Next Article in Journal
Resistance to Rhabdoviridae Infection and Subversion of Antiviral Responses
Next Article in Special Issue
Exosomes: Implications in HIV-1 Pathogenesis
Previous Article in Journal
Ultra Deep Sequencing of a Baculovirus Population Reveals Widespread Genomic Variations
Previous Article in Special Issue
Exosomes and Their Role in the Life Cycle and Pathogenesis of RNA Viruses
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Early Events in Chikungunya Virus Infection—From Virus CellBinding to Membrane Fusion

by
Mareike K. S. Van Duijl-Richter
,
Tabitha E. Hoornweg
,
Izabela A. Rodenhuis-Zybert
and
Jolanda M. Smit
*
Department of Medical Microbiology, University of Groningen and University Medical Center Groningen, 9700 RB Groningen, The Netherlands
*
Author to whom correspondence should be addressed.
Viruses 2015, 7(7), 3647-3674; https://doi.org/10.3390/v7072792
Submission received: 4 May 2015 / Revised: 24 June 2015 / Accepted: 29 June 2015 / Published: 7 July 2015
(This article belongs to the Special Issue Viruses and Exosomes)

Abstract

:
Chikungunya virus (CHIKV) is a rapidly emerging mosquito-borne alphavirus causing millions of infections in the tropical and subtropical regions of the world. CHIKV infection often leads to an acute self-limited febrile illness with debilitating myalgia and arthralgia. A potential long-term complication of CHIKV infection is severe joint pain, which can last for months to years. There are no vaccines or specific therapeutics available to prevent or treat infection. This review describes the critical steps in CHIKV cell entry. We summarize the latest studies on the virus-cell tropism, virus-receptor binding, internalization, membrane fusion and review the molecules and compounds that have been described to interfere with virus cell entry. The aim of the review is to give the reader a state-of-the-art overview on CHIKV cell entry and to provide an outlook on potential new avenues in CHIKV research.

1. Introduction

Chikungunya virus (CHIKV) is an arbovirus transmitted by mosquitoes of the Aedes (Ae.) species. Upon infection, about 75%–95% of the individuals develop Chikungunya fever, characterized by high fever, myalgia, joint pain, rash, and intense asthenia [1,2]. A common long-term complication (occurring in 12%–49% of patients) is severe, debilitating joint paint that can persist for months to years after infection [3]. Furthermore, in rare cases, encephalopathy, encephalitis, myocarditis, hepatitis, and circulatory failure is seen [4,5].
Previously, CHIKV caused small outbreaks in confined regions within Africa and Asia. This situation drastically changed by the end of 2004 when the first major CHIKV outbreak started [6]. Since then, the virus has spread globally with millions of people infected. To date, CHIKV is epidemic in large parts of Africa, Asia, and the tropical regions of the Americas [7]. Within the last 1.5 years, the virus has spread to more than 40 countries within Central America involving over 1 million CHIKV infections [8]. There are four CHIKV lineages—the West African (WA) lineage, the Asian lineage, the Eastern/Central/Southern Africa (ECSA) lineage, and the Indian Ocean lineage (IOL); the latter emerged from the ECSA lineage in 2004 [9,10]. Some IOL strains adapted to a new vector, Ae. albopictus, without significantly compromising viral fitness for the initial vector Ae. aegypti, thereby increasing the epidemic potential of the virus. Functional studies revealed that this is caused by adaptive mutations within the viral spike proteins E1 and E2 of CHIKV [11,12]. The IOL lineage caused the majority of CHIKV outbreaks in 2004–2012, whereas the Asian lineage and ECSA lineage are mainly responsible for the current outbreaks in the Americas [10,13].
There is currently no vaccine nor a specific antiviral treatment available to prevent or treat CHIKV infection. A potential antiviral strategy involves the inhibition of the cell entry process of the virus. CHIKV cell entry is based on a series of dynamic events between the viral glycoproteins E1 and E2 and the host cell, including virus-cell attachment, virus internalization, intracellular trafficking, and membrane fusion. In this review we will describe the current knowledge related to the cell tropism, the cell entry pathway of CHIKV and will discuss the molecules that have been identified to interfere with these processes.

2. Viral Structure

CHIKV belongs to the alphavirus genus within the Togaviridae family. It is a member of the antigenic Semliki Forest Complex, which include, amongst others, the closely related O’nyong-nyong virus (ONNV), Semliki Forest virus (SFV), and Ross River virus (RRV). Other alphaviruses are for example Sindbis virus (SINV) and Venezuelan or Eastern Equine Encephalitis Virus (VEEV and EEEV, respectively) [14,15]. To date, most studies have been performed with SFV, SINV, RRV, and VEEV. Alphaviruses are enveloped spherical particles with a diameter of 65–70 nm [16,17]. The alphavirus genome consists of a single-stranded positive-sensed 11.8 kB RNA molecule packaged by the C protein to form the nucleocapsid. This nucleocapsid is surrounded by a host-cell derived lipid bilayer with two inserted transmembrane glycoproteins, E1 and E2 [15]. The composition of the host-cell derived lipid bilayer strongly resembles the plasma membrane of the infected host cell. For mammalian-derived CHIKV virions, the membrane consists of cholesterol and phospholipids in a ratio of approximately 1:1 [18,19,20,21].
The E1 protein is 439 amino acids (aa) long and contains one conserved N-linked glycosylation site at position 141 [22]. E1 is anchored in the lipid bilayer with a 30 residue transmembrane helix at the carboxy-terminal end. The cytosolic region is only five residues in length and does not interact with the nucleocapsid [22,23]. The N-terminal ectodomain of E1 consists of 404 residues and is structurally divided into three β-barrel domains named DI, DII, and DIII (Figure 1a).
Figure 1. Structure of the E2/E1 dimer. (A) Ribbon diagram showing the ectodomains of the CHIKV E1 and E2 glycoprotein ([22]; PDB 3N41). The structural domains I, II, and III of E1 are shown in blue, red and yellow, respectively. E2 domain A, B, and C are designated in cyan, green, and pink, respectively. In the mature virion, the E1 fusion loop (E1-FL, orange), is covered by a binding groove between E2 domain A and B. The β-ribbon of E2 containing the acid-sensitive region is highlighted in dark purple. Within this region, the hydrogen bond between E2-H170 and E1-S57 stabilizes the E2/E1 dimer interaction at neutral pH [26,142]. E1-A/V226 and E1-V178 are important for lipid sensing before fusion [12,154]. The black arrow points towards the viral membrane; (B,C) Surface view (PDB 2XFC) of one virus spike from the side (B) and the top (C). E1 is depicted in the same colors as in the ribbon diagram, E2 is depicted in gray for clarity. This figure was prepared using the program PyMOL.
Figure 1. Structure of the E2/E1 dimer. (A) Ribbon diagram showing the ectodomains of the CHIKV E1 and E2 glycoprotein ([22]; PDB 3N41). The structural domains I, II, and III of E1 are shown in blue, red and yellow, respectively. E2 domain A, B, and C are designated in cyan, green, and pink, respectively. In the mature virion, the E1 fusion loop (E1-FL, orange), is covered by a binding groove between E2 domain A and B. The β-ribbon of E2 containing the acid-sensitive region is highlighted in dark purple. Within this region, the hydrogen bond between E2-H170 and E1-S57 stabilizes the E2/E1 dimer interaction at neutral pH [26,142]. E1-A/V226 and E1-V178 are important for lipid sensing before fusion [12,154]. The black arrow points towards the viral membrane; (B,C) Surface view (PDB 2XFC) of one virus spike from the side (B) and the top (C). E1 is depicted in the same colors as in the ribbon diagram, E2 is depicted in gray for clarity. This figure was prepared using the program PyMOL.
Viruses 07 02792 g001
DIII is situated at the C-terminus of the protein and closest to the envelope, followed by the central DI and DII at the tip, which contains a hydrophobic fusion peptide [22,24,25,26]. The E2 protein has a length of 423 aa and is N-glycosylated at positions 263 and 345 [22]. Blast analyses revealed that like for E1, the glycosylation sites of E2 are conserved between all four CHIKV lineages (GenBank accession numbers strain RSU1: HM045797.1; strain IbH35: HM045786.1; strain S27: AF369024.2; strain LR2006 OPY: DQ443544.2). At the C-terminus, a transmembrane helix of 26 residues is located, followed by a cytoplasmic domain of 33 residues. The cytoplasmic domain contacts the nucleocapsid and studies with other alphaviruses have shown that this interaction is important for the correct assembly and budding of progeny viruses from the plasma membrane of infected cells [22,27,28]. The ectodomain of E2 has a size of 364 aa and consists of three immunoglobulin-fold domains termed A, B, and C, which are connected by a long β-ribbon (Figure 1a) [26,29].
On a mature virion, 240 copies of E1 and E2 are arranged as 80 trimeric spikes; a single spike consisting of three E2/E1 heterodimers (Figure 1b,c). The spikes are positioned in an icosahedral T = 4 symmetry and form a continuous protein shell around the particle [22,26,30,31,32]. Within the E2/E1 heterodimer, E1 laterally contacts E2 along the central domain II and partially domain III. The E1 hydrophobic fusion peptide is buried in a groove between domain A and domain B of E2 (see Figure 1), thereby preventing pre-mature activation of the membrane fusion machinery of the virus [26].

3. Viral Tropism

Infection starts when a CHIKV-infected Ae. mosquito is feeding on a human host [33]. During feeding, CHIKV particles are thought to be released within the dermis and into the subcutaneous capillaries of the skin [34]. Within 2–4 days, the virus reaches the blood and disseminates to other parts of the body. Although CHIKV pathogenesis is still poorly understood, recent studies shed light onto the organs and cells involved in CHIKV replication (systematically reviewed by [35]). The CHIKV target organs include joints, muscle, skin, and less frequently, the liver, kidneys, eye and the central nervous system (CNS). Infection of these organs is frequently associated with a marked infiltration of mononuclear cells such as monocytes/macrophages. The virus tropism described within this section is mostly based on studies using ECSA and IOL strains. A few studies directly compared the infectivity of IOL, WA and ESCA on multiple cell lines and revealed that these viruses exhibit a comparable tropism [36,37,38]. However, more studies are required to determine the exact tropism for all four CHIKV lineages.

3.1. Viremia—Where Is the Virus Produced?

During the 7–12 days-long acute viremic period, CHIKV load can reach 109–1012 viral particles per milliliter [39,40,41]. The observation that CHIKV reaches a high titer in a relatively short time period is suggestive for replication in blood leukocytes [42]. Indeed, other alphaviruses replicate in immune cells including dendritic cells (e.g., SFV, RRV, and VEEV) and monocytes (e.g., RRV and VEEV) [36,43,44,45,46]. In contrast to the above-mentioned alphaviruses, peripheral blood mononuclear cells (PBMCs) do not seem to contribute significantly to the production of CHIKV progeny [36,47]. In fact, in vitro analysis revealed that most blood-derived cell types such as lymphocytes, dendritic cells, and natural killer cells are refractory to CHIKV infection [36,37]. Conflicting reports were published on the permissiveness of monocytes to CHIKV infection [36,42]. However, it is clear that even though monocytes might harbor CHIKV antigens, viral production supported by the primary cultures of monocytes cannot explain the titers detected in blood of acute phase patients. These observations suggest that local CHIKV replication in dermal fibroblasts, migrating monocytes/macrophages, and endothelial cells are pivotal for virus production. Indeed, in vitro studies revealed these cells are much more permissive to CHIKV infection [36,37,48,49].

3.2. Arthrotropism of CHIKV

Mononuclear cell infiltration and viral replication in the muscles (particularly skeletal muscle progenitor cells, not muscle fibers) and joints (in fibroblasts of the joint capsule and presumably in osteoblasts) are associated with debilitating arthralgia, myalgia, and in some cases, arthritis [50,51,52,53,54,55]. While the acute phase symptoms usually resolve within two weeks, the musculoskeletal pain may linger for weeks to months or even years [56,57,58,59,60]. Chronic disease has been linked to persistent virus replication in the target cells and/or the establishment of a self-sustained inflammatory mechanism that leads to the tissue damage (for more details on this topic see [3,61,62]). In vivo, synovial macrophages and satellite muscle cells have been shown to contain viral RNA or protein months after infection [63,64]. However, the exact mechanisms underlying CHIKV persistence in tissue sanctuaries are still ill-understood and more studies in animal models and clinical investigations are needed to address this issue.

3.3. Less Common Tropism of CHIKV

CHIKV is classified as an Old World arthritogenic alphavirus and hence is not expected to be neurotropic or encephalitogenic [14]. However, cases of Guillain-Barré syndrome and encephalitis have been reported following CHIKV infection [65,66,67]. CHIKV and anti-CHIKV IgM has been detected in the blood-cerebrospinal fluid of human neonates and adult patients with encephalopathy [68]. Studies in mice suggest that CHIKV particles can enter the CNS via the Virchow-Robin spaces and choroid plexuses [55]. Thereafter, replication occurs within choroid plexus epithelial cells, leptomeninges, and ependymal cells but not in brain parenchyma [4]. In vitro, however, parenchymal cells including neurons, astrocytes, microglial cells, and neuroblastoma cells were found to be permissive to all strains tested (including ECSA, WA, and IOL) [4,37,38,69,70,71]. The ability of CHIKV to infect brain endothelial cells is still under debate as human brain microvascular endothelial cells can be infected, while primary brain endothelial cells [55] and the brain endothelial cell line hCMEC/D3 [36] were found to be refractory to infection by IOL isolates. The mechanism leading to CHIKV infection of the CNS in humans is to be elucidated.

