• No results found

University of Groningen Sensing dengue and chikungunya virus (co-) infections Aguilar Briseño, Alberto

N/A
N/A
Protected

Academic year: 2021

Share "University of Groningen Sensing dengue and chikungunya virus (co-) infections Aguilar Briseño, Alberto"

Copied!
29
0
0

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

Hele tekst

(1)

Sensing dengue and chikungunya virus (co-) infections

Aguilar Briseño, Alberto

DOI:

10.33612/diss.172910464

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:

2021

Link to publication in University of Groningen/UMCG research database

Citation for published version (APA):

Aguilar Briseño, A. (2021). Sensing dengue and chikungunya virus (co-) infections. University of

Groningen. https://doi.org/10.33612/diss.172910464

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)

1

CHAPTER 1

(3)

Epidemiology of dengue virus and chikungunya virus

Dengue and chikungunya fever are two of the most important mosquito-borne viral diseases worldwide. Their etiological agents are dengue virus (DENV) and chikungunya virus (CHIKV), respectively. Both viruses are transmitted by mosquitoes of the genus Aedes, mainly Aedes

aegypti and Aedes albopictus. Globalization, increased urbanization and global warming are

considered the main factors that have contributed to the upsurge of cases and the geographical expansion of both viruses and mosquitoes (Figure 1)1–3.

Figure 1. Areas at risk of dengue and chikungunya virus infection. The map depicts the countries and regions currently

being at risk of dengue and chikungunya virus infections, as well the area of distribution of their mosquito vectors A.

albopictus and A. aegypti. Adapted from4,5. Figure made in Biorender.com

The oldest record of a dengue-like illness dates back to China (265- 420 A.D). Between 1779 and 1780, outbreaks of a disease comparable to dengue were reported in Africa, Asia and North America6. DENV was originally recognized as the causative agent of dengue in 1943 and since

then the incidence of the disease has dramatically increased7. Since then, four serotypes of

DENV (1-4) have been recognized and these are spread throughout tropical and subtropical regions of the world8. The largest number of cases ever reported worldwide occurred in 2019,

where in the Americas approximately 3.1 million cases were reported with more than 25,000 classifi ed as severe dengue9. To date, approximately 3.9 billion people in over 129 countries are

at risk of DENV infection. The most affected regions are the Americas, the Western Pacifi c and the South-East Asia, the latter representing 70% of the global burden of the disease2,10.

Chikungunya fever was fi rst described as a dengue-like disease during an outbreak in Tanzania (previously the Southern Province of Tanganyika) in 195311. The disease was mainly

characterized by fever, rash and arthralgia, with arthralgia being the main symptom of the disease. Due to the contorted postures of those experiencing severe arthralgia, the disease was

(4)

1

referred to as chikungunya, the Makonde term that translates to ‘that which bends up’3. For

the next 50 years, small and restricted episodes of transmission occurred only Africa. Later, during 2004-2005, a large outbreak hit Kenya and La Reunion Island, further spreading to India in 2006, where it infected more than 1 million people12. The virus then spread to Europe

where it caused restricted outbreaks in Italy, Croatia, France, Spain and Portugal. The first local episodes of transmission in the Americas were reported in Saint Martin in 2013, which was followed by a rapid spread to South, Central and North America13. To date, CHIKV is endemic

in over 60 countries worldwide, with Asia and the Americas being the most affected regions. The latest CHIKV outbreaks have been reported in Brazil, Pakistan, India, Sudan, Yemen and most recently Chad and Cambodia in 2020.14

Importantly, due to the geographic overlap in distribution of DENV, CHIKV and now Zika virus (ZIKV) and their respective vectors, cases of co-infections with these arboviruses have been reported15–19. The lack of vaccines or antivirals to ameliorate these diseases makes the search for

new tools that aid in development of effective therapies imperative to control arboviral infections.

The viruses

Dengue virus

DENV is an enveloped virus with a single-stranded, positive-sense RNA genome belonging to the Flaviviridae family20. The RNA genome of 11 kb encodes three structural glycoproteins:

envelope (E), precursor membrane (prM, of which the mature form is M), capsid (C); and seven non-structural proteins: NS1, NS2a, NS2b, NS3, NS4a, NS4b and NS521. One hundred and

eighty copies of E and M are embedded on the surface of the host-derived lipid membrane of mature virions. Multiple copies of another structural protein, capsid (C), encapsulate the RNA genome to form the nucleocapsid22–24 (Figure 2). The ectodomain of the E protein contains three

structurally defined domains. Domain I functions as the hinge region and is located between domain II and the immunoglobulin-like domain III25. Domain II harbors the fusion loop and

domain III has been postulated to be involved in virus receptor binding. Ninety homodimers of E are arranged in sets of three E head-to-tail dimers that lie in 30 rafts and form a herringbone pattern on the surface of the virus21. E and M are anchored to the viral membrane via their

transmembrane regions. The cytoplasmic tails of the E and M proteins do not interact with the nucleocapsid in the virion26.

(5)

Figure 2. DENV structure and genome composition. (A) Cross-sectional representation of a DENV virion depicting the

E dimer (yellow), the capsid protein (C, green), the M protein (red), the RNA genome and the viral membrane. (B) 3D model of a fully mature and immature DENV virion. (C) DENV genome and polyprotein composition. The scissors repre-sent the cleavage sites of furin, NS3 and signal peptidase proteases. Figure adapted from27,28 and made in Biorender.

Dengue virus replication cycle

The fi rst step in DENV virus replication is receptor binding and entry. To date, no specifi c receptor has been identifi ed for DENV infection in human cells29. However, several attachment

factors are known to mediate host cell binding and entry, including the ubiquitously expressed heat-shock proteins 70 and 90 and heparan sulfate30–32. Moreover, the more cell-type specifi c

C-type lectin receptors namely C-type lectin domain family 5 member A (CLEC5A), dendritic cell-specifi c intercellular adhesion molecule-3 grabbing non integrin (DC-SIGN, also known as CD209), L-SING and mannose receptor (CD206) and CD14 have been reported to serve as attachment factors for DENV infection33–37.

The E glycoprotein mediates cell entry via clathrin-mediated endocytosis. Acidifi cation of the endosomal compartments triggers a conformational change that dissociates the E protein dimers, thus exposing the domain II fusion loop. The E protein then inserts into the endosomal membrane, inducing the trimerization of E and subsequently driving fusion of the viral and

(6)

1

endosomal lipid membranes. Upon fusion, the nucleocapsid is released into the cytoplasm and C becomes degraded by ubiquitin-proteosome-dependent process38–40. Once genome uncoating

occurs, the positive sense RNA genome is translated into a single polyprotein. This polyprotein is further processed co- and post-translationally by cellular and viral proteases into 3 (E, prM and C) structural and 7 non-structural proteins (NS1, NS2a, NS2b, NS3, NS4a, NS4b and NS5)21 (Figure 2). The E and prM proteins dimerize and get oriented into the lumen of

the endoplasmic reticulum (ER). The non-structural glycoproteins assemble together to form the replication complex, in which NS5 functions as the RNA-dependent RNA polymerase21,41.

Replication of the genome occurs in ER-derived vesicles. The progeny RNA then exits the vesicles and associates with C proteins to form the nucleocapsid which buds into the ER thereby acquiring an E- and prM-containing lipid bilayer. The presence of the prM protein on newly assembled immature virions protects E from premature fusion from within the acidic compartments of exocytotic pathway. Virion maturation occurs while passing through the trans-Golgi network where the host enzyme furin cleaves prM into M and the pr peptide26,42,43. The

pr peptide remains attached to the mature virus progeny and dissociates from the virus particle after its release to the extracellular milieu42,44,45. Importantly, in case of DENV, the cleavage

of prM into M is generally inefficient and results in the release of a mixture of heterogeneous particles namely mature, partially mature and fully mature virions46. Depending of the infected

cell-type, up to 40% of released DENV virions still contain prM43,47. Generally, fully immature

virions are considered virtually non-infectious47–49.

Chikungunya virus

Chikungunya virus is a member of the family Togaviridae and genus Alphavirus to which Semliki forest virus, rubella virus and Sindbis virus also belong. CHIKV is an enveloped virus with a single-stranded positive-sense RNA genome of 11.8 kb. The RNA genome consists of two separate open reading frames (ORFs), the 5’ORF encodes for the four non-structural proteins (nsP1-4) and the 3’ORF encodes for the 5 structural proteins (C, E1, E2, E3 and 6K)50.

