Protein C is best known for its mild deficiency associated with venous thrombosis risk and severe deficiency associated with neonatal purpura fulminans. Activated protein C (APC) anticoagulant activity involves proteolytic inactivation of factors Va and VIIIa, and APC resistance is often caused by factor V Leiden. Less known is the clinical success of APC in reducing mortality in severe sepsis patients (PROWESS trial) that gave impetus to new directions for basic and preclinical research on APC. This review summarizes insights gleaned from recent in vitro and in vivo studies of the direct cytoprotective effects of APC that include beneficial alterations in gene expression profiles, anti-inflammatory actions, antiapoptotic activities, and stabilization of endothelial barriers. APC's cytoprotection requires its receptor, endothelial cell protein C receptor, and protease-activated receptor-1. Because of its pleiotropic activities, APC has potential roles in the treatment of complex disorders, including sepsis, thrombosis, and ischemic stroke. Although much about molecular mechanisms for APC's effects on cells remains unclear, it is clear that APC's structural features mediating anticoagulant actions and related bleeding risks are distinct from those mediating cytoprotective actions, suggesting the possibility of developing APC variants with an improved profile for the ratio of cytoprotective to anticoagulant actions.

Protein C circulates in plasma at 70 nM as the zymogen of the anticoagulant serine protease, activated protein C (APC), which averages 40 pM (∼ 2.3 ng/mL) in normal plasma.1  The physiologic importance of the protein C system is most clearly demonstrated by the massive, usually lethal, thrombotic complications occurring in infants with severe homozygous protein C deficiency and the significantly increased risk for venous thrombosis in heterozygous deficient adults.2,3  The most commonly identifiable hereditary risk factor for venous thrombosis among whites involves an APC-cleavage site mutation (Arg506Gln, factor V Leiden) that is the major target for factor Va inactivation by APC.4-6  Targeted deletion of the protein C gene in mice results in perinatal lethality.7,8  Nevertheless, in the absence of fetal protein C, embryogenesis and development occur, potentially due to maternal protein C in the fetus. Although protein C–null embryos develop at the expected Mendelian distribution until embryonic day 17.5, these embryos show extensive bleeding, coagulopathy, fibrin deposition, and liver necrosis.7,8  In moderately to severely deficient mice, protein C levels of 1% to 18% suffice for development and birth, although such mice are prone to early onset thrombosis and inflammation, indicating protein C's physiological role as an antithrombotic and anti-inflammatory protein.9 

Cell surfaces localize and promote a variety of biochemical transformations involving protein C and various cellular protein receptors (Figure 1A-C). These biochemical reactions include protein C activation, expression of APC anticoagulant activity, and initiation of APC's cytoprotective actions (see “The cytoprotective protein C pathway”).

Figure 1

Schematic models of PC activation and APC activities. Protein C (PC) activation occurs on the endothelial cell membrane and requires thrombomodulin (TM) (A). Thrombin (IIa) bound to thrombomodulin activates protein C. Binding of protein C to its endothelial receptor, EPCR, provides for the most efficient activation of protein C. Protein C and APC have a similar affinity for EPCR. Dissociation of APC from EPCR allows expression of APC's anticoagulant activity (B), whereas retention of APC bound to EPCR allows APC to express multiple direct cellular activities (C). APC conveys its anticoagulant activity (B) when bound to cell membrane surfaces, various microparticles, or lipoproteins (eg, HDL). As an anticoagulant, APC cleaves the activated cofactors Va (fVa) and VIIIa (fVIIIa) to yield the inactivated cofactors fVi and fVIIIi. Inactivation of factors Va and VIIIa by APC is enhanced by a number of different protein cofactors (eg, protein S, factor V) and various lipids cofactors (eg, phosphatidylserine, cardiolipin, glucosylceramide, HDL). Beneficial activities of APC that involve direct effects of APC on cells require the cellular receptors EPCR and PAR-1 (C). These activities include APC-mediated alteration of gene expression, anti-inflammatory activities, antiapoptotic activities, and protection of endothelial barrier functions. Collectively, these activities are referred to as APC's cytoprotective activities.

Figure 1

Schematic models of PC activation and APC activities. Protein C (PC) activation occurs on the endothelial cell membrane and requires thrombomodulin (TM) (A). Thrombin (IIa) bound to thrombomodulin activates protein C. Binding of protein C to its endothelial receptor, EPCR, provides for the most efficient activation of protein C. Protein C and APC have a similar affinity for EPCR. Dissociation of APC from EPCR allows expression of APC's anticoagulant activity (B), whereas retention of APC bound to EPCR allows APC to express multiple direct cellular activities (C). APC conveys its anticoagulant activity (B) when bound to cell membrane surfaces, various microparticles, or lipoproteins (eg, HDL). As an anticoagulant, APC cleaves the activated cofactors Va (fVa) and VIIIa (fVIIIa) to yield the inactivated cofactors fVi and fVIIIi. Inactivation of factors Va and VIIIa by APC is enhanced by a number of different protein cofactors (eg, protein S, factor V) and various lipids cofactors (eg, phosphatidylserine, cardiolipin, glucosylceramide, HDL). Beneficial activities of APC that involve direct effects of APC on cells require the cellular receptors EPCR and PAR-1 (C). These activities include APC-mediated alteration of gene expression, anti-inflammatory activities, antiapoptotic activities, and protection of endothelial barrier functions. Collectively, these activities are referred to as APC's cytoprotective activities.

Close modal

Physiological proteolytic activation of protein C by thrombin occurs on the surface of the endothelial cell and involves the 2 membrane receptors, thrombomodulin and endothelial protein C receptor (EPCR) (Figure 1A).10  Binding of thrombin to thrombomodulin on the endothelial surface shields thrombin's procoagulant exosite I and promotes its anticoagulant properties by activation of protein C by the thrombin-thrombomodulin complex. This reaction is augmented by localization of protein C on the endothelial surface by its binding to EPCR.11 

The anticoagulant actions of APC are primarily based on the irreversible proteolytic inactivation of factors Va and VIIIa with contributions by various cofactors (Figure 1B). These anticoagulant APC cofactors comprise both proteins and lipids, including protein S, factor V, high-density lipoprotein, anionic phospholipids (eg, phosphatidylserine, cardiolipin), and glycosphingolipids (eg, glucosylceramide). The reader is referred to excellent previous reviews for discussions of APC's anticoagulant mechanisms that are beyond the scope of this review.10,12,13 

In reactions mediated by EPCR10  and the effector receptor, protease-activated receptor-1 (PAR-114 ), APC remarkably acts directly on cells to exert multiple cytoprotective effects including (1) alteration of gene expression profiles; (2) anti-inflammatory activities; (3) antiapoptotic activity; and (4) protection of endothelial barrier function (Figure 1C).15-19  As clear from the schemes presented in Figure 1B-C, the pathways for anticoagulant and cytoprotective pathways differ not only in terms of biologic effects but also in terms of substrates and cofactors. APC's substrates for anticoagulant action are factors Va and VIIIa and for cytoprotective actions, PAR-1. We would like to emphasize making an explicit distinction between the protein C cytoprotective pathway and the protein C anticoagulant pathway because this distinction is useful for design and interpretation of preclinical and clinical results and of in vitro experiments. Moreover, this distinction may also be made based on genetic engineering of APC variants with selectively altered specificities for factor Va versus PAR-1. Such APC variants may facilitate experimental distinction of the relative importance of APC's direct effects on cells versus APC's anticoagulant activity (see “Roles for APC's anticoagulant activity versus APC's cytoprotective activities”).

The tissue distribution of EPCR and thrombomodulin is notably distinct and different, with higher levels of EPCR on the surface of large vessels versus smaller vessels, whereas thrombomodulin shows the opposite pattern.10,20-23  No studies have defined the tissue distribution or cellular localization of EPCR relative to thrombomodulin or PAR-1. It will be interesting to establish the extent to which thrombomodulin and EPCR colocalize in lipid membrane rafts in different tissues where protein C activation can be localized versus the colocalization of PAR-1 and EPCR. Such data may enable speculation about the tissue specificity for protein C activation versus expression of APC cytoprotective activities.

Both the anticoagulant and cytoprotective protein C pathways are important for APC's beneficial effects in various settings. As noted in “The anticoagulant protein C pathway,” the anticoagulant protein C pathway is centered on inactivation of factors Va and VIIIa by APC and involves protein and lipid cofactors (Figure 1B). Until recently, investigators often presumed that the systemic anticoagulant activity of APC mediated APC's beneficial anti-inflammatory effects because APC down-regulates generation of thrombin, which was recognized for its proinflammatory properties and because no other mechanism for APC's anti-inflammatory activity was apparent. However, both clinical trial data and basic lab findings have changed this viewpoint and stimulated research on the protein C cytoprotective pathway. In clinical trials, the ability of APC, but not of other anticoagulants such as antithrombin and tissue factor pathway inhibitor, to reduce mortality in severe sepsis patients in comparable large phase 3 trials implicates the less well-defined cytoprotective actions of APC as potentially very important for APC's pharmacologic success.24-26  In the laboratory, extensive in vivo and in vitro studies related to murine ischemic stroke models imply that APC's neuroprotective effects, at least in part, are independent of APC's anticoagulant activity,16  and recent studies show that APC variants with severely reduced anticoagulant activity prevent endotoxemia-induced death in mice (see “Roles for APC's anticoagulant activity versus APC's cytoprotective activities”).27  These various observations support the notion that APC's beneficial cytoprotective effects are beneficial in vivo and that they are, at least in part, independent of APC anticoagulant activity.27-29 

The direct effects of APC on cells, that is, APC's cytoprotective effects, include (1) alteration of gene expression profiles; (2) anti-inflammatory activities; (3) antiapoptotic activity; and (4) endothelial barrier stabilization. Although potentially interrelated, each of these activities of APC is distinct and may or may not involve the same intracellular mechanisms with their particular APC dose-response characteristics, depending on a cell's receptor profile and on a particular cell's location, condition, or particular properties. Much remains to be clarified about these issues.

As summarized under the next 4 subheadings, EPCR is required for most if not all of APC's known cytoprotective actions, and a requirement for PAR-1 is also implicated for many of APC's effects on cells.

APC-mediated alteration of gene expression profiles

Transcriptional profiling of alterations in gene expression patterns caused by APC's treatment of cells revealed modulation of gene expression for the major pathways of inflammation and apoptosis.15,16,19,30-33  The effects of APC generally involved down-regulation of proinflammatory and proapoptotic pathways and up-regulation of anti-inflammatory and antiapoptotic pathways (Table 1), thus helping to explain the anti-inflammatory and antiapoptotic activities of APC. APC suppresses nuclear transcription factor κB (NFκB)–modulated genes by directly reducing NFκB expression and functional activity; this causes inhibition of cytokine signaling and inhibition of tumor necrosis factor-α (TNFα)–dependent induction of adhesion molecules. APC up-regulates antiapoptotic gene products related to Bcl-2 and suppresses proapoptotic p53 and Bax expression.