4. Cell Entry and Membrane Fusion

4.1. Receptor Binding

The first step in infection involves binding of the virus to a host cell receptor [72,73]. Based on the wide range of cell types CHIKV infects in vivo and in vitro, the cellular receptor of CHIKV is likely to be ubiquitously expressed among species and cell types. Receptor binding is facilitated by the E2 glycoprotein of CHIKV [74,75]. Both domain A and domain B of the E2 protein contain putative receptor binding sites [26,29]. Furthermore, bioinformatic analysis revealed that E2 domain B contains a class III PDZ binding motif [76]. These motifs have been described to mediate protein-protein interactions [77,78].
To date, prohibitin (PHB), phosphatidylserine (PtdSer)-mediated virus entry-enhancing receptors (PVEERs), and glycosaminoglycans (GAGs) have been suggested as CHIKV receptor proteins in mammalian cells [69,79,80] and ATPsynthase β subunit in mosquito cells [81]. Notably, CHIKV infection can proceed in absence of these proteins, indicating that these proteins facilitate the initial interaction with the cell surface rather than virus uptake [82,83].

4.1.1. Prohibitin

Prohibitins (PHBs) are evolutionary conserved multifunctional membrane proteins, which are present in multiple cellular compartments [84]. PHBs play a role in for example mitochondrial integrity, cell proliferation, cell survival and endocytosis in white adipose tissue [85,86,87]. Importantly, PHBs are ubiquitously expressed at the cell surface of numerous mammalian cells [77,78]. Wintachai and co-workers showed that anti-PHB antibodies and siRNAs towards PHB reduced CHIKV infection of microglial cells up to two-fold. CHIKV was also found to bind to PHB in U937 cells, but despite this interaction the cells did not support a productive infection [69]. On the other hand, flavaglines, plant compounds that directly interact with PHB did inhibit CHIKV infection in HEK-293T cells for up to 50% [88]. Thus, it is clear that PHB facilitates virus-cell binding. PHB likely acts to capture and concentrate CHIKV particles at the cell surface. However, since the inhibiting compounds only moderately reduced infectivity and that U937 cells are refractory to CHIKV despite PHB binding demonstrates that other factors are required to mediate (efficient) infection. The precise role of PHB in CHIKV cell entry remains to be elucidated.

4.1.2. Phosphatidylserine (PtdSer)-Mediated Virus Entry-enhancing Receptors

T-cell immunoglobulin and mucin domain (TIM) family members are expressed on various immune cells and a range of mucosal epithelia, and are known to regulate immune cell activity [89,90,91,92]. Recently, TIM-1 was described to enhance the entry and infection of chimeric virus particles displaying the glycoproteins of CHIKV or other viruses in HEK293T cells [79]. TIM-1 binds to phosphatidylserine (PtdSer) in the viral envelope and functions to concentrate the virus at the cell surface. These receptors act on the basis of their long stalk region and PtdSer binding motif. Indeed, other unrelated proteins with a long stalk region and PtdSer motif were also able to support viral cell entry, demonstrating that virus-cell binding is not TIM-1 specific [93]. These results further indicate that CHIKV uses TIM-1 as an attachment factor but not as a specific receptor.

4.1.3. Glycosaminoglycans

Glycosaminoglycans (GAGs) are large complex carbohydrate molecules that are expressed at the cell surface of most mammalian cell types. GAGs include among others heparan sulfate, keratan sulfate, chondroitin sulfate, and dermatan sulfate [86]. These molecules can bind a wide variety of proteins and mainly function in cellular adhesion, growth, differentiation, and signaling [94]. Several alphaviruses are known to use GAGs for cell entry [95,96,97,98,99]. Natural isolates of EEEV and low passage strains of VEEV were found to depend on GAGs for efficient infection of cells [97,100]. For other alphaviruses, heparan sulfate binding was related to virus-cell culture adaptation [95,96,101] and an attenuated disease phenotype in mice [95,96].
For CHIKV, GAG expression was found to increase the binding and infection efficiency of both a clinical and a vaccine strain in CHO cells. However, GAG binding is not a property of all CHIKV strains, as CHIKV-LR replicon particles do not require cell-surface GAGs for infection [80,102]. Yet, CHIKV like other alphaviruses readily adapts to GAGs [103,104]. GAG utilization is facilitated by mutations to positively charged amino acids at E2-82 and E2-79 [80,104]. For example, an arginine at E2-82 or a lysine at E2-79 leads to enhanced infectivity in mammalian cells and attenuated virulence in mice [75,104,105]. Though, the observation that a clinical strain utilizes GAGs for cell entry suggests that GAGs might also play a role in natural infection.

4.1.4. ATP Synthase β Subunit

ATP synthase β subunit (ATPSβ) was recently found to interact with CHIKV in mosquito cells. Furthermore, ATPSβ-down-regulation significantly reduced viral entry and virus production [81]. The ATPSβ gene is widely conserved and is for example expressed in human endothelial and hepatic cells [106,107,108]. Although involved in F1/ATPase catalysis in the mitochondria, ATPS is also located at the surface of the plasma membrane. There, it can bind ligands as apolipoprotein A-I, apolipoprotein E and angiostatin [94]. Therefore, it is of interest to examine whether this protein is involved in CHIKV entry in mammalian cells and whether ATPSβ also exerts its function via increasing attachment of virions to the cell surface or whether other mechanisms are involved.

4.1.5. Other CHIKV Receptors Candidates

Another potential CHIKV receptor is the aV integrin (ITGAV) and b1 integrin (ITGB1) dimer, consisting of two members of the integrin superfamily. Integrin superfamily members form various transmembrane dimers, which function as cell adhesion receptors binding to different extracellular ligands [109]. The aV integrin (ITGAV) and b1 integrin (ITGB1) dimer was found to be differentially expressed in the brain proteasome of mice early in CHIKV infection. A direct effect of the integrin dimer on CHIKV infection has not been studied yet [110]. However, these protein dimers were previously reported as an adenovirus receptor [111,112] and other members of the integrin superfamily serve as receptors for RRV and West Nile virus [113,114]. As rare cases of neuropathology have been described, it would be of interest to investigate if CHIKV infection of the CNS is facilitated by integrin dimers.
Additionally, Heat shock protein 60 (HsP60) is a CHIKV receptor candidate. HsP60 is mainly known as a mitochondrial molecular chaperone which is involved in protein folding [115]. The protein has also been detected at the cell surface of murine monocytes/macrophage, B lymphocytes and T lymphocytes and human T lymphocytes [116,117]. HsP60 was found to interact with CHIKV by a two dimensional Virus Overlay Protein Binding Assay (2D-VOPBA) [69,118]. HsP60 was previously implicated in DENV infection [119], but, thus far, no functional proof on the role of HsP60 in CHIKV entry has been explored.

4.2. CHIKV Cell Entry and Membrane Fusion

Alphaviruses are generally internalized via clathrin-mediated endocytosis (CME), though also direct fusion with the plasma membrane has been described for SINV [120,121]. CME is a constitutive process within mammalian cells [82,122]. Invagination and scission of the membrane to form a virus-containing clathrin-coated vesicle occurs via a complex interplay of several proteins including adaptor protein-2, dynamin, clathrin, epsin, and Eps15 (see for extensive review [123]). Thereafter, the clathrin-coated vesicle is transported inside the cell after which the clathrin molecules dissociate and the virus is delivered to endosomes. The low-pH environment of the endosomes subsequently triggers conformational changes within the E1/E2 glycoproteins to mediate fusion of the viral membrane with the endosomal membrane (see Figure 2). Alphavirus fusion usually occurs from within mildly acidic early endosomes. VEEV fusion on the other hand has been described to fuse from within late endosomes [124,125,126,127].
For CHIKV, contradicting observations were reported. CHIKV infection was found to be dependent on dynamin [36], a large multidomain GTPase driving the pinching of endocytic vesicles from the plasma membrane [123]. Dynamin is an important mediator of CME and caveolar endocytosis [127]; and is also described to act in phagocytosis [128]. Furthermore, CHIKV infection was found to be mediated by Eps15 [129], a molecule essential for the assembly of the clathrin-coated pits [123]. Involvement of Eps15 can however not confirm entry via CME, as Eps15 has also been implicated in clathrin-independent entry pathways [130]. Specific inhibitors like siRNAs against the clathrin heavy chain did not inhibit CHIKV infection in HEK239T cells [129], but did show a marked reduction in infectivity in human umbilical vein endothelial cells (HUVEC), the cell line U-2 OS, and primary human umbilical vein endothelial cells [131]. Moreover, we recently showed that Pitstop2, a biochemical inhibitor of CME, reduced CHIKV cell entry in BS-C-1 cells [132]. Furthermore, using live-cell microscopy, we determined that approximately 90% of all particles that fused entered through CME. Taken together, thus far, a limited number of studies has been performed to investigate the cell entry pathway of CHIKV and most of these studies point towards entry through CME although clathrin-independent entry is also reported. The entry pathway taken by the virus maybe cell-specific. Alternatively, CHIKV has the capacity to infect cells via multiple pathways. The latter is supported by the fact that none of the inhibitor strategies applied so far completely blocked CHIKV infection [36,129,131,132]. In the absence of cellular perturbations, CHIKV was shown to enter via CME and therefore we hypothesize that CME, like for other alphaviruses, is the main pathway exploited by CHIKV.
Upon endocytosis, the virus is delivered to early endosomes. Approximately 40% of the particles fuse within 10 seconds after delivery to the endosome. More than 95% of all CHIKV fusion events occurred from within in early endosomal compartments [132]. This is in line with data from Bernard et al. who showed that CHIKV infection is dependent on early endosomes, but not on late endosomes [129]. In mosquito cells, however, CHIKV infection was dependent on the integrity of both Rab5 and Rab7-positive endosomes, which is suggestive of fusion from within maturing or late endosomes [133]. Variability in endosomal pH between cells [134] may explain this discrepancy. Biophysical analysis revealed that the pH threshold for CHIKV fusion lies—depending on the viral strain—between pH 6.2 and 5.9 [132,135]. Thus, we postulate that fusion is triggered once the pH of the endosomal lumen is below the threshold for membrane fusion. It is highly unlikely that other processes are involved, especially considering the rapid kinetics of membrane fusion within Rab5-positive endosomes.
Figure 2. Chikungunya virus cell entry and potential antiviral strategies. The viral life cycle starts with attachment of the virus particle to one of the ubiquitously expressed attachment factors or receptors at the cell surface (1); Subsequently, the virus is internalized into the cell via clathrin-mediated endocytosis (2); Then, clathrin-molecules dissociate from the vesicle and the virus is delivered to Rab5+ endosomes. Within the mildly acidic lumen of the endosome, the viral glycoproteins E2 and E1 undergo major conformational changes that lead to membrane fusion (3); Thereafter, the nucleocapsid core is released into the cytosol (4). The molecules and compounds that are known to interfere with entry are stated in the boxes.
Figure 2. Chikungunya virus cell entry and potential antiviral strategies. The viral life cycle starts with attachment of the virus particle to one of the ubiquitously expressed attachment factors or receptors at the cell surface (1); Subsequently, the virus is internalized into the cell via clathrin-mediated endocytosis (2); Then, clathrin-molecules dissociate from the vesicle and the virus is delivered to Rab5+ endosomes. Within the mildly acidic lumen of the endosome, the viral glycoproteins E2 and E1 undergo major conformational changes that lead to membrane fusion (3); Thereafter, the nucleocapsid core is released into the cytosol (4). The molecules and compounds that are known to interfere with entry are stated in the boxes.
Viruses 07 02792 g002

