• No results found

University of Groningen Functional genomics approach to understanding sepsis heterogeneity Le, Kieu

N/A
N/A
Protected

Academic year: 2021

Share "University of Groningen Functional genomics approach to understanding sepsis heterogeneity Le, Kieu"

Copied!
19
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

University of Groningen

Functional genomics approach to understanding sepsis heterogeneity

Le, Kieu

DOI:

10.33612/diss.98318779

IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite from

it. Please check the document version below.

Document Version

Publisher's PDF, also known as Version of record

Publication date:

2019

Link to publication in University of Groningen/UMCG research database

Citation for published version (APA):

Le, K. (2019). Functional genomics approach to understanding sepsis heterogeneity. University of

Groningen. https://doi.org/10.33612/diss.98318779

Copyright

Other than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons).

Take-down policy

If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim.

Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons the number of authors shown on this cover page is limited to 10 maximum.

(2)

CHAPTER

01

Why do organs fail differently

in patients with sepsis ?

Endothelial heterogeneity

and organ-specific failure

phenotypes in sepsis

(3)

Kieu T. T Le1* & Dayang Erna-Zulaikha2*, Grietje Molema2, Vinod Kumar1,3

Jan G. Zijlstra4, Jill Moser2,4‡

1 Department of Genetics, University of Groningen, University Medical Center Groningen,

Groningen, The Netherlands

2 Department of Pathology and Medical Biology, Medical Biology section,

University of Groningen, University Medical Center Groningen, Groningen, The Netherlands

3 Department of Internal Medicine and Radboud Center for Infectious Diseases (RCI),

Radboudumc, 6525 GA Nijmegen, The Netherlands

4 Department of Critical Care, University of Groningen, University Medical Center Groningen,

Groningen, The Netherlands *Shared First authorship

Corresponding author:

Jill Moser PhD E-mail: j.moser@umcg.nl

Keywords:

Sepsis, Endothelial cells (EC), endothelial heterogeneity, Acute Kidney Injury (AKI), Acute Respiratory Dysfunction Syndrome (ARDS), endothelial permeability, endothelial activation, inflammation.

(4)

Sepsis, a dysregulated host response due to infection is associated with high morbidity and mortality which is a result of multiple organ dysfunction syndrome (MODS). There are currently no therapeutic treatments available for patients with sepsis which is partly attributed to our incomplete understanding of why and how organs fail in patients with sepsis. Clinical manifestations of organ failure differ from one organ to another despite a systemic pathogenic insult. The lungs become drowned in fluids with extensive neutrophil sequestration, while the kidney appears to shut down with minimal morphological changes and limited leukocyte infiltration. For years, a dysregulated immune system was thought to be the central mediator of organ failure. However, the endothelium is gaining increasing attention due to its crucial role in maintaining organ homeostasis, vascular barrier, coagulation and inflammation. All of these processes are disrupted in the failing organs of sepsis patients, yet this occurs heterogeneously within and between organs.

In this review, we summarize recent advances in our understanding of heterogeneous endothelial responses which may mediate organ-specific failure phenotypes in patients with sepsis. We have focussed on sepsis-associated acute kidney injury (sepsis AKI) and acute respiratory dysfunction syndrome (ARDS) and consider differences and similarities of endothelial behaviour between the kidney and lung and how this information might be useful for guiding microvascular targeted therapy for patients with sepsis and multiple organ failure.

Background

Sepsis is a life-threatening condition characterized by an overwhelming, heterogeneous and progressive host dysregulation as a result of infection (Singer

et al., 2016). Despite recent advances in

early sepsis recognition and prompt antibiotic administration protocols, sepsis continues to be the leading cause of death among critically ill patients in intensive care units (ICU) worldwide (Mayr, Yende and Angus, 2014). The fatal outcome of patients with sepsis is due to multiple organ dysfunction syndrome (MODS), with mortality being directly associated with the number of failing organ systems and the degree of organ damage (Angus and van der Poll, 2013).

The pathophysiological mechanisms driving organ injury and failure in patients with sepsis is poorly understood. For a long time, decreased oxygen delivery and a dysregulated immune system were thought to be the main drivers of organ failure. However, three decades of clinical trials targeting these defective responses have failed to improve patient outcome (Seeley and Bernard, 2016). The complexity of sepsis, the heterogeneity between patients and the relatively late diagnosis underlie poor outcome and the failure of multiple one-size-fits-all clinical trials (Marshall, 2014). Current therapeutic options are still limited to antibiotic administration and supporting organ function in the ICU. Despite years of research, we still do not know exactly why and how organs fail in patients with sepsis. Likewise, why some patients develop MODS while others do not. However, predisposing factors that increase sepsis susceptibility such as age, having comorbid disease, immunodeficiency and genetic factors all likely contribute.

Intriguingly, the clinical manifestations of sepsis-associated organ dysfunction differ from one organ to another despite a similar systemic pathogenic insult (Table. 1). Lungs become drowned in fluids with massive neutrophil sequestration primarily due to ‘leaky’ microvasculature (Hendrickson and Matthay, 2018), while the kidney appears to shut down

(5)

with minimal morphological alterations and limited neutrophil infiltration (Lerolle et al., 2010; Takasu et al., 2013; Aslan et al., 2018). In other words, each organ responds differently to the same systemic insult. These observations have led us to question what regulates organ- specific responses in sepsis. Is there a common cell-type and/or response which dictates the way an organ fails in sepsis? If there is, how does that differ between organs? And more importantly, how will these organ-specific responses influence treatment options for patients with multiple organ failure?

When patients with sepsis enter the ICU, multiple cellular cascades encompassing inflammation, coagulation and metabolic changes run in parallel in various cells within different organs leading to impaired function (Angus and van der Poll, 2013). One cell type which is common to all organs is the endothelium, which is known to play an important role in maintaining organ homeostasis (Aird, 2007). While aberrant endothelial responses have been shown to predict mortality in critically ill patients (Duffy et al., 2011), endothelial cells (ECs) phenotypically differ between organs within vascular compartments of the same organ, and even between adjacent ECs within the same vessels (Dayang et al., 2019). Different pathogens and/or inflammatory stimuli can initiate endothelial inflammatory activation, leukocyte infiltration, endothelial permeability and coagulation yet this occurs heterogeneously within and between organs. Hence, what is the contribution of the inter- and intra-organ endothelial heterogeneous response to sepsis on the specific way organs fail in patients with multiple organ failure?

The aim of this review was to summarize recent advances in our understanding of inter-organ heterogeneous endothelial responses that may mediate organ-specific failure phenotypes in patients with sepsis.

Implications for patient care and potential directions for future treatment strategies will also be discussed.

The endothelium as a

heterogeneous entity of cells with

specialized functions

1. Endothelial phenotypes

The macrovasculature and organ microvasculature are lined by cells that tightly interact with each other to form a semi-permeable barrier called the endothelium. The endothelial lining maintains organ homeostasis by regulating various functions including the trafficking of fluid, solutes, hormones and macromolecules (Mehta and Malik, 2006). It is now well established that EC phenotypes vary between different organs, between different microvascular compartments within the same organ, and between neighbouring ECs of the same blood vessel type (Aird, 2007). Yet, how each organ determines the functional properties of its endothelium is currently not well understood. Endothelial properties are to a certain extent dependent on the surrounding tissue microenvironments, such as flow, and interaction with adjacent specialized cells (i.e. pericytes, podocytes, epithelial cells etc). Inflammatory cues and signals from parenchyma cells can induce posttranscriptional modifications since they are in direct contact with ECs. Most, but not, all organ-specific EC gene signatures are rapidly lost when they are removed from their in vivo microenvironment and grown in culture (Lacorre et al., 2004; Burridge and Friedman, 2010). However, certain heterogeneous gene expression profiles have also been observed in cultured organ-specific ECs suggesting that epigenetic control may mediate at least certain properties of ECs (Chi et al., 2003; Marcu et al., 2018).

