key: cord-0943269-ss8rksiq authors: Lewis, D.H.; Chan, D.L.; Pinheiro, D.; Armitage‐Chan, E.; Garden, O.A. title: The Immunopathology of Sepsis: Pathogen Recognition, Systemic Inflammation, the Compensatory Anti‐Inflammatory Response, and Regulatory T Cells date: 2012-03-17 journal: J Vet Intern Med DOI: 10.1111/j.1939-1676.2012.00905.x sha: 3cb3db1d96357bafc65e97ba5e4cb92c51dad4b9 doc_id: 943269 cord_uid: ss8rksiq Sepsis, the systemic inflammatory response to infection, represents the major cause of death in critically ill veterinary patients. Whereas important advances in our understanding of the pathophysiology of this syndrome have been made, much remains to be elucidated. There is general agreement on the key interaction between pathogen‐associated molecular patterns and cells of the innate immune system, and the amplification of the host response generated by pro‐inflammatory cytokines. More recently, the concept of immunoparalysis in sepsis has also been advanced, together with an increasing recognition of the interplay between regulatory T cells and the innate immune response. However, the heterogeneous nature of this syndrome and the difficulty of modeling it in vitro or in vivo has both frustrated the advancement of new therapies and emphasized the continuing importance of patient‐based clinical research in this area of human and veterinary medicine. S epsis, defined as the systemic inflammatory response to infection (Box 1), remains the major cause of death in critically ill human patients. [1] [2] [3] [4] [5] [6] Recent human studies estimate the annual incidence of sepsis to be 240-300 cases per 100,000 population, with associated costs of nearly $17 billion in the United States 7, 8 ; the rate of occurrence is also increasing at around 9% each year. 8 Although large-scale veterinary epidemiological studies are uncommon, a substantial proportion of the critically ill veterinary population is estimated to be septic. 9, 10 The case fatality rate associated with sepsis in a variety of veterinary species is reported to approach 50%, emphasizing the need for a greater understanding of the pathophysiology of sepsis to improve therapeutic practices. [11] [12] [13] The pathophysiology of sepsis remains incompletely understood. A multitude of cell types, inflammatory mediators, and coagulation factors are involved and recent research has focused on the contributions of the innate immune system and T cells in this complex syndrome. 2, [14] [15] [16] [17] [18] [19] [20] This review will present an account of the current understanding of the function of the immune system in sepsis, with emphasis on the interaction of pathogens with innate components of the immune system and the key role of the endothelium in triggering and propagating a pro-inflammatory state. In addition, the complex interplay between pro-and antiinflammatory cytokines and the spectrum of the host defense response will be discussed. Finally, the importance of regulatory T cells (Tregs) in maintaining the balance of host inflammatory mechanisms will be described. Living organisms face a constant barrage of potentially pathogenic microorganisms. Survival depends upon physical barriers to resist entry of pathogens, as well as the presence of a constitutive, or innate, immune system that can rapidly induce a defensive inflammatory response. Such a system can be found in virtually all species, suggesting that it is evolutionarily ancient and highly successful. 21, 22 The innate immune system further interacts with the adaptive immune system, based on a system of T and B lymphocytes that respond to specific epitopes. Whereas the keratinized epithelium of the dermis and the mucous lining of the body cavities discourage pathogenic colonization, various other innate defenses are employed to minimize penetration of the body wall by microorganisms, including antimicrobial peptides (AMPs; otherwise known as host-defense peptides) on mucosal surfaces. 23, 24 A comprehensive review of AMPs in veterinary species has recently been published. 25 In summary, AMPs comprise 3 main groups: digestive enzymes and peptides that disrupt the microbial cell membrane, peptides that bind essential elements, and peptides that act as decoys for microbial attachment. The 2 major classes of bactericidal AMPs in the mammalian immune system are the defensins and cathelicidins. Compared to eukaryotic cells, bacterial cell walls lack cholesterol-which acts to stabilize cell membranes-and negatively charged phospholipids. Instead, bacterial cell walls are rich in anions (basic amino acids) and thus attract the cationic defensins, which carpet the microbial membrane and institute channel formation. 26 These channels then lead to transmembrane pore formation, membrane destabilization, and microbial cell death, although recent studies suggest that their function could go beyond that of lipid bilayer perturbation. 27 Whereas a large number of AMPs form part of the non-oxidative killing mechanism within phagolysosomes in cells such as neutrophils and macrophages, a growing number are thought to be actively secreted onto epithelial surfaces of the gastrointestinal, respiratory and urinary tracts [28] [29] [30] [31] -including the ovine gastrointestinal tract 32 and the canine testis. 33 Recent work has revealed that important amounts of these AMPs are secreted not only by immune cells such as neutrophils and alveolar macrophages but also by atypical defense cells such as type II pneumocytes. 34 This theme is universal throughout the processes of the innate immune system and challenges the traditional view of a "standing army" of immune defense cells, replacing it with the concept of a body-wide, integrated community of cells contributing to pathogen vigilance. 35 Two factors are vital to the rapid ability of the innate immune system to respond to pathogen incursion: the presence of receptors against pathogen markers and the ubiquitous nature of these receptors in the body. 35 Individual receptors are genetically encoded and display strong homology within and between species. 22 These receptors are not only expressed on many effector cells of the immune system-including macrophages, neutrophils, dendritic cells, and lymphocytesbut are also found on epithelial cells, endothelial cells, and myocytes 36, 37 ; expression has also been detected in the bovine endometrium. 38 The major targets of these pattern recognition receptors (PRRs) are known as pathogen-associated molecular patterns (PAMPs), although the presence of these molecules in nonpathogenic and commensal bacteria has led to the suggestion that the term "microbial-associated molecular Although bacterial cell surface components such as LPS in the outer cell membrane of Gram-negative bacteria and lipotechoic acid (LTA) in the cell membrane of Gram-positive bacteria represent classic examples of PAMPs, recognition of "altered self" secondary to host cell colonization by viral pathogens is also likely. The innate immune system has thus developed the ability to detect markers of endogenous cell damage called "alarmins" or "danger-associated molecular pat-terns" (DAMPs) 39-41 Table 1 shows key PAMPs and their corresponding receptors, 42, 43 whereas Table 2 shows some of the confirmed interactions of DAMPs. 39,40,43-46 Nonrodent sepsis models, genetic approaches, and immunological studies have demonstrated the presence of a large number of PRRs in clinical veterinary species (Table 3) . 38, [47] [48] [49] [50] [51] [52] [53] [54] [55] Initiation of an inflammatory reaction to necrotic, rather than apoptotic, cell death would appear to be useful in host defense; however, the interaction of DAMPs and PAMPs with their receptors leads to increased case fatality rates in sepsis. 15, 19, 56 Owing to PRR cross-reactivity for both PAMPs and DAMPs, multiple positive feedback systems become established, leading to rapid progression of a global inflammatory response with consequent clinical signs (Fig 1) . 16, 19, 57 Functional interactions between PRRs, including synergy and cross-tolerance, also occur. 58, 59 If an inflammatory state persists, the very defensive mechanisms of the innate immune system designed to protect the host can lead to further tissue damage, as well as diminished antimicrobial activity that allows opportunistic secondary infections. 15, 20 Whereas the existence of such feedback systems in veterinary species can at present only be inferred, experimental data suggest PRR reactivity to both pathogen-and host-derived ligands in cattle, pigs, horses, and dogs. [60] [61] [62] [63] [64] [65] [66] Furthermore, the blunted PAMP-induced TNF, IL-6 and IL-10 response of whole blood in dogs with lymphoma is thought to underlie their higher risk of sepsis. 67 Evolutionary pressure has resulted in the encoding of a number of different host proteins within three distinct families-the Toll-like receptors (TLRs), the nucleotide-binding domain, leucine-rich repeat containing proteins (NLRs; previously designated as the nucleo-tide-binding oligomerization domain [NOD]-like receptors) and the retinoic acid-inducible gene-1 (RIG-1)-like receptors (RLRs). [68] [69] [70] Cooperation between the RLRsintracellular viral nucleic acid sensors-and the endosomal TLRs appears likely, 71-73 although little is currently known about the involvement of RLRs in the pathogenesis of sepsis. 14 This family of PRRs is therefore not considered further in this review. Whereas PRRs form a heterogenous group of proteins, certain characteristics-such as leucine-rich repeat domains, scavenger receptor cysteine-rich domains, and C-type lectin domains-can be commonly recognized. 68, 69 The Toll-Like Receptors The Toll gene was first identified in Drosophila melanogaster as encoding a transmembrane glycoprotein with a role in determination of dorsoventral polarity in the embryo. 74 Additional investigation revealed a large extracellular component with leucinerich repeats, with a cytoplasmic portion described as the TIR (Toll/interleukin-1 receptor [L-1R]) domain owing to its similarity to the intracellular region of the mammalian IL-1R. 75 Not only was the Toll protein discovered to play a role in the innate immune response to fungal infection in Drosophila, but a subsequent series of studies served to highlight the link between TLRs and the mammalian immune response. [76] [77] [78] [79] To date, 10 human and 12 murine functional TLRs have been identified by a combination of immunological techniques and examination of genetic databases. 42, 71 The cytoplasmic TIR domain of the TLRs interacts with a variety of TIR-domain-containing adaptors (Fig 2) . [80] [81] [82] [83] PAMP/DAMP ligation of the respective PRR activates a chain of kinases in the IL-1R associated kinase (IRAK) family, ultimately resulting in activation of the inhibitor of j B kinase (IKK) enzyme complex and the mitogen-activated protein kinase (MAPK) pathway (Fig 2) . [84] [85] [86] Whereas the connection of TLR activation to the NF-jB and MAPK signaling pathways is well recognized, a number of other pathways are also triggered by TLR stimulation, including Protein Kinase R/eukaryotic translation initiation New abbreviations (for previous abbreviations see Table 1 ): DAMP, danger-associated molecular pattern; RAGE, receptor for advanced glycation end products; HMGB, high mobility group box protein; HSP, heat shock protein; gp, glycoprotein; LDL, low density lipoprotein; Ig, immunoglobulin; AGEs, advanced glycation end products; S100s, S100 proteins (calgranulins). Binds free iron a1-antitrypsin Protease inhibitor a2-macroglobulin Protease inhibitor factor 2-a kinase 2 (PKR/elF2a), Notch, phosphoinositide 3-kinases (PI-3K), and small GTPases. [87] [88] [89] [90] [91] [92] TLRs are expressed by a number of different cells, including dendritic cells, macrophages, B cells, natural killer (NK) cells, endothelial cells, epithelial cells, and fibroblasts (Tables 1 and 2); their site of expression varies, with TLRs 1, 2, 4, 5, 6, and 11 present on the external plasma membrane and TLRs 3, 7, 8, and 9 in endosomes. 