Vous êtes sur la page 1sur 21

Molecular Biology of Inammation and Sepsis: A Primer (printe...

http://www.medscape.com/viewarticle/585997_print

www.medscape.com
From Critical Care Medicine

Molecular Biology of Inammation and Sepsis: A Primer


Ismail Cinel, MD, PhD; Steven M. Opal, MD Posted: 02/02/2009; Crit Care Med. 2009;37(1):291-304. 2009 Lippincott Williams & Wilkins

Abstract and Introduction


Abstract

Background: Remarkable progress has been made during the last decade in defining the molecular mechanisms that underlie septic shock. This rapidly expanding field is leading to new therapeutic opportunities in the management of severe sepsis. Aim: To provide the clinician with a timely summary of the molecular biology of sepsis and to better understand recent advances in sepsis research. Data Selection: Medline search of relevant publications in basic mechanisms of sepsis/severe sepsis/septic shock, and selected literature review of other manuscripts about the signalosome, inflammasome, apoptosis, or mechanisms of shock. Data Synthesis and Findings: The identification of the toll-like receptors and the associated concept of innate immunity based upon pathogen- or damage-associated molecular pattern molecules allowed significant advances in our understanding of the pathophysiology of sepsis. The essential elements of the inflammasome and signal transduction networks responsible for activation of the host response have now been characterized. Apoptosis, mitochondrial dysfunction, sepsis-related immunosuppression, late mediators of systemic inflammation, control mechanisms for coagulation, and reprogramming of immune response genes all have critical roles in the development of sepsis. Conclusions: Many of these basic discoveries have direct implications for the clinical management of sepsis. The translation of these bench-to-bedside findings into new therapeutic strategies is already underway. This brief review provides the clinician with a primer into the basic mechanisms responsible for the molecular biology of sepsis, severe sepsis, and septic shock.
Introduction

In sepsis, the expected and appropriate inflammatory response to an infectious process becomes amplified leading to organ dysfunction or risk for secondary infection. A continuum exists from a low grade systemic response associated with a self-limited infection to a marked systemic response with solitary or multiorgan dysfunction, i.e., severe sepsis. As a clinical syndrome, sepsis occurs when an infection is associated with the systemic inflammatory response. The complex toll-like receptor signaling and associated downstream regulators of immune cell functions play a crucial role in the [1] innate system as a first line of defense against pathogens. However, signaling is sometimes conflicting and a sustained inflammatory response can result in tissue damage. In addition, a reduction in their antimicrobial capacity may set the stage for [2] opportunistic and/or super infections. Severe sepsis may also be associated with an exaggerated procoagulant state. This may lead to ischemic cell injury, an effect that further amplifies the damage caused by inappropriate inflammation. A microvasculature injured by inflammation and ischemia, in turn, further deranges the host response by altering leukocyte trafficking, generating [3] apoptotic microparticles, and increasing cellular hypoxia. Mitochondrial dysfunction, an acquired intrinsic defect in cellular respiration termed cytopathic hypoxia, also has an important role by decreasing cellular oxygen consumption in this chaotic process. In this review, we highlight the current understanding of the basic molecular mechanisms that modulate these events so as to produce sepsis.

Pattern Recognition Receptors, Pathogen-Associated Molecular Patterns (PAMPs) and DangerAssociated Molecular Patterns (DAMPs)
The initiation of the host response during sepsis or tissue injury involves three families of pattern recognition receptors (PRRs): 1) toll-like receptors (TLRs); 2) nucleotide-oligomerization domain leucine-rich repeat (NOD-LRR) proteins; and 3) cytoplasmic caspase [4,5] activation and recruiting domain helicases such as retinoic-acid-inducible gene I (RIG-I)-like helicases (RLHs). These receptors initiate the innate immune response and regulate the adaptive immune response to infection or tissue injury. Gram-positive and Gram-negative bacteria, viruses, parasites, and fungi all possess a limited number of unique cellular constituents not found in vertebrate animals. These elements are now referred to as PAMPs, or more appropriately microbial-associated [6] molecular patterns, as these molecules are also common in nonpathogenic and commensal bacteria. PAMPs bind to PRRs, such

1 of 21

4/23/10 7:38 PM

Molecular Biology of Inammation and Sepsis: A Primer (printe...

http://www.medscape.com/viewarticle/585997_print
[7]

as TLRs, expressed on the surface of host cells. Cytoplasmic PRRs exist to detect invasive intracellular pathogens. The NOD proteins recognize common fragments of bacterial peptidoglycan. Diamino-pimelate from Gram-negative bacteria is the ligand for NOD1 and muramyl dipeptide from peptidoglycan is the ligand for NOD2 in the cytosol (see Figs. 1 and 2). The PRRs also recognized damage signals from the release of endogenous peptides and glycosaminoglycans from apoptotic or necrotic host [8-10] cells Caspase activation and recruiting domain helicases primarily recognize viral nucleic acids and activate antiviral measures including the type I interferons.

Figure 1. Specific host immune response to each pathogen is mediated by various sets of pathogen associated molecular patterns (PAMPs) and pattern recognition receptors (PRRs) as detailed in Figure 1. PRRs are essential for initiating the host's immune defenses against invading pathogens, yet they can also contribute to persistent and deleterious systemic inflammation. PRRs also serve as receptors for endogenous danger signals, hemodynamic changes in sepsis (tissue hypoperfusion and ischemia/reperfusion phenomenon), thus same signaling systems that alerts the highly advantageous, host defense mechanisms, also contributes to the disadvantageous, pathologic events of systemic inflammation, coagulation, tissue damage in target organs in sepsis. Heat shock proteins, fibrinogen, fibronectin, hyaluran, biglycans and high mobility group box-1 (HMGB-1) have been defined as danger associated molecular patterns (DAMPs) which are likely relevant for sepsis. Toll-like receptors (TLRs), especially TLR4, are involved in the recognition of these endogenous or harmful self-antigens ligands which are released during noninfectious injury, such as trauma or ischemia/reperfusion suggesting their function may not be restricted to the recognition of extrinsic pathogens. NOD-LRR, nucleotide-oligomerization domain leucine-rich repeat; ASC, apoptosisassociated speck-like protein containing caspase activation and recruiting domain; NF-, nuclear factor .

2 of 21

4/23/10 7:38 PM

Molecular Biology of Inammation and Sepsis: A Primer (printe...

http://www.medscape.com/viewarticle/585997_print

Figure 2. Binding of toll-like receptors (TLRs) activates intracellular signal-transduction pathways that lead to the activation of transcriptional activators such as interferon regulator factors, phosphoinositide 3-kinase (PI3K)/Akt, activator protein-1, and cytosolic nuclear factor-kappa (NF-). Activated NF- moves from the cytoplasm to the nucleus, binds to transcription sites and induces activation of an array of genes for acute phase proteins, inducible nitric oxide synthase, coagulation factors, proinflammatory cytokines, as well as enzymatic activation of cellular proteases. TLR9 DNA, TLR 3 dsRNA, and TLR7/8 ss RNA are endosomal. TLR 10 ligand is not defined and TLR1 forms heterodimers with TLR2. LPS, lipopolysaccharide; IRF, interferon regulatory factor; JNK; c Jun N-terminal kinase. TLR expression is significantly up-regulated in experimental models of sepsis and in patients with sepsis. Trauma including thermal injury generates danger-associated molecular patterns (i.e., high mobility group box-1, heat shock proteins, S100 proteins, hyaluran, etc.) that augment TLR expression like PAMPs. It also primes the innate immune system for enhanced TLR reactivity, [15] resulting in excess lipopolysaccharide (LPS)-induced mortality. Multiple positive feedback loops between danger-associated molecular patterns and PAMPs, and their overlapping receptors temporally and spatially drive these processes and may represent the molecular basis for the observation that infections, as well as nonspecific stress factors, can trigger flares in systemic inflammatory response. The degree to which TLR regulation mediates the ultimate outcome in sepsis in individual patients remains [13,16] elusive and is an active area of clinical investigation.
[11-14]

Inflammasome and Signalosome Pathways


TLRs induce pro-interleukin(IL)-1beta production and prime NLR-containing multiprotein complexes, termed inflammasomes, to [17,19] respond to bacterial products and products of damaged cells. This results in caspase-1 activation and the subsequent [19] processing of pro-IL-1 to its active extracellular form IL-1. Caspases are a set of cysteine proteases that alter the enzymatic activity of target proteins at specific peptide sequences adjacent to aspartate moieties. Caspases are important in process of apoptosis, cellular regulation, and inflammation (Figs. 1 and 3). One of the many targets of the caspase cascade is caspase activated DNase (CAD). CAD activation induces DNA fragmentation characteristic of programmed cell death (apoptosis).

3 of 21

4/23/10 7:38 PM

Molecular Biology of Inammation and Sepsis: A Primer (printe...

http://www.medscape.com/viewarticle/585997_print

Figure 3. Pathogenic mechanisms during sepsis or in response to tissue injury can lead to organ dysfunction. PRRs, pattern recognition receptors; TLRs, toll-like receptors; NOD-LRR, nucleotide-oligomerization domain leucine-rich repeat protein receptors; RLHs; retinoic-acid-inducible gene I (RIG-I)-like helicases; TNF-, tumor necrosis factor alpha; IL-1, interleukin 1; HMGB-1, high mobility group box-1; LPS, lipopolysaccharide; LTA, lipoteichoic acid; PGN, peptidoglycan; VSMCs, vascular smooth muscle cells; RBC, red blood cell; MPO, myeloperoxidase; XOR, xanthine oxidoreductase; iNOS, inducible nitric oxide synthase; COX-2, cyclooxygenase; RAGE, receptor for advanced glycation end products; ROCK; RhoA/Rho kinase; PARP-1, poly(ADP ribose) polymerase-1; PAR, protease activated receptor; ROS/RNS, reactive oxygen and nitrogen species; NF, nuclear factor; NADPH, nicotimanide adenosine dinucleotide phosphate; PAMPs, pathogen-associated molecular patterns; DAMPs, danger-associated molecular patterns. The posttranslational activation of caspase-1 is tightly regulated by inflammasome, the known components of which include caspase-1, ASC (apoptosis-associated speck-like protein containing a caspase activation and recruiting domain), NALP1 (NACHT, [20] leucine rich repeat and pyrin domain containing 1), and caspase-5. Additionally, alternate inflammasome constructions have been [21-23] suggested to contain pyrin, NALP3, and other members of the NOD-LRR family. ASC facilitates inflammasome assembly thus triggering caspase-1 activation and IL-1 processing. Interestingly, ASC can also regulate the nuclear factor (NF)-B pathway, thus

4 of 21

4/23/10 7:38 PM

Molecular Biology of Inammation and Sepsis: A Primer (printe...