The genome is packaged by 240 copies of the capsid protein (C) to form the nucleocapsid. The nucleocapsid is surrounded by a host-cell derived lipid bilayer in which 240 units of E1 and E2 glycoproteins are anchored. E1 and E2, are arranged into 80 spikes, each spike is a trimer of E2/ E1 heterodimers50,51 (Figure 3). The E1 glycoprotein is a type II membrane protein and has three

domains, DI, DII and DIII. The fusion loop is located at the distal point of DII. The E2 contains three immunoglobulin domains, DA, DB and DC. DB is located at the end of the spike, DA in the center and DC lay proximal to the viral membrane52.

(7)

Figure 3. CHIKV structure and genome composition. (A) Cross-sectional representation of a CHIKV virion depicting

the E1-E2 trimer (yellow), the capsid protein (C, green), the RNA genome and the viral membrane. (B) 3D model of a CHIKV virion. (C) CHIKV genome and polyproteins composition. The scissors represent the cleavage sites of furin, nsP2, signal peptidase and capsid proteases. Figure adapted from53,54 and made in Biorender.

Chikungunya virus replication cycle

The fi rst step of CHIKV replication cycle is attachment to the host-cell lipid membrane. Recently, Mxra8 (also known as DICAM or limitrin), which is expressed in myeloid, mesenchymal and epithelial cells, has been reported to be a receptor for several arthritogenic alphaviruses, including CHIKV55. However, several attachment factors, namely prohibitin,

phosphatidylserine-mediated virus entry-enhancing receptors and glycosaminoglycans, have been described to mediated CHIKV entry to the cell56. The E2 glycoprotein facilitates attachment

to the cell surface receptors which promotes entry to the cell mainly by clathrin-mediated endocytosis. Acidifi cation of the endosomal vesicles (pH below 6.2 or 5.9) destabilizes the E1/E2 heterodimer leading to the exposure of the fusion loop and the subsequent fusion of the viral and endosomal lipid membranes57. After fusion occurs, the nucleocapsid is delivered

into the cytoplasm. Disassembly of the capsid and subsequent release of the RNA genome is thought to happen due to its binding to the large ribosomal unit58. The 5’ORF of the viral

(8)

1

nsP2, nsP3 and nsP4. This polyprotein is further processed in cis by nsP2 generating P123 and nsP459,60 (Figure 3). The newly unstable replication complex will then synthetize

negative-sense RNA intermediates61. When the levels of P123/nsP4 reach a stoichiometric balance, the

polyprotein will be completely processed to its individuals’ non-structural proteins. Meanwhile, the negative-sense RNA intermediates serve as a template for the synthesis of genomic and subgenomic RNAs62. Translation of the subgenomic RNA generates a polyprotein precursor

containing the structural proteins. Auto-processing of the polyprotein precursor will generate the C protein63. Due to the presence of a signal sequence in E3, the rest of the polyprotein will

be directed and translocated to the ER64. Once in the ER, host proteases will cleave the rest

of the polyprotein to generate E1, 6K and the fused form of E3-E265 (Figure 3). Upon further

post-translational modification E2/3 and E1 are transported to the trans-Golgi network where the host protease furin will mediate the cleavage of E2/3 into E2 and soluble E3, leading to the formation of E2/E1 heterodimers, which are transported to the plasma membrane. The viral RNA is packaged by C proteins to form the nucleocapsid. Interactions between the nucleocapsid and E2 proteins drive virion assembly and release at the plasma membrane62.

DENV and CHIKV tropism

DENV and CHIKV viruses are transmitted by the same mosquito’s species, thus both share the same route of entry to the human body. At the site of the mosquito bite, dermal fibroblasts, keratinocytes and skin-resident Langerhans cells are the first targets for DENV and CHIKV virus infection66–70. Inflammatory responses elicited after the mosquito bite induce the recruitment

of mononuclear phagocytic cells, namely monocytes, macrophages and dendritic cells which are the main targets for DENV infection in blood 49,71–78. For CHIKV it is thought that blood

monocytes, which have low susceptibility to infection, contribute to viral dissemination in muscles, joints, eyes and skin79,80. Additionally, DENV antigen has been detected in B cells,

T cells and platelets 81–84. Hepatocytes and endothelial cells have also found to sustain DENV

replication in vitro and in patient material66,85–87.

The innate immune system

Sensing pathogens: cellular effectors

Innate immune cells together with epithelial and endothelial cells constitute the first line of defense against invading pathogens. Innate immunity is largely composed of myeloid cells that can phagocyte and destroy pathogens88. The myeloid cell repertoire includes polymorphonuclear

and mononuclear phagocytic cells. Neutrophils, basophils and eosinophils comprise the polymorphonuclear phagocytes and have key role in controlling infection. Of these cells, neutrophils are of great importance as they are equipped with a variety of weapons to efficiently destroy invading pathogens88,89. The mononuclear phagocytic cells are a heterogeneous

(9)

population of cells encompassing monocytes subsets in blood and tissues, dendritic cells (DCs) as well as macrophages 90. Monocytes in blood represent a dynamic cell population composed

of three subsets which differ in phenotype and function91–93. In humans, blood monocytes are

classified by the expression of CD14 and CD1692. Classical monocytes (CM, CD14++CD16-)

account for the 85% of the monocyte pool. The rest of the monocyte population consists of intermediate (IM, CD14++CD16+) and non-classical (NM, CD14+CD16++) monocytes91,94.

CM play a very important role in the initiation of the inflammation by producing a large set of inflammatory mediators such as cytokines, myeloperoxidase and superoxide95. IM, known

as inflammatory monocytes, are rapidly recruited to the sites of inflammation where they contribute with the production of inflammatory cytokines. Additionally, IM can differentiate into inflammatory macrophages96–98. NM, also known as patrolling monocytes, act as protectors of

the endothelium and once activated differentiate into anti-inflammatory macrophages to repair tissue damage99–101. Macrophages are a heterogenous population of cells comprising

tissue-resident macrophages and monocyte-derived macrophages. Their morphology and function vary depending on their microenvironment102. Exposure to cytokines and other stimuli drives what

is known as macrophage polarization. In general, M1, or classically activated macrophages, have an inflammatory profile and mediate pathogen clearance. M2, or alternatively activated macrophages, have an anti-inflammatory profile and promote wound healing and resolution of the inflammation. However, this classification seems insufficient to describe their unique expression profiles102,103. Importantly, macrophages release chemotactic cytokines that recruit

other innate immune cells to the site of infection and can initiate adaptive immunity to invading pathogens via antigen presentation88. Dendritic cells (DCs) are also a heterogenous population

of myeloid cells. They are known to stimulate and induce adaptive immune responses88,104.

Moreover, DCs contribute to inflammation by producing pro-inflammatory cytokines and interferons upon infection. The DC repertoire in blood was initially comprised by conventional DC1 and DC2, and plasmacytoid DCs (pDCs)104. However, with the current emergence of

single-cell technologies, such as single-cell RNA sequencing, this repertoire has now expanded to seven subsets105,106. Additionally, each tissue also contains its own collection of

tissue-resident dendritic cells104.

Sensing pathogens: pattern recognition receptors

Cells of the innate immune system perform immune surveillance by utilizing membrane-bound and cytoplasmic pattern recognition receptors (PRRs). During infection, several PRRs work together to sense specifically pathogen-associated molecular patterns (PAMP’s) and host-derived danger signals that are released by infected cells termed tissue-host-derived danger-associated molecular patterns (DAMP’s). PRRs are germline encoded, constitutively expressed and show different pattern of expression among cell types. Engagement of these PRRs in turn activates specific signaling cascades leading to the production of cytokines and chemokines 107,108. PRRs

can be classified in several families based on domain homology encompassing the membrane-bound Toll-like receptors (TLRs), C-type lectin receptors (CLRs) and the intracellularly located retinoic acid-inducible gene I (RIG-I)-like receptors (RLRs), nucleotide-binding domain,

(10)

1

leucine-rich repeat (LRR)-containing receptors (NLRs or NOD-like receptors), AIM2-like receptors (ALRs) and cGAS (Figure 4)107,109,110. From all these families of PRRs, TLRs and RLRs

have been widely study for their role in sensing RNA viruses111. ALRs and cGAS are cytoplasmic

DNA sensors, thus commonly sensing DNA viruses107. RLRs comprise three cytosolic proteins,

retinoic acid-inducible gene-I (RIG-I), melanoma differentiation gene 5 (MDA5) and laboratory of genetics and physiology 2 (LGP2). These receptors sense specific features of RNA genomes like the presence of 5’triphosphate RNA, single or double-stranded RNA and characteristic regions of genomes like the poly-uridine region of the hepatitis C virus107. Following the

sensing of the viral genomes, RLRs translocate to the mitochondria or the peroxisomes where the adaptor molecule mitochondrial antiviral protein signaling protein (MAVS) is present. Depending on MAVS localization, specific signaling pathways will be triggered leading to the activation of transcription factors nuclear factor kappa B (NF-κB) and interferon regulatory factors 3 and 7 (IRF3 and IRF7, respectively). Ergo, inducing the production of Type-1 IFNs and pro-inflammatory cytokines107,109,110. Binding of IFNs to their specific receptors will activate the

Janus activated kinase-signal transducer and activator of transcription (JAK-STAT) pathway and subsequent production of interferon stimulated genes (ISGs)107.