(6)

2. Glycocalyx

The endothelial barrier is comprised of inter-endothelial junctions and various non- cellular components, such as the glycocalyx and extracellular matrix (Weinbaum, Tarbell and Damiano, 2007). A complex network of macromolecules including proteoglycans and sialoproteins forms the 1 to 3 μm glycocalyx layer on the apical side of endothelium (Van den Berg, Vink and Spaan, 2003; Weinbaum, Tarbell and Damiano, 2007). Endothelial glycocalyx make-up in various organs are reported to be organ-specific and vascular bed specific (Schmidt et al., 2012). Pulmonary artery ECs were coated with α-galactose carbohydrates, while pulmonary microvascular ECs found to be enriched with α- and β-N- acetylgalactosamine carbohydrates (King et al., 2004). The implication of the glycocalyx composition differences on endothelial barrier susceptibility to sepsis is not known. The extent of which inflammatory mediators and leukocytes might disrupt the glycocalyx may partly drive organ failure such as acute kidney injury (AKI) or acute respiratory distress (ARDS). The exact mechanisms regulating glycocalyx degradation are not yet fully understood but may be organ and vascular bed-specific (Uchimido, Schmidt and Shapiro, 2019). Significant experimental evidence shows that damage of the glycocalyx in sepsis compromises the endothelium by promoting permeability (Colbert and Schmidt, 2016). Additionally, excessive fluid resuscitation in patients with sepsis is thought to degrade the endothelial glycocalyx (Chappell et al., 2014). Nevertheless, therapeutic strategies that aimed to preserve glycocalyx integrity have failed to improve the outcome of patients with sepsis (Chelazzi et al., 2015).

3. Barrier integrity

The cleft between ECs are sealed by tight- and adherens- junctions, which are both functionally coupled to the cell-matrix complex

(Wallez and Huber, 2008). Mechanical strength and stability of the barrier is provided by adherens junctions (Dejana and Orsenigo, 2013)(Dejana, Bazzoni and Lampugnani, 1999). Additionally, tight junctions such as Claudin- 5 are only expressed in peripheral endothelium (Komarova and Malik, 2010). Gap junctions, containing connexins (i.e. Cx43, Cx40, and Cx37) mainly function to regulate cell-to-cell communication (Bazzoni and Dejana, 2004). The Angiopoietin/Tie2 receptor system is also involved in the maintenance of barrier function. Angiopoietin 1 (Ang1) and Angiopoietin 2 (Ang2) ligands bind to the Tie2 receptor (Matijs van Meurs, Kümpers,

et al., 2009). In quiescent conditions, Ang1

is released by pericytes and binds to the Tie2 receptor ensuring barrier maintenance (Davis et al., 2007). However, during inflammation, Ang2 is produced by ECs and competitively binds to the Tie2 receptor disrupting the vascular barrier (Leligdowicz

et al., 2018). Disruption of the various systems

controlling vascular barrier integrity will lead to vascular leakage resulting in oedema, and other detrimental consequences. Additionally, recent studies found that the expression of endothelial junction molecules needed to maintain endothelial integrity, VE-cadherin, Occludin and Claudin-5 were distinct in various mouse organs (Aslan et al., 2017). Hence, endothelial phenotypes and junction molecule composition varies between different microvascular beds in different organ systems which can have a major impact on the dynamics and regulation of vascular permeability.

4. Coagulation

ECs also maintain the fluidity of the blood by governing factors that are involved in coagulation and fibrinolysis, such as tissue factors (TFs), Endothelial protein C receptor (EPCR), plasminogen activator inhibitor Type 1 (PAI-1) and von Willebrand factor

(7)

(vWF) (Aird, 2007). Endothelial coagulation factors are stored in endothelial storage granules known as Weibel-Palade bodies until needed (Levi, Van Der Poll and Schultz, 2012). vWF is heterogeneously distributed, with higher expression noted in larger vessels such as veins, compared to the capillaries (Pusztaszeri, Seelentag and Bosman, 2006; Yuan et al., 2016). Moreover, vWF is mosaically expressed within individual vessels (Yuan et

al., 2016). Although it is now established that

coagulation plays an important role in sepsis-induced organ dysfunction, it is not completely clear why specific organs such as the kidney and lung are more susceptible to developing (micro)vascular thrombosis than other organs such as the heart or liver (Levi, Van Der Poll and Schultz, 2012). Inflammation does not only lead to activation of coagulation, but coagulant activation can also result in inflammation. Pro- inflammatory cytokines activate the coagulant system and concomitantly downregulate anticoagulant pathways. A detailed overview of organ-specific responses of the coagulant system has been reviewed in detail elsewhere (Levi, Van Der Poll and Schultz, 2012). Incorporation of different extracellular signals, cell responses and EC signaling mechanisms in different vascular bed regions may explain why the pro- or anti-coagulant response of EC may differ between organs (Rosenberg and Aird, 1999).

5. Endothelial activation

The EC is well equipped for recognising and responding to hostile or toxic components in the blood stream. In sepsis, ECs recognize damage-associated molecular pattern (DAMPs) and pattern associated molecular pattern (PAMPs) by pattern recognition receptors. LPS is a major component of the Gram-negative bacteria cell wall found in the blood of patients with sepsis (Dauphinee and Karsan, 2006). For a long time, the molecular mechanisms controlling LPS-mediated

activation of ECs was thought to be controlled via similar mechanisms as those described for immune cell activation (Dauphinee and Karsan, 2006). However, in contrast to immune cell activation, recent studies have identified RIG-I as an additional receptor controlling LPS-mediated endothelial activation together with the well- known TLR4 receptor (Moser et

al., 2016; Yan et al., 2017).

During inflammation, the endothelium becomes activated and expresses adhesion molecules, such as E-selectin, VCAM-1, ICAM-1 which facilitates leukocyte rolling, adhesion, arrest, and transmigration (Ley et

al., 2007; Nourshargh, Hordijk and Sixt, 2010).

Initial rolling of leukocytes is also characterized by the binding of VCAM-1 and ICAM-1 to cognate leukocyte integrin, such as Very late antigen-4 (VLA-4), which results in leukocyte arrest and adhesion (Berlin et al., 1995). Once the EC-leukocyte interaction is robust, it can withstand circulatory shear stress, and undergo paracellular or transcellular transmigration (Phillipson et al., 2009). In the context of sepsis, most studies have focused on understanding endothelial responses to endotoxin such as LPS or proinflammatory cytokines such as TNFα, both found in the plasma of patients with sepsis. Previous studies have assumed that LPS and TNFα signalling downstream of their receptors converged into similar pathways. However, an increasing number of studies provide evidence that this may not be the case (Wang et al., 2014; Moser et al., 2016). Additionally, distinct EC subpopulations of HUVEC in vitro were identified each having distinct inflammatory phenotypes that were controlled by different regulatory signalling mechanisms (Dayang

et al., 2019). This appeared to be a common

response of ECs since similar findings were observed using ECs isolated from the lung (Dayang et al., 2019). A subpopulation of cells which remained quiescent despite exposure to LPS was identified (Dayang et al., 2019). Why

(8)

these cells remain quiescent in the presence of LPS or an inflammatory stimulus such as TNFα is currently unknown. Nevertheless, understanding the molecular mechanisms controlling this quiescent phenotype may be exploited in order to identify and generate therapeutic strategies to inhibit endothelial activation in the setting of sepsis.