[24] [25]

http://www.medscape.com/viewarticle/585997_print

linking the inflammasome to the signalosome (Fig. 1).

The inflammasome pathways contribute to the inflammatory response in sepsis. Caspase-1 knock out mice are protected from [26] sepsis while a naturally occurring polymorphism for human caspase-12, a putative regulator of caspase-1, has been linked to [27,28] [29] sepsis. Thus, caspase-1 activation appears to be a prerequisite for a competent immune response. Recently, the inflammasome components have been shown to be significantly lower in septic shock patients during the early stages of systemic [30] inflammatory response with elevated plasma cytokines levels. This monocyte deactivation process may be maladaptive in the later phases of sepsis and predispose to secondary infection. Caspase-1 and its proinflammatory cytokine products are likely to contribute to the pathogenesis of sepsis in overwhelming inflammation. However, like many other essential elements of innate immunity, caspase-1 also has a positive impact on host defense against several infections up-regulating microbial killing mechanisms such as the production of reactive oxygen and nitrogen species [31,32] [7] (ROS and RNS). Negative regulation of TLRs and TLR-induced programmed cell death has been defined. Apoptosis [33] induction is used by Pseudomonas aeruginosa to inhibit the secretion of immune and proinflammatory mediators by target cells. MyD88 protein and IL-1 receptor-associated kinase pathways activate NF-B during the innate immune response. NF-B is a transcription factor that is constituted by homo- or heterodimers of the Rel protein family with a pivotal role in inflammation, cell survival, and proliferation. In unstimulated cells, NF-B is maintained in a latent form in the cytoplasm by means of sequestration by inhibitory B (IB) proteins. NF-B activating stimuli, such as cytokines, viruses, and lipopolysaccharide (LPS), induce the degradation of inhibitory Bs by the proteasome, unmasking the nuclear localization signal of NF-B, resulting in its nuclear translocation, binding to NF-B motifs, and gene transcription. Dynamic redox control of NF- B through glutaredoxin-regulated [35] S-glutathionylation of inhibitory Bs has been demonstrated. NF-B, subject to regulation by redox changes, has been shown to be involved in the transcriptional regulation of more than 150 genes with a significant portion demonstrating proinflammatory [36] [37] properties. NF-B is readily activated upon intraperitoneally LPS challenge within 4 hrs in lung, liver, and spleen. A variant IL-1 receptor-associated kinase -1 haplotype has recently been demonstrated to affect the magnitude of NF-B activation and directly [38] correlates with an increased incidence of septic shock and significantly reduced survival rates. The impact of NF-B signaling is [39,40] tissue-specific as deficient NF-B activation in intestinal epithelium is associated with increased inflammation in vivo. Both studies demonstrate that defects of NF-B signaling cause immunosuppression which triggers and maintains inflammation. Thus, it can be suggested that inhibition of massive NF-B activation in vivo leads to reduced inflammatory responses, at least during certain phases and certain tissues (i.e., parenchymal tissues) in sepsis. On the other hand, this activation is followed by negative NF-B regulation favoring apoptosis in immune cells which may lead to immunosuppression and fatal outcome in severe sepsis. Recently, using specific gene-targeted deletions, it has been shown that deletion of MyD88 caused a worsened survival in a model of severe peritonitis, despite the marked decrease in sepsis-induced T and B lymphocyte apoptosis suggesting MyD88, like NF-B, is also [41] critical host survival in sepsis.
[34]

Toll-Like Receptor Signaling


The Role of Phosphoinositide 3-Kinase

Phosphoinositide 3-kinase (PI3K), a signal transduction enzyme, and the downstream serine/threonine kinase Akt (also known as [42] protein kinase B) have been reported in cellular activation, inflammatory responses, chemotaxis, and apoptosis (Figs. 2 and 3). PI3K can function either as a positive or negative regulator of TLR signaling. PI3Ks (three types) act at several steps downstream of TLRs, depending on the cell type and/or the engagement of a specific TLR regulating downstream signaling to NF-KB transactivation [43,44] or to mitogen-activated phosphokinase activation. As a positive mediator of TLR signaling, PI3K together with p38 and extracellular regulated kinase (ERK)1/2 mitogen-activated phosphokinases, lead to production of proinflammatory cytokines IL-1, [44,45] IL-6, and IL-8 upon microbial challenge. In detail, PI3K activation appeared to have a significantly promoting function for these [46] mediators in monocytes, whereas activation appeared to limit the LPS response for generation of these cytokines in neutrophils. However, PI3K inhibition resulted in impaired oxidative burst and phagocytosis activity in both neutrophils and monocytes. By limiting C5a-mediated effects on neutrophil cytokine generation, and promoting oxidative burst and phagocytosis, PI3K activation seems to be a therapeutic approach for limiting inflammation in sepsis. On the other hand, PI3K/Akt signaling pathway acts as an endogenous negative feedback mechanism that serves to limit [47] proinflammatory and apoptotic events as seen in monocytes in response to endotoxin. The suppressive role in inflammation and [48] coagulation was confirmed in mouse model of endotoxemia. It can also promote the generation of anti-inflammatory cytokine [49] [50] IL-10. PI3K/Akt has an effect in balancing Th1 vs. Th2 responses. Overexpression of Akt in lymphocytes decreases [51] lymphocyte apoptosis, a Th1 cytokine propensity, and improve outcome in cecal ligation and puncture-induced sepsis. Peptidemediated activation of Akt and extracellular regulated kinase signaling protects lymphocytes from numerous apoptotic stimuli both in

5 of 21

4/23/10 7:38 PM

Molecular Biology of Inammation and Sepsis: A Primer (printe...


[52]

http://www.medscape.com/viewarticle/585997_print

vitro and in vivo.

This suggests again the survival advantage of PI3K/Akt pathway activation for the later stages of sepsis.

The Role of Rho GTPases

The Rho family of small GTPases is one of the master regulators of cell motility, as they control actin cytoskeleton remodeling. RhoA, Rac1, and CDc42 are the well known family members which act molecular switches regulating responses of innate immune cells [53] related to pathogen sensing, intracellular uptake, and destruction. These GTPases play both unique and overlapping roles in [54] phagocyte functions including migration, chemotaxis, and optimal bacterial killing. It has been shown that PAMPs may utilize [55] GTPases through TLRs (i.e., TLR-2, -4, -3, and -9) in which Rac1 activation is required for PI3K activation upon TLR2 stimulation. RhoA regulates not only cytoskeletal events, which mediate neutrophil migration, but also contributes to NF--dependent [56] proinflammatory gene transcription (Fig. 3). It has been suggested that ROCK inhibition could attenuate cytoskeletal rearrangement of endothelial cells, leading to decreased neutrophil emigration into the lung parenchyma in LPS-induced lung [57] [58] injury. An important role for rho kinase in leukocyte recruitment is also supported in an endotoxemic liver injury model. The recently emerged connections between TLR signaling and small Rho-GTPases seem to provide new therapeutic avenue of research in sepsis.

Toll-Like Receptor Signaling and Regulatory T Cells (Tregs)


To balance self-tolerance and immunity against pathogens, the immune system depends on both up-regulatory and down-regulatory mechanisms. Recent studies have suggested that several lymphocyte subpopulations (i.e., CD4+CD25+ Foxp3+ T regulatory-cell [Tregs]) may have the capacity to actively suppress an adaptive immune response and may potentially be involved in septic immune dysfunction. Naturally occurring Tregs express the transcription factor forkhead box protein (FoxP3), which is induced by the [59,60] anti-inflammatory cytokine transforming growth factor-. TLR triggering induces dendritic cell maturation also, which is [61] essential for the induction of adaptive immune responses. Studies have recently highlightened the importance of TLRs on [62] Tregs. In this regard, the dominant role of TLR2 signaling on the Treg-mediated immune suppression has been demonstrated.
[63,64]

Tregs control inflammatory reactions to commensal bacteria and opportunist pathogens and play a major role in suppressing immune [65] [66] [67] reactivity, ranging from autoimmunity to infectious disease and to injury. Monneret et al observed that sepsis increases CD4+CD25+ T cells in the peripheral blood of septic patients. This was subsequently found to be a relative increase in Tregs due to [68] a decrease in the CD4+CD25- T effector cell populations. Furthermore, Treg-mediated induction of alternatively activated macrophages has been demonstrated suggesting as one of the causes of immune dysfunction in systemic inflammatory response [69,70] syndrome and sepsis. Targeting apoptosis of Tregs may be a new therapeutic approach in preventing the continuum of sepsis to severe sepsis. However, it has been reported that depletion of CD25+ cells before inducing sepsis did not alter septic mortality [71,72] pointing the need of more studies to clarify the significance of this cell population's expansion in sepsis morbidity. Adoptive transfer of Tregs before or following the initiation of polymicrobial sepsis improved survival by enhancing tumor necrosis [73] factor (TNF)- production and bacterial clearance. Tregs inhibit LPS-induced monocyte survival through the Fas/FasL dependent [74] proapoptotic mechanism, which might play a role in the resolution of exaggerated inflammation. The positive and negative effects of TLR on Tregs are curious in sepsis, and the dynamics of TLR expression on immune-suppressive Tregs upon inflammation or in relation to type of pathogen are needed to illuminate.

Neutrophils and Monocyte/Macrophages in Inflammation


Myeloid cells including neutrophils and elements of monocyte/macrophage lineage are heavily armed with large stores of proteolytic enzymes and with the capacity to rapidly generate ROS and RNS to degrade internalized pathogens. The highly proapoptotic nature of neutrophils is designed to maintain a balance between antimicrobial effectiveness and the potential for neutrophil-associated [75] damage to the host in septic challenge or in other injurious processes, such as trauma or ischemia/reperfusion. Host tissue damage in severe sepsis may arise via a variety of mechanisms including premature neutrophil activation during migration, extracellular release of cytotoxic molecules and toxins during microbial killing, removal of infected or damaged host cells or debris [76] during host tissue remodeling, and failure to terminate acute inflammatory responses. Therefore, to maximize host defense capabilities while minimizing damage to host tissues, neutrophil microbial responses are tightly regulated. On the other hand, quorum sensing (the ability of bacteria to assess their population density) has been found to have a crucial role in regulating tissue invasion [1] by bacterial pathogens, and inhibitors of quorum sensing system provide new avenues for intervention against invasive pathogens. Evidence now exists that quorum sensing system can even open up bidirectional lines of communication between bacteria and the human host. Leukocyte-endothelial interactions, which may also contribute to inflammation-mediated injury, involve two sets of adhesion

6 of 21

4/23/10 7:38 PM

Molecular Biology of Inammation and Sepsis: A Primer (printe...