TLRs were originally identified as homologues of the Toll receptor in the fruit fly Drosophila

melanogaster112. In mammals, these receptors function as sentinels of the innate immune system.

Ten functional TLRs are present in humans and 13 in mice. They are type I transmembrane proteins that contain three domains, an extracellular domain with long-leucine-rich repeat regions that mediates protein-protein or protein-PAMP sensing, a transmembrane domain and a cytoplasmic toll/interleukin-1 receptors (TIR) domain113. Plasma membrane-bound TLR1,

TLR2, TLR4, TLR5, TLR6 and TLR10 sense lipids, proteins, lipoproteins and glycoproteins while TLR3, TLR7, TLR8, TLR9, TLR11 and TLR13 are localized inside the endosomal compartments where they sense nucleic acids110,113. Upon PAMP recognition, sorting adaptors

molecules toll/interleukin-1 receptor domain-containing adaptor protein (TIRAP) and TIR domain-containing adaptor protein inducing IFNβ (TRIF)-related adaptor molecule (TRAM) will recruit and assemble supramolecular organizing centers (SMOCs), the myddosome and the triffosome, respectively. The myddosome consists of the myeloid differentiation primary response protein 88 (MyD88) and several copies of the IL-1 receptor-associated kinases (IRAK) serine-threonine kinases. The triffosome, on the other hand, consists of TRIFF which is believed to recruit TNF receptors-associated factor 3 (TRAF3) to form this complex110,113,114. Activation

of all TLRs will lead to the production of cytokines and chemokines through NF-κB and activator protein 1 (AP-1) activation via the myddosome. Production of interferons, however, is induced by TLRs localized in the endosomal compartments and its mediated by the interferon regulatory factors (IRF) family of transcription factors via the triffosome and myddosome, with the triffosome complex most associated with these responses113,114. The production of cytokines

and interferons will trigger a variety of inflammatory responses including the recruitment of immune cells and the induction of adaptive immunity114.

(11)

Figure 4. Pattern recognition receptors and innate immune signaling. General overview of the different families of

PRRs utilized by innate immune cells to perform immune surveillance. These families encompass surface bound and endosomal expressed TLRs, RIG-I-like receptors, NLRs, ALRs and the cGAS/STING pathway. Activation of these PRRs trigger signaling pathways leading to nuclear translocation of transcription factors IRFs, NF-κB and AP-1, and the subse-quent production of infl ammatory and antiviral cytokines. Figure made in Biorender.com

Endothelial cells

Blood vessels encompassing veins, venules, arteries, arterioles and capillaries are known as the vasculature. Their function is to transport the blood throughout the body115. The vascular

endothelium is a semi-permeable layer of endothelial cells (EC) that line the macro- and microvasculature. Its function is to provide a barrier between the vessel’s lumen and the underlying tissue116. EC are sealed to each other by tight junctions, adherens junctions and

VE-cadherin117. In addition, the glycocalyx barrier located over the EC regulates vascular

permeability, protects the EC against invading pathogens, infl ammatory mediators and controls the interactions between EC and immune cells in blood118. During homeostasis (Figure 5),

Angiopoietin 1 released by pericytes binds to Tie2 receptor ensuring barrier maintenance and endothelium integrity119. Activation of the endothelium can occur either due to direct sensing of

PAMPs and DAMPs by PRR expressed on EC or indirectly due to pre-established infl ammatory responses. Upon activation, EC will respond by producing angiopoietin 2, infl ammatory and chemoattractant cytokines (e.g. IL-6, IL-8 and MCP-1) and induce the enhanced surface expression of adhesion molecules (E-selectin, P-selectin, vascular cell adhesion protein 1 or VCAM-1, and intercellular adhesion molecule 1 or ICAM-1)120–122. This activation will

(12)

1

and subsequently induce endothelium disruption and vascular permeability122–125. Endothelium

activation and dysfunction are two independent processes, with the latter resulting in the EC not being able to restore the homeostatic healthy state. Exacerbated infl ammation and increased leukocyte traffi cking are some of the best-known factors that can contribute to the endothelium dysfunction. Importantly, the endothelial responses to infl ammation as well as its regulatory functions are organ specifi c126. Hence, understanding these responses and functions are crucial

for selecting the right specifi c treatments to alleviate endothelium-associated pathogenesis.

Figure 5. Endothelium during homeostasis. Endothelial cells form a semi-permeable layer that line the micro and macro

vasculature known as the endothelium. Keeping the integrity of the endothelium is crucial for the proper functioning and maintenance of organs. Several molecules work in consonance to maintain the endothelium in homeostasis. For instance, pericytes release angiopoetin-1 that binds the Tie2 receptor and promoting homeostasis. Tight and adherens junctions seal ECs making the endothelium a semi-permeable layer. The glycocalyx barrier protects EC from infl ammatory mediators and invading pathogens. Furthermore, ECs are equipped with PRRs that detect pathogens and/or their components, which will lead to the activation of the endothelium and the production of adhesion molecules and infl ammatory and antiviral cytokines. Figure made in Biorender.

Dengue disease pathogenesis and innate immune responses

Dengue viruses are responsible for an estimated number of 390 million of infections in which the disease is clinically apparent in 96 million cases8. Clinical manifestations range from an

acute self-limited dengue with or without warning sings (previously termed dengue fever) to a life-threatening severe dengue (previously known as dengue hemorrhagic fever and dengue shock syndrome)127,128. The clinical course of dengue is depicted in Figure 6. While most DENV

infections are asymptomatic (75%), few result in a mild febrile disease (25%) characterized by fever, muscular pain, headache, eye pain and rash. Moreover, an estimated 1-5% of patients further develop thrombocytopenia, vascular permeability and plasma leakage, hallmarks of severe disease. The dengue mortality rate is low (<1%), with most of the patients recovering after 7 days post-illness. To date, there are no specifi c treatments available to ameliorate severe disease.

(13)

0 Febrile Phase 1 2 3 4 5 6 Critical Phase

Time (days)

• Viremia • Host immune responses • Shock • Bleeding • Organ impairment 10 7 8 9 Recovery Phase • Resolution plasma leakage • No viremia • Antibody responses

Figure 6. Clinical course of dengue. The clinical course of DENV infection consist of three phases. The febrile phase

is characterized by high viral titers which coincides with the clinical manifestations of the disease. An exacerbated and uncontrolled immune response to the infection can lead to the development of vascular permeability and plasma leakage, clinical manifestations that hallmark severe dengue (critical phase). The critical phase is followed by the recovery phase in which plasma leakage is resolved and the virus is cleared. Adapted from129.

Factors like comorbidities, bacterial co-infections and/or translocations and the presence of cross-reactive infectious-enhancing antibodies (raised during previous heterologous infections) have been linked to severe disease development130–135. Myeloid cells like monocytes, macrophages and

dendritic cells produce Type-1 IFNs and pro-inflammatory cytokines during DENV infection that, on the one hand, can be protective for some patients or, on the other hand, can promote severe disease development. Among the three monocyte subsets in blood, increased frequencies of intermediate or inflammatory monocytes have been associated with onset of severe dengue 72,136,137.

Macrophages can contribute to the overall immune responses driving severe disease by producing inflammatory mediators upon DENV infection 138,139. DCs, have been shown to contribute to the

innate immune responses in course of DENV infection by inducing the production of Type-1 IFNs 140–142. Interestingly, the presence of heterologous antibodies from previous DENV infections

that bind to FCγR-expressing cells can enhance DENV entry and infection, a phenomenon called antibody-dependent enhancement (ADE). In vitro DENV infections in the presence of human serum have demonstrated that monocytes, macrophages and mature dendritic cells can support ADE49,77,139,143. Knowing the important role that monocytes, macrophages and DCs play

during DENV infection and pathogenesis, future research should focus in addressing the overall contribution of these cells to severe dengue development.