Table 1

Up-regulation and down-regulation of gene expression by APC

Gene*Official gene namePathway affectedReference sequenceOther protein nameSources
Up-regulation      
    BCL2A1 BCL2-related protein A1 Apoptosis NM_004049 A1 Joyce et al,15  Riewald and Ruf30  
    BMP2 Bone morphogenetic protein 2 Differentiation NM_001200 BMP-2 Riewald et al,19  Riewald and Ruf30  
    CCL2 Chemokine (C-C motif) ligand 2 Inflammation NM_002982 MCP-1 Riewald et al,19  Franscini et al33  
    CX3CL1 Chemokine (C-X3-C motif) ligand 1 Adhesion NM_002996 Fractalkine Joyce et al,15  Franscini et al,33  Brueckmann et al34  
    CXCL2 Chemokine (C-X-C motif) ligand 2 Inflammation NM_002089 MIP-2, GRO Riewald et al,19  Riewald and Ruf,30  Franscini et al33  
    DDX21 DEAD (Asp-Glu-Ala-Asp) box polypeptide 21 Transcription NM_004728 RNA helicase II/Gu-α Joyce et al15  
    F3 Coagulation factor III (thromboplastin, tissue factor) Coagulation NM_001993 TF, CD142 Riewald et al,19  Riewald and Ruf30  
    HBEGF Heparin-binding EGF-like growth factor Proliferation NM_001945 HBEGF Riewald et al19  
    IL1RI Interleukin 1 receptor, type 1 Inflammation NM_000877 IL-1R1, CD121a Riewald and Ruf,30  Franscini et al33  
    IL6 Interleukin 6 (interferon, beta 2) Inflammation NM_000600 IL-6 Franscini et al,33  Hooper et al35  
    IL8 Interleukin 8 Inflammation NM_000584 IL-8, CXCL-8 Franscini et al,33  Hooper et al,35  Brueckmann et al36  
    NFKB1 Nuclear factor of kappa light polypeptide gene enhancer 1 Transcription NM_003998 NF-κB Joyce et al,15  Joyce and Grinnell,32  Franscini et al33  
    NFKB2 Nuclear factor of kappa light polypeptide gene enhancer 2 Transcription NM_002502 NF-κB2 p100 Joyce et al,15  Riewald and Ruf,30  Franscini et al33  
    NOS3 Nitric oxide synthase 3 (endothelial cell) Inflammation NM_000603 eNOS Joyce et al15  
    NR4A1 Nuclear receptor subfamily 4, group A, member 1 Apoptosis NM_173157 TR3, Nur77, NA4A1 Riewald et al,19  Riewald and Ruf,30  Franscini et al33  
    NR4A2 Nuclear receptor subfamily 4, group A, member 2 Transcription NM_006186 NURR1, NA4A2 Riewald et al,19  Riewald and Ruf30  
    NR4A3 Nuclear receptor subfamily 4, group A, member 3 Transcription NM_006981 NOR-1, NA4A3 Riewald et al,19  Riewald and Ruf30  
    PCNA Proliferating cell nuclear antigen Cell cycle NM_002592 PCNA Joyce et al15  
    PTGS2 Prostaglandin-endoperoxide synthase 2 Inflammation NM_000963 COX-2 Riewald et al,19  Riewald and Ruf,30  Brueckmann et al37  
    SMAD3 SMAD, mothers against DPP homolog 3 (DrosophilaApoptosis NM_005902 SMAD-3 Joyce et al15  
    TNFAIP3 Tumor necrosis factor, alpha-induced protein 3 Apoptosis NM_006290 A20 Joyce et al,15  Riewald et al19  
Down-regulation      
    B2M Beta-2-microglobulin Inflammation NM_004048 B2M Joyce and Grinnell32  
    BIRC5 Baculoviral IAP repeat-containing 5 (survivin) Apoptosis NM_001168 Survivin Joyce and Grinnell,32  Franscini et al33  
    CALR Calreticulin Inflammation NM_004343 Ro, cC1qR Joyce et al,15  Joyce and Grinnell32  
    CMKOR1 Chemokine orphan receptor 1 Cell signaling NM_020311 RDC1 Joyce et al15  
    CX3CL1 Chemokine (C-X3-C motif) ligand 1 Adhesion NM_002996 Fractalkine Joyce et al,15  Franscini et al,33  Brueckmann et al34  
    EFNA1 Ephrin-A1 Chemoattract NM_004428 Ephrin-A1 Joyce et al15  
    ICAM1 Intercellular adhesion molecule 1 (CD54) Adhesion NM_000201 ICAM-1, CD54 Joyce et al,15  Franscini et al33  
    IL1R1 Interleukin 1 receptor, type I Inflammation NM_000877 IL-1R1, CD121a Riewald and Ruf,30  Franscini et al33  
    LTB Lymphotoxin beta (TNF superfamily, member 3) Inflammation NM_002341 LT-β Joyce and Grinnell32  
    MMP10 Matrix metallopeptidase 10 (stromelysin 2) Proteolysis NM_002425 MMP-10 Joyce et al15  
    NFKB1 Nuclear factor of kappa light polypeptide gene enhancer 1 Transcription NM_003998 NF-κB Joyce et al,15  Joyce and Grinnell,32  Franscini et al33  
    NFKB2 Nuclear factor of kappa light polypeptide gene enhancer 2 Transcription NM_002502 NF-κB2 p100 Joyce et al,15  Riewald and Ruf30  
    SELE Selectin E (endothelial adhesion molecule 1) Adhesion NM_000450 E-selectin, CD62E Joyce et al,15  Franscini et al33  
    SOD2 Superoxide dismutase 2, mitochondrial Oxidation NM_000636 MnSOD Joyce et al15  
    THBS1 Thrombospondin 1 Adhesion NM_003246 TSP-1 Riewald and Ruf30  
    TP53 Tumor protein p53 (Li-Fraumeni syndrome) Apoptosis NM_000546 p53 Cheng et al,16  Riewald and Ruf,30  Guo et al31  
    VCAM1 Vascular cell adhesion molecule 1 Adhesion NM_080682 VCAM-1, CD106 Joyce et al,15  Franscini et al33  
Gene*Official gene namePathway affectedReference sequenceOther protein nameSources
Up-regulation      
    BCL2A1 BCL2-related protein A1 Apoptosis NM_004049 A1 Joyce et al,15  Riewald and Ruf30  
    BMP2 Bone morphogenetic protein 2 Differentiation NM_001200 BMP-2 Riewald et al,19  Riewald and Ruf30  
    CCL2 Chemokine (C-C motif) ligand 2 Inflammation NM_002982 MCP-1 Riewald et al,19  Franscini et al33  
    CX3CL1 Chemokine (C-X3-C motif) ligand 1 Adhesion NM_002996 Fractalkine Joyce et al,15  Franscini et al,33  Brueckmann et al34  
    CXCL2 Chemokine (C-X-C motif) ligand 2 Inflammation NM_002089 MIP-2, GRO Riewald et al,19  Riewald and Ruf,30  Franscini et al33  
    DDX21 DEAD (Asp-Glu-Ala-Asp) box polypeptide 21 Transcription NM_004728 RNA helicase II/Gu-α Joyce et al15  
    F3 Coagulation factor III (thromboplastin, tissue factor) Coagulation NM_001993 TF, CD142 Riewald et al,19  Riewald and Ruf30  
    HBEGF Heparin-binding EGF-like growth factor Proliferation NM_001945 HBEGF Riewald et al19  
    IL1RI Interleukin 1 receptor, type 1 Inflammation NM_000877 IL-1R1, CD121a Riewald and Ruf,30  Franscini et al33  
    IL6 Interleukin 6 (interferon, beta 2) Inflammation NM_000600 IL-6 Franscini et al,33  Hooper et al35  
    IL8 Interleukin 8 Inflammation NM_000584 IL-8, CXCL-8 Franscini et al,33  Hooper et al,35  Brueckmann et al36  
    NFKB1 Nuclear factor of kappa light polypeptide gene enhancer 1 Transcription NM_003998 NF-κB Joyce et al,15  Joyce and Grinnell,32  Franscini et al33  
    NFKB2 Nuclear factor of kappa light polypeptide gene enhancer 2 Transcription NM_002502 NF-κB2 p100 Joyce et al,15  Riewald and Ruf,30  Franscini et al33  
    NOS3 Nitric oxide synthase 3 (endothelial cell) Inflammation NM_000603 eNOS Joyce et al15  
    NR4A1 Nuclear receptor subfamily 4, group A, member 1 Apoptosis NM_173157 TR3, Nur77, NA4A1 Riewald et al,19  Riewald and Ruf,30  Franscini et al33  
    NR4A2 Nuclear receptor subfamily 4, group A, member 2 Transcription NM_006186 NURR1, NA4A2 Riewald et al,19  Riewald and Ruf30  
    NR4A3 Nuclear receptor subfamily 4, group A, member 3 Transcription NM_006981 NOR-1, NA4A3 Riewald et al,19  Riewald and Ruf30  
    PCNA Proliferating cell nuclear antigen Cell cycle NM_002592 PCNA Joyce et al15  
    PTGS2 Prostaglandin-endoperoxide synthase 2 Inflammation NM_000963 COX-2 Riewald et al,19  Riewald and Ruf,30  Brueckmann et al37  
    SMAD3 SMAD, mothers against DPP homolog 3 (DrosophilaApoptosis NM_005902 SMAD-3 Joyce et al15  
    TNFAIP3 Tumor necrosis factor, alpha-induced protein 3 Apoptosis NM_006290 A20 Joyce et al,15  Riewald et al19  
Down-regulation      
    B2M Beta-2-microglobulin Inflammation NM_004048 B2M Joyce and Grinnell32  
    BIRC5 Baculoviral IAP repeat-containing 5 (survivin) Apoptosis NM_001168 Survivin Joyce and Grinnell,32  Franscini et al33  
    CALR Calreticulin Inflammation NM_004343 Ro, cC1qR Joyce et al,15  Joyce and Grinnell32  
    CMKOR1 Chemokine orphan receptor 1 Cell signaling NM_020311 RDC1 Joyce et al15  
    CX3CL1 Chemokine (C-X3-C motif) ligand 1 Adhesion NM_002996 Fractalkine Joyce et al,15  Franscini et al,33  Brueckmann et al34  
    EFNA1 Ephrin-A1 Chemoattract NM_004428 Ephrin-A1 Joyce et al15  
    ICAM1 Intercellular adhesion molecule 1 (CD54) Adhesion NM_000201 ICAM-1, CD54 Joyce et al,15  Franscini et al33  
    IL1R1 Interleukin 1 receptor, type I Inflammation NM_000877 IL-1R1, CD121a Riewald and Ruf,30  Franscini et al33  
    LTB Lymphotoxin beta (TNF superfamily, member 3) Inflammation NM_002341 LT-β Joyce and Grinnell32  
    MMP10 Matrix metallopeptidase 10 (stromelysin 2) Proteolysis NM_002425 MMP-10 Joyce et al15  
    NFKB1 Nuclear factor of kappa light polypeptide gene enhancer 1 Transcription NM_003998 NF-κB Joyce et al,15  Joyce and Grinnell,32  Franscini et al33  
    NFKB2 Nuclear factor of kappa light polypeptide gene enhancer 2 Transcription NM_002502 NF-κB2 p100 Joyce et al,15  Riewald and Ruf30  
    SELE Selectin E (endothelial adhesion molecule 1) Adhesion NM_000450 E-selectin, CD62E Joyce et al,15  Franscini et al33  
    SOD2 Superoxide dismutase 2, mitochondrial Oxidation NM_000636 MnSOD Joyce et al15  
    THBS1 Thrombospondin 1 Adhesion NM_003246 TSP-1 Riewald and Ruf30  
    TP53 Tumor protein p53 (Li-Fraumeni syndrome) Apoptosis NM_000546 p53 Cheng et al,16  Riewald and Ruf,30  Guo et al31  
    VCAM1 Vascular cell adhesion molecule 1 Adhesion NM_080682 VCAM-1, CD106 Joyce et al,15  Franscini et al33  

eNOS indicates endothelial nitric oxide synthase.

