absence of α5β1 during preimplantation development and early gastrulation, but are unable to do so as development proceeds.

Endothelial cell-cell interactions and vessel formation.

Angioblasts and ECs must contact like cells for vessels to sprout and lengthen. Such cell-cell adhesion is mediated by a distinct series of cell surface receptors that includes platelet EC adhesion molecule (PECAM-1),31 a member of the Ig superfamily, and vascular endothelial (VE)-cadherin.8 ECs express two isoforms of PECAM that mediate cell adhesion that differ in their requirement for divalent cations and sulfated proteoglycans. VE-cadherin, also known as cadherin-5, found almost exclusively on ECs, promotes cell-cell adhesion by a calcium-dependent homotypic mechanism,32 ie, in the presence of calcium, VE cadherin from one cell binds to the VE-cadherin expressed on an adjacent cell. As the vessel matures, more classic junctional complexes, such as tight junctions and gap junctions, form depending on the function of the particular vascular bed. Thus, during vascular development, junction formation initially involves rather weak adhesion complexes, likely required for cell-cell recognition, that facilitate the assembly of additional junctional complexes. However, the factors that determine the organ-specific nature of junction formation remain unknown.

Proteinases and vascular development.

Vascular development may be regulated by some of the same factors that are involved in the control of blood clotting and capillary formation. EC movement through the ECM is tightly regulated and requires integrin mediated cell-matrix adhesion complex formation and subsequent disassembly. This involves repetitive cycles of reversible integrin/matrix binding, assembly, and disassembly of cytoskeletal elements as well as matrix degradation restricted to the advancing edge of the moving cell.33 This is accomplished through the organization of active molecular complexes that approximate integrins and integrin ligands with matrix metalloproteinases and plasminogen activators with their respective substrates and inhibitors at sites of cell-matrix interaction.34 It has long been hypothesized that, when ECs are exposed to angiogenic stimuli, plasminogen activation is initiated through the binding of plasminogen and urokinase to their receptors. This leads to the formation of plasmin that activates prometalloproteases, degrades noncollagenous components of the matrix, provides a path for cell migration, and releases peptides that can promote or inhibit continued angiogenesis.35 Two peptides, angiostatin and endostatin,36,37 are currently being evaluated for clinical use in reversing tumor angiogenesis.36 However, much of the dogma regarding the role of plasmin in angiogenesis may have to be revisited in light of recent data indicating that transgenic animals with targeted disruption in the genes for tissue-type plasminogen activator (t-PA), urokinase (u-PA), the urokinase receptor (u-PAR), and plasminogen appear to develop a normal vasculature in the absence of trauma or other stressors (see below).

Perspective.

The induction of embryonic angioblasts to differentiate into ECs, organize into a vascular network, and subsequently populate the specialized vascular bed of an organ results from a complex genetic program, the details of which are only now emerging. This program is not only sensitive to the composition and structure of the ECM but is influenced by cell-cell contact as well as angiogenic and angiostatic growth factors and peptides generated by vascular expansion itself. Many of these same events are recapitulated after injury or as part of an inflammatory response and, if allowed to proceed unchecked (eg, tumor angiogenesis and diabetic retinopathy), can have serious consequences for the organism. Insight into the molecular and genetic programs involved in vascular differentiation may suggest better approaches to minimize ischemic tissue damage, avoid tissue rejection, stimulate wound healing, and inhibit tumor growth. This section has focused on the development of the vascular endothelium; however, the formation of the basement membrane, the induction and differentiation of vascular smooth muscle cells, and the complex process of assembling the elastic lamina are all essential to vessel formation and are under separate regulation. Ultimately, prevention of devastating effects from congenital abnormalities and facilitation of normal vascular function will, in a large part, be influenced by our ability to manipulate molecular mechanisms involved in vascular development.

Many human vascular diseases are exquisitely restricted to specific types of vessels. For example, the contribution of platelets to the pathogenesis of arterial and venous thrombosis differs as does the susceptibility of these two types of vessels to atherosclerosis. It is also common for vasculitis to show marked predilection for specific arteries, veins, or capillaries or for certain organs. Tumor cells may show similar predilection to metastasize through particular vascular beds.38 Even when systemic risk factors are clearly evident, such as is the case with inherited disorders of lipoprotein metabolism or proteins that control coagulation, there is marked regional variation in disease expression. Furthermore, clinical events such as thromboses are generally episodic and often localized to single vessels. The basis for variation is poorly understood, but may lie, in part, in the heterogeneity of ECs themselves (see Augustin et al2 and McCarthy et al39 for reviews).

To date, appreciation of EC function has been largely based on the behavior of cultured umbilical vein ECs (HUVECs). Indeed, it is remarkable that so many concepts in vascular biology have been predicated on the repertoire of umbilical ECs studied under such potentially unphysiologic in vitro conditions; this is especially true considering their derivation from a type of vessel that rarely, if ever, is affected by the most common human vascular disorders. More recently, there has been greater appreciation that EC heterogeneity may contribute both to the maintenance of adaptive processes and to the development of disorders restricted to specific vascular beds.

EC heterogeneity among and within tissues.

Variation in the appearance of capillary endothelium from different vascular sites has long been recognized and appears well suited to postulated differences in function (Fig 2). For example, the brain and retina are lined by continuous ECs connected by tight junctions that help to maintain the blood-brain barrier; the liver, spleen, and bone marrow sinusoids are lined by discontinuous ECs that allow cellular trafficking between intercellular gaps; while the intestinal villi, endocrine glands, and kidneys are lined by fenestrated ECs that facilitate selective permeability required for efficient absorption, secretion, and filtering (see Dejana32 for review). ECs from diverse tissues are also heterogeneous with respect to their surface phenotype and protein expression. For example, von Willebrand factor (vWF), used commonly as a marker for ECs, is not expressed uniformly on cells from all types of vessels,40,41 the expression of tissue type plasminogen activator is limited in vivo to approximately 3% of vascular ECs,42 and the constitutive expression of u-PA is reportedly confined to renal ECs,43,44 which are also uniquely susceptible to injury by verotoxin.45Microvascular ECs also differ in their susceptibility to undergo apoptosis induced by plasma from patients with thrombotic thrombocytopenic purpura.46 The induction of tissue factor after infusion of cytokines or endotoxin is similarly restricted to specific vessels,47 among many other examples of heterogeneity at the level of protein expression.

Fig. 2.

EC heterogeneity. (A) Electron micrograph showing the junction between two capillary ECs in a guinea pig pancreas (micrographs reprinted with permission from R.F. Bolender, The Journal of Cell Biology, 1974, vol. 61, p. 269). (B) Electron micrograph demonstrating the diversity of ECs from two types of capillaries: (1) vesicular invaginations (arrow) on both luminal and abluminal plasma membrane of a muscle capillary EC; (2) fenestrated capillary from the lamina propria of the colon with thin diaphragms (arrow) covering the plasma membrane pores (micrographs reprinted with permission from E. Weihe, Textbook of Histology, (ed 12), 1994, p. 391, courtesy of Chapman and Hall).

Fig. 2.

EC heterogeneity. (A) Electron micrograph showing the junction between two capillary ECs in a guinea pig pancreas (micrographs reprinted with permission from R.F. Bolender, The Journal of Cell Biology, 1974, vol. 61, p. 269). (B) Electron micrograph demonstrating the diversity of ECs from two types of capillaries: (1) vesicular invaginations (arrow) on both luminal and abluminal plasma membrane of a muscle capillary EC; (2) fenestrated capillary from the lamina propria of the colon with thin diaphragms (arrow) covering the plasma membrane pores (micrographs reprinted with permission from E. Weihe, Textbook of Histology, (ed 12), 1994, p. 391, courtesy of Chapman and Hall).

Close modal

One of the clearest examples of EC heterogeneity lies in the expression of homing receptors involved in cell trafficking. In the mouse, Lu-ECAM-1 (lung-specific EC adhesion molecule) is exclusively expressed by pulmonary postcapillary ECs and some splenic venules,48whereas Mad-CAM-1 (mucosal addressin cell adhesion molecule-1) is expressed primarily on high endothelial venules in Peyer's patches of the small intestine.49 Microvascular ECs derived from the bone marrow show an affinity for binding megakaryocytes and CD34+ progenitor cells and constitutively secrete hematopoietic stimulating factors such as Kit-ligand, granulocyte colony-stimulating factor, granulocyte-macrophage colony-stimulating factor, and interleukin-6 (IL-6), which help control trafficking, proliferation, and hematopoietic lineage-specific differentiation.50 Tumor cells may show clear preferential adhesion to the endothelium of specific organs paralleling their in vivo metastatic propensities (see McCarthy et al39 for review).

Microvascular ECs cultured from the brain, liver, and other organs each express distinct patterns of cell surface markers, protein transporters, and intracellular enzymes.51 These tissue-specific phenotypic differences can be maintained for some time under identical tissue culture conditions (eg, Grau et al52). Distinct subsets of ECs often exist within a single organ. In situ studies of adult human liver show two distinct sinusoidal EC phenotypes: hepatic periportal vessels express PECAM-1 and CD34, whereas sinusoidal intrahepatic ECs do not.53During the development of the human liver, ECs progress from a phenotype closely resembling adult hematopoietic sinusoidal bone marrow ECs, which supports fetal intrahepatic hematapoiesis, to one resembling adult hepatic sinusoidal ECs, including the expression of the T-lymphocyte marker CD4.54 

Environmental and genetic regulation of EC phenotype.

There is extensive evidence to indicate that heterogeneity develops in part as a result of variation in exposure of EC to environmental stimuli, some of which act only over short distances or even require cell-cell contact to effect change. Numerous exogenous factors affect EC phenotype, including mechanical forces, soluble growth promoters and inhibitors, cytokines, plasma lipids and proteins [eg, thrombin, plasmin, antibodies, Lp(a), etc], and contact with circulating and tissue-based cells (eg, smooth muscle cells and pericytes) and with the ECM, microbes, and their soluble products.

There are numerous examples of how the microenvironment can regulate the endothelial phenotype, a phenomenon that has been referred to as transdifferentiation.2 For example, aortic ECs cultured on extracellular matrix derived from the lung are induced to express Lu-ECAM-1,48 whereas the cells develop fenestrae when cultured on matrix derived from kidney-derived MDCK cells.55 Transplantation studies in the chick-quail system illustrate that ECs can take on the characteristics of the tissue into which they are transplanted in vivo,56 whereas other examples show that ECs acquire a different phenotype ex vivo.57 Studies in transgenic mice expressing the Lac Z reporter gene under control of 2,182 bp of the 5′ flanking sequence and the first exon and intron of the vWF gene suggest that expression is regulated by signals derived from the local microenvironment that influence pathways specific for particular vascular beds.58 

EC heterogeneity can thus arise as a consequence of local concentrations of exogenous effectors or due to intrinsic variations in responsiveness (reviewed in McCarthy et al39). ECs grown on extracts of basement membrane from different organs have been observed to develop preferential adhesivity for tumor cells prone to metastasize to that organ.59 ECs derived from saphenous vein have been reported to synthesize less prostaglandin I2(PGI2) than those from the internal mammary artery, a finding that may contribute to the rapidity with which pathogenic changes may develop in venous bypass grafts placed under arterial pressure.60 Possible genetic bases for EC diversity have only recently been considered and have not been studied in depth. Microvascular and macrovascular ECs differ in the fastidiousness of their growth, propsensity to form capillary-like structures, synthesis of PGI2, and expression of adhesion receptors for lymphocytes, among other properties (see Ades et al61 for review). Tissue-specific transcription factors or signal transduction molecules responsible for activating and/or de-repressing transcription apparati in a tissue-specific manner are only beginning to be understood. Identification of these control factors will be important in the design of vectors that will enable expression of EC proteins in a tissue-specific manner.

The effects of cell culture.

Only in the past few years has the technology become available that permits in situ study of EC behavior. These studies indicate that the constitutive phenotype of ECs is unstable and their behavior can change rapidly once explanted. Commonly used culture conditions may activate or otherwise alter the endothelial phenoytpe (eg, Grant et al62). There is, as yet, no model for generating the resting EC in vitro. Thus, all the information described in subsequent sections should be considered in the context of the cell source as well as the ex vivo culture conditions, including passaging, the presence/omission of shear forces, and factors released into blood that alter the behavior of the endothelium from that which occurs in healthy blood vessels in vivo. Thus, much may be gained in the future by a more critical consideration of EC heterogeneity both in terms of understanding homeostasis and vascular pathology, as well as in targeting the delivery of gene therapy, antithrombotic agents, and antitumor agents to an anatomically or functionally distinct endothelial region.63 

The endothelium not only provides a structural barrier between the circulation and surrounding tissue, but ECs also secrete mediators that influence vascular hemodynamics in the physiologic state (Table 1). ECs contribute to the regulation of blood pressure and blood flow by releasing vasodilators such as nitric oxide (NO) and prostacyclin (PGI2), as well as vasoconstrictors, including endothelin (ET) and platelet-activating factor (PAF). These chemically diverse compounds are not stored in intracellular granules. Rather, their major biologic effects are regulated by localization of specific receptors on vascular cells, through their rapid metabolism, or at the level of gene transcription. NO is constitutively secreted by ECs, but its production is modulated by a number of exogenous chemical and physical stimuli, whereas the other known mediators (PGI2, ET, and PAF) are synthesized primarily in response to changes in the external environment.

Table 1.

Vasoregulatory Substances Synthesized by the Endothelium

Substance Principal Effect Other Effects SecretionCompound Precursor Compound
NO (nitric oxide) Vasodilatation  Maintains basal tone of vessels; inhibits leukocyte adhesion; inhibits platelet adhesion, activation, secretion, and aggregation; promotes platelet disaggregation; inhibits smooth muscle cell migration and proliferation  Paracrine/Constitutive and induced by thrombin, ADP, bradykinin, substance P, muscarinic agonists, shear stress, cyclic strain, cytokines  Heterodiatomic free radical  L-arginine  
  
PGI2 (prostacyclin) Vasodilatation  Retards platelet aggregation and deposition Paracrine/Induced at sites of vascular perturbation Eicosanoid  Arachidonic acid  
  
PAF (plateletactivating factor)  Vasoconstriction  Promotes leukocyte adhesion at cell surface  Juxtacrine/Induced Phospholipid  Arachidonic acid  
  
ET-1 (endothelin-1)  Vasoconstriction  Mitogen for smooth muscle cells; modulates effect of numerous compounds  Paracrine/Induced by hypoxia, shear stress, and ischemia  21 Amino acid peptide Preproendothelin-1 (203 amino acids) 
Substance Principal Effect Other Effects SecretionCompound Precursor Compound
NO (nitric oxide) Vasodilatation  Maintains basal tone of vessels; inhibits leukocyte adhesion; inhibits platelet adhesion, activation, secretion, and aggregation; promotes platelet disaggregation; inhibits smooth muscle cell migration and proliferation  Paracrine/Constitutive and induced by thrombin, ADP, bradykinin, substance P, muscarinic agonists, shear stress, cyclic strain, cytokines  Heterodiatomic free radical  L-arginine  
  
PGI2 (prostacyclin) Vasodilatation  Retards platelet aggregation and deposition Paracrine/Induced at sites of vascular perturbation Eicosanoid  Arachidonic acid  
  
PAF (plateletactivating factor)  Vasoconstriction  Promotes leukocyte adhesion at cell surface  Juxtacrine/Induced Phospholipid  Arachidonic acid  
  
ET-1 (endothelin-1)  Vasoconstriction  Mitogen for smooth muscle cells; modulates effect of numerous compounds  Paracrine/Induced by hypoxia, shear stress, and ischemia  21 Amino acid peptide Preproendothelin-1 (203 amino acids) 

Principal regulatory compounds synthesized by the endothelium, their effects on the vasculature and other processes, their mode of secretion, and the nature of their chemical composition and precursor compound.

NO.

ECs elaborate NO, a heterodiatomic free radical product generated through the oxidation of L-arginine to L-citrulline by NO synthases.64 One isoform, eNOS or the Nos3 gene product, is constitutively active in ECs but is stimulated further by receptor-dependent agonists that increase intracellular calcium and perturb plasma membrane phospholipid asymmetry.65Receptor-dependent agonists that stimulate eNOS include thrombin, adenosine 5'-diphosphate, bradykinin, substance P, and muscarinic agonists, in addition to shear stress66 and cyclic strain.67 The increase in eNOS activity evoked by shear stress contributes to the phenomenon of flow-mediated vasodilatation, an important autoregulatory mechanism by which blood flow increases in response to exercise.68 This is in part a result of shear-induced transcriptional activation due to the presence of a shear response consensus sequence, GAGACC, in the promoter of the Nos3 gene.69 In addition to eNOS, cytokines have been shown to stimulate bovine microvascular endothelium in culture to express an inducible isoform of NO synthase, iNOS, or the Nos2 gene product.70 EC-derived NO has several important effects on the vasculature. First, NO maintains basal tone by relaxing vascular smooth muscle cells71 through the binding of NO to the heme prosthetic group of guanylyl cyclase. Endothelial-derived NO also inhibits platelet adhesion, activation, secretion, and aggregation and promotes platelet disaggregation, in part through a cyclic GMP-dependent mechanism.72 PGI2, which does not affect platelet adhesion,73 acts synergistically with NO to inhibit other steps in the platelet activation cascade.74NO also inhibits expression of P-selectin on platelets and, by inhibiting the agonist-dependent increase in intraplatelet calcium,72 suppresses the calcium-sensitive conformational change in the heterodimeric integrin glycoprotein αIIbβ3 (GP IIb-IIIa) required for fibrinogen binding.75 Additionally, NO appears to promote platelet disaggregation indirectly by impairing the activity of phosphoinositide 3-kinase, which normally supports conformational changes in αIIbβIIIa, rendering its association with fibrinogen effectively irreversible.76 

In addition to these effects on the vasculature, endothelial-derived NO inhibits leukocyte adhesion to the endothelium77,78 and inhibits smooth muscle cell migration79 and proliferation.80 These latter effects serve to limit neointimal proliferation that occurs after vascular injury and, combined with its stimulatory effect on EC migration and proliferation,81 suggest that NO helps to sustain vascular reparative mechanisms.

ECs also produce a less well-characterized compound known as endothelium-derived hyperpolarizing factor (EDHF) that promotes vascular smooth muscle relaxation (see Garland et al82 for review). Muscarinic agonists stimulate ECs to release EDHF, causing a transient hyperpolarization of the cell membrane. It has been proposed that EDHF exerts its vascular effects by activating ATP-sensitive potassium channels, smooth muscle sodium-potassium ATPase, or both,83 but its role in vascular (patho)physiology requires further study.

ET.

Remarkably, ECs produce not only the potent vasodilator NO, but also synthesize endothelin-1 (ET),84 the most potent vasoconstrictor identified to date. Endothelins comprise a family of 21-amino acid peptides produced by many cell types.84 ET-1 is not stored in granules85 but is formed after transcription of the gene encoding preproendothelin-1, the inactive precursor of ET-1, after stimulation by hypoxia, shear stress, and ischemia. ET-1 released from ECs binds to the abundant G-protein–coupled ET-A receptor expressed on vascular smooth muscle cells, which results in an increased intracellular calcium concentration and, in turn, increases vascular smooth muscle cell tone.86 Of interest, this effect of ET-1 persists after the hormone dissociates from its receptor through longer-lived effects on intracellular calcium. NO shortens the duration of these effects by accelerating the restoration of intracellular calcium to basal levels.87 The interplay between ET-1 and ET-A receptors likely contributes to basal vascular tone as well. ET-1 potentiates the vasoconstrictor actions of catecholamines, which, in turn, potentiate the actions of ET-1. In states of endothelial dysfunction, such as atherosclerosis, in which concentrations of bioactive NO are reduced, the relatively unopposed actions of ET-1 promote vasoconstriction and smooth muscle proliferation.88 

Prostacyclin (PGI2) and PAF.

The contribution of ECs to the regulation of vasomotor tone is even more finely regulated as evidenced by the production of additional vasoactive compounds such as prostacyclin (PGI2) and PAF. Prostacyclin and PAF factor provide an interesting contrast. Both are intercellular signaling molecules synthesized by stimulated ECs in vitro and in vivo.89 Both are lipids: PGI2being an eicosanoid and PAF being a phospholipid.90,91 Neither is constitutively present in resting human ECs nor stored within the cell. The synthesis of each is induced rapidly by humoral and mechanical stimuli via discrete, regulated pathways.90,91 Once formed, PGI2 and PAF have relatively short half lives, one of several features that limits the magnitude of their signals and exerts control over their biologic activities.92,93 

However, a major difference between the two factors lies in the range over which they exert their effects: PAF acts in a juxtacrine fashion, whereas PGI2 acts as a paracrine signaling molecule. PAF, expressed on the surface of the endothelium, remains cell-associated even in the presence of physiologic concentrations of albumin or other acceptor molecules91 and binds to and activates its receptor on leukocytes,94 fulfilling critical criteria of a juxtacrine signaling molecule. Consistent with this notion, PAF synthesized by cultured human ECs acts in concert with P-selectin (see below) to promote leukocyte adhesion.94 

In contrast, PGI2 is rapidly released from ECs,95 although the export mechanism has not been precisely defined. Thus, PAF and PGI2 have spatially differentiated realms of signaling, even though both derive from a common precursor and are synthesized concurrently.91,96 This feature may contribute to differences in their actions at the endothelial interface with the blood: PAF is specialized to signal leukocytes at the cell surface, whereas PGI2 acts primarily in solution to retard platelet aggregation and deposition. Both PGI2 and PAF also elicit autocrine effects on ECs,91,92 which may be important in modulating angiogenesis and controlling the synthesis of EC-derived mediators.

PGI2 was the first endothelial-derived vascular smooth muscle relaxing factor to be identified. PGI2, which was generated locally, and PGI2 or its analogs, which were infused systemically, caused vasodilatation and altered regional blood flow.93 A receptor for PGI2, the IP receptor, is present on vascular smooth muscle as well as on platelets,97 consistent with early experimental observations, indicating that PGI2 acts principally to modulate the function of these two cell types.98 Although IP receptors are present in the arterial vascular wall, PGI2 is not constitutively produced and does not appear to regulate basal systemic vascular tone.99 Rather, PGI2 synthesis is induced at sites of vascular perturbation, where it may regulate vasoconstriction and platelet deposition.66 Because of its effects on blood flow and relevant cell-cell interactions, PGI2 may influence local inflammatory responses as well. An important recent advance has been the identification of prostaglandin H synthase-II (PHS-II), an inducible form of a key enzyme in PGI2 formation providing a mechanism by which the production of PGI2 and other eicosanoids can be sustained in chronic states of inflammation and vascular injury.

The receptor for PAF, the first receptor characterized at a molecular level that recognizes a biologically-active lipid, is a member of the serpentine G-protein–linked family (reviewed in Whatley et al91). Intravascular infusion of PAF causes either vasodilatation or vasoconstriction, depending on the concentration administered, the time, and the specific vascular bed studied.93 Some hemodynamic effects of PAF in vivo are indirect and depend on the generation of eicosanoids or leukotrienes or mediators derived from activated leukocytes or platelets and on cardiac effects.91-93 In shock and other in vivo pathologic states, PAF acts concomitantly or sequentially with other classes of mediators, including leukotrienes and tumor necrosis factor-α (TNF-α).90,91,100 As with PGI2, it is unlikely that PAF is a circulating regulator of blood pressure under basal conditions, despite early studies suggesting that PAF-like activity is released from kidneys.92 

A crucial physiologic function of the endothelium is to facilitate blood flow by providing an antithrombotic surface that inhibits platelet adhesion and clotting. However, when the endothelium is perturbed by physical forces or by specific chemical factors, the cells undergo programmatic biochemical changes that culminate in their transformation to a prothrombotic surface. A dynamic equilibrium exists between these two states, modulated both at the level of gene transcription and at the level of the intact cell, that often permits the injured endothelium to return to its unperturbed state once the procoagulant stimulus has dissipated (Table2; see Bombeli et al101 for review). Although the fibrin clots formed as a consequence of procoagulant activity may serve a protective organ function by limiting vascular damage induced by trauma, infection, and inflammation, the loss of anticoagulant activity may predispose to several common thrombotic disorders discussed in the sections that follow.

Table 2.

Regulation of Hemostasis and Thrombosis by the Endothelium

Antithrombotic Prothrombotic
Coagulation protein binding sites  Glycosaminoglycans/ATIII  Binding sites for: fibrin, FIX, IXa, X, Xa, FXII, kallikrein  
 TFPI  Tissue factor  
 Thrombomodulin  Thrombin receptor  
  Receptor for protein C/APC  
 
Products produced and/or stored by platelets PGI2 NO ADPase  vWF PAF Fibrinogen FV FXI  
 
Fibrinolytic factors  t-PA production u-PA expression  PAI-1, PAI-2 PAI-3 (protein C Inhibitor)  
 u-PAR  TAFI activation  
 Plasminogen binding sites  
 Annexin II  
 
Vasomotor factors  NO PGI2 TxA2 Endothelin-1 
Antithrombotic Prothrombotic
Coagulation protein binding sites  Glycosaminoglycans/ATIII  Binding sites for: fibrin, FIX, IXa, X, Xa, FXII, kallikrein  
 TFPI  Tissue factor  
 Thrombomodulin  Thrombin receptor  
  Receptor for protein C/APC  
 
Products produced and/or stored by platelets PGI2 NO ADPase  vWF PAF Fibrinogen FV FXI  
 
Fibrinolytic factors  t-PA production u-PA expression  PAI-1, PAI-2 PAI-3 (protein C Inhibitor)  
 u-PAR  TAFI activation  
 Plasminogen binding sites  
 Annexin II  
 
Vasomotor factors  NO PGI2 TxA2 Endothelin-1 

Outline of the antithrombotic and prothrombotic properties of ECs detailing the binding sites expressed by ECs, the factors stored and/or secreted by ECs that affect platelet function, factors produced by the EC that influence the fibrinolytic state of the vasculature, and vasomotor substances secreted by the EC.

Abbreviations: ATIII, antithrombin III; PGI2, prostacyclin; TFPI, tissue factor pathway inhibitor; APC, activated protein C; PAF, platelet-activating factor; t-PA, tissue plasminogen activator; u-PA, urokinase plasminogen activator; u-PAR, urokinase plasminogen activator receptor; PAI, plasminogen activator inhibitor; TAFI, thrombin activatable fibrinolysis inhibitor; TxA2, thromboxane A2.

Anticoagulant mechanisms.