As previously mentioned, severe dengue is hallmarked by vascular permeability and plasma leakage, indicating a major role of the endothelium in severe disease development. Exacerbated inflammation due to cytokine storm is thought to be the principal contributor to these endothelial disturbances and the immunopathogenesis of severe dengue. Indeed, proinflammatory cytokines IL-1β and TNFα, which are known to induce endothelium activation, have been found to be elevated in patients with dengue129,144,145. During severe dengue, several systems in

(14)

1

charge of maintaining endothelium integrity are compromised. For instance, EC release high levels of angiopoietin 2 which in turn bind to Tie2 impeding its binding with angiopoietin 1, thus disrupting the EC integrity146,147. The coagulation system also gets affected as DENV

inhibits the conversion of prothrombin intro thrombin leading to the development of internal bleeding148. Additionally, shedding of the glycocalyx during severe dengue has been shown to

promote plasma leakage149.

PRR-expressing cells can indirectly contribute to endothelial activation and dysfunction by sensing specific features of DENV during the course of the infection, subsequently leading to the production of pro-inflammatory mediators. For instance, cytoplasmic RIG-I and MDA5, and endosomal-expressed TLR3 and TLR7, are known to sense DENV RNA genome leading to the production of Type-1 IFNs and proinflammatory cytokines. Interestingly, the production of Type-1 IFNs generally has a protective antiviral role impacting DENV replication. In contrast, signaling cascades triggered by these endosomal and cytoplasmatic sensors can also lead to the production of TNFα, IL-6 and IL-1β, which are cytokines associated with severe dengue. Moreover, DENV and presumably other flaviviruses can induce mitochondrial damage-associated mitochondrial DNA release leading of the activation of cGAS/STING and TLR9 pathways and the subsequent production of Type-I IFNs150–152. Importantly, DENV

non-structural proteins (NSs) target key molecules of these signaling pathways leading to the production of Type-1 IFNs as a mechanism of immune evasion. For instance, NS2B and NS2B3 target the cGAS/STING pathway to prevent mitochondrial (mt)DNA sensing during DENV infection150,151,153. NS4 blocks the interaction between adaptor molecule MAVS and

RIG-I. Furthermore, NS4B, NS2A and NS2B3 impede the phosphorylation and activation of IKKԑ and TBK1. With regard to the IFNα/β receptor signaling, NS4 and NS5 prevent the phosphorylation of STAT1 and NS5 also targets STAT2 for degradation via proteosome further inhibiting the production of ISGs. DENV virus immune evasion of TLRs pathways hasn’t been described yet. However, DENV NS1 is known to induce TLR activation and contributes to severe dengue development. TLR4 expressed on the surface of monocytes can sense DENV NS1, which is released from DENV infected cells, leading to the production of cytokines and further activation of the endothelium154,155. In addition, in vitro and ex vivo models have shown

that NS1 can induce vascular permeability and plasma leakage by directly interacting with glycocalyx components156–159. DENV can also be sensed by CLRs, indeed, it has been reported

that the C-type lectin domain 5 member A (CLEC5A) upon sensing of DENV particles can induce potent immune responses that contribute to severe disease in an in vivo model of ADE33.

Moreover, it has been recently described that CLEC5A in partnership with TLR2 can sense extracellular vesicles released by platelets in vivo, further promoting the release of cytokines and neutrophil extracellular traps (NETs) by neutrophils. The release of these inflammatory effectors by neutrophils results in vascular permeability160. Furthermore, breakdown of ECs

tight junctions leading to vascular leakage has been linked to the release of tryptase by mast cells upon sensing DENV161. Taken together, rapid antiviral responses coupled with controlled

inflammatory responses are crucial for the containment of DENV infection. The precise molecular mechanisms involved in the development of severe dengue are not fully understood

(15)

yet. However, our understanding of the role of PRRs in the sensing of DENV particles by innate immune cells as well ECs has increased over the years. Further research should focus on using agonists or antagonists of these receptors and their associated downstream signaling pathways in relevant settings to develop new therapies aiming to control severe dengue development.

Chikungunya disease pathogenesis and innate immune

responses

After the mosquito bite, CHIKV rapidly spreads to several tissues throughout the bloodstream. CHIKV titers in serum samples from infected patients have been reported to reach up to 109 CHIKV particles /mL162,163. The high viremia coincides with the occurrence of clinical

symptoms such as fever, rash, headache, myalgia and arthralgia which lasts up to 5-7 days164.

However, contrary to dengue, which is transmitted by the same mosquito’s species, 50 to 97% of the CHIKV-infected individuals will develop the aforementioned clinical manifestations163.

Interestingly, even though CHIKF is a mild self-limiting disease, severe cases requiring hospitalization have been reported165,166. During the outbreak in La Réunion island in 2006,

complications such as myocarditis, encephalitis and multiorgan failure following CHIKV infection in newborns, the elderly and diabetic patients were associated with disease mortality. In addition, atypical clinical manifestations such as Guillain-Barré syndrome have been reported in CHIKV-infected patients during La Réunion and French Polynesia outbreaks167,168.

Importantly, considering the fact that the clinical symptoms of CHIKF highly resemble those of dengue and Zika, the correct diagnosis of the disease remains challenging. To date, there is no vaccine and no specific therapeutic agent available to prevent or ameliorate CHIKF.

Nearly 40% of patients with acute CHIKF will develop chronic disease hallmarked by polyarthralgia, which can last up to five years after acute infection164 (Figure 7). Presence

of viral antigen and potent inflammation in the joints are thought to be the cause of chronic disease. Indeed, the presence of viral antigen in macrophages, monocytes and lymphocytes has been observed in synovial fluids and tissues of a one-year CHIKV chronically infected patient. Additionally, several inflammatory mediators such as MCP1, IP-10, MIG, IL-6, TNFα and IL-1β have been detected in plasma samples in acute CHIKV patients and non-human primates169. From those, MCP1, IL-6 and IL-8 have been detected in synovial fluids

of chronic CHIKV patients and their presence is indicative of chronic disease development (Figure 7)163. Importantly, the use of in vitro and in vivo mouse models and non-human primate

models to study CHIKV pathogenesis have increased the current understanding of the CHIKV pathogenesis and associated pathologies169. Those models have underscored the importance of

monocytes and macrophages in CHIKV chronic disease. Namely, both cell types were found to be recruited to inflammation sites via MCP1. There, monocytes and macrophages not only contribute to CHIKV pathogenesis by producing inflammatory mediators, but also serve as a reservoir for viral replication. In NHP, CHIKV RNA has been detected in joints and lymphoid organs after clearance in blood (up to 9 days post infection). Additionally, after acute infection

(16)

1

the virus is mainly detected in macrophages. The use of CCR2 defi cient mice, which function as a receptor for MCP1, has shown altered immune responses and pathogenesis following CHIKV infection as evidenced by the development of severe and prolonged arthritis. Furthermore, the use of Bindarit, a drug that specifi cally downregulates gene expression of MPC1170, has been

shown to prevent monocyte infi ltration and CHIKV-mediated arthritis in mice171. Moreover,

infl ammatory monocytes and macrophages infi ltrating into bone tissue can differentiate and/or interact with osteoblasts causing bone erosion, thus contributing with the severity of the CHIKV-mediated arthritis172. In CHIKV patients, this process has been associated with presence of high

levels of receptor activator of nuclear factor kappa-B ligand (RANKL), a protein that controls bone regeneration and remodeling171,173. Another cell type found present in high numbers in

infl amed joints of mice are natural killer cells (NKs). The importance of NKs in patients is evidenced by their strong activation during early CHIKV infection, which has been correlated with balanced adaptive immune responses to CHIKV antigens169,174,175.

CHIKV infection 2-4 days Acute Phase Months-yearsIFN responseMCP1, TNFα, IL-1β, IL-6 Chronic PhaseMonocyte/Macrophage infiltration to jointsArthritisBone loss Viremia 3-5 days SymptomsIFN responseMCP1, TNFα, IL-1β, IL-6 Viremia Symptoms

Figure 7. Clinical course of Chikungunya. After the mosquito bite there is an incubation period of 2-4 days followed by

a rapid increase in CHIKV titers in blood. This coincides with development of clinical manifestations in the acute phase which last 3-days. During this phase, immune cells like monocyte and macrophages produce huge amounts of infl ammatory mediators. Activated monocytes and macrophages then infi ltrate to joint and bone tissue causing infl ammation and promoting chronic disease-associated symptoms, such as polyarthralgia and bone loss, that can persist from several months up to several years. Adapted from169.