5. Molecular Mechanism of CHIKV Fusion

The molecular mechanisms involved in the membrane fusion process have been studied in great detail for SFV and SINV. The studies published thus far on CHIKV suggest that the molecular mechanisms involved in fusion are highly conserved between alphaviruses [26,135,136]. For example, like SFV and SINV [137,138,139], CHIKV can fuse with receptor-free liposomes, indicating that fusion is independent of a protein receptor [132,135]. The fusion process can be roughly divided in the following steps: (1) destabilization of the E2/E1 heterodimer, (2) integration of the E1 protein in the target membrane, (3) E1 trimerization, and (4) fusion pore formation (Figure 3).
Destabilization of the alphavirus E2/E1 heterodimer is triggered once the virus is exposed to the mildly acidic pH within the endosomes [26,140]. Histidines, which have a pkA of ~6–7, play a critical role in this process [25,141]. Recently, and in line with other alphaviruses [25], a series of highly conserved histidines within the envelope glycoproteins of CHIKV has been identified to control the pH-dependent conformational changes during fusion [136]. For example, E2-H170, a residue that is located within the acid sensitive region (ASR) of E2 (Figure 1, purple), becomes largely disordered at low pH [26,142]. Earlier work on SFV showed that protonation of this residue reduces the stability of the E2/E1 heterodimer by disabling the hydrogen bond with E1-S57. Once the E2/E1 interactions are loosened, the B domain of E2 moves away and the E1 fusion loop is exposed [26,29,143]. Thereafter, the E1 protein adopts an extended form and the hydrophobic fusion loop inserts into the target membrane (Figure 3) [17,144,145]. For SFV, this interaction is both low pH and cholesterol-dependent [144,146,147]. The presence of sphingomyelin strongly stimulates cholesterol- mediated E1 binding, but is not strictly required [147,148]. It is likely that these lipid interactions are similar for CHIKV, as cholesterol and sphingomyelin in the target membrane greatly enhance the fusion potential of CHIKV [132,135]. One of the amino acids important for lipid- and pH-sensing of SFV, SINV and CHIKV is situated at the E1-226 position [12,149,150,151]. This residue lies within the central DII domain in close proximity to the fusion loop (Figure 1) [25]. CHIKV strains with a valine instead of an alanine at the 226 position are more dependent on cholesterol and require a lower pH for infection [12,152] and fusion [132,153,154]. Another residue that has been reported to play a role in SFV lipid recognition is E1-V178 [155,156]. This residue is conserved among most alphaviruses, and experimental mutation of this residue to alanine leads to decreased cholesterol dependence of CHIKV fusion [154].
When the fusion peptide inserts into the target membrane, E2 is presumably still in association with the E1 molecules, as has been shown for SINV [17]. As the pH further decreases, the E2 molecules completely dissociate, which enables E1 trimerization [17,157]. A highly conserved histidine residue (E1-H3) is essential in regulating low-pH-induced trimerization [136,158,159]. For SFV, the first step in E1 trimerization involves the formation of a core trimer between DI and DII, which is dependent on low pH and most likely the presence of cholesterol and sphingomyelin in the target membrane [147,160,161,162]. Furthermore, sphingolipids have been proposed to play a role in stabilizing the E1 trimer [155]. After formation and stabilization of the core trimer, domain III re-folds back independently of pH towards the core trimer to form a hairpin-like homotrimer (Figure 3) [160,161]. This process brings the two opposing membranes together and forces merging of the outer membrane leaflets (hemifusion). Subsequently, a fusion pore is formed and expands, through which the nucleocapsid gains access to the cytosol [160,163]. For SFV, it has been shown that several E1 homotrimers assemble in a ring-like structure on the target membrane [158,160] with recent research indeed suggesting that for CHIKV fusion, several trimers need to act simultaneously to mediate membrane fusion [135].
Figure 3. Model of alphavirus membrane fusion [145]. (a) On a mature virion, 240 copies of E1 and E2 are arranged as 80 trimeric spikes; a single spike consisting of three E2/E1 heterodimers. Domains of E1 are colored as in figure 1; E2 is shown in gray. The E1 hydrophobic fusion loop (indicated as a star) is buried in a groove between domain A and domain B of E2; (b) Destabilization of the E2/E1 heterodimer is triggered once the virus is exposed to the mildly acidic pH. Domain B of E2 moves away and the E1 fusion loop is exposed; (c) Insertion of the fusion loop into the target membrane and dissociation of the E2 protein. Formation of a E1 core trimer between DI and DII; (d,e) Re-folding of E1 DIII and stem region to form a hairpin-like homotrimer, forcing the two opposing membranes together; (f) Merging of the opposing membrane leaflets (hemifusion); (g) Formation of the final stable homotrimer and opening of the fusion pore. (Figure reprinted with permission from Nature Reviews Microbiology).
Figure 3. Model of alphavirus membrane fusion [145]. (a) On a mature virion, 240 copies of E1 and E2 are arranged as 80 trimeric spikes; a single spike consisting of three E2/E1 heterodimers. Domains of E1 are colored as in figure 1; E2 is shown in gray. The E1 hydrophobic fusion loop (indicated as a star) is buried in a groove between domain A and domain B of E2; (b) Destabilization of the E2/E1 heterodimer is triggered once the virus is exposed to the mildly acidic pH. Domain B of E2 moves away and the E1 fusion loop is exposed; (c) Insertion of the fusion loop into the target membrane and dissociation of the E2 protein. Formation of a E1 core trimer between DI and DII; (d,e) Re-folding of E1 DIII and stem region to form a hairpin-like homotrimer, forcing the two opposing membranes together; (f) Merging of the opposing membrane leaflets (hemifusion); (g) Formation of the final stable homotrimer and opening of the fusion pore. (Figure reprinted with permission from Nature Reviews Microbiology).
Viruses 07 02792 g003

6. Inhibition of Early Events in Infection

Many studies have focused on the development of antiviral drugs to CHIKV. Antiviral treatment to reduce viremia is probably only possible in areas with hyper-endemic CHIKV activity. This as treatment should start quickly after disease onset given the short period of viremia. The development of antivirals is important as it may prevent the development of persistent disease. Indeed, mouse studies have indicated that antibody-based therapy might prevent persistent infection. Thus far, multiple inhibitors have been identified affecting different stages of the viral life cycle (extensively reviewed in [1,164]). Here we will discuss the compounds that specifically interfere with: (1) attachment of the virus to the target cell, (2) endocytosis, and (3) membrane fusion.

6.1. Interference with Virus-Receptor Binding

The green tea component epigallocatechin-3-gallate (EGCG) was found to inhibit CHIKV attachment and infection of HEK293T cells [165]. EGCG has a broad antiviral activity against numerous viruses, and presumably acts via binding competition with heparan sulfates and sialic acid [166]. Other plant-derived compounds that interfere with virus-receptor binding are flavagline, which act by binding to the CHIKV attachment factor PHB [88].
Another strategy involves the use of neutralizing monoclonal antibodies (MAbs) [167,168]. For CHIKV, the MAbs CHK-9, m242, and IM-CKV063 have been described to target the E2 putative receptor-binding domain A and have been proposed to prevent cellular binding [22,169]. Blocking infection through interference of virus-receptor binding is however challenging as CHIKV interacts with multiple attachments factors via distinct epitopes on E2 domain A and B. Furthermore, antibody-binding to CHIKV particles will target the immune-complex to Fc receptor-expressing cells which internalize the particle via interaction of the antibody to the Fc receptor. It remains to be investigated if antibodies that interfere with virus-receptor binding neutralize CHIKV infection in cells expressing the Fc receptor.

6.2. Interference with Endocytosis

In a large screen using a CHIKV replicon system, compounds with a 10-H phenothiazine structure including the licensed antipsychotic drug chlorpromazine were found active against CHIKV [170]. Chlorpromazine has been implicated to block the formation of clathrin-coated pits [171]. Since it is likely that CHIKV infects cells via this pathway it is of interest to evaluate if compounds with a 10-H phenothiazine structure have potential in antiviral treatment. However, since chlorpromazine is prescribed for various psychotic disorders, the psychological effects of these compounds should be monitored closely. In addition, single molecule therapy seems unlikely as CHIKV was found to enter cells via clathrin-dependent and clathrin-independent pathways. The anti-malaria drug chloroquine, which inhibits acidification of endosomes, was also found to hamper CHIKV infection [172,173,174,175]. Unfortunately, however, a double-blind placebo-controlled randomized clinical trial showed that chloroquine-treatment does not reduce viremia or the frequency of febrile arthralgia. In fact, an increased prevalence of persistent arthralgia was seen compared to the control group [172]. Given these results it seems doubtful that chloroquine-based therapy will be pursued in future studies.

6.3. Interference with Membrane Fusion

A pivotal step in membrane fusion is re-folding of the E1 DI/DII core trimer against the DIII stem [160,161]. Indeed, binding of exogenous recombinant E1-DIII proteins of SFV and CHIKV efficiently inhibit CHIKV membrane fusion and infection of BHK cells. Here, the presence of the stem region in the exogenous CHIKV E1-DIII proteins was a prerequisite for inhibition of membrane fusion [162]. Therefore, like for other alphaviruses, the E1 stem region has been proposed as a candidate target for small molecule inhibitors against CHIKV [162,176]. However, no specific drug that acts on this level has been described so far.
Another group of compounds that have been described to possess potent antiviral activity to CHIKV are arbidol (ARB) and it derivatives [177,178]. Arbidol (ARB) was originally licensed in Russia to treat and prevent Influenza infections. Escape mutant analysis and attachment assays revealed that ARB interferes with the early stages of CHIKV infection [177]. The precise mechanism underlying ARB activity remains to be elucidated, but earlier studies independent of CHIKV postulated that ARB functions through inhibition of virus-membrane fusion [179,180].
An alternative strategy to block membrane fusion is the use of MAbs. CHIKV MAb IM-CKV063 has been suggested to stabilize the E2/E1 trimeric spike as it binds to a conformational epitope spanning two E2 units [169]. Furthermore, the strongly neutralizing MAb C9 was found to target the ASR of E2 and has been predicted to prevent the conformational changes preceding membrane fusion [181]. MAb CHK-152 was shown to stabilize the E2 B domain, thereby inhibiting the exposure of the E1 fusion loop [22]. Indeed, functional studies revealed that CHK-152 abolishes membrane fusion activity of CHIKV. Importantly, this antibody also showed protective efficacy in mice and therefore is a candidate for antiviral therapy [182]. Thus, a variety of antibodies have been identified to interfere with infection and future studies should address which cocktail of antibodies has the largest therapeutic potential.

7. Future Perspectives and Concluding Remarks

Since the re-emergence of CHIKV about a decade ago, our understanding of the biology of the virus has greatly improved. Important progress has been made in defining the cells targeted and the pathways exploited by the virus to enter these cells. However, more research is required to fully elucidate the mechanisms of CHIKV infection, as some steps remain unclear.
For example, it will be important to identify the cells that serve as the main viral factories during the viremic period. Identification of the cells that produce most virus progeny will not only increase our understanding of CHIKV pathogenesis but will also guide the development of antiviral treatments. Furthermore, although we know that persistent CHIKV replication is associated with chronic arthralgia, the understanding of the underlying mechanism is poor. However, a growing body of evidence points towards an imperative role of an altered immune response initiated during acute infection in the development of chronic disease [55,62,183,184,185]. More in vivo studies in animal models will be required to gain more insight into the correlation of CHIKV tropism and viral persistence.
To date, several molecules have been described to facilitate CHIKV infection. However, most if not all molecules act as an attachment factor rather than an entry receptor. The entry receptor pivotal for CHIKV infection has yet to be identified. On the other hand, one can question whether such an entry receptor exists, considering the broad range of cells that can be infected and the fact that no receptor is required for membrane fusion. Based on the studies conducted thus far it becomes clear that CHIKV can utilize a variety of attachment factors and this may be sufficient to enter a cell.
The membrane fusion machinery of alphaviruses is largely defined. However, further fine-tuning is warranted as it may guide the development of compounds that interfere with infection. Mutagenesis studies already have enormously increased our understanding of the membrane fusion mechanism and will continue to do so. For dengue, an antibody has been identified that “traps” a fusion intermediate [186], and using autologous CHIKV antibodies represents another useful approach to further unravel the viral fusion mechanism.
CHIKV enters cells via clathrin-mediated endocytosis and fuses from within early endosomes. Interestingly, novel adaptations in the emerging IOL strains like the A226V mutation in E1 and substitutions in the acid-sensitive region of the E2 protein alter the pH-dependent membrane fusion properties of the virus. Furthermore, increased infection of mosquito midgut cells was observed, which likely led to the enhanced fitness of the virus in Ae. albopictus [153]. It will be interesting to investigate if there is a direct correlation between the higher infection rate and the altered pH-dependent membrane fusion properties. For example, the site of membrane fusion may be important for successful initiation of infection.
The only entry inhibitor tested so far in a clinical study, chloroquine, gave disappointing results [172]. Ongoing discovery of antiviral inhibitors is of utmost importance given the high burden of CHIKV infection. Antibody-based therapy is an attractive approach especially as it has been shown that it might prevent persistent infection in mice [187]. The proof-of-principle that antibody-based therapy is effective in humans is expected soon as a current clinical trial is evaluating the effect of anti-CHIKV serum antibodies in preventing severe disease in neonates [188]. This group of patients is of special interest for treatment as newborns are more likely to develop severe disease and can be treated during early stages of disease [1]. Antibody-based therapies likely always consist of a set of antibodies. This because CHIKV contains multiple receptor binding domains and it is unlikely that one antibody abolishes virus-receptor interaction in all CHIKV permissive cells. Furthermore, antibody-bound particles are internalized via Fc-receptors expressed on immune cells like macrophages, cells that facilitate CHIKV infection [37]. We hence postulate that antibodies targeting the membrane fusion machinery of the virus represent the most robust entry inhibitors. Therefore, focus should be laid on the further identification of neutralizing antibodies that interfere with membrane fusion.