Control of thrombin generation is a pivotal step in the balance between the natural antithrombotic and the induced procoagulant activities of the endothelium. Thrombin, a serine protease, serves diverse functions in coagulation, including the activation of platelets, several coagulation enzymes, and cofactors. Thrombin also stimulates procoagulant pathways on the ECs themselves. Therefore, it is not surprising that several highly regulated pathways have evolved to constrain the generation and activity of thrombin (see Rosenberg and Rosenberg102 for review), such that little enzyme activity is found in the plasma of healthy individuals.103 The matrix surrounding the endothelium contains heparan sulfate and related glycosaminoglycans (GAGs) that promote the activity of cell/matrix associated antithrombin III (AT-III)104; the subendothelium contains dermatan sulfate, which promotes the antithrombin activity of heparin cofactor II.105 ECs also prevent thrombin formation through the expression of tissue factor pathway inhibitor (TFPI), which binds to factor Xa within the tissue-factor/VIIa/Xa complex (see Broze106 for review). TFPI is released from its EC stores by heparin. TFPI and AT-III both contribute to physiologic hemostasis and can be depleted in acquired thrombotic states.107,108 

The endothelium also helps to contain thrombin activity through the expression of thrombomodulin (see Esmon and Fukudome109 for review). Binding of thrombin to TM facilitates the enzyme's ability to activate the anticoagulant protein C. In turn, the activity of activated protein C (APC) is enhanced by its cofactor protein S, which is synthesized by EC, among other cell types.110 ECs also express receptors for APC111 that regulate the activity of this pathway. APC, in turn, promotes the inactivation of activated factors V and VIII. Binding of thrombin to TM also dampens the enzyme's ability to activate platelets, factor V, factor XIII, and fibrinogen and promotes EC fibrinolytic activity (see below). TM also inhibits prothrombinase activity indirectly by binding factor Xa.112 Thrombin bound to TM is rapidly endocytosed and degraded.113 Various inflammatory cytokines downregulate TM gene transcription and accelerate TM internalization114,115while at the same time promote tissue factor expression (see below). Soluble TM is also shed into plasma and elevated plasma levels have been identified in various disorders associated with EC injury (see Cucurull and Gharavi116 and below).

Procoagulant mechanisms.

The pivotal step in transforming the EC membrane from an anticoagulant to a procoagulant surface is the induction of tissue factor (TF). TF dramatically accelerates factor VIIa-dependent activation of factors X and IX, so it is not surprising that TF is not expressed by unperturbed endothelium, at least in the adult organism.117Interruption of the gene for TF is associated with impaired vascular development and lethal embryonic bleeding,118-120 but the source and function of TF during development have not been elucidated. Synthesis of TF is induced in vitro by diverse agonists, including thrombin, endotoxin, several cytokines, shear, hypoxia, oxidized lipoproteins, and many other provocations (see Rapaport and Rao121 and Nemerson122 for reviews). Procoagulant activity is accelerated by exposure of anionic phospholipids that may occur as a consequence of apoptosis.108 TF is localized primarily beneath and between cultured ECs,123 although some evidence for expression on the cell surface has been presented.124 TF mRNA and protein levels decline despite continued exposure to agonists, a mechanism that may help contain the extent of fibrin formation. Cells in culture also shed microvesicles containing TF,125 and plasma levels of TF are elevated in patients with disseminated intravascular coagulation,126 although the cellular source has not been established. TF expression is rapidly induced after vascular injury,127 and TF is found associated with ECs within atherosclerotic plaque128,129 and in tumor-derived vessels.130 TF may also contribute to the regulation of angiogenesis and tumor metastases through mechanisms independent of coagulation.131,132 Yet, it has been difficult to demonstrate expression of TF by ECs in vivo even in response to potent provocations where expression was expected, another example of the dissociation between the behavior of these cells in culture and that seen in the whole organism.47 

Once ECs expressing TF are exposed to plasma, prothrombinase activity is generated and fibrin is formed on the surface of the cells.133 This implies that ECs express binding sites for factors IX, IXa, X, and Xa; thrombin; and fibrin.134 Yet, the identity and location of most of these binding sites is unknown, as is their role in either physiologic hemostasis or in thrombosis. Factor IX has recently been shown to bind type IV collagen in the EC matrix,135 although its cellular association site promoting assembly of the intrinsic FX activation complex has not been identified. Several candidate FX/Xa binding sites have been reported,136,137 whereas others may be induced as a result of exogenous stimuli.138 ECs also express receptors for proteins of the contact factor pathway,139 but their role in hemostasis is uncertain.

The most thoroughly characterized EC binding site for a coagulation protein is the thrombin receptor, also termed the protease-activated receptor-1 (PAR1). The thrombin receptor is a high-affinity G-protein–coupled protein140 that is activated when a fragment derived from the amino terminus of the protein, formed as a result of cleavage by thrombin, binds to the remaining cell-associated receptor fragment. Binding of thrombin leads to a wide array of changes in expression of prothrombotic and antithrombotic molecules by cultured ECs, including TF, PAI-1, NO, PAF, ET, and PGI2, among others (see Kanthou and Benzakour141 for review) and disruption of cell-cell contacts (see Garcia et al142 for review). Thrombin is also mitogenic for ECs, fibroblasts, and smooth muscle cells and is chemotactic for monocytes.

The seemingly normal phenotype of surviving adult mice with targeted disruptions in the thrombin receptor gene143 was unanticipated and raised questions about the physiologic role of this protein as well as other proteinase activated receptors expressed on ECs.144 The subsequent discovery of two additional protease-activated G-protein–coupled receptors, PAR-2 and PAR-3, helped to explain this observation.145 Both PAR-1 and PAR-2 are present on some human ECs, whereas PAR-1, but not PAR-2, is expressed on human platelets. PAR-3 is expressed by human bone marrow and mouse megakaryocytes, but its expression on ECs has not been established. Notable tissue- and species-specific differences in expression and cellular distribution of PARs have been described,145,146 making it difficult at present to relate the phenotype of the various murine knockouts to human physiology.

ECs also express several receptors for fibrin and specific fibrin degradation products,147 including a 130-kD glycoprotein,148 a tissue transglutaminase,149and the αvβ3 integrin, although evidence regarding their expression in vivo is only now emerging.150,151 Binding of fibrin promotes EC adhesion, spreading, proliferation, and migration; cell retraction; leukocyte adhesion; and inhibition of PGI2 synthesis. Cultured ECs also express glycoprotein Ib, which binds vWF secreted constitutively by ECs and, presumably, the ultralarge vWF multimers released from Weibel-Palade bodies in response to a number of agonists (see Wagner and Bonfanti152 for review). Expression of GPIbα by ECs is enhanced by TNF-α,153 but whether this glycoprotein participates in the physiologic or pathophysiologic binding of vWF in vivo requires additional study. The αvβ3integrin also binds vWF in the EC substratum. In vitro, conditions that decrease αvβ3 expression (TNF-α plus interferon-γ [IFN-γ] or arterial shear stress) increase GPIbα expression,154 suggesting that the EC state may affect the availability of adhesion receptors, although these finding require confirmation in vivo.

Undoubtedly, additional receptors for coagulation proteins with distinct functions will be characterized in the future. Genetic and acquired alterations in the structure, expression, and function of these EC receptors may contribute to hitherto unexplained hemorrhagic and thrombotic disorders. However, despite rather extensive study of EC procoagulant function in culture, the extent to which platelet adherence or fibrin formation actually occurs on the surface of the intact endothelium in vivo (as opposed to subendothelial matrix exposed to blood) remains unclear. Perhaps the best indirect evidence comes from animal models of Escherichia coli sepsis and cytokine infusion in which TF- and contact factor-dependent intravascular coagulation and multiorgan ischemic injury occurs in the absence of overt EC disruption.155 Nevertheless, the contribution of endothelium, platelets, monocytes, and other cell types in these models will require further study.

Experiments using cultured ECs have yielded a concept that the endothelial surface is profibrinolytic and thus helps maintain blood in its fluid state.156 However, experiments using animal models have shown this conceptually satisfying hypothesis does not accurately reflect the situation in vivo.42,157 It has also become clear that the contribution of ECs to fibrinolysis varies with their metabolic status (ie, quiescent or activated), their vascular derivation, and the concentration of other hemostatically active molecules in the local plasma milieu.

Plasminogen activators.

Studies with ECs cultured from various tissues have led to the widely held inference that t-PA production and secretion is a property of all ECs.156 However, a few studies have gone largely unnoticed in which PA activity, demonstrated by fibrin zymography, was observed only in association with the adventitia and not with lumenal ECs.157 More recent studies using in situ hybridization and immunohistochemistry have demonstrated t-PA antigen and mRNA only in a distinct subset of quiescent microvascular ECs of both primates and mice.42,158 Hence, contrary to assumptions based on work with cultured ECs, t-PA is associated only with a distinct subpopulation of the microvasculature, even after provocation. In both murine brain and lung, the percentage of microvascular ECs producing t-PA increases markedly upon exposure to pertinent stimuli; however, in both cases, t-PA production remains an exclusive property of microvascular ECs.42 Hence, invoking local EC production of t-PA in large vessels as a mechanism of maintaining blood flow may not adequately describe the in vivo situation.

t-PA production by cultured ECs is regulated by a variety of external stimuli at the level of gene transcription and cellular release.159 Measurements of plasma t-PA levels suggest such regulated production/secretion occurs in vivo as well.160Intracellular signaling pathways operative in stimulated t-PA release have been described in vitro.161 The mechanisms that control EC t-PA production in vivo are less well understood, but clearly such regulation occurs.162 Humans exhibit higher plasma levels of t-PA after exercise or venous compression, but the cellular source of this increase has not been established.161 

The other mammalian PA, u-PA, appears not to be produced by most quiescent ECs.43 Rather, it is expressed by ECs involved in wound repair or angiogenesis,163 consistent with the hypothesized importance of u-PA in cell migration and tissue remodeling. Yet, u-PA is obviously important to vascular homeostasis, because mice genetically deficient in u-PA develop inflammation induced thrombi164 and manifest thrombotic tissue injury in response to lipopolysaccharide (LPS).165However, the extrarenal source of u-PA in physiologic states has not been established.

PA receptors.

The presence of EC receptors for t-PA has been reported by several groups.166-171 Binding of t-PA to ECs has been reported to promote its fibrinolytic activity166,172 and to stimulate cell proliferation.173 Recently, one such t-PA binding site has been identified as annexin II, which is expressed on ECs174 and binds t-PA in a specific and saturable manner in vitro.172 However, the expression of annexin II on the endothelium in vivo has not yet been demonstrated.

The u-PA receptor (u-PAR) expressed by ECs appears identical to that expressed on other cell types.175,176 u-PAR is a three-domain protein linked to cell surfaces by a glycerophosphatidyl inositol anchor. Single-chain u-PA (the form found in plasma177) bound to cells via u-PAR exhibits increased plasminogen activating efficiency175,178 and is relatively protected from inhibition by PAI-1 and PAI-2.179,180 u-PAR may be expressed primarily on the surface of migrating ECs participating in angiogenesis, rather than on quiescent ECs lining normal vessels.181 Mice genetically lacking u-PAR develop normally and do not exhibit spontaneous vascular occlusion.182 Hence, u-PAR has yet to be shown to participate in maintaining physiologic blood fluidity, although it may be important in vascular repair.

Cells express diverse binding sites for plasminogen, among which are proteins that exhibit a carboxyterminal lysine (see Plow et al35 for review). Plasminogen binds to ECs in vitro with an affinity that would predict receptor occupancy at physiologic plasma concentrations.183 Cell-associated plasmin may be relatively protected from inhibition by α2-plasmin inhibitor.184 However, the exact identity of these EC plasminogen binding sites remains uncertain and their expression in vivo has not been established.185 Lp(a) competes for the binding of plasminogen to ECs,186 which may contribute to the prothrombotic effects of this lipoprotein.187 

Plasminogen activator inhibitors (PAIs).

ECs in culture produce abundant PAI-1 that is associated primarily with its extracellular matrix, resulting in stabilization of its activity.188 PAI-1 synthesis is stimulated by numerous agents, including thrombin, endotoxin, various cytokines, Lp(a), and oxidized LDL, among others.189 Yet, experiments in mice have shown that liver is the major source of plasma PAI-1 and that quiescent EC express little or no inhibitor.190 However, after exposure to inflammatory stimuli, ECs in virtually every tissue express PAI-1.190 PAI-2 is found normally in plasma only during pregnancy191 and is not synthesized by ECs to an appreciable extent. However, multiply passaged ECs express PAI-2 in response to some agonists that may point to a local effect in select settings.192 PAI-3 (also known as the protein C inhibitor) has a much lower affinity for u-PA and t-PA than does PAI-1, but it is present in plasma at much higher concentrations.193Production of PAI-3 by ECs has not been reported, but PAI-3 antigen can bind to heparan sulfate proteoglycan on the lumenal surface of ECs, thereby increasing its activity.194 

Thrombomodulin.

Binding of thrombin to thrombomodulin (see “The Role of the Endothelium in Coagulation” above) accelerates its capacity to activate a protein known as thrombin-activatable fibrinolysis inhibitor (TAFI).195 TAFI is a procarboxypeptidase-B–like molecule that, when activated, cleaves basic carboxyterminal residues within fibrin and other proteins. This results in the loss of plasminogen/plasmin and t-PA binding sites on fibrin such that fibrinolysis is retarded.195 Thus, through the regulated expression of thrombomodulin, ECs serve as potent templates to decrease the rate of intravascular fibrinolysis.

Although a simple balance between profibrinolytic (PAs) and antifibrinolytic (PAIs) pathways seemed an attractive mechanism to explain the clinical experience that unperturbed endothelium helps maintain blood fluidity, more recent in vivo data have shown that the mechanism may not be quite so straightforward. Indeed, ECs seem to express more antifibrinolytic than profibrinolytic activity in many settings studied to date. Clearly, more work will be required to clarify the contribution of quiescent and activated ECs to fibrinolysis.

Summary.

The first part of this two part series has focused on the development of the vasculature and the physiological functions of the endothelium as a gate-keeper regulating blood flow and hemostasis. Current insights into the generally unappreciated heterogeneity of endothelium from different vascular sites have been noted as potential discrepancies between the quiescent state of the endothelium in vivo and the behavior of these cells in culture. The second part of this review will concentrate on the mechanism by which the endothelium contributes to cell trafficking and the impact of endothelial injury on the development of several common human vascular disorders.

The endothelium, positioned at the interface between blood and tissue, is equipped to respond quickly to local changes in biological needs caused by trauma or inflammation. In the first part of this review, the capacity of the endothelium to move rapidly between an antithrombotic and prothrombotic state was discussed. In this second part, the mechanisms by which the endothelium regulates the trafficking of the cellular elements of the blood will be considered first, after which the impact of EC dysfunction on the pathogenesis of several common vascular disorders will be reviewed.

In addition to the above-mentioned contribution of the endothelium to regulating blood coagulation, ECs also express cell surface-molecules that orchestrate the trafficking of circulating blood cells. These cell-associated molecules help direct the migration of leukocytes into specific organs under physiologic conditions and accelerate migration towards sites of inflammation, eg, in response to IL-6196or IL-8,197 among many others. Recently, these pathways have also been implicated in the adhesion of platelets and erythrocytes in several common disorders associated with vascular occlusion.

Interactions of platelets and leukocytes with the vessel wall.

Flowing leukocytes and platelets may adhere to specific regions of the endothelium, to exposed subendothelial components, or to each other during the process of immune surveillance as well as in response to tissue injury or infection. These multicellular interactions are essential precursors of physiologic inflammation and hemostasis. Conversely, uncontrolled adhesion of leukocytes and platelets contributes to inflammatory and thrombotic disorders. Under shear forces, both platelets and leukocytes interact with vessel surfaces through a multistep process that includes (1) initial formation of usually reversible attachments; (2) activation of the attached cells; (3) development of stronger, shear-resistant adhesion; and (4) spreading, emigration, and other sequelae (Fig 3).

Fig. 3.

Physiologic interaction of leukocytes with the endothelium. Leukocyte adhesion and transmigration occurs during inflammation, usually at the postcapillary venules where shear stress is lowest.

Fig. 3.

Physiologic interaction of leukocytes with the endothelium. Leukocyte adhesion and transmigration occurs during inflammation, usually at the postcapillary venules where shear stress is lowest.

Close modal
Platelet adhesion during hemostasis.

Circulating platelets normally do not interact with the EC surface (see Schafer198 for review), in part due to the release of PGI2, the release of NO, and the recently described expression of an Ecto-ADPase (CD39).199 However, during hemorrhage, platelets adhere avidly to exposed subendothelial components, where they are rapidly activated. Circulating platelets interact with the adherent platelets, producing a hemostatic plug that promotes thrombin generation and development of a stable fibrin clot. Platelets adhere particularly efficiently to the subendothelium under high shear stress, accounting for the greater number of platelets in fibrin clots within arteries compared with those within veins.200 

Under the influence of arterial shear stress, unactivated platelets attach first to the subendothelium through interactions of the platelet glycoprotein (GP) Ib-IX-V complex with immobilized vWF, a large, multimeric protein with binding sites for several other molecules, including subendothelial collagen.201 The GPIb-IX-V complex consists of four proteins, each with one or more leucine-rich repeats: the disulfide-linked GPIbα and GPIbβ and the noncovalently associated GPIX and GPV. The binding site for vWF is located on GPIbα, between the amino-terminal leucine-rich repeats and the membrane-proximal O-glycan–rich domain. The region includes clustered tyrosines that must be sulfated for GPIbα to bind vWF.202GPIbα binds weakly to plasma vWF, but with high activity to immobilized vWF under conditions of high shear stress, which may favor binding by altering the conformation of GPIbα and/or vWF.200,203 Flowing platelets attach transiently to vWF, resulting in continuous movement of the cells along the surface.204 Such cellular interactions require very fast molecular rates for attachment and detachment; the fast dissociation rates are not significantly accelerated by shear stresses for detachment. Under the lower shear stresses found in veins, unactivated platelets use the integrin αIIbβ3 to attach to and immediately arrest on immobilized fibrinogen.204Under low shear conditions, platelets may also use integrins or other molecules to attach to subendothelial matrix components, such as fibronectin, laminin, and thrombospondin.205 

Once platelets adhere to either vWF or fibrinogen, they are activated by secreted products such as ADP or epinephrine or by surface molecules, such as collagen, that cross-link the integrin α2β1 and other platelet receptors. The activated platelets spread and adhere more avidly to the subendothelial surface, principally through binding of activated αIIbβ3 to fibrinogen, which recruits additional platelets into aggregates206; platelet αIIbβ3 also binds to a distinct site on vWF.201 Shear-resistant adhesion may be further enhanced by interactions of other integrins or receptors with laminin, fibronectin, and thrombospondin.200 As thrombin is generated, converting bound fibrinogen to fibrin, the aggregated platelets contract to strengthen the clot. Signaling through adhesion receptors, particularly integrins, may regulate the cytoskeletal-protein redistributions required for clot retraction.206 Increased bleeding is observed in patients with inherited defects in molecules that mediate platelet adhesion such as Bernard-Soulier disease (absence/dysfunction of GPIb-IX-V) and Glanzmann's thrombasthenia (absence/dysfunction of αIIbβ3), confirming their physiological function.203 

Leukocyte adhesion during inflammation.

During inflammation, leukocytes tether to and roll on the EC surface. The cells then arrest, spread, and finally emigrate between ECs to reach the underlying tissues. Unlike platelets, which typically attach to the subendothelium of arteries under high shear stresses, leukocytes usually attach to the ECs, where shear stresses are lowest, in the lining postcapillary venules.

In most circumstances, interactions with selectins, transmembrane glycoproteins that recognize cell-surface carbohydrate ligands found on leukocytes, initiate and mediate tethering and rolling of leukocytes on the EC surface.207 Selectins constitute a family of three known molecules, each of which has an amino-terminal Ca2+-dependent lectin domain, an EGF domain, a series of short consensus repeats, a transmembrane domain, and a cytoplasmic tail. L-selectin is expressed on most leukocytes and binds to ligands constitutively expressed on high endothelial venules of lymphoid tissues, to ligands induced on endothelium at sites of inflammation, and to ligands exposed on other leukocytes. E-selectin is expressed on activated ECs and leukocytes. P-selectin is rapidly redistributed from secretory granules to the surface of platelets and ECs stimulated with thrombin or other secretagogues. Like E-selectin, P-selectin binds to ligands on leukocytes. Leukocytes adherent to the endothelium can make contact with flowing leukocytes through the L-selectin molecule, resulting in amplification of leukocyte recruitment to sites of inflammation.208 At sites of hemorrhage, leukocytes tether to and roll on adherent platelets.209 Monocytes recruited in this manner may augment fibrin generation, perhaps by elaborating tissue factor after their activation.210 Selectin ligands expressed on high endothelial venules also mediate rolling of activated platelets and enhance accumulation of lymphocytes in lymph nodes.211 Thus, selectins initiate inflammatory, immune, and hemostatic responses by promoting transient multicellular interactions under conditions of shear stress.

The selectins bind weakly to sialylated and fucosylated oligosaccharides, such as sialyl Lewis x, a terminal component of glycans attached to many proteins and lipids on most leukocytes and some ECs. Strikingly, the selectins bind with higher affinity to only a few sialylated and fucosylated glycoproteins on target cells.207 E-selectin binds preferentially to ESL-1, a protein with at most five N-glycans and no described O- glycans. L-selectin and P-selectin bind preferentially to sialomucins whose recognition requires sulfation as well as sialylation and fucosylation. The sulfate esters are attached to O-glycans on GlyCAM-1, a ligand for L-selectin secreted by high endothelial venules.212 In contrast, the sulfate esters are attached to tyrosines near the amino terminus of PSGL-1, a ligand for selectins on leukocytes.213 Construction of some glycans may be restricted to specific sites on the polypeptide backbone of only a few proteins.214 Of the described glycoprotein ligands for selectins, only PSGL-1, a ligand for selectins on leukocytes, has been shown to mediate cell-cell interactions under shear conditions (reviewed in McEver and Cummings215). The α4β1 and α4β7integrins, which are expressed on mononuclear cells and eosinophils, but not on neutrophils, also mediate tethering and rolling and occasionally arrest the flow of leukocytes on ECs by binding to the Ig-ligands VCAM-1 and MAdCAM-1.216,217 Some lymphocytes use CD44 to roll on hyaluronate-bearing surfaces.218 

Under hydrodynamic flow, cell tethering and rolling requires bonds with sufficient mechanical strength between adhesion molecules and rapid rates of association and dissociation.219 Interestingly, attachment of leukocytes through selectins requires a threshold hydrodynamic shear force that may prevent leukocyte aggregation in regions of low flow.220 A higher shear threshold for L-selectin may reflect faster dissociation rates of L-selectin ligand bonds220 and/or adhesion-induced shedding of L-selectin from the cell surface.221 Because L-selectin, PSGL-1, and α4 integrins are concentrated on the tips of the leukocyte microvilli, the probability of rapid contact with PSGL-1 is increased and the repulsion is minimized between the charged glycocalyces of apposing cells.222 

The slow velocities of rolling leukocytes favor encounters with chemokines or lipid autacoids presented at or near the apical surface of the endothelium. These mediators transduce signals that cooperate with those produced by engagement of L-selectin or PSGL-1 to activate the leukocytes.223 This crucial activation event, coupled with the slow rolling velocities, enables the β2integrins on leukocytes to bind to Ig ligands such as ICAM-1 and ICAM-2 on the EC surfaces.224 Plasma fibrinogen also links leukocytes to the endothelium by binding simultaneously to αMβ2 and ICAM-1,225 two integrins on the vessel wall that provide shear-resistant attachments. Subsequently, leukocytes migrate between ECs into tissues by mechanisms that are not completely understood but are affected by gradients of chemokines with restricted specificities,49β1 and β2 integrins activation states, and homotypic interactions with the Ig-like receptor, PECAM-1.226 This may require disruption of homotypic interaction of cadherins at endothelial tight junctions.227 

Leukocyte recruitment to lymphoid tissues or inflammatory sites requires the coordinated expression of specific combinations of adhesion and signaling molecules. Diversity at each step of the multistep cascade ensures that the appropriate leukocytes accumulate for a restricted period in response to a specific challenge.49,224 Absence of P-selectin delays fatty streak formation in mice predisposed to developing atherosclerotic lesions.228 On the other hand, increased numbers of infections are observed in patients who are congenitally deficient in β2 integrins229 or in fucosylated ligands for selectins,230 confirming the physiologic significance of these molecules in immune and inflammatory responses. Increased susceptibility to infection combined with impaired leukocyte accumulation in mice rendered genetically deficient in selectins,231-234 in fucosyltransferases,235 in ICAM-1,236-239 or in α4integrins240 further support the overlapping functions of these molecules.

Endothelium in cell-mediated immunity (CMI).

CMI is defined as the protective set of immune reactions that can be adoptively transferred from a sensitized individual to an unimmunized host by a subset of T lymphocytes but not by antibodies. Vascular ECs may play two important roles in the evolution of CMI reactions: (1) antigen presentation to T cells (reviewed in Pober et al241) and (2) recruitment of inflammatory cells.49 Recall responses such as CMI reactions develop directly in peripheral tissues in which circulating memory T cells are activated by antigen presented on the surface of a resident cell population. This is in contrast to primary immunity that begins in the secondary lymphoid organs such as lymph node or spleen, where naive T cells encounter antigen on the surface of a specialized antigen-presenting cell. Once effector and memory cells develop in a secondary lymphoid organ, they may emigrate to the peripheral site via the blood stream, where reactivation by the antigenic stimulus is possible.

The two cell types that may present antigen to specific T cells in peripheral tissues are macrophages, resident in the tissues, or local microvascular ECs. In vitro, cultured human ECs from a variety of vascular beds constitutively express class I MHC molecules (used to present peptides derived from foreign proteins to CD8+ T cells). IFN-γ can induce ECs to express class II MHC molecules (used to present peptides derived from foreign proteins to CD4+ T cells). There is only limited information on antigen processing by cultured ECs, but indirect evidence (ie, the formation of functional peptide-MHC molecule complexes that can be recognized by cultured T-cell lines) suggests that EC are fully competent to perform this function. In vivo, microvascular ECs constitutively express both class I and class II MHC molecules, although the levels of both molecules can be increased further by cytokines (eg, IFN-γ).