The host immune response is crucial for the containment of CHIKV infection and mainly driven by the production of antiviral Type-1 IFNs. PRRs such as RIG-I-, MDA5- and TLR3 are most likely to sense CHIKV infection based on their function. Indeed, CHIKV genome- and replication intermediates-sensing by these PRRs have been shown to be involved in triggering the aforementioned antiviral responses in in vitro/in vivo models176–178. Importantly, experiments

in mice have provided evidence that an MyD88-dependent sensor such as TLR2, TLR4 or TLR7 may also control viral dissemination and spread177. However, the involvement of any of those

PRRs during CHIKV infection in patients or human in vitro systems hasn’t been described yet. Additionally, it was recently reported that mtDNA is released into the cytoplasm during CHIKV

(17)

infection, further inducing Type-1 interferon responses via the cGAS/STING pathway179.

Interestingly, CHIKV counteracts the cGAS/STING pathway by inducing the degradation of cGAS through an ATG7 autophagy-dependent mechanism and via direct interaction of nsp1 to STING resulting in signaling inhibition179. Furthermore, CHIKV nsp2 inhibits the JAK/STAT

signaling pathway by impeding the translocation of STAT1 into the nucleus, preventing the transcription of ISGs180.

Scope of the thesis

The research presented in this thesis focuses on the innate immune responses elicited by dengue and chikungunya virus infections. Although the immunopathologies of dengue and chikungunya virus infections have been described, the mechanisms underlying the exacerbated inflammation observed during severe dengue and CHIKV-mediated arthritis are not completely understood. Interestingly, monocytes in blood represent primary targets for DENV and CHIKV infections, and due to their associated role as innate immune sentinels, also contribute to disease pathogenesis by producing inflammatory and antiviral cytokines. Here, by utilizing human primary cell in vitro systems, we aim to decipher the contribution of monocytes in the immunopathology of dengue and CHIKV infections.

Due to the number of dengue and chikungunya cases augmenting over the last decades, the incidence of co-infections with both viruses has dramatically increased. In chapter 2, by using primary human PBMCs, we aimed to describe the viral kinetics and innate immune responses of DENV and CHIKV co-infections in vitro. By co-infecting with UV-inactivated particles, we delineate the contribution of viral replication in such responses. Moreover, multiplex analysis of cytokines allowed us to determine the innate immune signature of DENV and CHIKV mono-/co-infections. As PBMCs comprise several types of cells such as lymphocytes and monocytes, we investigated in the subsequent chapters which of these cells contribute to the immunopathogenesis of DENV and CHIKV infections.

Ex vivo analysis of PBMCs has previously shown than CHIKV infection increases the frequency

of intermediate inflammatory monocytes. In chapter 3, we used a previously established in vitro model of CHIKV infection to rewire PBMC responses and delineate the role of intermediate monocytes in virus replication and innate responses.

As sentinels of the innate immune system, monocytes use toll-like receptors to sense pathogens and induce innate immune responses. In chapter 4, by using human primary PBMCs and endothelial cells in vitro models, we identified the role of TLR2 as a sensor and regulator of the innate immune responses following DENV infection. Furthermore, by analyzing the TLR2 expression and monocyte subset frequencies in a DENV-infected pediatric cohort, we identified a prognostic value of TLR2 expression in disease pathogenesis.

(18)

1

Since virtually non-infectious immature DENV particles are released in large numbers by DENV infected cells, we aimed to investigate the contribution of these particles in the innate immune responses during dengue virus infection. In chapter 5, we indicate that the maturation status of the DENV is likely to influence the kinetics and extent of inflammatory responses during DENV infection. Additionally, we scrutinized the role of TLR2 in these responses using primary human PBMCs and endothelial cell in vitro systems.

(19)

References

1. Guzman, M. G. & Harris, E. Dengue. Lancet 385, 453–465 (2015).

2. Bhatt, S. et al. The global distribution and burden of dengue. Nature 496, 504–507 (2013).

3. Vu, D. M., Jungkind, D. & LaBeaud, A. D. Chikungunya Virus. Clinics in Laboratory Medicine vol. 37 371–382 (2017).

4. CDC. Geographic Distribution | Chikungunya virus | CDC. Geographic distribution 1 https://www.cdc.gov/ chikungunya/geo/index.html (2016).

5. DengueMap. https://www.healthmap.org/dengue/en/.

6. Gubler, D. J. Dengue and Dengue Hemorrhagic Fever. CLINICAL MICROBIOLOGY REVIEWS vol. 11 (1998). 7. Messina, J. P. et al. Global spread of dengue virus types: Mapping the 70 year history. Trends in Microbiology

(2014) doi:10.1016/j.tim.2013.12.011.

8. World Health Organization. Dengue control. http://www.who.int/denguecontrol/disease/en/ (2018). 9. WHO. Dengue and severe dengue. Who http://www.who.int/mediacentre/factsheets/fs117/en/ (2016). 10. Brady, O. J. et al. Refining the Global Spatial Limits of Dengue Virus Transmission by Evidence-Based

Consensus. PLoS Negl. Trop. Dis. 6, (2012).

11. Lumsden, W. H. R. An epidemic of virus disease in Southern Province, Tanganyika territory, in 1952–1953 II. General description and epidemiology. Trans. R. Soc. Trop. Med. Hyg. 49, 33–57 (1955).

12. Kaur, P. et al. Chikungunya outbreak, South India, 2006. Emerg. Infect. Dis. 14, 1623–1625 (2008). 13. Weaver, S. C. & Lecuit, M. Chikungunya virus and the global spread of a mosquito-borne disease. New

England Journal of Medicine vol. 372 1231–1239 (2015).

14. World Health Organisation. Chikungunya Fact sheet N°327. Who https://www.who.int/news-room/fact-sheets/detail/chikungunya (2020).

15. Waggoner, J. J. et al. Viremia and Clinical Presentation in Nicaraguan Patients Infected With Zika Virus, Chikungunya Virus, and Dengue Virus. Clin. Infect. Dis. 63, 1584–1590 (2016).

16. Villamil-Gómez, W. E., González-Camargo, O., Ayubi, J., Zapata-Serpa, D. & Rodriguez-Morales, A. J. Dengue, chikungunya and Zika co-infection in a patient from Colombia. Journal of Infection

and Public Health vol. 9 684–686 (2016).

17. Castellanos, J. E. et al. Dengue-chikungunya coinfection outbreak in children from Cali, Colombia in 2018– 2019. Int. J. Infect. Dis. (2020) doi:10.1016/j.ijid.2020.10.022.

18. Silva, M. M. O. et al. Concomitant Transmission of Dengue, Chikungunya, and Zika Viruses in Brazil: Clinical and Epidemiological Findings from Surveillance for Acute Febrile Illness. Clin. Infect. Dis. 69, 1353–1359 (2019).

19. Carrillo-Hernández, M. Y., Ruiz-Saenz, J., Villamizar, L. J., Gómez-Rangel, S. Y. & Martínez-Gutierrez, M. Co-circulation and simultaneous co-infection of dengue, chikungunya, and zika viruses in patients with febrile syndrome at the Colombian-Venezuelan border. BMC Infect. Dis. 18, (2018).

20. Harris, E., Holden, K. L., Edgil, D., Polacek, C. & Clyde, K. Molecular biology of flaviviruses. in Novartis

Foundation symposium vol. 277 23–39; discussion 40, 71–3, 251–3 (2006).

21. Kuhn, R. J. et al. Structure of dengue virus: Implications for flavivirus organization, maturation, and fusion.

Cell 108, 717–725 (2002).

22. Nybakken, G. E., Nelson, C. A., Chen, B. R., Diamond, M. S. & Fremont, D. H. Crystal Structure of the West Nile Virus Envelope Glycoprotein. J. Virol. 80, 11467–11474 (2006).

(20)

1

23. Modis, Y., Ogata, S., Clements, D. & Harrison, S. C. Variable Surface Epitopes in the Crystal Structure of

Dengue Virus Type 3 Envelope Glycoprotein. J. Virol. 79, 1223–1231 (2005).

24. Modis, Y., Ogata, S., Clements, D. & Harrison, S. C. A ligand-binding pocket in the dengue virus envelope glycoprotein. Proc. Natl. Acad. Sci. U. S. A. 100, 6986–6991 (2003).

25. Morrone, S. R. & Lok, S. M. Structural perspectives of antibody-dependent enhancement of infection of dengue virus. Current Opinion in Virology vol. 36 1–8 (2019).