*

Selected genes whose expression is modulated by APC. Genes were selected from available microarray data. Gene selection criteria are based on at least 2-fold up-regulation or down-regulation by APC and independent confirmation of the gene array results. The list is not a complete list of all potentially important changes in gene expression. Nomenclature according to HUGO gene nomenclature committee.

These genes are down-regulated by APC in TNFα-stimulated human umbilical vein endothelial cells (HUVECs) but up-regulated by APC when human coronary artery endothelial cells (HCAECs) are stimulated with a mixture of interleukin-1β, TNFα, and interferon-γ.15,33 

Modulation of the expression of thrombospondin-1 and p53 by APC in the microarray analysis was less than 2-fold. Nevertheless, APC-induced changes in these genes are shown to be functionally significant and are herein included (see “APC-mediated alteration of gene expression profiles”).

The effects of APC on gene expression profiles are at least in part due to the inhibitory effect of APC on transcription factor activity. For instance, APC suppressed inflammation-activated transcription factors of the activator protein-1 (AP-1) family c-Fos and FosB, which directly induce intercellular adhesion molecule-1 (ICAM-1) and monocyte chemoattractant protein-1 (MCP-1) in endothelial cells, as well as c-Rel, a member of the NFκB family.15 

The molecular mechanisms for APC's effects on gene expression remain to be elucidated, but much of APC's modulation of gene expression profiles is mediated by EPCR-dependent PAR-1 activation. Curiously, APC-induced alterations of gene expression that required PAR-1 were not remarkably different from those induced by thrombin when the effects of these 2 proteases on unperturbed endothelial cells were studied in vitro. However, after endothelial cells were stimulated by TNFα, some of the effects on gene expression were different for APC compared with thrombin.30  Up-regulation of p53 mRNA by thrombin was most notably reduced by APC.16  Similar results were observed for expression of thrombospondin-1 that was down-regulated by APC but not by thrombin.30  APC, but not thrombin, down-regulates both p53 mRNA and protein by APC.16,30 

Thus, for stressed endothelial cells, APC and thrombin clearly do not coincide in their transcriptome alteration effects, although they both can activate the same receptor, PAR-1. The basis for how 2 protease agonists can activate the same G-protein–coupled receptor, PAR-1, with different outcomes has not been clarified. Evidently, APC alters the transcriptome via mechanisms that are, at least in part, distinct from those used by thrombin.

APC anti-inflammatory activity

The anti-inflammatory vascular effects of APC can be divided into its effects on endothelial cells and its effects on leukocytes. APC's effects on endothelial cells include inhibition of inflammatory mediator release and down-regulation of vascular adhesion molecules, thereby reducing leukocyte adhesion and infiltration of tissues and limiting damage to underlying tissue (Figure 2). In addition, APC protects and maintains endothelial barrier function and reduces the chemotactic potential of several potent chemotactic agents.43-48 

Figure 2

The cytoprotective protein C pathway for APC anti-inflammatory activity. Anti-inflammatory effects of APC include APC's effects on vascular endothelial cells and APC's effects on leukocytes. Inhibition of inflammatory gene expression on endothelial cells by APC is EPCR and PAR-1 dependent (green arrow). APC also down-regulates expression of vascular adhesion molecules (such as ICAM-1, VCAM-1, and E-selectin) on the endothelial surface in the presence of inflammatory mediators (red block), thereby limiting leukocyte adhesion and infiltration. APC reduces proinflammatory mediator release (such as TNFα and IL-1β) from leukocytes (red block). The mechanisms by which APC acts on leukocytes are incompletely resolved and could involve the receptors EPCR and PAR-1, EPCR alone, or others (green arrows).38-41  The complex of soluble EPCR (sEPCR) and proteinase 3 (PR3) binds to the integrin complex CD11b/CD18 (αMβ2; Mac-1; CR3) on activated neutrophils. Although speculative at present (indicated by the question mark), binding of (A)PC to sEPCR is retained when sEPCR is bound to proteinase 3, suggesting that sEPCR might mediate APC cellular signals and/or activation of protein C on leukocytes.42 

Figure 2

The cytoprotective protein C pathway for APC anti-inflammatory activity. Anti-inflammatory effects of APC include APC's effects on vascular endothelial cells and APC's effects on leukocytes. Inhibition of inflammatory gene expression on endothelial cells by APC is EPCR and PAR-1 dependent (green arrow). APC also down-regulates expression of vascular adhesion molecules (such as ICAM-1, VCAM-1, and E-selectin) on the endothelial surface in the presence of inflammatory mediators (red block), thereby limiting leukocyte adhesion and infiltration. APC reduces proinflammatory mediator release (such as TNFα and IL-1β) from leukocytes (red block). The mechanisms by which APC acts on leukocytes are incompletely resolved and could involve the receptors EPCR and PAR-1, EPCR alone, or others (green arrows).38-41  The complex of soluble EPCR (sEPCR) and proteinase 3 (PR3) binds to the integrin complex CD11b/CD18 (αMβ2; Mac-1; CR3) on activated neutrophils. Although speculative at present (indicated by the question mark), binding of (A)PC to sEPCR is retained when sEPCR is bound to proteinase 3, suggesting that sEPCR might mediate APC cellular signals and/or activation of protein C on leukocytes.42 

Close modal

APC inhibits inflammatory mediator release by leukocytes as well as endothelial cells. APC diminishes cytokine release from leukocytes and thereby may attenuate initiation of systemic inflammatory responses. This APC action might reduce the so-called cytokine storm that is associated with sepsis. For example, APC inhibits lipopolysaccharide (LPS)–induced production of proinflammatory mediators by monocytes.49-52 

Mechanisms for APC's anti-inflammatory effects on leukocytes are not completely clear; however, evidence indicates an important role for EPCR38-41  that is on the surface of monocytes, CD56+ natural killer cells, neutrophils, and eosinophils, but not T cells and B cells.48,53-55  Of interest, soluble EPCR lacking the transmembrane helix of native EPCR interacts with the integrin CD11b/CD18 (Mac-1) (αMβ2) (CR3) on leukocytes, suggesting that binding of soluble EPCR to CD11b/CD18 might interfere with leukocyte adhesion (Figure 2).42  Proteinase-3 (PR3), a serine protease with elastase-like properties stored in granules of neutrophils, also binds both CD11b/CD18 and soluble EPCR. PR3 expression on the plasma membrane of neutrophils is increased upon neutrophil stimulation. Although one may speculate that PR3 can mediate binding of soluble EPCR to CD11b/CD18, proof for a functional role of the EPCR:PR3:CD11b/CD18 complex remains elusive. But since the soluble EPCR:PR3 complex binds protein C and APC, it is possible that this complex may mediate APC cellular signals and/or activation of protein C on leukocytes.42 

Animal injury studies are consistent with the hypothesis that APC inhibits leukocyte cytokine release, chemotaxis, and migration in vivo. In both rats and humans, APC inhibits endotoxin-induced pulmonary injury and inflammation, at least in part through inhibition of leukocyte accumulation and chemotaxis.44,45  In a rat compression-induced spinal cord injury model, APC reduces TNFα levels, neutrophil accumulation, and motor disturbances.56  It should be noted, however, that neutrophil accumulation is enhanced by fibrin. Thus, the anticoagulant action of APC, which decreases thrombin generation, may indirectly reduce neutrophil accumulation in tissues by reducing fibrin deposition.

APC prolongs the lifespan of in vivo circulating monocytes by inhibition of spontaneous monocyte apoptosis (normal half-life ∼ 24 hours).32,38  Inhibition of monocyte apoptosis by APC requires the receptors EPCR and PAR-1, consistent with results for inhibition of endothelial cell apoptosis by APC.38  Whether a lengthened survival of monocytes is beneficial during sepsis is controversial as the consequential prolonged proinflammatory mediator release may worsen tissue damage and augment the inflammatory response. On the other hand, immediate host responses to invading microorganisms and phagocytosis-mediated clearance of invading pathogens would benefit from a decreased monocyte turnover. Selective inhibition of proinflammatory mediator release in monocytes by APC while leaving the antimicrobial phagocytic capacity of monocytes intact suggests that APC antiapoptotic activity in monocytes is likely to contribute an overall anti-inflammatory effect.38 

APC antiapoptotic activity

Apoptosis is a tightly regulated process to execute an efficient cell death program.57-59  Apoptotic signals coming from inside the cell activate the intrinsic pathway in response to cellular stress (eg, hypoxia), and release of cytochrome c from mitochondria and subsequent procaspase-3 activation are key steps for the intrinsic pathway. Key regulators, such as the tumor-suppressor protein, p53, and the Bcl-2 family of proteins, determine the mitochondria's roles by balancing apoptotic versus survival signals. Activation of the extrinsic apoptotic pathway, which occurs via activation of initiator caspases, such as procaspase-8, is dependent on membrane-bound death receptors and sensors of the extracellular environment. The extrinsic apoptotic pathway may also use elements of the intrinsic pathway via activation of the proapoptotic members of the Bcl-2 family or may directly activate the common pathway by activation of procaspase-3. Whether activated via the intrinsic or the extrinsic pathway, caspase-3 cleaves specific cellular substrates, leading to the morphologic and biochemical features characteristic of apoptotic cell death.57-59 

APC manifests antiapoptotic activity both in vitro and in vivo, and this activity requires the enzymatic active site of APC and its receptors, EPCR and PAR-1 (Figure 3).15,16,18,60  In murine injury models, APC inhibits apoptosis while providing neuroprotection. Reduction of apoptosis is associated with improved survival in sepsis, suggesting a potential important role for APC's antiapoptotic activity in reducing mortality in sepsis.61 

Figure 3

The cytoprotective protein C pathway for APC antiapoptotic activity. APC antiapoptotic activity requires the APC receptors EPCR and PAR-1. APC antiapoptotic activity is at least partially dependent on modulation of gene expression (green arrow). APC down-regulates proapoptotic p53 and Bax protein (red blocks) and maintains protective antiapoptotic Bcl-2 protein levels (green arrow), thereby beneficially affecting the Bax/Bcl-2 ratio. APC inhibits activation (red block) of both initiator caspases (eg, tPA-induced caspase-8 activation60 ) as well as activation of effector caspases (eg, staurosporine or hypoxia/hypoglycemia-induced caspase-3 activation16,18 ). Further work is needed to clarify exactly how APC exerts antiapoptotic activity and what the relative contributions of APC's effects on gene expression, of APC-specific signaling, and of APC-specific proteolysis involving particular receptors or effectors might be for APC's antiapoptotic actions. Green arrows indicate stimulation by APC, red blocks indicate inhibition by APC, and black arrows/blocks denote stimulation/inhibition of (downstream) reactions.

Figure 3

The cytoprotective protein C pathway for APC antiapoptotic activity. APC antiapoptotic activity requires the APC receptors EPCR and PAR-1. APC antiapoptotic activity is at least partially dependent on modulation of gene expression (green arrow). APC down-regulates proapoptotic p53 and Bax protein (red blocks) and maintains protective antiapoptotic Bcl-2 protein levels (green arrow), thereby beneficially affecting the Bax/Bcl-2 ratio. APC inhibits activation (red block) of both initiator caspases (eg, tPA-induced caspase-8 activation60 ) as well as activation of effector caspases (eg, staurosporine or hypoxia/hypoglycemia-induced caspase-3 activation16,18 ). Further work is needed to clarify exactly how APC exerts antiapoptotic activity and what the relative contributions of APC's effects on gene expression, of APC-specific signaling, and of APC-specific proteolysis involving particular receptors or effectors might be for APC's antiapoptotic actions. Green arrows indicate stimulation by APC, red blocks indicate inhibition by APC, and black arrows/blocks denote stimulation/inhibition of (downstream) reactions.