How do ECs compare with tissue macrophages as antigen-presenting cells? This has been a difficult question to address experimentally in humans because the relevant cell populations (ie, ECs, macrophages, and T cells) are not readily isolated from one immunized individual. A commonly used indirect approach is to examine the response of T cells isolated from one donor to cultured ECs and monocytes isolated from a second donor. This allogeneic response of T cells is directed against complexes formed between peptides associated with allogeneic MHC molecules and results from a cross-reaction of T cells that are specific for a foreign peptide associated with a self MHC molecule; it is an excellent model of normal immunity that avoids the requirement for isolating ECs from immunized donors. ECs activate about one fifth as many allogeneic T cells (either CD8+ or CD4+) as monocytes, determined in limiting dilution analyses for production of IL-2. Some of this difference may arise from a larger number of different peptide-MHC molecule complexes displayed on freshly isolated monocytes compared with serially cultured ECs, but a critical contribution to this difference is that monocytes can activate both naive and memory T cells, whereas ECs can only activate memory T cells.242 It has recently been reported that ECs actually make naive T cells unresponsive to stimulation (ie, induce clonal anergy).243 The ability of ECs to activate memory but not naive T cells in vitro is consistent with a role in presentation of antigen as an initiating event in CMI, which is a memory T-cell response.

The differences in ability among cell types to activate resting naive or memory T cells are best explained by differences in expression of cell surface ligands, called costimulators, that provide antigen-independent signals that complement those provided by T-cell antigen-dependent recognition.241 Human ECs primarily provide costimulation to T cells through LFA-3 (CD58), which interacts with T-cell CD2. Monocytes and other professional antigen-presenting cells additionally provide costimulation to T cells through B7.1 (CD80) or B7.2 (CD86), which interact with T-cell CD28. Surprisingly, pig ECs express B7.2 and, more remarkably, pig B7.2 can functionally costimulate human T cells through CD28, a potential problem facing those who wish to use pigs as organ donors in human transplantation.244 

The recruitment of inflammatory cells is the second role played by ECs in CMI. Once memory T cells become activated by antigens, they, in turn, can activate a variety of EC functions that contribute to recruitment of inflammatory cells. The signals provided by T cells to activate ECs may involve contact-dependent signals (eg, T-cell CD40 ligand may engage EC CD40)245 or cytokines (eg, T-cell–secreted TNF/LT, IFN-γ, or IL-4 may act on the EC; reviewed in Pober and Cotran246). Several different responses of ECs to these signals contribute to inflammation, including production of vasodilators to increase delivery of leukocytes to the tissue, expression of adhesion molecules that tether and bind circulating leukocytes, synthesis of chemokines that contribute to transendothelial migration, and leakage of plasma proteins that form a provisional matrix in tissues for migration of extravasated leukocytes (reviewed in Pober and Cotran247).

The expression of various adhesion molecules on the endothelial surface changes over time, favoring neutrophil recruitment initially (eg, dependent on E-selectin expression) and recruitment of other leukocytes at later times (eg, dependent on VCAM-1 expression). The identity of chemokines may also change over time from synthesis of neutrophil-activating C-X-C chemokines early on to the subsequent synthesis of C-C chemokines that act on other leukocytes. The net result is that the composition of infiltrates change over time from neutrophil-rich to T-cell–rich and monocyte-rich delayed hypersensitivity (DTH) reactions or to T-cell–rich, eosinophil-rich, and basophil-rich late-phase reactions. The differences between DTH and late-phase inflammation appear to be attributable to local production of IFN-γ versus IL-4 and IL-5, respectively. The T cells recruited in DTH reactions are predominantly Th1-like cells (that mediate DTH), whereas those in late-phase reactions are predominantly Th2-like cells (that mediate late-phase reactions). The selective recruitment of Th1 cells into DTH reactions may be mediated by E-selectins and P-selectins.248 Interestingly, although Th1 and Th2 cells appear to express equivalent levels of PSGL-1, only Th1 cells are able to bind P-selectin.249 

Once infiltrates develop, a cytokine-rich milieu is generated that is sustained until the antigen is eliminated. Such chronic cytokine exposure has effects on ECs not seen at early times. For example, over the first few days, adhesion molecules that are initially expressed diffusely on the lumenal surface redistribute to inter-EC junctions.250 The basement membrane becomes enriched in sulfated glycosaminoglycans,251 and the cells assume an altered morphology characteristic of endothelium at sites of high lymphocyte extravasation, such as the high endothelial venules of lymph nodes.252 These features may promote leukocyte extravasation in acute settings. More chronic CMI reactions result in angiogenesis and tissue remodeling.

In general, CMI reactions do not produce endothelial injury, perhaps because T cells efficiently focus the response on the source of antigen, microbe-infected cells. An exception may be instances in which ECs are themselves infected by intracellular microbes (eg, viruses), so that cytolytic T lymphocytes (CTLs) kill the infected endothelium. Endothelial injury may also develop in transplantation, in which the immune system may perceive engrafted cells as self-cells that have been infected by virus. Some of the peptides recognized by graft-rejecting CTLs in association with allogeneic MHC molecules may be EC-specific and not found on leukocytes.253 In these instances, the CTL response may be directed at the endothelium. In addition, endothelium may be killed when CTL or natural killer (NK) cells are overstimulated by cytokines and lose their specificity. The actions of such lymphokine-activated killer (LAK) cells may contribute to the vascular leak syndrome associated with IL-2 or LAK therapy in cancer patients.254 It is increasingly appreciated that ECs may actively resist immune-mediated injury and that several of the resistance mechanisms involve cytokine-inducible genes.255Thus, the CMI response itself may protect ECs if the onset is sufficiently gradual (or if it is delayed by immunosuppression), so that the cells have had adequate time to acquire the resistant phenotype.

Erythrocyte-endothelial adherence.

Interactions between the endothelium and erythrocytes may contribute to the vascular complications of sickle cell anemia (SSA),256infection with Plasmodium falciparum malaria,257and diabetes.258 Red blood cell (RBC) adherence may initiate or promote intravascular sludging and occlusion leading to ischemic tissue and organ damage, retinopathy, dermal ulcers, strokes, and other infarctive pathologies. RBC adherence is dependent on EC surface molecules and is modulated by local hemodynamic factors. Recently, some of the RBC receptors, EC adhesion molecules, cytokines, and other vaso-active substances involved in adherence have been identified.

Sickle cell anemia.

Although the tendency for hemoglobin SS to polymerize at low oxygen tension is assumed to be the dominant factor in the pathogenesis of occlusive pain episodes, morphologic evidence of sickling is not seen immediately after hemoglobin is deoxygenated.259 Rather, adherence of SS-RBCs to vascular endothelium, which retards transit through the microvasculature, may be an important initiating event in this cascade.260,261 Adherence of SS-RBCs in vitro is sufficiently strong to withstand fluid shear forces typical of those seen in postcapillary venules.262 The resultant delay in capillary transit may allow time for sickle cells containing deoxygenated hemoglobin to deform leading to stable vascular obstruction263 and the resultant development of painful crises.

Adherence of SS-RBCs appears to result not only from the intrinsic membrane abnormalities induced in the erythrocytes, but also as a result of specific plasma factors and the state of EC activation (reviewed in Wick and Eckman264). For example, plasma from patients with SSA promotes RBC adherence in excess of that seen with normal plasma,265,266 and plasma collected during painful crises promotes adherence to an even greater extent.265Fibrinogen,265 fibronectin,262vWF,262 and thrombospondin267,268 have all been identified as factors in plasma that modulate SS-RBC adhesion.

Several adherence pathways have been described in vitro, including (1) bridging of GPIb-like molecules on sickle cells with their cognate receptors on ECs by unusually large vWF multimers released from activated platelets or by the stimulated endothelium itself262; (2) bridging by thrombospondin via CD36 on sickle reticulocytes269 and the αvβ3 integrin on large vessel ECs267 or αvβ3 and CD36 on microvascular endothelium268; (3) binding of sickle reticulocytes via α4β1receptors269,270 to VCAM-1 expressed on ECs stimulated by cytokines270,271 or by double-stranded RNA272; (4) binding of sickle reticulocyte via α4β1activated by phorbol ester or IL-8 to EC-associated fibronectin273; and (5) binding of SS-RBCs to E-selectin expressed on ECs stimulated by IL-1β.274 The expression of VLA-4 and CD36 on reticulocytes from sickle cell patients is reduced by treatment with hydroxyurea.275 However, the precise contribution of each of these or additional pathways to SS-RBC adherence in vivo requires further study.

In vitro, adherence of sickled RBCs to large venous vessels differs both qualitatively and quantitatively from adherence to microvascular endothelium.276 High molecular weight vWF multimers promote greater adherence to venous than to microvascular endothelium.276 Autologous plasma promotes greater adherence of sickle RBCs to microvascular endothelium than does plasma from individuals without SS disease.276 Sickle cell adherence is localized to postcapillary venules in ex vivo tissue perfusion studies, with no adherence observed in either capillaries or arterioles.263 These differences are likely due, at least in part, to variation in the expression of adhesion molecules and their receptors on the vasculature (see, eg, Swerlick et al277).

Sickle cell adherence is also dependent on local hemodynamic conditions (reviewed in Wick and Eckman264). Under static conditions, the dense SS-RBCs adhere most avidly, possibly due to intrinsic membrane alterations.261 In contrast, the least-dense sickle cells and reticulocytes are most adherent to the endothelium in vitro270,273,274 and ex vivo under flow conditions.230 Sickle cell adherence under flow is also more tenacious than under static conditions.278 Thus, adherence of reticulocytes expressing adhesion receptors may dominate in vivo in situations in which flow is maintained.264Adherence and trapping of membrane-damaged sickled erythrocytes may follow once flow has been further impeded, ultimately leading to complete vascular occlusion.279 

Painful crises frequently accompany ischemia, infarction, infection, or inflammation, situations in which the coagulation cascade may be activated as well.256 Activation of leukocytes and/or platelets may result in the generation and release of cytokines, adhesive proteins, or other factors that modulate endothelial or sickle cell adhesivity. Observations that activation of sickle cells273 and ECs with cytokines,270virus,272 or thrombogenic plasma proteins262,265,267,268 promotes RBC adherence suggests a mechanism by which infection and inflammation may initiate or propagate vaso-occlusion and episodic pain. Presumably, sickle cell adherence in vivo is most extensive at sites where the relevant adherence molecules are expressed most highly and shear stresses are sufficiently low to permit binding.264 Adherence of sickle cells may alter their metabolism,280 promote leukocyte adhesion,281 and contribute to their desquemation.281 

Malaria.

Plasmodium falciparum causes cerebral manifestations, perhaps the most serious complication of malaria. Maturation of the parasite within host RBCs induces membrane changes which promote adherence to cerebral microvascular endothelium in vitro and may lead to vascular congestion and hypoxia in vivo.257 It has been proposed that cytoadherence provides selective advantage to the invading parasites by facilitating their growth under the conditions of reduced oxygen tension found in the cerebral microcirculation and by enabling parasitized RBCs to avoid splenic filtration.257 

Parasitized RBCs bind to cell surface molecules on ECs, including CD36,257 ICAM-1,257 VCAM-1,282 and E-selectin.282 Additionally, thrombospondin allows bridging between parasitized RBCs and CD36 receptors on ECs.257 The various endothelial receptors act synergistically to slow and arrest parasitized RBCs under flow conditions and at shear stresses in the physiological range.283,284 Additional binding sites for parasitized RBCs may be induced as a consequence of the inflammatory response to Plasmodium falciparum infection257 due to leukocyte activation,285 cytokine release,286 and EC activation.285 

Binding to the endothelium occurs through knobs on RBCs induced by the parasite. These knobs contain Plasmodium falciparum erythrocyte membrane protein 1 that appears to participate in this process.257 Cytoadherence is inhibited by peptide fragments of the erythrocyte band 3 protein287 and by monoclonal antibodies that recognize band 3 on RBCs infected with mature parasites, suggesting involvement of cryptic regions of the protein exposed or altered during the course of infection.288 It has also been reported that some field isolates of Plasmodium falciparum promote adherence to chondroitin sulfate A.289 These data provide additional opportunities for antimalarial therapy based on inhibition of interactions between parasitized RBCs and ECs.

Diabetes mellitus.

Erythrocytes from patients with diabetes mellitus are more adherent to normal ECs than are RBCs from healthy donors.290 The extent of adhesion correlates with the severity of vascular complications290; greater adherence is observed in the absence of plasma,266 suggesting a defect intrinsic to the RBC. Adhesion is augmented further by plasma from diabetic patients as well as by fibrinogen,258 suggesting a mechanism by which acute-phase reactants may modulate vascular obstruction. Persistent exposure to hyperglycemia induces the formation of advanced glycation end products (AGEs) that modify structures on erythrocyte. AGE modification of erythrocytes allows them to engage a specific receptor, RAGE, (receptor for advanced glycation endproducts), which has been identified immunohistochemically and by in situ hybridization in the vasculature in vivo.258 Exposure of RBCs harvested from patients with diabetes to cultured endothelium results in increased adherence compared with those from euglycemic individuals due to erythrocyte-associated AGEs binding to endothelial RAGE.258Furthermore, it has been hypothesized that interaction of diabetic erythrocytes with endothelium through an AGE/RAGE linkage may promote oxidant stress leading to EC activation (see below).

It is currently believed that endothelium must remain in a resting or unperturbed state to optimize expression of anticoagulant activities which prevent thrombus formation (see “The Role of the Endothelium in Coagulation” above). However, the endothelium is a dynamic organ that responds to an array of agonists and environmental challenges by undergoing an activation process not unlike that of platelets, which eventuates in the loss of anticoagulant properties and/or acquisition of procoagulant function. Although the role of the endothelium in the pathogenesis of thrombosis in vivo remains unproven, accumulating evidence points towards dysregulation of EC function as pivotal in the development of several important thrombotic disorders. Plasma factors such as antibodies or lipoproteins that perturb EC function in vitro have been identified. It is likely that genetic differences in EC responsiveness to environmental pressures will be uncovered as contributors to the development of other common vascular diseases.

The endothelium in atherosclerosis.

Atherosclerosis is the most prevalent vascular disease in developed countries. The concept that atherosclerosis arises in response to endothelial injury was first proposed more than 20 years ago, when it was appreciated that irregularities in EC organization are often found overlying early fatty streaks, whereas overt endothelial denudation is seen only in the late stages of the disease (see Ross and Glomset291 for review; Fig 4). There is now extensive evidence that this morphologically abnormal endothelium is also dysfunctional and actually contributes to the propagation of lesions (see Ross292 and McGorisk and Treasure293 for reviews). These findings not only provide insight into the pathogenesis of atherosclerosis, but also suggest means to monitor the progression of lesions and effectiveness of treatment.

Fig. 4.

Very early development of atherosclerosis in a nonhuman primate. The focal origin of atherosclerosis is apparent beneath an intact endothelium. Scanning electron microscopy of ECs outlined by silver deposition in a thoracic aorta (photograph courtesy of Peter F. Davies, PhD, Institute for Medicine and Engineering, University of Pennsylvania, Philadelphia, PA).

Fig. 4.

Very early development of atherosclerosis in a nonhuman primate. The focal origin of atherosclerosis is apparent beneath an intact endothelium. Scanning electron microscopy of ECs outlined by silver deposition in a thoracic aorta (photograph courtesy of Peter F. Davies, PhD, Institute for Medicine and Engineering, University of Pennsylvania, Philadelphia, PA).

Close modal

Atherosclerosis is a multifactorial disease with numerous predisposing factors, including smoking, diabetes, hyperlipidemia, hypertension, mechanical stress, and inflammation. Such diverse and complex processes may perturb EC function through a common pathway. Alternatively, the endothelium may react to diverse stimuli with a limited repertoire of reparative, but ultimately dysfunctional, responses.

Oxidant stress.

Oxidant stress has been proposed as a mechanism common to diverse injuries such as unsaturated lipids that can be converted to cytotoxic lipid peroxidation products, various chemicals, radiation, and reactive oxygen metabolites released by leukocytes that migrate into the vasculature in response to infection and autoimmune injury. The pathways involved in the initiation and control of oxidant injury are receiving considerable study. Oxidized LDL and its peroxide derivative lysophosphatidylcholine stimulate protein kinase C activity, phosphoinositide turnover, and release of internal calcium; impair EC replication and angiogenesis; and induce apoptosis (reviewed in Henry294). Cytokines, such as TNF-α, can both induce reactive oxygen species in ECs and stimulate the ubiquitous transcription factor NF-κB, resulting in a transcriptional activation of other proatherogenic molecules such as VCAM-1 (reviewed in Larrick and Wright295). Oxidation reactions also promote the formation of AGEs that contribute to diabetic vasculopathy (reviewed in Schmidt et al296) and initiate transcriptional activation of VCAM-1297 and monocyte chemotactic protein-1 (MCP-1), which promotes monocyte entry into the vessel wall (reviewed in Gimbrone298).

Shear stress.

ECs are exposed continuously to fluid shear stresses that lead to a dynamic interaction between the cell and the substratum via focal contact sites. Shear-induced changes in transduced biomechanical forces can cause not only cytoskeletal rearrangement and altered morphology but changes in endothelial gene expression299,300(Fig 5). Most studies have examined primarily changes that occur within hours of initiating flow, which may best reflect the situation in vascular beds exposed to newly flowing blood such as postangioplasty, but the adaptive response of endothelium to shear forces is less well characterized.

Fig. 5.

EC alignment by directional steady flow in vitro (photographs courtesy of Peter F. Davies, PhD, Institute for Medicine and Engineering, University of Pennsylvania, Philadelphia, PA). (A) Before exposure to shear stress (no flow). (B) Twenty-four hours after exposure to flow (shear stress, 10 dynes/cm2).

Fig. 5.

EC alignment by directional steady flow in vitro (photographs courtesy of Peter F. Davies, PhD, Institute for Medicine and Engineering, University of Pennsylvania, Philadelphia, PA). (A) Before exposure to shear stress (no flow). (B) Twenty-four hours after exposure to flow (shear stress, 10 dynes/cm2).

Close modal

An effect of shear on vascular biology is suggested by the observation, eg, of decreased vasodilator function at coronary branch points that have a predilection for atherosclerosis.301 Consistent with this notion, a number of genes relevant to the development of atherosclerosis expressed by ECs have shear stress response elements that coordinate their induction. Shear modulates EC production of products regulating vasoconstriction (NO, endothelin-1), vessel growth (βFGF, platelet derived growth factor [PDGF]-A and -B, and TGF-β), fibrinolysis (t-PA), and cell adhesion (MCP-1, VCAM-1, and ICAM-1) (see Malek and Izumo,300 Ando and Kamiya,302 and Tsao et al303 for reviews). Shear has been reported to modulate the expression of thrombomodulin in a reversible manner299 and abrogate cytokine-induced EC tissue factor expression.304 

The mechanism(s) responsible for the modulation of gene expression by shear is under study. At least part of these shear-induced effects are mediated through modulation of gene transcription. A number of genes, such as PDGF-B, contain one or more shear stress responsive elements (SSREs) that include an NF-κB–responsive GAGACC promoter sequence305 in the 5′ upstream region. However, the induction of TGF-β and MCP-1 appear to occur through alternative sites, eg, a TRE/AP-1–responsive element.306 Levels of transcription factors NF-κB and AP-1 are increased in sheared EC.307 How shear-related transcription factors affect immediate and persistent gene transcription and how this inductive pathway differs from other injury-related responses remain to be elucidated.

Homocysteine.

Homocysteine is a sulfhydryl amino acid formed during the conversion of methionine to cysteine. Elevated plasma levels of homocysteine may result from deficiencies of cystathionine-β-synthase, deficiencies of enzymes involved in the folate-dependent pathway of homocysteine remethylation, or deficiencies of folate or vitamin B12 themselves (see Guba et al308 and Rees and Rodgers309 for reviews). Homozygous deficiency of cystathionine β-synthase leads to markedly elevated plasma concentrations of homocysteine and is associated with premature atherosclerosis and arterial thrombosis. The results of several recent large prospective and case-controlled studies suggest that even modestly elevated levels of homocysteine may pose a risk factor for atherosclerosis as well as for arterial and venous thrombosis (reviewed in Mayer et al310), although it is not clear that all pathways leading to hyperhomocysteinemia pose comparable risks. Several large prospective studies are underway to assess the magnitude of this risk.

These observations have led to the identification of several pathways by which homocysteine may affect the anticoagulant and procoagulant functions of cultured ECs (Table 2).311 It is important to consider that most studies to date have used cultured cells exposed for brief periods to concentrations of homocysteine that exceed that which is observed in vivo. Clearly, such concentrations of homocysteine induce formation of hydrogen peroxide312 and oxidized LDL313 and may be directly cytotoxic for ECs; NO is at least partially protective.314 However, induction of TF,315 activation of factor V,316 decreased binding of t-PA to annexin-II,317 inhibition of thrombomodulin,318 and reduced expression of heparan sulfate319 and possibly PGI2,320among many other changes, have all been observed in ECs exposed to high concentrations of homocysteine. Homocysteine is also mitogenic for rat aortic smooth muscle cells.321 That similar effects may occur in vivo is suggested by vascular dysfunction in monkeys with diet-induced hyperhomocysteinemia322 and elevated levels of thrombomodulin and vWF in the plasma of homocysteinemic patients with peripheral arterial disease. Therapy with pyridoxine and folic acid led to rapid reductions in the levels of both markers.323Studies in animal models in which moderately elevated levels of homocysteine is sustained may provide additional insight into the role of the endothelium and other pathogenic mechanisms of thrombosis and atherosclerosis.

Consequences of EC injury on endothelial-derived vasoactive factors.

Healthy human epicardial coronary arteries dilate when acetylcholine is infused,324 whereas atherosclerotic arteries constrict325 due to impaired release of NO (see “Vasoregulation” above) and PGI2 from the endothelium.326 Loss of vasodilitation in the absence of overt stenosis has been observed in patients with hypercholesterolemia,327 increased Lp(a),328diabetes,329 homocystinuria,330 and possibly hypertension (reviewed in Woodman331 and Heistad et al332). Similar effects have been observed with advancing age,333 exposure to cigarette smoke,330,334 and sedentary life style.335 Endothelial dysfunction is almost universal 12 to 24 months after cardiac transplantation.336Indeed, abnormal vasodilator function may be a more sensitive marker of coronary artery disease in some settings than is angiography.337 

The reduction in NO is attributable in part to lowered levels of eNOS in atherosclerotic vessels,338 but the reason for this is uncertain. One clue may be that impaired vasodilatation is especially marked at coronary branch points where flow and shear stress have been shown to stimulate NO production in healthy vessels.301 In addition, free oxygen radicals and hydrogen peroxide generated by EC exposed to high levels of LDL may inactivate NO.339 

Inhibition of NO activity accelerates atherosclerosis in animal models,340 whereas supplementation with L-arginine, the precursor of NO, diminishes lesion formation (reviewed in Cooke341) and reverses endothelial dysfunction in otherwise healthy young humans with hypercholesterolemia.342 The loss of NO may impact on multiple steps in the atherogenic process.341 NO may also counteract the proatherogenic effects of endothelin (reviewed in Mathew et al343). Oxidized LDL binds to a newly described receptor on ECs344and increases the production and secretion of ET in cultured ECs and in intact blood vessels.345 ET is also released by dysfunctional coronary arteries that constrict in response to acetylcholine.346 Plasma concentrations of ET are elevated in asymptomatic patients with hypercholesterolemia and increased plasma levels of both Lp(a)347 and ET have been reported to correlate with the severity of atherosclerosis.346 

Monitoring the reversal of EC dysfunction.

A significant reduction in cardiac events has been reported as a result of lipid lowering therapy, but the mechanism responsible for this benefit has not been elucidated. Little reduction in the cross-sectional area of preexisting lesions is seen, although progression of the lesions may be slowed348 and incidence of plaque rupture may be lessened. What is clear is that there is an improvement in various indirect measures of EC function in vivo possibly as a result of a more favorable profile of vasoactive agents that are produced at critical locations.349 Similar beneficial effects on surrogate markers of EC function have been seen with the use of antioxidants,350 whereas in other studies the acute administration of vitamin C has been associated with improved EC-dependent vasodilatation in chronic smokers351 and patients with coronary artery disease.352 Similar beneficial effects of angiotensin converting enzyme inhibitors have been confirmed in some353 but not in all354animal models (reviewed in Lonn et al355).

Thus, with the advent of potential means to treat atherosclerosis, the need for reliable, noninvasive surrogate markers of risk and vascular function have become apparent. Several candidate molecules have emerged. Abnormalities in serotonin-induced arterial vasodilatation, a process dependent on NO, precede the development of clinical disease and resolve within 12 weeks of the institution of cholesterol lowering therapy.356 Elevated plasma levels of PAI-1 and thrombomodulin revert towards normal as well.357 In contrast, little change has been seen in the elevated plasma levels of E-selectin, VCAM-1, and ICAM-1.358 There is also interest in measuring levels of 8-epi PGF, an isoprostane with potent vasoconstrictor activity in the pulmonary and renal circulations generated through free radical catalyzed peroxidation of arachidonic acid by ECs and other vascular cells. Levels of 8-epi PGF in the urine have been used as a marker of oxidant stress in vivo (reviewed in Morrow and Roberts359). Levels in smokers are elevated and fall with cessation of smoking or treatment with vitamin C but not with vitamin E or aspirin therapy.360 

Perspective.

Healthy ECs contribute to the prevention of atherosclerosis in medium to large arteries by inhibiting platelet activation, limiting the entry of cells and lipids into the vessel wall, maintaining a nonproliferative and biochemically quiescent intima,361 and secreting products under appropriate stimuli that limit potentially injurious responses that occur as a byproduct of host response to injury. These self-protective mechanisms are impaired as a result of oxidant, chemical, and shear stress, while at the same time the biochemical profile of the endothelium changes in a way that promotes inflammatory and fibroproliferative responses. The role of ECs in preventing or limiting the effects of plaque rupture and terminal thrombosis is little understood. The injurious processes that initiate atherosclerosis appear to persist throughout life in most individuals. Studies are now being conducted to determine the extent to which the biochemical and functional changes in the vessel wall can be reversed at different stages of the disease.

Endothelial perturbation and vascular dysfunction in diabetes.

Vascular dysfunction is a contributing factor in the etiology of several clinically important secondary complications of diabetes mellitus including retinopathy, accelerated atherosclerosis, microvascular disease, nephropathy, neuropathy, and impaired wound healing.362,363 The effects of hyperglycemia on EC function can be imparted through several pathways: (1) production of reactive oxygen intermediates; (2) direct activation of protein kinase C; (3) activation of the aldose reductase pathway resulting in an accumulation of sorbitol and diminished levels of myo-inositol; and (4) nonenyzmatic glycoxidation of long-lived macromolecules.362,363 Because glycoxidation of proteins and lipids occurs ubiquitously in patients with diabetes and is irreversible, its consequences are especially relevant to long-term vascular dysfunction. Initially, exposure of free amino groups to reducing sugars, such as glucose, results in the formation of early glycation products, Schiff bases, and Amadori products. These are reversibly modified species, such as hemoglobin A1c, used for long-term monitoring of blood sugar in diabetic patients. Further molecular rearrangements occur, in part due to oxidation, resulting in irreversible AGEs (Fig 6). The latter have pathophysiologic relevance in that AGE-modified proteins may not function normally and/or may perturb cellular properties in a manner distinct from that of the native molecule. This occurs when the AGE form of the molecule binds to cellular receptors which recognize AGEs, including RAGE364,365 and the macrophage scavenger receptor.366 RAGE is expressed by endothelium, monocytes, and smooth muscle cells and is likely to play a major role in the development of vascular disease in diabetics.365 

Fig. 6.