[77,78]

http://www.medscape.com/viewarticle/585997_print

molecules, selectins, and integrins. P-selectin is expressed on platelets; E- and P-selectins are expressed by endothelial cells, whereas L-selectin is expressed on leukocytes. Selectins mediate neutrophil rolling along activated endothelial surfaces as circulating neutrophils decelerate to engage endothelial receptors. The beta-2 intergrins (CD11/CD18 complexes) mediate tight [78] adhesion to endothelial membranes, allowing subsequent egress of neutrophils to extravascular sites of inflammation. Activated neutrophils stimulate transendothelial albumin transport through intracellular adhesion molecule-1 mediated, Scr-dependent caveolin [79-82] phosphorylation. The role of caveolin in inflammatory response in sepsis has been recently defined (Fig. 3). Neutrophils contribute to blood coagulation in localized inflammation and in generalized sepsis. During systemic inflammation, homeostatic mechanisms are compromised in the microcirculation including endothelial hyperactivity, fibrin deposition, microvascular occlusion, and cellular exudates that further impede adequate tissue oxygenation. Neutrophils participate in these rheologic changes [83] through their augmented binding to blood vessel walls and through the formation of platelet-leukocyte aggregates. Neutrophil elastase, other proteases, glycases and inflammatory cytokines degrade endogenous anticoagulant activity, and impair fibrinolysis on [1] endothelial surfaces favoring a procoagulant state.
The Role of Apoptosis
[1,3,75]

Sepsis-induced neutrophil-mediated tissue injury has been demonstrated in a variety of organs including the lungs, [87] [86] [88,89] [86] diaphragm, kidneys, intestine, and liver. Apoptosis is a counter-regulator of the initial inflammatory response in [90] sepsis. Neutrophils are constitutively proapoptotic and apoptosis is fundamental for the resolution of inflammation and cell turnover. Neutrophils can undergo apoptosis via intrinsic and extrinsic pathways; the latter also requires mitochondrial amplification [91] (Fig. 4). The role played by mitochondria in the regulation of neutrophil life span is more crucial than in other cell types in the [92] body. As neutrophils kill pathogens using ROS and RNS and a mixture of lytic enzymes, delayed clearance of neutrophils in sepsis can potentially contribute to cell/organ injury. Importantly, the phagocytosis of bacteria and fungi accelerates neutrophil apoptosis. Apoptotic cell clearance induces anti-inflammatory effects in tissues. It has been shown that intratracheal administration of killed E. coli attenuated lung injury and improved survival in an intestinal ischemia/reperfusion injury associated with marked [93] pulmonary neutrophil infiltration.

[84-86]

Figure 4. Summary of apoptotic signaling pathways as seen through activation of death receptor (extrinsic) or mitochondrial

7 of 21

4/23/10 7:38 PM

Molecular Biology of Inammation and Sepsis: A Primer (printe...

http://www.medscape.com/viewarticle/585997_print

(intrinsic) pathway. Extrinsic signals bind to their receptors and trigger intracellular signaling, leading to caspase-8 activation. Activation of caspase-8 by extrinsic stimuli (such as tumor necrosis factor- [TNF-], Fas ligand) involves mitochondriadependent signaling in type II cells. In type I cells, on the other hand, execution of apoptosis occurs without significant participation of mitochondria. MAPK, mitogen-activated phosphokinase; PI3K/Akt, phosphoinositide 3-kinase/Akt; APAF-1, apoptosis protease activating factor 1; ER, endoplasmic reticulum; IL, interleukin. Cytokine-induced prolonged neutrophil survival is accompanied by evidence of increased neutrophil activation, including augmented [90] respiratory burst activity. Neutrophils from patients with sepsis manifest markedly prolonged survival in vitro in association with [94] evidence of cellular activation. Phagocytosis of apoptotic neutrophils by macrophages inhibits the release of proinflammatory [95] cytokines and promotes the secretion of anti-inflammatory cytokines. In contrast, inefficient apoptotic cell clearance is [96] proinflammatory and immunogenic. The recognition of apoptotic cells by macrophages is largely dependent on the cell surface appearance of an anionic phospholipid, phosphatidylserine (PS), which is normally confined to the inner leaflet of the plasma [97] membrane. Asymmetric distribution of PS across the plasma membrane is mainly because of the activity of a specialized enzymatic mechanism, aminophospholipid translocase. S-nitrosylation of critical cysteine residues inhibits aminophospholipid translocase, leading to PS externalization. PS expression during apoptosis generates an eat-me signal (Fig. 5), which in turn triggers [98] clearance of apoptotic cells and suppresses the inflammatory response. It has been demonstrated that S-nitrosylation of critical cysteine residues in aminophospholipid translocase using a cell-permeable transnitrosylating agent, S-nitroso-acetyl- cysteine, [99] resulted in egression of PS to the outer surface of the plasma membrane, rendering these cells recognizable by macrophages. The therapeutic potential of regulating neutrophil life-span in sepsis remains to be determined. Care must be exercised in regulating of this pathway, as sepsis enhances the capacity of macrophages to clear expanded apoptotic populations, a mechanism [100] contributing to septic immune suppression.

Figure 5. The recognition of apoptotic cells by macrophages is largely dependent on the cell surface appearance of phosphatidylserine (PS). S-nitrosylation of critical cysteine residues inhibits aminophospholipid translocase (APLT), leading to PS externalization. It generates an eat me signal during apoptosis. iNOS, inducible nitric oxide synthase; NO, nitric oxide; O , superoxide; ONOO , peroxynitrite.
The Role of ROS and RNS
2

ROS and RNS exert several beneficial physiologic functions, such as intracellular signaling for several cytokines and growth factors, second messengers for hormones and redox regulation. Despite their importance as a defense mechanism against invading pathogens, an overwhelming production of ROS and RNS or a deficit in oxidant scavenger and antioxidant defenses result in [101,102] oxidative/nitrosative stress, a key element in the deleterious processes in sepsis (Fig. 6).

8 of 21

4/23/10 7:38 PM

Molecular Biology of Inammation and Sepsis: A Primer (printe...

http://www.medscape.com/viewarticle/585997_print

Figure 6. Proinflammatory/anti-inflammatory cytokines and reactive oxygen and nitrogen species have important effects within the microcirculatory unit: the arteriole, endothelial cell, capillary bed, and the venule. The arteriole is where the characteristic intractable vasodilation of sepsis occurs. The capillary bed is where the effects of endothelial cell activation/dysfunction are most pronounced and microvascular thromboses are formed. The postcapillary venule is where leukocyte trafficking is most disordered. All of these causes impair flow through the microcirculation leading to microcirculatory dysfunction. TLR, toll-like receptor; NOD-LRR, nucleotide-oligomerization domain leucine-rich repeat; RLH, (RIG-I)-like helicase; VSM, vascular smooth muscle; IL, interleukin; TGF, transforming growth factor; MIF, migration inhibitory factor; TNF, tumor necrosis factor; HMGB, high mobility group box; PAF, protease activating factor; sTNFR, soluble tumor necrosis factor receptor. Stimulated neutrophils produce ROS and RNS through the nicotinamide adenine dinucleotide phosphate oxidase complex, [103] myeloperoxidase and xanthine oxidoreductase and represent a defense mechanism against invading microorganisms. Lipopolysaccharide and other proinflammatory mediators activate nicotinamide adenine dinucleotide phosphate oxidase to produce superoxide radical (O ). In aqueous environments, superoxide radical is rapidly catalyzed by superoxide dismutase hydrogen
2

peroxide (H O ) and hydroxyl radicals. Myeloperoxidase from neutrophil azurophilic granules produces hypochlorous acid from
2 2

hydrogen peroxide (H O ) and chloride anion (Cl ) during respiratory burst. These radicals are highly cytotoxic, and neutrophils used
2 2

them to kill bacteria and other pathogens. Expression of human xanthine oxidoreductase is markedly up-regulated by hypoxia, ischemia/reperfusion, LPS, and TNF-. Increased activity of xanthine oxide, one the important contributors of ROS production, has [104] been reported in adult and pediatric patients with sepsis. O
2

in the presence of nitric oxide, generates peroxynitrite (ONOO ), a key player in the pathogenesis of sepsis-induced organ
-

dysfunction. ONOO can cause DNA strand breakage, which triggers the activation of DNA repair enzymes such as poly (adenine dinucleotide phosphate-ribose) polymerase (Fig. 3). Poly (adenine dinucleotide phosphate-ribose) polymerase inhibitors protect [87-89] against oxidative and nitrosative stress-induced organ dysfunction in endotoxemia. Recently, the potential role of poly (adenine dinucleotide phosphate-ribose) polymerase activation has been implicated in the pathogenesis of myocardial contractile dysfunction [105] associated with human septic shock.

Novel Cytokines in Inflammation


High Mobility Group Box-1 and Receptor for Advanced Glycation End-Products

High mobility group box-1 (HMGB-1) is a nonhistone, nuclear DNA-binding protein involved in nucleosome stabilization and gene transcription. However, when HMGB-1 is released in large quantities into the extracellular environment, it becomes a lethal mediator

9 of 21

4/23/10 7:38 PM

Molecular Biology of Inammation and Sepsis: A Primer (printe...