71, 93 TLR 2 appears to be of key importance, owing to its ability to recognize PAMPs as diverse as lipotechoic acid (from Gram-positive bacteria), peptidoglycan (from Gram-positive and Gram-negative bacteria), hemagglutinin (from measles virus), polysaccharides (from yeasts), lipoproteins (from E. coli, Borrelia burgdorferi, Mycoplasma spp. and Mycobacterium tuberculosis), as well as complete pathogens such as Clostridium spp., Chlamydophila spp., and herpes simplex virus, and a variety of endogenous ligands. [94] [95] [96] [97] Much of its wide-ranging influence stems from the formation of heterodimers with TLRs 1 and 6. 98-100 TLR 4 also has a significant role in triggering the innate immune response as it recognizes molecules such as LPS (from Gram-negative bacteria), various viral protein envelopes, and a large number of endogenous molecules ( Immune system cell apoptosis/down-regulation The pathophysiology of sepsis-an overview. A primary infectious insult-of bacterial, viral, protozoal or fungal origin-damages host tissues. The ligation of pattern-recognition receptors (TLRs and NLRs) by PAMPs and DAMPs promotes the release of both pro-and anti-inflammatory cytokines, as well as acute phase proteins, with a number of pathophysiological sequelae that ultimately lead to organ hypoperfusion, tissue hypoxia, the generation of ROS and RNS, mitochondrial dysfunction and cell death by necrosis, apoptosis, pyroptosis, and autophagy. Further DAMPs are generated, helping to perpetuate the inflammatory response, and the death of immune cells leaves the organism vulnerable to secondary, or opportunistic, infections. Abbreviations: PAMPs = pathogen-associated molecular patterns; DAMPs = danger-associated molecular patterns; TLRs = Toll-like receptors; NLRs = nucleotide-binding domain, leucine-rich repeat containing protein; RLR = retinoic acid-inducible gene-1 (RIG-1)-like receptor; NFjB = nuclear factor kappa B; PMN = polymorphonuclear cell; ROS = reactive oxygen species; RNS = reactive nitrogen species. 4 appears to be dependent on formation of a complex with other PRRs, myeloid differentiation protein-2 (MD2), membrane-bound CD14 (mCD14), and lipopolysaccharide-binding protein (LBP). 71, 104 Both TLR 2 and 4 have been extensively researched in human septic patients, in whom early upregulation of these TLRs can exacerbate the severity of illness and mortality. 19, 57 TLR 5 is expressed by epithelial cells of the respiratory and intestinal tract and is a PRR for flagellin, an important component of motile bacteria such as Salmonella spp. 105,106 TLR 5 is incriminated in sepsis, 107 although exactly how the PAMP and PRR interact in vivo remains unclear. 108 The intracellular TLRs 3, 7, 8, and 9 all share high sequence homology and recognize nucleotides. 71,109 TLR 3 is the only TLR yet discovered that does not initiate the MyD88 signaling pathway, interacting solely with TIR-domain-containing adaptor molecule 1 (TICAM1, also known as TRIF) (Fig 2) . 109,110 TLR 11, to date only identified in rats and mice, appears to recognize Toxoplasma gondii and certain pathogens of the urinary tract, although a role in sepsis has not so far been described. 111, 112 The Nucleotide-Binding Domain, Leucine-Rich Repeat Containing Proteins, and Inflammasomes The realization that TLRs could not account for the full range of PAMP recognition motivated the discovery of additional intracellular PRRs, including the NLRs. Of the NLRs, the 2 cytosolic receptors NOD1 and NOD2 were the first to be discovered; subsequent examination of genomic databases has suggested that there are at least 23 NLRs in humans and 34 in mice. [113] [114] [115] [116] [117] Common to all NLRs is their structure, comprising a leucine-rich repeat domain (thought to be the PAMP receptor region), a central NOD domain, and an N-terminal effector domain responsible for downstream signaling. [118] [119] [120] The NLRs are found in the cytosolic compartment of eukaryotic cells, although some recent evidence suggests that they can also be associated with the plasma membrane. 121, 122 The mechanisms underlying NLR contact with their respective PAMPs have yet to be confirmed in vivo, 123-125 but recognition of the PAMP triggers the activation of a series of kinases leading to the phosphorylation of Toll-like receptor (TLR) signaling pathways. The cytoplasmic Toll/interleukin-1 receptor (TIR) domain of the TLRs interacts with TIR-domain-containing adaptors, such as myeloid differentiation primary response gene 88 (MyD88), TIR-containing adaptor protein (TIRAP), TIR-domain-containing adaptor molecule 1 (TICAM1, also known as TRIF) and TIR-domain-containing adaptor molecule 2 (TICAM2, also known as TRAM). PAMP binding to the respective receptor results in the activation of either the MyD88 or TICAM1/TRIF signaling pathways. These pathways involve a series of kinases in the IL-1R associated kinase (IRAK) family, whose action eventually results in activation of the inhibitor of j B kinase (IKK) enzyme complex and the mitogen-activated protein kinase (MAPK) pathway. The IKK complex phosphorylates the inhibitory IjBa protein, thus freeing the nuclear transcription factor nuclear factor kappa B (NF-jB); triggering of the MAPK signaling pathway results in the activation of activator protein 1 (AP-1). Both NF-jB and AP-1 then initiate the transcription of cytokine genes. Abbreviations: TRAF = TNF receptor-associated factor; RIP = receptor interacting protein; TAK = TGF-b-activated kinase; NEMO = NF-jB essential modulator; TBK = TRAF family member-associated NF-jB activator (TANK)-binding kinase; IRF = interferon regulatory factor. IjB, mobilization of NF-jB, and activation of MAPK signaling (Fig 3) . 126-129 NOD1 (NLR family, caspase activation, and recruitment domain [CARD] containing 1; NLRC1) and NOD2 (NLRC2, also known as CARD15) are known to recognize muropeptides derived from peptidoglycan, a major structural component of both Gram-positive and Gram-negative bacterial cell walls (Table 1) . 115, 118 Whereas the role of NOD2 is well established in the pathogenesis of Crohn's disease, 130 little is known about the role of the NOD proteins in sepsis. However, synthetic NOD1 agonists administered to mice stimulate chemokine production, neutrophil recruitment, 131 and-in one model-shock and multiple organ dysfunction, 132 raising the possibility of NOD1 involvement in the pathogenesis of sepsis. A number of studies have been performed on the role of inflammatory cysteinyl aspartate-specific proteinases, or caspases, in murine and human sepsis. Caspase-1 has been the focus of particular attention and is activated by a macromolecular complex of around 700 kDa-one of a number of inflamma-somes 133 -comprising pro-caspase-1 together with various adapter proteins, including apoptosis-associated speck-like protein containing a carboxy-terminal CARD (ASC). 133-135 Activation of the currently identified inflammasomes depends upon the interaction of NLR family members (NLRC4; NLR family, pyrin domain containing 1 [NLRP1] and NLRP3) with their PAMP or DAMP ligands. [136] [137] [138] Caspase-1 is thought to be the executioner caspase in inflammation, important in the synthesis of active IL-1b and IL-18 and thus pathogen defense. Hence, mice deficient in caspase-1 are more susceptible to bacterial infections and sepsis [139] [140] [141] ; in contrast, mice deficient in caspase-12, which normally suppresses caspase-1 activity, show enhanced bacterial clearance and resistance to sepsis. 142 Parallel observations have been made in human patients possessing the full-length caspase-12 proenzyme-uniquely present in approximately 20% of people of African descent-which attenuates the innate immune response to endotoxin and is thought to be a risk factor for the development of severe sepsis. 143 Whereas caspase-1 is essential for host defense against Nucleotide-binding domain, leucine-rich repeat containing protein (NLR) signaling pathways. Upon PAMP ligation, NLRs recruit the serine-threonine kinase RIP-like interacting caspase-like apoptosis regulatory protein (CLARP) kinase (RICK), also known as receptor-interacting serine/threonine-protein kinase 2 (RIPK2 or RIP2), which binds to NF-jB essential modulator (NEMO), a subunit of IKK, resulting in the phosphorylation of IjB and the release of NF-jB; RICK also mediates the recruitment of transforming growth factor b-activated kinase 1 (TAK1) and together these molecules stimulate activation of the mitogen-activated protein kinase (MAPK) signaling pathway. In common with the TLR family, NLR activation leads to the transcription of inflammatory cytokine genes via the mobilization of NF-jB and AP-1. Another important consequence of NLR ligation is activation of the inflammasome, a macromolecular complex comprising pro-caspase-1 together with various adapter proteins. Caspase-1 is important in the synthesis of active IL-1b and IL-18, and induces a type of programmed cell death called pyroptosis. pathogens, its activity needs to be tightly controlled. Excessive caspase-1 activity and thus endotoxic shock was induced in mice given high doses of LPS, but mice deficient in caspase-1, 144 or ASC in another study, 145 were resistant to the lethal effects of LPS. Inflammasome activity has also been linked to the phenomenon of immunoparalysis. Thus, the expression of mRNA encoding the key inflammasome components ASC, NALP1, and caspase-1 was decreased in human monocytes during early septic shock, thought to reflect monocyte deactivation; furthermore, NALP1 mRNA abundance was linked to survival in patients with sepsis. 146 Very little is known about inflammasomes in the veterinary species, although inflammasome assembly mRNAs have been analyzed in the ovine jejunum 147 and induction of inflammasome genes in the spleen was documented in septic pigs infected with either Haemophilus parasuis 148 or Streptococcus suis. 149 After mapping of the gene sequences for human and murine PRRs, exploration of the genomic sequence of other species has allowed the identification of homologs of the majority of TLRs in dogs, 150-154 cats, 49 169, [182] [183] [184] [185] [186] [187] [188] [189] [190] [191] Other PRRs have also been identified in these species, although less is known about them than the TLRs (Table 3) . Toll-like receptors 1-10 have been identified in the bovine genome, 50, 161 with numerous studies documenting their expression in tissues as varied as the endometrium, 38,163 cornea, 158 mammary gland, 167,192 skin, 51 and lung. 168 Whereas most of these studies have utilized PCR techniques to identify TLR expression, 38, 51, 158, 163, 167 an increasing number are employing flow cytometry 192, 193 or immunohistochemistry. 168 TLR signaling in cattle is similar to that described in mice and humans (Fig 2) 162, 192, 194 and the PKR/elF2a pathway appears to be important in bovine viral diarrhea virus (BVDV) and rotavirus infections. 195, 196 Comparatively little is known about the NLR family in cattle, although mRNA encoding NOD 1 and 2 has been identified in bovine mammary tissue. 197, 198 Because of the financial implications of Johne's disease and the currently unconfirmed link between Mycobacterium avium ssp. paratuberculosis and Crohn's disease, a number of studies have examined the potential role of polymorphisms of candidate genes-including TLRs-in susceptibility to paratuberculosis. 199 Recently, TLRs 1-10 have also been cloned and sequenced in sheep 48 and goats. 174 Little is currently known about the implication of TLRs in sepsis of ruminants, although the pathways involved in triggering SIRS in sheep appear to be similar to those reported in other species. 169, 200 The impact of infectious agents upon commercial viability of pigs and the contribution of this species to human disease modelling has helped advance the characterization of porcine PRRs. 191 Thus, TLRs 1-10 have been cloned and sequenced [201] [202] [203] [204] [205] [206] [207] [208] ; surface and endosomal TLRs have been detected-both by PCR and immunohistochemistry-in a variety of porcine tissues and enhanced expression demonstrated in response to a number of infectious agents and PAMPs. 