A two-hit model of vascular perturbation. Stage 1 shows the interaction between AGEs, modified biochemical species and their receptors (RAGEs), transmembrane protein receptors of the Ig family present at low levels on a range of cells including ECs in addition to macrophages (Mfs), smooth muscle cells, and neurons. AGEs may perturb the normal physiologic function of the modified species and thus alter the EC's normal vascular functions. Stage 2 details subsequent perturbations resulting from the superimposed stimuli of accumulated lipoprotiens as seen in atherosclerotic lesions, foreign materials that may be seen in wound repair, and bacterial infection that may be seen in periodontal disease.

Fig. 6.

A two-hit model of vascular perturbation. Stage 1 shows the interaction between AGEs, modified biochemical species and their receptors (RAGEs), transmembrane protein receptors of the Ig family present at low levels on a range of cells including ECs in addition to macrophages (Mfs), smooth muscle cells, and neurons. AGEs may perturb the normal physiologic function of the modified species and thus alter the EC's normal vascular functions. Stage 2 details subsequent perturbations resulting from the superimposed stimuli of accumulated lipoprotiens as seen in atherosclerotic lesions, foreign materials that may be seen in wound repair, and bacterial infection that may be seen in periodontal disease.

Close modal

RAGE is a member of the Ig superfamily of cell surface molecules. It is composed of an extracellular domain with one V-type, followed by two C-type regions.364,365 There is a single transmembrane spanning domain and a short, highly charged cytosolic tail that likely transmits the signal of ligand occupancy by interacting cytosolic transduction molecules. The single RAGE gene is located on chromosome 6 in the major histocompatibility complex between genes for class II and class III molecules. This proximity of RAGE to genes contributing to the host response is in keeping with expression of the receptor. In mature animals, RAGE is present at low levels in a range of cells (endothelium, smooth muscle cells, mononuclear phagocytes, and neurons), but after perturbation, as in diabetes, immune/inflammatory disorders, or Alzheimer's disease, RAGE expression is dramatically upregulated. In diabetes, AGE-modified proteins appear to act as ligands for RAGE modulating a number of secondary messenger pathways. For example, AGE interaction with RAGE results in cellular oxidant stress eventuating in activation of p21ras, MAP kinases (erk's 1 and 2), and the transcription factor NF-κB.367 Such activation of NF-κB results from binding of p50/p65 heterodimers to DNA binding motifs, as in the gene for vascular cell adhesion molecule-1 (VCAM-1). RAGE-dependent enhanced VCAM-1 expression is observed both in cultured ECs and in vivo in mice infused with AGEs.368 Increased VCAM-1 levels are also observed in diabetic vasculature upon immunohistological analysis. In parallel with expression of VCAM-1 on the cell surface, ECs release a soluble form of VCAM-1 (sVCAM-1) into culture supernatants, potentially providing a means of monitoring cellular stress in vivo. Patients with diabetes and microalbuminuria, the latter considered a harbinger of impending future vascular complications, display higher plasma sVCAM-1 than those without microalbuminuria.369 

AGE-RAGE interaction likely underlies vascular hyperpermeability, another salient feature of diabetic vasculopathy. Such hyperpermeability is blocked by anti-RAGE IgG or by preventing AGE binding by infusion of a soluble form of the extracellular domain of RAGE (the latter is termed sRAGE). Vascular leakage of solutes in diabetic animals can be largely blocked by infusion of sRAGE. These data identify a reversible component of diabetic vascular dysfunction and suggest that AGE-RAGE-induced cellular perturbation may be a contributor. The principal insights to be gained from analysis of RAGE binding to nonenzymatically glycated ligands is probably in the setting of chronic vascular perturbation.365 Because AGE modification of proteins is irreversible, AGEs accumulated in the vessel wall are present for extended periods of time. Thus, the diabetic vascular milieu has properties that distinguish it from that in euglycemic subjects. A two-hit model can be envisioned (Fig6) in which tissue and blood AGEs interacting with RAGE provide a baseline state of vascular activation for the first-hit. The second stage comprises a superimposed stimulus, such as accumulated lipoproteins in atherosclerotic lesions, foreign material in wounds, and bacterial infection in periodontal disease. AGE-RAGE interaction provides a backdrop of chronic inflammation, with increased expression of proinflammatory cytokines, thrombogenic factors, cell adherence molecules, and vascular permeability, which aggravates and probably accelerates the development of vascular lesions.

Antibody-mediated EC injury: Solid organ transplantation.

Transplantation of vascularized organs, such as kidney, heart, lung, and liver, has become the treatment of choice for end-stage organ failure. The key limitations on clinical transplantation today are rejection of allografts posttransplantation and the shortage of available donor organs. The EC lining of graft vessels plays a prominent role in both of these clinical problems.

Despite the enormous advances in clinical immunosuppression of transplant recipients that have been made in the last 50 years of practice, the principal cause of graft failure is still rejection, ie, immunological reactions of the host against graft cell alloantigens that injure and destroy the graft. ECs play three crucial roles in the process of graft rejection: (1) ECs stimulate the host immune system by presenting alloantigens in an immunogenic form to host lymphocytes, thereby helping to initiate graft rejection; (2) ECs respond to host stimuli, eg, inflammatory cytokines, to promote intragraft inflammation and thrombosis that contribute to graft injury; and (3) ECs lining graft vessels are primary cellular targets of the host antigraft response. In addition, graft ECs are sensors and mediators of antigen-independent injury to which the graft is subjected during harvest, transport, and implantation, a dramatic instance of ischemia reperfusion. The mechanisms by which ECs present antigens to lymphocytes, promote inflammation in response to cytokines, are injured by the immune response, and respond to oxidant injury have been discussed in previous sections of this review and will not be duplicated in this section. We will focus here on specific issues related to immunologic allograft rejection.

Rejection reactions are commonly classified according to the time when they occur after surgical transplantation and by their histopathologic features.370 Hyperacute rejection occurs within minutes to hours of perfusion of the graft by host blood, ie, in the perioperative period. It is characterized by extensive intravascular thrombosis of graft vessels and consequent graft ischemic infarction. Hyperacute rejection is mediated by host antigraft antibodies, usually IgM antibodies reactive with graft endothelial carbohydrate epitopes such as ABO blood groups, and by complement activation that is initiated by the antibodies bound to graft ECs. Thrombosis results from EC lysis and desquamation, exposing thrombogenic subendothelial basement membrane, or, in cases of sublytic quantities of complement depositon, by loss of endothelial antithrombotic mechanisms (eg, shedding of cell surface anticoagulant heparan sulfate) combined with activation of endothelial prothrombotic mechanisms (eg, release of stored high molecular weight vWF, release of lipid procoagulants, and possibly induction of TF).371 Matching of donors and recipients for ABO blood groups has significantly reduced the incidence of hyperacute rejection in allografts.

Acute rejection reactions usually develop between the first and second week after transplantation. Although a variety of effector mechanisms may be involved, recent studies of human allograft biopsies have emphasized the primary role of cytolytic cells, principally CTLs.372,373 Some host CTLs recovered from rejecting grafts have shown specificity for graft ECs over graft leukocytes.374 Because the specificity of CTL is determined by the antigens involved in their induction (ie, the same antigen receptor mediates initial differentiation and subsequent effector function of CTL), graft ECs must have played a role in stimulating host CTL development by presenting antigen to precursor cells, ie, to resting CD8+ T cells recruited from the circulation into the graft. Lysis of microvascular ECs is a prominent and early component of acute cell-mediated rejection.375 More severe rejection reactions typically involve injury of larger graft vessels as well. Such vascular rejection is believed to start as a host CTL reaction against graft arterial or arteriolar ECs (called endothelialitis or intimal arteritis at this early stage373) and may progress to severe necrotizing, transmural vasculitis. The most severe vascular rejection reactions appear to involve host antigraft antibodies (both IgG and IgM) as well as cytolytic lymphocytes. Presensitized hosts (eg, resulting from a prior transplantaton procedure) who have expanded numbers of memory T or B cells reactive with the donor may show accelerated acute rejection during the first few days after transplantation370). These accelerated rejection reactions are similar to experimental second set rejection seen in animal models of retransplantation. The use of tissue typing (for renal transplantation) and of improved immunosupression (especially since the introduction of cyclosporin A) have reduced the incidence of graft loss due to acute and/or accelerated rejection to fewer than 10% of organs. ECs are considered prime targets for further improvements in immunosupression, eg, by targeting adhesion molecules such as ICAM-1 to reduce posttransplant ischemia-reperfusion, inflammation, and CTL effector functions.

With current advances in controlling acute rejection, the major cause of graft loss has become chronic rejection.376 Chronic rejection appears in biopsies as replacement fibrosis of graft parenchyma, developing over months to years. These changes are widely thought, at least in cardiac and renal transplantation, to be secondary to graft ischemia caused by progressive occlusion of the lumen of graft arteries (called graft arteriosclerosis). Graft arteriosclerosis is characterized by concentric, diffuse intimal hyperplasia of large, medium, and small graft arteries. These changes are accelerated compared with atherosclerosis in that they can develop into clinically significant lesions as early as 6 months to a few years posttransplantation. Despite all of the advances in immunosuppresion for acute rejection, the incidence of allograft failure to chronic rejection has remained at about 10% of grafts per year, with no evidence of improvement.

Graft arteriosclerosis is restricted to graft vessels, ie, it completely spares the host's vessels. Involved graft arteries contain increased numbers of intimal smooth muscle cells and deposition of extracellular matrix accompanied (or preceeded) by a sparse subendothelial infiltrate of host T cells and macrophages.377 Cytolytic effector cells are markedly fewer than in acute vascular rejection,378 and the endothelium shows only rare apoptotic cells.379 The major theories of pathogenesis are that graft arteriosclerosis results from chronic, low-level endothelial injury (ie, persistent endothelialitis caused by CTLs and/or alloantibodies), followed by fibroproliferative repair, or that graft arteriosclerosis results from a conversion of an acute cytolytic immune response to a chronic delayed-type hypersensitivity reaction in response to persistent immune stimulation by graft ECs. There is currently no therapy for graft arteriosclerosis except retransplantation. Future therapy may be targeted at preventing the intimal smooth muscle fibroplastic reaction rather than further increases in immunosuppression.

The next horizon in transplantation is to address the donor organ shortage by xenotransplantation of animal organs (eg, pigs) into human recipients.380 The use of pig organs has raised concerns about introducing new infectious agents into the human population, but the major practical problem is a very high incidence of hyperacute rejection. Two factors contribute to development of severe, uncontrollable hyperacute rejection of pig xenografts by human or old world monkey recipients. First, all mammals, except humans and old world monkeys, express a galactose α-1,3 galactose carbohydrate epitope instead of ABO blood groups on their ECs.381Moreover, virtually all humans have high levels of circulating natural IgM antibodies reactive with this alternative epitope, so that ABO matching cannot be used to evade hyperacute rejection in xenotransplantation. Second, the problem of abundant natural antibody is compounded by the fact that pig ECs express complement regulatory proteins, eg, DAF and CD59, that are unable to control the human complement system, ie, they are species specific for pig complement proteins.382 The combination of high levels of complement-activating IgM antibodies and limited EC resistance to human (or primates) complement proteins invariably leads to rapid and overwhelming intravascular graft thrombosis.

A major current effort is underway to produce transgenic pigs that will have reduced levels of galactose α-1,3 galactose and/or express human complement regulatory proteins. If successful, transplanters will still need to address later phases of the human-antipig rejection reaction (eg, potentially strong acute cellular rejection due to expression B7.2 costimulator molecules on pig ECs383 or to strong NK reactions384 due to possibly antibody binding and absence of self class I MHC molecules on pig ECs). Some of these problems may be partly ameliorated by the failure of pigs to respond to certain human cytokines (eg, to IFN-γ385), but additional studies will be needed to determine the significance of these differences.

Systemic lupus erythematosus (SLE) and the antiphospholipid antibody syndrome (aPS).

Immune-mediated endothelial dysfunction may contribute to the development of thrombosis in patients with SLE and the aPS (see McCrae and Cines386 for review). One mechanism by which endothelial damage and/or activation may occur is through the effects of EC-reactive antibodies. Several groups have demonstrated anti-EC antibodies (AECA) in sera of patients with SLE387and in patients with primary and secondary aPS (see McCrae et al,388 among others). Controversy remains as to whether antiphospholipid antibodies per se comprise the biologically important subpopulation of AECA (see Cines389 for review), although recent studies suggest that they may activate ECs through an effect on the plasma protein β2-glycoprotein I.390 AECA have been shown to alter the anticoagulant and procoagulant activities of cultured ECs in a number of ways. However, the in vivo importance of these antibodies in the pathogenesis of thrombosis remains unknown. Indeed, in only some studies has the presence of AECA correlated with thrombotic events or disease activity.388 Nevertheless, the fact that such antibodies induce the secretion of markers of EC injury/activation such as vWF from cultured ECs,388 when considered in light of reports demonstrating elevated levels of vWF in the plasma of patients with SLE,391 suggests that at least some of these in vitro effects may reflect processes that occur in vivo.

Heparin-induced thrombocytopenia and thrombosis (HITT).

Approximately 1% to 3% of patients who receive heparin develop severe thrombocytopenia, approximately 20% of whom also develop venous and/or arterial thrombi (see Arepally and Cines392and McCrae and Cines393 for reviews). Plasma from approximately 90% of patients with HITT contain antibodies that bind to complexes of heparin and platelet factor 4 (PF4).394 PF4 released from activated platelets may form complexes with heparin on the surface of activated platelets, targeting them for FcγRIIA-dependent activation by antiheparin-PF4 antibodies. However, the remarkable propensity of patients with HITT to develop thrombosis may also be promoted by the capacity of these antibodies to recognize PF4 bound to EC-heparan sulfate proteoglycans, which stimulate ECs to express TF and to bind platelets.395,396 

EC-reactive antibodies in other vasculitic disorders.

AECAs have been described in several other disorders in which their role in pathogenesis is even less certain. PR-3, a common target of antineutrophil cytoplasmic antibodies (ANCA) is expressed on activated ECs in vitro397 and has been implicated in vascular injury in an animal model of vasculitis.398 However, EC-specific antibodies, apparently distinct from ANCA, have also been described in many patients with Wegener's granulomatosis.399 Plasma from children with the hemolytic uremic syndrome contain lytic AECA that recognize an unidentified EC surface protein that is suppressed by IFN-γ in vitro.400 AECA have also been implicated in mixed connective tissue disease, rheumatoid arthritis, and scleroderma,401 in which EC apoptosis has been implicated,402 as well as in atherosclerosis,403 Kawaski's disease,404Bechet's disease,405 and various forms of vasculitis.406 In each of these conditions, elevated levels of EC-derived proteins have also been described in the plasma of affected patients implying endothelial activation or injury.407 However, a pathogenic role of AECA or immune complexes in human vasculitis has not been proven. It is also pertinent to note that complex changes in antigen expression occur when ECs have been activated by cytokines or other agonists in vitro (see Favaloro408 for review) and in vivo.409 This activation promotes cell mediated immunity, promotes leukocyte/platelet adhesion (see “EC Pertubation and Vascular Disease” above) and exposes cryptic autoantigens. Fcγ receptors capable of binding circulating immune complexes may exist410 or can be induced on ECs411 at certain vascular sites.

Complement-mediated EC activation.

The vascular endothelium is exposed to activated complement components as a consequence of antigen-antibody interactions that occur naturally in plasma and within the vessel wall in certain pathologic conditions.387 ECs also constitutively express several complement proteins412 and can be induced to synthesize others by various cytokines,413 at least in vitro. Complement deposition on the vasculature is controlled, in part, by concerted actions of the regulatory proteins C1-esterase inhibitor, which is secreted by ECs,414 and by decay-accelerating factor, membrane cofactor protein of complement, and homologous restriction factor, which are expressed on the cell surface.415-419 Complement that deposits on the endothelium when these containment mechanisms are exceeded can increase vascular permeability, stimulate procoagulant pathways, and recruit ECs to become active participants in the inflammatory processes.

For example, multiple components of the complement cascade act in concert to augment the recruitment of leukocytes to sites of vascular inflammation. Binding of C1q to specific EC receptors420augments expression of E-selectin and possibly ICAM-1 and VCAM-1,421 C5a upregulates the expression of P-selectin,422 and the terminal C5b-9 membrane attack complex (MAC) acts synergistically with TNF-α to stimulate expression of E-selectin and ICAM-1.423 Sublytic concentrations of MAC also activate the NF-κB pathway leading to secretion of IL-8 and MCP-1.424 

MAC also stimulates cultured ECs to express TF,425,426 to release large vWF multimers,427 and to release procoagulant microvesicles containing anionic phospholipids and binding sites for factor Va that accelerate prothrombinase activity.428 On the other hand, MAC also stimulates production of PGI2429 and the complement components C7 and C9 provide sites for the binding and activation of plasminogen.430 However, as with other areas of research in vascular biology, studies have been confined largely to cultured cells and additional studies are required to elucidate the relative contribution of complement-dependent responses in vivo.

Thrombotic disorders of uncertain cause: Thrombotic thrombocytopenic purpura (TTP) and the hemolytic uremic syndrome (HUS).

TTP and HUS are related-disorders characterized pathologically by the development of platelet microthrombi that occlude small arterioles and capillaries and clinically by microangiopathic hemolytic anemia and thrombocytopenia. Endothelial dysfunction plays a prominent role in the pathogenesis of both disorders (see Moake431and Heild432). Approximately 90% of cases of HUS occur in early childhood, often after an episode of bloody diarrhea caused by enteropathic strains of Escherichia coli that release an exotoxin, designated verotoxin-1 (VT-1), which is similar to the 70-kD Shiga toxin 18. VT-1 binds with high affinity to globotriosylceramide (Gb3) receptors expressed at the highest density on renal glomerular ECs.431 VT-1 is directly cytotoxic to ECs. In addition, VT-1 promotes neutrophil-mediated EC injury433,434 and induces the production of TNF-α by monocytes435 and cells within the kidney.436 In turn, TNF-α, in concert with IL-1, increases Gb3 expression and exacerbates the sensitivity of the endothelium to toxin-mediated437 and antibody-mediated404 cytotoxicity, promotes vWF release,438 and impairs fibrinolytic activity.44 In accord with this putative pathogenic mechanism, elevated plasma levels of PAI-1 have been reported to be a sign of a poor prognosis in childhood HUS.439 EC injury may also contribute to the pathogenesis of the microangiopathic syndromes that may follow the use of certain chemotherapeutic agents,440 cyclosporin,441 quinine/quinidine (see Gottschall et al442 for review), or after bone transplantation.440 

There is considerable evidence to suggest that EC injury plays a role in the pathogenesis of TTP. The most exhaustively studied protein in this regard is plasma vWF, which circulates in plasma as oligomers that range in size from 1 to 15 × 106 kD. The so-called unusually large vWF multimers (ULvWF) are normally found in subendothelial matrix, in the supernatant of cultured ECs,431 and in platelet releasates, but are not normally detected in plasma.431 Platelet microthrombi in TTP contain abundant vWF but little fibrinogen, in contrast to those seen in DIC. A subgroup of patients has been identified who suffer from chronic, relapsing TTP and whose plasma continues to contain elevated levels of ULvWF between relapses.443 Furthermore, plasma from patients with sporadic or isolated episodes of TTP often contains either ULvWF or decreased amounts of larger vWF multimers during periods of active disease. ULvWFs may exacerbate microvascular thrombosis through their ability to aggregate platelets at high levels of shear stress. The secretion of ULvWF by cultured ECs is stimulated by many agonists including Shiga toxin.431 However, elevated levels of vWF occur in other thrombotic microangiopathies, and their exact role in TTP/HUS requires further study. Reports of elevated levels of thrombomodulin,444 tissue-type plasminogen activator (t-PA), plasminogen activator inhibitor type 1 (PAI-1),445 ELAM-1,446 and decreased levels of PGI2447 and TFPI448 in plasma from patients with TTP provides additional support for the notion that endothelial damage plays a pivotal role in the pathogenesis of the disease.

The events that initiate TTP remain unknown. AECA have been described in TTP and HUS, but their role is uncertain. More recently, plasma from patients with TTP and HUS has been reported to induce apoptosis in microvascular ECs; it is of great interest that cells from dermal, renal, and cerebral origin were most susceptible, whereas pulmonary and coronary arterial cells were not.46 The plasma factors responsible for these changes remain to be identified.

Pregnancy-induced hypertension.

Pregnancy-induced hypertension (or preeclampsia) is the most common medical disorder of pregnancy, affecting 5% to 13% of all primaras. Although the clinical manifestations of preeclampsia are generally not evident until the third trimester, the pathogenesis of this disorder may involve a deficiency in placentation,449 the process in which fetal trophoblast cells remodel the maternal uterine spiral arteries early in pregnancy. Incomplete remodeling of the spiral arteries leads to compromised placental perfusion. Alterations in EC morphology occur within the placenta450 and in the glomerular capillaries (glomerular endotheliosis), the latter being characterized by EC swelling and lipid accumulation (see Ferris451 for review). Fibrin deposition in microvasculature is common. Affected women show increased responsiveness to the pressor effects of angiotensin II,314increased amounts of thromboxane A2 relative to PGI2 in their urine,452 elevated plasma levels of endothelin,453 and their umbilical vessels demonstrate less PGI2 synthesis and decreased NO release in response to bradykinin (see Ferris451 for review). Additional evidence suggesting endothelial damage is the reported findings of elevated plasma levels of EC-derived fibronectin,454,455vWF,455 and PAI-1455 in affected women. Importantly, increased levels of vWF456 and cellular fibronectin454 may be detected before the onset of clinical manifestations.

The pathophysiology of EC damage in preeclampsia remains a mystery. Some,457,458 but not all groups,459,460 have reported that plasma from affected women is cytotoxic for cultured ECs. Preeclamptic sera have also been reported to impair EC proliferation,461 stimulate fibronectin release,462 increase triglyceride accumulation,463 and increase PDGF synthesis464; variable effects on PGI2synthesis have been reported.463,465 A role for abnormalities of lipid peroxidation,466 for immunologic factors, and for underlying, but otherwise inapparent, maternal vascular disease (see Ness and Roberts467 for review) have also been advanced. Yet, it must be noted that all of these studies remain largely unconfirmed, the putative injurious plasma factor(s) has yet to be identified, and the pathophysiology and significance of EC injury outside of the terminal cases of eclampsia remains enigmatic.

The endothelium can no longer be viewed as a static physical barrier that simply separates blood from tissue. Rather, it is now clear that the endothelium helps to coordinate functions of differentiated tissues in a way that meets the requirements of the organism as a whole. In part, this is accomplished by the location of the endothelium at the interface with the blood and the capacity of these specialized cells to receive and transmit biochemical and physical information bidirectionally. Information sensed on the lumenal surface of the endothelium can be transmitted either by direct permeation or active transport of soluble mediators through the capillaries to deeper tissues or indirectly through the capacity of ECs to modulate the behavior of smooth muscle cells and other components of the vessel wall. In turn, physiologic and pathophysiologic events in tissue alter EC interactions with soluble and cellular blood components.

The endothelium, as with all cell types, displays an immediate and prototypic response to diverse agonists that is modulated in complex ways by subsequent events. In the case of the endothelium, this first response appears designed to prevent physical disruption of the vessel wall by trauma, microbial organisms, toxins, or other threats to the maintenance of intravascular volume and oxygen delivery. This protective response is accomplished by the rapid transformation of the endothelium to a procoagulant, vasoconstrictive, and proinflammatory state that has multiple effects on of its structure and behavior.

Several ramifications of this reflexive, adaptive response of the endothelium have now become evident. First, it is clear that ECs rapidly undergo some of these same biochemical and phenotypic changes soon after being placed in culture, as a consequence of which the behavior of the unperturbed endothelium cannot be reliable inferred from currently available in vitro techniques.

Second, an extensive experimental literature has emerged supporting the notion that several common human vascular diseases are in part a consequence of the same responses of the endothelium to stress; ie, that prolonged or exaggerated endothelial activation leads to dysfunction that is an early, often preclinical component of vascular disease. Unfortunately, it is generally impossible to access vascular tissue directly and sequentially during these preclinical stages of disease development; without such tissue, the EC contribution to disease development can only be inferred. As a consequence, most research in vascular biology continues to (1) focus on the footprints of disease by analyzing damaged vessels, generally at the endstage of the process; (2) link putative circulatory factors to disorders through their effect on cultured ECs, often derived from unaffected tissue; and (3) develop animal models that may simulate human diseases.

Third, it is now clear that the endothelium is not a homogeneous organ. ECs from different vascular beds show highly differentiated functions as a consequence of genetic diversity and the impact of specialized surroundings. These biochemical and phenotypic differences extend to their susceptibility to injury and effect function of the vasculature as a whole.

Fourth, there is remarkably little information on the potential contribution of genetic differences in EC behavior among individuals with respect to bleeding disorders, thrombosis, atherosclerosis, and vasculitis. Without such information, our current approach to studying these major vascular disorders can be compared with the study of anemia without appreciating the existence of hemoglobinopathies or the study of bleeding disorders without appreciating the contribution of genetic abnormalities in platelet function. It is hoped that future research will enable the direct study of EC behavior and thereby enhance our understanding of the contribution of the endothelium to vascular biology.

Fig. 7.

EC injury in a case of antiphospholipid antibody syndrome (photographs courtesy of Emma E. Furth, MD, Department of Pathology and Laboratory Medicine, University of Pennsylvania, Philadelphia, PA). EC injury seen in a duodenal biopsy from a 40-year-old woman who presented with profuse intestinal bleeding and was found to have a lupus anticoagulant and a markedly positive anticardiolipin antibody. (A) Elastic stain (original magnification × 200) highlighting a fresh thrombus (right) with the beginning stages of organization and EC ingrowth, an older organized thrombus with fibroblast proliferation (center), within a vessel showing vacuolated, injured, and disrupted ECs. (B) A trichrome stain (original magnification × 400) of the same field highlighting the thrombus material (red acellular material on the right) with early stages of organization. (C) A hematoxylin and eosin (H and E) stain (original magnification × 400) showing fibrinoid intimal necrosis (right) in the absence of an inflammatory reaction within a small vessel in the same duodenal biopsy as shown in (A) and (B). The ECs show marked vacuolization (left). The surrounding eosinophilic vascular cells are smooth muscle cells surrounded by fibroblasts. (D) A higher power view (original magnification × 600) of the same biopsy showing four capillaries with grossly vacuolated ECs and luminal effacement.