[106]

http://www.medscape.com/viewarticle/585997_print

of systemic inflammation (Fig. 3). Indeed, it has been recently shown that it has a weak proinflammatory activity by itself and [107] binding to bacterial substances including lipid molecules such as phophatidylserine strengthens its effects. Interestingly, phosphatidylserine has been implicated in the regulation of inflammation and HMGB-1 might thus regulate its anti-inflammatory [108] activities. HMGB-1 is released into the extracellular space through acetylation or phosphorylation. HMGB-1 is either passively released from necrotic cells, and a mechanism that represents a process adopted by the innate immune system to recognize damaged and [106] necrotic cells, or actively secreted by immune cells including macrophages and neutrophils to trigger inflammation. Recently, HMGB-1 release from macrophages has been shown during the course of apoptosis as well as necrosis and defined as a [111,112] downstream event of cell apoptosis during severe sepsis. After treatment with LPS or various cytokines such as TNF-, IL-1, or IFN-, HMGB-1 is released from activated macrophages within 4 hrs and reaches a plateau around 18-24 hrs. It binds to several transmembrane receptors such as receptor for advanced glycation end products (RAGE), TLR-2, and -4, activating NF-B [113,114] and extracellular regulated kinase 1/2. Although HMGB-1 was originally described as a late mediator of endotoxin-induced lethality, recent studies indicate a role for [116] HMGB-1 in angiogenesis, tissue repair, and regeneration. HMGB-1 appears to be a novel myocardial depressant factor upon release by resident myocardial cells following tissue injury. HMGB-1 might decrease energy utilization in ischemic tissue, thereby [117] preventing injured myocytes from worsening ATP depletion that eventuate in necrosis. On other hand, excessive release of HMGB-1 in sepsis might contribute to sustained inflammation and to profound myocardial depression. The cholinergic anti-inflammatory pathway is a neural mechanism that inhibits the expression of HMGB-1 and other cytokines. [118-121] Signals transmitted via the vagus nerve, the principal nerve of the parasympathetic nervous system, significantly attenuate [122] the release of HMGB-1 and other cytokines in inflammation in animal and human studies. The remarkable binding characteristics of HMGB-1 suggest another important role for this protein in the extracellular fluid. HMGB-1 [108,123] might serve as a shuttle platform for LPS and other microbial mediators for docking to CD14 and recognition by the TLRs or [124] through binding to host-derived proinflammatory mediators, such as IL-1beta. Elevated levels of HMGB-1 are measurable in the majority of patients up to 1 wk after the diagnosis of sepsis or septic shock and are [125,126] correlated with the degree of organ dysfunction. However, serum HMGB-1 levels do not consistently identify nonsurvivors [127] from survivors as a predictor of hospital mortality. As HMGB-1 is late inflammatory cytokine of sepsis, it provides a wide therapeutic time window for clinical intervention and remains an attractive target for sepsis treatment. RAGE, a member of the immunoglobulin superfamily, is a pattern-recognition receptor that binds diverse classes of endogenous molecules including HMGB-1. It has been defined as part of a newly appreciated component of the innate immune system referred to [128] as the danger associated molecular pattern system (Fig. 3). Membrane bound and soluble forms of RAGE (sRAGE) have been detected in plasma. Soluble RAGE, anti-RAGE antibody (Fab 2 fragment) or data from RAGE-/- animals have been shown to [129] decrease inflammation, reduce neutrophil extravasation, and reduce migration. Recently, the demonstration of survival benefit after delayed administration of anti-RAGE antibody in a murine model of polymicrobial sepsis, even when delayed up to 24 hrs, [130] provides a therapeutic rationale for the use of anti-RAGE mAb as a salvage therapy for established severe sepsis.
Macrophage Migration Inhibitory Factor
[115] [109,110]

Migration inhibitory factor (MIF) acts as a stress response mediator and proinflammatory cytokine upon induction by [131] glucocorticoids. This protein is readily measurable in patients with sepsis, and MIF probably contributes to the pathogenesis of sepsis. Inhibition of MIF or its targeted deletion attenuates TNF- and IL-1 expression and protects mice from lethality is [132,133] [133] experimental sepsis. Systemic challenge of animals with MIF increases LPS-related lethality. MIF promotes the [134] expression of TLR4 on macrophages, and thereby sensitizes these immune effector cells to LPS. The immunoregulatory effects of MIF might be crucial to the control and resolution of the inflammatory response as a consequence [135] of its ability to regulate activation-induced apoptosis. High MIF levels delay the removal of activated monocytes/macrophages by apoptosis. This prolongs monocyte/macrophage survival, increases cytokine production, and sustains an ongoing proinflammatory response.

Mitochondrial Dysfunction in Inflammation


Although microvascular flow abnormalities occur, findings of decreased oxygen consumption
[136]

and elevated tissue oxygen

10 of 21

4/23/10 7:38 PM

Molecular Biology of Inammation and Sepsis: A Primer (printe...


[137]

http://www.medscape.com/viewarticle/585997_print
[138]

tension, yet minimal cell death despite functional and biochemical derangements, suggest that the problem lies more in [139] cellular oxygen utilization rather than a problem with oxygen delivery. It is postulated that prolonged and systemic inflammatory insult is accompanied by a basic tissue survival response mediated by switching off its energy-consuming biophysiological [140,141] processes. Recent evidence suggests that sepsis and septic shock severely impair the mitochondria, and the severity and [142] outcome of organ dysfunction could be related to mitochondrial dysfunction. Depleted levels of reduced glutathione, an important intramitochondrial antioxidant, in combination with excess generation of ROS and RNS severely inhibit oxidative phosphorylation and [142] ATP generation. This acquired intrinsic derangement in cellular energy metabolism which has also been termed cytopathic hypoxia, contributes to reduced activities of mitochondrial electron transport chain enzyme complexes and impaired ATP [141,143,144] biosynthesis, potentially organ dysfunction in sepsis. Sepsis-related derangements in mitochondrial function can activate the ubiquitin proteolytic pathway in skeletal muscle of septic [145] patients. Mitochondrial permeability transition pore seems to be involved in sepsis-induced mitochondrial damage, since its [146] inhibition significantly improved organ function and reduced mortality in rodents. Whether sepsis-related reduction in energy supply could result in a state of cellular shutdown analogous to myocardial stunning (or hibernation) following coronary occlusion, allowing for eventual restoration of organ function and survival, has not yet been determined.

Impact of Inflammation on Coagulation


The clotting system is almost invariably activated by systemic microbial invasion. Clotting is one most prominent features of sepsis. Coagulation contributes significantly to the outcome in sepsis with concurrent down-regulation of anticoagulant systems and [148] fibrinolysis (Fig. 7). Inflammation-induced coagulation in turn contributes to further inflammation. Indeed, collaboration between clotting and inflammation accounts for the basic survival strategy of walling off the damaged and infected tissues from the rest of the [149] host. The key determinant of survival in sepsis is to limit excess systemic inflammatory and coagulopathic damage while [147,150] retaining the benefits of controlled antimicrobial clearance and localized clot formation.
[147]

Figure 7. The extrinsic pathway (tissue factor pathway) is the primary mechanism by which thrombin is generated in sepsis. The intrinsic cascade (contact factor pathway) primarily serves an accessory role in amplifying the prothrombotic events that are initiated in sepsis. Thrombin, factor Xa and the TF-factor VIIa complex interact with the PAR (protease activated receptors) system and directly activate endothelial cells, platelets and white blood cells, and induce a proinflammatory response. Plateletderived microparticles (MPs) express functional adhesion receptors including P-selectin on their surface, attach to the site of injury on the vessel wall, and support the rolling of leukocytes in the presence of shear stress or severe insult. When bearing appropriate counter-ligands, MPs can transfer their procoagulant potential to target cells. Platelet-derived MPs can bind to

11 of 21

4/23/10 7:38 PM

Molecular Biology of Inammation and Sepsis: A Primer (printe...

http://www.medscape.com/viewarticle/585997_print

soluble and immobilized fibrinogen, thus delivering procoagulant entities to the thrombus via the formation of aggregates. In vitro, interaction between endothelial MPs and monocytes promotes TF mRNA expression and TF-dependent procoagulant activity. Activated platelets and platelet-derived MPs thus amplify leukocyte-mediated tissue injury in thrombotic and inflammatory disorders. EPCR, endothelial protein C receptor; PMN, polymorphonuclear leukocyte; APC, activated protein C; ADAMST13, a disintegrin and metalloproteinase with a thrombospondin type 1 motifs 13; ULVWF, unusually large von Willebrand factor; ICAM, intracellular adhesion molecule; VCAM, vascular cell adhesion molecule-1; IL, interleukin; TNF, tumor necrosis factor; IFN, interferon. The inflammatory reaction to tissue injury activates the clotting system, and coagulation promotes inflammation. The role of [153-155] procoagulant apoptotic microparticles have also been demonstrated in sepsis. Linkage of coagulation enzymes with their serine protease activity with protease activated receptors (PAR 1-4) on endothelial surfaces increases P-selectin, cytokine [156,157] production, and adhesion molecule expression, leading to microcirculatory dysfunction in severe sepsis.
Activated Protein C.
[3,151,152]

Activated protein C (APC) is derived from its zymogen protein C in contact with thrombin: thrombomodulin complexes on endothelial [158] surfaces. Protein C was originally thought to be synthesized exclusively by the liver. It has recently shown that it is strongly [159] expressed by the endothelium and keratinocytes. The conversion to APC is augmented by endothelial PC receptor (EPCR) [160,161] which is present on endothelial cells, neutrophils, monocytes, and keratinocytes whereas soluble EPCR inhibits APC [162] [163] anticoagulant activity. Soluble EPCR is released constitutively and levels increase in patients with Gram-negative sepsis. [164,165] Levels of APC, protein C, and its cofactor protein S are depleted in sepsis. Furthermore, peripheral conversion of Protein C by the thrombin: thrombomodulin complex is impaired in sepsis, further contributing to microvascular thrombosis and vascular [166] leakage. APC is a central endogenous anticoagulant protein with antithrombotic, anti-inflammatory, antiapoptotic, and profibrinolytic activities [167,168] (Fig. 8). Although an anti-inflammatory role for APC may be an indirect consequence of its ability to reduce thrombin [169] generation, APC also has direct anti-inflammatory properties. APC can cleave and activate PAR1-dependent cellular [170] 3 4 pathways. APC competes for PAR-1 binding with thrombin, but in vitro studies suggest that APC is 10 - to 10 -fold less potent that thrombin in cleaving PAR-1. The critical receptors required for both PC activation and APC cellular signaling (i.e., [171] thrombomodulin, EPCR and PAR-1) are co-localized in lipid rafts on endothelial cells. EPCR is associated with caveolin-1 on lipid rafts and EPCR binding to the gamma-carboxyglutamic acid domain of protein C/APC leads to its dissociation from caveolin-1 (Fig. 7). APC then engages PAR-1 generating a protective signaling pathway through coupling of PAR-1 to the pertussis toxinsensitive G(i)-protein. Thus, when EPCR is bound by protein C, the PAR-1-dependent protective signaling responses in endothelial cells can be mediated by either thrombin or APC. These results explain how PAR-1 and EPCR participate in protective signaling [172] events in endothelial cells.