168, 183, [209] [210] [211] Members of the NLR family have also been identified [212] [213] [214] [215] [216] and current research suggests that porcine PRR signaling pathways are similar to those of other mammalian species. 169, 217 Various members of the TLR family have been identified in horses, including TLR 2, 3, 4, 5, and 9. 47, 53 Whereas many of these have been characterized by PCR, [176] [177] [178] [179] TLRs 4 and 9 have also been localized by immunohistochemistry and immunogold electron microscopy 168, 179, 181 -and a recent study demonstrated TLR 5 expression by equine neutrophils by flow cytometry. 218 Stimulation by PAMPs has increased gene expression of TLRs 2, 3, and 4 in vitro, 177 and TLR 9 in vivo, 181 whereas clinical studies have reported increased TLR 4 gene expression in both foals and adult horses with SIRS/sepsis, but no differences in expression between survivors and nonsurvivors. 219, 220 Complementary studies have demonstrated increased plasma endothelin-1 concentrations and decreased long-term survival in horses with severe versus mild-to-moderate endotoxemia. 221 As for cattle and pigs, initial investigations have indicated that the downstream signaling pathways instigated by PAMP stimulation of equine TLRs 2-4 are similar to those identified in other species. 222 To date, there have been no published reports on members of the NLR family in the horse. Less is currently known about PRRs in small animals. [53] [54] [55] Various PCR studies have confirmed the presence of members of the TLR family in both dogs, including TLRs 2, 4, 7, and 9, 61,151-154,223 and cats, including TLRs 1-9. 49, 151, 155, 156 Comparative examination of the canine genome has also identified the presence of genes encoding members of the NLRP family, integral to inflammasome and thus caspase activation. 224 A number of canine TLRs have also been detected by immunohistochemical staining and flow cytometry. 168, 209, 225, 226 The use of feline models for investigation of human diseases as diverse as type 2 diabetes mellitus and human immunodeficiency virus infection has yielded additional data on PRRs in cats, including the expression of functional TLRs by the endocrine pancreas 155 and modulation of TLR signaling by retroviral pathogens. 156, 227 Enhanced expression of canine TLRs has been observed in clinical cases of osteoarthritis 60 and cystic endometrial hyperplasia/ pyometra 61, 226 ; however, the majority of published research in dogs concerns the expression of PRRs in the intestinal tract, particularly in relation to inflammatory bowel disease (IBD). 150,223,228 Ongoing research is attempting to identify whether or not certain single nucleotide polymorphisms (SNPs) of PRRs can be related to the propensity for particular canine breeds to develop IBD and other immune-mediated diseases. 229- 231 The analysis of SNPs in PRR-encoding genetic sequences is also an exciting field of research in large animals, 164, 213, 232 laying the foundation for the breeding of livestock with enhanced disease resistance and the design of vaccines better able to target dendritic cells. 191, 233 Finally, recent work has documented the expression of an ortholog of human "triggering receptor expressed on myeloid cells-1" (TREM-1) by canine neutrophils. 234 Expression of TREM-1 was upregulated by microbial agonists of TLR1/2, TLR2/6, and TLR4/ MD2. 234 This receptor, which has shown promise as a biomarker of sepsis in humans, amplifies pro-inflammatory responses to microbial products. 59, 235 Inflammatory Mediators: Cytokines and Chemokines PRR ligation triggers signaling cascades that culminate in the activation of NF-jB and AP-1 via MyD88 or TICAM1/TRIF (Fig 2) . [236] [237] [238] NF-jB and AP-1 enter the nucleus and activate transcription sites for a variety of genes, including acute phase proteins, inducible nitric oxide synthase (iNOS), coagulation factors, and pro-inflammatory cytokines and chemokines, such as tumor necrosis factor (TNF)-a and ILs-1, 6, 8, and 12. The TICAM1/TRIF pathway results in the phosphorylation of interferon regulatory factors 3 and 7 (IRF3, IRF7), which likewise enter the nucleus and stimulate the transcription of genes encoding interferon (IFN)a, IFNb, and other type 1 IFN-inducible genes. [236] [237] [238] Serum concentrations of TNF-a correlate with death in certain types of human sepsis. 239, 240 Studies of naturally occurring sepsis in veterinary patients have yielded similar results, although this pattern appears not to be universal: increased serum concentrations of TNF-a correlate with mortality in canine parvovirus and neonatal septicemia in cattle and horses, but not in septic cats. 11,241,242 TNF-a is predominantly produced by activated macrophages and T cells-but also by mast cells, B cells, NK cells, neutrophils, endothelial cells, myocytes, osteoblasts, and fibroblasts-as a 26 kDa precursor (pro-TNF) expressed on the plasma membrane; there it is cleaved by TNF-converting enzyme (TACE/ADAM17) to yield a 17 kDa soluble form, both the soluble and membrane-bound forms appearing to be active. 243 TNF-a exerts its effects by interaction with one of 2 receptors, TNF receptors 1 and 2 (TNFR1, TNFR2). 244, 245 Activation of TNFR1 appears to mediate the proinflammatory and apoptotic pathways associated with inflammation, whereas TNFR2 plays a role in the promotion of tissue repair and angiogenesis. 245 However, the complexity of signaling networks operating in sepsis is underlined by the observation that NF-jB activity induces molecules that block apoptosis mediated by TNFR1, suggesting that integration of the various signals occurs in vivo. 246 Some "cooperation" between the 2 receptor types, particularly at low TNF-a concentrations, is also likely and stimulation of both TNFRs leads to further NF-jB and AP-1 release. 244, 247 Many of the classical features of inflammation can be attributed to the actions of TNF-a upon the endothelium, with increased production of iNOS and cyclo-oxygenase 2 (COX-2) leading to vasodilatation and local slowing of blood flow 248 ; TNF-a also stim-ulates the expression of endothelial adhesion molecules such as E-selectin, intercellular adhesion molecule 1 (ICAM-1), and vascular cell adhesion molecule 1 (VCAM-1). 249, 250 These 3 molecules lead to the tethering of leukocytes to the endothelial wall and their transmigration into the interstitium, accompanied by fluid and plasma macromolecules. 19 There is also evidence for the upregulation of TNF-a and other proinflammatory cytokines in a canine model of sepsis involving the intravenous infusion of low doses of LPS: increased serum concentrations of TNF-a, IL-1b, and IL-6 [251] [252] [253] were accompanied in one of the studies by increased expression of pulmonary E-selectin and ICAM-1 and the influx of neutrophils-although whether the LPS induced the expression of the adhesion molecules directly or via induction of the proinflammatory cytokines, or both, remained unclear. 251 A similar phenomenon of up-regulation of proinflammatory cytokines, including TNF-a, has also been observed in cats 254 and horses 255-257 treated with intravenous LPS, as well as the transcription of neutrophil chemoattractants by equine endothelial cells stimulated by Th2 cytokines. 258 Furthermore, concentrations of TNF-a, IL-6, and endotoxin were all higher in the blood and peritoneal fluid of horses with colic than in a healthy control group. 259 The stimulation of feline whole blood with various PAMPs has also elicited the synthesis of proinflammatory cytokines, including TNF, IL-1b, and CXCL-8. 260 Plasma nitrite and nitrate, oxidation products of NO, have been examined in canine sepsis, in which they were present at higher concentrations than in dogs with SIRS alone. 261 An additional study showed that the inflammatory response to the intravenous administration of LPS, examined by measuring serum concentrations of TNF-a, IL-1 and IL-6, was mitigated in dogs fed a diet rich in fish oils, suggesting that diet may be an adjunct to conventional antiinflammatory treatment. 262 In addition to the upregulation of iNOS, COX-2, and adhesion molecules, TNF-a also induces the expression of procoagulant proteins such as tissue factor (TF)-and down-regulates anti-coagulant factors such as thrombomodulin-leading to activation of the coagulation cascade. 263 Despite the important role for TNF-a in endothelial activation, experimental evidence suggests that direct stimulation of TLRs expressed by endothelial and vascular smooth muscle cells may provide an alternative pathway for the vascular dysfunction seen in sepsis. [264] [265] [266] In addition to TNF-a, NF-jB activation results in the transcription of a number of proinflammatory interleukins, such as IL-1, IL-6, CXCL-8 (IL-8), and IL-12 (Fig 1) . IL-1 acts in a synergistic manner with TNF-a in the "hyperacute" period after innate immune stimulation in sepsis. 267, 268 Two proinflammatory forms of IL-1 (IL-1a and IL-1b) have been identified and induce the synthesis of adhesion molecules and cytokines by endothelial cells, encouraging leukocyte activation, endothelial tethering, and transmigration into the interstitium. 269 ,270 IL-1 also upregulates iNOS and COX2 production, acts as the major endogenous pyrogen in fever, and increases corticosteroid release via hypothalamic effects. 269, 271 Another important proinflammatory cytokine in sepsis is IL-6: as well as stimulatory effects upon leukocyte activation and myeloid progenitor cell proliferation, IL-6 also triggers the acute phase response and is a powerful pyrogen. 272,273 Like TNF-a and IL-1, plasma concentrations of IL-6 are increased in sepsis and may be predictive of progression to multiple organ dysfunction and death. 17, 274, 275 Whereas enhanced gene expression of IL-6 correlates with death in septic foals, the opposite relationship appears to hold for serum IL-6 concentrations. 219,276 However, increased serum concentrations of IL-6 and IL-1b do correlate with death in reports of naturally occurring sepsis in dogs and cats. 11, 13 Serum concentrations of anti-inflammatory cytokines, most notably IL-10, also increase in sepsis. 277, 278 This acts not only to inhibit the release of TNF-a, IL-1b, and IL-6 from monocytes and macrophages, but also to induce the production of IL-1 receptor antagonist protein (IRAP-1) and soluble TNFR, thus reducing circulating concentrations of these cytokines. 279 The critical role of IL-10 in mediating the balance between pro-and anti-inflammatory processes can be seen in experimental models: IL-10 knockout mice are profoundly susceptible to sepsis, whereas IL-10 administration prevents these consequences. 280 In clinical situations, however, the pattern appears to be more complex, with increased serum IL-10 concentrations associated with mortality in septic foals and humans, 281,282 but a lower prevalence of feline infectious peritonitis in cats infected with feline coronavirus. 283 Two additional cytokines have recently been identified as being critical in sepsis. Macrophage migratory inhibitory factor (MIF), produced by the anterior pituitary gland, is present at increased concentrations in SIRS and sepsis 284, 285 ; serum concentrations have not only correlated with mortality, but inhibition of MIF appears to be protective. [286] [287] [288] Although the trigger for MIF release in vivo is unclear, it is thought to delay apoptosis of activated monocytes and macrophages, thus helping to perpetuate a proinflammatory state. 288 High mobility group box protein 1 (HMGB-1) is an endogenous protein involved in nuclear DNA stabilization. 289 However, after necrotic or apoptotic cell death, it is released into the circulation where it has direct pro-inflammatory actions. 290 In addition, HMGB-1 appears to potentiate the effect of certain PAMPs and DAMPs upon TLR-2, TLR-4, and the receptor for advanced glycation end products (RAGE). 291 HMGB-1 was initially reported as a late inflammatory mediator in sepsis, 292 although ongoing research has highlighted a number of roles in tissue repair and angiogenesis. 