26. Mukhopadhyay, S., Kuhn, R. J. & Rossmann, M. G. A structural perspective of the Flavivirus life cycle.

Nature Reviews Microbiology vol. 3 13–22 (2005).

27. Simmonds, P. et al. ICTV virus taxonomy profile: Flaviviridae. J. Gen. Virol. 98, 2–3 (2017).

28. Jarmer, J. et al. Variation of the Specificity of the Human Antibody Responses after Tick-Borne Encephalitis Virus Infection and Vaccination. J. Virol. 88, 13845–13857 (2014).

29. Rodenhuis-Zybert, I. A., Wilschut, J. & Smit, J. M. Dengue virus life cycle: Viral and host factors modulating infectivity. Cellular and Molecular Life Sciences vol. 67 2773–2786 (2010).

30. Reyes-del Valle, J., Chávez-Salinas, S., Medina, F. & del Angel, R. M. Heat Shock Protein 90 and Heat Shock Protein 70 Are Components of Dengue Virus Receptor Complex in Human Cells. J. Virol. 79, 4557– 4567 (2005).

31. Chen, Y. et al. Dengue virus infectivity depends on envelope protein binding to target cell heparan sulfate.

Nat. Med. 3, 866–871 (1997).

32. Germi, R. et al. Heparan sulfate-mediated binding of infectious dengue virus type 2 and yellow fever virus.

Virology 292, 162–168 (2002).

33. Chen, S. T. et al. CLEC5A is critical for dengue-virus-induced lethal disease. Nature 453, 672–676 (2008). 34. Miller, J. L. et al. The mannose receptor mediates dengue virus infection of macrophages. PLoS Pathog. 4, (2008). 35. Dejnirattisai, W. et al. Lectin switching during dengue virus infection. J. Infect. Dis. 203, 1775–1783 (2011). 36. Navarro-Sanchez, E. et al. Dendritic-cell-specific ICAM3-grabbing non-integrin is essential for the

productive infection of human dendritic cells by mosquito-cell-derived dengue viruses. EMBO Rep. 4, 723– 728 (2003).

37. Chen, Y. C., Wang, S. Y. & King, C. C. Bacterial lipopolysaccharide inhibits dengue virus infection of primary human monocytes/macrophages by blockade of virus entry via a CD14-dependent mechanism. J.

Virol. 73, 2650–2657 (1999).

38. Van Der Schaar, H. M. et al. Dissecting the cell entry pathway of dengue virus by single-particle tracking in living cells. PLoS Pathog. 4, (2008).

39. Krishnan, M. N. et al. Rab 5 Is Required for the Cellular Entry of Dengue and West Nile Viruses. J. Virol. 81, 4881–4885 (2007).

40. Byk, L. A. et al. Dengue virus genome uncoating requires ubiquitination. MBio 7, (2016).

41. Zhang, Y. et al. Conformational changes of the flavivirus E glycoprotein. Structure 12, 1607–1618 (2004). 42. Yu, I. M. et al. Structure of the immature dengue virus at low pH primes proteolytic maturation. Science (80-.

). 319, 1834–1837 (2008).

43. Junjhon, J. et al. Differential Modulation of prM Cleavage, Extracellular Particle Distribution, and Virus Infectivity by Conserved Residues at Nonfurin Consensus Positions of the Dengue Virus pr-M Junction. J.

Virol. 82, 10776–10791 (2008).

44. Yu, I.-M. et al. Association of the pr Peptides with Dengue Virus at Acidic pH Blocks Membrane Fusion. J.

(21)

45. Wengler, G. & Wengler, G. Cell-associated West Nile flavivirus is covered with E+pre-M protein heterodimers which are destroyed and reorganized by proteolytic cleavage during virus release. J. Virol. 63, 2521–6 (1989).

46. Rodenhuis-Zybert, I. A., Wilschut, J. & Smit, J. M. Partial maturation: An immune-evasion strategy of dengue virus? Trends in Microbiology vol. 19 248–254 (2011).

47. Zybert, I. A., van der Ende-Metselaar, H., Wilschut, J. & Smit, J. M. Functional importance of dengue virus maturation: Infectious properties of immature virions. J. Gen. Virol. 89, 3047–3051 (2008).

48. Rodenhuis-Zybert, I. A. et al. Immature dengue virus: A veiled pathogen? PLoS Pathog. 6, e1000718 (2010). 49. Richter, M. K. S. et al. Immature dengue virus is infectious in human immature dendritic cells via interaction

with the receptor molecule DC-SIGN. PLoS One 9, (2014).

50. Solignat, M., Gay, B., Higgs, S., Briant, L. & Devaux, C. Replication cycle of chikungunya: A re-emerging arbovirus. Virology vol. 393 183–197 (2009).

51. Rupp, J. C., Sokoloski, K. J., Gebhart, N. N. & Hardy, R. W. Alphavirus RNA synthesis and non-structural protein functions. J. Gen. Virol. 96, 2483–2500 (2015).

52. Schnierle, B. S. Cellular attachment and entry factors for chikungunya virus. Viruses vol. 11 (2019). 53. Yap, M. L. et al. Structural studies of Chikungunya virus maturation. Proc. Natl. Acad. Sci. U. S. A. 114,

13703–13707 (2017).

54. Firth, A. E., Chung, B. Y. W., Fleeton, M. N. & Atkins, J. F. Discovery of frameshifting in Alphavirus 6K resolves a 20-year enigma. Virol. J. 5, (2008).

55. Zhang, R. et al. Mxra8 is a receptor for multiple arthritogenic alphaviruses. Nature 557, 570–574 (2018). 56. van Duijl-Richter, M. K. S., Hoornweg, T. E., Rodenhuis-Zybert, I. A. & Smit, J. M. Early events in

chikungunya virus infection—from virus cell binding to membrane fusion. Viruses vol. 7 3647–3674 (2015). 57. Van Duijl-Richter, M. K. S., Blijleven, J. S., van Oijen, A. M. & Smit, J. M. Chikungunya virus fusion

properties elucidated by single-particle and bulk approaches. J. Gen. Virol. 96, 2122–2132 (2015). 58. Singh, I. & Helenius, A. Role of ribosomes in Semliki Forest virus nucleocapsid uncoating. J. Virol. 66,

7049–7058 (1992).

59. Strauss, J. H. & Strauss, E. G. The alphaviruses: Gene expression, replication, and evolution. Microbiological

Reviews vol. 58 491–562 (1994).

60. Scholte, F. E. M. et al. Characterization of Synthetic Chikungunya Viruses Based on the Consensus Sequence of Recent E1-226V Isolates. PLoS One 8, (2013).

61. Utt, A. et al. Versatile trans-replication systems for chikungunya virus allow functional analysis and tagging of every replicase protein. PLoS One 11, (2016).

62. Schwartz, O. & Albert, M. L. Biology and pathogenesis of chikungunya virus. Nature Reviews Microbiology vol. 8 491–500 (2010).

63. Thomas, S. et al. Functional dissection of the alphavirus capsid protease: Sequence requirements for activity.

Virol. J. 7, 1–7 (2010).

64. Okeoma, C. M. Chikungunya virus: Advances in biology, pathogenesis, and treatment. Chikungunya Virus:

Advances in Biology, Pathogenesis, and Treatment (2016). doi:10.1007/978-3-319-42958-8.

65. Wong, K. Z. & Chu, J. J. H. The interplay of viral and host factors in chikungunya virus infection: Targets for antiviral strategies. Viruses (2018) doi:10.3390/v10060294.

66. Jessie, K., Fong, M. Y., Devi, S., Lam, S. K. & Wong, K. T. Localization of dengue virus in naturally infected human tissues, by immunohistochemistry and in situ hybridization. J. Infect. Dis. 189, 1411–1418 (2004).

(22)

1

67. Limon-Flores, A. Y. et al. Dengue virus inoculation to human skin explants: An effective approach to assess in

situ the early infection and the effects on cutaneous dendritic cells. Int. J. Exp. Pathol. 86, 323–334 (2005). 68. Wu, S. J. L. et al. Human skin Langerhans cells are targets of dengue virus infection. Nat. Med. (2000)

doi:10.1038/77553.

69. Couderc, T. & Lecuit, M. Chikungunya virus pathogenesis: From bedside to bench. Antiviral Research vol. 121 120–131 (2015).

70. Sourisseau, M. et al. Characterization of reemerging chikungunya virus. PLoS Pathog. 3, 0804–0817 (2007). 71. Schmid, M. A. & Harris, E. Monocyte Recruitment to the Dermis and Differentiation to Dendritic Cells

Increases the Targets for Dengue Virus Replication. PLoS Pathog. 10, (2014).