Fig. 7.

EC injury in a case of antiphospholipid antibody syndrome (photographs courtesy of Emma E. Furth, MD, Department of Pathology and Laboratory Medicine, University of Pennsylvania, Philadelphia, PA). EC injury seen in a duodenal biopsy from a 40-year-old woman who presented with profuse intestinal bleeding and was found to have a lupus anticoagulant and a markedly positive anticardiolipin antibody. (A) Elastic stain (original magnification × 200) highlighting a fresh thrombus (right) with the beginning stages of organization and EC ingrowth, an older organized thrombus with fibroblast proliferation (center), within a vessel showing vacuolated, injured, and disrupted ECs. (B) A trichrome stain (original magnification × 400) of the same field highlighting the thrombus material (red acellular material on the right) with early stages of organization. (C) A hematoxylin and eosin (H and E) stain (original magnification × 400) showing fibrinoid intimal necrosis (right) in the absence of an inflammatory reaction within a small vessel in the same duodenal biopsy as shown in (A) and (B). The ECs show marked vacuolization (left). The surrounding eosinophilic vascular cells are smooth muscle cells surrounded by fibroblasts. (D) A higher power view (original magnification × 600) of the same biopsy showing four capillaries with grossly vacuolated ECs and luminal effacement.

Close modal

Address reprint requests to Douglas B. Cines, MD, Department of Pathology and Laboratory Medicine, University of Pennsylvania, 513A Stellar-Chance, 422 Curie Blvd, Philadelphia, PA 19104; e-mail:dcines@mail.med.upenn.edu.