293 Circulating concentrations of HMGB-1 appear to correlate with mortality in canine SIRS patients, 294 but were unable to predict hospital mortality in 1 study of septic human patients. 295 The Acute Phase Response In addition to the release of cytokines and chemokines from activated immune cells, triggering of PRRs also bring about the release of large quantities of acute phase proteins (APPs) from hepatocytes; these proteins have a variety of functions designed to re-establish homeostasis, assisting in pathogen elimination and subduing inflammation. 296, 297 The acute phase response is characterized by fever, neutrophilia, activation of the coagulation, and complement cascades (classical, alternative, and mannose-binding lectin pathways), serum iron and zinc binding, enhanced gluconeogenesis, increased muscle catabolism, and altered lipid metabolism. [297] [298] [299] In general, 2 groups of APPs are recognized: type I, induced by IL-1a, IL-1b, and TNF-a, and type II, induced by IL-6. 298 As a consequence of the up-regulation of APP production, concentrations of other plasma proteins such as albumin, protein C, protein S, and antithrombin (collectively known as negative APPs) decrease. 299, 300 Granulocyte colony-stimulating factor (G-CSF) released from monocytes is also an important component of the acute phase response, thought to mediate a protective role against bacterial infection by virtue of its impact on neutrophils. 301 A number of APPs have been characterized in veterinary species, including dogs, cats, cattle, sheep, horses, and chicken, 297, 302 and they have been used as biomarkers of inflammation in both research and clinical arenas in these species. 303, 304 Recent studies have identified adiponectin and insulin-like growth factor-1 as negative acute phase proteins in a canine model of endotoxemia. 305 Much research has been conducted to determine whether or not serum levels of APPs are predictive of survival in sepsis, with procalcitonin and C-reactive protein (CRP) the focus of particular attention. 306 Plasma procalcitonin concentrations appear to correlate with bacteremia and organ dysfunction in human clinical studies. 306, 307 Whereas it also appears to be a canine APP, it does not allow the discrimination of inflammatory or infectious from neoplastic disease when measured as a whole blood PCR assay 308 ; however, extrathyroidal procalcitonin gene expression was documented in dogs with SIRS but not in healthy animals in a preliminary observational study. 309 The utility of CRP measurements has been assessed in a number of studies examining infectious, 310,311 inflammatory, 312-319 neoplastic, 320, 321 and endocrine 322 diseases in the dog, including critically ill dogs 323 : in general, serum CRP concentrations provide a sensitive but nonspecific means of measuring inflammation, offering diagnostic and prognostic information in some disorders but not others. A recent study of SIRS and sepsis in dogs demonstrated a correlation between decreasing serum CRP concentration and recovery from disease, suggesting its use as a prognostic biomarker in this context. 324 Despite not classically considered part of the innate immune response, the prevalence of coagulation disorders and disseminated intravascular coagulation (DIC) in sepsis underlines the intimate link between the inflammatory and coagulation pathways. 325,326 On a local scale, activation of coagulation may act defensively to impede the dispersal of pathogens and inflammatory mediators from the site of insult. 327 Clinical studies have supported experimental models demonstrating increased activation of coagulation, as well as downregulation of anticoagulant mechanisms and reduced fibrinolysis, in human SIRS and sepsis patients. 266, 328 Similar findings have also been reported in dogs 329 and horses. 330 Whereas DIC 331 and pulmonary thromboembolism 332 have both been recorded in association with sepsis in cats, experimental models of endotoxin infusion in this species have yielded variable results, 333,334 a recent study failing to elicit any biologically significant alterations in coagulation parameters, 334 underlining the multifactorial pathogenesis of coagulopathies in clinical patients. A detailed description of the complexities of the interaction between inflammation and coagulation is beyond the scope of this article; however, several comprehensive reviews of hemostasis in SIRS and sepsis in veterinary species have recently been published and an overview of key interactions is presented in Fig 4 [335] [336] [337] In brief, key to the triggering of the coagulation pathway in sepsis is tissue factor (TF), which initiates coagulation via the contact activation (extrinsic) pathway. In health, lack of TF exposure within the vascular system and the presence of various circulating proteinssuch as protein C, antithrombin, and tissue factor plasminogen inhibitor-modulate coagulation by the prevention of TF activation. 266, 335 The expression of TF by monocytes or macrophages and tissue parenchymal cells is activated by various inflammatory cytokines, CRP, and PAMPs such as LPS 335,338 -a phenomenon that has also been documented in cats 339 and horses (earlier studies in this species citing "procoagulant activity" rather than TF per se). [340] [341] [342] Although findings differ between species, large numbers of TF-expressing microparticles have been identified in blood samples from septic human patients and may correlate with mortality. 343, 344 These microparticles are released from a variety of activated or apoptotic cells-such as platelets, monocytes, erythrocytes, and endothelial cells-and their interaction with endothelial cells and platelets drives the coagulation pathway. [345] [346] [347] Although only an indirect measure, plasma von Willebrand factor concentrations were higher in septic dogs than those in healthy control animals, suggesting that endothelial cell activation also occurs in canine sepsis. 348 Interestingly, platelets enhanced endotoxin-induced equine monocyte TF activity in vitro 341although whether microparticles derived from platelets or other sources play a role in the pathogenesis of equine sepsis in vivo remains unknown. Additional work in an equine model of endotoxemia has shown that large volume resuscitation has no impact on coagulation parameters beyond the changes attributed to endotoxemia, providing useful additional data to inform the treatment of sepsis in this species. 349 Whereas much research has been directed toward the inhibition of coagulation in sepsis, only activated protein C has shown any benefit in human clinical trials. 350, 351 The antithrombotic and anti-inflammatory properties of recombinant human activated protein C led to recommendations for its use in severe sepsis in 2004 352 and 2008, 353 but a recent meta-analysis found no evidence in support of its administration in the treatment of severe sepsis or septic shock. 354 Indeed, there is still debate about its mechanism of action and, as yet, little experience of its use in veterinary medicine, despite encouraging pharmacological data in experimental models. 266, 355, 356 In addition to the role of TF in initiating coagulation in the presence of inflammation, activated platelets and endothelial cells-as well as bacterial surfaces-also trigger the contact phase system, leading to the formation of kallikrein and bradykinin. 357, 358 Bradykinin in turn enhances vasodilatation and increases vascular permeability, as well as reducing platelet function 359 ; kallikrein accelerates fibrinolysis by conversion of plasminogen to plasmin and causes additional activation of Factor XII, leading to stimulation of the classical complement pathway. 360, 361 A final connection between coagulation and inflammation in sepsis has become apparent with the recent exploration of the role of "a disintegrin-like and metalloproteinase with thrombospondin type-I motifs-13" (ADAMTS-13). ADAMTS-13 is produced by the stellate (Ito) cells of the liver and acts to cleave ultra-large von Willebrand's Factor (vWF) multimers into smaller multimers. 362 These ultra-large vWF multimers are released from endothelial stores after inflammation and lead to platelet activation and aggregation; the ensuing microthrombi further compromise tissue blood flow, leading to additional propagation of the proinflammatory state. 363 Although not yet identified in clinical veterinary species, decreased plasma ADAM-TS-13 activity is associated with a poor prognosis in human sepsis patients [364] [365] [366] ; decreased activity is attributed to both a diminution of hepatic production and an increase in breakdown by plasma proteases. 366 The Compensatory Anti-Inflammatory Response Syndrome and Cell Death Ongoing investigation of the molecular mechanisms of the SIRS response, as well as the notable failure of therapeutic blockade of proinflammatory mediators, led to the realization that the mortality associated with sepsis could not be explained solely by an uncontrollable "cytokine storm". 367, 368 This resulted in the concept of an opposing "compensatory anti-inflammatory response syndrome" (CARS), thought to be an adaptive response to the excessive proinflammatory process in SIRS and sepsis. 367, 369 Whereas an appropriate balance is struck in the vast majority of instances of host-defense challenge, this is lost in sepsis-leading either to an uncontrolled proinflammatory reaction to infection, resulting in organ dysfunction, an undesirable compromise of the immune system permitting opportunistic infection (the "second hit"), or a combination of both. 367 In the absence of inflammation, various anticoagulant mechanisms exist to prevent activation of coagulation. Thrombomodulin and endothelial protein C receptor interact with thrombin to activate protein C (aPC), which in turn interacts with protein S to inactivate factor V. Endothelial cells also express tissue factor pathway inhibitor (TFPI) and antithrombin on their luminal surface and secrete tissue plasminogen activator (tPA). However, the majority of these anticoagulant factors are negative acute phase proteins and thus plasma concentrations falling during sepsis. Furthermore, proinflammatory cytokines released during sepsis activate endothelial cells, platelets and circulating white blood cells, which begin to express tissue factor (TF) and lose surface anticoagulant proteins. Microparticle formation is increased, together with expression of functional adhesion molecules on platelets, microparticles, monocytes and endothelial cells. Plasma antithrombin, aPC, protein S, and TFPI concentrations fall; ADAMTS-13 activity also drops, leading to greater persistence of ultralarge von Willebrand factor, which further enhances platelet and microparticle activation and adhesion. After triggering of the contact activation (TF) pathway, the interaction of microparticles with endothelial cells and platelets further drives the coagulation pathway. The presence of coagulation pathway components (TF-VIIa complex, Xa, Va, thrombin) and other circulating proteases activates endothelial protease-activated receptors (PARs), which propagate the inflammatory process by stimulating the further release of mediators (TNFa, IL-6 and CXCL-8), platelet activation and cell death. Triggering of the contact phase system activates factor XII (Hageman factor), leading to the formation of kallikrein (KK) and bradykinin (BK). Rather than a sequential or compensatory change, as was initially proposed, SIRS and CARS appear to occur simultaneously, and act to balance the host's need to maintain defense while minimizing self-induced tissue damage. 370 Serum concentrations of both proand anti-inflammatory mediators increase early on in sepsis 370 ; likewise, concentrations of a variety of both types of mediators (eg, IL-6, IRAP-1) are predictive of septic morbidity or mortality. 370-372 Thus the changes in cytokine profile are both dynamic and heterogeneous, indicating that prescriptive immunomodulatory therapies are unlikely to meet with success. 