The endothelium can be exposed to different types of infectious pathogens bacteria, fungi or viruses. The impact of these pathogens on endothelial activation has not been extensively studied. However, recent work from our laboratory found that some pathogens such as S. pneumoniae and C. albicans could not induce endothelial activation directly (Kieu Le, unpublished observations). Instead, these pathogens likely interact with immune cells initiating the release of various proinflammatory mediators which subsequently activates the endothelium. It remains to be elucidated whether this also happens in vivo. Interestingly, EC responses initiated by mediators released from PBMCs stimulated with different pathogens were not dependant on the type of pathogen (Kieu Le, unpublished observations). Hence, humoral factors released by PBMCs stimulated with different pathogens induce similar endothelial responses.

Organ-specific failure phenotypes in

sepsis and the role of the endothelium

A role for aberrant endothelial behaviour and leukocyte influx promoting organ failure in sepsis is well established (Singbartl and Ley, 2004), yet the precise mechanisms and organ-specific responses poorly understood. In order to understand the pathophysiology of organ dysfunction in sepsis, animal models and/or cell culture models are extensively used. As described above in vivo observations show that the extent endothelial inflammatory activation differs between the microvasculature of various organs. Likewise,

leukocyte recruitment is organ- and stimulus-specific (Liu and Kubes, 2003; Devi et al., 2013; Rossaint and Zarbock, 2013). Each organ has specialized capillaries by which EC-leukocyte interactions can take place. Both the kidney and lungs are particularly vulnerable in patients with sepsis which often leads to Acute Kidney Injury (AKI) and Acute Respiratory Distress Syndrome (ARDS) respectively. Examination of post-mortem organs from patients with sepsis and multiple organ failure revealed some sequestration and aggregation of neutrophils in renal microvascular compartments (Nuytinck

et al., 1988; Thijs and Thijs, 1998; Brealey

and Singer, 2000; Aslan et al., 2018) yet major infiltration of neutrophils was found within the pulmonary microvasculature of patients (Brealey and Singer, 2000). However, other organs and cellular systems are also compromised, with different endothelial-regulated processes being affected (Table. 1). The microvasculature plays an important role in mediating the failure of all organs yet how these endothelial responses are mediated clearly differs per organ. The role of the endothelium in mediating kidney and lung failure phenotypes will be discussed further in more detail.

(9)

Organ Clinical features Pathophysiology Available treatment Kidney Proteinuria ↑ blood Creatinine ↓ Urine Output ↓ Glomerular filtration rate (GFR)

Disturbed microvascular integrity Epithelial cell injury/cycle arrest/apoptosis/ necrosis, ↓ metabolism due to excessive inflammation.

Capillary leukocyte infiltration

Renal Replacement Therapy (RRT) Lung Impaired oxygenation Pulmonary oedema ↓ compliance Fibrosis Microvascular hyperpermeability Acute inflammatory infiltrates and neutrophil accumulation

Disrupted lung alveolar-capillary barrier Perivascular oedema

Fibrosis

Mechanical ventilation with low tidal volume and positive end-expiratory pressure (PEEP) Liver Jaundice Cholestasis Hypoxic hepatitis

Impaired bile and bile acid transport Adhesion of neutrophils to SECs Thrombi formation & microvascular hypoperfusion

↑ cytokines & Fibrin deposition

Hepatocyte injury, Endotoxin and bacteria spill over None Heart Myocardial depression Severe biventricular dysfunction Arrhythmia Abnormal calcium homeostasis-cardiomyocyte injury Focal mitochondrial injury Defects in cardiomyocyte

coupling due to abnormal cardiac gap junctions

Impaired coronary microcirculation Endothelial activation Inotropic agents Betablockers Brain /CNS Confusion / Delirium Impaired cognition Amnesia

Impaired endothelial integrity (microvasculature & BBB), Hyperinflammation leading to microglial activation Light sedation Early rehabilitation Cardiovascular Ventricular dilatation ↓ ejection fraction ↓ contractility Vasoplegia Myocardial depression Impaired calcium homeostasis,

↑ NO & phosphate production Endothelial inflammatory activation

Inotropic agents Betablockers

(10)

Sepsis-associated Acute Kidney Injury

(Sepsis-AKI).

Renal endothelial activation,

inflammation and leukocyte infiltration

Approximately 50% of patients with sepsis admitted to the intensive care develop AKI which is associated with high mortality (Poston and Koyner, 2019). Recent advances in our understanding of AKI have identified that aberrant renal microvascular responses play a major role in driving renal failure in patients with sepsis (Kellum and Prowle, 2018). The kidney microvasculature composed of arterioles, venules, and two capillary networks, namely the glomerulus and the peritubular capillary plexus (Ince et al., 2016) regulate blood flow within the kidney and mediate inflammation, permeability and coagulation. Glomerular ECs are highly fenestrated and covered by a rich glycocalyx layer that allow filtration of water and small solutes as well as supporting podocyte structure (Singh et al., 2007;

Haraldsson, Nystrom and Deen, 2008). The peritubular capillaries are also fenestrated and contribute to tubular reabsorption as well as supporting renal tubule cell function (Jen, Haragsim and Laszik, 2011).

In sepsis, the renal microvascular endothelial compartments produce adhesion molecules from the selectin family (P-selectin and E-selectin), and integrin family (VCAM-1 and ICAM-(VCAM-1) mediating leukocytes rolling, tethering, adherence and subsequent infiltration into the tissues (Nourshargh and Alon, 2014). However, an association between leukocyte adherence and the distinct adhesion molecule expression profiles in specific renal microvascular compartments is difficult to establish (Matijs van Meurs, Kurniati, et al., 2009; Molema, 2010; Asgeirsdottir et al., 2012). Although not extensive, leukocytes, predominantly neutrophils, guided by endothelial activation, were found to localize predominantly in the glomerular and

Gastrointestinal

Mucosal bleeding Paralytic ileus

Endothelial cell injury and microcirculation dysfunction Epithelial hyperpermeability Altered microbiome Proton pump inhibitor Early enteral nutrition Probiotics Selective digestive decontamination Coagulation Bleeding Microthrombi Tissue ischemia Diffuse intravascular coagulation (DIC) Intravascular coagulation Endothelial injury

Systemic thrombin generation

Antithrombin Thrombomodulin Concentrated platelets Immune system Hyperinflammation Immunosuppression Secondary infection Virus reactivation

Balance between hyper- and

(11)

peritubular capillaries (Kuligowski, Kitching and Hickey, 2006; Kitching, Holdsworth and Hickey, 2008; Aslan et al., 2018).

In mice, LPS and TNF-α induced expression of E-selectin was predominantly found in the glomeruli and to a lesser extent in the arterioles (Dayang et al., 2019). In contrast, VCAM-1 is highly expressed in the arterioles and venules, while expressed to a lesser extent in the glomeruli (M. van Meurs et

al., 2009; Asgeirsdottir et al., 2012; Dayang et al., 2019). E- selectin was predominantly found

expressed in the glomeruli and to a lesser extent in the arterioles upon LPS challenge (Dayang

et al., 2019). In contrast, VCAM-1 was highly

expressed in the arterioles and venules, while expressed to a lesser extent in the glomeruli (M. van Meurs et al., 2009; Asgeirsdottir et

al., 2012; Dayang et al., 2019) Despite high

VCAM-1 expression in the arterioles during sepsis, comparable to levels in the venules, leukocyte adherence on the surface of renal arterioles rarely occurs. In a mouse model of experimental sepsis, cecal ligation and puncture (CLP), E-selectin and P-selectin expression was found in the glomeruli and peritubular capillaries. ICAM-1 expression was observed in the peritubular capillaries but not in the glomeruli (Herter et al., 2014). Recent work from our laboratory revealed that the inflammatory response of neighbouring ECs within a specific renal microvascular compartment were distinct (Dayang et al., 2019). The molecular mechanisms controlling renal heterogeneous inter- and intra-vascular endothelial activation in vivo are currently poorly understood but may involve regulatory non-coding RNAs, such as miR-126 in the regulation of VCAM-1 in the renal vasculature (Asgeirsdottir et al., 2012).