1
Fishman
AP
Endothelium: A distributed organ of diverse capabilities.
Ann NY Acad Sci
401
1982
1
2
Augustin
HG
Kozian
DH
Johnson
RC
Differentiation of endothelial cells: Analysis of the constitutive and activated endothelial cell phenotypes.
Bioessays
16
1994
901
3
Jaffe
EA
Nachman
RL
Becker
CG
Minick
CR
Culture of human endothelial cells derived from umbilical veins: Identification by morphologic criteria.
J Clin Invest
52
1973
2745
4
Gimbrone
MA
Cotran
RS
Folkman
J
Human vascular endothelial cells in culture. Growth and DNA synthesis.
J Cell Biol
60
1974
673
5
Lewis
LJ
Hoak
JC
Maca
RD
Fry
GL
Replication of human endothelial cells in culture.
Science
181
1973
453
6
Fishman
MC
Stainier
DYR
Cardiovascular development: Prospects for a genetic approach.
Circ Res
74
1994
757
7
Hanahan
D
Folkman
J
Patterns and emerging mechanisms of the angiogenic switch during tumorigenesis.
Cell
86
1996
353
8
Risau
W
Flamme
I
Vasculogenesis.
Annu Rev Cell Dev Biol
11
1995
73
9
Palis
J
McGrath
KE
Kingsley
PD
Initiation of hematopoiesis and vasculogenesis in murine yolk sac explants.
Blood
86
1995
156
10
Flamme
I
Baranowski
A
Risau
W
A new model of vasculogenesis and angiogenesis in vitro as compared with vascular growth in the avian area vasculosa.
Anat Rec
237
1993
49
11
Asahara
T
Murohara
T
Sullivan
A
Silver
M
van der Zee
R
Li
T
Witzenbichler
B
Schatteman
G
Isner
JM
Isolation of putative progenitor endothelial cells for angiogenesis.
Science
275
1997
964
12
Risau
W
Differentiation of endothelium.
Endothelium Differ
9
1995
926
13
Buck
CA
Edelman
JM
Buck
CE
Kennedy
G
Baldwin
HS
Expression patterns of adhesion receptors in the developing mouse lung: Function implications.
Cell Adhesion Commun
4
1996
69
14
Pardanaud
L
Yassine
F
Dieterlen-Lievre
F
Relatonship between vasculogenesis, angiogenesis and haemopoiesis during avian ontogeny.
Development
105
1989
473
15
Bielinska
M
Narita
N
Heikinheimo
M
Porter
SB
Wilson
DB
Erythropoiesis and vasculogenesis in embryoid bodies lacking visceral yolk sac endoderm.
Blood
88
1996
3720
16
Beck L Jr, D'Amore P: Vascular development: Cellular and molecular regulation. FASEB J 11:365, 1997
17
Breier
G
Risau
W
The role of vascular endothelial growth factor in blood vessel formation.
Trends Cell Biol
6
1996
454
18
Shalaby
R
Rossant
J
Yamaguchi
TP
Gertsenstein
M
Wu
X-F
Breitman
ML
Schuh
AC
Failure of blood-island formation and vasculogenesis in FLK-1-deficient mice.
Nature
376
1995
62
19
Sato
TN
Tozawa
Y
Deutsch
U
Wolburg-Buchholz
K
Fujiwara
Y
Gendron-Maguire
M
Gridley
T
Wolburg
H
Risau
W
Qin
Y
Distinct roles of the receptor tyrosine kinases Tie-1 and Tie-2 in blood vessel formation.
Nature
376
1995
70
20
Mustonen
T
Alitalo
K
Endothelial receptor tyrosine kinases involved in angiogenesis.
J Cell Biol
129
1995
895
21
Davis
S
Aldrich
TH
Jones
PF
Acheson
A
Compton
DL
Jain
V
Ryan
TE
Bruno
J
Radziejewski
C
Maisonpierre
PC
Yancopoulos
GD
Isolation of angiopoietin-1, a ligand for the TIE2 receptor, by secretion-trap expression cloning.
Cell
87
1996
1161
22
Maisonpierre
P
Suri
G
Jones
P
Bartunkova
S
Wiedgand
S
Radziejewski
C
Compton
D
McClain
J
Aldrich
T
Papadonopoulos
N
Daly
T
Davis
S
Sato
T
Yankopoulos
GD
Angiopoietin-2, a natural antagonist for Tie-2 that disrupts in vivo angiogenesis.
Science
277
1997
55
23
Suri
C
Jones
PF
Patan
S
Bartunkova
S
Maisonpierre
PC
Davis
S
Sato
TN
Yancopoulos
GD
Requisite role of angiopoietin-1, a ligand for the Tie-2 receptor during embryonic angiogenesis.
Cell
87
1996
1171
24
Folkman
J
D'Amore
PA
Blood vessel formation: What is the molecular basis?
Cell
87
1996
1153
25
Vikkula
M
Boou
L
Carraway
K
Calvert
J
DIamond
A
Gooumnerov
B
Pasyk
K
Marchuk
D
Warman
M
Cantley
L
Mulliken
J
Olsen
BR
Vascular dysmorphogenesis caused by an activating mutation in the receptor tyrosine kinase.
Cell
87
1996
1181
26
Stainier
DYR
Fishman
MC
The zebrafish as a model system to study cardiovascular development.
Trends Cardiovasc Med
4
1994
207
27
Sankar
S
Mahooti-Brooks
N
Bensen
L
McCarthy
TL
Centrella
M
Madri
JA
Modulation of transforming growth factor β receptor levels on microvascular endothelial cells during in vitro angiogenesis.
J Clin Invest
97
1996
1436
28
Schwartz
MA
Schaller
MD
Ginsburg
MH
Integrins: emerging paradigms of signal transduction.
Annu Rev Cell Develop Biol
11
1995
549
29
Varner
JA
Brooks
PC
Cheresh
DA
Review: The integrin αvβ3: Angiogenesis and apoptosis.
Cell Adhesion Commun
3
1995
367
30
Hynes
RO
Targeted mutations in cell adhesion genes: What have we learned from them?
Dev Biol
180
1996
402
31
Albelda
SM
Muller
WA
Buck
CA
Newman
PJ
Molecular and cellular properties of PECAM-1 (endoCAM/CD31): A novel vascular cell-cell adhesion molecule.
J Cell Biol
114
1991
1059
32
Dejana
E
Endothelial adherens junctions: Implications in the control of vascular permeability and angiogenesis.
J Clin Invest
98
1996
1949
33
Bretscher
MS
Moving membrane up to the front of migrating cells.
Cell
85
1996
465
34
Carmeliet P, Collen D: Gene targeting and gene transfer studies of the plasminogen/plasmin system: Implications in thrombosis, hemostasis, neointima formation, and atherosclerosis. FASEB J 9:934, 1995
35
Plow EF, Herren T, Redlitz A, Miles LA, Hoover-Plow JL: The cell biology of the plasminogen system. FASEB J 9:939, 1995
36
O'Reilly
MS
Holmgren
L
Shing
Y
Chen
C
Rosenthan
RA
Moses
M
Lane
WS
Cao
Y
Sage
EH
Folkman
J
Angiostatin: A novel angiogenesis inhibitor that mediates the suppression of metastases by a Lewis lung carcinoma.
Cell
79
1994
315
37
O'Reilly
MS
Boehm
T
Shing
Y
Fukai
N
Vasios
G
Lane
WS
Flynn
E
Birkhead
JR
Olsen
BR
Folkman
J
Endostatin: An endogenous inhibitor of angiogenesis and tumor growth.
Cell
88
1997
277
38
Zetter
BR
On target with tumor blood vessels.
Nat Biotechnol
15
1997
1243
39
McCarthy
SA
Kuzu
I
Gatter
KC
Bicknell
R
Heterogeneity of the endothelial cell and its role in organ preference of tumour metastasis.
TIPS
12
1991
462
40
Kumar
S
West
DC
Ager
A
Heterogeneity in endothelial cells from large vessels and microvessels.
Differentiation
36
1987
57
41
Turner
RR
Beckstead
JH
Warnke
RA
Wood
GA
Endothelial cell phenotypic diversity. In situ demonstration of immunologic and enzymatic heterogeneity that correlates with specific morphologic subtypes.
Am J Clin Pathol
87
1985
569
42
Levin
EG
Osborn
KG
The expression of endothelial cell tissue plasminogen activator in vivo: A function defined by vessel size and anatomic location.
J Cell Sci
11
1997
139
43
Wojta
J
Hoover
RL
Daniel
TO
Vascular origin determines plaminogen activator expression in human endothelial cells. Renal endothelial cells produce large amounts of single chain urokinase type plasminogen activator.
J Biol Chem
264
1989
2846
44
Louise
CB
Obrig
TG
Human renal microvascular endothelial cells as a potential target in the development of the hemolytic uremic syndrome as related to fibrinolysis factor expression, in vitro.
Microvasc Res
47
1994
377
45
Obrig
TG
Louise
CB
Lingwood
CA
Boyd
B
Daniel
TO
Endothelial cell heterogeneity in shiga toxin receptors and responses.
J Biol Chem
268
1993
15484
46
Mitra
D
Jaffe
EA
Weksler
B
Hajjar
KA
Doserland
C
Laurence
J
Thrombotic thrombocytopenic purpura and sprodic hemolytic-uremic syndrome plasmas induce apoptosis in restricted lineages of human microvascular endothelial cells.
Blood
89
1997
1224
47
Drake
TA
Cheng
J
Chang
A
Taylor FB Jr
Expression of tissue factor, thrombomodulin, and E-selectin in baboons with lethal Eschericia coli sepsis.
Am J Pathol
142
1993
1458
48
Zhu
D
Cheng
C-F
Pauli
BU
Mediation of lung metastasis of murine melanoma by a lung-specific endothelial cell adhesion molecule.
Proc Natl Acad Sci USA
88
1991
9568
49
Butcher
EC
Picker
LJ
Lymphocyte homing and homeostasis.
Science
272
1996
60
50
Rafii
S
Shapiro
F
Pettengall
R
Ferris
B
Nachman
RL
Moore
MA
Asch
AS
Human bone marrow microvascular endothelial cells support long-term proliferation and differentiation of myeloid and megakaryocyte progenitors.
Blood
86
1995
3353
51
Owman
C
Hardbeo
JE
Functional heterogeneity of cerebrovascular endothelium.
Brain Behav Evol
32
1988
65
52
Grau
GE
de Moreloose
P
Bulla
O
Lou
J
Lei
Z
Reber
R
Mili
N
Ricou
B
Morel
DR
Suter
PM
Haemostatic properties of human pulmonary and cerebral microvascular endothelial cells.
Thromb Haemost
77
1997
585
53
Morin
O
Patry
P
Lafleur
L
Heterogeneity of endothelial cells of adult rat liver as resolved by sedimentation velocity and flow cytometry.
J Cell Physiol
119
1984
327
54
Couveland
A
Scoazec
J-Y
Dauge
M-C
Bringer
A-F
Potet
FGF
Structural and functional differentiation of sinusoidal endothelial cells during liver organogenesis in humans.
Blood
87
1996
4568
55
Milici
AJ
Furie
MB
Carley
WW
The formation of fenestrations and channels by capillary endothelium in vitro.
Proc Natl Acad Sci USA
82
1985
6181
56
Pardanaud
L
Altmann
C
Kitos
P
Dieterlen-Lievre
F
Buck
CA
Vasculogenesis in the early quail blastodisc as studied with a monoclonal antibody recognizing endothelial cells.
Development
100
1987
339
57
Grau
GE
Mili
N
Lou
JN
Morel
DR
Ricou
B
Lucas
R
Suter
PM
Phenotypic and functional analysis of pulmonary microvascular endothelial cells from patients with acute respiratory distress syndrome.
Lab Invest
74
1996
761
58
Aird
WC
Edelberg
JM
Weiler-Guettler
H
Simmons
WW
Smith
TW
Rosenberg
RD
Vascular bed-specific expression of an endothelial cell gene is programmed by the tissue microenviromnent.
J Cell Biol
138
1997
1117
59
Pauli
BU
Lee
C-L
Organ preference of metastasis. The role of organ-specifically modulated endothelial cells.
Lab Invest
58
1988
379
60
Subramanian
VA
Hernandez
Y
Tack-Goldman
K
Grabowski
EF
Weksler
BB
Prostacyclin production by internal mammary artery as a factor in coronary artery bypass grafts.
Surgery
100
1986
376
61
Ades
EW
Candal
FJ
Swerlick
RA
George
VG
Summers
S
Bosse
DC
Lawley
TJ
HMEC-1: Establishment of an immortalized human microvascular endothelial cell line.
J Invest Dermatol
99
1992
683
62
Grant
DS
Tashiro
K-I
Segui-Real
B
Yamada
Y
Martin
GR
Kleinman
HK
Two different laminin domains mediate the differentiation of human endothelial cells into capillary-like structures in vitro.
Cell
58
1989
933
63
Gibbons
GH
Dzau
VJ
Molecular therapies for vascular diseases.
Science
272
1996
689
64
Stamler
JS
Singel
DJ
Loscalzo
J
Biochemistry of nitric oxide and its redox-activated forms.
Science
258
1992
1898
65
Venema
RC
Sayegh
HS
Arnal
JF
Harrison
DG
Role of the enzyme calmodulin-binding domain in membrane association and phospholipid inhibition of endothelial nitric oxide synthase.
J Biol Chem
270
1995
14705
66
Topper
JN
Cai
J
Falb
D
Gimbrone MA Jr
Identification of vascular endothelial genes differentially responsive to fluid mechanical stimuli: Cyclooxygenase-3, manganese superoxide dismutase, and endothelial cell nitric oxide synthase are selectively up-regulated by steady laminar shear stress.
Proc Natl Acad Sci USA
93
1996
10417
67
Awolesi
MA
Widmann
MD
Sessa
WC
Sumpio
BE
Cyclic strain increases endothelial nitric oxide synthase activity.
Surgery
116
1994
439
68
Loscalzo
J
Vita
JA
Ischemia, hyperemia, exercise, and NO: Complex physiology and complex molecular adaptations.
Circulation
90
1994
2556
69
Venema
RC
Nishida
K
Alexander
RW
Harrison
DG
Murphy
TJ
Organization of the bovine gene encoding the endothelial nitric oxide synthase.
Biochem Biophys Acta
1218
1994
413
70
Balligand
JL
Ungureanu-Longroid
D
Simmons
WW
Kobzik
L
Lowenstein
CJ
Lamas
S
Kelly
RA
Smith
TW
Michel
T
Induction of nitric oxide synthase in rat cardiac microvascular endothelial cells by IL-beta and IFN-gamma.
Am J Physiol
268
1995
H1293
71
Loscalzo
J
Welch
G
Nitric oxide and its role in the cardiovascular system.
Prog Cardiovasc Dis
38
1995
87
72
Mendelsohn
ME
O'Neill
S
George
D
Loscalzo
J
Inhibition of fibrinogen binding to human platelets by S-nitroso-N-acetylcysteine.
J Biol Chem
265
1990
19028
73
Radomski
MW
Palmer
RMJ
Moncada
S
Comparative pharmacology of endothelium-derived relaxing factor, nitric oxide, and prostacyclin in platelets.
Br J Pharmacol
92
1987
181
74
Stamler
J
Vaughan
DE
Loscalzo
J
Synergistic disaggregation of platelets by tissue plasminogen activator, prostaglandin E1, and nitroglycerin.
Circ Res
65
1989
769
75
Michelson
AD
Benoit
SE
Furman
MI
Breckwoldt
WL
Rohrer
MJ
Barnard
MR
Loscalzo
J
Effects of endothelium-derived relaxing factor/nitric oxide on platelet surface glycoproteins.
Am J Physiol
39
1996
H1640
76
Pigazzi A, Folli F, Heydrich S, Loscalzo J: Nitric oxide inhibits thrombin-induced phosphoinositide 3-kinase activity in human platelets. Circulation (in press)
77
Kubes
P
Suzuki
M
Granger
DN
Nitric oxide: An endogenous modulator of leukocyte adhesion.
Proc Natl Acad Sci USA
88
1991
4651
78
De Caterina
R
Libby
P
Peng
HB
Thannickal
VJ
Rajavashisth
TB
Gimbrone MA Jr
Shin
WS
Liao
JK
Nitric oxide decreases cytokine-induced endothelial activation. Nitric oxide selectively reduces endothelial expression of adhesion molecules and proinflammatory cytokines.
J Clin Invest
96
1995
60
79
Marks
DS
Vita
JA
Folts
JD
Keaney JF Jr
Welch
GN
Loscalzo
J
Inhibition of neointimal proliferation in rabbits following vascular injury by a single treatment with a protein adduct of nitric oxide.
J Clin Invest
96
1995
2630
80
Garg
UC
Hassid
A
Nitric oxide (NO) and 8-bromo-cyclic GMP inhibit mitogenesis and proliferation of cultured rat vascular smooth muscle cells.
J Clin Invest
83
1989
1974
81
Fukuo
K
Inoue
T
Morimoto
S
Nakahasi
T
Yasuda
O
Kitanno
S
Sasanda
A
Ogihara
T
Nitric oxide mediates cytoxicity and basic fibroblast growth factor release in cultured vascular smooth muscle cells.
J Clin Invest
95
1995
669
82
Garland
CJ
Plane
F
Kemp
BK
Cocks
TM
Endothelium-dependent hyperpolarization: A role in the control of vascular tone.
Trends Biol Sci
16
1995
23
83
Feletou
M
Vanhoutte
PM
Endothelium-dependent hyperpolarization of canine coronary smooth muscle.
Br J Pharmacol
93
1988
515
84
Levin
ER
Endothelins.
N Engl J Med
333
1996
356
85
Nakamura
S
Naruse
M
Maruse
K
Demura
H
Uemura
H
Immunocytochemical localization of endothelin in cultured bovine endothelial cell.
Histochemistry
94
1990
475
86
Simonson MS, Dunn MJ: Cellular signaling by peptides of the endothelin gene family. FASEB J 4:2989, 1990
87
Goligorsky
MS
Tsukahara
H
Magazine
H
Andersen
TT
Malik
AB
Bahou
WF
Termination of endothelin signaling: Role of nitric oxide.
J Cell Physiol
158
1994
485
88
Lopez
JA
Armstrong
ML
Piegors
DJ
Heistad
DD
Vascular responses to endothelin-1 in atherosclerotic primates.
Arteriosclerosis
10
1990
1113
89
Zimmerman
GA
Whatley
RE
Benson
DE
McIntyre
TM
Prescott
SM
Endothelial cells for studies of platelet-activating factor and arachidonate metabolites.
Methods Enzymol
187
1990
520
90
Meade
EA
Jones
DA
Zimmerman
GA
McIntyre
TM
Prescott
SM
Prostaglandins and related compounds. Lipid messengers with many actions.
Handbook Lipid Res
8
1996
285
91
Whatley
RE
Zimmerman
GA
Prescott
SM
McIntyre
TM
Platelet-activating factor and PAF-like mimetics.
Handbook Lipid Res
8
1996
239
92
Imaizumi
T
Stafforini
DM
Yamada
Y
McIntyre
TM
Prescott
SM
Zimmerman
GA
Platelet-activating factor: A mediator for clinicians.
J Intern Med
238
1995
5
93
Campbell WB, Halushka PV: Eicosanoids and platelet-activating factor, in Hardman JG, Limbird LE, Molinoff PM, Ruddon RW, Gilman AG (eds): Goodman and Gilman's The Pharmacologic Basis of Therapeutics. New York, NY, McGraw-Hill, 1996, p 601
94
Lorant
DE
Zimmerman
GA
McIntyre
TM
Prescott
SM
Platelet-activating factor mediates procoagulant activity on the surface of endothelial cells by promoting leukocyte adhesion.
Semin Cell Biol
6
1995
295
95
Weksler
BB
Marcus
AJ
Jaffe
EA
Synthesis of prostaglandin I2 (prostacyclin) by cultured human and bovine endothelial cells.
Proc Natl Acad Sci USA
74
1977
3922
96
McIntyre
TM
Zimmerman
GA
Satoh
K
Prescott
SM
Cultured endothelial cells synthesize both platelet-activating factor and prostacyclin in response to histamine, bradykinin and ATP.
J Clin Invest
76
1985
271
97
Coleman
RA
Smith
WL
Narumiya
S
Classification of prostanoid receptors: Properties, distribution, and structure of the receptors and their subtypes.
Pharmacol Rev
46
1994
205
98
Majerus
PW
Arachidonate metabolism in vascular disorders.
J Clin Invest
72
1983
1521
99
FitzGerald
GA
Pedersen
AK
Patrono
C
Analysis of prostacyclin and thromboxane biosynthesis in cardiovascular disease.
Circulation
67
1983
1174
100
Chen
X-S
Sheller
JR
Johnson
EN
Funk
CD
Role of leukotrienes revealed by targeted distruption of the 5-lipoxygenase gene.
Nature
372
1992
179
101
Bombeli
T
Meller
M
Haeberli
A
Anticoaguant properties of the vascular endothelium.
Thromb Haemost
77
1997
408
102
Rosenberg
RD
Rosenberg
JS
Natural anticoagulant mechanisms.
J Clin Invest
74
1984
1
103
Bauer
KA
Weiss
LM
Sparrow
D
Vakonas
PS
Rosenberg
RD
Aging-associated changes in indices of thrombin generation and protein C activation in human. Normative aging study.
J Clin Invest
80
1987
1527
104
Marcum
JA
Rosenberg
RD
Anticoagulantly active heparin-like molecules from vascular tissue.
Biochemistry
23
1984
1730
105
Tollefsen
DM
Pestka
CA
Heparin cofactor II activity in patients with disseminated intravascular coagulation and hepatic failure.
Blood
66
1985
769
106
Broze GJ Jr
Tissue factor pathway inhibitor.
Thromb Haemost
74
1995
90
107
Jesty
J
Lorenz
A
Rodriguez
J
Wun
T-C
Initiation of the tissue factor pathway of coagulation in the presence of heparin: Control by antithrombin III and tissue factor pathway inhibitor.
Blood
87
1996
2301
108
Bombeli
T
Karsan
A
Tait
JF
Harlan
JM
Apoptotic vascular endothelial cells become procoagulant.
Blood
90
1997
2429
109
Esmon
CT
Fukudome
K
Cellular regulation of the protein C pathway.
Semin Cell Biol
6
1995
259
110
Fair
DS
Marlar
RA
Levin
EG
Human endothelial cells synthesize protein S.
Blood
67
1986
1168
111
Fukudome
K
Kurosawa
S
Stearns-Kurosawa
D-J
He
X
Rezaie
AR
Esmon
CT
The endothelial cell protein C receptor. Cell surface expression and direct ligand binding by the soluble receptor.
J Biol Chem
271
1996
17491
112
Thompson
EA
Salem
HH
Inhibition by human thrombomodulin of factor Xa-mediated cleavage of prothrombin.
J Clin Invest
78
1986
13
113
Esmon
CT
Molecular events that control the protein C anticoagulant pathway.
Thromb Haemost
70
1993
29
114
Nawroth
PP
Stern
DM
Modulation of endothelial cell hemostatic properties by tumor necrosis factor.
J Exp Med
163
1986
740
115
Mantovani
A
Sozzani
S
Vecchi
A
Introna
M
Allavena
P
Cytokine activation of endothelial cells: new molecules for an old paradigm.
Thromb Haemost
78
1997
406
116
Cucurull
E
Gharavi
AE
Thrombomodulin: A new frontier in lupus research?
Clin Exp Rheumatol
15
1997
1
117
Drake
TA
Morrissey
JH
Edgington
TS
Selective cellular expression of tissue factor in human tissues. Implications for disorders of hemostasis and thrombosis.
Am J Pathol
134
1989
1087
118
Carmeliet
P
Mackman
N
Moons
L
Luther
T
Gressens
P
van Vlaenderen
I
Demunck
H
Kaspar
M
Breier
G
Evrard
P
Muller
M
Risau
W
Edginston
T
Collen
D
Role of tissue factor in embryonic blood vessel development.
Nature
383
1996
73
119
Bugge
TH
Xiao
Q
Kormbrinck
KW
Flick
MJ
Holmback
K
Danton
MJS
Colbert
MC
Witte
DP
Fujikawa
K
Davie
E
Degen
JL
Fatal embryonic bleeding events in mice lacking tissue factor, the cell-associated initiation of blood coagulation.
Proc Natl Acad Sci USA
93
1996
6258
120
Toomey
JR
Kratzer
KE
Lasky
NM
Stanton
JJ
Broze GJ Jr
Targeted disruption of the murine tissue factor gene results in embryonic lethality.
Blood
88
1996
1583
121
Rapaport
SI
Rao
VM
The tissue factor pathway: How it has become a “Prima Ballerina”.
Thromb Haemost
74
1995
7
122
Nemerson
Y
Tissue factor: Then and now.
Thromb Haemost
74
1995
180
123
Ryan
J
Brett
J
Tijburg
P
Bach
RR
Kisiel
W
Stern
D
Tumor necrosis factor-induced endothelial tissue factor is associated with subendothelial matrix vesicles but is not expressed on the apical surface.
Blood
80
1992
966
124
Narahara
N
Enden
T
Wiiger
M
Prydz
H
Polar expression of tissue factor in human umbilical vein endothelial cells.
Arterioscler Thromb
14
1994
1815
125
Bona
R
Lee
E
Rickles
F
Tissue factor apoprotein: Intracellular transport and expression in shed membrane vesicles.
Thromb Res
48
1987
487
126
Shimura
M
Wada
H
Waskita
Y
Nakse
T
Hiyoyama
K
Nagaya
S
Shiku
H
Plasma tissue factor and tissue factor pathway inhibitor levels in patients with disseminated intravascular coagulation.
Am J Pathol
52
1996
165
127
Marmur
JD
Rossikhina
M
Guha
A
Fyle
B
Friedrich
V
Mendlowitz
M
Nemerson
Y
Taubman
MB
Tissue factor is rapidly induced in arterial smooth muscle after balloon injury.
J Clin Invest
91
1993
2253
128
Thiruvikraman
SV
Guha
A
Roboz
J
Taubman
MB
Nemerson
Y
Fallon
JT
In situ localization of tissue factor in human atherosclerotic plaques by binding of digoxigenin-labeled factors VIIa and X.
Lab Invest
75
1996
451
129
Hatakeyama
K
Asada
Y
Marutsuka
K
Sato
Y
Kamikubo
Y
Sumiyoashi
A
Localization and activity of tissue factor in human aortic atherosclerotic lesions.
Atherosclerosis
133
1997
213
130
Contrino
J
Hair
G
Kreutzer
DL
Rickles
FR
In situ detection of tissue factor in vascular endothelial cells: Correlation with the malignant phenotype of human breast disease.
Nat Med
2
1996
209
131
Sato
Y
Asada
Y
Marutsuka
K
Hatakeyama
K
Sumiyoshi
A
Tissue factor induces migration of cultured aortic smooth muscle cells.
Thromb Haemost
75
1996
389
132
Zhang
Y
Deng
Y
Luther
T
Muller
M
Ziegler
R
Waldherr
R
Stern
DM
Nawroth
PP
Tissue factor controls the balance of angiogenic and antiangiogenic properties of tumor cells in mice.
J Clin Invest
94
1994
1320
133
Stern
D
Nawroth
P
Handley
D
Kisiel
W
An endothelial cell-dependent pathway of coagulation.
Proc Natl Acad Sci USA
82
1985
2523
134
Brinkman
H-JM
Mertens
K
Holthius
J
Zwart-Huinink
LA
Grijm
K
van Mourik
JA
The activation of blood coagulation factor X on the surface of endothelial cells: A comparison with various vascular cells, platelets and monocytes.
Br J Haematol
87
1994
332
135
Cheung
WF
van den Born
J
Kuhn
K
Kjellen
L
Hudson
BG
Stafford
DW
Identification of the endothelial cell binding site for factor IX.
Proc Natl Acad Sci USA
80
1996
7001
136
Altieri DC: Xa receptor EPR-1. FASEB J 9:860, 1995
137
Bono
F
Herault
J-P
Avril
C
Schaeffer
P
Lormeau
J-C
Herbert
J-M
Human umbilical vein endothelial cells express high affinity receptors for factor Xa.
J Cell Physiol
172
1997
36
138
Etingen
OR
Silverstein
RL
Hajjar
DP
Identification of a monocyte receptor on herpes-infected endothelial cells.
Proc Natl Acad Sci USA
88
1991
7200
139
Schmaier
AH
Kuo
A
Cines
DB
High molecular weight kininogen expression on human endothelial cells.
J Biol Chem
31
1988
16327
140
Woolkanis
MJ
DeMelfi TM Jr
Blanchard
N
Hoxie
JA
Brass
LF
Regulation of thrombin receptors on HUVECs.
J Biol Chem
270
1995
9686
141
Kanthou
C
Benzakour
O
Cellular effects of thrombin and their signalling pathways.
Cell Pharmacol
2
1995
293
142
Garcia
JGN
Pavalko
FM
Patterson
CE
Vascular endothelial cell activation and permeability responses to thrombin.
Blood Coag Fibrinol
6
1995
609
143
Connolly
AJ
Ishihara
H
Kahn
MI
Ferese RV Jr
Coughlin
SR
Role of the thrombin receptor in development and evidence for a second receptor.
Nature
381
1996
516
144
Mirtza
H
Yatsula
V
Bahou
WF
The proteinase activated receptor-2 (PAR-2) mediates mitogenic responses in human vascular endothelial cells. Molecular characterization and evidence for functional coupling to the thrombin receptor.
J Clin Invest
97
1996
1705
145
Ishihara
H
Connoly
AJ
Zeng
D
Kahn
ML
Zheng
YW
Timmons
C
Tram
T
Coughlin
SR
Protease-activated receptor 3 is a second thrombin receptor in humans.
Nature
386
1997
502
146
Brass
LF
Molino
M
Protease-activated G protein-coupled receptors on human platelets and endothelial cells.
Thromb Haemost
78
1997
234
147
Odrijn
TM
Francis
CW
Sporn
LA
Bunce
LA
Marder
VJ
Simpson-Haidaris
PJ
Heparin-binding domain of fibrin mediates its binding to endothelial cells.
Arterioscler Thromb Vasc Biol
16
1996
1544
148
Erban
JK
Wagner
DD
A 130-kDa protein on endothelial cells binds to amino acids 15-42 of the Bβ chain of fibrinogen.
J Biol Chem
267
1992
2451
149
Greenburg CS, Birckbichler PJ, Rice RH: Transglutaminases: Multifunctional cross-linking enzymes that stabilize tissues. FASEB J 5:3071, 1991
150
Okada
Y
Copeland
BR
Hamann
GF
Koziol
JA
Cheresh
DA
Del Zoppo
GJ
Integrin αvβ3 is expressed in selected microvessels after focal cerebral ischemia.
Am J Pathol
149
1996
37
151
Ruosiahti
E
Engvail
E
Integrins and vascular extracellular matrix assembly.
J Clin Invest
99
1997
1149
152
Wagner
DD
Bonfanti
R
Von Willebrand factor and the endothelium.
Mayo Clin Proc
66
1991
621
153
Rajagopalan
V
Essex
DW
Shapiro
SS
Konkle
BA
Tumor necrosis factor-α modulation of glycoprotein Ibα expression in human endothelial and erythroleukemia cells.
Blood
80
1992
153
154
(abstr, suppl)
Beacham
DA
Tran
LP
Ludlow
LB
Huang
R
Konkle
BA
Shapiro
SS
Endothelial glycoprotein Ibα expression is upregulated by arterial shear stress.
Blood
86
1995
2188a
155
Taylor FB Jr
Studies on the inflammatory-coagulant axis in the baboon response to E. coli: Regulatory role of proteins C, S, C4bBP and of inhibitors of tissue factor.
Prog Clin Biol Res
388
1994
175
156
Hirsh J, Salzman EW, Marder VJ, Colman RW: Overview of the thrombotic process and its therapy, in Colman RW, Hirsh J, Marder VJ, Salzman EW (eds): Hemostasis and Thrombosis. Basic Principles and Clinical Pratice. Philadelphia, PA, Lippincott, 1994, p 1151
157
Todd
AS
Histological localization of fibrinolysin activator.
J Pathol Bacteriol
78
1959
281
158
Pandolfi
M
Histochemistry of tissue plasminogen activator.
Thromb Diath Haemorrh
34
1975
661
159
Levin
EG
Marotti
KR
Santell
L
Protein kinase C and the stimulation of tissue plasminogen activator release from human endothelial cells.
J Biol Chem
264
1989
16030
160
Cugno
M
Uziel
L
Fabreizi
I
Bottasso
B
Maggiolini
F
Agostoni
A
Fibrinolytic response in normal subjects to venous occlusion and DDAVP infusion.
Thromb Res
56
1989
625
161
Levin
EG
Santell
L
Thrombin- and histamine-induced signal transduction in human endothelial cells. Stimulation and agonist-dependent desensitization of protein phosphorylation.
J Biol Chem
266
1991
174
162
McCrae
KR
Aceto
JF
Cines
DB
Clinical aspects of the plasminogen activator system.
Adv Pathol Lab Med
7
1994
467
163
Bacharach
E
Itin
A
Keshet
E
In vivo patterns of expression of urokinase and its inhibitor PAI-1 suggest a concerted role in regulating physiologic angiogenesis.
Proc Natl Acad Sci USA
89
1992
10686
164
Carmeliet
P
Schoonjans
L
Kieckens
L
Ream
B
Degan
J
Bronson
R
De Vos
R
van den Oord
JJ
Collen
D
Mulligan
RC
Physiological consequences of loss of plasminogen activator gene function in mice.
Nature
369
1994
419
165
Yamamoto
K
Loskutoff
DJ
Fibrin deposition in tissues from endotoxin-treated mice correlates with decreases in the expression of urokinase-type but not tissue type plasminogen activator.
J Clin Invest
97
1996
2440
166
Barnathan
ES
Kuo
A
van der Keyl
H
McCrae
KM
Cines
DB
Binding of tissue type plasminogen activator to human endothelial cells: Evidence for two distinct binding sites.
J Biol Chem
263
1988
7792
167
Beebe
DP
Miles
LA
Plow
EF
A linear amino acid sequence involved in the interaction of t-PA with its endothelial cell receptor.
Blood
74
1989
2034
168
Fukao
H
Hagiya
Y
Nonaka
T
Okada
K
Matsuo
O
Analysis of binding protein for tissue-type plasminogen activator in human endothelial cells.
Biochem Biophys Res Commun
187
1992
956
169
Ramakrishnan
V
Sinicropi
DV
Dere
R
Darbonne
WC
Bechtol
KB
Baker
JB
Interaction of wild-type and catalytically inactive mutant forms of tissue-type plasminogen activator with human umbilical vein endothelial cell monolayers.
J Biol Chem
265
1990
2755
170
Cheng
X-F
Pohl
G
Back
O
Wallen
P
Characterization of receptors interacting specially with the B-chain of tissue plasminogen activator on endothelial cells.
Fibrinolysis
10
1996
167
171