370,373 Furthermore, apoptosis of lymphocytes, hepatocytes, gastrointestinal epithelial cells and endothelial cells is increased, whereas that of neutrophils is decreased. 374,375 Neutrophil function is also altered: both the migration of neutrophils to infected tissues 376, 377 and their antimicrobial function 378 is diminished; moreover, peritoneal neutrophils are a potential source of IL-10, suppressing inflammatory monocytes in a model of polymicrobial sepsis. 379 Although monocyte survival appears unaltered, sepsis results in the production of molecules such as IL-1 receptor associated kinase M (IRAK-M), MyD88 short variant (MyD88s), and A20-binding inhibitor of NF-jB activation3 (ABIN-3) , which reduce activation of the NF-jB signaling pathway and therefore dampen the response to PAMP recognition. 380, 381 Monocyte function is impaired in sepsis, with decreased expression of the MHC class II molecule human leukocyte antigen-DR (HLA-DR) 382, 383 and decreased inflammasome expression. 146 Longitudinal observational studies in human patients indicate that a failure to regain >70% normal monocytic expression of HLA-DR is associated with an increased risk of secondary bacterial infection and decreased survival. 384 In addition to these changes in antigen-presenting function of monocytes, dendritic cells also undergo increased apoptosis in sepsis, further impairing the host's ability to respond to pathogens. 385 Although apoptosis, or type 1 programmed cell death (PCD), is responsible for the majority of immune cell death in sepsis and is implicated in immunoparalysis, alternative pathways also play a role. 386, 387 Autophagy is a cellular mechanism that primarily acts as a cytoplasmic "clean-up" process, as well as assisting in delivery of proteins to antigen presentation pathways; however, it may also mediate type II PCD and interact with apoptosis. 388, 389 Much current interest has been directed toward the role of autophagy in trauma and sepsis, although whether it acts in a cytoprotective role or as a mechanism of PCD, or both, remains unclear. 386, 390 A final mechanism of interest in the pathogenesis of sepsis is pyroptosis, a term used to describe the process of caspase-1-mediated PCD, which is distinct from death mediated by the apoptotic caspases 3, 6, and 8. 391 Although some question whether pyroptosis is truly a unique cell death mechanism, or simply a special case of apoptosis or necrosis (oncosis), 392 it is characterized by rapid plasma membrane rupture and release of proinflamma-tory intracellular contents, some of which may act as DAMPs. 393 One of the targets of caspase-1 during sepsis is the glycolytic pathway. 394 A number of studies also suggest an important role for apoptosis in the pathogenesis of SIRS, sepsis, and infectious disease in veterinary species. Apoptotic cells were observed in the liver, kidney, thymus, stomach, and lymphocyte population of endotoxemic piglets 395 and the primary and secondary lymphoid organs of pigs infected with classical swine fever virus. 396 LPS and TNF-a both induced apoptosis of bovine glomerular endothelial cells, modeling a potential pathomechanism of acute renal failure in Gram-negative sepsis 397 ; this phenomenon could be potently inhibited by glucocorticoids in vitro. 398 Haemophilus somnus, a Gram-negative pathogen of cattle that causes sepsis and vasculitis, induces caspases 3 and 8, and subsequent apoptosis, of endothelial cells in vitro 399 ; furthermore, the bacterium stimulates platelets that are in turn able to induce endothelial cell apoptosis by a contact-dependent mechanism involving the activation of caspases 8 and 9 and the synthesis of reactive oxygen species. 400 The activity of matrix metalloproteinases 2 and 9 and the expression of phosphorylated Paxillin showed positive correlation with cardiomyocyte apoptosis in an ovine model of endotoxemic shock, 401 whereas apoptosis of peripheral blood mononuclear cells and splenocytes of sheep infected with bluetongue virus was thought to contribute to immunosuppression in this disease. 402 Ileal epithelial apoptosis was documented in one feline model of endotoxemia, 403 although a second demonstrated lymphocyte apoptosis in the spleen and Peyer's patches. 334 Apoptosis of T cells also contributes to the immunosuppression characteristic of feline immunodeficiency virus infection. 404 Finally, apoptosis of intestinal epithelial cells was observed in a canine model of sepsis induced by the intravenous infusion of E. coli. 405 One of the key features of the adaptive immune system is its ability to generate antigen-specific receptors with an enormous diversity of specificities, some of which may recognize host-derived epitopes. 406, 407 The majority of potentially autoaggressive thymocytes are deleted in the thymic medulla in a process called negative selection. 408, 409 However, this process of central tolerance is imperfect and underlines the importance of a number of peripheral tolerance mechanisms, including clonal deletion, functional inactivation (anergy), and phenotypic skewing. 410-412 Although a subject of debate for many years, a population of Tregs is now known to play a major role in peripheral tolerance, complementing the preceding (intrinsic) mechanisms. [413] [414] [415] Little is currently known about tolerance mechanisms in veterinary species, 416 but Tregs have been identified in cats, [417] [418] [419] [420] [421] [422] [423] dogs, [424] [425] [426] [427] [428] [429] [430] [431] [432] [433] pigs, [434] [435] [436] [437] [438] [439] [440] [441] [442] cows, [443] [444] [445] [446] sheep, 447, 448 and horses. [449] [450] [451] [452] [453] [454] Both naturally occurring and peripherally induced Tregs have been characterized-the former the product of a pathway of thymic differentiation called altered negative selection and the latter the product of peripheral activation of conventional T cells in the context of an environment rich in transforming growth factor b (TGF-b) or IL-10. [455] [456] [457] Naturally occurring Tregs have been identified by their constitutive expression of the IL-2 receptor a chain (CD25) and Forkhead box P3 (FOXP3), a transcription factor that plays a pivotal role in both their ontogeny and peripheral function. FOXP3 acts to stabilize the Treg transcriptome by repressing a number of pro-inflammatory and growthpromoting genes-for example, IL-2 and IFNG-while activating others encoding key molecules involved in Treg function-for example, CTLA4 and CD25. 458, 459 Naturally occurring Tregs are known to interact with cells of both the innate and adaptive immune systems-including monocytes, macrophages, natural killer cells, neutrophils, mast cells, dendritic cells, and both T and B cells-generally mediating a suppressive function to prevent the development of autoaggressive responses and maintain the population of peripheral CD4 + T cells, thus contributing to immune system homeostasis. 456, 460, 461 Tregs are also known to express a variety of TLRs, stimulation of which has augmented or abolished regulatory function in various studies. [462] [463] [464] [465] The molecular mechanisms of immune suppression mediated by naturally occurring Tregs have not been fully elucidated, but involve cell contactdependent interactions, induction of cell death, and secretion of immunosuppressive cytokines, including IL-10 and TGF-b. 416, 456 Various studies have documented increased proportions of Tregs in human sepsis during the phase of immunoparalysis, [466] [467] [468] [469] but the role of this change remains unclear because depletion of Tregs in murine models of sepsis has yielded variable conclusions between models, either improving, enhancing, or bearing no influence on mortality. [470] [471] [472] Given the ability of Tregs to induce the alternative activation pathway of macrophages 473 and to inhibit the LPS-induced survival of monocytes through a proapoptotic mechanism involving the Fas/FasL pathway, 474 they may make a potentially significant contribution to immune system dysfunction in human septic patients, 468, 469, 475, 476 although this is still a controversial area. 477 Additional research is required to elucidate the role of Tregs in sepsis, as well as to identify their mechanisms of action in this context. Moreover, to the authors' knowledge no veterinary studies have interrogated Treg number or function in septic patients to date. Pharmacological manipulation of Treg activity continues to be explored experimentally and may eventually translate to clinical cases; however, therapeutic interventions to alter the resistance of immune cells to Treg suppression may prove an equally valid alternative approach. 478, 479 Conclusions As is apparent from Figure 1 , a broad concept of the immunopathological mechanisms underlying sepsis is now generally accepted. However, many of the molecular details are both complex and incompletely elucidated. Ongoing research has also indicated the existence of multiple redundant pathways within the innate immune response, potentially explaining the failure of many highly selective therapeutic interventions. 14, 19 A single "magic bullet" for the treatment of sepsis is highly unlikely to exist: an individually tailored set of therapies, based on point-of-care assessment of the immunopathological status of the patient, is the likely future-albeit a distant one. 373 Various human studies have resulted in the publication of consensus statements regarding the treatment of sepsis, 352,353 but current veterinary evidence to substantiate these interventions is thin on the ground. What has become evident is that outside the experimental laboratory, the heterogeneous nature of the septic patient population means that clinical trials must be carefully designed to obtain meaningful data. [480] [481] [482] Ongoing basic research into the immunopathology of sepsis in clinical veterinary species is equally important, to elucidate the underlying molecular mechanisms and thus direct clinical studies to those aspects of disease likely to benefit the greatest number of patients. Definitions for sepsis and organ failure and guidelines for the use of innovative therapies in sepsis The pathophysiology and treatment of sepsis International Sepsis Definitions Conference International Sepsis Definitions Conference Sepsis in European intensive care units: Results of the SOAP study The sepsis syndrome. Definition and general approach to management Epidemiology of severe sepsis in the United States: Analysis of incidence, outcome, and associated costs of care The epidemiology of sepsis in the United States from 1979 through 2000 Systemic inflammatory response syndrome, sepsis, and multiple organ dysfunction Prognostic variables for survival of neonatal foals under intensive care Calcium regulating hormones and serum calcium and magnesium concentrations in septic and critically ill foals and their association with survival Plasma interleukin-6 response is predictive for severity and mortality in canine systemic inflammatory response syndrome and sepsis Pro-inflammatory mechanisms in sepsis Molecular biology of inflammation and sepsis: A primer The immunopathogenesis of sepsis Recent advances in sepsis research: Novel biomarkers and therapeutic targets The immunology of sepsis The pathogenesis of sepsis Antimicrobial peptides of phagocytes and epithelia Phylogenetic perspectives in innate immunity Antimicrobial peptides in health and disease Antimicrobial peptides in animals and their role in host defences Innate immunity and host defense peptides in veterinary medicine Mechanism of the binding, insertion and destabilization of phospholipid bilayer membranes by alpha-helical antimicrobial and cell non-selective membrane-lytic peptides Antibiotic activities of host defense peptides: More to it than lipid bilayer perturbation Maintaining a sterile urinary tract: The role of antimicrobial peptides Host defense peptides in the oral cavity and the lung: Similarities and differences Gastroduodenal mucosal defense Antimicrobial peptides in the airway Antimicrobial peptide expression is developmentally regulated in the ovine