Acknowledgments

This work was supported by the Dutch Organization for Scientific Research (NWO - Earth and Life Sciences; ASPASIA grant to J.M.S. and Veni 863.13.015 grant to I.A.R.-Z.) and by the University Medical Center Groningen (grant to M.K.S.v.D.-R.).

Author Contributions

M.K.S.v.D.-R., T.E.H., I.A.R.-Z., and J.M.S. wrote the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Thiberville, S.D.; Moyen, N.; Dupuis-Maguiraga, L.; Nougairede, A.; Gould, E.A.; Roques, P.; de Lamballerie, X. Chikungunya Fever: Epidemiology, Clinical Syndrome, Pathogenesis and Therapy. Antiviral Res. 2013, 99, 345–370. [Google Scholar] [CrossRef] [PubMed]
  2. Morrison, T.E. Reemergence of Chikungunya Virus. J. Virol. 2014, 88, 11644–11647. [Google Scholar] [CrossRef] [PubMed]
  3. Dupuis-Maguiraga, L.; Noret, M.; Brun, S.; Le Grand, R.; Gras, G.; Roques, P. Chikungunya Disease: Infection-Associated Markers from the Acute to the Chronic Phase of Arbovirus-Induced Arthralgia. PLoS Negl. Trop. Dis. 2012, 6, e1446. [Google Scholar] [CrossRef] [PubMed]
  4. Das, T.; Jaffar-Bandjee, M.C.; Hoarau, J.J.; Krejbich Trotot, P.; Denizot, M.; Lee-Pat-Yuen, G.; Sahoo, R.; Guiraud, P.; Ramful, D.; Robin, S.; et al. Chikungunya Fever: CNS Infection and Pathologies of a Re-Emerging Arbovirus. Prog. Neurobiol. 2010, 91, 121–129. [Google Scholar] [CrossRef] [PubMed]
  5. Pellot, A.S.; Alessandri, J.L.; Robin, S.; Samperiz, S.; Attali, T.; Brayer, C.; Pasquet, M.; Jaffar-Bandjee, M.C.; Benhamou, L.S.; Tiran-Rajaofera, I.; et al. Severe Forms of Chikungunya Virus Infection in a Pediatric Intensive Care Unit on Reunion Island. Med. Trop. (Mars) 2012, 72, 88–93. [Google Scholar] [PubMed]
  6. Staples, J.E.; Breiman, R.F.; Powers, A.M. Chikungunya Fever: An Epidemiological Review of a Re-Emerging Infectious Disease. Clin. Infect. Dis. 2009, 49, 942–948. [Google Scholar] [CrossRef] [PubMed]
  7. Weaver, S.C.; Lecuit, M. Chikungunya Virus and the Global Spread of a Mosquito-Borne Disease. N. Engl. J. Med. 2015, 372, 1231–1239. [Google Scholar] [PubMed]
  8. Pan American Health Organization. Number of Reported Cases of Chikungunya Fever in the Americas, by Country Or Territory 2013–2015 (to Week Noted) Epidemiological Week/EW 17. (Updated as of 1 May 2015)Available online: https://clinicaltri-als.gov/ct2/show/NCT02230163?term=Chikungunya&rank=5 or http://www.webcita-tion.org/6YE7aebST. (accessed on 2 May 2015).
  9. Volk, S.M.; Chen, R.; Tsetsarkin, K.A.; Adams, A.P.; Garcia, T.I.; Sall, A.A.; Nasar, F.; Schuh, A.J.; Holmes, E.C.; Higgs, S.; et al. Genome-Scale Phylogenetic Analyses of Chikungunya Virus Reveal Independent Emergences of Recent Epidemics and various Evolutionary Rates. J. Virol. 2010, 84, 6497–6504. [Google Scholar] [CrossRef] [PubMed]
  10. Weaver, S.C.; Forrester, N.L. Chikungunya: Evolutionary History and Recent Epidemic Spread. Antiviral Res. 2015, 120, 32–39. [Google Scholar] [CrossRef] [PubMed]
  11. Tsetsarkin, K.A.; Chen, R.; Yun, R.; Rossi, S.L.; Plante, K.S.; Guerbois, M.; Forrester, N.; Perng, G.C.; Sreekumar, E.; Leal, G.; et al. Multi-Peaked Adaptive Landscape for Chikungunya Virus Evolution Predicts Continued Fitness Optimization in Aedes Albopictus Mosquitoes. Nat. Commun. 2014, 5, 4084. [Google Scholar] [CrossRef] [PubMed]
  12. Tsetsarkin, K.A.; Vanlandingham, D.L.; McGee, C.E.; Higgs, S. A Single Mutation in Chikungunya Virus Affects Vector Specificity and Epidemic Potential. PLoS Pathog. 2007, 3, e201. [Google Scholar] [CrossRef] [PubMed]
  13. Weaver, S.C. Arrival of Chikungunya Virus in the New World: Prospects for Spread and Impact on Public Health. PLoS Negl Trop. Dis. 2014, 8, e2921. [Google Scholar] [CrossRef] [PubMed]
  14. Powers, A.M.; Brault, A.C.; Shirako, Y.; Strauss, E.G.; Kang, W.; Strauss, J.H.; Weaver, S.C. Evolutionary Relationships and Systematics of the Alphaviruses. J. Virol. 2001, 75, 10118–10131. [Google Scholar] [CrossRef] [PubMed]
  15. Jose, J.; Snyder, J.E.; Kuhn, R.J. A Structural and Functional Perspective of Alphavirus Replication and Assembly. Future Microbiol. 2009, 4, 837–856. [Google Scholar] [CrossRef] [PubMed]
  16. Simizu, B.; Yamamoto, K.; Hashimoto, K.; Ogata, T. Structural Proteins of Chikungunya Virus. J. Virol. 1984, 51, 254–258. [Google Scholar] [PubMed]
  17. Cao, S.; Zhang, W. Characterization of an Early-Stage Fusion Intermediate of Sindbis Virus using Cryoelectron Microscopy. Proc. Natl. Acad. Sci. USA 2013, 110, 13362–13367. [Google Scholar] [CrossRef] [PubMed]
  18. Allan, D.; Quinn, P. Membrane Phospholipid Asymmetry in Semliki Forest Virus Grown in BHK Cells. Biochim. Biophys. Acta Biomembr. 1989, 987, 199–204. [Google Scholar] [CrossRef]
  19. Renkonen, O.; Kaarainen, L.; Simons, K.; Gahmberg, C.G. The Lipid Class Composition of Semliki Forest Virus and Plasma Membranes of the Host Cells. Virology 1971, 46, 318–326. [Google Scholar] [CrossRef]
  20. Van Meer, G.; Simons, K.; Op den Kamp, J.A.; van Deenen, L.M. Phospholipid Asymmetry in Semliki Forest Virus Grown on Baby Hamster Kidney (BHK-21) Cells. Biochemistry 1981, 20, 1974–1981. [Google Scholar] [CrossRef] [PubMed]
  21. Kalvodova, L.; Sampaio, J.L.; Cordo, S.; Ejsing, C.S.; Shevchenko, A.; Simons, K. The Lipidomes of Vesicular Stomatitis Virus, Semliki Forest Virus, and the Host Plasma Membrane Analyzed by Quantitative Shotgun Mass Spectrometry. J. Virol. 2009, 83, 7996–8003. [Google Scholar] [CrossRef] [PubMed]
  22. Sun, S.; Xiang, Y.; Akahata, W.; Holdaway, H.; Pal, P.; Zhang, X.; Diamond, M.S.; Nabel, G.J.; Rossmann, M.G. Structural Analyses at Pseudo Atomic Resolution of Chikungunya Virus and Antibodies show Mechanisms of Neutralization. eLife 2013, 2, e00435. [Google Scholar] [CrossRef] [PubMed]
  23. Mukhopadhyay, S.; Zhang, W.; Gabler, S.; Chipman, P.R.; Strauss, E.G.; Strauss, J.H.; Baker, T.S.; Kuhn, R.J.; Rossmann, M.G. Mapping the Structure and Function of the E1 and E2 Glycoproteins in Alphaviruses. Structure 2006, 14, 63–73. [Google Scholar] [CrossRef] [PubMed]
  24. Lescar, J.; Roussel, A.; Wien, M.W.; Navaza, J.; Fuller, S.D.; Wengler, G.; Wengler, G.; Rey, F.A. The Fusion Glycoprotein Shell of Semliki Forest Virus: An Icosahedral Assembly Primed for Fusogenic Activation at Endosomal pH. Cell 2001, 105, 137–148. [Google Scholar] [CrossRef]
  25. Roussel, A.; Lescar, J.; Vaney, M.C.; Wengler, G.; Wengler, G.; Rey, F.A. Structure and Interactions at the Viral Surface of the Envelope Protein E1 of Semliki Forest Virus. Structure 2006, 14, 75–86. [Google Scholar] [CrossRef] [PubMed]
  26. Voss, J.E.; Vaney, M.C.; Duquerroy, S.; Vonrhein, C.; Girard-Blanc, C.; Crublet, E.; Thompson, A.; Bricogne, G.; Rey, F.A. Glycoprotein Organization of Chikungunya Virus Particles Revealed by X-ray Crystallography. Nature 2010, 468, 709–712. [Google Scholar] [CrossRef] [PubMed]
  27. Jose, J.; Przybyla, L.; Edwards, T.J.; Perera, R.; Burgner, J.W., 2nd; Kuhn, R.J. Interactions of the Cytoplasmic Domain of Sindbis Virus E2 with Nucleocapsid Cores Promote Alphavirus Budding. J. Virol. 2012, 86, 2585–2599. [Google Scholar] [CrossRef] [PubMed]
  28. Zhang, R.; Hryc, C.F.; Cong, Y.; Liu, X.; Jakana, J.; Gorchakov, R.; Baker, M.L.; Weaver, S.C.; Chiu, W. 4.4 A Cryo-EM Structure of an Enveloped Alphavirus Venezuelan Equine Encephalitis Virus. EMBO J. 2011, 30, 3854–3863. [Google Scholar] [CrossRef] [PubMed]
  29. Li, L.; Jose, J.; Xiang, Y.; Kuhn, R.J.; Rossmann, M.G. Structural Changes of Envelope Proteins during Alphavirus Fusion. Nature 2010, 468, 705–708. [Google Scholar] [CrossRef] [PubMed]
  30. Vogel, R.H.; Provencher, S.W.; von Bonsdorff, C.H.; Adrian, M.; Dubochet, J. Envelope Structure of Semliki Forest Virus Reconstructed from Cryo-Electron Micrographs. Nature 1986, 320, 533–535. [Google Scholar] [CrossRef] [PubMed]
  31. Pletnev, S.V.; Zhang, W.; Mukhopadhyay, S.; Fisher, B.R.; Hernandez, R.; Brown, D.T.; Baker, T.S.; Rossmann, M.G.; Kuhn, R.J. Locations of Carbohydrate Sites on Alphavirus Glycoproteins show that E1 Forms an Icosahedral Scaffold. Cell 2001, 105, 127–136. [Google Scholar] [CrossRef]
  32. Cheng, R.H.; Kuhn, R.J.; Olson, N.H.; Rossmann, M.G.; Choi, H.K.; Smith, T.J.; Baker, T.S. Nucleocapsid and Glycoprotein Organization in an Enveloped Virus. Cell 1995, 80, 621–630. [Google Scholar] [CrossRef]
  33. Reiter, P.; Fontenille, D.; Paupy, C. Aedes Albopictus as an Epidemic Vector of Chikungunya Virus: Another Emerging Problem? Lancet Infect. Dis. 2006, 6, 463–464. [Google Scholar] [CrossRef]
  34. Ramasubramanian, M.K.; Barham, O.M.; Swaminathan, V. Mechanics of a Mosquito Bite with Applications to Microneedle Design. Bioinspir. Biomim. 2008, 3. [Google Scholar] [CrossRef] [PubMed]
  35. Weaver, S.C.; Osorio, J.E.; Livengood, J.A.; Chen, R.; Stinchcomb, D.T. Chikungunya Virus and Prospects for a Vaccine. Expert Rev. Vaccines 2012, 11, 1087–1101. [Google Scholar] [CrossRef] [PubMed]
  36. Sourisseau, M.; Schilte, C.; Casartelli, N.; Trouillet, C.; Guivel-Benhassine, F.; Rudnicka, D.; Sol-Foulon, N.; Le Roux, K.; Prevost, M.C.; Fsihi, H.; et al. Characterization of Reemerging Chikungunya Virus. PLoS Pathog. 2007, 3, e89. [Google Scholar] [CrossRef] [PubMed]
  37. Solignat, M.; Gay, B.; Higgs, S.; Briant, L.; Devaux, C. Replication Cycle of Chikungunya: A Re-Emerging Arbovirus. Virology 2009, 393, 183–197. [Google Scholar] [CrossRef] [PubMed]
  38. Wikan, N.; Sakoonwatanyoo, P.; Ubol, S.; Yoksan, S.; Smith, D.R. Chikungunya Virus Infection of Cell Lines: Analysis of the East, Central and South African Lineage. PLoS ONE 2012, 7, e31102. [Google Scholar] [CrossRef] [PubMed]