Blocking E-selectin, P-selectin or both in septic mice significantly diminished neutrophil infiltration into the kidney and preserved kidney morphology (Herter et al., 2014). Moreover, neutrophil depletion was

found to protect CLP-induced renal injury in mice (Herter et al., 2014), highlighting the significance of endothelial activation and subsequent neutrophil infiltration in relation to sepsis-associated renal failure. Similar findings have been observed in human post-mortem kidney biopsies from sepsis patients with acute kidney injury (AKI). Monocytes were found to aggregate in the glomerular capillaries, cortical and medullary capillaries which was not observed in critically ill trauma patients with systemic inflammation (Lerolle et

al., 2010; Aslan et al., 2018).

Activated ECs also produce cytokines and chemokines that can initiate and orchestrate inflammation. In a rat model of acute renal microvascular injury, in situ hybridization of renal tissue revealed co-localization of endothelially produced IP-10/ CXCL10 mRNA with infiltrating T cells in the tubulointerstitial compartment (Panzer et al., 2006). In the same kidney, glomerular IP10/ CXCL10 expression was absent, while MCP-1/CCL2 was detectable and associated with monocyte infiltration (Panzer et al., 2006). In CLP mice, monocytes were found to undergo CX3CR1-dependent crawling 6 hours after CLP induction yet no monocytes were found to transmigrate into the parenchyma. Interestingly, monocyte inhibition exacerbated CLP-induced renal injury suggesting that monocytes may mediate anti-inflammatory effects (Chousterman et al., 2016). The concept that monocytes may play an anti-inflammatory role in the kidney was supported by Hato and colleagues, which showed macrophages elicited a renal protective effect by low-dose LPS preconditioning (Hato et al., 2015). Organ-specific inflammatory models recently revealed that leukocyte activation and recruitment differs between the kidney and lung and that endothelial activation is differentially regulated in these organs, which contradicts assumptions that leukocyte recruitment into organs is regulated similarly

(12)

for each organ. Recently reviewed in detail elsewhere (Maas, Soehnlein and Viola, 2018).

Together these studies imply that 1) the interaction of distinct immune cell subsets with the endothelium is characterized by EC heterogeneous expression of adhesion molecules and specific chemokine signals within different renal microvascular compartments, and 2) in certain circumstances, leukocyte-endothelial interactions do not necessarily promote inflammation causing functional and structural failure but that their role in the resolution of inflammation may be as important.

Renal vascular barrier dysfunction

As mentioned above, minimal histological changes were observed in renal biopsies from septic patients, likewise in the kidneys from experimental sepsis mice (Maiden et al., 2016), which corroborates the observation that the microvascular barrier in the kidney of sepsis patients is uncompromised (Aslan

et al., 2017). Loss of the glycocalyx leads

to functional impairment of the glomerular barrier in sepsis (Ince et al., 2016). In CLP-induced rats, an increase in urine albumin was correlated with a decrease in glycocalyx components, syndecan-1, hyaluronic acid, and sialic acid (Adembri et al., 2011). Likewise, loss of the glycocalyx was observed in LPS and TNFα-challenged mice (Xu et al., 2014). TNF-/- mice were found to be resistant to LPS-induced glomerular permeability and were not able to induce heparanase expression in response to LPS injection therefore protecting the barrier from heparan sulphate degradation (Xu et al., 2014), a common glycocalyx degradation pathway (Schmidt et al., 2012). The interaction of TNF-α and syndecans leads to structural rearrangement of ECs that increases paracellular permeability, allowing extravasation of macromolecules, such as albumin (Christaki and Opal, 2008).

In normal healthy mice, the expression level of tight junction molecules Occludin and Claudin-5 were 20-fold and 100-fold higher in the lungs compared to kidney, respectively (Aslan et al., 2017), which implies that a different endothelial barrier regulation exists in different organs under quiescent conditions. Moreover, Occludin and Claudin-5 mRNA induction was greatly increased in the kidney, and to a lesser extent in the lungs of the same LPS-challenged mice. Additionally, expression of adherence junction molecule, VE-cadherin was unaltered in the kidney but was increased in the lungs of LPS challenged mice (Aslan et

al., 2017). VE-Cadherin plasma protein levels

are associated with AKI in septic patients (Yu

et al., 2019).

The Ang/Tie2 system also plays an important role in maintenance of vascular integrity. In human sepsis renal biopsies, Renal Ang1 mRNA levels were lower in human sepsis patient biopsies compared to control patients therefore disturbing the Ang2/ Ang1 ratio (Aslan et al., 2014). Loss of plasma Ang1 in patients with sepsis is associated with poor outcome (Fang et al., 2018). Moreover, plasma Ang1 levels and the Ang2/Ang1 ratio have recently been identified to characterize AKI and ARDS patient phenotypes (Bos et al., 2017; Famous et al., 2017). In mice, kidney and lung Tie2 mRNA and protein was diminished in response to LPS challenge (Matijs van Meurs, Kurniati, et al., 2009). However, the temporary loss of both Tie2 mRNA and protein was not associated with major changes in glomerular function which suggests that other factors regulate glomerular microvascular integrity (Haraldsson, Nystrom and Deen, 2008). Structurally, fenestrations in the glomerular EC contribute to permeability. In sepsis, the size and number of fenestrations are reduced leading to glomerular EC permeability. VEGF, a potent molecule known to increase endothelial permeability, is increased during

(13)

sepsis (Pierrakos and Vincent, 2010). VEGF produced in the kidney is suggested to play a role in the regulation of glomerular permeability (Hippenstiel et al., 1998). VEGF release during septic AKI may also contribute to distant organ injury and multiple organ failure (Chelazzi et al., 2015).

In summary, loss of the glycocalyx, altered fenestrations, altered expression of junction molecules, a dysbalanced Ang/Tie2 system all contribute to impaired endothelial integrity in the kidney associated with sepsis.

Sepsis-associated Acute Respiratory

Dysfunction Syndrome (ARDS)

Lung endothelial activation,

inflammation and leukocyte infiltration

The primary function of the lung is gas exchange for which it is equipped with a large surface of epithelial cells separated from the blood stream by ECs. The lung is directly exposed to infectious and toxic agents from the environment and is therefore equipped with an extensive innate immune system (Chen et al., 2018). Approximately 40% of septic patients suffer from acute respiratory distress syndrome (ARDS) or acute lung injury (ALI), which is associated with a poor outcome (Fujishima et al., 2016). In ARDS, gas exchange is hampered due to thickening of the alveolar membrane and flooding of cells and protein-rich exudate into the alveolus leading to oedema (Ware and Matthay, 2000). Often patients with sepsis and ARDS need to be ventilated. However, the force exercised by mechanical ventilation can aggravate lung inflammation and is therefore also detrimental to the lungs. Ultimately the lung fails due to decreased ventilation and diffusion which cannot be compensated for by supportive mechanical ventilation.