Hajjar
KA
Hamel
NM
Harpel
PC
Nachman
RL
Binding of tissue plasminogen activator to cultured human endothelial cells.
J Clin Invest
80
1987
1712
172
Hajjar
KA
Hamel
NM
Identification and characterization of human endothelial cell membrane binding sites for tissue plasminogen activator and urokinase.
J Biol Chem
265
1990
2908
173
Weling
TH
Huber
TS
Messina
LM
Stanley
JC
Tissue plasminogen activator increases canine endothelial cell proliferation rate through a plasmin-independent, receptor-mediated mechanism.
J Surg Res
66
1996
36
174
Hajjar
KA
Jacovina
AT
Checko
J
An endothelial cell receptor for plasminogen/tissue plasminogen activator.
J Biol Chem
269
1994
21191
175
Barnathan
ES
Kuo
A
Rosenfeld
L
Kariko
K
Leski
M
Robbiati
F
Nolli
ML
Henkin
J
Cines
DB
Interaction of single chain urokinase type plasminogen activator with human endothelial cells.
J Biol Chem
265
1990
2865
176
Barnathan
ES
Kuo
A
Kariko
K
Rosenfeld
L
Murray
SC
Behrendt
N
Ronne
E
Weiner
D
Henkin
J
Cines
DB
Characterization of human endothelial cell urokinase-type plasminogen activator receptor protein and mRNA.
Blood
76
1990
1795
177
Wun
TC
Schleuning
W-D
Reich
E
Isolation and characterization of urokinase from human plasma.
J Biol Chem
257
1982
3276
178
Manchanda
N
Schwartz
BS
Single chain urokinase. Augmentation of enzymatic activity upon binding to monocytes.
J Biol Chem
266
1991
14580
179
Schwartz
BS
Differential inhibition of soluble and cell surface receptor bound scu-PA by plasminogen activator inhibitor.
J Biol Chem
269
1994
8319
180
Higazi
AA-R
Mazar
A
Wang
J
Reilly
R
Henkin
J
Kniss
D
CInes
D
Single chain urokinase-type plasminogen activator bound to its receptor is relatively resistant to plasminogen activator inhibitor type 1.
Blood
87
1996
3545
181
Pepper
MS
Sappino
A-P
Stocklin
R
Montesano
R
Orci
L
Vassalli
J-D
Upregulation of urokinase receptor expression on migrating endothelial cells.
J Cell Biol
122
1993
673
182
Bugge
TH
Flick
MJ
Danton
MJS
Daugherty
CC
Romer
J
Dano
K
Carmeliet
P
Collen
D
Degen
JL
Urokinase-type plasminogen activator is effective in fibrin clearance in the absence of its receptor or tissue-type plasminogen activator.
Proc Natl Acad Sci USA
93
1996
5899
183
Miles
LA
Levin
EG
Plescia
J
Collen
D
Plow
EF
Plasminogen receptors, urokinase receptors, and their modulation on human endothelial cells.
Blood
72
1988
628
184
Plow
EF
Freaney
DE
Plescia
J
Miles
LA
The plasminogen system and cell surfaces: Evidence for plasminogen and urokinase receptors on the same cell type.
J Cell Biol
103
1986
2411
185
Ploplis
VA
Carmeliet
P
Vazirzadeh
S
van Vlaenderen
I
Moons
L
Plow
EF
Collen
D
Effects of disruption of the plasminogen gene on thrombosis, growth, and health in mice.
Circulation
92
1995
2585
186
Miles
LA
Fless
GM
Levin
EG
Scanu
AM
Plow
EF
A potential basis for the thrombotic risks associated with lipoprotein(a).
Nature
339
1989
301
187
Grainger
DJ
Kemp
PR
Metcalfe
JC
Liu
AC
Lawn
RM
Williams
NR
Grace
AA
Schofield
PM
Chauhan
A
The serum concentration of active transforming growth factor-β is severely depressed in advanced atherosclerosis.
Nat Med
1
1995
74
188
Levin
EG
Santell
L
Association of a plasminogen activator inhibitor (PAI-1) with the growth substratum and membrane of human endothelial cells.
J Cell Biol
105
1987
2543
189
Loskutoff DJ, Sawdey M, Mimuro J: Type 1 plasminogen activator inhibitor, in Coller BS (ed): Progress in Hemostasis and Thrombosis. Philadelphia, PA, Saunders, 1989, p 87
190
Sawdy
MS
Loskutoff
DJ
Regulation of murine type 1 plasminogen activator inhibitor gene in vivo. Tissue specificity and induction by lipopolysaccharide, tumor necrosis factor-α, and transforming growth factor-β.
J Clin Invest
88
1991
1346
191
Astedt
B
Hagerstrand
I
Lecander
I
Cellular localization in placenta of placental type plasminogen activator inhibitor.
Thromb Haemost
56
1986
63
192
Scarpati
EM
Sadler
JE
Regulation of endothelial cells coagulant properties.
J Biol Chem
264
1989
20705
193
Suzuki
K
Nishioka
J
Hashimoto
S
Protein C inhibitor. Purification from human plasma and characterization.
J Biol Chem
258
1983
163
194
Geiger
M
Prilinger
U
Griffin
JH
Binder
BR
Urinary protein C inhibitor. Glycosaminoglycans synthesized by the epithelial cell line TC1-598 enhance its interaction with urokinase.
J Biol Chem
266
1991
11851
195
Bajzar
L
Morser
J
Nesheim
M
TAFI, or plasma procarboxypeptidase B, couples the coagulation and fibrinolytic cascades through the thrombin-thrombomodulin complex.
J Biol Chem
271
1996
16603
196
Romano
M
Sironi
M
Toniatti
C
Polentarutti
N
Fruscella
P
Ghezzi
P
van Hinsbergh
V
Sozzani
S
Bussolino
F
Poli
V
Cilberto
G
Mantovani
A
Role of IL-6 and its soluble receptor in induction of chemokines and leukocyte recruitment.
Immunity
6
1997
315
197
Middleton
J
Neil
S
Wintle
J
Clark-Lewis
I
Moore
H
Lam
C
Auer
M
Hub
E
Rot
A
Transcytosis and surface presentation of IL-8 by venular endothelial cells.
Cell
91
1997
385
198
Schafer
AI
Vascular endothelium: In defense of blood fluidity.
J Clin Invest
99
1997
1143
199
Marcus
AJ
Broekman
MJ
Drosopoulos
JHF
Islam
N
Alyonycheva
TN
Safier
LB
Hajjar
KA
Posnett
DN
Shoenborn
MA
Schoolet
KA
Gayle
RB
Maliszewski
CR
The endothelial cell ecto-ADPase responsible for inhibition of platelet function is CD39.
J Clin Invest
99
1997
1351
200
Roth
GJ
Platelets and blood vessels: The adhesion event.
Immunol Today
13
1992
100
201
Ruggeri Z, Ware J: von Willebrand factor. FASEB J 7:308, 1993
202
Dong
J-f
Hyun
W
Lopez
JA
Aggregation of mammalian cells expressing the platelet glycoprotein (GP) Ib-IX complex and the requirement for tyrosine sulfation of GPIbα.
Blood
86
1995
4175
203
Lopez
JA
The platelet glycoprotein Ib-IX complex.
Blood Coagul Fibrinol
5
1994
97
204
Savage
B
Saldivar
E
Ruggeri
ZM
Initiation of platelet adhesion by arrest onto fibrinogen or translocation on von Willebrand factor.
Cell
84
1996
289
205
Kroll
MH
Hellums
JD
McIntire
LV
Schafer
AI
Moake
JL
Platelets and shear stress.
Blood
88
1996
1525
206
Calvete
JJ
Clues for understanding the structure and function of a prototypic human integrin: The platelet glycoprotein IIb/IIIa complex.
Thromb Haemost
72
1994
1
207
McEver
RP
Moore
KL
Cummings
RD
Leukocyte trafficking mediated by selectin-carbohydrate interactions.
J Biol Chem
270
1995
11025
208
Walcheck
B
Moore
KL
McEver
RP
Kishimoto
TK
Neutrophil-neutrophil interactions under hydrodynamic shear stress involve L-selectin and PSGL-1: A mechanism that amplifies initial leukocyte accumulation on P-selectin in vitro.
J Clin Invest
98
1996
1081
209
Buttrum
SM
Hatton
R
Nash
GB
Selectin-mediated rolling of neutrophils on immobilized platelets.
Blood
82
1993
1165
210
Palabrica
T
Lobb
R
Furie
BC
Aronovitz
M
Benjamin
C
Hsu
Y-M
Sajer
SA
Furie
B
Leukocyte accumulation promoting fibrin deposition is mediated in vivo by P-selectin on adherent platelets.
Nature
359
1992
848
211
Diacovo
TG
Puri
KD
Warnock
RA
Springer
TA
Von Andrian
UH
Platelet-mediated lymphocyte delivery to high endothelial venules.
Science
273
1996
252
212
Hemmerich
S
Leffler
H
Rosen
SD
Structure of the O-glycans in GlyCAM-1, an endothelial-derived ligand for L-selectin.
J Biol Chem
270
1995
12035
213
Wilkins
PP
Moore
KL
McEver
RP
Cummings
RD
Tyrosine sulfation of P-selectin glycoprotein ligand-1 is required for high affinity binding to P-selectin.
J Biol Chem
270
1995
22677
214
Wilkins
PP
McEver
RP
Cummings
RD
Structures of the O-glycans on P-selectin glycoprotein ligand-1 from HL-60 cells.
J Biol Chem
271
1996
18732
215
McEver
RP
Cummings
RD
Role of PSGL-1 binding to selectins in leukocyte recruitment.
J Clin Invest
100
1997
485
216
Berlin
C
Bargatz
RF
Campbell
JJ
Von Andrian
UH
Szabo
MC
Hasslen
SR
Nelson
RD
Berg
EL
Erlandsen
SL
Butcher
EC
α4 integrins mediate lymphocyte attachment and rolling under physiologic flow.
Cell
80
1995
413
217
Alon
R
Kassner
PD
Carr
MW
Finger
EB
Hemler
ME
Springer
TA
The integrin VLA-4 supports tethering and rolling in flow on VCAM-1.
J Cell Biol
128
1995
1243
218
DeGrendele
HC
Estess
P
Picker
LJ
Siegelman
MH
CD44 and its ligand hyaluronate mediate rolling under physiologic flow: A novel lymphocyte-endothelial cell primary adhesion pathway.
J Exp Med
183
1996
1119
219
Alon
R
Hammer
DA
Springer
TA
Lifetime of the P-selectin: Carbohydrate bond and its response to tensile force in hydrodynamic flow.
Nature
374
1995
539
220
Finger
EB
Puri
KD
Alon
R
Lawrence
MB
Von Andrian
UH
Springer
TA
Adhesion through L-selectin requires a threshold hydrodynamic shear.
Nature
379
1996
266
221
Walcheck
B
Kahn
J
Fisher
JM
Wang
BB
Fisk
RS
Payan
DG
Feehan
C
Betageri
R
Darlak
K
Spatola
AF
Kishimoto
TK
Neutrophil rolling altered by inhibition of L-selectin shedding in vitro.
Nature
380
1996
720
222
Patel
KD
Nollert
MU
McEver
RP
P-selectin must extend a sufficient length from the plasma membrane to mediate rolling of neutrophils.
J Cell Biol
131
1995
1893
223
Weyrich
AS
McIntyre
TM
McEver
RP
Prescott
SM
Zimmerman
GA
Monocyte tethering by P-selectin regulates monocyte chemotactic protein-1 and tumor necrosis factor-α secretion.
J Clin Invest
95
1995
2297
224
Springer
TA
Traffic signals on endothelium for lymphocyte recirculation and leukocyte emigration.
Annu Rev Physiol
57
1995
827
225
Languino
LR
Plescia
J
Duperray
A
Brian
AA
Plow
EF
Geltosky
JE
Altieri
DC
Fibrinogen mediates leukocyte adhesion to vascular endothelium through an ICAM-1-dependent pathway.
Cell
73
1993
1423
226
Muller
WA
Weigl
SA
Monocyte-selective transendothelial migration: Dissection of the binding and transmigration phases by an in vitro assay.
J Exp Med
176
1992
819
227
Dejana E, Corada M, Lampugnani MG: Endothelial cell-to-cell junctions. FASEB J 9:910, 1995
228
Johnson
RC
Chapman
SM
Dong
ZM
Ordovas
JM
Mayadas
TN
Herz
J
Hynes
RO
Schaefer
EJ
Wagner
DD
Absence of P-selectin delays fatty streak formation in mice.
J Clin Invest
99
1997
1037
229
Anderson
DC
Springer
TA
Leukocyte adhesion deficiency: An inherited defect in the Mac-1, LFA-1, and p150,95 glycoproteins.
Annu Rev Med
38
1987
175
230
Etzioni
A
Frydman
M
Pollack
S
Avidor
I
Phillips
ML
Paulson
JC
Gershoni-Baruch
R
Brief report: Recurrent severe infections caused by a novel leukocyte adhesion deficiency.
N Engl J Med
327
1992
1789
231
Ley
K
Tedder
TF
Leukocyte interactions with vascular endothelium: New insights into selectin-mediated attachment and rolling.
J Immunol
155
1995
525
232
Frenette
PS
Mayadas
TN
Rayburn
H
Hynes
RO
Wagner
DD
Susceptibility to infection and altered hematopoiesis and mice deficient in both P- and E-selectin.
Cell
84
1996
563
233
Bullard
DC
Kunkel
EJ
Kubo
H
Hicks
MJ
Lorenzo
I
Doyle
NA
Doerschuk
CM
Ley
K
Beaudet
AL
Infectious susceptibility and severe deficiency of leukocyte rolling and recruitment in E-selectin and P-selectin double mutant mice.
J Exp Med
183
1996
2329
234
Frenette
PS
Mayadas
TN
Rayburn
H
Hynes
RO
Wagner
DD
Susceptibility to infection and altered hematopoiesis in mice deficient in both P- and E-selectins.
Cell
84
1996
563
235
Maly
P
Thall
AD
Petryniak
B
Rogers
CE
Smith
PL
Marks
RM
Kelly
RJ
Gersten
KM
Cheng
G
Saunders
TL
Camper
SA
Camphausen
RT
Sullivan
FX
Isogai
Y
Hindsgaul
O
von Andrian
UH
Lowe
JB
The α(1,3)fucosyltranferase Fuc-TVII controls leukocyte trafficking through an essential role in L-, E-, and P-selectin ligand biosynthesis.
Cell
86
1996
643
236
Kunkel
EJ
Jung
U
Bullard
DC
Norman
KE
Wolitzky
BA
Vestweber
D
Beaudet
AL
Ley
K
Absence of trauma-induced leukocyte rolling in mice deficient in both P-selectin and intercellular adhesion molecule 1.
J Exp Med
183
1996
571
237
Schmits
R
Kundig
TM
Baker
IM
Shumaker
G
Simard
JJL
Duncan
G
Wakeham
A
Shaninian
A
van der Heiden
A
Bachman
MF
Ohashi
PS
Mak
TW
Hickstein
DD
LFA-1-deficient mice show normal CTL responses to virus but fail to reject immunogenic tumor.
J Exp Med
183
1996
1415
238
Coxon
A
Rieu
P
Barkalow
FJ
Askari
S
Sharpe
AH
von Andrian
UH
Arnaout
MA
Mayadas
TN
A novel role for the beta 2 integrin CD11b/CD18 in neutrophil apoptosis: A homeostatic mechanism in inflammation.
Immunity
5
1996
653
239
Lu
H
SMith
CW
Perrard
J
Bullard
D
Tang
L
Shappell
SB
Entman
ML
Beaudet
AL
Ballantyne
CM
LFA-1 is sufficient in mediating neutrophil emigration in MAC-1 deficient mice.
J Clin Invest
99
1997
1340
240
Arroyo
AG
Yang
JT
Rayburn
H
Hynes
RO
Differential requirements for α4 integrins during fetal and adult hematopoiesis.
Cell
85
1996
997
241
Pober
JS
Orosz
CG
Rose
ML
Savage
COS
Can graft endothelial cells initiate a host anti-graft immune response?
Transplant
61
1996
343
242
Murray
AG
Libby
P
Pober
JS
Human vascular smooth muscle cells poorly costimulate and actively inhibit allogeneic CD4+ T cell proliferation in vitro.
J Immunol
154
1995
151
243
Marcelli-Berg
FM
Hargreaves
REG
Carmichael
P
Dorling
A
Lombardi
G
Lechler
RI
Major histocompatibility complex class II-expressing endothelial cells induce allospecific non-responsiveness in naive T cells.
J Exp Med
183
1996
1603
244
Murray
AG
Khodadoust
MM
Pober
JS
Bothwell
ALM
Porcine aortic endothelial cells strongly activate human T cells: Direct presentation of swine MHC antigens and effective costimulation by swine ligands for human CD2 and CD28.
Immunity
1
1994
57
245
Karman
K
Hughes
CCW
Schechner
J
Fnaslow
WC
Pober
JS
CD40 on human endothelial cells: Inducibility by cytokines and functional regulation of adhesion molecule expression.
Proc Natl Acad Sci USA
92
1995
4342
246
Pober
JS
Cotran
RS
Cytokines and endothelial cell biology.
Physiol Rev
70
1990
427
247
Pober
JS
Cotran
RS
Overview: The role of endothelial cells in inflammation.
Transplant
50
1990
537
248
Austrup
F
Vestweber
D
Borges
E
Lohning
M
Grauer
R
Herz
V
Renz
H
Hallmann
R
Scheffold
A
Rudbruch
A
Hamann
A
P- and E-slectin mediate recruitment of T helper 1 but not T helper 2 cells into inflammed tissues.
Nature
385
1997
81
249
Borges
E
Tietz
W
Steegmaier
M
Moll
T
Hallmann
R
Hamann
A
Vestweber
D
P-selectin glycoprotein ligand-1 (PSGL-1) on T helper 1 but not T helper 2 cells binds to P-selectin and supports migration into inflammed skin.
J Exp Med
185
1997
573
250
Bradley
JR
Pober
JS
Prolonged cytokine exposure causes a dynamic redistribution of endothelial cell adhesion molecules to intercellular junctions.
Lab Invest
75
1996
463
251
Montesano
R
Orci
L
Vassalli
P
Human endothelial cell cultures: Phenotypic modulation by leukocyte interleukins.
J Cell Physiol
122
1985
424
252
Stolpen
AH
Guinan
EC
Fiers
W
Pober
JS
Recombinant tumor necrosis factor and immune interferon act singly and in combination to reorganize human vascular endothelial cell monolayers.
Am J Pathol
123
1986
16
253
Jutte
NHP
Knoof
CJ
Heijse
P
Balk
AHMM
Mochtar
B
Claas
FHJ
Weimanr
W
Human heart endothelial-cell-restricted allorecognition.
Transplant
62
1996
403
254
Damle
NK
Doyle
LV
Bender
JR
Bradley
EC
Interleukin 2-activated human lymphocytes exhibit enhanced adhesion to normal vascular endothelial cells and cause their lysis.
J Immunol
138
1987
1779
255
Bach
FH
Ferran
C
Hechenlettner
P
Mark
W
Koyamada
N
Miyatake
T
Winkler
H
Badrichani
A
Candinas
D
Hancock
WW
Accomodation of vascularized xenografts: Expression of “protective genes” by donor endothelial cells in a host TH2 cytokine environment.
Nat Med
3
1997
196
256
Platt AF Jr
Eckman
JR
The multidisciplinary management of pain in patients with sickle cell syndromes.
J Am Acad Phys Assist
2
1989
104
257
Howard
RJ
Gilladoga
AD
Molecular studies related to the pathogenesis of cerebral malaria.
Blood
74
1989
2603
258
Wautier
JL
Wautier
MP
Schmidt
A-M
Anderson
GM
Hori
O
Zoukourian
C
Capron
L
Chappey
O
Yan
S-D
Brett
J
Guillausseau
P-J
Stern
D
Advanced glycation end products (AGEs) on the surface of diabetic erythrocytes bind to the vessel wall via a specific receptor inducing oxidant stress in the vasculature: A link between surface-associated AGEs and diabetic complications.
Proc Natl Acad Sci USA
91
1994
7742
259
Mozarelli
A
Hofrichter
J
Eaton
WA
Delay time of hemoglobin S polymerization prevents most cells from sickling in vivo.
Science
237
1987
500
260
Eaton
WA
Hofrichter
J
Ross
PD
Delay time of gelation: A possible determinant of clinical severity of sickle cell disease.
Blood
47
1976
621
261
Hebbel
RP
Yamada
O
Moldow
CF
Jacob
HS
White
JG
Eaton
JW
Abnormal adherence of sickle erythrocytes to cultured vascular endothelium. Possible mechansism for microvascular occlusion.
J Clin Invest
65
1980
154
262
Wick
TM
Moake
JL
Udden
MM
Eskin
SG
Sean
DA
McIntire
LV
Unusually large von Willebrand factor multimers increase adhesion of sickle erythrocytes to human endothelial cells under controlled flow.
J Clin Invest
80
1987
905
263
Kaul
DK
Fabry
ME
Nagel
RL
Microvascular sites and characteristics of sickle cell adhesion to vascular endothelium in shear flow conditions: Pathophysiological implications.
Proc Natl Acad Sci USA
86
1989
3356
264
Wick
TM
Eckman
JR
Molecular basis of sickle cell-endothelial cell interactions.
Curr Opin Hematol
3
1996
118
265
Hebbel
RP
Moldow
CF
Steinberg
MH
Modulation of erythrocyte-endothelial interactions and the vaso-occlusive severity of sickling disorders.
Blood
58
1981
947
266
Wautier
JL
Wautier
MP
Pintigny
D
Galacteros
F
Courillon
A
Passa
P
Caen
JP
Factors involved in cell adhesion to vascular endothelium.
Blood Cells
9
1983
221
267
Sugihara
K
Sugihara
T
Mohandas
N
Hebbel
RP
Thrombospondin mediates adherence of CD36+ sickle reticulocytes to endothelial cells.
Blood
80
1992
2634
268
Brittain
HA
Eckman
JR
Howard
RJ
Wick
TM
Thrombospondin from activated platelets promotes sickle erythrocyte adherence to human microvascular endothelium under physiologic flow: A potential role for platelet activation in sickle cell vaso-occlusion.
Blood
81
1993
2137
269
Joneckis
CC
Ackley
RL
Orringer
EP
Wayner
EA
Parise
LV
Integrin α4β1 and glycoprotein IV (CD36) are expressed on circulating reticulocytes in sickle cell anemia.
Blood
82
1993
3548
270
Swerlick
RA
Eckman
JR
Kumar
A
Jeitler
M
Wick
TM
α4β1-integrin expression on sickle reticulocytes: Vascular cell adhesion molecule-dependent binding to endothelium.
Blood
82
1993
1891
271
Setty
BNY
Stuart
MJ
Vascular adhesion molecule-1 is involved in mediating hypoxia-induced sickle red blood cell adherence to endothelium: Potential role in sickle cell disease.
Blood
88
1996
2311
272
Smolinski
PA
Offermann
MK
Eckman
JR
Wick
TM
Synthetic double stranded RNA increases sickle red blood cell adherence to endothelium via a VCAM-1/VLA-4 pathway.
Blood
85
1995
2945
273
Kumar
A
Eckman
JR
Swerlick
RA
Wick
TM
Phorbol ester stimulation increases sickle erythrocyte adherence to endothelium: A novel pathway involving α4β1 integrin receptors on sickle reticulocytes and fibronectin.
Blood
88
1996
4348
274
Natarajan
M
Udden
MM
McIntire
LV
Adhesion of sickle red blood cells and damage to interleukin-1β stimulated endothelial cells under flow in vitro.
Blood
87
1996
4845
275
Styles
LA
Lubin
B
Vichinsky
E
Lawrence
S
Hua
M
Test
S
Kuypers
F
Decrease of very late activation antigen-4 and CD36 on reticulocytes in sickle cell patients treated with hydroxyurea.
Blood
89
1997
2554
276
Brittain
HA
Eckman
JR
Wick
TM
Sickle erythrocytic adherence to large vessel and microvascular endothelium under physiologic flow is qualitatively different.
J Lab Clin Med
19
1992
538
277
Swerlick
RA
Lee
KH
Wick
TM
Lawley
TJ
Human dermal microvascular endothelial but not human umbilical vein endothelial cells express CD36 in vivo and in vitro.
J Immunol
148
1992
78
278
(abstr, suppl 1)
Smolinski
PA
Eckman
JR
Wick
TM
Increased tenacity of sickle erythrocyte-endothelial adherence by involvement of multiple receptor-ligand interactions.
Blood
86
1995
298a
279
Kaul
DK
Fabry
ME
Nagel
RL
Vaso-occulusion by sickle cells: Evidence for selective trapping of dense cells.
Blood
68
1986
1162
280
Setty
BNY
Chen
D
Stuart
MJ
Sickle red blood cells stimulate endothelial cell production of eicosanoids and diacyglycerol.
J Lab Clin Med
128
1996
313
281
Solovey
A
Lin
Y
Browne
P
Choong
S
Wayner
E
Hebbel
RP
Circulating activated endothelial cells in sickle cell anemia.
N Engl J Med
337
1997
1584
282
Ockenhouse
CF
Tegodhi
T
Maeno
Y
Benjamin
C
Ho
M
Kan
KE
Thway
Y
Win
K
Aikawa
M
Lobb
R
Human vascular endothelial cell adhesion receptors for Plasmodium falciparum-infected erythrocytes: Roles for endothelial leukocyte adhesion molecule 1 and vascular cell adhesion molecule 1.
J Exp Med
176
1992
1183
283
Cooke
BM
Berendt
AR
Craig
AG
MacGergor
J
Newbold
CI
Nash
GB
Rolling and stationary cytoadhesion of red blood cells parasitized by Plasmodium falciparum: Separate roles for ICAM-1, CD36 and thrombospondin.
Br J Haematol
87
1994
162
284
McCormick
CJ
Craig
A
Roberts
D
Newbold
CI
Berendt
AR
Intercellular adhesion molecule-1 and CD36 synergize to mediate adherence of Plasmodium falciparum-infected erythrocytes to cultured human microvascular endothelial cells.
J Clin Invest
100
1997
2521
285
Elhassan
IM
Hviid
L
Satti
G
Akerstrom
B
Jakobsen
PH
Jensen
JB
Theander
TG
Evidence of endothelial inflammation, T cell activation and T cell reallocation in uncomplicated Plasmodium falciparum malaria.
Am J Trop Med Hyg
51
1994
372
286
Jakobsen
PH
Bate
CAW
Taverne
J
Playfair
JHL
Malaria: Toxins, cytokines and disease.
Parasite Immunol
17
1995
223
287
Crandall
I
Collins
WE
Gysin
J
Sherman
IW
Synthetic peptides based on motifs present in human band 3 protein inhibit cytoadherence/sequestration of the malaria parasite Plasmodium falciparum.
Proc Natl Acad Sci USA
90
1993
4703
288
Crandall
I
Sherman
IW
Cytoadherence-related neoantigens on Plasmodium falciparum (human malaria)-infected erythrocytes result from the exposure of normally cryptic regions of the band 3 protein.
Parasitology
108
1994
257
289
Chaiyoroj
S
Angkasekwinai
P
Buranakiti
A
Looareesuwan
S
Rogerson
SJ
Brown
GV
Cytoadherence characteristics of Plasmodium falciparum isolates from Thailand: Evidence for chondroitin sulfate A as a cytoadherence receptor.
Am J Trop Med Hyg
55
1996
76
290
Wautier
JL
Panton
RC
Wautier
MP
Pintigny
D
Abadie
E
Passa
P
Caen
JP
Increased adhesion of erythrocytes to endothelial cells in diabetes mellitus and its relation to vascular complications.
N Engl J Med
305
1981
237
291
Ross
R
Glomset
JA
The pathogenesis of atherosclerosis (second of two parts).
N Engl J Med
295
1976
420
292
Ross
R
The pathogenesis of atherosclerosis: a perspective for the 1990s.
Nature
362
1993
801
293
McGorisk
GM
Treasure
CB
Endothelial dysfunction in coronary heart disease.
Curr Opin Cardiol
11
1996
341
294
Henry
PD
Hyperlipidemic endothelial injury and angiogenesis.
Basic Res Cardiol
1
1994
107
295
Larrick JW, Wright SC: Hypertension-enhanced monocyte adhesion in experimental atherosclerosis. FASEB J 4:3215, 1990
296
Schmidt
AM
Hori
O
Brett
J
Yan
SD
Wautier
JL
Stern
D
Cellular receptors for advanced glycation end products. Implications for induction of oxidant stress and cellular dysfunction in the pathogensis of vascular lesions.
Arterioscler Thromb
14
1994
1521
297
Marui
N
Offermann
MK
Swerlick
R
Kunsch
C
Rosen
CA
Ahmad
M
Alexander
RW
Medford
RM
Vascular cell adhesion molecule-1 (VCAM-1) gene transcription and expression are regulated through an antioxidant-sensitive mechanism in human vascular endothelial cells.
J Clin Invest
92
1993
1866
298
Gimbrone MA Jr: Vascular endothelium: An integrator of pathophysiologic stimuli in atherosclerosis. Am J Cardiol 75:67B, 1995
299
Davies
P
Flow-mediated endothelial mechanotransduction.
Physiol Rev
75
1995
519
300
Malek
AM
Izumo
S
Control of endothelial cell gene expression by flow.
J Biomechanics
28
1995
1515
301
McLenachan
JM
Vita
J
Fish
DR
Treasure
CB
Cox
DA
Ganz
P
Selwyn
AP
Early evidence of endothelial vasodilator dysfunction at coronary branch points.
Circulation
82
1990
1169
302
Ando
J
Kamiya
A
Flow-dependent regulation of gene expression in vascular endothelial cells.
Heart J
37
1996
19
303
Tsao
PS
Buitrago
R
Chan
JR
Cooke
JP
Fluid flow inhibits endothelial adhesiveness: Nitric oxide and transcriptional regulation of VCAM-1.
Circulation
94
1996
1682
304
Grabowski
EF
Lam
FP
Endothelial cell function, including tissue factor expression, under flow conditions.
Thromb Haemost
74
1995
123
305
Khachigian
LM
Resnick
N
Gimbrone MA Jr
Collins
T
Nuclear factor κB interacts functionally with the platelet-derived growth factor B-chain shear-stress response element in vascular endothelial cells exposed to fluid shear stress.
J Clin Invest
92
1993
1706
306
Ohno
M
Cooke
JP
Dzau
VJ
Gibbons
GH
Fluid shear stress induces endothelial transforming growth factor beta-1 transcription and production.
J Clin Invest
96
1995
1169
307
Lan
Q
Mercurius
KO
Davies
PF
Stimulation of transcription factors NFκV and AP1 in endothelial cells subjected to shear stress.
Biochem Biophys Res Commun
201
1994
950
308
Guba
SC
Fink
LM
Fonseca
V
Hyperhomocysteinemia. An emerging and important risk factor for thromboembolic and cardiovascular disease.
Am J Clin Pathol
105
1996
709
309
Rees
MM
Rodgers
GM
Homocysteinemia: Association of a metabolic disorder with vascular disease and thrombosis.
Thromb Res
71
1993
337
310
Mayer
EL
Jacobsen
DW
Robinson
K
Homocysteine and coronary atherosclerosis.
J Am Coll Cardiol
27
1996
517
311
Harpel
PC
Zhang
X
Borth
W
Homocysteine and hemostasis: Pathogenetic mechanisms predisposing to thrombosis.
J Nutrition
71
1996
1285S
312
Starkebaum
G
Harlan
JM
Endothelial cell injury due to copper-catalyzed hydrogen peroxide generation from homocysteine.
J Clin Invest
77
1993
1370
313
Heinecke
JW
Rosen
H
Suzuki
LA
Chait
A
The role of sulfur-containing amino acids in superoxide production and modification of low density lipoprotein by arterial smooth muscle cells.
J Biol Chem
262
1987
10098
314
Stamler
JS
Osborne
JA
Jaraki
O
Rabbani
LE
Mullins
M
Singel
D
Loscalzo
J
Adverse vascular effects of homocytsteine are modulated by endothelium-derived relaxing factor and related oxides of nitrogen.
J Clin Invest
91
1993
308
315
Fryer
RH
Wilson
BD
Gubler
DB
Fitzgerald
LA
Rodgers
GM
Homocysteine, a risk factor for premature vascular disease and thrombosis, induces tissue factor activity in endothelial cells.
Arterioscler Thromb
13
1993
1327
316
Rodgers
GM
Kane
WH
Activation of endogenous factor V by a homocysteine-induced vascular endothelial cell activator.
J Clin Invest
77
1986
1909
317
Hajjar
KA
Homocysteine-induced modulation of tissue plasminogen activator binding to its endothelial cell membrane receptor.
J Clin Invest
91
1993
2873
318
Lentz
SR
Sadler
JE
Inhibtion of thrombomodulin surface expression and protein C activation by the thrombogenic agent homocysteine.
J Clin Invest
88
1991
1906
319
Nishinaga
M
Ozawa
T
Shimada
K
Homocysteine, a thrombogenic agent, suppresses anticoagulant heparan sulfate expression in cultured porcine aortic endothelial cells.
J Clin Invest
92
1992
381
320
Wang
J
Dudman
NPD
Wilken
DE
Effects of homocysteine and related compounds on prostacyclin production by cultured vascular endothelial cells.
Thromb Haemost
6
1993
1047
321
Tsai
JC
Perella
MA
Yoshizumi
M
Hsieh
CM
Haber
E
Schlegel
R
Lee
ME
Promotion of vascular smooth muscle cell growth by homocysteine: A link to atherosclerosis.
Proc Natl Acad Sci USA
91
1994
6369
322
Lentz
SR
Sobey
CG
Piegors
DJ
Bhopatkar
MY
Faraci
FM
Malinow
MR
Heistad
DD
Vascular dysfunction in monkeys with diet-induced hyperhomocyst(e)inemia.
J Clin Invest
98
1996
24
323
Van Den Berg
M
Boers
GHJ
Franken
DG
Blom
HJ
van Kamp
GJ
Jakobs
C
Rauwerda
JA
Kluft
C
Stehouwert
CDA
Hyperhomocysteinemia and endothelial dysfunction in young patients with peripheral arterial occlusive disease.
Eur J Clin Invest
25
1995
176
324
Gryglewski
RJ
Chlopicki
S
Swies
J
Niezabitowski
P
Prostacyclin, nitric oxide, and atherosclerosis.
Ann NY Acad Sci
748
1995
194
325
Ludmer
PL
Selwyn
AP
Shook
TL
Wayne
RR
Mudge
GH
Alexander
RW
Ganz
P
Paradoxical vasoconstriction induced by acetylcholine in atherosclerotic coronary arteries.
N Engl J Med
315
1986
1046
326
Siegel
G
Ruckborn
K
Schnalke
F
Muller
J
Endothelial dysfunction in human atherosclerotic coronary arteries.