Close modal

APC reduces many characteristic apoptotic features, including DNA degradation, caspase-3 activation, and phosphatidylserine translocation to the outer cell membrane.15,16,18,60  At this time, specific intracellular targets for APC that mediate inhibition of apoptosis have not been identified. However, in stressed human brain endothelial cells (BECs), APC reduces the amounts of p53 protein and mRNA induced by hypoxia (Figure 3).16  Curiously, only a small number of gene products, including p53 and thrombospondin-1, show PAR-1–dependent down-regulation by APC in contrast to their up-regulation by thrombin.30  During hypoxic stress of brain endothelial cells, APC reduces up-regulation of proapoptotic Bax and maintains levels of protective Bcl-2 protein, thereby blunting stimulation of the intrinsic apoptotic pathway.16 

The antiapoptotic effects of APC are not limited to the intrinsic apoptotic pathway as APC counteracts the neurotoxicity of tissue plasminogen activator (tPA), which exerts proapoptotic activity via the extrinsic pathway.60  In the setting of tPA-induced apoptosis of brain cells, APC inhibits caspase-8 activation, showing APC's antiapoptotic effects when either the extrinsic pathway or the intrinsic pathway is activated in the brain. Thus, the antiapoptotic effects of APC have been shown to be broadly cytoprotective both in vitro and in vivo.

Further work is needed to clarify exactly how APC exerts antiapoptotic activity and whether this activity is limited to APC's effects on gene expression or whether other additional effects, such as APC-specific signaling or APC-specific proteolysis involving particular receptors or effectors, might play direct roles.

APC-mediated endothelial barrier stabilization

Breakdown of the endothelial barrier is a key factor in the pathogenesis of inflammation. The consequences of increased endothelial permeability involve swelling, hypotension, and promotion of inflammation, and these processes can contribute to the physiopathology of sepsis, acute lung injury, and organ failure (see reviews62-64 ). APC induces potent barrier protective effects via EPCR-dependent PAR-1 activation, induction of sphingosine kinase-1 (SphK-1), and up-regulation of sphingosine-1-phosphate (S1P) formation via sphingosine kinase (Figure 4).43,46  S1P is a biologically active sphingolipid that signals via the S1P receptor-1 (S1P1), a G-protein–coupled receptor belonging to the endothelial differentiation gene (Edg) family. Activation of the S1P1 by S1P reduces endothelial permeability and stabilizes the cellular cytoskeleton via stabilization of cytoskeletal elements that are dependent on modulation of Rho family GTPases and mitogen-activated protein kinases (MAPKs) (see reviews62,64,65 ).

Figure 4

The cytoprotective protein C pathway for APC-mediated endothelial barrier protection. Protection of endothelial barrier function by APC requires EPCR and PAR-1.43,46  Activation of PAR-1 by APC stimulates sphingosine kinase-1 (SphK-1) to form sphingosine-1-phosphate (S1P) from sphingosine.43  S1P either released from activated platelets or formed intracellularly by SphK-1 activates the sphingosine-1-phosphate receptor 1 (S1P1) to promote increased endothelial barrier protection. A direct interaction of EPCR with S1P1 was postulated, but a mechanistic contribution to barrier protective effects remains unclear (indicated by question mark).46  In addition to barrier protective effects, S1P also promotes cell survival. In contrast, ceramide and sphingosine contribute to signals inducing cell death. The dynamic equilibrium between the formation and degradation of ceramide, sphingosine, and S1P is a sphingolipid rheostat for the cell, and the setting of this rheostat balances death-initiating and death-preventing signaling. It is currently unclear whether the S1P cell survival signals contribute to APC antiapoptotic activity or other cytoprotective activities.

Figure 4

The cytoprotective protein C pathway for APC-mediated endothelial barrier protection. Protection of endothelial barrier function by APC requires EPCR and PAR-1.43,46  Activation of PAR-1 by APC stimulates sphingosine kinase-1 (SphK-1) to form sphingosine-1-phosphate (S1P) from sphingosine.43  S1P either released from activated platelets or formed intracellularly by SphK-1 activates the sphingosine-1-phosphate receptor 1 (S1P1) to promote increased endothelial barrier protection. A direct interaction of EPCR with S1P1 was postulated, but a mechanistic contribution to barrier protective effects remains unclear (indicated by question mark).46  In addition to barrier protective effects, S1P also promotes cell survival. In contrast, ceramide and sphingosine contribute to signals inducing cell death. The dynamic equilibrium between the formation and degradation of ceramide, sphingosine, and S1P is a sphingolipid rheostat for the cell, and the setting of this rheostat balances death-initiating and death-preventing signaling. It is currently unclear whether the S1P cell survival signals contribute to APC antiapoptotic activity or other cytoprotective activities.

Close modal

APC-enhanced endothelial barrier protection requires PAR-1. Curiously, robust signaling of PAR-1 by thrombin leads to endothelial barrier destabilization, whereas PAR-1 activation by APC leads to barrier stabilization.43  Differences in barrier effects mediated by thrombin and APC are dependent on the level of S1P induction; S1P concentration-dependent selectivity for S1P receptor–mediated second messenger Rac (protective) and Rho (destabilizing) signaling; and recruitment of S1P receptors to caveolin-rich microdomains.65  In vitro studies suggest an apparent functional interaction of EPCR with S1P1 associated with endothelial barrier stabilization.46  These protective effects of APC appear to be more effectively achieved by endogenously generated APC than by exogenously added APC (Figure 1C).66  This finding highlights a novel potential mechanistic link between the protein C cytoprotective pathway and pathways for S1P-mediated barrier modulation (Figure 4).43,46,66 

S1P not only promotes endothelial barrier stabilization but also mediates antiapoptotic signals. The sphingolipids, ceramide and sphingosine, are in dynamic metabolic equilibrium with S1P. The ratio of ceramide to sphingosine, sometimes referred to as a sphingolipid rheostat, appears to mediate proapoptotic signaling. Whether APC-induced S1P up-regulation and concomitant ceramide and sphingosine down-regulation contribute to APC antiapoptotic activity is currently unclear. Glucosylceramide, a major metabolic product of ceramide, enhances APC anticoagulant activity,67,68  suggesting that interactions between the sphingolipid regulatory pathway and the protein C pathways might be both more complicated and more significant than previously appreciated.

Protease-activated receptors (PARs) are G-protein–coupled receptors and are activated by proteolytic cleavage of an extracellular N-terminal tail that forms an intramolecular tethered ligand that triggers intracellular signaling.69  In reactions requiring EPCR, APC can activate PAR-1 on endothelial cells and both PAR-1 and PAR-2 on transfected cells, although to date only the activation of PAR-1 has been shown to alter cellular functional activity.16-19  In the presence of EPCR, APC induces PAR-1–dependent MAPK phosphorylation, increases intracellular calcium fluxes, and modulates PAR-1–specific gene products in endothelial cells.17,19,30  Dependent on EPCR and PAR-1, APC inhibits hypoxia/hypoglycemia or staurosporine-induced apoptosis in various endothelial cells types.15,16,18  This indicates that in the presence of EPCR, APC can induce biologically relevant intracellular signaling transduction through PAR-1 on endothelial cells (Figure 1C).

A number of the remarkable in vivo anti-inflammatory and neuroprotective effects of APC require PAR-1 and EPCR. Using murine APC and mice with targeted gene deletions of PARs or mice that were severely deficient in EPCR, it was shown that PAR-1 and EPCR are required for pharmacological beneficial effects of APC in vivo in mouse models for ischemic stroke.16,31,60,70  On murine neurons, PAR-3 seems to play a supporting role in APC-mediated cytoprotective effects in addition to PAR-1.31  Thrombin activates human platelets via PAR-1 and PAR-4, whereas it activates mouse platelets via PAR-3 and PAR-4.71,72  Curiously, murine PAR-3 is not directly activated by thrombin to generate cellular signaling, but rather it is hypothesized to serve as a cofactor for thrombin's proteolytic activation of murine PAR-4.72  These species differences might complicate comparison of results for the mechanistic roles of specific PARs based on studies using murine injury models compared with human cells or with results observed for humans. Nonetheless, it is clear that the beneficial pharmacologic effects of APC in various injury models require PAR-1 where the mechanism has been carefully assessed.

The observation that the beneficial pharmacologic effects of APC require PAR-1 may seem somewhat surprising as PAR-1, the archetype thrombin receptor, is often described as mediating proinflammatory signals.73  Curiously, a gene expression profiling study of TNFα-perturbed endothelial cells demonstrated some specific PAR-1–dependent effects of APC that were opposite those of PAR-1–dependent effects of thrombin.30  Strikingly, both p53 and thrombospondin-1 expression were down-regulated by APC but up-regulated by thrombin, showing that this same receptor on the same cell type can mediate opposite biologic effects by 2 different plasma serine proteases.30  One may wonder whether the cytoprotective effects of APC might also be achieved by other PAR-1 agonists or whether they are specific to APC's interactions with PAR-1. A PAR-1 agonist peptide showed partial reduction of hypoxia-induced apoptosis of human brain endothelial cells and of N-methyl-D-aspartate (NMDA)–induced apoptosis of murine cortical neurons, but it failed to inhibit staurosporine-induced apoptosis of transformed endothelial cells.16,18  Furthermore, thrombin did not protect against staurosporine-induced endothelial cell apoptosis.18  These various observations illustrate that the molecular mechanisms for the effects caused by PAR-1 activation on cells are incompletely resolved; although the requirement for PAR-1 in the pharmacologic actions of APC has been proven, a role for PAR-1 in the physiologic actions of APC is a subject of current debate.74-76 

Beneficial effects of APC have been demonstrated in severe sepsis trials in humans and a variety of animal injury model systems (Table 2). As discussed in the following subsections, current and potential therapeutic applications for APC's beneficial effects include, among others, severe sepsis, lung injury and inflammation, ischemic stroke, and angiogenesis and wound healing.