gastrointestinal tract CD208/dendritic cell-lysosomal associated membrane protein is a marker of normal and transformed type II pneumocytes Friendly and dangerous signals: Is the tissue in control? Recognition of commensal microflora by toll-like receptors is required for intestinal homeostasis Toll-like receptor and antimicrobial peptide expression in the bovine endometrium Tolerance, danger, and the extended family Pathogen recognition and innate immunity Innate immune sensing of pathogens and danger signals by cell surface Toll-like receptors CD24 and Siglec-10 selectively repress tissue damage-induced immune responses Emerging paradigm: Toll-like receptor 4-sentinel for the detection of tissue damage Characterization of equine and other vertebrate TLR3, TLR7, and TLR8 genes Molecular cloning and characterization of Toll-like receptors 1-10 in sheep Toll-like receptor expression in feline lymphoid tissues Radiation hybrid mapping of all 10 characterized bovine Toll-like receptors Identification and expression of Toll-like receptors 1-10 in selected bovine and ovine tissues Toll-like receptor family in domestic animal species Pattern recognition receptors in companion and farm animals The key to unlocking the door to animal disease? Variation matters: TLR structure and species-specific pathogen recognition Modulation of tissue Toll-like receptor 2 and 4 during the early phases of polymicrobial sepsis correlates with mortality Role of Toll-like receptors in the development of sepsis Synergy and cross-tolerance between Toll-like receptor (TLR) 2-and TLR4-mediated signaling pathways TREM and TREM-like receptors in inflammation and disease Expression of Tolllike receptors 2 and 4 in stifle joint synovial tissues of dogs with or without osteoarthritis Gene transcription of TLR2, TLR4, LPS ligands and prostaglandin synthesis enzymes are up-regulated in canine uteri with cystic endometrial hyperplasia-pyometra complex Early time course of altered leukocyte response to lipopolysaccharide and peptidoglycan in porcine gunshot injury Age-associated changes to pathogen-associated molecular pattern-induced inflammatory mediator production in dogs Interactions between lipopolysaccharides and blood factors on the stimulation of equine polymorphonuclear neutrophils Pathogen associated molecular pattern motifs from Gram-positive and Gramnegative bacteria induce different inflammatory mediator profiles in equine blood Blunted pathogenassociated molecular pattern motif induced TNF, IL-6 and IL-10 production from whole blood in dogs with lymphoma Host innate immune receptors and beyond: Making sense of microbial infections Innate immune recognition: Mechanisms and pathways Systems biology of innate immunity Toll-like receptors and their crosstalk with other innate receptors in infection and immunity Differential role of TLR-and RLR-signaling in the immune responses to influenza A virus infection and vaccination Roles of caspase-8 and caspase-10 in innate immune responses to double-stranded RNA The Toll gene of Drosophila, required for dorsal-ventral embryonic polarity, appears to encode a transmembrane protein Host defense mechanisms triggered by microbial lipoproteins through Toll-like receptors The dorsoventral regulatory gene cassette spatzle/Toll/cactus controls the potent antifungal response in Drosophila adults Differential roles of TLR2 and TLR4 in recognition of gram-negative and grampositive bacterial cell wall components TIRAP: An adapter molecule in the Toll signaling pathway TIR-containing adaptors in Tolllike receptor signalling MyD88 is an adaptor protein in the hToll/IL-1 receptor family signaling pathways The family of five: TIR-domaincontaining adaptors in Toll-like receptor signalling Sequential control of Toll-like receptor-dependent responses by IRAK1 and IRAK2 Essential function for the kinase TAK1 in innate and adaptive immune responses Inhibition of phosphoinositide 3-kinase enhances TRIF-dependent NFkappa B activation and IFN-beta synthesis downstream of Tolllike receptor 3 and 4 Bruton's tyrosine kinase together with PI 3-kinase are part of Toll-like receptor 2 multiprotein complex and mediate LTA induced Tolllike receptor 2 responses in macrophages Phosphatidylinositol 3-kinase is involved in Toll-like receptor 4-mediated cytokine expression in mouse macrophages PKR regulates TLR2/TLR4-dependent signaling in murine alveolar macrophages Mastoparan, a G protein agonist peptide, differentially modulates TLR4-and TLR2-mediated signaling in human endothelial cells and murine macrophages pendent Th1 differentiation through Delta 4 Notch-like ligand in response to bacterial LPS Non-LPS components of Chlamydia pneumoniae stimulate cytokine production through Toll-like receptor 2-dependent pathways Peptidoglycan-and lipoteichoic acid-induced cell activation is mediated by toll-like receptor 2 TLR2 -promiscuous or specific? A critical re-evaluation of a receptor expressing apparent broad specificity Discrimination of bacterial lipoproteins by Toll-like receptor 6 Cutting edge: Role of Toll-like receptor 1 in mediating immune response to microbial lipoproteins Cutting edge: Toll-like receptor 4 (TLR4)-deficient mice are hyporesponsive to lipopolysaccharide: Evidence for TLR4 as the Lps gene product LPS/TLR4 signal transduction pathway Toll-like receptors in innate immunity The structural basis of lipopolysaccharide recognition by the TLR4-MD-2 complex A conserved surface on Toll-like receptor 5 recognizes bacterial flagellin Regulation of humoral and cellular gut immunity by lamina propria dendritic cells expressing Toll-like receptor 5 Bacterial ligands generated in a phagosome are targets of the cytosolic innate immune system Role of adaptor TRIF in the MyD88-independent Toll-like receptor signaling pathway TLR11 activation of dendritic cells by a protozoan profilin-like protein A Toll-like receptor that prevents infection by uropathogenic bacteria An essential role for NOD1 in host recognition of bacterial peptidoglycan containing diaminopimelic acid The plant immune system Intracellular NOD-like receptors in host defense and disease NOD-LRR proteins: Role in host-microbial interactions and inflammatory disease NLRX1 is a regulator of mitochondrial antiviral immunity Membrane recruitment of NOD2 in intestinal epithelial cells is essential for nuclear factor-kappaB activation in muramyl dipeptide recognition Pannexin-1-mediated intracellular delivery of muramyl dipeptide induces caspase-1 activation via cryopyrin/NLRP3 independently of Nod2 Nod1 mediates cytoplasmic sensing of combinations of extracellular bacteria Nod1 responds to peptidoglycan delivered by the Helicobacter pylori cag pathogenicity island A critical role of RICK/RIP2 polyubiquitination in Nod-induced NF-kap-paB activation Nod1, an Apaf-1-like activator of caspase-9 and nuclear factor-kappaB TAK1 is a central mediator of NOD2 signaling in epidermal cells Nod2, a Nod1/ Apaf-1 family member that is restricted to monocytes and activates NF-kappaB NOD2, an intracellular innate immune sensor involved in host defense and Crohn's disease Nod1 acts as an intracellular receptor to stimulate chemokine production and neutrophil recruitment in vivo Selective NOD1 agonists cause shock and organ injury/dysfunction in vivo The inflammasome: A molecular platform triggering activation of inflammatory caspases and processing of proIL-beta Cytosolic flagellin requires Ipaf for activation of caspase-1 and interleukin 1beta in salmonella-infected macrophages Gout-associated uric acid crystals activate the NALP3 inflammasome Immune recognition of Pseudomonas aeruginosa mediated by the IPAF/NLRC4 inflammasome Roles of caspase-1 in Listeria infection in mice Innate immunity against Francisella tularensis is dependent on the ASC/caspase-1 axis Caspase-1 regulates Escherichia coli sepsis and splenic B cell apoptosis independently of interleukin-1beta and interleukin-18 Enhanced bacterial clearance and sepsis resistance in caspase-12-deficient mice Differential modulation of endotoxin responsiveness by human caspase-12 polymorphisms Mice deficient in IL-1 beta-converting enzyme are defective in production of mature IL-1 beta and resistant to endotoxic shock Differential activation of the inflammasome by caspase-1 adaptors ASC and Ipaf Inflammasome mRNA expression in human monocytes during early septic shock Developmental control of the Nlrp6 inflammasome and a substrate, IL-18, in mammalian intestine Understanding Haemophilus parasuis infection in porcine spleen through a transcriptomics approach Response of swine spleen to Streptococcus suis infection revealed by transcription analysis Full-length cDNA cloning of Toll-like receptor 4 in dogs and cats Cloning of canine Toll-like receptor 9 and its expression in dog tissues Molecular cloning and tissue expression of canine Toll-like receptor 2 (TLR2) Cloning of canine Toll-like receptor 7 gene and its expression in dog tissues Feline pancreatic islet-like clusters and insulin producing cells express functional Toll-like receptors (TLRs) Altered bone marrow dendritic cell cytokine production to Toll-like receptor and CD40 ligation during chronic feline immunodeficiency virus infection Characterisation of antibodies to bovine Toll-like receptor (TLR)-2 and cross-reactivity with ovine TLR2 Primary bovine colonic cells: A model to study strain-specific responses to Escherichia coli Identification of full length bovine TLR1 and functional characterization of lipopeptide recognition by bovine TLR2/1 heterodimer Kinome analysis of Toll-like receptor signaling in bovine monocytes Expression of genes associated with immunity in the endometrium of cattle with disparate postpartum uterine disease and fertility Toll-like receptors TLR1, TLR2 and TLR4 gene mutations and natural resistance to Mycobacterium avium subsp. paratuberculosis infection in cattle Characterization of bovine Toll-like receptor 8: Ligand specificity, signaling essential sites and dimerization Mastitis increases mammary mRNA abundance of beta-defensin 5, Tolllike-receptor 2 (TLR2), and TLR4 but not TLR9 in cattle Expression of Toll-like receptor 9 in lungs of pigs, dogs and cattle Comparative genomics of Toll-like receptor signalling in five species Characterization of ovine Tolllike receptor 9 protein coding region, comparative analysis, detection of mutations and maedi visna infection Toll-like receptor (TLR)6 and TLR1 differentiation in gene expression studies of Johne's disease Toll-like receptors and agonist responses in the developing fetal sheep lung Neonatal goats display a stronger TH1-type cytokine response to TLR ligands than adults Expression of Toll-like receptors 2, 3, 4, 6, 9, and MD-2 in the normal equine cornea, limbus, and conjunctiva Differential induction of Toll-like receptor gene expression in equine monocytes activated by Toll-like receptor ligands or TNF-alpha Elevated amount of Toll-like receptor 4 mRNA in bronchial epithelial cells is associated with airway inflammation in horses with recurrent airway obstruction Expression of Toll-like receptor 4 and 2 in horse lungs Genotyping of Toll-like receptor 4, myeloid differentiation factor 2 and CD-14 in the horse: An investigation into the influence of genetic polymorphisms on the LPS induced TNF-alpha response in equine whole blood Expression of Toll-like receptor 9 in horse lungs Porcine Toll-like receptors: Recognition of Salmonella enterica serovar Choleraesuis and influence of polymorphisms Urinary bladder tissue engineering using natural scaffolds in a porcine model: Role of Toll-like receptors and impact of biomimetic molecules. Cells Tissues Differential activation and maturation of two porcine DC populations following TLR ligand stimulation Polymorphism distribution and structural conservation in RNA-sensing Toll-like receptors 3, 7, and 8 in pigs Expression of porcine Toll-like receptor 2, 4 and 9 gene transcripts in the presence of lipopolysaccharide and Salmonella enterica serovars Typhimurium and Choleraesuis Expression of Toll-like receptor mRNA and cytokines in pigs infected with porcine reproductive and