Infiltration of inflammatory cells is a major hallmark of sepsis induced acute respiratory distress syndrome (ARDS) (Ware

et al., 2004; Abraham et al., 2006). In sepsis,

various leukocyte subtypes but predominantly neutrophils were shown to infiltrate lung tissues. The expression of endothelial adhesion molecules and associated infiltrating neutrophils correlated with organ function impairment in lung biopsies from patients with ARDS (Windsor et al., 1993). Leukocytes tend to adhere to the post-capillary venules, but some studies have indicated that the lung capillaries are the preferred site of adherence and extravasation (Rossaint and Zarbock, 2013). Neutrophils located in the capillaries of 2-15μm diameter were retained longer allowing the neutrophils to be in direct physical contact with ECs yet contact alone does not initiate neutrophil transmigration (Rossaint and Zarbock, 2013; Margraf, Ley and Zarbock, 2019). Neutrophils were shown to migrate across inflamed ECs via their interaction with platelets in murine models of Escherichia coli,

Klebsiella- and LPS- induced pneumonia.

Platelets and inflamed vascular endothelium communicate and attract neutrophils via interactions between P-selectin on platelets and P-selectin ligand (PSGL-1) on neutrophils (Zarbock, Singbartl and Ley, 2006; Zarbock, Polanowska- Grabowska and Ley, 2007; Grommes et al., 2012). Platelet depletion reduced neutrophil accumulation in lung tissue indicating that neutrophils require aggregation with activated platelets to establish an interaction with ECs during bacterial infection (Zarbock, Polanowska- Grabowska and Ley, 2007). A large pool of neutrophils normally reside in the lung parenchyma (Kreisel et al., 2010) which upon bacteria-challenge extravasated immediately, forming extravascular clusters that co-localized with monocytes in the parenchymal area. Correspondingly, monocyte depletion resulted in reduced neutrophil recruitment and transendothelial migration (Kreisel et al., 2010). How neutrophils are recruited and the

(14)

role of the adhesion molecules seems to be inflammatory stimulus dependent as reviewed elsewhere (Margraf, Ley and Zarbock, 2019).

Alveolar macrophages were the first cells to detect LPS upon intra-tracheal LPS delivery rather than the ECs (Hollingsworth

et al., 2005). However, pulmonary ECs are

suggested to play an essential role in regulating responses to bacterial products, such as LPS. Andonegui and colleagues found endothelial TLR4 expression was sufficient to sequester neutrophils into the lungs after LPS challenge (Andonegui et al., 2003). These observations demonstrate the uniqueness of leukocyte infiltration in the lungs in sepsis, which differs from the rest of the organs mainly due to major differences in the microvascular make-up, with EC heterogeneity playing a major role (Margraf, Ley and Zarbock, 2019).

Lung vascular barrier dysfunction

The clinical presentation of lungs ‘drowning’ with fluids in sepsis is a clear indicator of a compromised vascular barrier integrity. Plasma from septic patients increases pulmonary EC permeability in vitro (Leligdowicz et al., 2018). As already mentioned, vascular barrier integrity is regulated by endothelial junctional molecules. During sepsis, leukocytes interact with the glycocalyx, resulting in its degradation (Schmidt et al., 2012) leading to pulmonary oedema (Maniatis and Orfanos, 2008), which is the hallmark of acute lung injury in sepsis.

The endothelial glycocalyx is pivotal in regulating the alveolar endothelial integrity. In an LPS-induced murine model, lung ECs express high levels of heparanase, which mediates degradation of glycocalyx in the lung microvascular beds, resulting in an increased vascular leakage (Uchimido, Schmidt and Shapiro, 2019). In human biopsies, low levels of heparanase are found in normal healthy lung vascular beds, whereas during damage the levels of heparanase are increased. A

heparanase inhibitor was found to attenuate pulmonary endothelial hyperpermeability after sepsis onset in a CLP-induced murine model (Schmidt et al., 2012). Inhibition of gap junction molecule connexin43 (Cx43) induced EC permeability in an experimental lung injury model (Parthasarathi, 2012). VE-cadherin expression in mice is 200 times higher in the lungs compared to the kidney and following LPS injection, VE-cadherin expression in lungs was reduced, but not in the kidney (Aslan et al., 2017). Endocytosis of VE-cadherin induces gaps between ECs in the lungs, leading to increased permeability (Lee and Slutsky, 2010). High mobility group protein B1 (HMGB1) was also shown to induce gaps in between ECs (Wang, 1999), and is implicated in sepsis- associated ARDS (Ueno et al., 2004).

(15)

REFERENCES

Abraham, E. et al. (2006) ‘Peripheral blood neutrophil activation patterns are associated with pulmonary inflammatory responses to lipopolysaccharide in humans.’, Journal of

immunology (Baltimore, Md. : 1950), 176(12), pp.

7753–60.

Adembri, C. et al. (2011) ‘Sepsis induces albuminuria and alterations in the glomerular filtration barrier: a morphofunctional study in the rat.’, Critical care (London, England), 15(6), p. R277. doi: 10.1186/cc10559.

Aird, W. C. (2007) ‘Phenotypic heterogeneity of the endothelium: I. Structure, function, and mechanisms’, Circulation Research. doi: 10.1161/01. RES.0000255691.76142.4a.

Andonegui, G. et al. (2003) ‘Endothelium-derived toll-like receptor-4 is the key molecule in LPS-induced neutrophil sequestration into lungs’,

Journal of Clinical Investigation, 111(7), pp. 1011–

1020. doi: 10.1172/JCI200316510.

Angus, D. C. and van der Poll, T. (2013) ‘Severe Sepsis and Septic Shock’, New England

Journal of Medicine, 369(9), pp. 840–851. doi:

10.1056/NEJMra1208623.

Asgeirsdottir, S. A. et al. (2012) ‘MicroRNA-126 contributes to renal microvascular heterogeneity of VCAM-1 protein expression in acute inflammation’,

AJP: Renal Physiology, 302(12), pp. F1630–F1639.

doi: 10.1152/ajprenal.00400.2011.

Aslan, A. et al. (2014) ‘The renal angiopoietin/ Tie2 system in lethal human sepsis.’, Critical care

(London, England), 18(2), p. 423. doi: 10.1186/

cc13806.

Aslan, A. et al. (2017) ‘Organ-Specific Differences in Endothelial Permeability- Regulating Molecular Responses in Mouse and Human Sepsis’,

Shock. doi: 10.1097/SHK.0000000000000841.

Aslan, A. et al. (2018) ‘Kidney histopathology in lethal human sepsis’, Critical Care. doi: 10.1186/ s13054-018-2287-3.

Bazzoni, G. and Dejana, E. (2004) ‘Endothelial cell-to-cell junctions: molecular organization and role in vascular homeostasis’,

Physiol Rev., 84(0031-9333 (Print)), pp. 869– 901.

doi: 10.1152/physrev.00035.2003.

Van den Berg, B. M., Vink, H. and Spaan, J. A. E. (2003) ‘The endothelial glycocalyx protects against myocardial edema’, Circulation Research. doi: 10.1161/01.RES.0000065917.53950.75.

Berlin, C. et al. (1995) ‘α4 integrins mediate lymphocyte attachment and rolling under physiologic flow’, Cell, 80(3), pp. 413–422. doi: 10.1016/0092-8674(95)90491-3.