Table 2

In vivo APC beneficial effects

Model, hostTriggerRead outAPC effectSources
Sepsis     
    Human Sepsis (high risk of death) Death Inhibits Bernard et al,24  Bernard et al77  
    Human Sepsis (low risk of death) Death No effect Abraham et al78  
    Human LPS Thrombosis/inflammation No effect Kalil et al79  
    Baboon* E coli Death Inhibits Taylor et al80  
    Hamster LPS Inflammation Inhibits Hoffmann et al81  
    Rat LPS Inflammation Inhibits Iba et al82  
    Rat LPS Arterial blood pressure Increases Isobe et al83  
    Rat LPS Cardiovascular dysfunction/inflammation Inhibits Favory et al84  
Pulmonary inflammation and injury     
    Human LPS Thrombosis Inhibits Van der Poll et al85  
    Human LPS Inflammation Inhibits Nick et al45  
    Rat LPS Inflammation Inhibits Murakami et al44  
    Rat Acid aspiration Inflammation Inhibits Jian et al86  
    Rat Smoke Inflammation Inhibits Wong et al87  
    Mouse* Bleomycin Fibrosis/inflammation Inhibits Shimizu et al,40  Yasui et al88  
    Mouse* Ovalbumin Inflammation Inhibits Yuda et al41  
Stroke     
    Rabbit I/R Neurologic injury Inhibits Yamauchi et al89  
    Rat I/R Neurologic injury/thrombosis/inflammation Inhibits Zlokovic et al90  
    Mouse* I/R Neurologic injury/thrombosis/inflammation Inhibits Cheng et al,16  Cheng et al,70  Zlokovic et al,90  Fernández et al,91  Shibata et al92  
    Mouse NMDA Apoptosis/neurologic injury Inhibits Guo et al,31  Liu et al60  
Wound healing/angiogenesis     
    Rat Biopsy Wound closure Promotes Xue et al,93  Jackson et al94  
    Mouse None Capillary formation Promotes Uchiba et al95  
    Chick None Capillary formation Promotes Xue et al93  
Miscellaneous     
    Intestinal     
        Rat I/R Inflammation Inhibits Schoots et al96  
    Acute pancreatitis     
        Rat Taurocholate Inflammation Inhibits Yamenel et al97  
    Gastric injury     
        Rat Stress Thrombosis/inflammation Inhibits Isobe et al98  
    Spinal cord injury     
        Rat Compression Motor disturbance/inflammation Inhibits Taoka et al56  
    Islet transplantation     
        Mouse Transplantation Apoptosis/inflammation Inhibits Contreras et al99  
Model, hostTriggerRead outAPC effectSources
Sepsis     
    Human Sepsis (high risk of death) Death Inhibits Bernard et al,24  Bernard et al77  
    Human Sepsis (low risk of death) Death No effect Abraham et al78  
    Human LPS Thrombosis/inflammation No effect Kalil et al79  
    Baboon* E coli Death Inhibits Taylor et al80  
    Hamster LPS Inflammation Inhibits Hoffmann et al81  
    Rat LPS Inflammation Inhibits Iba et al82  
    Rat LPS Arterial blood pressure Increases Isobe et al83  
    Rat LPS Cardiovascular dysfunction/inflammation Inhibits Favory et al84  
Pulmonary inflammation and injury     
    Human LPS Thrombosis Inhibits Van der Poll et al85  
    Human LPS Inflammation Inhibits Nick et al45  
    Rat LPS Inflammation Inhibits Murakami et al44  
    Rat Acid aspiration Inflammation Inhibits Jian et al86  
    Rat Smoke Inflammation Inhibits Wong et al87  
    Mouse* Bleomycin Fibrosis/inflammation Inhibits Shimizu et al,40  Yasui et al88  
    Mouse* Ovalbumin Inflammation Inhibits Yuda et al41  
Stroke     
    Rabbit I/R Neurologic injury Inhibits Yamauchi et al89  
    Rat I/R Neurologic injury/thrombosis/inflammation Inhibits Zlokovic et al90  
    Mouse* I/R Neurologic injury/thrombosis/inflammation Inhibits Cheng et al,16  Cheng et al,70  Zlokovic et al,90  Fernández et al,91  Shibata et al92  
    Mouse NMDA Apoptosis/neurologic injury Inhibits Guo et al,31  Liu et al60  
Wound healing/angiogenesis     
    Rat Biopsy Wound closure Promotes Xue et al,93  Jackson et al94  
    Mouse None Capillary formation Promotes Uchiba et al95  
    Chick None Capillary formation Promotes Xue et al93  
Miscellaneous     
    Intestinal     
        Rat I/R Inflammation Inhibits Schoots et al96  
    Acute pancreatitis     
        Rat Taurocholate Inflammation Inhibits Yamenel et al97  
    Gastric injury     
        Rat Stress Thrombosis/inflammation Inhibits Isobe et al98  
    Spinal cord injury     
        Rat Compression Motor disturbance/inflammation Inhibits Taoka et al56  
    Islet transplantation     
        Mouse Transplantation Apoptosis/inflammation Inhibits Contreras et al99  

LPS indicates lipopolysaccharide; I/R, ischemia/reperfusion; and NMDA, N-methyl-D-aspartate.

*

EPCR required.

PAR-1 required.

PAR-3 required.

APC in severe sepsis

The landmark Protein C Evaluation in Severe Sepsis (PROWESS) study24  was the first positive study in recent years of clinical sepsis trials that demonstrated a significant reduction of 28-day all-cause mortality in patients with severe sepsis. In this randomized, double-blind, placebo-controlled multicenter trial, drotrecogin alpha-activated (recombinant APC) reduced the risk of death by 6.1% (relative risk reduction, 19.4% [95% confident interval, 6.6% to 30.5%]) when given for 4 days by infusion at 24 μg/kg per hour (plasma APC levels averaging ∼ 50 ng/mL). These results were especially remarkable since 2 other potent anticoagulant proteins, tissue factor pathway inhibitor (TFPI) and antithrombin III (ATIII), failed to reduce mortality in similar phase 3 clinical trials (OPTIMIST and KyberSept).25,26  This reasonably implies that APC anticoagulant activity alone does not explain APC's success in reducing mortality. Thus, this implies that APC's anti-inflammatory and antiapoptotic activities are relevant for reduction of mortality.

The PROWESS trial results for severe sepsis patients were reported to be confirmed in the ENHANCE US trial.77  Nevertheless, the absence of an effect of APC on mortality in patients with sepsis with a low risk of death in the ADDRESS trial and in a pediatric trial clearly indicates that the currently used APC therapeutic regimen has its limitations.78  A low but significant increased risk of serious bleeding is associated with the PROWESS regimen for APC infusion (24 μg/kg per hour for 96 hours). It is unclear whether APC administration at higher doses and/or for shorter periods might have less risk of serious bleeding and equivalent or greater efficacy. Almost none of the protocols for APC's beneficial effects in animal injury models used a protocol like that of the PROWESS trial, and serious bleeding was not a notable finding in any of the reported animal injury studies. This suggests that APC dosage regimens that have not yet been tested in humans, for example, bolus or bolus plus brief infusions (0.5 to 4 hours) at APC doses higher than 24 μg/kg per hour, might have less side effects and equivalent or greater efficacy. Moreover, as discussed in “Roles for APC's anticoagulant activity versus APC's cytoprotective activities,” genetically engineered variants of APC with reduced anticoagulant activity but with normal cytoprotective activity might provide superior properties in clinical situations, such as severe sepsis; such APC variants with attenuated anticoagulant activity might permit higher doses given by bolus or short infusions at different times with less risk of serious bleeding.

APC in lung injury and inflammation

APC's beneficial actions against inflammatory injury were shown in models of lung injury. Clinical studies show that reduced plasma protein C levels and impaired APC generation in the intra-alveolar space and bronchoalveolar lavage fluid in patients with lung injury and airway inflammation were associated with increased collagen deposition in the lung and worse clinical outcome.100,101  In a mouse bleomycin-induced lung injury model, APC inhibited development of lung fibrosis and, in a mouse ovalbumin-induced bronchial asthma model, APC prevented the development of an allergic inflammation.41,88  Furthermore, in the lungs of bleomycin-challenged mice, APC inhibited platelet-derived growth factor effects that include potent mitogenic action, chemoattractant activity, and induction of matrix-related gene expression (eg, collagen and fibronectin).40  When APC is given at pharmacologic doses, part of APC's protective effects on lung injury and inflammation is mediated through APC-dependent inhibition of leukocyte accumulation and chemotaxis.44,45  These studies may provide novel promising directions for treatments of lung inflammation, fibrosis, and asthma. Of interest, airway epithelial cells express protein C, thrombomodulin, and EPCR, and these cells support generation of APC in the presence of thrombin.102  Thus, it is reasonable that the endogenous protein C system, as well as pharmacologically administered APC, can help protect the lung against injury.

APC in ischemic stroke

Based on preclinical animal studies and observational human studies, a very promising untested area for potential APC human therapy involves ischemic stroke.103  The prospective epidemiologic Atherosclerosis Risk in Communities (ARIC) study reported that plasma protein C appeared protective against ischemic stroke (OR = 0.65 [0.4-1.0]).104  In murine models of focal cerebral ischemia, APC provided remarkable anti-inflammatory and neuroprotective effects.16,91,92  The neuroprotective action of APC was apparent in both rat and murine stroke models even when APC was given 6 hours after onset of brain ischemia.90 

tPA, one of the few available treatment options in ischemic stroke, effectively promotes plasmin-dependent lysis of thrombotic occlusions, but this drug has notable side effects. Not only is tPA directly neurotoxic to neurovascular cells as it induces apoptosis in brain endothelial cells and neurons, but tPA also aggravates the risk for hemorrhagic transformation due to up-regulation and activation of pro–matrix metalloprotease 9 (pro-MMP9), especially when tPA is administered after 3 hours of onset of stroke. A potential breakthrough for ischemic stroke treatment is also implicit in recent in vivo studies of combined APC and tPA given for neuroprotection during murine ischemic stroke and for prevention of excitotoxic injury following administration of NMDA.60,70  In both rat and murine stroke models, the cytoprotective activities of APC effectively counteract tPA's neurotoxic effects on endothelial cells, the blood-brain barrier, and neurons, thus improving the overall beneficial activity of tPA.60,70  APC prevents tPA-induced bleeding by specifically counteracting tPA's up-regulation of pro-MMP9 expression in stressed human brain endothelium in vitro and in hypoxic murine brain in vivo.70  In summary, APC appears promising for treatment of ischemic stroke in the absence as well as in the presence of tPA.

APC in angiogenesis, wound healing, and inflammation

APC can induce endothelial cell proliferation and angiogenesis in vitro and in vivo, which is dependent on EPCR and PAR-1.95  APC can stimulate keratinocyte proliferation and migration and APC can contribute to extracellular matrix degradation by activation of the gelatinase, matrix metalloproteinase-2 (MMP-2).105,106  Thus, beneficial effects of APC in regenerative processes (eg, angiogenesis and wound healing) have been demonstrated. Furthermore, APC may induce wound repair and promote endothelial cell migration and proliferation via up-regulation of endothelial cytokines, as was shown for interleukin-6 (IL-6) and IL-8, or chemokines such as MCP-1.19,35,49,107  In certain conditions, APC can up-regulate MCP-1 expression at both the RNA and protein levels by suppression of nitric oxide (NO) synthesis and stabilization of mRNA.36,107  However, a clear application for pharmacologic APC in the setting of angiogenesis or wound healing is less obvious than in the settings of severe sepsis or ischemic stroke (see “APC in severe sepsis” and “APC in ischemic stroke”). The intimate relationships between inflammation and host defense and between inflammation and wound healing are complex, and further work is needed to clarify the physiologic or pharmacologic roles for APC in these processes.

Given the widely observed beneficial effects of APC in humans and in various animal injury models (Table 2) and given the multiplicity of APC activities (Figure 1), one wonders, what are the mechanistic secrets to APC's successes? For example, is the anticoagulant activity of APC essential both for success in the PROWESS trial and also for the increased risk for serious bleeding in severe sepsis patients, especially during the 4-day period of APC infusion.24,108  Alternatively, given the failure of other anticoagulants (eg, antithrombin and tissue factor pathway inhibitor) in severe sepsis trials, is one or more of the direct actions of APC on cells critical to reducing mortality in patients? And in various kinds of injury, what activities of APC are critical and which are dispensable? To help address these questions about APC's mechanisms of action and, ultimately, to obtain safer yet effective APC variants, we have undertaken a series of in vitro and in vivo studies of engineered APC variants.29 

Our first goal was to construct potentially safer APC variants with reduced risk of bleeding due to reduced anticoagulant activity that retained normal direct actions on cells, so we altered factor Va binding exosites in APC without affecting exosites that recognize PAR-1. The anticoagulant action of APC critically involves a cleavage site at Arg506 in factor Va, and this cleavage depends on certain positively charged residues in surface loops on APC's protease domain, including loop 37 (protein C residues 190-193, equivalent to chymotrypsin [CHT] residues 36-39), the Ca++-binding loop (residues 225-235, CHT residues 70-80), and the autolysis loop (residues 301-316, CHT residues 142-153) (Figure 5A).112,114-119  Two APC variants were made with Ala mutations in 2 surface loops and involved Ala replacements of Arg229 and Arg230 and of Lys191, Lys192, and Lys193 (Figures 5A,C). Each APC variant had little anticoagulant activity (4% to 14% compared with wild-type APC) (Figure 5B), whereas each APC variant retained normal antiapoptotic activity, requiring PAR-1 and EPCR (Figure 5D).29  When the ratio of cytoprotective to anticoagulant activities of these APC variants was compared with wild-type APC (normalized to a ratio of 1:1), the 2 APC variants designated 3K3A-APC and 229/230-APC exhibited 25-times and 7-times greater antiapoptotic activity relative to that of wild-type APC (Figure 5E). Thus, genetic engineering of APC can reduce anticoagulant activity while preserving APC's direct actions on cells that are EPCR-dependent and that require PAR-1.