respiratory syndrome virus Toll-like receptor expressions in porcine alveolar macrophages and dendritic cells in responding to poly IC stimulation and porcine reproductive and respiratory syndrome virus (PRRSV) infection Bacterial lipopolysaccharide induces increased expression of Toll-like receptor (TLR) 4 and downstream TLR signaling molecules in bovine mammary epithelial cells Binding of different forms of lipopolysaccharide and gene expression in bovine blood neutrophils Functional characterization of bovine TIRAP and MyD88 in mediating bacterial lipopolysaccharide-induced endothelial NF-kappaB activation and apoptosis Modulation of PKR activity in cells infected by bovine viral diarrhea virus Muramyl dipeptide synergizes with Staphylococcus aureus lipoteichoic acid to recruit neutrophils in the mammary gland and to stimulate mammary epithelial cells Candidate gene and genome-wide association studies of Mycobacterium avium subsp. paratuberculosis infection in cattle and sheep: A review Toll-like receptor 4 inhibitor TAK-242 attenuates acute kidney injury in endotoxemic sheep Molecular cloning, characterization and tissue expression of porcine Toll-like receptor 4 Tolllike receptor 7 and MyD88 knockdown by lentivirus-mediated RNA interference to porcine dendritic cell subsets Toll-like receptor 4 and cytokine expression involved in functional immune response in an originally established porcine intestinal epitheliocyte cell line Swine Tolllike receptor 9(1) recognizes CpG motifs of human cell stimulant Biased distribution of single nucleotide polymorphisms (SNPs) in porcine Toll-like receptor 1 (TLR1), TLR2, TLR4, TLR5, and TLR6 genes Porcine Toll-like receptor 1, 6, and 10 genes: Complete sequencing of genomic region and expression analysis Genomic structure, promoter analysis and expression of the porcine (Sus scrofa) TLR4 gene Expression of inflammatory cytokines and Toll-like receptors in the brain and respiratory tract of pigs infected with porcine reproductive and respiratory syndrome virus Molecular identification and functional expression of porcine Toll-like receptor (TLR) 3 and TLR7 Influence of polymorphisms in porcine NOD2 on ligand recognition Differences in distribution of single nucleotide polymorphisms among intracellular pattern recognition receptors in pigs Porcine reproductive and respiratory syndrome virus (PRRSV) suppresses interferon-beta production by interfering with the RIG-I signaling pathway Molecular cloning and functional characterization of porcine nucleotide-binding oligomerization domain-1 (NOD1) recognizing minimum agonists, meso-diaminopimelic acid and meso-lanthionine Molecular cloning and functional characterization of porcine nucleotide-binding oligomerization domain-2 (NOD2) Disparities in TLR5 expression and responsiveness to flagellin in equine neutrophils and mononuclear phagocytes Expression of inflammation-associated genes in circulating leukocytes collected from horses with gastrointestinal tract disease Digital blood flow and plasma endothelin concentration in clinically endotoxemic horses Differential induction of MyD88-and TRIF-dependent pathways in equine monocytes by Toll-like receptor agonists Evaluation of mucosal bacteria and histopathology, clinical disease activity and expression of Toll-like receptors in German Shepherd Dogs with chronic enteropathies Evolution and functional divergence of NLRP genes in mammalian reproductive systems Antibodies specific for human or murine Toll-like receptors detect canine leukocytes by flow cytometry Differential expression of Toll-like receptor 4 (TLR4) in healthy and infected canine endometrium Leukemia virus long terminal repeat activates NFkappaB pathway by a TLR3-dependent mechanism Kathrani A, House A, Catchpole B, et al. Polymorphisms in the TLR4 and TLR5 gene are significantly associated with inflammatory bowel disease in German Shepherd Dogs Characterization of ovine TLR7 and TLR8 protein coding regions, detection of mutations and Maedi Visna virus infection Targeting the porcine immune system-particulate vaccines in the 21st century Expression and function of triggering receptor expressed on myeloid cells-1 (TREM-1) on canine neutrophils Yamamoto M, Takeda K. Current views of Toll-like receptor signaling pathways The essential roles of Tolllike receptor signaling pathways in sterile inflammatory diseases Tumor necrosis factor and interleukin-1 serum levels during severe sepsis in humans Tumor necrosis factor and interleukin-1 in the serum of children with severe infectious purpura Serum concentrations of tumor necrosis factor-alpha in neonatal calves with presumed septicemia Endotoxemia and tumor necrosis factor activity in dogs with naturally occurring parvoviral enteritis Vinay DS, Kwon BS. The tumour necrosis factor/TNF receptor superfamily: Therapeutic targets in autoimmune diseases NF-kappaB signals induce the expression of c-FLIP Ligand passing: the 75-kDa tumor necrosis factor (TNF) receptor recruits TNF for signaling by the 55-kDa TNF receptor Tumor necrosis factor-alpha induces cyclooxygenase-2 expression and prostaglandin release in brain microvessel endothelial cells Two distinct monokines, interleukin 1 and tumor necrosis factor, each independently induce biosynthesis and transient expression of the same antigen on the surface of cultured human vascular endothelial cells Tumor necrosis factor and interferon-gamma induce distinct patterns of endothelial activation and associated leukocyte accumulation in skin of Papio anubis Effects of continuous low-dose infusion of lipopolysaccharide on expression of E-selectin and intercellular adhesion molecule-1 messenger RNA and neutrophil accumulation in specific organs in dogs Changes of serum cytokine activities and other parameters in dogs with experimentally induced endotoxic shock Effect of lipopolysaccharide infusion on gene expression of inflammatory cytokines in normal horses in vivo IL-4 stimulates the expression of CXCL-8, E-selectin, VEGF, and inducible nitric oxide synthase mRNA by equine pulmonary artery endothelial cells Tumor necrosis factor and interleukin-6 activity and endotoxin concentration in peritoneal fluid and blood of horses with acute abdominal disease Plasma nitrate/nitrite concentrations in dogs with naturally developing sepsis and non-infectious forms of the systemic inflammatory response syndrome Recombinant tumor necrosis factor induces procoagulant activity in cultured human vascular endothelium: Characterization and comparison with the actions of interleukin 1 Elucidation of Toll-like receptor and adapter protein signaling in vascular dysfunction induced by gram-positive Staphylococcus aureus or gram-negative Escherichia coli Inflammation, endothelium, and coagulation in sepsis Dinarello CA. Proinflammatory and anti-inflammatory cytokines as mediators in the pathogenesis of septic shock Differential cytokine responses in human and mouse lymphatic endothelial cells to cytokines in vitro Analytic review: Interleukin-6 in surgery, trauma, and critical care: Part I: Basic science The proand anti-inflammatory properties of the cytokine interleukin-6 Plasma concentrations of cytokines, their soluble receptors, and antioxidant vitamins can predict the development of multiple organ failure in patients at risk Six at six: Interleukin-6 measured 6 h after the initiation of sepsis predicts mortality over 3 days Interleukin-10: A complex role in the pathogenesis of sepsis syndromes and its potential as an anti-inflammatory drug Interleukin-10 differentially regulates cytokine inhibitor and chemokine release from blood mononuclear cells and fibroblasts Interleukin-10 prevents death in lethal necrotizing pancreatitis in mice Cytokines and metabolic patterns in pediatric patients with critical illness Expression of molecular markers in blood of neonatal foals with sepsis Natural feline coronavirus infection: Differences in cytokine patterns in association with the outcome of infection Macrophage migration inhibitory factor in acute lung injury: Expression, biomarker, and associations Macrophage migration inhibitory factor: A regulator of innate immunity Targeted disruption of migration inhibitory factor gene reveals its critical role in sepsis Macrophage migration inhibitory factor (MIF) and manganese superoxide dismutase (MnSOD) as early predictors for survival in patients with severe sepsis or septic shock Macrophage migration inhibitory factor (MIF) sustains macrophage proinflammatory function by inhibiting p53: Regulatory role in the innate immune response Role of Toll-like receptors 2 and 4, and the receptor for advanced glycation end products in high-mobility group box 1-induced inflammation in vivo HMGB1, a potent proinflammatory cytokine in sepsis Cutting edge: Extracellular high mobility group box-1 protein is a proangiogenic cytokine High-mobility group box 1 as a surrogate prognostic marker in dogs with systemic inflammatory response syndrome HMGB1 as a predictor of organ dysfunction and outcome in patients with severe sepsis The systemic pro-inflammatory response in sepsis Moshage H. Cytokines and the hepatic acute phase response Acute phase reaction and acute phase proteins The systemic reaction during inflammation: The acute-phase proteins Relationship of granulocyte colony stimulating factor with other acute phase reactants in man Acute phase proteins in dogs and cats: Current knowledge and future perspectives Acute phase proteins: Biomarkers of infection and inflammation in veterinary medicine Assay validation and diagnostic applications of major acute-phase protein testing in companion animals Adiponectin and IGF-1 are negative acute phase proteins in a dog model of acute endotoxaemia Procalcitonin and C-reactive protein during systemic inflammatory response syndrome, sepsis and organ dysfunction Diagnostic value of procalcitonin, interleukin-6, and interleukin-8 in critically ill patients admitted with suspected sepsis Canine procalcitonin messenger RNA expression Evaluation of CALC-I gene (CALCA) expression in tissues of dogs with signs of the systemic inflammatory response syndrome Serum C-reactive protein concentration in benign and malignant canine spirocercosis C-reactive protein in canine babesiosis caused by Babesia rossi and its association with outcome Acute phase protein concentrations in dogs with nasal disease The role of acute phase proteins in diagnosis and management of steroidresponsive meningitis arteritis in dogs Concentrations of acute-phase proteins in dogs with steroid responsive meningitis-arteritis C-reactive protein concentration in canine idiopathic polyarthritis Changes in CRP, SAA and haptoglobin produced in response to ovariohysterectomy in healthy bitches and those with pyometra Evaluation of serum haptoglobin and C-reactive protein in dogs with mammary tumors Serum acute phase protein concentrations in dogs with hyperadrenocorticism with and without concurrent inflammatory conditions Relationship among plasma amino acids, C-reactive protein, illness severity, and outcome in critically ill dogs Use of C-reactive protein to predict outcome in dogs with systemic inflammatory response syndrome or sepsis Sepsis, coagulation and anticoagulants The interaction between pathogens and the host coagulation system Leukocyte activation: The link between inflammation and coagulation during heatstroke. A study of patients during the 2003 heat wave in Paris Prospective evaluation of coagulation in critically ill neonatal foals Disseminated intravascular coagulation in cats The role of intravascular coagulation in feline endotoxin shock Systemic response to low-dose endotoxin infusion in cats The role of tissue factor and tissue factor pathway inhibitor in health and disease states Coagulopathy of the critically ill equine patient An updated view of hemostasis: Mechanisms of hemostatic dysfuntion associated with sepsis Cellular sources of tissue factor in endotoxemia and sepsis Evaluation of tissue factor procoagulant activity on the surface of feline leukocytes in response to treatment with lipopolysaccharide and heat-inactivated fetal