Bos, L. D. et al. (2017) ‘Identification and validation of distinct biological phenotypes in

SUMMARY

Understanding the characteristics of different microvascular compartments between and within organs in the disease setting will enhance our understanding of why the microvasculature of different organs behaves in a specific manner upon a similar pathogenic insult. Moreover, therapeutic strategies designed to attenuate endothelial inflammatory activation will prevent leukocyte adhesion and transmigration which will lead to diminished organ failure in patients with sepsis. Differences in endothelial glycocalyx shedding, activation, inflammation, permeability, coagulation, metabolism, and epigenetic make-up all contribute to the organ-specific endothelial responses in sepsis. This heterogeneous response is part of the explanation why organs fail in phenotypically different ways. One of the consequences of these phenotypically different endothelial responses is that therapeutic intervention targeting the microvasculature will have different effects in different organs, such as being beneficial in one organ yet detrimental in another. Our knowledge on inter- and intra-organ endothelial heterogeneity has expanded in recent years, yet how these heterogenic responses are molecularly controlled in vivo remains elusive. Future studies are needed to fully understand endothelial responses in the setting of sepsis if we want to identify targeted organ-specific therapeutic options to counteract the aberrant endothelial responses driving organ failure in patients with sepsis.

(16)

Dejana, E., Bazzoni, G. and Lampugnani, M. G. (1999) ‘Vascular endothelial (VE)- cadherin: only an intercellular glue?’, Experimental cell research, 252(1), pp. 13–19. doi: 10.1006/excr.1999.4601.

Dejana, E. and Orsenigo, F. (2013) ‘Endothelial adherens junctions at a glance’, Journal of Cell

Science, 126(12).

Devi, S. et al. (2013) ‘Multiphoton imaging reveals a new leukocyte recruitment paradigm in the glomerulus.’, Nature medicine, 19(1), pp. 107–12. doi: 10.1038/nm.3024.

Duffy, M. J. et al. (2011) ‘Impaired endothelium-dependent vasodilatation is a novel predictor of mortality in intensive care’, Critical Care Medicine. doi: 10.1097/CCM.0b013e318206bc4a.

Famous, K. R. et al. (2017) ‘Acute respiratory distress syndrome subphenotypes respond differently to randomized fluid management strategy’,

American Journal of Respiratory and Critical Care Medicine. doi: 10.1164/rccm.201603-0645OC.

Fang, Y. et al. (2018) ‘The role of biomarkers of endothelial activation in predicting morbidity and mortality in patients with severe sepsis and septic shock in intensive care: A prospective observational study’, Thrombosis Research. doi: 10.1016/j. thromres.2018.09.059.

Fujishima, S. et al. (2016) ‘Infection site is predictive of outcome in acute lung injury associated with severe sepsis and septic shock.’, Respirology

(Carlton, Vic.), 21(5), pp. 898– 904. doi: 10.1111/

resp.12769.

Grommes, J. et al. (2012) ‘Disruption of platelet-derived chemokine heteromers prevents neutrophil extravasation in acute lung injury’,

American Journal of Respiratory and Critical Care Medicine, pp. 628–636. doi:

10.1164/rccm.201108-1533OC.

Haraldsson, B., Nystrom, J. and Deen, W. M. (2008) ‘Properties of the Glomerular Barrier and Mechanisms of Proteinuria’, Physiological

Reviews, 88(2), pp. 451–487. doi: 10.1152/

physrev.00055.2006.

Hato, T. et al. (2015) ‘The macrophage mediates the renoprotective effects of endotoxin preconditioning.’, Journal of the American Society

of Nephrology : JASN, 26(6), pp. 1347–62. doi:

10.1681/ASN.2014060561.

Hendrickson, C. M. and Matthay, M. A. (2018) ‘Endothelial biomarkers in human sepsis: pathogenesis and prognosis for ARDS’, Pulmonary

Circulation. doi: 10.1177/2045894018769876.

Herter, J. M. et al. (2014) ‘Adhesion molecules involved in neutrophil recruitment during sepsis-induced acute kidney injury’, Journal

of Innate Immunity, 6(5), pp. 597–606. doi:

10.1159/000358238.

Hippenstiel, S. et al. (1998) ‘VEGF induces hyperpermeability by a direct action on endothelial patients with acute respiratory distress syndrome

by cluster analysis’, Thorax. doi: 10.1136/ thoraxjnl-2016-209719.

Brealey, D. and Singer, M. (2000) ‘Multi-organ dysfunction in the critically ill: effects on different organs.’, Journal of the Royal College of Physicians

of London, 34(5), pp. 428–431.

Burridge, K. A. and Friedman, M. H. (2010) ‘Environment and vascular bed origin influence differences in endothelial transcriptional profiles of coronary and iliac arteries’, American

Journal of Physiology-Heart and Circulatory Physiology, 299(3), pp. H837–H846. doi: 10.1152/

ajpheart.00002.2010.

Chappell, D. et al. (2014) ‘Hypervolemia increases release of atrial natriuretic peptide and shedding of the endothelial glycocalyx’, Critical Care. doi: 10.1186/s13054-014-0538-5.

Chelazzi, C. et al. (2015) ‘Glycocalyx and sepsis-induced alterations in vascular permeability’,

Critical Care, 19(1), p. 26. doi:

10.1186/s13054-015-0741-z.

Chen, L. et al. (2018) ‘Inflammatory responses and inflammation-associated diseases in organs’,

Oncotarget. doi: 10.18632/oncotarget.23208.

Chi, J.-T. et al. (2003) ‘Endothelial cell diversity revealed by global expression profiling.’, Proceedings

of the National Academy of Sciences of the United States of America, 100(19), pp. 10623–10628. doi:

10.1073/pnas.1434429100.

Chousterman, B. G. et al. (2016) ‘Ly6Chigh Monocytes Protect against Kidney Damage during Sepsis via a CX3CR1-Dependent Adhesion Mechanism’, Journal of the American Society of

Nephrology, 27(3), pp. 792–803. doi: 10.1681/

ASN.2015010009.

Christaki, E. and Opal, S. M. (2008) ‘Is the mortality rate for septic shock really decreasing?’,

Current Opinion in Critical Care, 14(5), pp. 580–586.

doi: 10.1097/MCC.0b013e32830f1e25.

Colbert, J. F. and Schmidt, E. P. (2016) ‘Endothelial and Microcirculatory Function and Dysfunction in Sepsis’, Clinics in Chest Medicine. doi: 10.1016/j.ccm.2016.01.009.

Dauphinee, S. M. and Karsan, A. (2006) ‘Lipopolysaccharide signaling in endothelial cells’,

Laboratory Investigation, 86(1), pp. 9–22. doi:

10.1038/labinvest.3700366.

Davis, B. et al. (2007) ‘Podocyte-specific expression of angiopoietin-2 causes proteinuria and apoptosis of glomerular endothelia.’, Journal of the

American Society of Nephrology : JASN, 18(8), pp.

2320–2329. doi: 10.1681/ASN.2006101093. Dayang, E.-Z. et al. (2019) ‘Identification of LPS-Activated Endothelial Subpopulations With Distinct Inflammatory Phenotypes and Regulatory Signaling Mechanisms’, Frontiers in Immunology, 10(May), pp. 1–12. doi: 10.3389/fimmu.2019.01169.

(17)

cells.’, The American journal of physiology, 274 (5 Pt 1), pp. L678–L684. doi: 10.1074/jbc.272.24.15442.

Hollingsworth, J. W. et al. (2005) ‘The critical role of hematopoietic cells in lipopolysaccharide-induced airway inflammation’, American Journal of

Respiratory and Critical Care Medicine, 171(8), pp.

806–813. doi: 10.1164/rccm.200407-953OC. Ince, C. et al. (2016) ‘The endothelium in sepsis’,

Shock. doi: 10.1097/SHK.0000000000000473.

Jen, K.-Y., Haragsim, L. and Laszik, Z. G. (2011) ‘Kidney Microvasculature in Health and Disease’, in Contributions to Nephrology, pp. 51–72. doi: 10.1159/000313945.

Kellum, J. A. and Prowle, J. R. (2018) ‘Paradigms of acute kidney injury in the intensive care setting’, Nature Reviews Nephrology. doi: 10.1038/nrneph.2017.184.