12 of 21

4/23/10 7:38 PM

Molecular Biology of Inammation and Sepsis: A Primer (printe...

http://www.medscape.com/viewarticle/585997_print

Figure 8. The homeostatic balance between thrombin and activated protein C (APC) in coagulation and inflammation is given. APC is a central endogenous anticoagulant protein with antithrombotic, anti-inflammatory, anti-apoptotic and pro-fibrinolytic activities. The ligand occupancy of endothelial protein C receptor (EPCR) switches the protease-activated receptor 1 (PAR-1)-dependent signaling specificity of thrombin from a permeability-enhancing to a barrier-protective response. Other pleiotropic effects are also explained. TAFI, thrombin activatable fibrinolysis inhibitor; NF-, nuclear factor . Therapeutic administration of recombinant human APC (rhAPC) is currently in use as a treatment strategy for severe sepsis patients [173,174] with a high risk of death. Genetically engineered variants of APC have been designed with greater antiapoptotic activity and reduced anticoagulant activity relative to wild-type APC to increase the risk/benefit ratio of rhAPC regarding the bleeding [175] [176] complication. A nonanticoagulant from of APC reduces mortality in experimental models of endotoxemia and sepsis.
von Willebrand Factor, A Disintegrin-Like and Metalloproteinase with Thrombospondin Type-1 Motifs 13 and Ashwell Receptor

The discovery of a disintegrin-like and metalloproteinase with thrombospondin type-1 motifs 13 (ADAMTS-13) has provided new insights in pathogenesis of thrombosis in sepsis. ADAMTS-13, the principal physiologic modulator of von Willebrand factor (VWF) is [177] produced mainly stellate cells in the liver. VWF is synthesized in vascular endothelial cells and released into the plasma as unusually large VWF multimers which are rapidly degraded into smaller VWF multimers by ADAMTS-13. Deficiency of the ADAMTS-13, as observed in most forms of thrombotic thrombocytopenic purpura, increases the level of unusually large VWF multimers in plasma and leads to platelet aggregation and/or thrombus formation, especially in small arterioles, resulting in [178-180] microvascular failure. Recently, inflammation associated-ADAMTS-13 deficiency has been described in patients with [181-185] systemic inflammatory response syndrome and severe sepsis (Fig. 7). Decreased levels of ADAMTS-13 have been reported [186] in healthy volunteers following endotoxin infusion. Furthermore, reduced ADAMTS-13 levels are associated with differences in [182,187] morbidity, mortality, and variables of inflammation and endothelial dysregulation in severe sepsis patients. The Ashwell receptor, which is the major lectin of hepatocytes, modulates VWF homeostasis. It has been recently demonstrated that the marked thrombocytopenia associated with S. pneumoniae sepsis is the result of Ashwell receptor-dependent clearance of [188] platelets. The ensuing reduction in platelet counts at the onset of sepsis protects the host against the development of disseminated intravascular coagulation. Homeostatic adaptation by this receptor moderates the onset and severity of disseminated [189] intravascular coagulation during sepsis suggesting the improvement in host survival probability. At present, it remains unclear whether blocking the Ashwell receptor may have a beneficial effect in severe sepsis.

Conclusion

13 of 21

4/23/10 7:38 PM

Molecular Biology of Inammation and Sepsis: A Primer (printe...

http://www.medscape.com/viewarticle/585997_print

A unifying concept of innate immunity is based upon pathogen-, or damage-associated molecular pattern molecules and downstream signaling pathways. This has facilitated significant advances in our understanding of the pathophysiology of sepsis and led to a multitude of clinical trials ( Table 1 ). However, further identification of the critical elements in severe sepsis that drives the transition from localized inflammation to deleterious host response remains incompletely illuminated. The limitations in our current knowledge [190] of the molecular mechanisms in sepsis have made the design of intervention trials in clinical sepsis challenging. New findings as to biomarkers and measures to detect genetic signatures of sepsis are now making their way into clinical trial designs. It is anticipated that such innovations will improve the outlook for successful development of new sepsis treatments tailored to individual patient needs.
Table 1. Ongoing and Upcoming Clinical Studies with Molecular Targets

References

1. van der Poll T, Opal SM: Host-pathogen interactions in sepsis. Lancet Infect Dis 2008; 8:32-43 2. Cinel I, Dellinger RP: Making a difference in outcome in severe sepsis: The organism, the host response and the treating clinician. Emerg Med Crit Care Rev 2007; 1:1-3 3. Cinel I, Dellinger RP: Advances in pathogenesis and management of sepsis. Curr Opin Infect Dis 2007; 20:345-352 4. Uematsu S, Akira S: Toll-like receptors and innate immunity. J Mol Med 2007; 84:712-725 5. Creagh EM, O'Neill LA: TLRs, NLRs and RLRs: A trinity of pathogen sensors that co-operate in innate immunity. Trends Immunol 2006; 27:352-357 6. Granucci F, Foti M, Ricciardi-Castagnoli P: Dendritic cell biology. Adv Immunol 2005; 88:193-233 7. Liew FY, Xu D, Brint EK, et al: Negative regulation of toll-like receptor-mediated immune responses. Nat Rev Immunol 2005; 5:446-458 8. Akira S, Uematsu S, Takeuchi O: Pathogen recognition and innate immunity. Cell 2006; 124:783-801 9. Bianchi ME: DAMPs PAM. Ps and alarmins: All we need to know about danger. J Leukoc Biol 2007; 81:1-5 10. Mollen KP, Anand RJ, Tsung A, et al: Emerging paradigm: Toll-like receptor 4-sentinel for the detection of tissue damage. Shock 2006; 26:430-437 11. Tsujimoto H, Ono S, Hiraki S, et al: Hemoperfusion with polymyxin B-immobilized fibers reduced the number of CD16+ CD14+ monocytes in patients with septic shock. J Endotoxin Res 2004; 10:229-237 12. Armstrong L, Medford AR, Hunter KJ, et al: Differential expression of toll-like receptor (TLR)-2 and TLR-4 on monocytes in human sepsis. Clin Exp Immunol 2004; 136:312-319 13. Tsujimoto H, Ono S, Majima T, et al: Neutrophil elastase, MIP-2, and TLR-4 expression during human and experimental

14 of 21

4/23/10 7:38 PM

Molecular Biology of Inammation and Sepsis: A Primer (printe...

http://www.medscape.com/viewarticle/585997_print

sepsis. Shock 2005; 23:39-44 14. Williams DL, Ha T, Li C, et al: Modulation of tissue toll-like receptor 2 and 4 during the early phases of polymicrobial sepsis correlates with mortality. Crit Care Med 2003; 31:1808-1818 15. Paterson HM, Murphy TJ, Purcell EJ, et al: Injury primes the innate immune system for enhanced toll-like receptor reactivity. J Immunol 2003; 171:1473-1483 16. Tsujimoto H, Ono S, Efron PA, et al: Role of toll like receptors in the development of sepsis. Shock 2008; 29:315-321 17. Ogura Y, Sutterwala FS, Flavell RA: The inflammasome: First line of the immune response to cell stress. Cell 2006; 126:659-662 18. Drenth JP, van der Meer JW: The inflammasome-A linebacker of innate defense. N Engl J Med 2006; 355:730-732 19. Dinarello CA: Interleukin-1, interleukin-18, and the interleukin-1 converting enzyme. Ann NY Acad Sci 1998; 856:1-11 20. Martinon F, Burns K, Tschopp J: The inflammasome. A molecular platform triggering activation of inflammatory caspases and processing of proIL-1. MolCell 2002; 10:417-426 21. Yu JW, Wu J, Zhang Z, et al: Cryopyrin and pyrin activate caspase-1, but not NF-kappaB, via ASC oligomerization. Cell Death Differ 2005; 13:236-249 22. Seshadri S, Duncan MD, Hart JM, et al: Pyrin levels in human monocytes and monocyte-derived macrophages regulate IL-1 processing and release. J Immunol 2007; 179:1274-1281 23. Tschopp J, Martinon F, Burns K: NALPs: A novel protein family involved in inflammation. NatRevMol Cell Biol 2003; 4:95-104 24. Sarkar A, Duncan M, Hart J, et al: ASC directs NF-B activation by regulating receptor interacting protein-2 (RIP2) caspase-1 interactions. J Immunol 2006; 176:4979-4986 25. Scott AM, Saleh M: The inflammatory caspases: Guardians against infections and sepsis. Cell Death Differ 2007; 14:23-31 26. Sarkar A, Hall MW, Exline M, et al: Caspase-1 regulates E. coli sepsis and splenic B cell apoptosis independently of IL-1 and IL-18. Am J Respir Crit Care Med 2006;174:1003-1010 27. Saleh M, Vaillancourt JP, Graham RK, et al: Differential modulation of endotoxin responsiveness by human caspase-12 polymorphisms. Nature 2004; 429:75-79 28. Saleh M, Mathison JC, Wolinski MK, et al: Enhanced bacterial clearance and sepsis resistance in caspase-12-deficient mice. Nature 2006; 440:1064-1068 29. Joshi VD, Kalvakolanu DV, Hasday JD, et al: IL-18 levels and the outcome of innate immune response to lipopolysaccharide: Importance of a positive feedback loop with caspase-1 in IL-18 expression. J Immunol 2002; 169:2536-2544 30. Fahy RJ, Exline MC, Gavrilin MA, et al: Inflammasome mRNA expression in human monocytes during early septic shock. Am J Resp Crit Care Med 2008; 177:983-988. 31. Lara-Tejero M, Sutterwala FS, Ogura Y, et al: Role of the caspase-1 inflammasome in Salmonella typhimurium pathogenesis. J Exp Med 2006; 203:1407-1412 32. Brightbill HD, Libraty DH, Krutzik SR, et al: Host defense mechanisms triggered by microbial lipoproteins through toll-like receptors. Science 1999; 285:732-736 33. Bruno TF, Woods DE, Mody CH: Exoenzyme S from Pseudomonas aeruginosa induces apoptosis in T lymphocytes. J Leukoc Biol 2000; 67:808-816 34. Wesche H, Henzel WJ, Shillinglaw W, et al: MyD88: An adapter that recruits IRAK to the IL-1 receptor complex. Immunity 1997; 7:837-847 35. Reynaert NL, van der Vliet A, Guala AS, et al: Dynamic redox control of NF-kappaB through glutaredoxin-regulated S-glutathionylation of inhibitory kappaB kinase beta. Proc Natl Acad Sci USA 2006; 103:13086-13091 36. Pahl HL: Activators and target genes of rel/NF-kappaB transcription factors. Oncogene 1999; 18:6853-6866 37. Blackwell TS, Yull FE, Chen CL, et al: Multiorgan nuclear factor kappa B activation in a transgenic mouse model of systemic inflammation. Am J Respir Crit Care Med 2000; 162:1095-1101 38. Arcaroli J, Silva E, Maloney JP, et al: Variant IRAK-1 haplotype is associated with increased nuclear factor-kappaB activation and worse outcomes in sepsis. Am J Respir Crit Care Med 2006; 173:1335-1341 39. Zaph C, Troy AE, Taylor BC, et al: Epithelial cell-intrinsic IKK-beta expression regulates intestinal immune homeostasis. Nature 2007; 446:552-556 40. Liu YJ, Soumelis V, Watanabe N, et al: TSLP: An epithelial cell cytokine that regulates T cell differentiation by conditioning dendritic cell maturation. Annu Rev Immunol 2007; 25:193-219 41. Peck-Palmer OM, Unsinger J, Chang KC, et al: Deletion of MyD88 markedly attenuates sepsis-induced T and B lymphocyte apoptosis but worsens survival. J Leukoc Biol 2008; 83:1009-1018 42. Cantley LC. The phosphoinositide 3-kinase pathway. Science 2002; 296:1655-1657 43. Okugawa S, Ota Y, Kitazawa T, et al: Janus kinase2 is involved in lipopolysaccharide-induced activation of macrophages. Am J Physiol Cell Physiol 2003; 285:C399-C408 44. Ojaniemi M, Glumoff V, Harju K, et al: Phosphatidylinositol kinase is involved in Toll-like receptor 4-mediated cytokine