bovine serum Spontaneous and lipopolysaccharide-induced expression of procoagulant activity by equine lung macrophages in comparison with blood monocytes and blood neutrophils Platelets enhance endotoxin-induced monocyte tissue factor (TF) activity in the horse Clinical relevance of monocyte procoagulant activity in horses with colic Induction of microparticle-and cell-associated intravascular tissue factor in human endotoxemia Levels of endothelial and platelet microparticles and their interactions with leukocytes negatively correlate with organ dysfunction and predict mortality in severe sepsis Bench-to-bedside review: Circulating microparticles-a new player in sepsis? The role of endothelial interleukin-8/NADPH oxidase 1 axis in sepsis Role of tissue factor and protease-activated receptors in a mouse model of endotoxemia Effects of small-and large-volume resuscitation on coagulation and electrolytes during experimental endotoxemia in anesthetized horses Efficacy and safety of recombinant human activated protein C for severe sepsis Endotoxemia and sepsis mortality reduction by non-anticoagulant activated protein C Surviving Sepsis Campaign guidelines for management of severe sepsis and septic shock Surviving Sepsis Campaign: International guidelines for management of severe sepsis and septic shock Pharmacological profile of recombinant, human activated protein C (LY203638) in a canine model of coronary artery thrombosis Human protein C induces anticoagulation and increased fibrinolytic activity in the cat Contact system activation in severe infectious diseases The dual role of the contact system in bacterial infectious disease Contact system. New concepts on activation mechanisms and bioregulatory functions Activation of the plasma kallikrein/kinin system on endothelial cell membranes Cleavage of ultra-large von Willebrand factor by ADAMTS-13 under flow conditions The balance between von Willebrand factor and its cleaving protease ADAMTS13: Biomarker in systemic inflammation and development of organ failure? AD-AMTS-13, von Willebrand factor and related parameters in severe sepsis and septic shock Decreased ADAM-TS-13 (A disintegrin-like and metalloprotease with thrombospondin type 1 repeats) is associated with a poor prognosis in sepsisinduced organ failure Severe secondary deficiency of von Willebrand factor-cleaving protease (ADAM-TS13) in patients with sepsis-induced disseminated intravascular coagulation: Its correlation with development of renal failure Circulating cytokine/inhibitor profiles reshape the understanding of the SIRS/CARS continuum in sepsis and predict mortality Early postoperative compensatory anti-inflammatory response syndrome is associated with septic complications after major surgical trauma in patients with cancer Procalcitonin, interleukin 6 and systemic inflammatory response syndrome (SIRS): Early markers of postoperative sepsis after major surgery Platelets, endothelium and shear join forces to mislead neutrophils in sepsis Sepsis induces early alterations in innate immunity that impact mortality to secondary infection Rapid up-regulation of IRAK-M expression following a second endotoxin challenge in human monocytes and in monocytes isolated from septic patients ABIN-3 is a novel lipopolysaccharide-inducible inhibitor of NF-kappaB activation Impaired antigen presentation by human monocytes during endotoxin tolerance Monocyte deactivation in septic patients: Restoration by IFN-gamma treatment Coincidence of pro-and anti-inflammatory responses in the early phase of severe sepsis: Longitudinal study of mononuclear histocompatibility leukocyte antigen-DR expression, procalcitonin, Creactive protein, and changes in T-cell subsets in septic and postoperative patients Leukocyte apoptosis and its significance in sepsis and shock When apoptosis meets autophagy: Deciding cell fate after trauma and sepsis Cell death during sepsis: Integration of disintegration in the inflammatory response to overwhelming infection Autophagy in immunity and inflammation Conversation between apoptosis and autophagy Sepsis induces extensive autophagic vacuolization in hepatocytes: A clinical and laboratory-based study Pyroptosis A cell death modality of its kind? Apoptosis, pyroptosis, and necrosis: mechanistic description of dead and dying eukaryotic cells The caspase-1 digestome identifies the glycolysis pathway as a target during infection and septic shock Apoptosis in the lymphatic organs of piglets inoculated with classical swine fever virus Glucocorticoids potently block tumour necrosis factor-alpha-and lipopolysaccharide-induced apoptotic cell death in bovine glomerular endothelial cells upstream of caspase 3 activation Caspase activation during Haemophilus somnus lipooligosaccharide-mediated apoptosis of bovine endothelial cells Endothelial cell apoptosis induced by bacteria-activated platelets requires caspase-8 and -9 and generation of reactive oxygen species Roles of MMP-2/-9 in cardiac dysfunction during early multiple organ failure in an ovine animal model Endotoxin-induced ileal mucosal injury and nitric oxide dysregulation are temporally dissociated Toward a detailed characterization of feline immunodeficiency virus-specific T cell immune responses and mediated immune disorders Fluid replacement with hypertonic or isotonic solutions guided by mixed venous oxygen saturation in experimental hypodynamic sepsis Nikolich-Zugich J, Slifka MK, Messaoudi I. The many important facets of T-cell repertoire diversity Parish IA, Heath WR. Too dangerous to ignore: Self-tolerance and the control of ignorant autoreactive T cells Tolerance: An overview and perspectives Mechanisms maintaining peripheral tolerance Human FoxP3 + CD4 + regulatory T cells: Their knowns and unknowns All creatures great and small: Regulatory T cells in mice, humans, dogs and other domestic animal species Transforming growth factor-beta/transforming growth factor-betaRII signaling may regulate CD4 + CD25 + T-regulatory cell homeostasis and suppressor function in feline AIDS lentivirus infection In vivo depletion of CD4 + CD25 + regulatory T cells in cats Feline immunodeficiency virus infection phenotypically and functionally activates immunosuppressive CD4 + CD25 + T regulatory cells CD4 + CD25 + T regulatory cells from FIV + cats induce a unique anergic profile in CD8 + lymphocyte targets CD4 + CD25 + T regulatory cells inhibit CD8 + IFN-gamma production during acute and chronic FIV infection utilizing a membrane TGF-betadependent mechanism T-regulatory cells infected with feline immunodeficiency virus up-regulate programmed death-1 (PD-1) Use of FoxP3 expression to identify regulatory T cells in healthy dogs and dogs with cancer Quantitation of canine regulatory T cell populations, serum interleukin-10 and allergen-specific IgE concentrations in healthy control dogs and canine atopic dermatitis patients receiving allergen-specific immunotherapy Changes in regulatory T cells in dogs with cancer and associations with tumor type Increase of regulatory T cells in the peripheral blood of dogs with metastatic tumors Crossreactivity of antibodies to canine CD25 and Foxp3 and identification of canine CD4 + CD25 + Foxp3 + cells in canine peripheral blood A novel monoclonal antibody specific for canine CD25 (P4A10): Selection and evaluation of canine Tregs Cloning and expression of canine CD25 for validation of an anti-human CD25 antibody to compare T regulatory lymphocytes in healthy dogs and dogs with osteosarcoma Relationship between regulatory and type 1 T cells in dogs with oral malignant melanoma Low-dose cyclophosphamide selectively decreases regulatory T cells and inhibits angiogenesis in dogs with soft tissue sarcoma Detection of Foxp3 protein expression in porcine T lymphocytes Phenotypic and functional characterisation of porcine CD4 + CD25 high regulatory T cells Molecular characterisation of porcine Forkhead-box p3 (Foxp3) Induction of inducible CD4 + CD25 + Foxp3 + regulatory T lymphocytes by porcine reproductive and respiratory syndrome virus (PRRSV) Localized Th1-, Th2-, T regulatory cell-, and inflammation-associated hepatic and pulmonary immune responses in Ascaris suum-infected swine are increased by retinoic acid Induction of porcine regulatory cells by mycophenolic acid-treated dendritic cells Current knowledge on porcine regulatory T cells Porcine regulatory T cells: Mechanisms and T-cell targets of suppression Sensitive detection of Foxp3 expression in bovine lymphocytes by flow cytometry Subpopulations of bovine WC1 + gammadelta T cells rather than CD4 + CD25 high Foxp3 + T cells act as immune regulatory cells ex vivo Development of monoclonal antibodies to detect bovine FOXP3 in PBMCs exposed to a staphylococcal superantigen Antigen-specific regulatory T cells in bovine paratuberculosis Infestation of sheep with Psoroptes ovis, the sheep scab mite, results in recruitment of Foxp3(+) T cells into the dermis Identification of CD4(+)CD25(high) Foxp3(+) T cells in ovine peripheral blood Characterization of monoclonal antibodies to equine interleukin-10 and detection of T regulatory 1 cells in horses Increased IL-4 and decreased regulatory cytokine production following relocation of Icelandic horses from a high to low endoparasite environment CD4 + CD25 + T cells expressing FoxP3 in Icelandic horses affected with insect bite hypersensitivity Interferon-gamma, interleukin-4 and interleukin-10 production by T helper cells reveals intact Th1 and regulatory TR1 cell activation and a delay of the Th2 cell response in equine neonates and foals Regulatory T cells: The suppressor arm of the immune system Regulatory T cells: How do they suppress immune responses? The development and function of regulatory T cells FOXP3 modifies the phenotypic and functional properties of regulatory T cells Making sense of regulatory T cell suppressive function Regulatory T cells selectively express Toll-like receptors and are activated by lipopolysaccharide Regulatory T cells and Toll-like receptors: What is the missing link? Toll-like receptor 2 signaling modulates the functions of CD4 + CD25 + regulatory T cells Modulation of regulatory T cells in health and disease: Role of toll-like receptors Marked elevation of human circulating CD4 + CD25 + regulatory T cells in sepsisinduced immunoparalysis Sepsis is characterized by the increases in percentages of circulating CD4 + CD25 + regulatory T cells and plasma levels of soluble CD25 Increased circulating regulatory T cells (CD4 + CD25 + CD127 -) contribute to lymphocyte anergy in septic shock patients Increased percentage of CD4 + CD25 + regulatory T cells during septic shock is due to the decrease of CD4 + CD25-lymphocytes Adoptive transfer of in vitro-stimulated CD4 + CD25 + regulatory T cells increases bacterial clearance and improves survival in polymicrobial sepsis The contribution of CD4 + CD25 + T-regulatory-cells to immune suppression in sepsis Foxp3 + regulatory T cells induce alternative activation of human monocytes/macrophages Human CD4 + CD25 + regulatory T lymphocytes inhibit lipopolysaccharide-induced monocyte survival through a Fas/Fas ligand-dependent mechanism Increased T regulatory cells and decreased Th1 pro-inflammatory cytokines correlate with culture-positive infection in febrile neutropenia childhood oncology patients Association between regulatory T cell activity and sepsis and outcome of severely burned patients: a prospective, observational study The relationship between CD4 + CD25 + CD127-regulatory T cells and inflammatory response and outcome during shock states Toll-like receptor signalling on Tregs: To suppress or not to suppress? Regulatory T cells overturned: The effectors fight back Sepsis: Rethinking the approach to clinical research Pharmacological treatment of sepsis Adjunctive therapies for severe sepsis The authors acknowledge funding from the Biotechnology and Biological Sciences Research Council, Novartis Animal Health, the European College of Veterinary Internal Medicine, and the American College of Veterinary Emergency and Critical Care for work on canine regulatory T cells and sepsis.