72. Singla, M. et al. Immune Response to Dengue Virus Infection in Pediatric Patients in New Delhi, India— Association of Viremia, Inflammatory Mediators and Monocytes with Disease Severity. PLoS Negl. Trop.

Dis. 10, (2016).

73. Wati, S., Li, P., Burrell, C. J. & Carr, J. M. Dengue virus (DV) replication in monocyte-derived macrophages is not affected by tumor necrosis factor alpha (TNF-alpha), and DV infection induces altered responsiveness to TNF-alpha stimulation. J. Virol. 81, 10161–71 (2007).

74. Michlmayr, D., Andrade, P., Gonzalez, K., Balmaseda, A. & Harris, E. CD14(+)CD16(+) monocytes are the main target of Zika virus infection in peripheral blood mononuclear cells in a paediatric study in Nicaragua.

Nat. Microbiol. (2017) doi:10.1038/s41564-017-0035-0.

75. Lo, Y.-L. et al. Dengue Virus Infection Is through a Cooperative Interaction between a Mannose Receptor and CLEC5A on Macrophage as a Multivalent Hetero-Complex. PLoS One 11, e0166474 (2016). 76. Schmid, M. A., Diamond, M. S. & Harris, E. Dendritic cells in dengue virus infection: Targets of virus

replication and mediators of immunity. Front. Immunol. 5, (2014).

77. Boonnak, K. et al. Role of Dendritic Cells in Antibody-Dependent Enhancement of Dengue Virus Infection.

J. Virol. 82, 3939–3951 (2008).

78. Sun, P. et al. Functional characterization of ex vivo blood myeloid and plasmacytoid dendritic cells after infection with dengue virus. Virology 383, 207–215 (2009).

79. Gardner, J. et al. Chikungunya Virus Arthritis in Adult Wild-Type Mice. J. Virol. 84, 8021–8032 (2010). 80. Her, Z. et al. Active Infection of Human Blood Monocytes by Chikungunya Virus Triggers an Innate

Immune Response. J. Immunol. 184, 5903–5913 (2010).

81. Correa, A. R. V. et al. Dengue virus directly stimulates Polyclonal B cell activation. PLoS One 10, (2015). 82. Lin, Y.-W. et al. Virus Replication and Cytokine Production in Dengue Virus-Infected Human B

Lymphocytes. J. Virol. 76, 12242–12249 (2002).

83. Silveira, G. F. et al. Human T Lymphocytes Are Permissive for Dengue Virus Replication. J. Virol. 92, e02181-17 (2018).

84. Simon, A. Y., Sutherland, M. R. & Pryzdial, E. L. G. Dengue virus binding and replication by platelets.

Blood 126, 378–385 (2015).

85. Avirutnan, P., Malasit, P., Seliger, B., Bhakdi, S. & Husmann, M. Dengue virus infection of human endothelial cells leads to chemokine production, complement activation, and apoptosis. J. Immunol. 161, 6338–46 (1998).

86. Suksanpaisan, L., Cabrera-Hernandez, A. & Smith, D. R. Infection of human primary hepatocytes with dengue virus serotype 2. J. Med. Virol. 79, 300–307 (2007).

87. Couvelard, A. et al. Report of a fatal case of dengue infection with hepatitis: Demonstration of dengue antigens in hepatocytes and liver apoptosis. Hum. Pathol. 30, 1106–1110 (1999).

(23)

88. Beutler, B. Innate immunity: An overview. Mol. Immunol. 40, 845–859 (2004).

89. Simon, A. K., Hollander, G. A. & McMichael, A. Evolution of the immune system in humans from infancy to old age. Proceedings of the Royal Society B: Biological Sciences vol. 282 (2015).

90. Dutertre, C.-A. et al. Single-Cell Analysis of Human Mononuclear Phagocytes Reveals Subset-Defining Markers and Identifies Circulating Inflammatory Dendritic Cells. Immunity 51, 573-589.e8 (2019). 91. Passlick, B., Flieger, D. & Ziegler-Heitbrock, L. Identification and characterization of a novel monocyte

subpopulation in human peripheral blood. Blood 74, 2527–2534 (1989).

92. Ziegler-Heitbrock, L. et al. Nomenclature of monocytes and dendritic cells in blood. Blood vol. 116 (2010). 93. Patel, A. A. et al. The fate and lifespan of human monocyte subsets in steady state and systemic inflammation.

J. Exp. Med. 214, 1913–1923 (2017).

94. Wong, K. L. et al. Gene expression profiling reveals the defining features of the classical, intermediate, and nonclassical human monocyte subsets. Blood 118, 16–32 (2011).

95. Yasaka, T., Manitich, N. M., Boxer, L. A. & Baehner, R. L. Functions of human monocyte and lymphocyte subsets obtained by countercurrent centrifugal elutriation: Differing functional capacities of human monocyte subsets. J. Immunol. 127, 1515–1518 (1981).

96. Kurihara, T., Warr, G., Loy, J. & Bravo, R. Defects in macrophage recruitment and host defense in mice lacking the CCR2 chemokine receptor. J. Exp. Med. 186, 1757–1762 (1997).

97. Shi, C. & Pamer, E. G. Monocyte recruitment during infection and inflammation. Nature Reviews

Immunology vol. 11 762–774 (2011).

98. Yang, J., Zhang, L., Yu, C., Yang, X. F. & Wang, H. Monocyte and macrophage differentiation: Circulation inflammatory monocyte as biomarker for inflammatory diseases. Biomarker Research vol. 2 (2014). 99. Thomas, G., Tacke, R., Hedrick, C. C. & Hanna, R. N. Nonclassical Patrolling Monocyte Function in the

Vasculature. Arterioscler. Thromb. Vasc. Biol. 35, 1306–1316 (2015).

100. Auffray, C. et al. Monitoring of blood vessels and tissues by a population of monocytes with patrolling behavior. Science (80-. ). 317, 666–670 (2007).

101. Olingy, C. E. et al. Non-classical monocytes are biased progenitors of wound healing macrophages during soft tissue injury. Sci. Rep. 7, (2017).

102. Arora, S., Dev, K., Agarwal, B., Das, P. & Syed, M. A. Macrophages: Their role, activation and polarization in pulmonary diseases. Immunobiology vol. 223 383–396 (2018).

103. Martinez, F. O. & Gordon, S. The M1 and M2 paradigm of macrophage activation: Time for reassessment.

F1000Prime Rep. 6, (2014).

104. Rhodes, J. W., Tong, O., Harman, A. N. & Turville, S. G. Human dendritic cell subsets, ontogeny, and impact on HIV infection. Frontiers in Immunology vol. 10 1088 (2019).

105. Villani, A. C. et al. Single-cell RNA-seq reveals new types of human blood dendritic cells, monocytes, and progenitors. Science (80-. ). 356, (2017).

106. Bourdely, P. et al. Transcriptional and Functional Analysis of CD1c+ Human Dendritic Cells Identifies a CD163+ Subset Priming CD8+CD103+ T Cells. Immunity 53, 335-352.e8 (2020).

107. Brubaker, S. W., Bonham, K. S., Zanoni, I. & Kagan, J. C. Innate Immune Pattern Recognition: A Cell Biological Perspective. Annu. Rev. Immunol. 33, 257–290 (2015).

108. Lee, C. C., Avalos, A. M. & Ploegh, H. L. Accessory molecules for Toll-like receptors and their function.

Nature Reviews Immunology vol. 12 168–179 (2012).

109. Chow, J., Franz, K. M. & Kagan, J. C. PRRs are watching you: Localization of innate sensing and signaling regulators. Virology 479–480, 104–109 (2015).

(24)

1

110. Pandey, S., Kawai, T. & Akira, S. Microbial Sensing by Toll-Like Receptors and Intracellular Nucleic Acid

Sensors. Cold Spring Harb Perspect Biol 7, a016246 (2015).

111. Jensen, S. & Thomsen, A. R. Sensing of RNA Viruses: a Review of Innate Immune Receptors Involved in Recognizing RNA Virus Invasion. J. Virol. 86, 2900–2910 (2012).

112. Stein, D., Roth, S., Vogelsang, E. & Nüsslein-Volhard, C. The polarity of the dorsoventral axis in the drosophila embryo is defined by an extracellular signal. Cell 65, 725–735 (1991).

113. Fitzgerald, K. A. & Kagan, J. C. Toll-like Receptors and the Control of Immunity. Cell vol. 180 1044–1066 (2020).

114. Kieser, K. J. & Kagan, J. C. Multi-receptor detection of individual bacterial products by the innate immune system. Nat. Rev. Immunol. 17, 376–390 (2017).