15 of 21

4/23/10 7:38 PM

Molecular Biology of Inammation and Sepsis: A Primer (printe...

http://www.medscape.com/viewarticle/585997_print

expression in mouse macrophages. Eur J Immunol 2003; 335:97-605 45. Guillot L, Le GofficR, BlochS, et al: Involvement of toll-like receptor 3 in the immune response of lung epithelial cells to double-stranded RNA and influenza A virus. J Biol Chem 2005; 280:5571-5580 46. Wrann CD, Tabriz NA, Barkhausen T, et al: The phosphatidylinositol 3-kinase signaling pathway exerts protective effects during sepsis by controlling C5a-mediated activation of innate immune functions. J Immunol 2007; 178:5940-5948 47. Guha M, Mackman N: The PI3K-Akt pathway limits LPS activation of signaling pathways and expression of inflammatory mediators in human monocytic cells. J Biol Chem 2002; 277:32124-32132 48. Schabbaue G, Tencati M, Pedersen B, et al: PI3K-Akt pathway suppresses coagulation and inflammation in endotoxemic mice. Arterioscler Thromb Vasc Biol 2004; 24:1963-1969 49. Pengal RA, Ganesan LP, Wei G, et al: Lipopolysaccharide-induced production of interleukin-10 is promoted by the serin threonine kinase Akt. Mol Immunol 2006; 43:1557-1156 50. Fukao T, Koyasu S: PI3K and negative regulation of TLR signaling. Trends Immunol 2003; 24:358-363 51. Bommhardt U, Chang KC, Swanson PE, et al: Akt decreases lymphocyte apoptosis and improves survival in sepsis. J Immunol 2004; 172:7583-7591 52. Mc Dunn JE, Muenzer JT, Rachdi L, et al: Peptide-mediated activation of Akt and extracellular regulated kinase signaling prevents lymphocyte apoptosis. FASEB J 2008; 22:561-568 53. Hall A: Rho GT. Pases and the actin cytoskeleton. Science 1998; 279:509-514 54. Ruse M, Knaus UG: New players in TLR-mediated innate immunity. Immunologic Res 2006; 34:33-48 55. Arbibe L, Mira JP, Teusch N, et al: Toll-like receptor 2-mediated NF-kappa B activation requires a Racl dependenpt pathway. Nat Immunol 2000; 1:533-540 56. Honing H, van den Berg TK, van der Pol SM, et al: Rho activation promotestransendothelial migration of monocytes via ROCK. J Leukoc Biol 2004; 75:523-528 57. Tasaka S, Koh H, Yamada W, et al: Attenuation of endotoxin-induced acute lung injury by the Rho-associated kinase inhibitor, Y-27632. Am J Respir Cell Mol Biol 2005; 32:504-510 58. Thorlacius K, Slotta JE, Laschke MW, et al: Protective effect of fasudil, a Rho-kinase inhibitor, on chemokine expression, leukocyte recruitment, and hepatocellular apoptosis in septic liver injury. J Leukoc Biol 2006; 79:923-931 59. Hori S, Nomura T, Sakaguchi S: Control of regulatory T cell development by the transcription factor FoxP3. Science 2003; 299:1057-1061 60. Zheng SG, Wang JH, Koss MN, et al: CD4+ and CD8+ regulatory T cells generated ex vivo with IL-2 and TGF-beta suppress a stimulatory Graft-versus- Host disease with a Lupus-like syndrome. J Immunol 2004; 172:1531-1539 61. Ausubel FM: Are innate immune signaling pathways in plants and animals conserved? Nat Immunol 2005; 6:973-979 62. Caramalho I, Lopes-Carvalho T, Ostler D, et al: Regulatory T cells selectively express toll-like receptors and are activated by lipopolysaccharide. J Exp Med 2003; 197:403-411 63. Sutmuller RPM, den Brok MH, Kramer M, et al: Toll-like receptor 2 controls expansion and function of regulatory T cells. J Clin Invest 2006; 116:485-494 64. Liu H, Komai-Koma M, Xu D, et al: Toll-like receptor 2 signaling modulates the functions of CD4+CD25+ regulatory T cells. Proc Natl Acad Sci USA 2006; 103:7048-7053 65. McHugh RS, Shevach EM: The role of suppressor T cells in regulation of immune responses. J Allergy Clin Immunol 2002; 110:693-702 66. Murphy TJ, Choileain NN, Zang Y, et al: CD4+CD25+ regulatory T cells control innate immune reactivity after injury. J Immunol 2005; 174:2957-2963 67. Monneret G, Debard AL, Venet F, et al: Marked elevation of human circulating CD4+CD25+ regulatory T cells in sepsisinduced immunoparalysis. Crit Care Med 2003; 31:2068-2071 68. Venet F, Pachot A, Debard AL, et al: Increased percentage of CD4+CD25+ regulatory T cells during septic shock is due to the decrease of CD4+CD25-lymphocytes. Crit Care Med 2004; 32:2329-2331 69. Takahashi H, Tsuda Y, Takeuchi D, et al: Influence of systemic inflammatory response syndrome on host resistance against bacterial infections. Crit Care Med 2004; 32:1879-1885 70. Tiemessen MM, Jagger AL, Evans HG, et al: CD4+CD25+Foxp3+ regulatory T cells induce alternative activation of human monocytes/macrophages. Proc Natl Acad Sci USA 2007; 104:19446-19451 71. Wisnoski N, Chung CS, Chen Y, et al: The contribution of CD4+ CD25+ T-regulatory-cells to immune suppression in sepsis. Shock 2007; 27:251-257 72. van Maren WW, Jacobs JF, de Vries IJ, et al: Toll-like receptor signalling on Tregs: To suppress or not to suppress? Immunology 2008; 124:445-452. 73. Heuer JG, Zhang T, Zhao J, et al: Adoptive transfer of in vitro-stimulated CD4+CD25+ regulatory T cells increases bacterial clearance and improves survival in polymicrobial sepsis. J Immunol 2005; 174:7141-7146

16 of 21

4/23/10 7:38 PM

Molecular Biology of Inammation and Sepsis: A Primer (printe...

http://www.medscape.com/viewarticle/585997_print

74. Venet F, Pachot A, Debard AL, et al: Human CD4+CD25+ regulatory T lymphocytes inhibit lipopolysaccharide-induced monocyte survival through a Fas/Fas ligand-dependent mechanism. J Immunol 2006; 177:6540-6547 75. Brown KA, Brain SD, Pearson JD, et al: Neutrophils in development of multiple organ failure in sepsis. Lancet 2006; 368:157-169 76. Marshal JC: The pathogenesis and molecular biology of sepsis. Crit Care Resusc 2006; 8:227-229 77. Chandra A, Enkhbaatar P, Nakano Y, et al: Sepsis: Emerging role of nitric oxide and selectins. Clinics 2006; 61:71-76 78. Hu G, Predescu D, Vogel SM: ICAM-1 dependent neutrophil adhesion to endothelial cells increases caveolae-mediated pulmonary vascular albumin permeability. Crit Care Med 2007; 34 (Suppl 12): A31 79. Garrean S, Gao X-P, Brovkovych V, et al: Caveolin-1 regulates NF-kB activation and lung inflammatory response to sepsis induced by lipopolysaccharide. J Immunol 2006 177: 4853-4860 80. Wang XM, Kim HP, Song R, et al: Caveolin-1 confers antiinflammatory effects in murine macrophages via the MKK3/p38 MAPK pathway. Am J Respir Cell Mol Biol 2006; 34:434-442 81. Medina FA, de Almeida CJ, Dew E, et al: Caveolin-1 deficient mice show defects in innate immunity and inflammatory immune response during Salmonella enterica serovar typhimurium infection. Infect Immun 2006; 74:6665-6674 82. Medina FA, Cohen AW, de Almeida CJ, et al: Immune dysfunction in caveolin-1 null mice following infection with Trypanosoma cruzi (Tulahuen strain). Microbes Infect 2007; 9:325-333 83. Astiz ME, DeGent GE, Lin RY, et al: Microvascular function and rheologic changes in hyperdynamic sepsis. Crit Care Med 1995; 23:265-271 84. Ozdulger A, Cinel I, Koksel O, et al: The protective effect of N-acetylcysteine on apoptotic lung injury in cecal ligation and puncture-induced sepsis model. Shock 2003; 19:366-372 85. Stehr SN, Knels L, Weissflog C, et al: Effects of IgM enriched solution on polymorphonuclear neutrophil function, bacterial clearance, and lung histology in endotoxemia. Shock 2007; 29:167-172 86. Kabay B, Kocaefe C, Baykal A, et al: Interleukin-10 gene transfer: Prevention of multiple organ injury in a murine cecal ligation and puncture model of sepsis. World J Surg 2007; 31:105-115 87. Ozdulger A, Cinel I, Unlu A, et al: Poly (ADP-ribose) synthetase inhibition prevents lipopolysaccharide-induced peroxynitrite mediated damage in diaphragm. PharmacolRes 2002; 46:67-73 88. Taner S, Cinel I, Ozer L, et al: Poly (ADP-ribose) synthetase inhibition reduces bacterial translocation in rats after endotoxin challenge. Shock 2001; 16:159-162 89. Cinel I, Buyukafsar K, Cinel L, et al: The role of poly (ADP-ribose) synthetase inhibition in preventing endotoxemia-induced intestinal epithelial apoptosis. Pharmacol Res 2002; 46:119-127 90. Savill J: Apoptosis in resolution of inflammation. J Leukoc Biol 1997; 61:375-380 91. Maianski NA, Geissler J, Srinivasula SM, et al: Functional characterization of mitochondria in neutrophils: A role restricted to apoptosis. Cell Death Differ 2004; 11:143-153 92. van Raam BJ, Verhoeven AJ, Kuijpers TW: Mitochondria in neutrophil apoptosis. Int J Hematol 2006; 84:199-204 93. Sookhai S, Wang JJ, McCourt M, et al: A novel therapeutic strategy for attenuating neutrophil-mediated lung injury in vivo. Ann Surg 2002; 235:285-291 94. Taneja R, Parodo J, Kapus A, et al: Delayed neutrophil apoptosis in sepsis is associated with maintenance of mitochondrial transmembrane potential and reduced caspase-9 activity. Crit Care Med 2004; 32:1460-1469 95. Fadok VA, Bratton DL, Konowal A, et al: Macrophages that have ingested apoptotic cells in vitro inhibit proinflammatory cytokine production through autocrine/paracrine mechanisms involving TGF, PGE2, and PAF. J Clin Invest 1998; 101:890-898 96. Sanford AN, Suriano AR, Herche D, et al: Abnormal apoptosis in chronic granulomatous disease and autoantibody production characteristic of lupus. Rheumatology 2006; 45:178-181 97. Daleke DL: Regulation of transbilayer plasma membrane phospholipid asymmetry. J Lipid Res 2003; 44:233-242 98. Bayir H, Kagan VE: Bench-to-bedside review: Mitochondrial injury, oxidative stress and apoptosis-there is nothing more practical than a good theory. Crit Care 2008; 12:206 99. Tyurina YY, Basova LV, Konduru NV, et al: Nitrosative stress inhibits the aminophospholipid translocase resulting in phosphatidylserine externalization and macrophage engulfment: Implications for the resolution of inflammation. J Biol Chem 2007; 282:8498-8509 100. Swan R, Chung CS, Albina J, et al: Polymicrobial sepsis enhances clearance of apoptotic immune cells by splenic macrophages. Surgery 2007; 142:253-261 101. Macdonald J, Galley HF, Webster NR: Oxidative stress and gene expression in sepsis. Br J Anaest 2003; 90:221-232 102. Matejovic M, Krouzecky A, Rokyta R Jr, et al: Effects of combining inducible nitric oxide synthase inhibitor and radical scavenger during porcine bacteremia. Shock 2007; 27:61-68 103. Fialkow L, Wang Y, Downey GP: Reactive oxygen and nitrogen species as signaling molecules regulating neutrophil function. Free Radic Biol Med 2007; 42:153-164