  39. Chow, A.; Her, Z.; Ong, E.K.; Chen, J.M.; Dimatatac, F.; Kwek, D.J.; Barkham, T.; Yang, H.; Renia, L.; Leo, Y.S.; et al. Persistent Arthralgia Induced by Chikungunya Virus Infection is Associated with Interleukin-6 and Granulocyte Macrophage Colony-Stimulating Factor. J. Infect. Dis. 2011, 203, 149–157. [Google Scholar] [CrossRef] [PubMed]
  40. Naze, F.; Le Roux, K.; Schuffenecker, I.; Zeller, H.; Staikowsky, F.; Grivard, P.; Michault, A.; Laurent, P. Simultaneous Detection and Quantitation of Chikungunya, Dengue and West Nile Viruses by Multiplex RT-PCR Assays and Dengue Virus Typing using High Resolution Melting. J. Virol. Methods 2009, 162, 1–7. [Google Scholar] [CrossRef] [PubMed]
  41. Ng, L.C.; Tan, L.K.; Tan, C.H.; Tan, S.S.; Hapuarachchi, H.C.; Pok, K.Y.; Lai, Y.L.; Lam-Phua, S.G.; Bucht, G.; Lin, R.T.; et al. Entomologic and Virologic Investigation of Chikungunya, Singapore. Emerg. Infect. Dis. 2009, 15, 1243–1249. [Google Scholar] [CrossRef] [PubMed]
  42. Her, Z.; Malleret, B.; Chan, M.; Ong, E.K.; Wong, S.C.; Kwek, D.J.; Tolou, H.; Lin, R.T.; Tambyah, P.A.; Renia, L.; et al. Active Infection of Human Blood Monocytes by Chikungunya Virus Triggers an Innate Immune Response. J. Immunol. 2010, 184, 5903–5913. [Google Scholar] [CrossRef] [PubMed]
  43. Fraser, J.R.; Becker, G.J. Mononuclear Cell Types in Chronic Synovial Effusions of Ross River Virus Disease. Aust. N. Z. J. Med. 1984, 14, 505–506. [Google Scholar] [CrossRef] [PubMed]
  44. Gardner, C.L.; Burke, C.W.; Tesfay, M.Z.; Glass, P.J.; Klimstra, W.B.; Ryman, K.D. Eastern and Venezuelan Equine Encephalitis Viruses Differ in their Ability to Infect Dendritic Cells and Macrophages: Impact of Altered Cell Tropism on Pathogenesis. J. Virol. 2008, 82, 10634–10646. [Google Scholar] [CrossRef] [PubMed]
  45. Hidmark, A.S.; McInerney, G.M.; Nordstrom, E.K.; Douagi, I.; Werner, K.M.; Liljestrom, P.; Karlsson Hedestam, G.B. Early Alpha/Beta Interferon Production by Myeloid Dendritic Cells in Response to UV-Inactivated Virus Requires Viral Entry and Interferon Regulatory Factor 3 but Not MyD88. J. Virol. 2005, 79, 10376–10385. [Google Scholar] [CrossRef] [PubMed]
  46. Shabman, R.S.; Morrison, T.E.; Moore, C.; White, L.; Suthar, M.S.; Hueston, L.; Rulli, N.; Lidbury, B.; Ting, J.P.; Mahalingam, S.; et al. Differential Induction of Type I Interferon Responses in Myeloid Dendritic Cells by Mosquito and Mammalian-Cell-Derived Alphaviruses. J. Virol. 2007, 81, 237–247. [Google Scholar] [CrossRef] [PubMed]
  47. Schilte, C.; Couderc, T.; Chretien, F.; Sourisseau, M.; Gangneux, N.; Guivel-Benhassine, F.; Kraxner, A.; Tschopp, J.; Higgs, S.; Michault, A.; et al. Type I IFN Controls Chikungunya Virus Via its Action on Nonhematopoietic Cells. J. Exp. Med. 2010, 207, 429–442. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Vanlandingham, D.L.; Hong, C.; Klingler, K.; Tsetsarkin, K.; McElroy, K.L.; Powers, A.M.; Lehane, M.J.; Higgs, S. Differential Infectivities of O’Nyong-Nyong and Chikungunya Virus Isolates in Anopheles Gambiae and Aedes Aegypti Mosquitoes. Am. J. Trop. Med. Hyg. 2005, 72, 616–621. [Google Scholar] [PubMed]
  49. Kumar, S.; Jaffar-Bandjee, M.C.; Giry, C.; Connen de Kerillis, L.; Merits, A.; Gasque, P.; Hoarau, J.J. Mouse Macrophage Innate Immune Response to Chikungunya Virus Infection. Virol. J. 2012, 9. [Google Scholar] [CrossRef] [PubMed]
  50. Issac, T.H.; Tan, E.L.; Chu, J.J. Proteomic Profiling of Chikungunya Virus-Infected Human Muscle Cells: Reveal the Role of Cytoskeleton Network in CHIKV Replication. J. Proteomics 2014, 108, 445–464. [Google Scholar] [CrossRef] [PubMed]
  51. Couderc, T.; Lecuit, M. Focus on Chikungunya Pathophysiology in Human and Animal Models. Microbes Infect. 2009, 11, 1197–1205. [Google Scholar] [CrossRef] [PubMed]
  52. Noret, M.; Herrero, L.; Rulli, N.; Rolph, M.; Smith, P.N.; Li, R.W.; Roques, P.; Gras, G.; Mahalingam, S. Interleukin 6, RANKL, and Osteoprotegerin Expression by Chikungunya Virus-Infected Human Osteoblasts. J. Infect. Dis. 2012, 206, 457–459. [Google Scholar] [CrossRef] [PubMed]
  53. Chen, W.; Foo, S.S.; Rulli, N.E.; Taylor, A.; Sheng, K.C.; Herrero, L.J.; Herring, B.L.; Lidbury, B.A.; Li, R.W.; Walsh, N.C.; et al. Arthritogenic Alphaviral Infection Perturbs Osteoblast Function and Triggers Pathologic Bone Loss. Proc. Natl. Acad. Sci. USA 2014, 111, 6040–6045. [Google Scholar] [CrossRef] [PubMed]
  54. Ozden, S.; Huerre, M.; Riviere, J.P.; Coffey, L.L.; Afonso, P.V.; Mouly, V.; de Monredon, J.; Roger, J.C.; El Amrani, M.; Yvin, J.L.; et al. Human Muscle Satellite Cells as Targets of Chikungunya Virus Infection. PLoS ONE 2007, 2, e527. [Google Scholar] [CrossRef] [PubMed]
  55. Couderc, T.; Chretien, F.; Schilte, C.; Disson, O.; Brigitte, M.; Guivel-Benhassine, F.; Touret, Y.; Barau, G.; Cayet, N.; Schuffenecker, I.; et al. A Mouse Model for Chikungunya: Young Age and Inefficient Type-I Interferon Signaling are Risk Factors for Severe Disease. PLoS Pathog. 2008, 4, e29. [Google Scholar] [CrossRef] [PubMed]
  56. Brighton, S.W.; Simson, I.W. A Destructive Arthropathy Following Chikungunya Virus Arthritis—A Possible Association. Clin. Rheumatol. 1984, 3, 253–258. [Google Scholar] [CrossRef] [PubMed]
  57. Marimoutou, C.; Ferraro, J.; Javelle, E.; Deparis, X.; Simon, F. Chikungunya Infection: Self-Reported Rheumatic Morbidity and Impaired Quality of Life Persist 6 Years Later. Clin. Microbiol. Infect. 2015, 21, 688–693. [Google Scholar] [CrossRef] [PubMed]
  58. Mathew, A.J.; Goyal, V.; George, E.; Thekkemuriyil, D.V.; Jayakumar, B.; Chopra, A.; Trivandrum COPCORD Study Group. Rheumatic-Musculoskeletal Pain and Disorders in a Naive Group of Individuals 15 Months Following a Chikungunya Viral Epidemic in South India: A Population Based Observational Study. Int. J. Clin. Pract. 2011, 65, 1306–1312. [Google Scholar] [CrossRef] [PubMed]
  59. Foissac, M.; Javelle, E.; Ray, S.; Guerin, B.; Simon, F. Post-Chikungunya Rheumatoid Arthritis, Saint Martin. Emerg. Infect. Dis. 2015, 21, 530–532. [Google Scholar] [CrossRef] [PubMed]
  60. Queyriaux, B.; Simon, F.; Grandadam, M.; Michel, R.; Tolou, H.; Boutin, J.P. Clinical Burden of Chikungunya Virus Infection. Lancet Infect. Dis. 2008, 8, 2–3. [Google Scholar] [CrossRef]
  61. Tang, B.L. The Cell Biology of Chikungunya Virus Infection. Cell. Microbiol. 2012, 14, 1354–1363. [Google Scholar] [CrossRef] [PubMed]
  62. Chen, W.; Foo, S.S.; Sims, N.A.; Herrero, L.J.; Walsh, N.C.; Mahalingam, S. Arthritogenic Alphaviruses: New Insights into Arthritis and Bone Pathology. Trends Microbiol. 2015, 23, 35–43. [Google Scholar] [CrossRef] [PubMed]
  63. Labadie, K.; Larcher, T.; Joubert, C.; Mannioui, A.; Delache, B.; Brochard, P.; Guigand, L.; Dubreil, L.; Lebon, P.; Verrier, B.; et al. Chikungunya Disease in Nonhuman Primates Involves Long-Term Viral Persistence in Macrophages. J. Clin. Investig. 2010, 120, 894–906. [Google Scholar] [CrossRef] [PubMed]
  64. Malvy, D.; Ezzedine, K.; Mamani-Matsuda, M.; Autran, B.; Tolou, H.; Receveur, M.C.; Pistone, T.; Rambert, J.; Moynet, D.; Mossalayi, D. Destructive Arthritis in a Patient with Chikungunya Virus Infection with Persistent Specific IgM Antibodies. BMC Infect. Dis. 2009, 9. [Google Scholar] [CrossRef] [PubMed]
  65. Wielanek, A.C.; Monredon, J.D.; Amrani, M.E.; Roger, J.C.; Serveaux, J.P. Guillain-Barre Syndrome Complicating a Chikungunya Virus Infection. Neurology 2007, 69, 2105–2107. [Google Scholar] [CrossRef] [PubMed]
  66. Lebrun, G.; Chadda, K.; Reboux, A.H.; Martinet, O.; Gauzere, B.A. Guillain-Barre Syndrome After Chikungunya Infection. Emerg. Infect. Dis. 2009, 15, 495–496. [Google Scholar] [CrossRef] [PubMed]
  67. Chusri, S.; Siripaitoon, P.; Hirunpat, S.; Silpapojakul, K. Case Reports of Neuro-Chikungunya in Southern Thailand. Am. J. Trop. Med. Hyg. 2011, 85, 386–389. [Google Scholar] [CrossRef] [PubMed]
  68. Grivard, P.; Le Roux, K.; Laurent, P.; Fianu, A.; Perrau, J.; Gigan, J.; Hoarau, G.; Grondin, N.; Staikowsky, F.; Favier, F.; et al. Molecular and Serological Diagnosis of Chikungunya Virus Infection. Pathol. Biol. (Paris) 2007, 55, 490–494. [Google Scholar] [CrossRef] [PubMed]
  69. Wintachai, P.; Wikan, N.; Kuadkitkan, A.; Jaimipuk, T.; Ubol, S.; Pulmanausahakul, R.; Auewarakul, P.; Kasinrerk, W.; Weng, W.Y.; Panyasrivanit, M.; et al. Identification of Prohibitin as a Chikungunya Virus Receptor Protein. J. Med. Virol. 2012, 84, 1757–1770. [Google Scholar] [CrossRef] [PubMed]
  70. Abraham, R.; Mudaliar, P.; Padmanabhan, A.; Sreekumar, E. Induction of Cytopathogenicity in Human Glioblastoma Cells by Chikungunya Virus. PLoS ONE 2013, 8, e75854. [Google Scholar] [CrossRef] [PubMed]
  71. Lim, P.J.; Chu, J.J. A Polarized Cell Model for Chikungunya Virus Infection: Entry and Egress of Virus Occurs at the Apical Domain of Polarized Cells. PLoS Negl. Trop. Dis. 2014, 8, e2661. [Google Scholar] [CrossRef] [PubMed]
  72. Haywood, A.M. Virus Receptors: Binding, Adhesion Strengthening, and Changes in Viral Structure. J. Virol. 1994, 68, 1–5. [Google Scholar] [PubMed]
  73. Mercer, J.; Schelhaas, M.; Helenius, A. Virus Entry by Endocytosis. Annu. Rev. Biochem. 2010, 79, 803–833. [Google Scholar] [CrossRef] [PubMed]
  74. Smith, T.J.; Cheng, R.H.; Olson, N.H.; Peterson, P.; Chase, E.; Kuhn, R.J.; Baker, T.S. Putative Receptor Binding Sites on Alphaviruses as Visualized by Cryoelectron Microscopy. Proc. Natl. Acad. Sci. USA 1995, 92, 10648–10652. [Google Scholar] [CrossRef] [PubMed]