King, J. et al. (2004) ‘Structural and functional characteristics of lung macro- and microvascular endothelial cell phenotypes’, Microvascular

Research. doi: 10.1016/j.mvr.2003.11.006.

Kitching, A. R., Holdsworth, S. R. and Hickey, M. J. (2008) ‘Targeting leukocytes in immune glomerular diseases’, Current Medicinal

Chemistry, 15(5), pp. 448–458. doi: http://dx.doi.

org/10.2174/092986708783503230.

Komarova, Y. and Malik, A. B. (2010)

Regulation of endothelial permeability via paracellular and transcellular transport pathways., Annual review of physiology. doi:

10.1146/annurev-physiol-021909-135833.

Kreisel, D. et al. (2010) ‘In vivo two-photon imaging reveals monocyte-dependent neutrophil extravasation during pulmonary inflammation’,

Proceedings of the National Academy of Sciences, 107(42), pp. 18073–18078. doi: 10.1073/

pnas.1008737107.

Kuligowski, M. P., Kitching, A. R. and Hickey, M. J. (2006) ‘Leukocyte Recruitment to the Inflamed Glomerulus: A Critical Role for Platelet-Derived P-Selectin in the Absence of Rolling’, The Journal of

Immunology, 176(11), pp. 6991–6999. doi: 10.4049/

jimmunol.176.11.6991.

Lacorre, D. A. et al. (2004) ‘Plasticity of endothelial cells: Rapid dedifferentiation of freshly isolated high endothelial venule endothelial cells outside the lymphoid tissue microenvironment’,

Blood. doi: 10.1182/blood-2003-10-3537.

Lee, W. L. and Slutsky, A. S. (2010) ‘Sepsis and Endothelial Permeability’, New England Journal

of Medicine, 363(7), pp. 689–691. doi: 10.1056/

NEJMcibr1007320.

Leligdowicz, A. et al. (2018) ‘Endothelial Activation: The Ang/Tie Axis in Sepsis’, Frontiers

in Immunology, 9, p. 838. doi: 10.3389/

fimmu.2018.00838.

Lerolle, N. et al. (2010) ‘Histopathology of septic shock induced acute kidney injury: Apoptosis

and leukocytic infiltration’, Intensive Care Medicine. doi: 10.1007/s00134-009- 1723-x.

Levi, M., Van Der Poll, T. and Schultz, M. (2012) ‘Systemic versus localized coagulation activation contributing to organ failure in critically ill patients’,

Seminars in Immunopathology, pp. 167–179. doi:

10.1007/s00281-011-0283-7.

Ley, K. et al. (2007) ‘Getting to the site of inflammation: The leukocyte adhesion cascade updated’, Nature Reviews Immunology, pp. 678– 689. doi: 10.1038/nri2156.

Liu, L. and Kubes, P. (2003) ‘Molecular mechanisms of leukocyte recruitment: Organ- specific mechanisms of action’, Thrombosis

and Haemostasis, pp. 213–220. doi: 10.1267/

THRO03020213.

Maas, S. L., Soehnlein, O. and Viola, J. R. (2018) ‘Organ-specific mechanisms of transendothelial neutrophil migration in the lung, liver, kidney, and aorta’, Frontiers in Immunology. doi: 10.3389/fimmu.2018.02739.

Maiden, M. J. et al. (2016) ‘Structure and function of the kidney in septic shock: A prospective controlled experimental study’, American Journal of

Respiratory and Critical Care Medicine, 194(6), pp.

692–700. doi: 10.1164/rccm.201511-2285OC. Maniatis, N. a and Orfanos, S. E. (2008) ‘The endothelium in acute lung injury/acute respiratory distress syndrome.’, Current opinion

in critical care, 14(1), pp. 22–30. doi: 10.1097/

MCC.0b013e3282f269b9.

Marcu, R. et al. (2018) ‘Human Organ-Specific Endothelial Cell Heterogeneity’, iScience, 4, pp. 20– 35. doi: 10.1016/j.isci.2018.05.003.

Margraf, A., Ley, K. and Zarbock, A. (2019) ‘Neutrophil Recruitment: From Model Systems to Tissue-Specific Patterns’, Trends in Immunology. doi: https://doi.org/10.1016/j.it.2019.04.010.

Marshall, J. C. (2014) ‘Why have clinical trials in sepsis failed?’, Trends in Molecular Medicine. doi: 10.1016/j.molmed.2014.01.007.

Mayr, F. B., Yende, S. and Angus, D. C. (2014) ‘Epidemiology of severe sepsis’, Virulence. Taylor & Francis, 5(1), pp. 4–11. doi: 10.4161/viru.27372.

Mehta, D. and Malik, A. B. (2006) ‘Signaling mechanisms regulating endothelial permeability.’,

Physiological reviews, 86(1), pp. 279–367. doi:

10.1152/physrev.00012.2005.

van Meurs, Matijs, Kümpers, P., et al. (2009) ‘Bench-to-bedside review: Angiopoietin signalling in critical illness – a future target?’, Critical Care. doi: 10.1186/cc7153.

van Meurs, Matijs, Kurniati, N. F., et al. (2009) ‘Shock-induced stress induces loss of microvascular endothelial Tie2 in the kidney which is not associated with reduced glomerular barrier function.’, American

journal of physiology. Renal physiology, 297(2), pp.

(18)

Rosenberg, R. D. and Aird, W. C. (1999) ‘Vascular-Bed–Specific Hemostasis and Hypercoagulable States’, New England Journal of

Medicine, 340(20), pp. 1555–1564. doi: 10.1056/

NEJM199905203402007.

Rossaint, J. and Zarbock, A. (2013) ‘Tissue-specific neutrophil recruitment into the lung, liver, and kidney’, Journal of Innate Immunity. doi: 10.1159/000345943.

Schmidt, E. P. et al. (2012) ‘The pulmonary endothelial glycocalyx regulates neutrophil adhesion and lung injury during experimental sepsis.’, Nature medicine, 18(8), pp. 1217–23. doi: 10.1038/nm.2843.

Seeley, E. J. and Bernard, G. R. (2016) ‘Therapeutic Targets in Sepsis: Past, Present, and Future’, Clinics in Chest Medicine. doi: 10.1016/j. ccm.2016.01.015.

Singbartl, K. and Ley, K. (2004) ‘Leukocyte recruitment and acute renal failure’, Journal of

Molecular Medicine, pp. 91–101. doi: 10.1007/

s00109-003-0498-8.

Singer, M. et al. (2016) ‘The third international consensus definitions for sepsis and septic shock (sepsis-3)’, JAMA - Journal of the American Medical

Association, 315(8), pp. 801–810. doi: 10.1001/

jama.2016.0287.

Singh, A. et al. (2007) ‘Glomerular Endothelial Glycocalyx Constitutes a Barrier to Protein Permeability’, Journal of the American Society of

Nephrology, 18(11), pp. 2885– 2893. doi: 10.1681/

ASN.2007010119.

Takasu, O. et al. (2013) ‘Mechanisms of cardiac and renal dysfunction in patients dying of sepsis’, American Journal of Respiratory and Critical

Care Medicine. doi: 10.1164/rccm.201211-1983OC.

Thijs, A. and Thijs, L. G. (1998) ‘Pathogenesis of renal failure in sepsis.’, Kidney international.

Supplement, 66, pp. S34–S37.

Uchimido, R., Schmidt, E. P. and Shapiro, N. I. (2019) ‘The glycocalyx: A novel diagnostic and therapeutic target in sepsis’, Critical Care. doi: 10.1186/s13054-018-2292-6.