17 of 21

4/23/10 7:38 PM

Molecular Biology of Inammation and Sepsis: A Primer (printe...

http://www.medscape.com/viewarticle/585997_print

104. Galley HF, Davies MJ, Webster NR: Xanthine oxidase activity and free radical generation in patients with sepsis syndrome. Crit Care Med 1996; 24:1649-1653 105. Soriano FG, Nogueira AC, Caldini EG, et al: Potential role of poly (adenosine 5'-diphosphate-ribose) polymerase activation in the pathogenesis of myocardial contractile dysfunction associated with human septic shock. Crit Care Med 2006; 34:1073-1079 106. Ulloa L, Tracey KJ: The cytokine profile: A code for sepsis. Trends Mol Med 2005; 11:56-63 107. Zimmermann K, Volkel D, Pable S, et al: Native versus recombinant high-mobility group B1 proteins: functional activity in vitro. Inflammation 2004; 28:221-229 108. Rouhiainen A, Tumova S, Valmu L, et al: Pivotal advance: Analysis of proinflammatory activity of highly purified eukaryotic recombinant HMGB1 (amphoterin). J Leukocyte Biol 2007; 81:49-58 109. Bonaldi T, Talamo F, Scaffidi P, et al: Monocytic cells hyperacetylate chromatin protein HMGB1 to redirect it towards secretion. EMBO J 2003; 22:5551-5560 110. Youn JH, Shin JS: Nucleocytoplasmic shuttling of HMGB1 is regulated by phosphorylation that redirects it toward secretion. J Immunol 2006; 177:7889-7897 111. Bell CW, Jiang W, Reich CF III, et al: The extracellular release of HMGB1 during apoptotic cell death. Am J Physiol Cell Physiol 2006; 291:C1318-C1325 112. Qin S, Wang H, Yuan R, et al: Role of HMGB1 in apoptosis-mediated sepsis lethality. J Exp Med 2006; 203:1637-1642 113. Hori O, Brett J, Slattery T, et al: The receptor for advanced glycation end products (RAGE) is a cellular binding site for amphoterin: Mediation of neurite outgrowth and co-expression of rage and amphoterin in the developing nervous system. J Biol Chem 1995; 270:25752-25761 114. Park JS, Svetkauskaite D, He Q, et al: Involvement of toll-like receptors 2 and 4 in cellular activation by high mobility group box 1 protein. J Biol Chem 2004; 279:7370-7377 115. Wang H, Bloom O, Zhang M, et al: HMG-1 as a late mediator of endotoxin lethality in mice. Science 1999; 285:248-251 116. Mitola S, Belleri M, Urbinati C, et al: Cutting edge: Extracellular high mobility group box-1 protein is a proangiogenic cytokine. J Immunol 2006; 176:12-15 117. Tzeng HP, Fan J, Vallejo JG, et al: Negative inotropic effects of high-mobility group box 1 protein in isolated contracting cardiac myocytes. Am J Physiol Heart Circ Physiol 2008; 294:H1490-H1496 118. Wang H, Yu M, Ochani M, et al: Nicotinic acetylcholine receptor 7 receptor is an essential regulator of inflammation. Nature 2003; 421:384-388 119. Pavlov VA, Wang H, Czura CJ, et al: The cholinergic anti-inflammatory pathway: A missing link in neuromodulation. Mol Med 2003; 9:125-134 120. Wang H, Liao H, Ochani M, et al: Cholinergic agonists inhibit HMGB1 release and improve survival in experimental sepsis. Nat Med 2004; 10:1216-1221 121. Huston JM, Gallowitsch-Puerta M, Ochani M, et al: Transcutaneous vagus nerve stimulation reduces serum high mobility group box 1 levels and improves survival in murine sepsis. Crit Care Med 2007; 35:2762-2768 122. Goldstein RS, Bruchfeld A, Yang L, et al: Cholinergic anti-inflammatory pathway activity and High Mobility Group Box-1 (HMGB1) serum levels in patients with rheumatoid arthritis. Mol Med 2007; 13(3-4):210-215 123. Youn JH, Oh YJ, Kim ES, et al: High Mobility Group Box 1 protein binding to lipopolysaccharide facilitates transfer of lipopolysaccharide to CD14 and enhances lipopolysaccharide-mediated TNF-{alpha} production in human monocytes. J Immunol 2008; 180:5067-5074 124. Sha Y, Zmijewski J, Xu Z, et al: HMGB1 develops enhanced proinflammatory activity by binding to cytokines. J Immunol 2008; 180:2531-2537 125. Sunden-Cullberg J, Norrby-Teglund A, Rouhiainen A, et al: Persistent elevation of high mobility group box-1 protein (HMGB1) in patients with severe sepsis and septic shock. Crit Care Med 2005; 33: 564-573 126. Gibot S, Massin F, Cravoisy A, et al: High-mobility group box 1 protein plasma concentrations during septic shock. Intensive Care Med 2007; 33:1347-1353 127. Karlsson S, Pettila V, Tenhunen J, et al: HMGB1 as a predictor of organ dysfunction and outcome in patients with severe sepsis. Intensive Care Med 2008; 34:1046-1053. 128. Harris HE, Raucci A: Alarmin(g) news about danger. EMBO Rep 2006; 7:774-778 129. Chavakis T, Bierhaus A, Al-Fakhri N, et al: The pattern recognition receptor (RAGE) is a counterreceptor for leukocyte integrins: A novel pathway for inflammatory cell recruitment. J Exp Med 2003; 198:1507-1515 130. Lutterloh EC, Opal SM, Pittman DD, et al: Inhibition of the RAGE products increases survival in experimental models of severe sepsis and systemic infection. Crit Care 2007; 11:R122 131. Calandra T, Bernhagen J, Metz CN, et al: MIF as a glucocorticoid-induced modulator of cytokine production. Nature 1995; 377:68-71

18 of 21

4/23/10 7:38 PM

Molecular Biology of Inammation and Sepsis: A Primer (printe...