  75. Ashbrook, A.W.; Burrack, K.S.; Silva, L.A.; Montgomery, S.A.; Heise, M.T.; Morrison, T.E.; Dermody, T.S. Residue 82 of the Chikungunya Virus E2 Attachment Protein Modulates Viral Dissemination and Arthritis in Mice. J. Virol. 2014, 88, 12180–12192. [Google Scholar] [CrossRef] [PubMed]
  76. Asnet Mary, J.; Paramasivan, R.; Tyagi, B.K.; Surender, M.; Shenbagarathai, R. Identification of Structural Motifs in the E2 Glycoprotein of Chikungunya Involved in Virus-Host Interaction. J. Biomol. Struct. Dyn. 2013, 31, 1077–1085. [Google Scholar] [CrossRef] [PubMed]
  77. Ye, F.; Zhang, M. Structures and Target Recognition Modes of PDZ Domains: Recurring Themes and Emerging Pictures. Biochem. J. 2013, 455, 1–14. [Google Scholar] [CrossRef] [PubMed]
  78. Yan, Y.; Vasudevan, S.; Nguyen, H.; Bork, U.; Sitaraman, S.; Merlin, D. Extracellular Interaction between hCD98 and the PDZ Class II Domain of hCASK in Intestinal Epithelia. J. Membr. Biol. 2007, 215, 15–26. [Google Scholar] [CrossRef] [PubMed]
  79. Moller-Tank, S.; Kondratowicz, A.S.; Davey, R.A.; Rennert, P.D.; Maury, W. Role of the Phosphatidylserine Receptor TIM-1 in Enveloped-Virus Entry. J. Virol. 2013, 87, 8327–8341. [Google Scholar] [CrossRef] [PubMed]
  80. Silva, L.A.; Khomandiak, S.; Ashbrook, A.W.; Weller, R.; Heise, M.T.; Morrison, T.E.; Dermody, T.S. A Single-Amino-Acid Polymorphism in Chikungunya Virus E2 Glycoprotein Influences Glycosaminoglycan Utilization. J. Virol. 2014, 88, 2385–2397. [Google Scholar] [CrossRef] [PubMed]
  81. Fongsaran, C.; Jirakanwisal, K.; Kuadkitkan, A.; Wikan, N.; Wintachai, P.; Thepparit, C.; Ubol, S.; Phaonakrop, N.; Roytrakul, S.; Smith, D.R. Involvement of ATP Synthase Beta Subunit in Chikungunya Virus Entry into Insect Cells. Arch. Virol. 2014, 159, 3353–3364. [Google Scholar] [CrossRef] [PubMed]
  82. Marsh, M.; Helenius, A. Virus Entry: Open Sesame. Cell 2006, 124, 729–740. [Google Scholar] [CrossRef] [PubMed]
  83. Salvador, B.; Zhou, Y.; Michault, A.; Muench, M.O.; Simmons, G. Characterization of Chikungunya Pseudotyped Viruses: Identification of Refractory Cell Lines and Demonstration of Cellular Tropism Differences Mediated by Mutations in E1 Glycoprotein. Virology 2009, 393, 33–41. [Google Scholar] [CrossRef] [PubMed]
  84. Chowdhury, I.; Thompson, W.E.; Thomas, K. Prohibitins Role in Cellular Survival through Ras-Raf-MEK-ERK Pathway. J. Cell. Physiol. 2014, 229, 998–1004. [Google Scholar] [CrossRef] [PubMed]
  85. Hossen, M.N.; Kajimoto, K.; Akita, H.; Hyodo, M.; Harashima, H. Vascular-Targeted Nanotherapy for Obesity: Unexpected Passive Targeting Mechanism to Obese Fat for the Enhancement of Active Drug Delivery. J. Control. Release 2012, 163, 101–110. [Google Scholar] [CrossRef] [PubMed]
  86. Kolonin, M.G.; Saha, P.K.; Chan, L.; Pasqualini, R.; Arap, W. Reversal of Obesity by Targeted Ablation of Adipose Tissue. Nat. Med. 2004, 10, 625–632. [Google Scholar] [CrossRef] [PubMed]
  87. Merkwirth, C.; Langer, T. Prohibitin Function within Mitochondria: Essential Roles for Cell Proliferation and Cristae Morphogenesis. Biochim. Biophys. Acta 2009, 1793, 27–32. [Google Scholar] [CrossRef] [PubMed]
  88. Wintachai, P.; Thuaud, F.; Basmadjian, C.; Roytrakul, S.; Ubol, S.; Desaubry, L.; Smith, D.R. Assessment of Flavaglines as Potential Chikungunya Virus Entry Inhibitors. Microbiol. Immunol. 2015, 59, 129–141. [Google Scholar] [CrossRef] [PubMed]
  89. Umetsu, S.E.; Lee, W.L.; McIntire, J.J.; Downey, L.; Sanjanwala, B.; Akbari, O.; Berry, G.J.; Nagumo, H.; Freeman, G.J.; Umetsu, D.T.; et al. TIM-1 Induces T Cell Activation and Inhibits the Development of Peripheral Tolerance. Nat. Immunol. 2005, 6, 447–454. [Google Scholar] [CrossRef] [PubMed]
  90. Zhu, C.; Anderson, A.C.; Schubart, A.; Xiong, H.; Imitola, J.; Khoury, S.J.; Zheng, X.X.; Strom, T.B.; Kuchroo, V.K. The Tim-3 Ligand Galectin-9 Negatively Regulates T Helper Type 1 Immunity. Nat. Immunol. 2005, 6, 1245–1252. [Google Scholar] [CrossRef] [PubMed]
  91. Kobayashi, N.; Karisola, P.; Pena-Cruz, V.; Dorfman, D.M.; Jinushi, M.; Umetsu, S.E.; Butte, M.J.; Nagumo, H.; Chernova, I.; Zhu, B.; et al. TIM-1 and TIM-4 Glycoproteins Bind Phosphatidylserine and Mediate Uptake of Apoptotic Cells. Immunity 2007, 27, 927–940. [Google Scholar] [CrossRef] [PubMed]
  92. Kondratowicz, A.S.; Lennemann, N.J.; Sinn, P.L.; Davey, R.A.; Hunt, C.L.; Moller-Tank, S.; Meyerholz, D.K.; Rennert, P.; Mullins, R.F.; Brindley, M.; et al. T-Cell Immunoglobulin and Mucin Domain 1 (TIM-1) is a Receptor for Zaire Ebolavirus and Lake Victoria Marburgvirus. Proc. Natl. Acad. Sci. USA 2011, 108, 8426–8431. [Google Scholar] [CrossRef] [PubMed]
  93. Moller-Tank, S.; Albritton, L.M.; Rennert, P.D.; Maury, W. Characterizing Functional Domains for TIM-Mediated Enveloped Virus Entry. J. Virol. 2014, 88, 6702–6713. [Google Scholar] [CrossRef] [PubMed]
  94. Gandhi, N.S.; Mancera, R.L. The Structure of Glycosaminoglycans and their Interactions with Proteins. Chem. Biol. Drug Des. 2008, 72, 455–482. [Google Scholar] [CrossRef] [PubMed]
  95. Klimstra, W.B.; Ryman, K.D.; Johnston, R.E. Adaptation of Sindbis Virus to BHK Cells Selects for use of Heparan Sulfate as an Attachment Receptor. J. Virol. 1998, 72, 7357–7366. [Google Scholar] [PubMed]
  96. Bernard, K.A.; Klimstra, W.B.; Johnston, R.E. Mutations in the E2 Glycoprotein of Venezuelan Equine Encephalitis Virus Confer Heparan Sulfate Interaction, Low Morbidity, and Rapid Clearance from Blood of Mice. Virology 2000, 276, 93–103. [Google Scholar] [CrossRef] [PubMed]
  97. Gardner, C.L.; Choi-Nurvitadhi, J.; Sun, C.; Bayer, A.; Hritz, J.; Ryman, K.D.; Klimstra, W.B. Natural Variation in the Heparan Sulfate Binding Domain of the Eastern Equine Encephalitis Virus E2 Glycoprotein Alters Interactions with Cell Surfaces and Virulence in Mice. J. Virol. 2013, 87, 8582–8590. [Google Scholar] [CrossRef] [PubMed]
  98. Smit, J.M.; Waarts, B.L.; Kimata, K.; Klimstra, W.B.; Bittman, R.; Wilschut, J. Adaptation of Alphaviruses to Heparan Sulfate: Interaction of Sindbis and Semliki Forest Viruses with Liposomes Containing Lipid-Conjugated Heparin. J. Virol. 2002, 76, 10128–10137. [Google Scholar] [CrossRef] [PubMed]
  99. Bear, J.S.; Byrnes, A.P.; Griffin, D.E. Heparin-Binding and Patterns of Virulence for Two Recombinant Strains of Sindbis Virus. Virology 2006, 347, 183–190. [Google Scholar] [CrossRef] [PubMed]
  100. Wang, E.; Brault, A.C.; Powers, A.M.; Kang, W.; Weaver, S.C. Glycosaminoglycan Binding Properties of Natural Venezuelan Equine Encephalitis Virus Isolates. J. Virol. 2003, 77, 1204–1210. [Google Scholar] [CrossRef] [PubMed]
  101. Heil, M.L.; Albee, A.; Strauss, J.H.; Kuhn, R.J. An Amino Acid Substitution in the Coding Region of the E2 Glycoprotein Adapts Ross River Virus to Utilize Heparan Sulfate as an Attachment Moiety. J. Virol. 2001, 75, 6303–6309. [Google Scholar] [CrossRef] [PubMed]
  102. Gardner, C.L.; Burke, C.W.; Higgs, S.T.; Klimstra, W.B.; Ryman, K.D. Interferon-Alpha/Beta Deficiency Greatly Exacerbates Arthritogenic Disease in Mice Infected with Wild-Type Chikungunya Virus but Not with the Cell Culture-Adapted Live-Attenuated 181/25 Vaccine Candidate. Virology 2012, 425, 103–112. [Google Scholar] [CrossRef] [PubMed]
  103. Levitt, N.H.; Ramsburg, H.H.; Hasty, S.E.; Repik, P.M.; Cole, F.E., Jr.; Lupton, H.W. Development of an Attenuated Strain of Chikungunya Virus for use in Vaccine Production. Vaccine 1986, 4, 157–162. [Google Scholar] [CrossRef]
  104. Gardner, C.L.; Hritz, J.; Sun, C.; Vanlandingham, D.L.; Song, T.Y.; Ghedin, E.; Higgs, S.; Klimstra, W.B.; Ryman, K.D. Deliberate Attenuation of Chikungunya Virus by Adaptation to Heparan Sulfate-Dependent Infectivity: A Model for Rational Arboviral Vaccine Design. PLoS Negl. Trop. Dis. 2014, 8, e2719. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Gorchakov, R.; Wang, E.; Leal, G.; Forrester, N.L.; Plante, K.; Rossi, S.L.; Partidos, C.D.; Adams, A.P.; Seymour, R.L.; Weger, J.; et al. Attenuation of Chikungunya Virus Vaccine Strain 181/Clone 25 is Determined by Two Amino Acid Substitutions in the E2 Envelope Glycoprotein. J. Virol. 2012, 86, 6084–6096. [Google Scholar] [CrossRef] [PubMed]
  106. National Center for Biotechnology Information. ATP5B ATP Synthase, H+ Transporting, Mitochondrial F1 Complex, Beta Polypeptide [Homo Sapiens (Human)]; National Center for Biotechnology Information: Bethesda, MD, USA, 2015; Available online: http://www.ncbi.nlm.nih.gov/gene/506 or http://www.webcitation.org/6YE9a0jmy; (accessed on 2 May 2015). [Google Scholar]
  107. Fu, Y.; Hou, Y.; Fu, C.; Gu, M.; Li, C.; Kong, W.; Wang, X.; Shyy, J.Y.; Zhu, Y. A Novel Mechanism of Gamma/Delta T-Lymphocyte and Endothelial Activation by Shear Stress: The Role of Ecto-ATP Synthase Beta Chain. Circ. Res. 2011, 108, 410–417. [Google Scholar] [CrossRef] [PubMed]
  108. Champagne, E.; Martinez, L.O.; Collet, X.; Barbaras, R. Ecto-F1Fo ATP Synthase/F1 ATPase: Metabolic and Immunological Functions. Curr. Opin. Lipidol. 2006, 17, 279–284. [Google Scholar] [CrossRef] [PubMed]
  109. Takada, Y.; Ye, X.; Simon, S. The Integrins. Genome Biol. 2007, 8, e215. [Google Scholar] [CrossRef] [PubMed]
  110. Fraisier, C.; Koraka, P.; Belghazi, M.; Bakli, M.; Granjeaud, S.; Pophillat, M.; Lim, S.M.; Osterhaus, A.; Martina, B.; Camoin, L.; et al. Kinetic Analysis of Mouse Brain Proteome Alterations Following Chikungunya Virus Infection before and After Appearance of Clinical Symptoms. PLoS ONE 2014, 9, e91397. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Mathias, P.; Wickham, T.; Moore, M.; Nemerow, G. Multiple Adenovirus Serotypes use Alpha V Integrins for Infection. J. Virol. 1994, 68, 6811–6814. [Google Scholar] [PubMed]