Ueno, H. et al. (2004) ‘Contributions of high mobility group box protein in experimental and clinical acute lung injury’, Am J Respir Crit Care

Med, 170. doi: 10.1164/rccm.200402- 188OC.

Wallez, Y. and Huber, P. (2008) ‘Endothelial adherens and tight junctions in vascular homeostasis, inflammation and angiogenesis’,

Biochimica et Biophysica Acta - Biomembranes,

pp. 794–809. doi: 10.1016/j.bbamem.2007.09.003. Wang, C. et al. (2014) ‘Rapamycin antagonizes TNF induction of VCAM-1 on endothelial cells by inhibiting mTORC2’, The Journal of Experimental

Medicine. doi: 10.1084/jem.20131125.

Wang, H. (1999) ‘HMG-1 as a Late Mediator of Endotoxin Lethality in Mice’, Science, 285(5425), pp. 248–251. doi: 10.1126/science.285.5425.248. van Meurs, M. et al. (2009) ‘Shock-induced

stress induces loss of microvascular endothelial Tie2 in the kidney which is not associated with reduced glomerular barrier function’, AJP: Renal

Physiology, 297(2), pp. F272–F281. doi: 10.1152/

ajprenal.00137.2009.

Molema, G. (2010) ‘Heterogeneity in responses of microvascular endothelial cells during inflammation’, Endothelial Dysfunction and

Inflammation, 31, pp. 15–35. doi:

10.1007/978-3-0346-0168-9_2.

Moser, J. et al. (2016) ‘Intracellular RIG-I Signaling Regulates TLR4-Independent Endothelial Inflammatory Responses to Endotoxin’, The Journal

of Immunology, 196(11), pp. 4681–4691. doi:

10.4049/jimmunol.1501819.

Nourshargh, S. and Alon, R. (2014) ‘Leukocyte Migration into Inflamed Tissues’, Immunity. doi: 10.1016/j.immuni.2014.10.008.

Nourshargh, S., Hordijk, P. L. and Sixt, M. (2010) ‘Breaching multiple barriers: leukocyte motility through venular walls and the interstitium’,

Nature Reviews Molecular Cell Biology, 11(5), pp.

366–378. doi: 10.1038/nrm2889.

Nuytinck, H. K. et al. (1988) ‘Whole-body inflammation in trauma patients. An autopsy study.’, Archives of surgery (Chicago, Ill. :

1960), 123(12), pp. 1519–1524. doi: 10.1001/

archsurg.1988.01400360089016.

Panzer, U. et al. (2006) ‘Compartment-Specific Expression and Function of the Chemokine IP-10/CXCL10 in a Model of Renal Endothelial Microvascular Injury’, Journal of the

American Society of Nephrology. doi: 10.1681/

asn.2005040364.

Parthasarathi, K. (2012) ‘Endothelial connexin43 mediates acid-induced increases in pulmonary microvascular permeability’, AJP: Lung

Cellular and Molecular Physiology, 303(1), pp.

L33–L42. doi: 10.1152/ajplung.00219.2011.

Phillipson, M. et al. (2009) ‘Vav1 Is Essential for Mechanotactic Crawling and Migration of Neutrophils out of the Inflamed Microvasculature’,

The Journal of Immunology, 182(11), pp. 6870–

6878. doi: 10.4049/jimmunol.0803414.

Pierrakos, C. and Vincent, J.-L. (2010) ‘Sepsis biomarkers: a review’, Critical Care, 14(1), p. R15. doi: 10.1186/cc8872.

Poston, J. T. and Koyner, J. L. (2019) ‘Sepsis associated acute kidney injury’, BMJ (Online). doi: 10.1136/bmj.k4891.

Pusztaszeri, M. P., Seelentag, W. and Bosman, F. T. (2006) ‘Immunohistochemical expression of endothelial markers CD31, CD34, von Willebrand factor, and Fli-1 in normal human tissues.’,

Immunohistochemical expression of endothelial markers CD31, CD34, von Willebrand factor, and Fli-1 in normal human tissues, 54(4), pp. 385–95. doi:

(19)

Ware, L. B. et al. (2004) ‘Significance of Von Willebrand factor in septic and nonseptic patients with acute lung injury’, American Journal

of Respiratory and Critical Care Medicine. doi:

10.1164/rccm.200310-1434OC.

Ware, L. B. and Matthay, M. A. (2000) ‘The acute respiratory distress syndrome.’, The New

England journal of medicine, 342(18), pp. 1334–

1349. doi: 10.1056/NEJM200005043421806. Weinbaum, S., Tarbell, J. M. and Damiano, E. R. (2007) ‘The Structure and Function of the Endothelial Glycocalyx Layer’, Annual Review of

Biomedical Engineering. doi: 10.1146/annurev.

bioeng.9.060906.151959.

Windsor, A. C. et al. (1993) ‘Role of the neutrophil in adult respiratory distress syndrome.’,

The British journal of surgery, 80(1), pp. 10–17. doi:

10.1002/bjs.1800800106.

Xu, C. et al. (2014) ‘TNF-mediated damage to glomerular endothelium is an important determinant of acute kidney injury in sepsis.’, Kidney international, 85(1), pp. 72–81. doi: 10.1038/ki.2013.286.

Yan, R. et al. (2017) ‘Endothelial Interferon Regulatory Factor 1 Regulates Lipopolysaccharide-Induced VCAM-1 Expression Independent of NFKB’, Journal of Innate Immunity, 9(6), pp. 546–550. doi:

10.1159/000477211.

Yu, W.-K. et al. (2019) ‘Vascular endothelial cadherin shedding is more severe in sepsis patients with severe acute kidney injury’, Critical Care, 23(1), p. 18. doi: 10.1186/s13054-019-2315-y.

Yuan, L. et al. (2016) ‘A role of stochastic phenotype switching in generating mosaic endothelial cell heterogeneity’, Nature

Communications. doi: 10.1038/ncomms10160.

Zarbock, A., Polanowska-Grabowska, R. K. and Ley, K. (2007) ‘Platelet-neutrophil- interactions: Linking hemostasis and inflammation’, Blood

Reviews, pp. 99–111. doi: 10.1016/j.blre.2006.06.001.

Zarbock, A., Singbartl, K. and Ley, K. (2006) ‘Complete reversal of acid-induced acute lung injury by blocking of platelet-neutrophil aggregation’,

Journal of Clinical Investigation, pp. 3211–3219. doi:

Referenties

GERELATEERDE DOCUMENTEN

Moreover, sepsis studies should be designed to observe the global view of host responses, for example accounting for immune responses and their interaction with different

Nevertheless, eQTL mapping shows that 33% suggestive sepsis-associated loci can affect expression levels of 55 potential causal genes and some of these genes are

These data demonstrate that EC in different microvascular compartments, and also within the same microvascular compartment, exert a heterogeneous phenotype which can be identified

Pharmacological inhibition with FAK1 inhibitor FAK14 attenuated LPS-induced endothelial inflammatory activation and leukocyte adhesion partly via blockade of NF-ljB

Additionally, despite the expression of E-selectin and VCAM-1 in the kidney microvascular compartments, neutrophil infiltration here was relatively limited compared to that of

In hoofdstuk 3 beschreven we de kinetiek en locatie van de expressie van moleculen die er op EC voor zorgen dat witte bloedcellen worden aangetrokken (E-selectin en VCAM-1) in

Exposure of endothelial cells to lipopolysaccharide in vitro leads to the formation of four distinct cell subpopulations based on E-selectin and VCAM-1 expression, each

Acknowledging differences in Acute Kidney Injury: a complex clinical syndrome in critically ill patients.. University