Figure 5

APC cytoprotective activities mediated by the cytoprotective protein C pathway are independent of APC anticoagulant activity. APC anticoagulant activity requires binding of the APC Gla-domain (green) (A) to negatively charged phospholipids on cell and/or platelet membranes or microparticles, whereas APC cytoprotective activity requires binding of the APC Gla-domain (green) to EPCR109  (purple) (C). EPCR residues important for protein C/APC binding, as determined by alanine mutagenesis, are indicated in red (C).110  Interactions of APC with its macromolecular substrates (factor Va, factor VIIIa, and PAR-1) are dictated by exosite interactions of APC protease domain surface loops (A,C). Important exosite loops in APC for interactions with factor Va are the autolysis loop, the 37-loop (yellow), and the calcium-binding loop (yellow). In particular, residues Lys191, Lys192, and Lys193 (red; A,C) in the 37-loop and residues Arg229 and Arg230 (red; A,C) in the calcium-binding loop contribute significantly to interactions with factor Va, and mutation of these residues to Ala severely compromises APC anticoagulant activity (B). Exosite specificity is illustrated by the fact that both the 37-loop residues (red/yellow) and the calcium-binding loop residues (red/yellow) (A,C) are required for normal cleavage of factor Va at Arg506 and thus for normal anticoagulant activity (B), while they are not involved in proteolytic activation of PAR-1 that generates antiapoptotic, cytoprotective activity (D). Based on the exosite specificity, a molecular engineering approach of APC surface loops was used to design APC variants with reduced risk of bleeding due to reduced anticoagulant activity but with full cytoprotective activity.29  Two such APC variants, 3K3A-APC and 229/230-APC, displayed increased ratios of cytoprotective (antiapoptotic) to anticoagulant activities (7- to 25-fold) compared with the reference ratio of 1.0 for recombinant wild-type APC (rwt-APC) (E). The ratio for a catalytically inactive mutant, S360A-APC (active site residue Ser360 mutated to Ala), is 0.0 (E). Data (B,D,E) were from Mosnier et al.29  Deep View Swiss PDB viewer (http://www.expasy.org/spdbv/) was used for the structural alignment. (A) The model of full-length APC111  is based on the serine protease domain structure of APC (Protein Data Bank entry 1AUT112 ). (C) The Gla-domain of the model of full-length APC was aligned with the protein C Gla-domain peptide crystallized in complex with soluble EPCR (1LQV).111,113  The model of full-length EPCR (kindly provided by Dr B. Dahlbäck, http://www.klkemi.mas.lu.se/dahlback/index.php) was aligned with the crystal structure of soluble EPCR to obtain a model representation of APC bound to EPCR on the endothelial cell membrane. (B, D) Error bars indicate SEM.

Figure 5

APC cytoprotective activities mediated by the cytoprotective protein C pathway are independent of APC anticoagulant activity. APC anticoagulant activity requires binding of the APC Gla-domain (green) (A) to negatively charged phospholipids on cell and/or platelet membranes or microparticles, whereas APC cytoprotective activity requires binding of the APC Gla-domain (green) to EPCR109  (purple) (C). EPCR residues important for protein C/APC binding, as determined by alanine mutagenesis, are indicated in red (C).110  Interactions of APC with its macromolecular substrates (factor Va, factor VIIIa, and PAR-1) are dictated by exosite interactions of APC protease domain surface loops (A,C). Important exosite loops in APC for interactions with factor Va are the autolysis loop, the 37-loop (yellow), and the calcium-binding loop (yellow). In particular, residues Lys191, Lys192, and Lys193 (red; A,C) in the 37-loop and residues Arg229 and Arg230 (red; A,C) in the calcium-binding loop contribute significantly to interactions with factor Va, and mutation of these residues to Ala severely compromises APC anticoagulant activity (B). Exosite specificity is illustrated by the fact that both the 37-loop residues (red/yellow) and the calcium-binding loop residues (red/yellow) (A,C) are required for normal cleavage of factor Va at Arg506 and thus for normal anticoagulant activity (B), while they are not involved in proteolytic activation of PAR-1 that generates antiapoptotic, cytoprotective activity (D). Based on the exosite specificity, a molecular engineering approach of APC surface loops was used to design APC variants with reduced risk of bleeding due to reduced anticoagulant activity but with full cytoprotective activity.29  Two such APC variants, 3K3A-APC and 229/230-APC, displayed increased ratios of cytoprotective (antiapoptotic) to anticoagulant activities (7- to 25-fold) compared with the reference ratio of 1.0 for recombinant wild-type APC (rwt-APC) (E). The ratio for a catalytically inactive mutant, S360A-APC (active site residue Ser360 mutated to Ala), is 0.0 (E). Data (B,D,E) were from Mosnier et al.29  Deep View Swiss PDB viewer (http://www.expasy.org/spdbv/) was used for the structural alignment. (A) The model of full-length APC111  is based on the serine protease domain structure of APC (Protein Data Bank entry 1AUT112 ). (C) The Gla-domain of the model of full-length APC was aligned with the protein C Gla-domain peptide crystallized in complex with soluble EPCR (1LQV).111,113  The model of full-length EPCR (kindly provided by Dr B. Dahlbäck, http://www.klkemi.mas.lu.se/dahlback/index.php) was aligned with the crystal structure of soluble EPCR to obtain a model representation of APC bound to EPCR on the endothelial cell membrane. (B, D) Error bars indicate SEM.

Close modal

The availability of APC variants such as 3K3A-APC and 229/230-APC with reduced anticoagulant activity but normal cytoprotective activities for preclinical and clinical studies can help clarify whether the anticoagulant activity or the cytoprotective actions of APC or both are crucial for its various beneficial properties. In preliminary studies, the recombinant murine 3K3A-APC variant appeared as active as wild-type murine APC in preventing death from endotoxemia in mice,27,28  and this variant was also similar to wild-type APC in its potency to provide neuroprotection in murine ischemic stroke models (B.V.Z., L.O.M., J.H.G., unpublished results, March 2005). These data suggest a primary role for APC's in vivo beneficial activities in these injury models. If these preliminary findings for the mechanism of APC's beneficial actions in sepsis and stroke are confirmed to be of primary importance in additional studies, then one might use such APC variants for therapy and minimize risk of serious bleeding. Moreover, safer APC variants with less bleeding risk may allow studies of higher APC doses for shorter times to evaluate the possibility that higher doses of APC may stabilize stressed cells at risk for excessive inflammation or apoptosis.

Inflammation, apoptosis, and thrombosis are intimately connected in the tangled web of host defense. Many inflammatory and thrombotic feedback loops cooperate to exacerbate pathogenic reactions. Both basic, preclinical and clinical studies have provided an expanding body of research on the cytoprotective protein C pathway that is distinct from the anticoagulant protein C pathway. Although much remains to be characterized about the multiple activities of APC, the pharmacologic use of this fascinating enzyme holds remarkable promise. Therapeutic APC variants may be engineered with decreased anticoagulant activity but normal cytoprotective activity, and such APC variants hold the promise of reducing risks for serious bleeding while retaining this agent's beneficial cytoprotective properties.

Conflict-of-interest disclosure: B.V.Z. is scientific founder of Socratech LLC, a startup biotech company with intellectual property involving activated protein C. The Scripps Research Institute has patents or patent applications related to activated protein C. The other authors declare no competing financial interests.

Correspondence: John H. Griffin, Department of Molecular and Experimental Medicine (MEM-180), The Scripps Research Institute, 10550 North Torrey Pines Rd, La Jolla, CA; e-mail: jgriffin@scripps.edu.

This work was supported by National Heart, Lung, and Blood Institute (NHLBI) grants (HL21544, HL31950, HL52246, and HL87618) and by a Basic Research Scholar Award from the American Society of Hematology (L.O.M.).

We gratefully acknowledge members of the Griffin and Zlokovic laboratories for their helpful discussions and ongoing research contributions to the field of the cytoprotective protein C pathway. We apologize to our respected colleagues whose important publications we were not able to cite or discuss due to space limitations.