http://www.medscape.com/viewarticle/585997_print

132. Calandra T, Echtenacher B, Roy DL, et al: Protection from septic shock by neutralization of macrophage migration inhibitory factor. Nat Med 2000; 6:164-170 133. Bozza M, Satoskar AR, Lin G, et al: Targeted disruption of migration inhibitory factor gene reveals its critical role in sepsis. J Exp Med 1999; 189:341-346 134. Roger T, David J, Glauser MP, et al: MIF regulates innate immune responses through modulation of Toll-like receptor 4. Nature 2001; 414:920-924 135. Mitchell RA, Liao H, Chesney J, et al: Macrophage migration inhibitory factor (MIF) sustains macrophage proinflammatory function by inhibiting p53: Regulatory role in the innate immune response. Proc Natl Acad Sci USA 2002; 99:345-350 136. Kreymann G, Grosser S, Buggisch P, et al: Oxygen consumption and resting metabolic rate in sepsis, sepsis syndrome, and septic shock. Crit Care Med 1993; 21:1012-1019 137. Rosser DM, Stidwill RP, Jacobson D, et al: Oxygen tension in the bladder epithelium rises in both high and low cardiac output endotoxemic sepsis. J Appl Physiol 1995; 79:1878-1882 138. Hotchkiss RS, Swanson PE, Freeman BD, et al: Apoptotic cell death in patients with sepsis, shock, and multiple organ dysfunction. Crit Care Med 1999; 27:1230-1251 139. Brealey D, et al: Mitochondrial dysfunction in a long-term rodent model of sepsis and organ failure. Am J Physiol Regul Integr Comp Physiol 2004; 286:R491-R497 140. Crouser ED: Mitochondrial dysfunction in septic shock and multiple organ dysfunction syndrome. Mitochondrion 2004; 4:729-741 141. Fink MP: Cytopathic hypoxia. Mitochondrial dysfunction as mechanism contributing to organ dysfunction in sepsis. Crit Care Clinics 2001; 17:219-237 142. Brealey D, Brand M, Hargreaves I, et al: Association between mitochondrial dysfunction and severity and outcome of septic shock. Lancet 2002; 360:219-223 143. Crouser ED, Julian MW, Blaho DV, et al: Endotoxin-induced mitochondrial damage correlates with impaired respiratory activity. Crit Care Med 2002; 30:276-284 144. Levy RJ, Vijayasarathy C, Raj NR, et al: Competitive and noncompetitive inhibition of myocardial cytochrome C oxidase in sepsis. Shock 2004; 21:110-114 145. Rabuel C, Renaud E, Brealey D, et al: Human septic myopathy: Induction of cyclooxygenase, heme oxygenase and activation of the ubiquitin proteolytic pathway. Anesthesiology 2004; 101:583-590 146. Larche J, Lancel S, Hassoun SM, et al: Inhibition of mitochondrial permeability transition prevents sepsis-induced myocardial dysfunction and mortality. J Am Coll Cardiol 2006; 48:377-385 147. Levi M, Ten Cate H: Disseminated intravascular coagulation. N Engl J Med 1999; 341:586-592 148. Schouten M, Wiersinga WJ, Levi M, et al: Inflammation, endothelium, and coagulation in sepsis. J Leukoc Biol 2008; 83:536-545 149. Opal SM: The nexus between systemic inflammation and disordered coagulation in sepsis. J Endotoxin Res 2004; 10:125-129 150. Esmon CT: Are natural anticoagulants candidates for modulating the inflammatory response to endotoxin? Blood 2000; 95:1113-1116 151. Johnson K, Choi Y, DeGroot E, et al: Potential mechanisms for a proinflammatory vascular cytokine response to coagulation activation. J Immunol 1998; 160:5130-5135 152. Stouthard JM, Levi M, Hack CE, et al: Interleukin-6 stimulates coagulation, not fibrinolysis, in humans. Thromb Haemost 1996; 76:738-742 153. Aupeix K, Hugel B, Martin T, et al: The significance of shed membrane particles during programmed cell death in vitro, and in vivo, in HIV-1 infection. J Clin Invest 1997; 99:1546-1554 154. Soriano AO, Jy W, Chirinos JA, et al: Levels of endothelial and platelet microparticles and their interactions with leukocytes negatively correlate with organ dysfunction and predict mortality in severe sepsis. Crit Care Med 2005; 33:2540-2546 155. Fujimi S, Ogura H, Tanaka H, et al: Activated polymorphonuclear leukocytes enhance production of leukocyte microparticles with increased adhesion molecules in patients with sepsis. J Trauma 2002; 52:443-448 156. Amaral A, Opal SM, Vincent JL: Coagulation in sepsis. Intensive Care Med 2004; 30:1032-1040 157. Vervloet MG, Thijs LG, Hack CE: Derangements of coagulation and fibrinolysis in critically ill patients with sepsis and septic shock. Semin Thromb Hemost 1998; 24:33-44 158. Yan SB, Dhainut JF: Activated protein C versus protein C in severe sepsis. Crit Care Med 2001; 29(Suppl):S69-S74 159. Xue M, Campbell D, Jackson CJ: Protein C is an autocrine growth factor for human skin keratinocytes, J Biol Chem 2007; 282:13610-13616 160. Xue M, March L, Sambrook PN, et al: Endothelial protein C receptor is overexpressed in rheumatoid arthritic (RA) synovium and mediates the anti-inflammatory effects of activated protein C in RA monocytes. Ann Rheum Dis 2007; 66:1574-1580 161. Regan LM, Stearns-Kurosawa DJ, Kurosawa S, et al: The endothelial cell protein C receptor: Inhibition of activated protein C

19 of 21

4/23/10 7:38 PM

Molecular Biology of Inammation and Sepsis: A Primer (printe...

http://www.medscape.com/viewarticle/585997_print

anticoagulant function without modulation of reaction with proteinase inhibitors. J Biol Chem 1996; 271:17499-17503 162. Kurosawa S, Stearns-Kurosawa DJ, Carson CW, et al: Plasma levels of endothelial cell protein C receptor are elevated in patients with sepsis and systemic lupus erythematosus: Lack of correlation with thrombomodulin suggests involvement of different pathological processes. Blood 1998; 91:725-727 163. Yan SB, Helterbrand JD, Hartman DL, et al: Low levels of protein C are associated with poor outcome in severe sepsis. Chest 2001; 120:915-922 164. LaRosa SP, Opal SM, Utterback B, et al: Decreased protein C, protein S, and antithrombin levels are predictive of poor outcome in Gram-negative sepsis caused by Burkholderia pseudomallei. Int J Infect Dis 2006; 10:25-31 165. Yan SB, Dhainaut JF: Activated protein C versus protein C in severe sepsis. Crit Care Med 2001; 29(Suppl):S69-S74 166. Joyce DE, Gelbert L, Ciaccia A, et al: Gene expression profile of antithrombotic protein C defines new mechanisms modulating inflammation and apoptosis. J Biol Chem 2001; 14:11199-111203 167. Jackson CJ, Xue M: Activated protein C-An anticoagulant that does more than stop clots. Int J Biochem Cell Biol 2008; 40:2692-2697 168. Dutt T, Toh CH: The yin-yang of thrombin and activated protein C. Br J Haematology 2008; 140:505-515 169. Riewald M, 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 170. Bae JS, Yang L, Rezaie AR: Receptors of the protein C activation and activated protein C signaling pathways are colocalized in lipid rafts of endothelial cells. Proc Natl Acad Sci USA 2007; 104:2867-2872 171. Bae JS, Yang L, Manithody C, et al: The ligand occupancy of endothelial protein C receptor switches the protease-activated receptor 1-dependent signaling specificity of thrombin from a permeability-enhancing to a barrier-protective response in endothelial cells. Blood 2007; 110:3909-3916 172. Shorr AF, Bernard GR, Dhainaut JF, et al: Protein C concentrations in severe sepsis: An early directional change in plasma levels predicts outcome. Crit Care 2006; 10:R92 173. Cinel I, Dellinger RP: Current treatment of severe sepsis. Curr Infect Dis Rep 2006; 8:358-365 174. Dellinger RP, Levy MM, Carlet JM, et al: Surviving Sepsis Campaign: International guidelines for management of severe sepsis and septic shock: 2008. Crit Care Med 2008; 36:296-327 175. Bae JS, Yang L, Manithody C, et al: Engineering a disulfide bond to stabilize the calcium binding loop of activated protein C eliminates its anticoagulant but not protective signaling properties. J Biol Chem 2007; 282:9251-9259 176. Kerschen EJ, Fernandez JA, Cooley BC, et al: Endotoxemia and sepsis mortality reduction by non-anticoagulant activated protein C. J Exp Med 2007; 204:2439-2448 177. Moake JL: Thrombotic microangiopathies. N Engl J Med 2002; 347:589-600 178. Ruggeri ZM: Von Willebrand factor, platelets and endothelial cell interactions. J Thromb Haemost 2003; 1:1335-1342 179. Furlan M, Robles R, Galbusera M, et al: von Willebrand factor-cleaving protease in thrombotic thrombocytopenic purpura and the hemolytic-uremic syndrome. New Engl J Med 1998; 339:1578-1584 180. Arya M, Anvari B, Romo GM, et al: Ultralarge multimers of von Willebrand factor form spontaneous high-strength bonds with the platelet glycoprotein Ib-IX complex: Studies using optical tweezers. Blood 2002; 99:3971-3977 181. Bianchi V, Robles R, Alberio L, et al: Von Willebrand factor-cleaving protease (ADAMTS13) in thrombocytopenic disorders: A severely deficient activity is specific for thrombotic thrombocytopenic purpura. Blood 2002; 100:710-713 182. Ono T, Mimuro J, Madoiwa S, et al: Severe secondary deficiency of von Willebrand factor-cleaving protease (ADAMTS13) in patients with sepsis-induced disseminated intravascular coagulation: Its correlation with development of renal failure. Blood 2006; 107:528-534 183. Nguyen TC, Liu A, Liu L, et al: AcquiredADAMTS-13 deficiency in pediatric patients with severe sepsis. Haematologica 2007; 92:121-124 184. Kremer Hovinga JA, Zeerleder S, Kessler P, et al: ADAMTS-13, von Willebrand factor and related parameters in severe sepsis and septic shock. J Thromb Haemost 2007; 5:2284-2290 185. Bockmeyer CL, Claus RA, Budde U, et al: Inflammation-associated ADAMTS13 deficiency promotes formation of ultra-large von Willebrand factor. Haematologica 2008; 93:137-140 186. Reiter RA, Varadi K, Turecek PL, et al: Changes in ADAMTS13 (von-Willebrand factor-cleaving protease) activity after induced release of von Willebrand factor during acute systemic inflammation. Thromb Haemost 2005; 93:554-558 187. Martin K, Borgel D, Lerolle N, et al: Decreased ADAMTS-13 (A disintegrin-like and metalloprotease with thrombospondin type 1 repeats) is associated with a poor prognosis in sepsis-induced organ failure. Crit Care Med 2007; 35:2375-2382 188. Grewal PK, Uchiyama S, Ditto D, et al: The Ashwell receptor mitigates the lethal coagulopathy of sepsis. Nat Med 2008; 14:648-655 189. van't Veer C, van der Poll T. Keeping blood clots at bay in sepsis. Nat Med 2008; 14:606-608 190. Treziack S, Cinel I, Dellinger RP, et al: Resuscitating the microcirculation in severe sepsis: The central role of nitric oxide,

20 of 21

4/23/10 7:38 PM

Molecular Biology of Inammation and Sepsis: A Primer (printe...

http://www.medscape.com/viewarticle/585997_print

emerging concepts for novel therapies, and challenges for clinical trials. Acad Emerg Med 2008; 15:1-15

Authors and Disclosures


Ismail Cinel, MD, PhD and Steven M. Opal, MD, from the Division of Critical Care Medicine (IC), The Robert Wood Johnson School of Medicine, The University of Medicine and Dentistry of New Jersey, Camden NJ; and The Infectious Disease Division (SMO), Memorial Hospital of RI, The Warren Alpert School of Medicine of Brown University, Providence, RI.

Disclaimer Dr. Opal has received grants from Wyeth and Eisai Inc. Dr. Cinel has not disclosed any potential conflicts of interest. Reprint Address For information regarding this article, E-mail: cinel- ismail@cooperhealth.edu Crit Care Med. 2009;37(1):291-304. 2009 Lippincott Williams & Wilkins

21 of 21

4/23/10 7:38 PM

Vous aimerez peut-être aussi