  112. Li, E.; Brown, S.L.; Stupack, D.G.; Puente, X.S.; Cheresh, D.A.; Nemerow, G.R. Integrin Alpha(V)Beta1 is an Adenovirus Coreceptor. J. Virol. 2001, 75, 5405–5409. [Google Scholar] [CrossRef] [PubMed]
  113. La Linn, M.; Eble, J.A.; Lubken, C.; Slade, R.W.; Heino, J.; Davies, J.; Suhrbier, A. An Arthritogenic Alphavirus Uses the alpha1beta1 Integrin Collagen Receptor. Virology 2005, 336, 229–239. [Google Scholar] [CrossRef] [PubMed]
  114. Schmidt, K.; Keller, M.; Bader, B.L.; Korytar, T.; Finke, S.; Ziegler, U.; Groschup, M.H. Integrins Modulate the Infection Efficiency of West Nile Virus into Cells. J. Gen. Virol. 2013, 94, 1723–1733. [Google Scholar] [CrossRef] [PubMed]
  115. Nakamura, H.; Minegishi, H. HSP60 as a Drug Target. Curr. Pharm. Des. 2013, 19, 441–451. [Google Scholar] [CrossRef] [PubMed]
  116. Soltys, B.J.; Gupta, R.S. Cell Surface Localization of the 60 kDa Heat Shock Chaperonin Protein (hsp60) in Mammalian Cells. Cell Biol. Int. 1997, 21, 315–320. [Google Scholar] [CrossRef] [PubMed]
  117. Belles, C.; Kuhl, A.; Nosheny, R.; Carding, S.R. Plasma Membrane Expression of Heat Shock Protein 60 in vivo in Response to Infection. Infect. Immun. 1999, 67, 4191–4200. [Google Scholar] [PubMed]
  118. Apte-Deshpande, A.D.; Paingankar, M.S.; Gokhale, M.D.; Deobagkar, D.N. Serratia Odorifera Mediated Enhancement in Susceptibility of Aedes Aegypti for Chikungunya Virus. Indian J. Med. Res. 2014, 139, 762–768. [Google Scholar] [PubMed]
  119. Padwad, Y.S.; Mishra, K.P.; Jain, M.; Chanda, S.; Karan, D.; Ganju, L. RNA Interference Mediated Silencing of Hsp60 Gene in Human Monocytic Myeloma Cell Line U937 Revealed Decreased Dengue Virus Multiplication. Immunobiology 2009, 214, 422–429. [Google Scholar] [CrossRef] [PubMed]
  120. Vancini, R.; Wang, G.; Ferreira, D.; Hernandez, R.; Brown, D.T. Alphavirus Genome Delivery Occurs Directly at the Plasma Membrane in a Time- and Temperature-Dependent Process. J. Virol. 2013, 87, 4352–4359. [Google Scholar] [CrossRef] [PubMed]
  121. Paredes, A.M.; Ferreira, D.; Horton, M.; Saad, A.; Tsuruta, H.; Johnston, R.; Klimstra, W.; Ryman, K.; Hernandez, R.; Chiu, W.; et al. Conformational Changes in Sindbis Virions Resulting from Exposure to Low pH and Interactions with Cells Suggest that Cell Penetration may Occur at the Cell Surface in the Absence of Membrane Fusion. Virology 2004, 324, 373–386. [Google Scholar] [CrossRef] [PubMed]
  122. Benmerah, A.; Lamaze, C. Clathrin-Coated Pits: Vive La Difference? Traffic 2007, 8, 970–982. [Google Scholar] [CrossRef] [PubMed]
  123. Kirchhausen, T.; Owen, D.; Harrison, S.C. Molecular Structure, Function, and Dynamics of Clathrin-Mediated Membrane Traffic. Cold Spring Harb. Perspect. Biol. 2014, 6. [Google Scholar] [CrossRef] [PubMed]
  124. Colpitts, T.M.; Moore, A.C.; Kolokoltsov, A.A.; Davey, R.A. Venezuelan Equine Encephalitis Virus Infection of Mosquito Cells Requires Acidification as Well as Mosquito Homologs of the Endocytic Proteins Rab5 and Rab7. Virology 2007, 369, 78–91. [Google Scholar] [CrossRef] [PubMed]
  125. Marsh, M.; Bolzau, E.; Helenius, A. Penetration of Semliki Forest Virus from Acidic Prelysosomal Vacuoles. Cell 1983, 32, 931–940. [Google Scholar] [CrossRef]
  126. Kolokoltsov, A.A.; Fleming, E.H.; Davey, R.A. Venezuelan Equine Encephalitis Virus Entry Mechanism Requires Late Endosome Formation and Resists Cell Membrane Cholesterol Depletion. Virology 2006, 347, 333–342. [Google Scholar] [CrossRef] [PubMed]
  127. Yamauchi, Y.; Helenius, A. Virus Entry at a Glance. J. Cell. Sci. 2013, 126, 1289–1295. [Google Scholar] [CrossRef] [PubMed]
  128. Gold, E.S.; Underhill, D.M.; Morrissette, N.S.; Guo, J.; McNiven, M.A.; Aderem, A. Dynamin 2 is Required for Phagocytosis in Macrophages. J. Exp. Med. 1999, 190, 1849–1856. [Google Scholar] [CrossRef] [PubMed]
  129. Bernard, E.; Solignat, M.; Gay, B.; Chazal, N.; Higgs, S.; Devaux, C.; Briant, L. Endocytosis of Chikungunya Virus into Mammalian Cells: Role of Clathrin and Early Endosomal Compartments. PLoS ONE 2010, 5, e11479. [Google Scholar] [CrossRef] [PubMed]
  130. Sigismund, S.; Woelk, T.; Puri, C.; Maspero, E.; Tacchetti, C.; Transidico, P.; Di Fiore, P.P.; Polo, S. Clathrin-Independent Endocytosis of Ubiquitinated Cargos. Proc. Natl. Acad. Sci. USA 2005, 102, 2760–2765. [Google Scholar] [CrossRef] [PubMed]
  131. Ooi, Y.S.; Stiles, K.M.; Liu, C.Y.; Taylor, G.M.; Kielian, M. Genome-Wide RNAi Screen Identifies Novel Host Proteins Required for Alphavirus Entry. PLoS Pathog. 2013, 9, e1003835. [Google Scholar] [CrossRef] [PubMed]
  132. Hoornweg, T.E.; van Duijl-Richter, M.K.S.; Ayala Nuñez, N.V.; van Hemert, M.J.; Smit, J.M. Dynamics of Chikungunya Virus Cell Entry Unraveled by Single Virus Tracking in Living Cells. Submitted.
  133. Lee, R.C.; Hapuarachchi, H.C.; Chen, K.C.; Hussain, K.M.; Chen, H.; Low, S.L.; Ng, L.C.; Lin, R.; Ng, M.M.; Chu, J.J. Mosquito Cellular Factors and Functions in Mediating the Infectious Entry of Chikungunya Virus. PLoS Negl. Trop. Dis. 2013, 7, e2050. [Google Scholar] [CrossRef] [PubMed]
  134. The Murphy Lab. Differences in Early Endosomal pH between Cell Types. Available online: http://murphylab.web.cmu.edu/projects/endosomal-pH-references.html or http://www.webcita-tion.org/6YFhQFTB1. (accessed on 3 May 2015).
  135. Van Duijl-Richter, M.; Blijleven, J.; van Oijen, A.; Smit, J. Chikungunya Virus Fusion Properties Elucidated by Single-Particle and Bulk Approaches. J. Gen. Virol. 2015. [Google Scholar] [CrossRef] [PubMed]
  136. Zeng, X.; Mukhopadhyay, S.; Brooks, C.L., 3rd. Residue-Level Resolution of Alphavirus Envelope Protein Interactions in pH-Dependent Fusion. Proc. Natl. Acad. Sci. USA 2015, 112, 2034–2039. [Google Scholar] [CrossRef] [PubMed]
  137. White, J.; Helenius, A. pH-Dependent Fusion between the Semliki Forest Virus Membrane and Liposomes. Proc. Natl. Acad. Sci. USA 1980, 77, 3273–3277. [Google Scholar] [CrossRef] [PubMed]
  138. Bron, R.; Wahlberg, J.M.; Garoff, H.; Wilschut, J. Membrane Fusion of Semliki Forest Virus in a Model System: Correlation between Fusion Kinetics and Structural Changes in the Envelope Glycoprotein. EMBO J. 1993, 12, 693–701. [Google Scholar] [PubMed]
  139. Smit, J.M.; Bittman, R.; Wilschut, J. Low-pH-Dependent Fusion of Sindbis Virus with Receptor-Free Cholesterol- and Sphingolipid-Containing Liposomes. J. Virol. 1999, 73, 8476–8484. [Google Scholar] [PubMed]
  140. Wahlberg, J.M.; Boere, W.A.; Garoff, H. The Heterodimeric Association between the Membrane Proteins of Semliki Forest Virus Changes its Sensitivity to Low pH during Virus Maturation. J. Virol. 1989, 63, 4991–4997. [Google Scholar] [PubMed]
  141. Kampmann, T.; Mueller, D.S.; Mark, A.E.; Young, P.R.; Kobe, B. The Role of Histidine Residues in Low-pH-Mediated Viral Membrane Fusion. Structure 2006, 14, 1481–1487. [Google Scholar] [CrossRef] [PubMed]
  142. Fields, W.; Kielian, M. A Key Interaction between the Alphavirus Envelope Proteins Responsible for Initial Dimer Dissociation during Fusion. J. Virol. 2013, 87, 3774–3781. [Google Scholar] [CrossRef] [PubMed]
  143. Hammar, L.; Markarian, S.; Haag, L.; Lankinen, H.; Salmi, A.; Cheng, R.H. Prefusion Rearrangements Resulting in Fusion Peptide Exposure in Semliki Forest Virus. J. Biol. Chem. 2003, 278, 7189–7198. [Google Scholar] [CrossRef] [PubMed]
  144. Gibbons, D.L.; Ahn, A.; Liao, M.; Hammar, L.; Cheng, R.H.; Kielian, M. Multistep Regulation of Membrane Insertion of the Fusion Peptide of Semliki Forest Virus. J. Virol. 2004, 78, 3312–3318. [Google Scholar] [CrossRef] [PubMed]
  145. Kielian, M.; Rey, F.A. Virus Membrane-Fusion Proteins: More than One Way to make a Hairpin. Nat. Rev. Microbiol. 2006, 4, 67–76. [Google Scholar] [CrossRef] [PubMed]
  146. Ahn, A.; Gibbons, D.L.; Kielian, M. The Fusion Peptide of Semliki Forest Virus Associates with Sterol-Rich Membrane Domains. J. Virol. 2002, 76, 3267–3275. [Google Scholar] [CrossRef] [PubMed]
  147. Klimjack, M.R.; Jeffrey, S.; Kielian, M. Membrane and Protein Interactions of a Soluble Form of the Semliki Forest Virus Fusion Protein. J. Virol. 1994, 68, 6940–6946. [Google Scholar] [PubMed]
  148. Nieva, J.L.; Bron, R.; Corver, J.; Wilschut, J. Membrane Fusion of Semliki Forest Virus Requires Sphingolipids in the Target Membrane. EMBO J. 1994, 13, 2797–2804. [Google Scholar] [PubMed]
  149. Vashishtha, M.; Phalen, T.; Marquardt, M.T.; Ryu, J.S.; Ng, A.C.; Kielian, M. A Single Point Mutation Controls the Cholesterol Dependence of Semliki Forest Virus Entry and Exit. J. Cell Biol. 1998, 140, 91–99. [Google Scholar] [CrossRef] [PubMed]
  150. Chatterjee, P.K.; Vashishtha, M.; Kielian, M. Biochemical Consequences of a Mutation that Controls the Cholesterol Dependence of Semliki Forest Virus Fusion. J. Virol. 2000, 74, 1623–1631. [Google Scholar] [CrossRef] [PubMed]
  151. Lu, Y.E.; Cassese, T.; Kielian, M. The Cholesterol Requirement for Sindbis Virus Entry and Exit and Characterization of a Spike Protein Region Involved in Cholesterol Dependence. J. Virol. 1999, 73, 4272–4278. [Google Scholar] [PubMed]
  152. Gay, B.; Bernard, E.; Solignat, M.; Chazal, N.; Devaux, C.; Briant, L. pH-Dependent Entry of Chikungunya Virus into Aedes Albopictus Cells. Infect. Genet. Evol. 2012, 12, 1275–1281. [Google Scholar] [CrossRef] [PubMed]