Eur Heart J
1
1993
99
327
Sorensen
KE
Celermajer
DS
Georgakopoulos
D
Hatcher
G
Betteridge
DJ
Deanfield
JE
Impairment of endothelium-dependent dilation is an early event in children with familial hypercholesterolemia and is related to the lipoprotein(a) level.
J Clin Invest
93
1994
50
328
Tsurumi
Y
Nagashima
H
Ichikawa
K
Sumiyoshi
T
Hosoda
S
Influence of plasma lipoprotein(a) levels on coronary vasomotor response to acetylcholine.
J Am Coll Cardiol
26
1995
1242
329
Steinberg
HO
Chaker
H
Leaming
R
Johnson
A
Brechtel
G
Baron
AD
Obesity/insulin resistance is associated with endothelial dysfunction.
J Clin Invest
97
1996
2601
330
Celermajer
DS
Sorensen
K
Ryalls
M
Robinson
J
Thomas
O
Leonard
JV
Deanfield
JE
Impaired endothelial function occurs in the systemic arteries of children with homozygous homocystinuria but not in their heterozygous parents.
J Am Coll Cardiol
22
1993
854
331
Woodman
OL
Modulation of vasoconstriction by endothelium-derived nitric oxide: The influence of vascular disease.
Clin Exp Pharmacol Physiol
22
1995
585
332
Heistad
DD
Baumbach
GL
Faraci
FM
Armstrong
ML
Sick vessel syndrome: Vascular changes in hypertension and atherosclerosis.
J Hum Hypertens
9
1995
449
333
Celermajer
DS
Sorensen
KE
Spiegelhalter
DJ
Georgakopoulos
D
Robinson
J
Deanfield
JE
Aging is associated with endothelial dysfunction in healthy men years before the age-related decline in women.
J Am Coll Cardiol
24
1994
471
334
Celermajer
DS
Sorensen
KE
Georgakopoulos
D
Bull
C
Thomas
O
Robinson
J
Deanfield
JE
Cigarette smoking is associated with dose-related and potentially reversible impairment of endothelium-dependent dilation in healthy young adults.
Circulation
88
1993
2149
335
Kingwell
BA
Tran
B
Cameron
JD
Jennings
GL
Dart
AM
Enhanced vasodilation to acetylcholine in athletes is associated with lower plasma cholesterol.
Am J Physiol
270
1996
H2008
336
Fish
RD
Nabel
EG
Selwyn
AP
Ludmer
PL
Mudge
GH
Kirshenbaum
JM
Schoen
FJ
Alexander
RW
Ganz
P
Responses of coronary arteries of cardiac transplant patients to acetylcholine.
J Clin Invest
81
1988
21
337
Reddy
KG
Nair
RN
Sheehan
HM
Hodgson
JM
Evidence that selective endothelial dysfunction may occur in the absence of angiographic or ultrasound atherosclerosis in patients with risk factors for atherosclerosis.
J Am Coll Cardiol
23
1994
833
338
Buttery
LD
Chester
AH
Springall
DR
Borland
JA
Michel
T
Yacoub
MH
Polak
JM
Atherogenic levels of low-density lipoprotein increase hydrogen peroxide generation in cultured human endothelial cells: Possible mechanism of heightened endocytosis.
J Pathol
179
1996
197
339
Holland
JA
Ziegler
LM
MEyer
JW
Atherogenic levels of low-density lipoprotein increase hydrogen peroxide generation in cultured human endothelial cells: Possible mechanism of heightened endocytosis.
J Cell Physiol
166
1996
144
340
Naruse
K
Shimizu
K
Muramatsu
M
Toki
Y
Miyazaki
Y
Okumura
K
Hashimoto
H
Ito
T
Long-term inhibition of NO synthesis promotes atherosclerosis in the hypercholesterolemic rabbit thoracic aorta. PGH2 does not contribute to impaired endothelium-dependent relaxation.
Arterioscler Thromb
14
1994
746
341
Cooke
JP
Role of nitric oxide in progression and regression of atherosclerosis.
West J Med
164
1996
419
342
Creager
MA
Gallagher
SJ
Girerd
XJ
Coleman
SM
Dzau
VJ
Cooke
JP
L-arginine improves endothelium-dependent vasodilation in hypercholesterolemic humans.
J Clin Invest
90
1992
1248
343
Mathew
V
Hasdai
D
Lerman
A
The role of endothelin in coronary atherosclerosis.
Mayo Clin Proc
71
1996
769
344
Sawamura
T
Kume
N
Aoyami
T
Morlwaki
H
Hoshikawa
H
Alba
Y
Tanaka
R
Miwa
S
Katsura
Y
Kita
T
Masaki
T
An endothelial receptor for oxidized low-density lipoprotein.
Nature
386
1997
73
345
Boulanger
CM
Tanner
FC
Bea
ML
Hahn
AW
Werner
A
Luscher
TF
Oxidized low density lipoproteins induce mRNA expression and release of endothelin from human and porcine endothelium.
Circ Res
70
1992
1191
346
Lerman
A
Holmes DR Jr
Bell
MR
Garratt
KN
Nishimura
RA
Burnett JC Jr
Endothelin in coronary endothelial dysfunction and early atherosclerosis in humans.
Circulation
94
1995
2426
347
Haak
T
Marz
W
Jungmann
E
Hausser
S
Siekmeier
R
Gross
W
Usadel
KH
Elevated endothelin levels in patients with hyperlipoproteinemia.
Clin Invest
72
1994
580
348
Effect of simvastatin on coronary atheroma: The Multicentre Anti-Atheroma Study (MAAS). Lancet 344:633, 1994
349
Treasure
CB
Klein
JL
Weintraub
WS
Talley
JD
Stillabower
ME
Kosinski
AS
Zhang
J
Boccuzzi
SJ
Cedarholm
JC
Alexander
RW
Beneficial effects of cholesterol-lowering therapy on the coronary endothelium in patients with coronary artery disease.
N Engl J Med
332
1995
4811
350
Hennig
B
Diana
JN
Toborek
M
McClain
CJ
Influence of nutrients and cytokines on endothelial cell metabolism.
J Am Coll Nutr
13
1994
224
351
Heitzer
T
Just
H
Munzel
T
Antioxidant vitamin C improves endothelial dysfunction in chronic smokers.
Circulation
94
1996
6
352
Levine
GN
Frei
B
Koulouris
SN
Gerhard
MD
Kenaey JR Jr
Vita
JA
Ascorbic acid reverses endothelial vasomotor dysfunction in patients with coronary artery disease.
Circulation
93
1996
1107
353
Chobanian
AV
Haudenschild
CC
Nickerson
C
Hope
S
Trandolapril inhibits atherosclerosis in the Watanabe heritable hyperlipidemic rabbit.
Hypertension
20
1992
473
354
Overturf
ML
Sybers
HD
Smith
SA
Captopril-induced hyperreninemia in cholesterol-fed rabbits.
Res Commun Chem Pathol Pharmacol
47
1985
229
355
Lonn
EM
Yusuf
S
Jha
P
Montague
TJ
Teo
KK
Benedict
CR
Pitt
B
Emerging role of angiotensin-converting enzyme inhibitors in cardiac and vascular protection.
Circulation
90
1994
2056
356
Stroes
ESG
Koomans
HA
deBruin
TWA
Rabelink
TJ
Vascular function in the forearm of hypercholesterolaemic patients on and off lipid-lowering medication.
Lancet
346
1995
467
357
Wada
H
Mori
Y
Kaneko
T
Wakita
Y
Nakase
T
Minamikawa
K
Ohiwa
M
Tamaki
S
Tanigawa
M
Kayegama
S
Deguchi
K
Nakano
T
Shirakawa
S
Suzuki
K
Elevated levels of vascular endothelial cell markers in patients with hypercholesterolemia.
Am J Hematol
44
1993
112
358
Hackman
A
Abe
Y
Insull
W
Jr
Pownall
H
Smith
L
Dunn
K
Gotto AM Jr
Ballantyne
CM
Levels of soluble cell adhesion molecules in patients with dyslipidemia.
Circulation
93
1996
1334
359
Morrow
JD
Roberts
LJ
The isoprostanes. Current knowledge and directions for future research.
Biochem Pharm
51
1996
1
360
Reilly
M
Delanty
N
Lawson
JA
FitzGerald
GA
Modulation of oxidant stress in vivo in chronic cigarette smokers.
Circulation
94
1996
19
361
Antonov
AS
Munn
DH
Kolodgie
FD
Virmani
R
Gerrity
RG
Aortic endothelial cells regulate proliferation of human monocytes in vitro via a mechanism synergistic with macrophgage colony-stimulating factor. Convergence at the cyclin E/p27kip1 regulatory checkpoint.
J Clin Invest
99
1997
2867
362
Ruderman A, Williamson N, Brownlee M: Glucose and diabetic vascular disease. FASEB J 6:2905, 1992
363
Browlee
M
Cerami
A
Viassara
H
AGEs in tissues and the biochemical basis of diabetic complications.
N Engl J Med
318
1989
1315
364
Schmidt
A-M
Vianna
M
Gerlach
M
Brett
J
Ryan
J
Kao
J
Esposito
C
Hegarty
H
Hurley
W
Clauss
M
Wang
F
Pan
Y-C
Tsang
T
Stern
D
Isolation and characterization of binding proteins for AGEs from lung tissue which are present on the endothelial cell surface.
J Biol Chem
267
1992
14987
365
Schmidt
A-M
Hori
O
Brett
J
Yan
S-D
Wautier
S-D
Stern
D
Cellular receptors for AGEs: Implications for induction of oxidant stress and cellular dysfunction in the pathogenesis of vascular lesions.
Arterioscler Thromb
1521
1994
1528
366
Khoury
J
Thomas
C
Loike
J
Hickman
S
Cao
L
Silverstein
S
Macrophages adhere to glucose-modified basement membrane via their scavenger receptors.
J Biol Chem
269
1994
10197
367
Lander
HL
Tauras
JM
Ogiste
JS
Moss
RA
Schmidt
A-M
Activation of the receptor for advanced glycation endproducts trigger a MAP kinase pathway regulated by oxidant stress.
J Biol Chem
272
1997
17810
368
Schmidt
A-M
Hori
O
Chen
J
Li
J
Crandall
J
Zhang
J
Cao
R
Yan
S-D
Brett
J
Stern
D
AGEs interaction with their endothelial cell receptor induce expression of VCAM-1 in cultured human endothelial cells and in mice.
J Clin Invest
96
1995
1395
369
Schmidt
A-M
Crandall
J
Cao
R
Hori
P
Lakatta
E
Elevated plasma levels of vascular cell adhesion molecules-1 (VCAM-1) in diabetic patients with microalbuminuria: A marker of vascular dysfunction and progressive vascular disease.
Br J Haematol
92
1996
747
370
Solez
K
Axelsen
RA
Bendiktsson
H
Burdick
JF
Cohen
AF
Colvin
RB
Croker
BP
Droz
D
Dunnill
MS
Halloran
PF
Hayry
P
Jennette
JC
Keown
PA
Marcussen
N
Mihatsch
MJ
Morozumi
K
Myers
ND
Nast
CC
Olsen
S
Racusen
LC
Ramos
EC
Rosen
S
Sachs
DH
Salomon
DR
Sanfilippo
F
Verani
R
von WIllebrand
E
Yamaguch
Y
International standardization of criteria for the histologic diagnosis of renal allograft rejection: The Banff working classification of kidney transplant pathology.
Kidney Int
44
1993
411
371
Pober JS: Vascular endothelium in transplantation, in Tilney NC, Strom TB, Paul LC (eds): Transplantation Biology: Cellular and Molecular Aspects. Philadelphia, PA, Lipincott-Raven, 1996, p 127
372
Strehlau
J
Pavlakis
M
Lipman
M
Shapiro
M
Vascincellos
L
Harmon
W
Strom
TB
Quantitative detection of immune activation transcripts as a diagnostic tool in kidney transplantation.
Proc Natl Acad Sci USA
94
1997
695
373
Meehan
SM
McClusky
RT
Pascual
M
Preffer
FI
Anderson
P
Schlossman
SF
Colvin
RB
Cytotoxicity and apoptosis in human renal allograft: Identification, distribution and quantitation of cells with a cytotokic granule protein GMP 17 (TIA-1) and cells with fragmented nuclear DNA.
Lab Invest
76
1997
639
374
Jutte
NHPM
Knoop
CJ
Heijse
P
Balk
AHMM
Mochtar
B
Claas
FHJ
Weimar
W
Human heart endothelial cell-restricted allorecognition.
Transplant
62
1996
403
375
Dvorak
HF
Mihm MC Jr
Dvorak
AM
Barnes
BA
Manseau
EJ
Galli
SJ
Rejection of first-set skin allografts in man. The microvasculature is the critical target of the immune response.
J Exp Med
150
1981
322
376
Gao
S-Z
Alderman
EL
Schroeder
JS
Silverman
JF
Hunt
SA
Accelerated coronary vascular disease in the heart transplant patient: Coronary angiographic findings.
J Am Coll Cardiol
12
1988
334
377
Salomon
RN
Hughes
CCW
Schoen
FJ
Payne
DD
Pober
JS
Libby
P
Human coronary transplantation-associated arteriosclerosis: evidence for a chronic immune reaction to activated graft endothelial cells.
Am J Pathol
138
1991
791
378
Hurban
RH
Beschorner
WE
Baumgartner
WA
Augustine
SM
Ren
H
Reitz
BA
Hutchins
GM
Accelerated arteriosclerosis in heart transplant recipients is associated with a T-lymphocyte-mediated endothelialitis.
Am J Pathol
137
1990
871
379
Dong
C
Wilson
JE
Winters
GL
McManus
BM
Human transplant coronary artery disease: Pathologic evidence for fas-mediated cytotoxicity in allograft arteriopathy.
Lab Invest
74
1996
921
380
d'Apice AJ, Pearse MJ: Xenotransplantation, in Tilney NC, Strom TB, Paul LC (eds): Transplanation Biology: Cellular and Molecular Aspects. Philadelphia, PA, Lippincott-Raven, 1996, p 701
381
Sandrin
MS
Vaughan
HA
Dabkowski
PL
McKenzie
IFC
Anti-pig IgM antibodies in human serum react predominantly with Gal (α1-3) Gal epitopes.
Proc Natl Acad Sci USA
90
1993
11391
382
Rooney
IA
Liszeuski
MK
Atkinson
JP
Using membrane band complement regulatory proteins to inhibit rejection.
Xeno
1
1993
29
383
Murray
AG
Khodadoust
MM
Pober
JS
Bothwell
ALM
Porcine aortic endothelial cells strongly activate T cells: Direct presentation of swine MHC antigens and effective costimulation by swine ligands for human CD2 and CD28.
Immunity
1
1994
57
384
Blakely
ML
Van Der Werf
WJ
Berndly
MC
Dalmasso
AP
Bach
FH
Hancock
WW
Activation of intragraft endothelial and mononuclear cells during discordant xenograft rejection.
Transplant
58
1994
1059
385
Sultan
P
Murray
AG
McNiff
JM
Lorber
MI
Askenase
PW
Bothwell
ALM
Pober
JS
Pig but not human interferon-γ initiated human endothelial cell-mediated rejection of pig tissue in vivo.
Proc Natl Acad Sci USA
94
1997
8767
386
McCrae KR, Cines DB: Immunologic properties of antibodies to endothelial cells: Relationship to antiphospholipid antibodies, in Cervera R, Khamashta MA, Hughes GRV (eds): Antibodies to Endothelial Cells and Vascular Damage. Boca Raton, FL, CRC, 1994, p 47
387
Cines
DB
Lyss
AP
Reeber
MJ
Bina
M
DeHoratius
R
Presence of complement-fixing anti-endothelial cell antibodies in systemic lupus erythematosus.
J Clin Invest
73
1984
611
388
McCrae
KR
DeMichele
A
Samuels
P
Roth
D
Kuo
A
Meng
Q-H
Rauch
J
Cines
DB
Detection of endothelial cell-reactive immunoglobulin in patients with anti-phospholipid antibodies.
Br J Haematol
79
1991
595
389
Cines
DB
Disorders associated with antibodies to endothelial cells.
Rev Infect Dis
11
1989
705
390
Simantov
R
LaSala
JM
Lo
SK
Gharavi
AE
Sammaritano
LR
Salmon
JE
Silverstein
RL
Activaton of cultured vascular endothelial cells by antiphospholipid antibodies.
J Clin Invest
96
1995
2211
391
Angles-Cano
E
Sultan
Y
Clauvel
JP
Predisposing factors to thrombosis in systemic lupus erythematosus. Possible relation to endothelial cell damage.
J Lab Clin Med
94
1979
312
392
Arepally G, Cines DB: Heparin-associated antibodies, in Shoenfeld Y, Peter J (eds): Autoantibodies. Amsterdam, The Netherlands, Elsevier, 1996, p 343
393
McCrae KR, Cines DB: Drug-induced thrombocytopenias, in Loscalzo J, Schafer AI (eds): Thrombosis and Hemorrhage. Boston, MA, Blackwell Scientific, 1994, p 545
394
Amiral
J
Bridey
F
Dreyfus
M
Vissac
AM
Fressinaud
E
Wolf
M
Meyer
D
Platelet factor 4 complexed to heparin is the target for antibodies generated in heparin-induced thrombocytopenia.
Thromb Haemost
68
1992
95
395
Visentin
GP
Ford
SE
Scott
JP
Aster
RH
Antibodies from patients with heparin induced thrombocytopenia/thrombosis are specific for platelet factor 4 complexed with heparin or bound to endothelial cells.
J Clin Invest
93
1994
81
396
Cines
DB
Tomaski
A
Tannenbaum
S
Immune endothelial cell injury in heparin-associated thrombocytopenia.
N Engl J Med
316
1987
581
397
Mayet W-J, Csernok E, Szymkowiak C, Gross WL, Meye zum Buschenfelde K-H: Human endothelial cells express proteinase 3, the target antigen of anticytoplasmic antibodies in Wegener's granulomatosis. Blood 82:1221, 1993
398
Damianovich
M
Gilburd
B
George
J
Del Papa
N
Afek
A
Goldberg
I
Kopolovic
Y
Roth
D
Barkai
G
Meroni
P-L
Shoenfeld
Y
Pathogenic role of anti-endothelial cell antibodies in vasculitis. An idiotypic experimental model.
J Immunol
156
1996
4946
399
Savage
COS
Pottinger
BE
Gaskin
G
Lockwood
CM
Pusey
CD
Pearson
JD
Vascular damage in Wegener's granulomatosis and microscopic polyarteritis: Presence of anti-endothelial cell antibodies and their relation to anti-neutrophil cytoplasm antibodies.
Clin Exp Immunol
85
1991
14
400
Leung
DYM
Havens
PL
Moake
JL
Kim
M
Pober
JS
Lytic anti-endothelial cell antibodies in haemolytic-uremic syndrome.
Lancet
2
1988
183
401
Carvalho
D
Savage
COS
Black
CM
Pearson
JD
IgG antiendothelial cell autoantibodies from scleroderma patients induce leukocyte adhesion to human vascular endothelial cell in vitro.
J Clin Invest
97
1996
111
402
Sgonc
R
Gruschwitz
MS
Dietrich
H
Recheis
H
Gershwin
ME
Wick
G
Endothelial cell apoptosis is a primary pathogenic event underlying skin lesions in avian and human scleroderma.
J Clin Invest
98
1996
785
403
Schett
G
Xu
Q
Amberger
A
van der Zee
R
Recheis
H
Willeit
J
Wick
G
Autoantibodies against heat shock protein 60 mediates endothelial cytotoxicity.
J Clin Invest
96
1995
2569
404
Leung
DYM
Collins
T
Lapierre
LA
Geha
RS
Pober
JS
Immunoglobulin M antibodies present in the acute phase of Kawaski syndrome lyse cultured endothelial cells stimulated by gamma interferon.
J Clin Invest
77
1986
1428
405
Cervera
R
Navarro
M
Lopez-Soto
A
Cid
M-C
Font
J
Esparza
J
Reverter
JC
Monteagudo
J
Ingelmo
M
Urbano-Marquez
A
Antibodies to endothelial cells in Behcet's disease: Cell-binding heterogeneity and association with disease activity.
Ann Rheumatol Dis
53
1994
265
406
Brasile
L
Zerbe
T
Rabin
B
Clarke
J
Abrams
A
Cerilli
J
Identification of the antibody to vascular endothelial cells in patients undergoing cardiac transplantation.
Transplant
40
1985
672
407
Boehme
MWJ
Schmitt
WH
Youinou
P
Stremmel
WR
Gross
WL
Clinical relevance of elevated serum thrombomodulin and soluble E-selectin with Wegerner's granulomatosis and other systemic vasculitides.
Am J Med
101
1996
387
408
Favaloro
EJ
Differential expression of surface antigens on activated endothelium.
Immunol Cell Biol
71
1993
571
409
Cotran
RS
Gimbrone MA Jr
Bevilacqua
MP
Mendrick
DL
Pober
JS
Induction and detection of a human endothelial activation antigen in vivo.
J Exp Med
164
1986
661
410
Groger
MG
Sarmay
E
Fiebirger
E
Wolff
K
Petzelbauer
P
Dermal microvascular endothelial cells express CD32 receptors in vivo and in vitro.
J Immunol
156
1996
1549
411
Cines
DB
Lyss
AP
Bina
M
COrkey
R
Kefalides
NA
Friedman
HM
Herpes simplex virus enhances the expression of receptors for C3 and the Fc portion of immunoglobulin G and human endothelial cells.
J Clin Invest
69
1982
123
412
Johnson
E
Hetland
G
Human umbilical vein endothelial cells synthesize functional C3, C5, C6, C8 and C9 in vitro.
Scand J Immunol
33
1991
667
413
Brooimanns
RA
Van der Ark
AAJ
Buurman
WA
van Es
LA
Daha
MR
Differential regulation of complement factor H and C3 production in human umbilical vein endothelial cells by IFN-γ and IL-1.
J Immunol
144
1990
3835
414
Schmaier
AH
Murray
SC
Heda
GH
Farber
A
Kuo
A
McCrae
K
Cines
DB
Synthesis and expression of C1 inhibitor by human umbilical vein endothelial cells.
J Biol Chem
264
1989
18173
415
Brooimans
RA
Van der Ark
AAJ
Tomita
M
van Es
LA
Daha
MR
CD59 expressed by human endothelial cells functions as a protective molecule against complement-mediated lysis.
Eur J Immunol
22
1992
791
416
Venneker
GT
van den Hoogen
FHJ
Boerbooms
AMT
Bos
JD
Asghar
SS
Aberrant expression of membrane cofactor protein and decay-accelerating factor in the endothelium of patients with systemic sclerosis. A possible mechanism of vascular damage.
Lab Invest
70
1994
830
417
Asch
AS
Kinoshita
T
Jaffe
EA
Nussenzweig
V
Decay-accelerating factor is present on cultured human umbilical vein endothelial cells.
J Exp Med
163
1986
221
418
McNearney
T
Ballard
L
Seya
T
Atkinson
JP
Membrane cofactor protein of complement is present on human fibroblast, epithelial, and endothelial cells.
J Clin Invest
84
1989
538
419
Nose
M
Katoh
M
Okada
N
Kyogoku
M
Okada
H
Tissue distribution of HRF-20, a novel factor preventing the membrane attack of homologous complement, and its predominant expression on endothelial cells in vivo.
Immunology
70
1990
145
420
Daha
MR
Miltenburg
AMM
Hiemstra
PS
Klar-Mohamad
N
van Es
LA
Van Hinsburg
VWM
The complement subcomponent C1q mediated binding of immune complexes and aggregates to endothelial cells in vitro.
Eur J Immunol
18
1988
783
421
Lozada
C
Levin
RI
Huie
M
Hirschhorn
R
Naime
D
Whitlow
M
Recht
PA
Golden
B
Cronstein
BN
Identification of C1q as the heat-labile serum cofactor required for immune complexes to stimulate endothelial expression of the adhesion molecules E-selecton and intercellular and vascular adhesion molecules 1.
Proc Natl Acad Sci USA
92
1995
8378
422
Foreman
KE
Vaporciyan
BK
Bonish
ML
Jones
KJ
Glovsky
MM
Eddy
SM
Ward
PA
C5a induced expression of P-selectin in endothelial cells.
J Clin Invest
94
1994
1147
423
Kilgore
KS
Shen
JP
Miller
BF
Ward
P
Warren
JS
Enhancement by the complement membrane attack complex of tumor necrosis factor-α-induced endothelial cell expression of E-selectin and ICAM-1.
J Immunol
155
1995
1434
424
Kilgore
KS
Schmid
E
Shanley
TP
Flory
CM
Maheswari
V
Tramontini
NL
Cohen
H
Ward
PA
Friedl
HP
Warren
JS
Sublytic concentrations of the membrane attack complex of complement induce endothelial cell interleukin-8 and monocyte chemoattractamt protein-1 through nuclear factor-κB activation.
Am J Pathol
150
1997
2019
425
Saadi
S
Holznecht
RA
Patte
CP
Stern
DM
Platt
JL
Complement-mediated regulation of tissue factor activity in endothelium.
J Exp Med
182
1995
1807
426
Tedesco
F
Pausa
M
Nardon
E
Introna
M
Mantovani
A
Dobrina
A
The cytolytically inactive terminal complement complex activates endothelial cells to express adhesion molecules and tissue factor procoagulant activity.
J Exp Med
185
1997
1619
427
Hattori
R
Hamilton
KK
McEver
RP
Sims
PJ
Complement proteins C5b-9 induce secretion of high molecular weight multimers of endothelial von Willebrand factor and translocation of granule membrane protein GMP-140 to the cell surface.
J Biol Chem
264
1989
9053
428
Hamilton
KK
Hattori
R
Esmon
CT
Sims
PJ
Complement proteins C5b-9 induce vesiculation of the endothelial plasma membrane and expose catalytic surface for assembly of the prothrombinase enzyme complex.
J Biol Chem
265
1990
3809
429
Suttorp
N
Seeger
W
Zinsky
S
Bhakdi
S
Complement complex C5b-8 induces PGI2 formation in cultured endothelial cells.
Am J Physiol
253
1987
C13
430
Christiansen
VJ
Sims
PJ
Hamilton
KK
Complement C5b-9 increases plasminogen binding and activation on human endothelial cells.
Arterioscler Thromb Vasc Biol
17
1997
164
431
Moake
JL
Haemolytic-uraemic syndrome: Basic science.
Lancet
343
1994
393
432
Heild
GH
Haemolytic-uremic syndrome in practice.
Lancet
343
1994
398
433
Forsyth
KD
Simpson
AC
Fitzpatrick
MM
Barratt
TM
Levinsky
RJ
Neutrophil-mediated endothelial cell injury in haemolytic uraemic syndrome.
Lancet
2
1989
411
434
Morigi
M
Micheletti
G
Figliuzzi
M
Imberti
B
Karmali
MA
Remuzzi
A
Remuzzi
G
Zoja
C
Verotoxin-1 promotes leukocyte adhesion to cultured endothelial cells under physiologic flow conditions.
Blood
86
1995
4553
435
van Setten
PA
Monnens
LAH
Verstrate
RGG
van den Heuvel
LPWJ
Van Hinsbergh
VWM
Effects of verotoxin-1 on nonadherent monocytes: binding characteristics, protein synthesis, and induction of cytokine release.
Blood
88
1996
174
436
Harel
Y
Silva
M
Weinberg
A
Cleary
TB
Beutler
B
A reporter transgene indicates renal-specific induction of tumor necrosis factor (TNF) by shiga-like toxin.
J Clin Invest
92
1993
2110
437
Keusch
GT
Acheson
DWK
Aaldering
L
Erban
J
Jacewicz
MS
Comparison of the effects of Shiga-like toxin 1 on cytokine- and butyrate-treated human umbilical and saphenous vein endothelial cells.
J Infect Dis
173
1996
1164
438
Kaye
SA
Obrig
TG
Effect of TNF-α, Shiga toxin and calcium ionophore on Weibel-Palade body content of endothelial cells: Possible implications for the hemolytic uremic syndrome.
Thromb Res
79
1995
415
439
Bergstein
JM
Riley
M
Bang
NU
Role of plasminogen-activator inhibitor type 1 in the pathogenesis and outcome of the hemolytic uremic syndrome.
N Engl J Med
327
1992
755
440
Moake
JL
Byrnes
JJ
Thrombotic microangiopathies associated with drugs and bone marrow transplantation.
Hematol Oncol Clin North Am
10
1996
485
441
Zoja
C
Furci
L
Ghilardi
F
Zilio
P
Benigni
A
Remuzzi
G
Cyclosporin-induced endothelial cell injury.
Lab Invest
56
1986
455
442
Gottschall
JL
Neahring
B
McFarland
JG
Wu
G-G
Weitekamp
LA
Aster
RH
Quinine-induced immune thrombocytopenia with hemolytic uremic syndrome: Clinical and serological findings in nine patients and review of the literature.
Am J Hematol
47
1994
283
443
Moake
JL
Rudy
CK
Troll
JH
Weinstein
MJ
Colannino
NM
Azocar
J
Seder
RH
Hong
SL
Deykin
D
Unusually large plasma factor VIII: von Willebrand factor multimers in chronic relapsing thrombotic thrombocytopenic purpura.
N Engl J Med
307
1982
1432
444
Wada
H
Ohiwa
M
Kaneko
T
Tamaki
S
Tanigawa
M
Shirakawa
S
Koyama
M
Hayashi
T
Suzuki
K
Plasma thrombomodulin as a marker of vascular disorders in thrombotic thrombocytopenic purpura and disseminated intravascular coagulation.
Am J Hematol
39
1992
20
445
Glas-Greenwalt
P
Hall
JM
Panke
TW
Kant
KS
Allen
CM
Pollak
VE
Fibrinolysis in health and disease: Abnormal levels of plasminogen activator, plasminogen activator inhibitor, and protein C in thrombotic thrombocytopenic purpura.
J Lab Clin Med
108
1986
415
446
Katayama
M
Handoa
M
Araki
Y
Ambo
H
Kawal
Y
Watanabe
K
Ikeda
Y
Soluble P-selectin is present in normal circulation and its plasma level is elevated in patients with thrombotic thrombocytopenic purpura and haemolytic uraemic syndrome.
Br J Haematol
84
1993
702
447
Wu
K
Hall
ER
Rosai
EC
Papp
AC
Serum prostacyclin binding defects in thrombotic thrombocytopenic purpura.
J Clin Invest
75
1985
168
448
Kobayashi
M
Wada
H
Wakita
Y
Shimura
M
Nakase
T
Hiyoyama
K
Nagaya
S
Minami
N
Nakano
T
Shiku
H
Decreased plasma tissue factor pathway inhibitor levels in patients with thrombotic thrombocytopenic purpura.
Am J Hematol
73
1995
10
449
Khong
TY
De Wolf
F
Robertson
WB
Brosens
I
Inadequate maternal vascular response to placentation in pregnancies complicated by preeclampsia and by small for gestational age infants.
Am J Obstetr Gynecol
157
1987
360
450
Khong
TY
Sawyer
IH
Heryet
AR
An immunohistologic study of endothelialization of uteroplacental vessels in human pregnancy—Evidence that the endothelium is focally disrupted by trophoblast in preeclampsia.
Am J Obstetr Gynecol
167
1992
751
451
Ferris
TF
Preeclampsia and postpartum renal failure: Examples of pregnancy-induced microangiopathy.
Am J Med
99
1995
343
452
Fitzgerald
DJ
Entmann
SSW
Mulloy
K
Fitzgerald
GA
Decreased prostacyclin biosynthesis preceding the clinical manifestations of pregnancy induced hypertension.
Circulation
75
1987
956
453
Taylor
RN
Varma
M
Teng
NN
Roberts
JM
Women with pre-eclampsia have higher plasma endothelin levels than women with normal pregnancy.
J Clin Endocrinol Metabol
71
1990
1675
454
Lockwood
CJ
Peters
JH
Increased plasma levels of EDI+ cellular fibronectin precede the clinical signs of preeclampsia.
Am J Obstetr Gynecol
162
1990
358
455
Friedman
SA
Schiff
E
Emeris
JJ
Dekker
GA
Sibai
BM
Biochemical corroboration of endothelial involvement in severe preeclampsia.
Am J Obstetr Gynecol
172
1995
202
456
Redman
CWG
Beilin
LJ
Denson
DWE
Bolton
FG
Stirrat
GM
Factor VIII consumption in pre-eclampsia.
Lancet
2
1977
1249
457
Rogers
GM
Taylor
RN
Roberts
JM
Preeclampsia is associated with a serum factor cytotoxic to human endothelial cells.
Am J Obstetr Gynecol
159
1988
906
458
Tsukimori
K
Maeda
H
Shingu
M
Koyanagi
T
Nounaga
M
Nakano
H
The possible role of endothelial cells in hypertensive disorders during pregnancy.
Obstetr Gynecol
80
1992
229
459
Endresen
MJ
Tosti
E
Lorentzen
B
Henricksen
T
Sera of preeclamptic women are not cytotoxic to endothelial cells in culture.
Am J Obstetr Gynecol
172
1995
196
460
Zammit
VC
Whitworth
JA
Brown
MA
Preeclampsia: The effects of serum on endothelial cell prostacyclin, endothelin and cell membrane integrity.
Am J Obstetr Gynecol
174
1996
737
461
Smarason
AK
Sargent
IL
Redmna
CWG
Endothelial cell proliferation is suppressed by plasma but not by serum from women with preeclampsia.
Am J Obstetr Gynecol
174
1996
787
462
Taylor
RN
Casal
DC
Jones
LA
Varma
M
Martin JN Jr
Roberts
JM
Selective effects of preeclamptic sera on human endothelial cell procoagulant protein expression.
Am J Obstetr Gynecol
165
1991
1705
463
Lorentzen
B
Endresen
MJ
Hovig
T
Haug
E
Henriksen
T
Sera from preeclamptic women increase the content of triglycerides and reduce the release of prostacyclin in cultured endothelial cells.
Thromb Res
63
1991
363
464
Roberts
JM
Edep
ME
Goldfein
A
Taylor
RN
Sera from preeclamptic women specifically activate human umbilical vein endothelial cells in vitro: Morphologic and biochemical evidence.
Am J Reprod Immunol
27
1992
101
465
Gallery
EDM
Rowe
J
Campbell
S
Hawkins
T
Effect of serum on secretion of prostacyclin and endothelin-1 by decidual endothelial cells from normal and preeclamptic pregnancies.
Am J Obstetr Gynecol
173
1995
918
466
Poranen
A-K
Ekblad
U
Uotila
P
Ahotupa
M
Lipid peroxidation and antioxidants in normal and pre-eclamptic pregnancies.
Placenta
17
1996
401
467
Ness
RB
Roberts
JR
Heterogenous causes constituting the single syndrome of preeclampsia: A hypothesis and its implications.
Am J Obstetr Gynecol
175
1996
1365
Sign in via your Institution