1
Gruber A and Griffin JH. Direct detection of activated protein C in blood from human subjects.
Blood
1992
;
79
:
2340
–2348.
2
Griffin JH, Evatt B, Zimmerman TS, Kleiss AJ, Wideman C. Deficiency of protein C in congenital thrombotic disease.
J Clin Invest
1981
;
68
:
1370
–1373.
3
Branson HE, Katz J, Marble R, Griffin JH. Inherited protein C deficiency and coumarin-responsive chronic relapsing purpura fulminans in a newborn infant.
Lancet
1983
;
2
:
1165
–1168.
4
Dahlbäck B, Carlsson M, Svensson PJ. Familial thrombophilia due to a previously unrecognized mechanism characterized by poor anticoagulant response to activated protein C: prediction of a cofactor to activated protein C.
Proc Natl Acad Sci U S A
1993
;
90
:
1004
–1008.
5
Greengard JS, Sun X, Xu X, et al. Activated protein C resistance caused by Arg506Gln mutation in factor Va.
Lancet
1994
;
343
:
1361
–1362.
6
Bertina RM, Koeleman BPC, Koster T, et al. Mutations in blood coagulation factor V associated with resistance to activated protein C.
Nature
1994
;
369
:
64
–67.
7
Jalbert LR, Rosen ED, Moons L, et al. Inactivation of the gene for anticoagulant protein C causes lethal perinatal consumptive coagulopathy in mice.
J Clin Invest
1998
;
102
:
1481
–1488.
8
Castellino FJ. Gene targeting in hemostasis: protein C.
Front Biosci
2001
;
6
:
D807
–D819.
9
Lay AJ, Liang Z, Rosen ED, Castellino FJ. Mice with a severe deficiency in protein C display prothrombotic and proinflammatory phenotypes and compromised maternal reproductive capabilities.
J Clin Invest
2005
;
115
:
1552
–1561.
10
Esmon CT. The protein C pathway.
Chest
2003
;
124
:
26S
–32S.
11
Stearns-Kurosawa DJ, Kurosawa S, Mollica JS, Ferrell GL, Esmon CT. The endothelial cell protein C receptor augments protein C activation by the thrombin-thrombomodulin complex.
Proc Natl Acad Sci U S A
1996
;
93
:
10212
–10216.
12
Dahlbäck B and Villoutreix BO. Regulation of blood coagulation by the protein C anticoagulant pathway: novel insights into structure-function relationships and molecular recognition.
Arterioscler Thromb Vasc Biol
2005
;
25
:
1311
–1320.
13
Mosnier LO and Griffin JH. Protein C anticoagulant activity in relation to anti-inflammatory and anti-apoptotic activities.
Front Biosci
2006
;
11
:
2381
–2399.
14
Vu TK, Hung DT, Wheaton VI, Coughlin SR. Molecular cloning of a functional thrombin receptor reveals a novel proteolytic mechanism of receptor activation.
Cell
1991
;
64
:
1057
–1068.
15
Joyce DE, Gelbert L, Ciaccia A, DeHoff B, Grinnell BW. Gene expression profile of antithrombotic protein C defines new mechanisms modulating inflammation and apoptosis.
J Biol Chem
2001
;
276
:
11199
–11203.
16
Cheng T, Liu D, Griffin JH, et al. Activated protein C blocks p53-mediated apoptosis in ischemic human brain endothelium and is neuroprotective.
Nat Med
2003
;
9
:
338
–342.
17
Domotor E, Benzakour O, Griffin JH, et al. Activated protein C alters cytosolic calcium flux in human brain endothelium via binding to endothelial protein C receptor and activation of protease activated receptor-1.
Blood
2003
;
101
:
4797
–4801.
18
Mosnier LO and Griffin JH. Inhibition of staurosporine-induced apoptosis of endothelial cells by activated protein C requires protease activated receptor-1 and endothelial cell protein C receptor.
Biochem J
2003
;
373
:
65
–70.
19
Riewald M, Petrovan RJ, Donner A, Mueller BM, Ruf W. Activation of endothelial cell protease activated receptor 1 by the protein C pathway.
Science
2002
;
296
:
1880
–1882.
20
Mollica LR, Crawley JT, Liu K, et al. Role of a 5′-enhancer in the transcriptional regulation of the human endothelial cell protein C receptor gene.
Blood
2006
;
108
:
1251
–1259.
21
Crawley JT, Gu JM, Ferrell G, Esmon CT. Distribution of endothelial cell protein C/activated protein C receptor (EPCR) during mouse embryo development.
Thromb Haemost
2002
;
88
:
259
–266.
22
Laszik Z, Mitro A, Taylor FB Jr, Ferrell G, Esmon CT. Human protein C receptor is present primarily on endothelium of large blood vessels: implications for the control of the protein C pathway.
Circulation
1997
;
96
:
3633
–3640.
23
Balazs AB, Fabian AJ, Esmon CT, Mulligan RC. Endothelial protein C receptor (CD201) explicitly identifies hematopoietic stem cells in murine bone marrow.
Blood
2005
;
107
:
2317
–2321.
24
Bernard GR, Vincent JL, Laterre PF, et al. Efficacy and safety of recombinant human activated protein C for severe sepsis.
N Engl J Med
2001
;
344
:
699
–709.
25
Abraham E, Reinhart K, Opal S, et al. Efficacy and safety of tifacogin (recombinant tissue factor pathway inhibitor) in severe sepsis: a randomized controlled trial.
JAMA
2003
;
290
:
238
–247.
26
Warren BL, Eid A, Singer P, et al. High-dose antithrombin III in severe sepsis: a randomized controlled trial.
JAMA
2001
;
286
:
1869
–1878.
27
Kerschen EJ, Cooley BC, Castellino FJ, et al. Mechanisms for mortality reduction by activated protein C in severe sepsis [abstract].
Blood
2006
;
108
:
5a
Abstract 1.
28
Kerschen EJ, Cooley BC, Castellino FJ, Griffin JH, Weiler H. Protective effect of activated protein C in murine endotoxemia: mechanism of action [abstract].
Blood
2005
;
106
:
12a
Abstract 26.
29
Mosnier LO, Gale AJ, Yegneswaran S, Griffin JH. Activated protein C variants with normal cytoprotective but reduced anticoagulant activity.
Blood
2004
;
104
:
1740
–1745.
30
Riewald M and Ruf W. Protease-activated receptor-1 signaling by activated protein C in cytokine perturbed endothelial cells is distinct from thrombin signaling.
J Biol Chem
2005
;
280
:
19808
–19814.
31
Guo H, Liu D, Gelbard H, et al. Activated protein C prevents neuronal apoptosis via protease activated receptors 1 and 3.
Neuron
2004
;
41
:
563
–572.
32
Joyce DE and Grinnell BW. Recombinant human activated protein C attenuates the inflammatory response in endothelium and monocytes by modulating nuclear factor-kappaB.
Crit Care Med
2002
;
30
:
S288
–S293.
33
Franscini N, Bachli EB, Blau N, et al. Gene expression profiling of inflamed human endothelial cells and influence of activated protein C.
Circulation
2004
;
110
:
2903
–2909.
34
Brueckmann M, Nahrup AS, Lang S, et al. Recombinant human activated protein C upregulates the release of soluble fractalkine from human endothelial cells.
Br J Haematol
2006
;
133
:
550
–557.
35
Hooper WC, Phillips DJ, Renshaw MA, Evatt BL, Benson JM. The up-regulation of IL-6 and IL-8 in human endothelial cells by activated protein C.
J Immunol
1998
;
161
:
2567
–2573.
36
Brueckmann M, Lang S, Borggrefe M, Huhle G, Haase KK. Drotrecogin alfa (activated) prolongs half-life time of messenger RNA encoding for interleukin-8 and monocyte chemoattractant protein-1.
Thromb Haemost
2003
;
90
:
1223
–1226.
37
Brueckmann M, Horn S, Lang S, et al. Recombinant human activated protein C upregulates cyclooxygenase-2 expression in endothelial cells via binding to endothelial cell protein C receptor and activation of protease-activated receptor-1.
Thromb Haemost
2005
;
93
:
743
–750.
38
Stephenson DA, Toltl LJ, Beaudin S, Liaw PC. Modulation of monocyte function by activated protein C, a natural anticoagulant.
J Immunol
2006
;
177
:
2115
–2122.
39
Shu F, Kobayashi H, Fukudome K, et al. Activated protein C suppresses tissue factor expression on U937 cells in the endothelial protein C receptor-dependent manner.
FEBS Lett
2000
;
477
:
208
–212.
40
Shimizu S, Gabazza EC, Taguchi O, et al. Activated protein C inhibits the expression of platelet-derived growth factor in the lung.
Am J Respir Crit Care Med
2003
;
167
:
1416
–1426.
41
Yuda H, Adachi Y, Taguchi O, et al. Activated protein C inhibits bronchial hyperresponsiveness and Th2 cytokine expression in mice.
Blood
2004
;
103
:
2196
–2204.
42
Kurosawa S, Esmon CT, Stearns-Kurosawa DJ. The soluble endothelial protein C receptor binds to activated neutrophils: involvement of proteinase-3 and CD11b/CD18.
J Immunol
2000
;
165
:
4697
–4703.
43
Feistritzer C and Riewald M. Endothelial barrier protection by activated protein C through PAR1-dependent sphingosine 1-phosphate receptor-1 crossactivation.
Blood
2005
;
105
:
3178
–3184.
44
Murakami K, Okajima K, Uchiba M, et al. Activated protein C attenuates endotoxin-induced pulmonary vascular injury by inhibiting activated leukocytes in rats.
Blood
1996
;
87
:
642
–647.
45
Nick JA, Coldren CD, Geraci MW, et al. Recombinant human activated protein C reduces human endotoxin-induced pulmonary inflammation via inhibition of neutrophil chemotaxis.
Blood
2004
;
104
:
3878
–3885.
46
Finigan JH, Dudek SM, Singleton PA, et al. Activated protein C mediates novel lung endothelial barrier enhancement: role of sphingosine 1-phosphate receptor transactivation.
J Biol Chem
2005
;
280
:
17286
–17293.
47
Zeng W, Matter WF, Yan SB, et al. Effect of drotrecogin alfa (activated) on human endothelial cell permeability and Rho kinase signaling.
Crit Care Med
2004
;
32
:
S302
–S308.
48
Feistritzer C, Sturn DH, Kaneider NC, Djanani A, Wiedermann CJ. Endothelial protein C receptor-dependent inhibition of human eosinophil chemotaxis by protein C.
J Allergy Clin Immunol
2003
;
112
:
375
–381.
49
Brueckmann M, Hoffmann U, De Rossi L, et al. Activated protein C inhibits the release of macrophage inflammatory protein-1-alpha from THP-1 cells and from human monocytes.
Cytokine
2004
;
26
:
106
–113.
50
Grey ST, Tsuchida A, Hau H, et al. Selective inhibitory effects of the anticoagulant activated protein C on the responses of human mononuclear phagocytes to LPS, IFN-gamma, or phorbol ester.
J Immunol
1994
;
153
:
3664
–3672.
51
White B, Schmidt M, Murphy C, et al. Activated protein C inhibits lipopolysaccharide-induced nuclear translocation of nuclear factor kappaB (NF-kappaB) and tumour necrosis factor alpha (TNF-alpha) production in the THP-1 monocytic cell line.
Br J Haematol
2000
;
110
:
130
–134.
52
Yuksel M, Okajima K, Uchiba M, Horiuchi S, Okabe H. Activated protein C inhibits lipopolysaccharide-induced tumor necrosis factor-alpha production by inhibiting activation of both nuclear factor-kappa B and activator protein-1 in human monocytes.
Thromb Haemost
2002
;
88
:
267
–273.
53
Joyce DE, Nelson DR, Grinnell BW. Leukocyte and endothelial cell interactions in sepsis: relevance of the protein C pathway.
Crit Care Med
2004
;
32
:
S280
–S286.
54
Galligan L, Livingstone W, Volkov Y, et al. Characterization of protein C receptor expression in monocytes.
Br J Haematol
2001
;
115
:
408
–414.
55
Sturn DH, Kaneider NC, Feistritzer C, et al. Expression and function of the endothelial protein C receptor in human neutrophils.
Blood
2003
;
102
:
1499
–1505.
56
Taoka Y, Okajima K, Uchiba M, et al. Activated protein C reduces the severity of compression-induced spinal cord injury in rats by inhibiting activation of leukocytes.
J Neurosci
1998
;
18
:
1393
–1398.
57
Boatright KM and Salvesen GS. Mechanisms of caspase activation.
Curr Opin Cell Biol
2003
;
15
:
725
–731.
58
Reed JC. Proapoptotic multidomain Bcl-2/Bax-family proteins: mechanisms, physiological roles, and therapeutic opportunities.
Cell Death Differ
2006
;
13
:
1378
–1386.
59
Green DR. Apoptotic pathways: ten minutes to dead.
Cell
2005
;
121
:
671
–674.
60
Liu D, Cheng T, Guo H, et al. Tissue plasminogen activator neurovascular toxicity is controlled by activated protein C.