  153. Tsetsarkin, K.A.; Weaver, S.C. Sequential Adaptive Mutations Enhance Efficient Vector Switching by Chikungunya Virus and its Epidemic Emergence. PLoS Pathog. 2011, 7, e1002412. [Google Scholar] [CrossRef] [PubMed]
  154. Kuo, S.C.; Chen, Y.J.; Wang, Y.M.; Tsui, P.Y.; Kuo, M.D.; Wu, T.Y.; Lo, S.J. Cell-Based Analysis of Chikungunya Virus E1 Protein in Membrane Fusion. J. Biomed. Sci. 2012, 19, e44. [Google Scholar] [CrossRef] [PubMed]
  155. Chatterjee, P.K.; Eng, C.H.; Kielian, M. Novel Mutations that Control the Sphingolipid and Cholesterol Dependence of the Semliki Forest Virus Fusion Protein. J. Virol. 2002, 76, 12712–12722. [Google Scholar] [CrossRef] [PubMed]
  156. Teissier, E.; Pecheur, E.I. Lipids as Modulators of Membrane Fusion Mediated by Viral Fusion Proteins. Eur. Biophys. J. 2007, 36, 887–899. [Google Scholar] [CrossRef] [PubMed]
  157. Wahlberg, J.M.; Bron, R.; Wilschut, J.; Garoff, H. Membrane Fusion of Semliki Forest Virus Involves Homotrimers of the Fusion Protein. J. Virol. 1992, 66, 7309–7318. [Google Scholar] [PubMed]
  158. Zheng, Y.; Sanchez-San Martin, C.; Qin, Z.L.; Kielian, M. The Domain I-Domain III Linker Plays an Important Role in the Fusogenic Conformational Change of the Alphavirus Membrane Fusion Protein. J. Virol. 2011, 85, 6334–6342. [Google Scholar] [CrossRef] [PubMed]
  159. Qin, Z.L.; Zheng, Y.; Kielian, M. Role of Conserved Histidine Residues in the Low-pH Dependence of the Semliki Forest Virus Fusion Protein. J. Virol. 2009, 83, 4670–4677. [Google Scholar] [CrossRef] [PubMed]
  160. Gibbons, D.L.; Vaney, M.C.; Roussel, A.; Vigouroux, A.; Reilly, B.; Lepault, J.; Kielian, M.; Rey, F.A. Conformational Change and Protein-Protein Interactions of the Fusion Protein of Semliki Forest Virus. Nature 2004, 427, 320–325. [Google Scholar] [CrossRef] [PubMed]
  161. Sanchez-San Martin, C.; Sosa, H.; Kielian, M. A Stable Prefusion Intermediate of the Alphavirus Fusion Protein Reveals Critical Features of Class II Membrane Fusion. Cell Host Microbe 2008, 4, 600–608. [Google Scholar] [CrossRef] [PubMed]
  162. Sanchez-San Martin, C.; Nanda, S.; Zheng, Y.; Fields, W.; Kielian, M. Cross-Inhibition of Chikungunya Virus Fusion and Infection by Alphavirus E1 Domain III Proteins. J. Virol. 2013, 87, 7680–7687. [Google Scholar] [CrossRef] [PubMed]
  163. Wengler, G.; Koschinski, A.; Wengler, G.; Repp, H. During Entry of Alphaviruses, the E1 Glycoprotein Molecules Probably Form Two Separate Populations that Generate either a Fusion Pore Or Ion-Permeable Pores. J. Gen. Virol. 2004, 85, 1695–1701. [Google Scholar] [CrossRef] [PubMed]
  164. Kaur, P.; Chu, J.J. Chikungunya Virus: An Update on Antiviral Development and Challenges. Drug Discov. Today 2013, 18, 969–983. [Google Scholar] [CrossRef] [PubMed]
  165. Weber, C.; Sliva, K.; von Rhein, C.; Kummerer, B.M.; Schnierle, B.S. The Green Tea Catechin, Epigallocatechin Gallate Inhibits Chikungunya Virus Infection. Antiviral Res. 2015, 113, 1–3. [Google Scholar] [CrossRef] [PubMed]
  166. Colpitts, C.C.; Schang, L.M. A Small Molecule Inhibits Virion Attachment to Heparan Sulfate- or Sialic Acid-Containing Glycans. J. Virol. 2014, 88, 7806–7817. [Google Scholar] [CrossRef] [PubMed]
  167. Couderc, T.; Khandoudi, N.; Grandadam, M.; Visse, C.; Gangneux, N.; Bagot, S.; Prost, J.F.; Lecuit, M. Prophylaxis and Therapy for Chikungunya Virus Infection. J. Infect. Dis. 2009, 200, 516–523. [Google Scholar] [CrossRef] [PubMed]
  168. Fric, J.; Bertin-Maghit, S.; Wang, C.I.; Nardin, A.; Warter, L. Use of Human Monoclonal Antibodies to Treat Chikungunya Virus Infection. J. Infect. Dis. 2013, 207, 319–322. [Google Scholar] [CrossRef] [PubMed]
  169. Fong, R.H.; Banik, S.S.; Mattia, K.; Barnes, T.; Tucker, D.; Liss, N.; Lu, K.; Selvarajah, S.; Srinivasan, S.; Mabila, M.; et al. Exposure of Epitope Residues on the Outer Face of the Chikungunya Virus Envelope Trimer Determines Antibody Neutralizing Efficacy. J. Virol. 2014, 88, 14364–14379. [Google Scholar] [CrossRef] [PubMed]
  170. Pohjala, L.; Utt, A.; Varjak, M.; Lulla, A.; Merits, A.; Ahola, T.; Tammela, P. Inhibitors of Alphavirus Entry and Replication Identified with a Stable Chikungunya Replicon Cell Line and Virus-Based Assays. PLoS ONE 2011, 6, e28923. [Google Scholar] [CrossRef] [PubMed]
  171. Gastaminza, P.; Whitten-Bauer, C.; Chisari, F.V. Unbiased Probing of the Entire Hepatitis C Virus Life Cycle Identifies Clinical Compounds that Target Multiple Aspects of the Infection. Proc. Natl. Acad. Sci. USA 2010, 107, 291–296. [Google Scholar] [CrossRef] [PubMed]
  172. De Lamballerie, X.; Boisson, V.; Reynier, J.C.; Enault, S.; Charrel, R.N.; Flahault, A.; Roques, P.; Le Grand, R. On Chikungunya Acute Infection and Chloroquine Treatment. Vector Borne Zoonotic Dis. 2008, 8, 837–839. [Google Scholar] [CrossRef] [PubMed]
  173. Khan, M.; Santhosh, S.R.; Tiwari, M.; Lakshmana Rao, P.V.; Parida, M. Assessment of in vitro Prophylactic and Therapeutic Efficacy of Chloroquine Against Chikungunya Virus in Vero Cells. J. Med. Virol. 2010, 82, 817–824. [Google Scholar] [CrossRef] [PubMed]
  174. Delogu, I.; de Lamballerie, X. Chikungunya Disease and Chloroquine Treatment. J. Med. Virol. 2011, 83, 1058–1059. [Google Scholar] [CrossRef] [PubMed]
  175. Nuckols, J.T.; McAuley, A.J.; Huang, Y.J.; Horne, K.M.; Higgs, S.; Davey, R.A.; Vanlandingham, D.L. pH-Dependent Entry of Chikungunya Virus Fusion into Mosquito Cells. Virol. J. 2014, 11. [Google Scholar] [CrossRef] [PubMed]
  176. Liao, M.; Kielian, M. Domain III from Class II Fusion Proteins Functions as a Dominant-Negative Inhibitor of Virus Membrane Fusion. J. Cell Biol. 2005, 171, 111–120. [Google Scholar] [CrossRef] [PubMed]
  177. Delogu, I.; Pastorino, B.; Baronti, C.; Nougairede, A.; Bonnet, E.; de Lamballerie, X. In vitro Antiviral Activity of Arbidol Against Chikungunya Virus and Characteristics of a Selected Resistant Mutant. Antiviral Res. 2011, 90, 99–107. [Google Scholar] [CrossRef] [PubMed]
  178. Di Mola, A.; Peduto, A.; La Gatta, A.; Delang, L.; Pastorino, B.; Neyts, J.; Leyssen, P.; de Rosa, M.; Filosa, R. Structure-Activity Relationship Study of Arbidol Derivatives as Inhibitors of Chikungunya Virus Replication. Bioorg. Med. Chem. 2014, 22, 6014–6025. [Google Scholar] [CrossRef] [PubMed]
  179. Villalain, J. Membranotropic Effects of Arbidol, a Broad Anti-Viral Molecule, on Phospholipid Model Membranes. J. Phys. Chem. B 2010, 114, 8544–8554. [Google Scholar] [CrossRef] [PubMed]
  180. Boriskin, Y.S.; Leneva, I.A.; Pecheur, E.I.; Polyak, S.J. Arbidol: A Broad-Spectrum Antiviral Compound that Blocks Viral Fusion. Curr. Med. Chem. 2008, 15, 997–1005. [Google Scholar] [CrossRef] [PubMed]
  181. Selvarajah, S.; Sexton, N.R.; Kahle, K.M.; Fong, R.H.; Mattia, K.A.; Gardner, J.; Lu, K.; Liss, N.M.; Salvador, B.; Tucker, D.F.; et al. A Neutralizing Monoclonal Antibody Targeting the Acid-Sensitive Region in Chikungunya Virus E2 Protects from Disease. PLoS Negl. Trop. Dis. 2013, 7, e2423. [Google Scholar] [CrossRef] [PubMed]
  182. Pal, P.; Dowd, K.A.; Brien, J.D.; Edeling, M.A.; Gorlatov, S.; Johnson, S.; Lee, I.; Akahata, W.; Nabel, G.J.; Richter, M.K.; et al. Development of a Highly Protective Combination Monoclonal Antibody Therapy Against Chikungunya Virus. PLoS Pathog. 2013, 9, e1003312. [Google Scholar] [CrossRef] [PubMed]
  183. Long, K.M.; Whitmore, A.C.; Ferris, M.T.; Sempowski, G.D.; McGee, C.; Trollinger, B.; Gunn, B.; Heise, M.T. Dendritic Cell Immunoreceptor Regulates Chikungunya Virus Pathogenesis in Mice. J. Virol. 2013, 87, 5697–5706. [Google Scholar] [CrossRef] [PubMed]
  184. Linn, M.L.; Aaskov, J.G.; Suhrbier, A. Antibody-Dependent Enhancement and Persistence in Macrophages of an Arbovirus Associated with Arthritis. J. Gen. Virol. 1996, 77, 407–411. [Google Scholar] [CrossRef] [PubMed]
  185. Ryman, K.D.; Klimstra, W.B. Closing the Gap between Viral and Noninfectious Arthritis. Proc. Natl. Acad. Sci. USA 2014, 111, 5767–5768. [Google Scholar] [CrossRef] [PubMed]
  186. Zhang, X.; Sheng, J.; Austin, S.K.; Hoornweg, T.E.; Smit, J.M.; Kuhn, R.J.; Diamond, M.S.; Rossmann, M.G. Structure of Acidic pH Dengue Virus Showing the Fusogenic Glycoprotein Trimers. J. Virol. 2015, 89, 743–750. [Google Scholar] [CrossRef] [PubMed]
  187. Hawman, D.W.; Stoermer, K.A.; Montgomery, S.A.; Pal, P.; Oko, L.; Diamond, M.S.; Morrison, T.E. Chronic Joint Disease Caused by Persistent Chikungunya Virus Infection is Controlled by the Adaptive Immune Response. J. Virol. 2013, 87, 13878–13888. [Google Scholar] [CrossRef] [PubMed]
  188. U.S. National Institutes of Health. Clinical Evaluation of Anti-CHIKV Hyperimmune Intravenous Immunoglobulins (CHIKVIVG-01). Available online: https://clinicaltri-als.gov/ct2/show/NCT02230163?term=Chikungunya&rank=5 or http://www.webcita-tion.org/6YE7aebST . (accessed on 2 May 2015).

Share and Cite

MDPI and ACS Style

Van Duijl-Richter, M.K.S.; Hoornweg, T.E.; Rodenhuis-Zybert, I.A.; Smit, J.M. Early Events in Chikungunya Virus Infection—From Virus CellBinding to Membrane Fusion. Viruses 2015, 7, 3647-3674. https://doi.org/10.3390/v7072792

AMA Style

Van Duijl-Richter MKS, Hoornweg TE, Rodenhuis-Zybert IA, Smit JM. Early Events in Chikungunya Virus Infection—From Virus CellBinding to Membrane Fusion. Viruses. 2015; 7(7):3647-3674. https://doi.org/10.3390/v7072792

Chicago/Turabian Style

Van Duijl-Richter, Mareike K. S., Tabitha E. Hoornweg, Izabela A. Rodenhuis-Zybert, and Jolanda M. Smit. 2015. "Early Events in Chikungunya Virus Infection—From Virus CellBinding to Membrane Fusion" Viruses 7, no. 7: 3647-3674. https://doi.org/10.3390/v7072792

Article Metrics

Back to TopTop