115. Molema, G. Heterogeneity in endothelial responsiveness to cytokines, molecular causes, and pharmacological consequences. Seminars in Thrombosis and Hemostasis vol. 36 246–264 (2010).

116. Mehta, D. & Malik, A. B. Signaling mechanisms regulating endothelial permeability. Physiological Reviews vol. 86 279–367 (2006).

117. Dejana, E., Tournier-Lasserve, E. & Weinstein, B. M. The Control of Vascular Integrity by Endothelial Cell Junctions: Molecular Basis and Pathological Implications. Developmental Cell vol. 16 209–221 (2009). 118. Uchimido, R., Schmidt, E. P. & Shapiro, N. I. The glycocalyx: A novel diagnostic and therapeutic target in

sepsis. Critical Care vol. 23 (2019).

119. Leligdowicz, A., Richard-Greenblatt, M., Wright, J., Crowley, V. M. & Kain, K. C. Endothelial activation: The Ang/Tie axis in sepsis. Frontiers in Immunology vol. 9 (2018).

120. Aird, W. C. The role of the endothelium in severe sepsis and multiple organ dysfunction syndrome. Blood vol. 101 3765–3777 (2003).

121. Joffre, J., Hellman, J., Ince, C. & Ait-Oufella, H. Endothelial Responses in Sepsis. Am. J. Respir. Crit. Care

Med. (2020) doi:10.1164/rccm.201910-1911tr.

122. Fiedler, U. et al. Angiopoietin-2 sensitizes endothelial cells to TNF-α and has a crucial role in the induction of inflammation. Nat. Med. 12, 235–239 (2006).

123. Vestweber, D. How leukocytes cross the vascular endothelium. Nature Reviews Immunology vol. 15 692– 704 (2015).

124. Harlan, J. M. & Winn, R. K. Leukocyte-endothelial interactions: Clinical trials of anti-adhesion therapy. in

Critical Care Medicine vol. 30 (2002).

125. Van Buul, J. D., Kanters, E. & Hordijk, P. L. Endothelial signaling by Ig-like cell adhesion molecules.

Arteriosclerosis, Thrombosis, and Vascular Biology vol. 27 1870–1876 (2007).

126. Aird, W. C. Phenotypic heterogeneity of the endothelium: I. Structure, function, and mechanisms. Circulation

Research vol. 100 158–173 (2007).

127. WHO (World Health Organisation). Dengue haemorrhagic fever: diagnosis, treatment, prevention and control. 2nd edition. Geneva World Heal. Organ. 103–117 (1997).

128. WHO. Dengue: guidelines for diagnosis, treatment, prevention and control; 2009. Geneva World Heal.

Organ. Google Sch. (2009).

129. Yacoub, S., Wertheim, H., Simmons, C. P., Screaton, G. & Wills, B. Cardiovascular manifestations of the emerging dengue pandemic. Nature Reviews Cardiology vol. 11 335–345 (2014).

130. van de Weg, C. A. M. et al. Lipopolysaccharide levels are elevated in dengue virus infected patients and correlate with disease severity. J. Clin. Virol. 53, 38–42 (2012).

(25)

131. van de Weg, C. A. M. et al. Microbial Translocation Is Associated with Extensive Immune Activation in Dengue Virus Infected Patients with Severe Disease. PLoS Negl. Trop. Dis. 7, e2236 (2013).

132. Teixeira, M. G. et al. Arterial Hypertension and Skin Allergy Are Risk Factors for Progression from Dengue to Dengue Hemorrhagic Fever: A Case Control Study. PLoS Negl. Trop. Dis. 9, (2015).

133. Toledo, J. et al. Relevance of Non-communicable Comorbidities for the Development of the Severe Forms of Dengue: A Systematic Literature Review. PLoS Negl. Trop. Dis. 10, (2016).

134. Lee, I. K., Hsieh, C. J., Lee, C. Te & Liu, J. W. Diabetic patients suffering dengue are at risk for development of dengue shock syndrome/severe dengue: Emphasizing the impacts of co-existing comorbidity(ies) and glycemic control on dengue severity. J. Microbiol. Immunol. Infect. 53, 69–78 (2018).

135. Khandia, R. et al. Modulation of Dengue/Zika Virus pathogenicity by antibody-dependent enhancement and strategies to protect against enhancement in Zika Virus infection. Frontiers in Immunology vol. 9 (2018). 136. Kwissa, M. et al. Dengue virus infection induces expansion of a CD14+CD16 + monocyte population that

stimulates plasmablast differentiation. Cell Host Microbe 16, 115–127 (2014).

137. Azeredo, E. L. et al. Differential regulation of toll-like receptor-2, toll-like receptor-4, CD16 and human leucocyte antigen-DR on peripheral blood monocytes during mild and severe dengue fever. Immunology 130, 202–216 (2010).

138. Arboleda Alzate, J. F., Rodenhuis-Zybert, I. A., Hernández, J. C., Smit, J. M. & Urcuqui-Inchima, S. Human macrophages differentiated in the presence of vitamin D3restrict dengue virus infection and innate responses by downregulating mannose receptor expression. PLoS Negl. Trop. Dis. 11, (2017).

139. Flipse, J. et al. Antibody-Dependent Enhancement of Dengue Virus Infection in Primary Human Macrophages; Balancing Higher Fusion against Antiviral Responses. Sci. Rep. 6, (2016).

140. Décembre, E., Assil, S., Hillaire, M. L. B., Dejnirattisai, W. & Mongkolsapaya, J. Sensing of Immature Particles Produced by Dengue Virus Infected Cells Induces an Antiviral Response by Plasmacytoid Dendritic Cells. PLoS Pathog 10, 1004434 (2014).

141. Sun, P. et al. Functional characterization of ex vivo blood myeloid and plasmacytoid dendritic cells after infection with dengue virus. Virology 383, 207–215 (2009).

142. Gandini, M. et al. Dengue Virus Activates Membrane TRAIL Relocalization and IFN-α Production by Human Plasmacytoid Dendritic Cells In Vitro and In Vivo. PLoS Negl. Trop. Dis. 7, (2013).

143. Halstead, S. B. & O’Rourke, E. J. Dengue viruses and mononuclear phagocytes. I. Infection enhancement by non-neutralizing antibody. J. Exp. Med. 146, 201–17 (1977).

144. Chunhakan, S., Butthep, P., Yoksan, S., Tangnararatchakit, K. & Chuansumrit, A. Vascular leakage in dengue hemorrhagic fever is associated with dengue infected monocytes, monocyte activation/exhaustion, and cytokines production. Int. J. Vasc. Med. 2015, 1–9 (2015).

145. Hottz, E. D. et al. Platelet activation and apoptosis modulate monocyte inflammatory responses in dengue. J

Immunol 193, 1864–1872 (2014).

146. Mapalagamage, M. et al. High levels of serum angiopoietin 2 and angiopoietin 2/1 ratio at the critical stage of dengue hemorrhagic fever in patients and association with clinical and biochemical parameters. J. Clin.

Microbiol. 58, (2020).

147. Malavige, G. N. & Ogg, G. S. Pathogenesis of vascular leak in dengue virus infection. Immunology vol. 151 261–269 (2017).

148. Lin, S. W. et al. Dengue virus nonstructural protein NS1 binds to prothrombin/thrombin and inhibits prothrombin activation. J. Infect. 64, 325–334 (2012).

149. Tang, T. H. C. et al. Increased Serum Hyaluronic Acid and Heparan Sulfate in Dengue Fever: Association with Plasma Leakage and Disease Severity. Sci. Rep. 7, 1–9 (2017).

Referenties

GERELATEERDE DOCUMENTEN

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

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

Considering there is no adequate animal model that mimics DENV and CHIKV infections in humans, combining in vitro studies in human peripheral blood cells with cohort studies will be

Chapter 2 Suppression of chikungunya virus replication and differential innate responses of human peripheral blood mononuclear cells during co- infection with dengue

To our knowledge this is the first systematic study that addresses the viral kinetics and innate immune responses upon DENV and CHIKV co-infection performed in primary human

Here, we used a previously established in vitro model of CHIKV infection in peripheral blood mononuclear cells to elucidate the mechanism and effect of IM expansion in the acute

Surprisingly, blocking TLR2 on PBMCs prior to exposure to DENV infection abolished inflammatory responses and significantly attenuated endothelial activation as

They serve as a niche for viral replication, vehicles for viral spread and contribute to severe disease development by sensing DENV and CHIKV infection leading to the production of