Nat Med
2004
;
10
:
1379
–1383.
61
Hotchkiss RS, Chang KC, Swanson PE, et al. Caspase inhibitors improve survival in sepsis: a critical role of the lymphocyte.
Nat Immunol
2000
;
1
:
496
–501.
62
McVerry BJ and Garcia JG. Endothelial cell barrier regulation by sphingosine 1-phosphate.
J Cell Biochem
2004
;
92
:
1075
–1085.
63
Taylor FB Jr. Staging of the pathophysiologic responses of the primate microvasculature to Escherichia coli and endotoxin: examination of the elements of the compensated response and their links to the corresponding uncompensated lethal variants.
Crit Care Med
2001
;
29
:
S78
–S89.
64
Burridge K and Wennerberg K. Rho and Rac take center stage.
Cell
2004
;
116
:
167
–179.
65
Singleton PA, Dudek SM, Chiang ET, Garcia JG. Regulation of sphingosine 1-phosphate-induced endothelial cytoskeletal rearrangement and barrier enhancement by S1P1 receptor, PI3 kinase, Tiam1/Rac1, and alpha-actinin.
FASEB J
2005
;
19
:
1646
–1656.
66
Feistritzer C, Schuepbach RA, Mosnier LO, et al. Protective signaling by activated protein C is mechanically linked to protein C activation on endothelial cells.
J Biol Chem
2006
;
281
:
20077
–20084.
67
Deguchi H, Fernández JA, Pabinger I, Heit JA, Griffin JH. Plasma glucosylceramide deficiency as potential risk factor for venous thrombosis and modulator of anticoagulant protein C pathway.
Blood
2001
;
97
:
1907
–1914.
68
Yegneswaran S, Deguchi H, Griffin JH. Glucosylceramide, a neutral glycosphingolipid anticoagulant cofactor, enhances the interaction of human- and bovine-activated protein C with negatively charged phospholipid vesicles.
J Biol Chem
2003
;
278
:
14614
–14621.
69
Coughlin SR. Thrombin signaling and protease-activated receptors.
Nature
2000
;
407
:
258
–264.
70
Cheng T, Petraglia AL, Li Z, et al. Activated protein C inhibits tissue plasminogen activator-induced brain hemorrhage.
Nat Med
2006
;
12
:
1278
–1285.
71
Kahn ML, Nakanishi-Matsui M, Shapiro MJ, Ishihara H, Coughlin SR. Protease-activated receptors 1 and 4 mediate activation of human platelets by thrombin.
J Clin Invest
1999
;
103
:
879
–887.
72
Nakanishi-Matsui M, Zheng YW, Sulciner DJ, et al. PAR3 is a cofactor for PAR4 activation by thrombin.
Nature
2000
;
404
:
609
–613.
73
Coughlin SR and Camerer E. PARticipation in inflammation.
J Clin Invest
2003
;
111
:
25
–27.
74
Ruf W. Is APC activation of endothelial cell PAR1 important in severe sepsis? yes.
J Thromb Haemost
2005
;
3
:
1912
–1914.
75
Esmon CT. Is APC activation of endothelial cell PAR1 important in severe sepsis? no.
J Thromb Haemost
2005
;
3
:
1910
–1911.
76
Ludeman MJ, Kataoka H, Srinivasan Y, et al. PAR1 cleavage and signaling in response to activated protein C and thrombin.
J Biol Chem
2005
;
280
:
13122
–13128.
77
Bernard GR, Margolis BD, Shanies HM, et al. Extended evaluation of recombinant human activated protein C United States trial (ENHANCE US): a single-arm, phase 3B, multicenter study of drotrecogin alfa (activated) in severe sepsis.
Chest
2004
;
125
:
2206
–2216.
78
Abraham E, Laterre PF, Garg R, et al. Drotrecogin alfa (activated) for adults with severe sepsis and a low risk of death.
N Engl J Med
2005
;
353
:
1332
–1341.
79
Kalil AC, Coyle SM, Um JY, et al. Effects of drotrecogin alfa (activated) in human endotoxemia.
Shock
2004
;
21
:
222
–229.
80
Taylor FB Jr, Chang AC, Esmon CT, et al. Protein C prevents the coagulopathic and lethal effects of Escherichia coli infusion in the baboon.
J Clin Invest
1987
;
79
:
918
–925.
81
Hoffmann JN, Vollmar B, Laschke MW, et al. Microhemodynamic and cellular mechanisms of activated protein C action during endotoxemia.
Crit Care Med
2004
;
32
:
1011
–1017.
82
Iba T, Kidokoro A, Fukunaga M, et al. Activated protein C improves the visceral microcirculation by attenuating the leukocyte-endothelial interaction in a rat lipopolysaccharide model.
Crit Care Med
2005
;
33
:
368
–372.
83
Isobe H, Okajima K, Uchiba M, et al. Activated protein C prevents endotoxin-induced hypotension in rats by inhibiting excessive production of nitric oxide.
Circulation
2001
;
104
:
1171
–1175.
84
Favory R, Lancel S, Marechal X, Tissier S, Neviere R. Cardiovascular protective role for activated protein C during endotoxemia in rats.
Intensive Care Med
2006
;
32
:
899
–905.
85
van der Poll T, Levi M, Nick JA, Abraham E. Activated protein C inhibits local coagulation after intrapulmonary delivery of endotoxin in humans.
Am J Respir Crit Care Med
2005
;
171
:
1125
–1128.
86
Jian MY, Koizumi T, Tsushima K, Fujimoto K, Kubo K. Activated protein C attenuates acid-aspiration lung injury in rats.
Pulm Pharmacol Ther
2005
;
18
:
291
–296.
87
Wong SS, Sun NN, Hyde JD. Drotrecogin alfa (activated) prevents smoke-induced increases in pulmonary microvascular permeability and proinflammatory cytokine IL-1beta in rats.
Lung
2004
;
182
:
319
–330.
88
Yasui H, Gabazza EC, Tamaki S, et al. Intratracheal administration of activated protein C inhibits bleomycin-induced lung fibrosis in the mouse.
Am J Respir Crit Care Med
2001
;
163
:
1660
–1668.
89
Yamauchi T, Sakurai M, Abe K, Takano H, Sawa Y. Neuroprotective effects of activated protein C through induction of insulin-like growth factor-1 (IGF-1), IGF-1 receptor, and its downstream signal phosphorylated serine-threonine kinase after spinal cord ischemia in rabbits.
Stroke
2006
;
37
:
1081
–1086.
90
Zlokovic BV, Zhang C, Liu D, et al. Functional recovery after embolic stroke in rodents by activated protein C.
Ann Neurol
2005
;
58
:
474
–477.
91
Fernández JA, Xu X, Liu D, Zlokovic BV, Griffin JH. Recombinant murine-activated protein C is neuroprotective in a murine ischemic stroke model.
Blood Cells Mol Dis
2003
;
30
:
271
–276.
92
Shibata M, Kumar SR, Amar A, et al. Anti-inflammatory, antithrombotic, and neuroprotective effects of activated protein C in a murine model of focal ischemic stroke.
Circulation
2001
;
103
:
1799
–1805.
93
Xue M, Thompson P, Sambrook PN, March L, Jackson CJ. Activated protein C stimulates expression of angiogenic factors in human skin cells, angiogenesis in the chick embryo and cutaneous wound healing in rodents.
Clin Hemorheol Microcirc
2006
;
34
:
153
–161.
94
Jackson CJ, Xue M, Thompson P, et al. Activated protein C prevents inflammation yet stimulates angiogenesis to promote cutaneous wound healing.
Wound Repair Regen
2005
;
13
:
284
–294.
95
Uchiba M, Okajima K, Oike Y, et al. Activated protein C induces endothelial cell proliferation by mitogen-activated protein kinase activation in vitro and angiogenesis in vivo.
Circ Res
2004
;
95
:
34
–41.
96
Schoots IG, Levi M, van Vliet AK, et al. Inhibition of coagulation and inflammation by activated protein C or antithrombin reduces intestinal ischemia/reperfusion injury in rats.
Crit Care Med
2004
;
32
:
1375
–1383.
97
Yamenel L, Mas MR, Comert B, et al. The effect of activated protein C on experimental acute necrotizing pancreatitis.
Crit Care
2005
;
9
:
R184
–R190.
98
Isobe H, Okajima K, Harada N, Liu W, Okabe H. Activated protein C reduces stress-induced gastric mucosal injury in rats by inhibiting the endothelial cell injury.
J Thromb Haemost
2004
;
2
:
313
–320.
99
Contreras JL, Eckstein C, Smyth CA, et al. Activated protein C preserves functional islet mass after intraportal transplantation: a novel link between endothelial cell activation, thrombosis, inflammation, and islet cell death.
Diabetes
2004
;
53
:
2804
–2814.
100
Yasui H, Gabazza EC, Taguchi O, et al. Decreased protein C activation is associated with abnormal collagen turnover in the intraalveolar space of patients with interstitial lung disease.
Clin Appl Thromb Hemost
2000
;
6
:
202
–205.
101
Ware LB, Fang X, Matthay MA. Protein C and thrombomodulin in human acute lung injury.
Am J Physiol Lung Cell Mol Physiol
2003
;
285
:
L514
–L521.
102
Hataji O, Taguchi O, Gabazza EC, et al. Activation of protein C pathway in the airways.
Lung
2002
;
180
:
47
–59.
103
Griffin JH, Fernández JA, Liu D, et al. Activated protein C and ischemic stroke.
Crit Care Med
2004
;
32
:
S247
–S253.
104
Folsom AR, Rosamond WD, Shahar E, et al. Prospective study of markers of hemostatic function with risk of ischemic stroke: The Atherosclerosis Risk in Communities (ARIC) Study Investigators.
Circulation
1999
;
100
:
736
–742.
105
Xue M, Thompson P, Kelso I, Jackson C. Activated protein C stimulates proliferation, migration and wound closure, inhibits apoptosis and upregulates MMP-2 activity in cultured human keratinocytes.
Exp Cell Res
2004
;
299
:
119
–127.
106
Nguyen N, Arkell J, Jackson CJ. Activated protein C directly activates human endothelial gelatinase A.
J Biol Chem
2000
;
275
:
9095
–9098.
107
Hooper WC, Phillips DJ, Renshaw MA. Activated protein C induction of MCP-1 in human endothelial cells: a possible role for endothelial cell nitric cxide synthase.
Thromb Res
2001
;
103
:
209
–219.
108
Bernard GR, Macias WL, Joyce DE, et al. Safety assessment of drotrecogin alfa (activated) in the treatment of adult patients with severe sepsis.
Crit Care
2003
;
7
:
155
–163.
109
Preston RJ, Ajzner E, Razzari C, et al. Multifunctional specificity of the protein C/activated protein C GLA domain.
J Biol Chem
2006
;
281
:
28850
–28857.
110
Liaw PCY, Mather T, Oganesyan N, Ferrell GL, Esmon CT. Identification of the protein C/activated protein C binding sites on the endothelial cell protein C receptor: implications for a novel mode of ligand recognition by a major histocompatibility complex class 1- type receptor.
J Biol Chem
2001
;
276
:
8364
–8370.
111
Pellequer JL, Gale AJ, Getzoff ED, Griffin JH. Three-dimensional model of coagulation factor Va bound to activated protein C.
Thromb Haemost
2000
;
84
:
849
–857.
112
Mather T, Oganessyan V, Hof P, et al. The 2.8 Å crystal structure of Gla-domainless activated protein C.
EMBO J
1996
;
15
:
6822
–6831.
113
Oganesyan V, Oganesyan N, Terzyan S, et al. The crystal structure of the endothelial protein C receptor and a bound phospholipid.
J Biol Chem
2002
;
277
:
24851
–24854.
114
Gale AJ and Griffin JH. Characterization of a thrombomodulin binding site on protein C and its comparison to an activated protein C binding site for factor Va.
Proteins
2004
;
54
:
433
–441.
115
Gale AJ, Tsavaler A, Griffin JH. Molecular characterization of an extended binding site for coagulation factor Va in the positive exosite of activated protein C.
J Biol Chem
2002
;
277
:
28836
–28840.
116
Gale AJ, Heeb MJ, Griffin JH. The autolysis loop of activated protein C interacts with factor Va and differentiates between the Arg506 and Arg306 cleavage sites.
Blood
2000
;
96
:
585
–593.
117
Shen L, Villoutreix BO, Dahlbäck B. Involvement of lys 62(217) and lys 63(218) of human anticoagulant protein C in heparin stimulation of inhibition by the protein C inhibitor.
Thromb Haemost
1999
;
82
:
72
–79.
118
Friedrich U, Nicolaes GAF, Villoutreix BO, Dahlbäck B. Secondary substrate-binding exosite in the serine protease domain of activated protein C important for cleavage at Arg-506 but not at Arg-306 in factor Va.
J Biol Chem
2001
;
276
:
23105
–23108.
119
Rezaie AR. Exosite-dependent regulation of the protein C anticoagulant pathway.
Trends Cardiovasc Med
2003
;
13
:
8
–15.
Sign in via your Institution